Vous êtes sur la page 1sur 236

Actinides

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Wed, 07 May 2014 06:47:20 UTC
Contents
Articles
Overview 1
Actinide 1

Elements 27
89 – Actinium 27

90 – Thorium 35

91 – Protactinium 49

92 – Uranium 60

93 – Neptunium 80

94 – Plutonium 88

95 – Americium 108

96 – Curium 124

97 – Berkelium 138

98 – Californium 150

99 – Einsteinium 160

100 – Fermium
173
101 – Mendelevium
181
102 – Nobelium
186
103 – Lawrencium
195

Miscellany 203
Actinides in the environment 203
Major actinide 206
Minor actinide 206

References
Article Sources and Contributors 208
Image Sources, Licenses and Contributors 212

Article Licenses
License 215
1

Overview

Actinide
Actinides in the periodic table

The actinide /ˈæktɨnaɪd/ or actinoid /ˈæktɨnɔɪd/ (IUPAC


nomenclature) series encompasses the 15 metallic chemical
elements with atomic numbers from 89 to 103, actinium
[2][3]
through lawrencium.
The actinide series derives its name from the group 3
element actinium. The informal chemical symbol An is
used in general discussions of actinide chemistry to refer to
any actinide. All but one of the actinides are f-block
elements, corresponding to the filling of the 5f electron
shell; lawrencium, a d-block element, is also generally
considered an actinide. In comparison with the lanthanides,
also mostly f-block elements, the actinides show much more
variable valence.

Of the actinides, primordial thorium and uranium occur


naturally in substantial quantities and small amounts of
persisting natural plutonium have also been identified. The
radioactive decay of uranium produces transient amounts of The atomic bomb dropped on Nagasaki had a plutonium
[1]
actinium and protactinium, and atoms of neptunium, charge.
americium, curium, berkelium and californium are
occasionally produced from transmutation reactions in uranium ores. The other actinides are purely synthetic
[4]
elements. Nuclear weapons tests have released at least six actinides heavier than plutonium into the environment;
analysis of debris from a 1952 hydrogen bomb explosion showed the presence of americium, curium, berkelium,
californium, einsteinium and fermium.

All actinides are radioactive and release energy upon radioactive decay; naturally occurring uranium and thorium,
and synthetically produced plutonium are the most abundant actinides on Earth. These are used in nuclear reactors
and nuclear weapons. Uranium and thorium also have diverse current or historical uses, and americium is used in the
ionization chambers of most modern smoke detectors.
Actinide 1

In presentations of the periodic table, the lanthanides and the actinides are customarily shown as two additional
rows below the main body of the table, with placeholders or else a selected single element of each series
(either lanthanum or lutetium, and either actinium or lawrencium, respectively) shown in a single cell of the
main table, between barium and hafnium, and radium and rutherfordium, respectively. This convention is
entirely a matter of aesthetics and formatting practicality; a rarely used wide-formatted periodic table inserts the
lanthanide and actinide series in their proper places, as parts of the table's sixth and seventh rows (periods).

Actinides

ctinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrenciu
89 90 91 92 93 94 95 96 97 98 99 101 102 103
Ac Th Pa U Np Pu Am Cm Bk Cf 100 Md No Lr

Es Fm

Primordial From decay Synthetic Border shows natural occurrence of the element

Discovery, isolation and synthesis

Synthesis of transuranium elements[5][6]


Element Year Method

Neptunium 1940 238


Bombarding U by neutrons

Plutonium 1941 238


Bombarding U by deuterons

Americium 1944 239


Bombarding Pu by neutrons

Curium 1944 239


Bombarding Pu by α-particles

Berkelium 1949 241


Bombarding Am by α-particles

Californium 1950 242


Bombarding Cm by α-particles

Einsteinium 1952 As a product of nuclear explosion

Fermium 1952 As a product of nuclear explosion

Mendelevium 1955 253


Bombarding Es by α-particles

Nobelium 1965 243


Bombarding Am by
15 238
N or U with α-
particles
Lawrencium 1961–1971 252 10
Bombarding Cf by B or
11 243 18
B and of Am with O

Like the lanthanides, the actinides form a family of elements with similar properties. Within the actinides, there are
two overlapping groups: transuranium elements, which follow uranium in the periodic table—and transplutonium
elements, which follow plutonium. Compared to the lanthanides, which (except for promethium) are found in nature
in appreciable quantities, most actinides are rare. The most abundant, or easy to synthesize actinides are uranium
[7]
and thorium, followed by plutonium, americium, actinium, protactinium and neptunium.
The existence of transuranium elements was suggested by Enrico Fermi based on his experiments in 1934. However,
even though four actinides were known by that time, it was not yet understood that they formed a family similar to
lanthanides. The prevailing view that dominated early research into transuranics was that they were regular elements
in the 7th period, with thorium, protactinium and uranium corresponding to 6th-period hafnium, tantalum and
tungsten, respectively. Synthesis of transuranics gradually undermined this point of view. By 1944 an observation
that curium failed to exhibit oxidation states above 4 (whereas its supposed 6th period neighbor, platinum, can reach
oxidation state of 7) prompted Glenn Seaborg to formulate a so-called "actinide hypothesis". Studies of known
actinides and discoveries of further transuranic elements provided more data in support of this point of view, but the
phrase "actinide hypothesis" (the implication being that "hypothesis" is something that has not been decisively
proven) remained in active use by scientists through the late 1950s.
At present, there are two major methods of producing isotopes of transplutonium elements: irradiation of the lighter
elements with either neutrons or accelerated charged particles. The first method is most important for applications, as
only neutron irradiation using nuclear reactors allows the production of sizeable amounts of synthetic actinides;
however, it is limited to relatively light elements. The advantage of the second method is that elements heavier than
[8]
plutonium, as well as neutron-deficient isotopes, can be obtained, which are not formed during neutron irradiation.
In 1962–1966, there were attempts in the United States to produce transplutonium isotopes using a series of six
underground nuclear explosions. Small samples of rock were extracted from the blast area immediately after the test
to study the explosion products, but no isotopes with mass number greater than 257 could be detected, despite
predictions that such isotopes would have relatively long half-lives of α-decay. This inobservation was attributed to
spontaneous fission owing to the large speed of the products and to other decay channels, such as neutron emission
[9]
and nuclear fission.

From actinium to neptunium


Uranium and thorium were the first actinides discovered. Uranium was
identified in 1789 by the German chemist Martin Heinrich Klaproth in
pitchblende ore. He named it after the planet Uranus, which had been
discovered only eight years earlier. Klaproth was able to precipitate a
yellow compound (likely sodium diuranate) by dissolving pitchblende
in nitric acid and neutralizing the solution with sodium hydroxide. He
then reduced the obtained yellow powder with charcoal, and extracted
a black substance that he mistook for metal. Only 60 years later, the
French scientist Eugène-Melchior Péligot identified it with uranium
oxide. He also isolated the first sample of uranium metal by heating
uranium tetrachloride with potassium. The atomic mass of uranium
was then calculated as 120, but Dmitri Mendeleev in 1872 corrected it
to 240 using his periodicity laws. This value was confirmed
[10]
experimentally in 1882 by K. Zimmerman.
Enrico Fermi suggested the existence
of transuranium elements in 1934. Thorium oxide was discovered by Friedrich Wöhler in the mineral,
[11]
which was found in Norway (1827). Jöns Jacob Berzelius
characterized this material in more detail by in 1828. By reduction of thorium tetrachloride with potassium, he
[12]
isolated the metal and named it thorium after the Norse god of thunder and lightning Thor. The same isolation
method was later used by Péligot for uranium.
Actinium was discovered in 1899 by André-Louis Debierne, an assistant of Marie Curie, in the pitchblende waste
left after removal of radium and polonium. He described the substance (in 1899) as similar to titanium and (in 1900)
as similar to thorium. The discovery of actinium by Debierne was however questioned in 1971 and 2000, arguing
that Debierne's publications in 1904 contradicted his earlier work of 1899–1900. The name actinium comes from the
Greek aktis, aktinos (ακτίς, ακτίνος), meaning beam or ray. This metal was discovered not by its own radiation but
[13]
by the radiation of the daughter products. Owing to the close similarity of actinium and lanthanum and low
abundance, pure actinium could only be produced in 1950. The term actinide was probably introduced by Victor
Goldschmidt in 1937.
Protactinium was possibly isolated in 1900 by William Crookes. It was first identified in 1913, when Kasimir Fajans
234m
and Oswald Helmuth Göhring encountered the short-lived isotope Pa (half-life 1.17 minutes) during their
238
studies of the U decay. They named the new element brevium (from Latin brevis meaning brief); the name was
changed to protoactinium (from Greek πρῶτος + ἀκτίς meaning "first beam element") in 1918 when two groups of
scientists, led by the Austrian Lise Meitner and Otto Hahn of Germany and Frederick Soddy and John Cranston of
231
Great Britain, independently discovered Pa. The name was shortened to Protactinium in 1949. This element was
little characterized until 1960, when A. G. Maddock and co-workers in UK produced 130 grams of protactinium
[14]
from 60 tonnes of waste left after extraction of uranium from its ore.
Neptunium (named for the planet Neptune, the next planet out from Uranus, after which uranium was named)
was discovered by Edwin McMillan and Philip H. Abelson in 1940 in Berkeley, California. They produced the
239
Np isotope (half-life 2.4 days) by bombarding uranium with slow neutrons. It was the first transuranium
element produced synthetically.

Plutonium and above


Transuranium elements do not occur in sizeable quantities in nature
and are commonly synthesized via nuclear reactions conducted with
nuclear reactors. For example, under irradiation with reactor neutrons,
uranium-238 partially converts to plutonium-239:

Glenn T. Seaborg and his group at the University


of California at Berkeley synthesized Pu, Am,
Cm, Bk, Cf, Es, Fm, Md, No and element 106,
which was later named seaborgium in his honor
while he was still living. They also synthesized
more than 100 atomic actinide isotopes.

In this way, Enrico Fermi with collaborators, using the first nuclear reactor Chicago Pile-1, obtained significant
amounts of plutonium-239, which were then used in nuclear weapons.
Actinides with the highest mass numbers are synthesized by bombarding uranium, plutonium, curium and
californium with ions of nitrogen, oxygen, carbon, neon or boron in a particle accelerator. So, nobelium was
produced by bombarding uranium-238 with neon-22 as
.
First isotopes of transplutonium elements, americium-241 and curium-242, were synthesized in 1944 by Glenn T.
Seaborg, Ralph A. James and Albert Ghiorso. Curium-242 was obtained by bombarding plutonium-239 with 32-
MeV α-particles
.
The americium-241 and curium-242 isotopes also were produced by irradiating plutonium in a nuclear reactor.
The latter element was named after Marie Curie and her husband Pierre who are noted for discovering radium
[15]
and for their work in radioactivity.
245
Bombarding curium-242 with α-particles resulted in an isotope of californium Cf (1950), and a similar procedure
yielded in 1949 berkelium-243 from americium-241. The new elements were named after Berkeley, California, by
analogy with its lanthanide homologue terbium, which was named after the village of Ytterby in Sweden.
In 1945, B. B. Cunningham obtained the first bulk chemical compound of a transplutonium element, namely
[16]
americium hydroxide. Over the next three to four years, milligram quantities of americium and microgram
amounts of curium were accumulated that allowed production of isotopes of berkelium (Thomson, 1949) and
californium (Thomson, 1950). Sizeable amounts of these elements were produced only in 1958 (Burris B.
[17]
Cunningham and Stanley G. Thomson), and the first californium compound (0.3 µg of CfOCl) was obtained only
[18]
in 1960 by B. B. Cunningham and J. C. Wallmann.
Einsteinium and fermium were identified in 1952–1953 in the fallout from the "Ivy Mike" nuclear test (1 November
1952), the first successful test of a hydrogen bomb. Instantaneous exposure of uranium-238 to a large neutron flux
resulting from the explosion produced heavy isotopes of uranium, including uranium-253 and uranium-255, and their
β-decay yielded einsteinium-253 and fermium-255. The discovery of the new elements and the new data on neutron
capture were initially kept secret on the orders of the U.S. military until 1955 due to Cold War tensions.
Nevertheless, the Berkeley team were able to prepare einsteinium and fermium by civilian means, through the
neutron bombardment of plutonium-239, and published this work in 1954 with the disclaimer that it was not the first
studies that had been carried out on the elements. The "Ivy Mike" studies were declassified and published in 1955.
The first significant (submicrograms) amounts of einsteinium were produced in 1961 by Cunningham and
colleagues, but this has not been done for fermium yet.
256
The first isotope of mendelevium, Md (half-life 87 min), was synthesized by Albert Ghiorso, Glenn T. Seaborg,
253
Gregory R. Choppin, Bernard G. Harvey and Stanley G. Thompson when they bombarded an Es target with alpha
particles in the 60-inch cyclotron of Berkeley Radiation Laboratory; this was the first isotope of any element to be
synthesized one atom at a time.
There were several attempts to obtain isotopes of nobelium by Swedish (1957) and American (1958) groups, but the
256
first reliable result was the synthesis of No by the Russian group (Georgy Flyorov et al.) in 1965, as
acknowledged by the IUPAC in 1992. In their experiments, Flyorov et al. bombarded uranium-238 with neon-22.
In 1961, Ghiorso et al. obtained the first isotope of lawrencium by irradiating californium (mostly californium-
252) with boron-10 and boron-11 ions. The mass number of this isotope was not clearly established (possibly
256 243 18
258 or 259) at the time. In 1965, Lr was synthesized by Flyorov et al. from Am and O. Thus IUPAC
recognized the nuclear physics teams at Dubna and Berkeley as the co-discoverers of lawrencium.
Isotopes

Nuclear properties of isotopes of the most important transplutonium isotopes[19]


Isotope Half-life Probability of Emission energy, MeV (yield in %) [20]
Specific of
spontaneous fission in %
α γ α, β-particles, Bq/kg fission, Bq/kg

241 432.2(7) years −10 5.485 (84.8) 0.059 (35.9) 14 546.1


Am 4.3(18)×10 1.27×10
5.442 (13.1) 0.026 (2.27)
5.388 (1.66)

243 3 −9 5.275 (87.1) 0.074 (67.2) 12 273.3


Am 7.37(4)×10 years 3.7(2)×10 7.39×10
5.233 (11.2) 0.043 (5.9)
5.181 (1.36)

242 162.8(2) days −6 6.069 (25.92) 17 9


Cm 6.2(3)×10 0.044 (0.04) 1.23×10 7.6×10
6.112 (74.08) −3
0.102 (4×10 )

244 18.10(2) years −4 5.762 (23.6) 15 9


Cm 1.37(3)×10 0.043 (0.02) 2.96×10 4.1×10
5.804 (76.4) −3
0.100 (1.5×10 )

245 3 −7 5.529 (0.58) 0.175 (9.88) 12 4


Cm 8.5(1)×10 years 6.1(9)×10 6.35×10 3.9×10
5.488 (0.83) 0.133 (2.83)
5.361 (93.2)

246 3 0.02615(7) 5.343 (17.8) 0.045 (19) 13 9


Cm 4.76(4)×10 years 1.13×10 2.95×10
5.386 (82.2)

247 7 — 5.267 (13.8) 0.402 (72) 9 —


Cm 1.56(5)×10 years 3.43×10
5.212 (5.7) 0.278 (3.4)
5.147 (1.2)

248 5 8.39(16) 5.034 (16.52) — 11 10


Cm 3.48(6)×10 years 1.40×10 1.29×10
5.078 (75)

249 330(4) days −8 −3 −5 16 7


Bk 4.7(2)×10 5.406 (1×10 ) 0.32 (5.8×10 ) 5.88×10 2.76×10
−4
5.378 (2.6×10 )

249 351(2) years −7 6.193 (2.46) 0.388 (66) 14 5


Cf 5.0(4)×10 1.51×10 7.57×10
6.139 (1.33) 0.333 (14.6)
5.946 (3.33)

250 13.08(9) years 0.077(3) 5.988 (14.99) 0.043 15 12


Cf 4.04×10 3.11×10
6.030 (84.6)

251 900(40) years ? 6.078 (2.6) 0.177 (17.3) 13 —


Cf 5.86×10
5.567 (0.9) 0.227 (6.8)
5.569 (0.9)

252 2.645(8) years 3.092(8) 6.075 (15.2) −2 16 14


Cf 0.042 (1.4×10 ) 1.92×10 6.14×10
6.118 (81.6) −2
0.100 (1.3×10 )

254 60.5(2) days ≈100 — 14 17


Cf 5.834 (0.26) 9.75×10 3.13×10
−2
5.792 (5.3×10 )

253 20.47(3) days −6 6.540 (0.85) 17 10


Es 8.7(3)×10 0.387 (0.05) 9.33×10 8.12×10
6.552 (0.71) −3
0.429 (8×10 )
6.590 (6.6)

254 275.7(5) days −6 6.347 (0.75) 0.042 (100) 16 —


Es < 3×10 6.9×10
6.358 (2.6) 0.034 (30)
6.415 (1.8)
255 39.8(12) days 0.0041(2) 6.267 (0.78) — 17 13
Es 4.38×10 (β) 1.95×10
6.401 (7) 16
3.81×10 (α)

255 20.07(7) hours −5 7.022 (93.4) 0.00057 (19.1) 19 12


Fm 2.4(10)×10 2.27×10 5.44×10
6.963 (5.04) 0.081 (1)
6.892 (0.62)

256 157.6(13) min 91.9(3) 6.872 (1.2) — 20 19


Fm 1.58×10 1.4×10
6.917 (6.9)

257 100.5(2) days 0.210(4) 6.752 (0.58) 0.241 (11) 17 14


Fm 1.87×10 3.93×10
6.695 (3.39) 0.179 (8.7)
6.622 (0.6)

256 77(2) min — 7.142 (1.84) — 20 —


Md 3.53×10
7.206 (5.9)

257 5.52(5) hours — 7.074 (14) 0.371 (11.7) 19 —


Md 8.17×10
0.325 (2.5)

258 51.5(3) days — 6.73 — 17 —


Md 3.64×10

255 3.1(2) min — 8.312 (1.16) 0.187 (3.4) 21 —


No 8.78×10
8.266 (2.6)
8.121 (27.8)

259 58(5) min — 7.455 (9.8) — 20 —


No 4.63×10
7.500 (29.3)
7.533 (17.3)

256 27(3) s < 0.03 8.319 (5.4) — 22 —


Lr 5.96×10
8.390 (16)
8.430 (33)

257 646(25) ms — 8.796 (18) — 24 —


Lr 1.54×10
8.861 (82)
Thirty-one isotopes of actinium and eight
excited isomeric states of some of its
nuclides were identified by 2010. Three
225 227 228
isotopes, Ac, Ac and Ac, were
found in nature and the others were
produced in the laboratory; only the three
natural isotopes are used in applications.
Actinium-225 is a member of radioactive
[21]
neptunium series; it was first
discovered in 1947 as a fission product
of uranium-233, it is an α-emitter with a
half-life of 10 days. Actinium-225 is less
available than actinium-228, but is more
promising in radiotracer applications.
Actinium-227 (half-life 21.77 years)
occurs in all uranium ores, but in small
quantities. One gram of uranium (in
radioactive equilibrium) contains only
−10 227
2×10 gram of Ac. Actinium-228 is a
member of radioactive thorium series
228 –
formed by the decay of Ra; it is a β
emitter with a half-life of 6.15 hours. In
−8
one tonne of thorium there is 5×10
228 Actinides have 89 to 103 protons and usually have 117 to 159 neutrons.
gram of Ac. It was discovered by Otto
Hahn in 1906.

Twenty nine isotopes of protactinium are known with mass numbers 212–240 as well as three excited isomeric
231 234
states. Only Pa and Pa have been found in nature. All the isotopes have short lifetime, except for
231 233
protactinium-231 (half-life 32,760 years). The most important isotopes are Pa and Pa, which is an intermediate
233
product in obtaining uranium-233 and is the most affordable among artificial isotopes of protactinium. Pa has
convenient half-life and energy of γ-radiation, and thus was used in most studies of protactinium chemistry.
Protactinium-233 is a β-emitter with a half-life of 26.97 days.

Uranium has the highest number (25) of both natural and synthetic isotopes. They have mass numbers of 217–242,
234 235 238
and three of them, U, U and U, are present in appreciable quantities in nature. Among others, the most
233 232 233
important is U, which is a final product of transformations of Th irradiated by slow neutrons. U has a very
235
higher fission efficiency by low-energy (thermal) neutrons, compared e.g. with U. Most uranium chemistry
9
studies were carried out on uranium-238 owing to its long half-life of 4.4×10 years.
There are 19 isotopes of neptunium with mass numbers from 225 to 244; they are all highly radioactive. The most
237 6 239 238
popular among scientists are long-lived Np (t = 2.20×10 years) and short-lived Np, Np ~ 2 days).
½
(t
½
241
Sixteen isotopes of americium are known with mass numbers from 232 to 248. The most important are Am and
243
Am, which are alpha-emitters and also emit soft, but intense γ-rays; both of them can be obtained in an
241 243
isotopically pure form. Chemical properties of americium were first studied with Am, but later shifted to Am,
243
which is almost 20 times less radioactive. The disadvantage of Am is production of the short-lived daughter
239 [22]
isotope Np, which has to be considered in the data analysis.
242 244
Among 19 isotopes of curium, the most accessible are Cm and Cm; they are α-emitters, but with much shorter
lifetime than the americium isotopes. These isotopes emit almost no γ-radiation, but undergo spontaneous fission
Actinide 9

245–248
with the associated emission of neutrons. More long-lived isotopes of curium ( Cm, all α-emitters) are formed
as a mixture during neutron irradiation of plutonium or americium. Upon short irradiation, this mixture is dominated
248
by curium-246, and then curium-248 begins to accumulate. Both of these isotopes, especially Cm, have a longer
5 242 244
half-life (3.48×10 years) and are much more convenient for carrying out chemical research than Cm and Cm,
247
but they also have a rather high rate of spontaneous fission. Cm has the longest lifetime among isotopes of curium
7
(1.56×10 years), but is not formed in large quantities because of the strong fission induced by thermal neutrons.
249
Fourteen isotopes of berkelium were identified with mass numbers 238–252. Only Bk is available in large
quantities; it has a relatively short half-life of 330 days and emits mostly soft β-particles, which are inconvenient for
−3
detection. Its alpha radiation is rather weak (1.45×10 % with respect to β-radiation), but is sometimes used to detect
247
this isotope. Bk is an alpha-emitter with a long half-life of 1,380 years, but it is hard to obtain in appreciable
quantities; it is not formed upon neutron irradiation of plutonium because of the β-stability of isotopes of curium
isotopes with mass number below 248.
Isotopes of californium with mass numbers 237–256 are formed in nuclear reactors; californium-253 is a β-emitter
250 252 254
and the rest are α-emitters. The isotopes with even mass numbers ( Cf, Cf and Cf) have a high rate of
254
spontaneous fission, especially Cf of which 99.7% decays by spontaneous fission. Californium-249 has a
relatively long half-life (352 years), weak spontaneous fission and strong γ-emission that facilitates its identification.
249 249
Cf is not formed in large quantities in a nuclear reactor because of the slow β-decay of the parent isotope Bk
and a large cross section of interaction with neutrons, but it can be accumulated in the isotopically pure form as the
249
β-decay product of (pre-selected) Bk. Californium produced by reactor-irradiation of plutonium mostly consists
250 252
of Cf and Cf, the latter being predominant for large neutron fluences, and its study is hindered by the strong
[23]
neutron radiation.

Properties of some transplutonium isotope pairs[24]


Parent isotope t Daughter t Time to establish
½ ½
isotope radioactive equilibrium

243 7370 years 239 2.35 days 47.3 days


Am Np

245 8265 years 241 14 years 129 years


Cm Pu

247 7 243 4.95 hours 7.2 days


Cm 1.64×10 years Pu

254 270 days 250 3.2 hours 35.2 hours


Es Bk

255 39.8 days 255 22 hours 5 days


Es Fm

257 79 days 253 17.6 days 49 days


Fm Cf

253
Among the 16 known isotopes of einsteinium with mass numbers from 241 to 257 the most affordable is Es. It is
an α-emitter with a half-life of 20.47 days, a relatively weak γ-emission and small spontaneous fission rate as
254
compared with the isotopes of californium. Prolonged neutron irradiation also produces a long-lived isotope Es
(t = 275.5 days).
½
254 255 256
Nineteen isotopes of fermium are known with mass numbers of 242–260. Fm, Fm and Fm are α-emitters
257
with a short half-life (hours), which can be isolated in significant amounts. Fm (t = 100 days) can accumulate
½
upon prolonged and strong irradiation. All these isotopes are characterized by high rates of spontaneous fission.
256
Among the 15 known isotopes of mendelevium (mass numbers from 245 to 260), the most studied is Md, which
mainly decays through the electron capture (α-radiation is ≈10%) with the half-life of 77 minutes. Another alpha
258 253 255
emitter, Md, has a half-life of 53 days. Both these isotopes are produced from rare einsteinium ( Es and Es
respectively), that limits their so their availability.
Actinide 1

Long-lived isotopes of nobelium and isotopes of lawrencium (and of heavier elements) have relatively small
half-lives. For nobelium 11 isotopes are known with mass numbers 250–260 and 262. Chemical properties of
255 256
nobelium and lawrencium were studied with No (t = 3 min) and Lr (t = 35 s). The longest-lived nobelium
½ ½
259
isotope No has a half-life of 1.5 hours.

Distribution in nature
Thorium and uranium are the most abundant actinides in nature with the
−3 −4
respective mass concentrations of 1.6×10 % and 4×10 %. Uranium mostly
occurs in the Earth's crust as a mixture of its oxides in the minerals uraninite,
which is also called pitchblende because of its black color. There are several
dozens of other uranium minerals such as carnotite (KUO VO ·3H O) and
2 4 2
autunite (Ca(UO ) (PO ) ·nH O). The isotopic composition of natural
2 2 4 2 2
238 235 234
uranium is U (relative abundance 99.2742%), U (0.7204%) and U
238 9
(0.0054%); of these U has the largest half-life of 4.51×10 years. The
worldwide production of uranium in 2009 amounted to 50,572 tonnes, of
which 27.3% was mined in Kazakhstan. Other important uranium mining
Unprocessed uranium
ore countries are Canada (20.1%), Australia (15.7%), Namibia (9.1%), Russia
(7.0%), and Niger (6.4%).

Content of plutonium in uranium and thorium ores


Ore Location Uranium Mass Ratio
content, 239 12
ratio Pu/U (×10 )
% 239
Pu/ore

Uraninite Canada 13.5 −12 7.1


9.1×10

Uraninite Congo 38 −12 12


4.8×10

Uraninite Colorado, US 50 −12 7.7


3.8×10

Monazite Brazil 0.24 −14 8.3


2.1×10

Monazite North Carolina, US 1.64 −14 3.6


5.9×10

Fergusonite - 0.25 −14 <4


<1×10

Carnotite - 10 −14 <0.4


<4×10

The most abundant thorium minerals are thorianite (ThO ), thorite (ThSiO ) and monazite, ((Th,Ca,Ce)PO ). Most
2 4 4
thorium minerals contain uranium and vice versa; and they all have significant fraction of lanthanides. Rich deposits
of thorium minerals are located in the United States (440,000 tonnes), Australia and India (~300,000 tonnes each)
[25]
and Canada (~100,000 tonnes).
−15
The abundance of actinium in the Earth's crust is only about 5×10 %. Actinium is mostly present in uranium-
containing, but also in other minerals, though in much smaller quantities. The content of actinium in most natural
235
objects corresponds to the isotopic equilibrium of parent isotope U, and it is not affected by the weak Ac
−12
migration. Protactinium is more abundant (10 %) in the Earth's crust than actinium. It was discovered in the
235
uranium ore in 1913 by Fajans and Göhring. As actinium, the distribution of protactinium follows that of U.
237
The half-life of the longest-lived isotope of neptunium, Np, is negligible compared to the age of the Earth.
Thus neptunium is present in nature in negligible amounts produced as intermediate decay products of other
isotopes.
239
Traces of plutonium in uranium minerals were first found in 1942, and the more systematic results on Pu are
summarized in the table (no other plutonium isotopes could be detected in those samples). The upper limit of
244 −20
abundance of the longest-living isotope of plutonium, Pu, is 3×10 %. Plutonium could not be detected in
samples of lunar soil. Owing to its scarcity in nature, most plutonium is produced synthetically. Negligible amounts
of americium, curium, berkelium and californium are produced by neutron capture reactions and beta decay in very
highly concentrated uranium-bearing deposits.

Extraction
Owing to the low abundance of actinides, their extraction is a complex,
multistep process. Fluorides of actinides are usually used because they
are insoluble in water and can be easily separated with redox reactions.
[26]
Fluorides are reduced with calcium, magnesium or barium:

Among the actinides, thorium and uranium are the easiest to isolate.
Monazite—a major thorium mineral
Thorium is extracted mostly from monazite: thorium diphosphate
(Th(PO ) ) is reacted with nitric acid, and the produced thorium nitrate
42
treated with tributyl phosphate. Rare-earth impurities are separated by increasing the pH in sulfate solution.
In another extraction method, monazite is decomposed with a 45% aqueous solution of sodium hydroxide at
140 °C. Mixed metal hydroxides are extracted first, filtered at 80 °C, washed with water and dissolved with
concentrated hydrochloric acid. Next, the acidic solution is neutralized with hydroxides to pH = 5.8 that results in
precipitation of thorium hydroxide (Th(OH) ) contaminated with ~3% of rare-earth hydroxides; the rest of rare-
earth hydroxides
4
remains in solution. Thorium hydroxide is dissolved in an inorganic acid and then purified from the rare earth
elements. An efficient method is the dissolution of thorium hydroxide in nitric acid, because the resulting solution
can be purified by extraction with organic solvents:
Th(OH) + 4 HNO → Th(NO ) + 4
4 3 34
H O
2
Metallic thorium is separated from
the anhydrous oxide, chloride or
fluoride by reacting it with calcium in
an inert atmosphere:
ThO + 2 Ca → 2 CaO + Th
2
Sometimes thorium is extracted by
electrolysis of a fluoride in a mixture
of sodium and potassium chloride at
700–800 °C in a graphite crucible.
Highly pure thorium can be extracted
from its iodide with the crystal bar
process.

Uranium is extracted from its ores in


various ways. In one method, the ore is
burned and then reacted with nitric
acid to convert uranium into a
dissolved state. Treating the solution
with a solution of tributyl phosphate
(TBP) in kerosene transforms uranium
[27] into an organic form
Separation of uranium and plutonium from nuclearfuel
UO (NO ) (TBP) . The insoluble
2 3 2 2
impurities are filtered and the uranium is extracted by reaction with hydroxides as (NH ) U O or with hydrogen
4 2 2 7
peroxide as UO ·2H O.
4 2

When the uranium ore is rich in such minerals as dolomite, magnesite, etc., those minerals consume much acid. In
this case, the carbonate method is used for uranium extraction. Its main component is an aqueous solution of sodium
4–
carbonate, which converts uranium into a complex [UO (CO ) ] , which is stable in aqueous solutions at low
2 33
concentrations of hydroxide ions. The advantages of the sodium carbonate method are that the chemicals have low
corrosivity (compared to nitrates) and that most non-uranium metals precipitate from the solution. The disadvantage
is that tetravalent uranium compounds precipitate as well. Therefore, the uranium ore is treated with sodium
carbonate at elevated temperature and under oxygen pressure:
2 UO2 + O2 + 6 CO2−
4–
3 → 2 [UO (CO ) ]
2 33
This equation suggests that the best solvent for the uranium carbonate processing is a mixture of carbonate with
bicarbonate. At high pH, this results in precipitation of diuranate, which is treated with hydrogen in the presence of
nickel yielding an insoluble uranium tetracarbonate.
Another separation method uses polymeric resins as a polyelectrolyte. Ion exchange processes in the resins result in
separation of uranium. Uranium from resins is washed with a solution of ammonium nitrate or nitric acid that yields
uranyl nitrate, UO (NO ) ·6H O. When heated, it turns into UO , which is converted to with hydrogen:
UO
2 3 2 2 3 2
UO + H → UO + H O
3 2 2 2
Reacting uranium dioxide with fluoric acid changes it to uranium tetrafluoride, which yields uranium metal upon
reaction with magnesium metal:
4 HF + UO → UF + 2 H O
2 4 2
To extract plutonium, neutron-irradiated uranium is dissolved in nitric acid, and a reducing agent (FeSO , or H O )
4 2 2
is added to the resulting solution. This addition changes the oxidation state of plutonium from +6 to +4, while
uranium remains in the form of uranyl nitrate (UO (NO ) ). The solution is treated with a reducing agent and
2 32
4+
neutralized with ammonium carbonate to pH = 8 that results in precipitation of Pu
compounds.
4+
In another method, Pu and UO2+
2 are first extracted with tributyl phosphate, then reacted with hydrazine washing out the recovered plutonium.

The major difficulty in separation of actinium is the similarity of its properties with those of lanthanum. Thus
actinium is either synthesized in nuclear reactions from isotopes of radium or separated using ion-exchange
procedures.

Properties
Actinides have similar properties to lanthanides. The 6d and 7s electronic shells are filled in actinium and
thorium, and the 5f shell is being filled with further increase in atomic number; the 4f shell is filled in the
lanthanides. The first experimental evidence for the filling of the 5f shell in actinides was obtained by McMillan
and Abelson in 1940. As in lanthanides (see lanthanide contraction), the ionic radius of actinides monotonically
[28]
decreases with atomic number (see also Aufbau principle).

Properties of actinides (the mass of the most long-lived isotope is in square brackets)[29]
Property Ac Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr

Core charge 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103

atomic mass [227] 232.0381 231.03588 238.02891 [237] [244] [243] [247] [247] [251] [252] [257] [258] [259] [262]

Number of 3 9 5 9 4 5 5 8 2 5 — — — — —
natural
isotopes

Natural 225, 226–232, 231, 232–240 237–240 238–240, 241–245 242–249 249–250 249–253 — — — — —
isotopes 227–228 234–235 233–236 242, 244

Longest-lived 227 232 231 238 237 244 243 247 247 251 252 257 258 259 262
isotope

Half-life of 21.8 14 32,500 4.47 2.14 million 80.8 7,370 15.6 1,400 years 900 1.29 100.5 52 58 min 261 min
the years billion years billion years million years million years years days days
longest-lived years years years years
isotope

Electronic 1 2 2 2 2 1 2 3 1 2 4 1 2 6 2 7 2 7 1 2 9 2 10 2 11 2 12 2 13 2 14 2 14 2 1
6d 7s 6d 7s 5f 6d 7s or 5f 6d 7s 5f 6d 7s or 5f 7s 5f 7s 5f 6d 7s 5f 7s or 5f 7s 5f 7s 5f 7s 5f 7s 5f 7s 5f 7s 7p
configuration 1 2 2 5 2 8 1 2
5f 6d 7s 5f 7s 5f 6d 7s
in the ground
state

Oxidation 2, 3 2, 3, 4 2, 3, 4, 5 2, 3, 4, 5, 3, 4, 5, 6, 7 3, 4, 5, 2, 3, 4, 2, 3, 4, 6, 2, 3, 4 2, 3, 4 2, 3, 4 2, 3 2, 3 2, 3 3
states 6 6, 7, 8 5, 6, 7 8

Metallic 0.203 0.180 0.162 0.153 0.150 0.162 0.173 0.174 0.170 0.186 0.186 — — — —
radius, nm

— 0.114 0.104 0.103 0.101 0.100 0.099 0.099 0.097 0.096 0.085 0.084 0.084 0.084 0.083
Ionic radius,
0.126 — 0.118 0.118 0.116 0.115 0.114 0.112 0.110 0.109 0.098 0.091 0.090 0.095 0.088
nm:
4+
An
3+
An
Temperature, 1050 1842 1568 1132.2 639 639.4 1176 1340 986 900 860 1530 830 830 1630
°C: 3198 4788 ? 4027 4131 ? 4174 3228 2607 3110 2627 ? 1470 ? 996 — — — —
melting
boiling

10.07 11.78 15.37 19.06 20.45 19.84 11.7 13.51 14.78 15.1 8.84
Density,
3
g/cm

— −1.83 −1.47 −1.38 −1.30 −1.25 −0.90 −0.75 −0.55 −0.59 −0.36 −0.29 — — —
Standard
−2.13 — — −1.66 −1.79 −2.00 −2.07 −2.06 −1.96 −1.97 −1.98 −1.96 −1.74 −1.20 −2.10
electrode
potential, V:

4+ 0
(An /An )

3+ 0
(An /An )

— Colorless Yellow Green Yellow-green Brown Red Yellow Beige Green — — — — —


Color
4+ Colorless Blue Dark blue Purple Purple Violet Rose Colorless Yellow-green Green Pink — — — —
[M(H O) ]
3+
[M(H O) ]
2 n

Approximate colors of actinide ions in aqueous solution[30]


Oxidation state 89 90 91 92 93 94 95 96 97 98 99

+3 3+ 3+ 3+ 3+ 3+ 3+ 3+ 3+ 3+ 3+ 3+
Ac Th Pa U Np Pu Am Cm Bk Cf Es

+4 4+ 4+ 4+ 4+ 4+ 4+ 4+ 4+ 4+
Th Pa U Np Pu Am Cm Bk Cf

+5 PaO+ UO+ NpO+ PuO+ AmO+


2 2 2 2 2

+6 UO2+ NpO2+ PuO2+ AmO2+


2 2 2 2

+7 NpO3+ PuO3+ 5-
[AmO ]
6
2 2

Physical properties

Major crystal structures of some actinides vs. temperature Metallic and ionic radii of actinides
Actinides are typical metals. All of them are soft and have a silvery
[31]
color (but tarnish in air), relatively high density and plasticity.
Some of them can be cut with a knife. Their electrical resistivity varies
between 15 and 150 µOhm·cm. The hardness of thorium is similar to
that of soft steel, so heated pure thorium can be rolled in sheets and
pulled into wire. Thorium is nearly half as dense as uranium and
plutonium, but is harder than either of them. All actinides are
radioactive, paramagnetic, and, with the exception of actinium, have
several crystalline phases: plutonium has seven, and uranium,
238
neptunium and californium three. Crystal structures of protactinium, A pellet of PuO to be used in a
radioisotope
uranium, neptunium and plutonium do not have clear analogs among 2
thermoelectric generator for either the Cassini or
the lanthanides and are more similar to those of the 3d-transition
Galileo mission. The pellet produces 62 watts of
metals. heat and glows because of the heat generated
by the radioactive decay (primarily α). Photo is
All actinides are pyrophoric, especially when finely divided, that is, taken after insulating the pellet under a graphite
they spontaneously ignite upon reaction with air. The melting point of blanket for minutes and removing the blanket.
actinides does not have a clear dependence on the number of f-
electrons. The unusually low melting point of neptunium and plutonium
(~640 °C) is explained by hybridization of 5f and 6d orbitals and the
formation of directional bonds in these metals.

Californium

Comparison of ionic radii of lanthanides and actinides[32]


Lanthanides 3+ Actinides 3+ 4+
Ln , Å An , Å An , Å

Lanthanum 1.061 Actinium 1.11 –

Cerium 1.034 Thorium 1.08 0.99

Praseodymium 1.013 Protactinium 1.05 0.93

Neodymium 0.995 Uranium 1.03 0.93

Promethium 0.979 Neptunium 1.01 0.92

Samarium 0.964 Plutonium 1.00 0.90

Europium 0.950 Americium 0.99 0.89

Gadolinium 0.938 Curium 0.98 0.88

Terbium 0.923 Berkelium - -

Dysprosium 0.908 Californium - -

Holmium 0.894 Einsteinium - -

Erbium 0.881 Fermium - -


Thulium 0.869 Mendelevium - -

Ytterbium 0.858 Nobelium - -

Lutetium 0.848 Lawrencium - -

Chemical properties
Like the lanthanides, all actinides are highly reactive with halogens and chalcogens; however, the actinides react
more easily. Actinides, especially those with a small number of 5f-electrons, are prone to hybridization. This is
explained by the similarity of the electron energies at the 5f, 7s and 6d shells. Most actinides exhibit a larger variety
of valence states, and the most stable are +6 for uranium, +5 for protactinium and neptunium, +4 for thorium and
[33]
plutonium and +3 for actinium and other actinides.
Chemically, actinium is similar to lanthanum, which is explained by their similar ionic radii and electronic
structure. Like lanthanum, actinium has oxidation of +3, but it is less reactive and has more pronounced basic
3+
properties. Among other trivalent actinides Ac is least acidic, i.e. has the weakest tendency to hydrolyze in
aqueous solutions.
Thorium is rather active chemically. Owing to lack of electrons on 6d and 5f orbitals, the tetravalent thorium
4+
compounds are colorless. At pH < 3, the solutions of thorium salts are dominated by the cations [Th(H O) ] .
The 2 8
4+
Th ion is relatively large, and depending on the coordination number can have a radius between 0.95 and 1.14 Å.
As a result, thorium salts have a weak tendency to hydrolyse. The distinctive ability of thorium salts is their high
solubility, not only in water, but also in polar organic solvents.
Protactinium exhibits two valence states; the +5 is stable, and the +4 state easily oxidizes to protactinium(V).
Thus tetravalent protactinium in solutions is obtained by the action of strong reducing agents in a hydrogen
atmosphere. Tetravalent protactinium is chemically similar to uranium(IV) and thorium(IV). Fluorides,
phosphates, hypophosphate, iodate and phenylarsonates of protactinium(IV) are insoluble in water and dilute
acids. Protactinium forms soluble carbonates. The hydrolytic properties of pentavalent protactinium are close to
those of tantalum(V) and niobium(V). The complex chemical behavior of protactinium is a consequence of the
start of the filling of the 5f shell in this element.
Uranium has a valence from 3 to 6, the last being most stable. In the hexavalent state, uranium is very similar to the
group 6 elements. Many compounds of uranium(IV) and uranium(VI) are non-stoichiometric, i.e. have variable
composition. For example, the actual chemical formula of uranium dioxide is UO , where x varies between −0.4
2+x
and 0.32. Uranium(VI) compounds are weak oxidants. Most of them contain the linear "uranyl" group, UO2+
2. Between 4 to 6 ligands can be accommodated in an equatorial plane perpendicular to the uranyl group. The uranyl

group acts as a hard acid and forms stronger complexes with oxygen-donor ligands than with nitrogen-donor ligands.
NpO2+
2 and
PuO2+
2 are also the common form of Np and Pu in the +6 oxidation state. Uranium(IV) compounds exhibit reducing

properties, e.g., they are easily oxidized by atmospheric oxygen. Uranium(III) is a very strong reducing agent.
Owing to the presence of d-shell, uranium (as well as many other actinides) forms organometallic compounds, such
III IV [34]
as U (C H ) and U (C H ) .
5 5 3 5 5 4
Neptunium has valence states from 3 to 7, which can be simultaneously observed in solutions. The most stable state
in solution is +5, but the valence +4 is preferred in solid neptunium compounds. Neptunium metal is very reactive.
Ions of neptunium are prone to hydrolysis and formation of coordination compounds.
Plutonium also exhibits the valence between 3 and 7, and thus is chemically similar to neptunium and uranium.
It is highly reactive, and quickly forms an oxide film in air. Plutonium reacts with hydrogen even at temperatures
as low as 25–50 °C; it also easily forms halides and intermetallic compounds. Hydrolysis reactions of plutonium
ions of different oxidation states are quite diverse. Plutonium(V) can enter polymerization reactions.
The largest chemical diversity among actinides is observed in americium, which can have valence between 2
and 6. Divalent americium is obtained only in dry compounds and non-aqueous solutions (acetonitrile).
Oxidation states +3,
+5 and +6 are typical for aqueous solutions, but also in the solid state. Tetravalent americium forms stable solid
compounds (dioxide, fluoride and hydroxide) as well as complexes in aqueous solutions. It was reported that in
alkaline solution americium can be oxidized to the heptavalent state, but these data proved erroneous. The most
[35]
stable valence of americium is 3 in the aqueous solutions and 3 or 4 in solid compounds.
Valence 3 is dominant in all subsequent elements up to lawrencium (with the possible exception of nobelium).
Curium can be tetravalent in solids (fluoride, dioxide). Berkelium, along with a valence of +3, also shows the
valence of +4, more stable than that of curium; the valence 4 is observed in solid fluoride and dioxide. The stability
4+ 4+
of Bk in aqueous solution is close to that of Ce . Only valence 3 was observed for californium, einsteinium and
fermium. The divalent state is proven for mendelevium and nobelium, and in nobelium it is more stable than the
trivalent state. Lawrencium shows valence 3 both in solutions and solids.

The redox potential increases from −0.32 V in uranium, through 0.34 V (Np) and 1.04 V (Pu) to 1.34 V
4+
in americium revealing the increasing reduction ability of the An ion from americium to uranium. All actinides
form
AnH hydrides of black color with salt-like properties. Actinides also produce carbides with the general formula of
3
AnC or AnC (U C for uranium) as well as sulfides An S and AnS .
2 2 3 2 3 2

Uranyl nitrate (UO (NO ) ) Aqueous solutions Aqueous solutions of neptunium Aqueous solutions of
of
2 3 2
uranium III, IV, V, VI III, IV, V, VI, VII plutonium III, IV, V, VI, VII
salts salts salts

Uranium Uranium hexafluoride U O (yellowcake)


3 8
tetrachloride

Compounds

Oxides and hydroxides


Oxides of actinides[36]
Compound Color Crystal symmetry, type Lattice constants, Å 3 Temperature, °C
Density, g/cm
a b c

Ac O White Hexagonal, La O 4.07 - 6.29 9.19 –


2 3 2 3

PaO - Cubic, CaF 5.505 - - - -


2 2

Pa O White cubic, CaF 5.446 - - - 700


2 5 2
Cubic 10.891 - 10.992 700–1100
Tetragonal 5.429 - 5.503 1000
Hexagonal 3.817 - 13.22 1000–1200
Rhombohedral 5.425 - - 1240–1400
Orthorhombic 6.92 4.02 4. 18 –

ThO Colorless Cubic 5.59 - - 9.87 –


2

UO Black-brown Cubic 5.47 - - 10.9 –


2

NpO Greenish-brown Cubic, CaF 5.424 - - 11.1 –


2 2

PuO Black Cubic, NaCl 4.96 - - 13.9 –

PuO Olive green Cubic 5.39 - - 11.44 –


2

Am O Red-brown Cubic, Mn O 11.03 - - 10.57 –


2 3 2 3
Red-brown Hexagonal, La O 3.817 5.971 11.7
2 3

AmO Black Cubic, CaF 5.376 - - - -


2 2

Cm O [37] Cubic, Mn O 11.01 - - 11.7 –


2 3 White 2 2
Hexagonal, LaCl 3.80 - 6
- 3
Monoclinic, Sm O 14.28 3.65 8.9
- 2 3

CmO Black Cubic, CaF 5.37 - - - -


2 2

Bk O Light brown Cubic, Mn O 10.886 - - - -


2 3 2 3

BkO Red-brown Cubic, CaF 5.33 - - - -


2 2

[38] Colorless Cubic, Mn O 10.79 - - - -


2 3
Cf2O3 Monoclinic, Sm O
Yellowish 14.12 3.59 8.80
2 3
- Hexagonal, La O 3.72 - 5.96
2 3

CfO Black Cubic 5.31 - - - -


2

Es O - Cubic, Mn O 10.07 - - - -
2 3 2 3
Monoclinic 14.1 3.59 8.80
Hexagonal, La O 3.7 - 6
2 3

Approximate colors of actinide oxides (most stable are bolded)[39]


Oxidation state 89 90 91 92 93 94 95 96 97 98 99

+3 Pu O
Am O Cm O Bk O Cf Es O
O

+4 ThO 2 3 2 3 2 3 2 3 2 3 2 3

PaO UO NpO PuO AmO CmO BkO CfO


2 2 2 2 2 2 2 2 2
+5 Pa O UO
Np O
2 5 2 5 2 5

+6 U O
3 8
UO
3
Dioxides of some actinides
Chemical formula ThO PaO UO NpO PuO AmO CmO BkO CfO
2 2 2 2 2 2 2 2 2

CAS-number 1314-20-1 12036-03-2 1344-57-6 12035-79-9 12059-95-9 12005-67-3 12016-67-0 12010-84-3 12015-10-0

Molar mass 264.04 263.035 270.03 269.047 276.063 275.06 270–284** 279.069 283.078

Melting point 3390 °C 2865 °C 2547 °C 2400 °C 2175 °C

Crystal structure

4+ 2−
An : / O :
Space group Fm3m

Coordination number An[8], O[4]

An – actinide
**Depending on the isotopes

Some actinides can exists in several oxide forms such as An O , AnO , An O and AnO . For all actinides, oxides
2 3 2 2 5 3
AnO are amphoteric and An O , AnO and An O are basic, they easily react with water, forming bases:
3 2 3 2 2 5
An O + 3 H O → 2 An(OH) .
2 3 2 3
These bases are poorly soluble in water and by their activity are close to the hydroxides of rare-earth metals. The
strongest base is of actinium. All compounds of actinium are colorless, except for black actinium sulfide (Ac S ).
2 3
Dioxides of tetravalent actinides crystallize in the cubic system, same as in calcium fluoride.
Thorium reacting with oxygen exclusively forms dioxide:

Thorium dioxide is a refractory material with the highest melting point among any known oxide (3390 °C). Adding
0.8–1% ThO to tungsten stabilizes its structure, so the doped filaments have better mechanical stability to
2
vibrations. To dissolve ThO in acids, it is heated to 500–600 °C; heating above 600 °C produces a very resistant to
2
acids and other reagents form of ThO . Small addition of fluoride ions catalyses dissolution of thorium dioxide in
2
acids.
Two protactinium oxides were obtained: PaO (black) and Pa O (white); the former is isomorphic with ThO and
2 2 5 2
the latter is easier to obtain. Both oxides are basic, and is a weak, poorly soluble base.
Pa(OH)
5
Decomposition of certain salts of uranium, for example UO (NO )·6H O in air at 400 °C, yields orange or yellow
2 3 2
UO . This oxide is amphoteric and forms several hydroxides, the most stable being UO (OH) . Reaction of
3 2 2
uranium(VI) oxide with hydrogen results in uranium dioxide, which is similar in its properties with ThO . This oxide
2
is also basic and corresponds to the uranium hydroxide (U(OH) ).
4
Plutonium, neptunium and americium form two basic oxides: An O and AnO . Neptunium trioxide is unstable;
2 3 2
thus, only Np O could be obtained so far. However, the oxides of plutonium and neptunium with the chemical
3 8
formula AnO and An O are well characterized.
2 2 3
Salts

Trichlorides of some actinides[40]


Chemical formula AcCl UCl NpCl PuCl AmCl CmCl BkCl CfCl
3 3 3 3 3 3 3 3

CAS-number 22986-54-5 10025-93-1 20737-06-8 13569-62-5 13464-46-5 13537-20-7 13536-46-4 13536-90-8

Molar mass 333.386 344.387 343.406 350.32 349.42 344–358** 353.428 357.438

Melting point 837 °C 800 °C 767 °C 715 °C 695 °C 603 °C 545 °C

Boiling point 1657 °C 1767 °C 850 °C

Crystal structure

3+ −
An : / Cl :
Space group P6 /m
3

Coordination number An*[9], Cl [3]

Lattice constants a = 762 pm a = 745.2 a = 739.4 a = 738.2 a = 726 pm a = 738.2 a = 738 pm


c = 455 pm pm pm pm c = 414 pm pm c = 409 pm
c = 432.8 pm c = 424.3 pm c = 421.4 pm c = 412.7 pm

*An – actinide
**Depending on the isotopes

Actinide fluorides[41]
Compound Color Crystal symmetry, type Lattice constants, Å 3
Density, g/cm
a b c

AcF White Hexagonal, LaF 4.27 - 7.53 7.88


3 3

PaF Dark brown Monoclinic 12.7 10.7 8.42 –


4

PaF Black Tetragonal, β-UF 11.53 - 5.19 –


5 5

ThF Colorless Monoclinic 13 10.99 8.58 5.71


4

UF Reddish-purple Hexagonal 7.18 - 7.34 8.54


3

UF Green Monoclinic 11.27 10.75 8.40 6.72


4

α-UF Bluish Tetragonal 6.52 - 4.47 5.81


5

β-UF Bluish Tetragonal 11.47 - 5.20 6.45


5

UF Yellowish Orthorhombic 9.92 8.95 5.19 5.06


6

NpF Black or purple Hexagonal 7.129 - 7.288 9.12


3

NpF Light green Monoclinic 12.67 10.62 8.41 6.8


4

NpF Orange Orthorhombic 9.91 8.97 5.21 5


6

PuF Violet-blue Trigonal 7.09 - 7.25 9.32


3
PuF Pale brown Monoclinic 12.59 10.57 8.28 6.96
4

PuF Red-brown Orthorhombic 9.95 9.02 3.26 4.86


6

AmF Pink or light beige hexagonal, LaF 7.04 - 7.255 9.53


3 3

AmF Orange-red Monoclinic 12.53 10.51 8.20 –


4

CmF From brown to white Hexagonal 4.041 - 7.179 9.7


3

CmF Yellow Monoclinic, UF 12.51 10.51 8.20 –


4 4

BkF Yellow-green Trigonal, LaF 6.97 - 7.14 10.15


3 3
Orthorhombic, YF 6.7 7.09 4.41 9.7
3

BkF - Monoclinic, UF 12.47 10.58 8.17 –


4 4

CfF - Trigonal, LaF 6. 94 - 7.10 –


3 3
- Orthorhombic, YF 6.65 7.04 4.39
3

CfF - Monoclinic, UF 1.242 1.047 8.126 –


4 4
- Monoclinic, UF 1.233 1.040 8.113
4

Actinides easily react with halogens forming salts with the formulas MX and MX
3 4
(X = halogen). So the first berkelium compound, BkCl , was synthesized in 1962
3
with an amount of 3 nanograms. Like the halogens of rare earth elements, actinide
chlorides, bromides, and iodides are water soluble, and fluorides are insoluble.
Uranium easily yields a colorless hexafluoride, which sublimates at a temperature of
56.5 °C; because of its volatility, it is used in the separation of uranium isotopes
with gas centrifuge or gaseous diffusion. Actinide hexafluorides have properties
close to anhydrides. They are very sensitive to moisture and hydrolyze forming
[42]
AnO 2F 2. The pentachloride and black hexachloride of uranium were synthesized,
Einsteinium triiodide glowing in
but they are both unstable.
the dark

Action of acids on actinides yields salts, and if the acids are non-oxidizing then the
actinide in the salt is in low-valence state:
U+2H → U (SO +2H
SO )
2 4 4 2 2
2 Pu + 6 HCl → 2 PuCl + 3 H
3 2
However, in these reactions the regenerating hydrogen can react with the metal, forming the corresponding hydride.
Uranium reacts with acids and water much more easily than thorium.
Actinide salts can also be obtained by dissolving the corresponding hydroxides in acids. Nitrates, chlorides, sulfates
and perchlorates of actinides are water soluble. When crystallizing from aqueous solutions, these salts forming a
hydrates, such as Th(NO ) ·6H O, Th(SO ) ·9H O and Pu (SO ) ·7H O. Salts of high-valence actinides easily
3 4 2 4 2 2 4 3 2
2
hydrolyze. So, colorless sulfate, chloride, perchlorate and nitrate of thorium transform into basic salts with formulas
Th(OH) SO and Th(OH) NO . The solubility and insolubility of trivalent and tetravalent actinides is like that of
2 4 3 3
lanthanide salts. So phosphates, fluorides, oxalates, iodates and carbonates of actinides are weakly soluble in water;
they precipitate as hydrates, such as ThF ·3H O and Th(CrO ) ·3H O.
4 2 4 2 2
2+ 2– 2–
Actinides with oxidation state +6, except for the AnO -type cations, form [AnO ] , [An O ] and other complex
2 4 2 7
anions. For example, uranium, neptunium and plutonium form salts of the Na UO (uranate) and (NH ) U O
2 4 42 2 7
(diuranate) types. In comparison with lanthanides, actinides more easily form coordination compounds, and this
ability increases with the actinide valence. Trivalent actinides do not form fluoride coordination compounds,
whereas tetravalent thorium forms K ThF , KThF , and even K ThF complexes. Thorium also forms the
2 6 5 5 9
corresponding sulfates (for example Na SO ·Th (SO ) ·5H O), nitrates and thiocyanates. Salts with the general
2 4 4 2 2
formula An Th(NO ) ·nH O are of coordination nature, with the coordination number of thorium equal to 12. Even
2 3 6 2
easier is to produce complex salts of pentavalent and hexavalent actinides. The most stable coordination compounds
of actinides – tetravalent thorium and uranium – are obtained in reactions with diketones, e.g. acetylacetone.

Applications
While actinides have some established daily-life applications, such as
[43][44] []
in smoke detectors (americium) and gas mantles (thorium),
they are mostly used in nuclear weapons and use as a fuel in nuclear
reactors. The last two areas exploit the property of actinides to release
enormous energy in nuclear reactions, which under certain conditions
may become self-sustaining chain reaction.

Interior of a smoke detector containing


americium-241.

The most important isotope for nuclear power applications is uranium-235. It is


used in the thermal reactor, and its concentration in natural uranium does not
exceed 0.72%. This isotope strongly absorbs thermal neutrons releasing much
235
energy. One fission act of 1 gram of U converts into about 1 MW·day. Of
235 [45]
importance, is that U emits more neutrons than it absorbs; upon reaching
235
the critical mass, U enters into a self-sustaining chain reaction. Typically,
uranium nucleus is divided into two fragments with the release of 2–3 neutrons,
for example:

Other promising actinide isotopes for nuclear power are thorium-232 and its
Self-illumination of a nuclear product from the thorium fuel cycle, uranium-233.
reactor by Cherenkov
radiation.

[46]
Nuclear reactor
The core of any nuclear reactor contains a set of hollow metal rods, usually made of zirconium alloys, filled with nuclear fuel cells – mostly oxide,
carbide, nitride or monosulfide of uranium, plutonium or thorium, or their mixture (the so-called MOX fuel). The most common fuel is oxide of
uranium-235.

Fast neutrons are slowed by moderators, which contain water, carbon, deuterium, or beryllium, as thermal neutrons to increase the efficiency of
their interaction with uranium-235. The rate of nuclear reaction is controlled by introducing additional rods made of boron or cadmium or a liquid
absorbent, usually boric acid. Reactors for plutonium production are called breeder reactor or breeders; they have a different design and use fast
neutrons.

Emission of neutrons during the fission of uranium is important not only for maintaining the nuclear chain
reaction, but also for the synthesis of the heavier actinides. Uranium-239 converts via β-decay into plutonium-
239, which, like uranium-235, is capable of spontaneous fission. The world's first nuclear reactors were built not
for energy, but for producing plutonium-239 for nuclear weapons.
About half of the produced thorium is used as the light-emitting material of gas mantles. Thorium is also added into
multicomponent alloys of magnesium and zinc. So the Mg-Th alloys are light and strong, but also have high melting
point and ductility and thus are widely used in the aviation industry and in the production of missiles. Thorium also
has good electron emission properties, with long lifetime and low potential barrier for the emission. The relative
content of thorium and uranium isotopes is widely used to estimate the age of various objects, including stars (see
radiometric dating).
The major application of plutonium has been in nuclear weapons, where the isotope plutonium-239 was a key
component due to its ease of fission and availability. Plutonium-based designs allow reducing the critical mass to
about a third of that for uranium-235. The "Fat Man"-type plutonium bombs produced during the Manhattan Project
used explosive compression of plutonium to obtain significantly higher densities than normal, combined with a
central neutron source to begin the reaction and increase efficiency. Thus only 6.2 kg of plutonium was needed for
an explosive yield equivalent to 20 kilotons of TNT. (See also Nuclear weapon design.) Hypothetically, as little as
4 kg of plutonium—and maybe even less—could be used to make a single atomic bomb using very sophisticated
assembly designs.
Plutonium-238 is potentially more efficient isotope for nuclear reactors, since it has smaller critical mass than
[47]
uranium-235, but it continues to release much thermal energy (0.56 W/g) by decay even when the fission chain
reaction is stopped by control rods. Its application is limited by the high price (about 1000 USD/g). This isotope has
been used in thermopiles and water distillation systems of some space satellites and stations. So Galileo and Apollo
[48]
spacecraft (e.g. Apollo 14 ) had heaters powered by kilogram quantities of plutonium-238 oxide; this heat is also
transformed into electricity with thermopiles. The decay of plutonium-238 produces relatively harmless alpha
particles and is not accompanied by gamma-irradiation. Therefore this isotope (~160 mg) is used as the energy
source in heart pacemakers where it lasts about 5 times longer than conventional batteries.
Actinium-227 is used as a neutron source. Its high specific energy (14.5 W/g) and the possibility of obtaining
significant quantities of thermally stable compounds are attractive for use in long-lasting thermoelectric generators
228
for remote use. Ac is used as an indicator of radioactivity in chemical research, as it emits high-energy electrons
228 228
(2.18 MeV) that can be easily detected. Ac- Ra mixtures are widely used as an intense gamma-source in
industry and medicine.
Development of self-glowing actinide-doped materials with durable crystalline matrices is a new area of actinide
utilization as the addition of alpha-emitting radionuclides to some glasses and crystals may confer luminescence.

Toxicity
Radioactive substances can harm
human health via (i) local skin
contamination, (ii) internal exposure
due to ingestion of radioactive
isotopes, and (iii) external
overexposure by β-activity and γ-
radiation. Together with radium and
Schematic illustration of penetration of radiation
transuranium elements, actinium is one through sheets of paper, aluminium and lead
of the most dangerous radioactive brick
poisons with high specific α-activity.
The most important feature of actinium
is its ability to accumulate and remain
in the surface layer of skeletons. At the
initial stage of poisoning, actinium
accumulates in the liver. Another
danger of actinium is that it undergoes
radioactive decay faster than being
excreted. Adsorption from the
digestive tract is much smaller
(~0.05%) for actinium than radium.

Protactinium in the body tends to


accumulate in the kidneys and bones.
The maximum safe dose of
protactinium in the human body is 0.03
µCi that corresponds to 0.5 Periodic table with elements colored according to the half-life of their most stable isotope.
231
micrograms of Pa. This isotope, Elements which contain at least one stable isotope. Slightly radioactive elements: the
which might be present in the air as most stable isotope is very long-lived, with a half-life of over four million
8 years. Moderately radioactive elements: the most stable isotope has half-life between 800
aerosol, is 2.5×10 times more toxic and 34.000 years. Really radioactive elements: the most stable isotope has half-life
than hydrocyanic acid. between one day and 103 years. Highly radioactive elements: the most stable isotope
has half-life between several minutes and one day. Extremely radioactive elements: the
Plutonium, when entering the body most stable isotope has half-life less than several minutes. Very little is known about
through air, food or blood (e.g. a these elements due to their extreme instability and radioactivity.
wound), mostly settles in the lungs,
liver and bones with only about 10% going to other organs, and remains there for decades. The long residence
time of plutonium in the body is partly explained by its poor solubility in water. Some isotopes of plutonium emit
ionizing
α-radiation, which damages the surrounding cells. The median lethal dose (LD ) for 30 days in dogs after
50
intravenous injection of plutonium is 0.32 milligram per kg of body mass, and thus the lethal dose for humans is

approximately 22 mg for a person weighing 70 kg; the amount for respiratory exposure should be approximately four
times greater. Another estimate assumes that plutonium is 50 times less toxic than radium, and thus permissible
Actinide 1

content of plutonium in the body should be 5 µg or 0.3 µCi. Such amount is nearly invisible in under microscope.
After trials on animals, this maximum permissible dose was reduced to 0.65 µg or 0.04 µCi. Studies on animals also
revealed that the most dangerous plutonium exposure route is through inhalation, after which 5–25% of inhaled
substances is retained in the body. Depending on the particle size and solubility of the plutonium compounds,
plutonium is localized either in the lungs or in the lymphatic system, or is absorbed in the blood and then transported
to the liver and bones. Contamination via food is the least likely way. In this case, only about 0.05% of soluble
0.01% insoluble compounds of plutonium absorbs into blood, and the rest is excreted. Exposure of damaged skin to
plutonium would retain nearly 100% of it.
Using actinides in nuclear fuel, sealed radioactive sources or advanced materials such as self-glowing crystals has
many potential benefits. However, a serious concern is the extremely high radiotoxicity of actinides and their
migration in the environment. Use of chemically unstable forms of actinides in MOX and sealed radioactive sources
is not appropriate by modern safety standards. There is a challenge to develop stable and durable actinide-bearing
materials, which provide safe storage, use and final disposal. A key need is application of actinide solid solutions in
durable crystalline host phases.

References and notes


[1] The Manhattan Project. An Interactive History (http://web.archive.org/web/20101122185847/http:// www.cfo.doe. gov/me70/
manhattan/nagasaki.htm). US Department of Energy
[2] Actinide element (http://www.britannica.com/EBchecked/topic/4354/actinoid-element), Encyclopædia Britannica on-line
[3] Although "actinoid" (rather than "actinide") means "actinium-like" and therefore should exclude actinium, that element is usually included in
the series.
[4] Greenwood, p. 1250
[5] Greenwood, p. 1252
[6] Nobelium and lawrencium were almost simultaneously discovered by Soviet and American scientists
[7] Myasoedov, p. 7
[8] Myasoedov, p. 9
[9] Myasoedov, p. 14
[10] Zimmerman, Ann., 213, 290 (1882); 216, 1 (1883); Ber. 15 (1882) 849
[11] Golub, p. 214
[12] (modern citation: Annalen der Physik, vol. 92, no. 7, pages 385–415)
[13] Golub, p. 213
[14] Greenwood, p. 1251
[15] Myasoedov, p. 8
[16] Wallace W. Schulz (1976) The Chemistry of Americium (http://www.osti.gov/bridge/purl. cover.
jsp;jsessionid=99C379B4BBA56BB186AAD989333D2B5E?purl=/7232133-fyKvqE/), U.S. Department of Commerce, p. 1
[17] S. G. Thompson, B. B. Cunningham: "First Macroscopic Observations of the Chemical Properties of Berkelium and Californium",
supplement to Paper P/825 presented at the Second Intl. Conf., Peaceful Uses Atomic Energy, Geneva, 1958
[18] Darleane C. Hoffman, Albert Ghiorso, Glenn Theodore Seaborg The transuranium people: the inside story, Imperial College Press, 2000
ISBN 1-86094-087-0, pp. 141–142
[19] Myasoedov, pp. 19–21
[20] Specific activity is calculated by given in the table half-lives and the probability of spontaneous fission
[21] Greenwood, p. 1254
[22] Myasoedov, p. 18
[23] Myasoedov, p. 22
[24] Myasoedov, p. 25
[25] Thorium (http:// minerals.usgs.gov/minerals/pubs/ commodity/thorium/mcs-2010-thori.pdf), USGS Mineral Commodities
[26] Golub, pp. 215–217
[27] Greenwood, pp. 1255, 1261
[28] Golub, pp. 218–219
[29] Greenwood, p. 1263
[30] Greenwood, p. 1265
[31] Greenwood, p. 1264
[32] Myasoedov, pp. 30–31
[33] Golub, pp. 222–227
[34] Greenwood, p. 1278
[35] Myasoedov, pp. 25–29
[36] Myasoedov, p. 88
[37] According to other sources, cubic sesquioxide of curium is olive-green. See
[38] The atmosphere during the synthesis affects the lattice parameters, which might be due to non-stoichiometry as a result of oxidation or
reduction of the trivalent californium. Main form is the cubic oxide of californium(III).
[39] Greenwood, p. 1268
[40] Greenwood, p. 1270
[41] Myasoedov, pp. 96–99
[42] Greenwood, p.1269
[43] Smoke Detectors and Americium (http://web.archive.org/web/19960101-re_/http://www.uic.com.au/ nip35.htm), Nuclear Issues
Briefing Paper 35, May 2002
[44] Greenwood, p. 1262
[45] Golub, pp. 220–221
[46] Greenwood, pp. 1256–1261
[47] John Holdren and Matthew Bunn Nuclear Weapons Design & Materials (http://web.archive.org/web/20101105035505/http://www.nti.
org/e_research/cnwm/overview/technical2.asp). Project on Managing the Atom (MTA) for NTI. 25 November 2002
[48] Apollo 14 Press Kit – 01/11/71 (http://www.hq.nasa. gov/ alsj/a14/ A14_PressKit. pdf), NASA, pp. 38–39

Bibliography
• Golub, A. M. (1971). Общая и неорганическая химия (General and Inorganic Chemistry) 2.
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
ISBN 0080379419.
• Myasoedov, B. (1972). Analytical chemistry of transplutonium elements. Moscow: Nauka. ISBN 0-470-62715-8.

External links
• Lawrence Berkeley Laboratory image of historic periodic table by Seaborg showing actinide series for the
first time (http://imglib.lbl.gov/ImgLib/COLLECTIONS/BERKELEY-LAB/SEABORG-ARCHIVE/index/
96B05654.html)Wikipedia:Link rot
• Lawrence Livermore National Laboratory, Uncovering the Secrets of the Actinides (http://www.llnl.gov/str/
pdfs/06_00. 2.pdf#search="actinide series")
• Los Alamos National Laboratory, Actinide Research Quarterly (http://arq.lanl.gov/)
1

Elements

89 – Actinium
Actinium
Ac
89

radium ← actinium → thoriumLa



Ac

(Uqu)

Actinium in the periodic table


Appearance

silvery-white, glowing with an eerie blue light

General properties

Name, symbol, number actinium, Ac, 89

Pronunciation /ækˈtɪniəm/
ak-TIN-nee-əm

Element category actinide


sometimes considered a transition metal

Group, period, block n/a, 7, f

Standard atomic weight (227)

Electron configuration 1 2
[Rn] 6d 7s
2, 8, 18, 32, 18, 9, 2

Physical properties

Phase solid

Density (near r.t.) −3


10 g·cm

Melting point (circa) 1323 K, 1050 °C, 1922 °F

Boiling point 3471 K, 3198 °C, 5788 °F

Heat of fusion −1
14 kJ·mol
89 Actinium 1

Heat of vaporization −1
400 kJ·mol

Molar heat capacity −1 −1


27.2 J·mol ·K

Atomic properties

Oxidation states 3
(neutral oxide)

Electronegativity 1.1 (Pauling scale)

Ionization energies −1
1st: 499 kJ·mol

−1
2nd: 1170 kJ·mol

Covalent radius 215 pm

Miscellanea

Crystal structure face-centered cubic

Thermal conductivity −1 −1
12 W·m ·K

CAS registry number 7440-34-8

History

Discovery Friedrich Oskar Giesel (1902)

First isolation Friedrich Oskar Giesel (1902)

Most stable isotopes

Main article: Isotopes of actinium

iso NA half-life DM DE (MeV) DP

225 trace 10 d α 5.935 221


Ac Fr

226 syn 29.37 h − 1.117 226


Ac β Th
ε 0.640 226
Ra
α 5.536 222
Fr

227 trace 21.772 y − 0.045 227


Ac β Th
α 5.042 223
Fr

• v
• t
[1]
• e

Actinium is a radioactive chemical element with symbol Ac (not to be confused with the abbreviation for an
acetyl group) and atomic number 89, which was discovered in 1899. It was the first non-primordial radioactive
element to be isolated. Polonium, radium and radon were observed before actinium, but they were not isolated
until 1902. Actinium gave the name to the actinide series, a group of 15 similar elements between actinium and
lawrencium in the periodic table.
A soft, silvery-white radioactive metal, actinium reacts rapidly with oxygen and moisture in air forming a white
coating of actinium oxide that prevents further oxidation. As with most lanthanides and actinides, actinium
assumes
oxidation state +3 in nearly all its chemical compounds. Actinium is found only in traces in uranium ores as the
227
isotope Ac, which decays with a half-life of 21.772 years, predominantly emitting beta particles. One tonne of
uranium ore contains about 0.2 milligrams of actinium. The close similarity of physical and chemical properties of
actinium and lanthanum makes separation of actinium from the ore impractical. Instead, the element is prepared, in
226
milligram amounts, by the neutron irradiation of Ra in a nuclear reactor. Owing to its scarcity, high price and
radioactivity, actinium has no significant industrial use. Its current applications include a neutron source and an
agent for radiation therapy targeting cancer cells in the body.

History
André-Louis Debierne, a French chemist, announced the discovery of a new element in 1899. He separated it from
pitchblende residues left by Marie and Pierre Curie after they had extracted radium. In 1899, Debierne described the
substance as similar to titanium and (in 1900) as similar to thorium. Friedrich Oskar Giesel independently discovered
actinium in 1902 as a substance being similar to lanthanum and called it "emanium" in 1904. After a comparison of
the substances half-lives determined by Debierne, Hariett Brooks in 1904, and Otto Hahn and Otto Sackur in 1905,
Debierne's chosen name for the new element was retained because it had seniority.
Articles published in the 1970s and later suggest that Debierne's results published in 1904 conflict with those
reported in 1899 and 1900. This has led some authors to advocate that Giesel alone should be credited with the
discovery. A less confrontational vision of scientific discovery is proposed by Adloff. He suggests that hindsight
criticism of the early publications should be mitigated by the nascent state of radiochemistry: highlighting the
prudence of Debierne's claims in the original papers, he notes that nobody can contend that Debierne's substance did
not contain actinium. Debierne, who is now considered by the vast majority of historians as the discoverer, lost
interest in the element and left the topic. Giesel, on the other hand, can rightfully be credited with the first
preparation of radiochemically pure actinium and with the identification of its atomic number 89.
The name actinium originates from the Ancient Greek aktis, aktinos (ακτίς, ακτίνος), meaning beam or ray. Its
symbol Ac is also used in abbreviations of other compounds that have nothing to do with actinium, such as acetyl,
acetate and sometimes acetaldehyde.

Properties
[2]
Actinium is a soft, silvery-white, radioactive, metallic element. Its estimated shear modulus is similar to that of
[3]
lead. Owing to its strong radioactivity, actinium glows in the dark with a pale blue light, which originates from the
surrounding air ionized by the emitted energetic particles. Actinium has similar chemical properties as lanthanum
and other lanthanides, and therefore these elements are difficult to separate when extracting from uranium ores.
Solvent extraction and ion chromatography are commonly used for the separation.
The first element of the actinides, actinium gave the group its name, much as lanthanum had done for the
lanthanides. The group of elements is more diverse than the lanthanides and therefore it was not until 1945 that
Glenn T. Seaborg proposed the most significant change to Dmitri Mendeleev's periodic table, by introducing the
actinides.
Actinium reacts rapidly with oxygen and moisture in air forming a white coating of actinium oxide that prevents
3+
further oxidation. As with most lanthanides and actinides, actinium exists in the oxidation state +3, and the Ac
1 2
ions are colorless in solutions. The oxidation state +3 originates from the 6d 7s electronic configuration of
actinium, that is it easily donates 3 electrons assuming a stable closed-shell structure of the noble gas radon.
The rare oxidation state +2 is only known for actinium dihydride (AcH ).
2
Chemical compounds
Only a limited number of actinium compounds are known including AcF , AcCl , AcBr , AcOF, AcOCl, AcOBr,
3 3 3
Ac S , Ac O and AcPO . Except for AcPO , they are all similar to the corresponding lanthanum compounds and
2 3 2 3 4 4
contain actinium in the oxidation state +3. In particular, the lattice constants of the analogous lanthanum and
actinium compounds differ by only a few percent.

Formula color symmetry space group No Pearson symbol a (pm) b (pm) c (pm) Z density,
3
g/cm

Ac silvery fcc Fm3m 225 cF4 531.1 531.1 531.1 4 10.07

AcH cubic Fm3m 225 cF12 567 567 567 4 8.35


2

Ac O white trigonal P3m1 164 hP5 408 408 630 1 9.18


2 3

Ac S cubic I43d 220 cI28 778.56 778.56 778.56 4 6.71


2 3

AcF [4] hexagonal P3c1 165 hP24 741 741 755 6 7.88
3 white

AcCl hexagonal P6 /m 165 hP8 764 764 456 2 4.8


3 3

AcBr white hexagonal P6 /m 165 hP8 764 764 456 2 5.85


3 3

AcOF white cubic Fm3m 593.1 8.28

AcOCl tetragonal 424 424 707 7.23

AcOBr tetragonal 427 427 740 7.89

AcPO ·0.5H O hexagonal 721 721 664 5.48


4 2

Here a, b and c are lattice constants, No is space group number and Z is the number of formula units per unit cell.
Density was not measured directly but calculated from the lattice parameters.

Oxides
Actinium oxide (Ac O ) can be obtained by heating the hydroxide at 500 °C or the oxalate at 1100 °C, in vacuum. It
2 3
crystal lattice is isotypic with the oxides of most trivalent rare-earth
metals.

Halides
Actinium trifluoride can be produced either in solution or in solid reaction. The former reaction is carried out at
room temperature, by adding hydrofluoric acid to a solution containing actinium ions. In the latter method,
actinium metal is treated with hydrogen fluoride vapors at 700 °C in an all-platinum setup. Treating actinium
trifluoride with ammonium hydroxide at 900–1000 °C yields oxyfluoride AcOF. Whereas lanthanum oxyfluoride
can be easily obtained by burning lanthanum trifluoride in air at 800 °C for an hour, similar treatment of actinium
[]
trifluoride yields no AcOF and only results in melting of the initial product.
AcF + 2 NH + H O → AcOF + 2 NH F
3 3 2 4
Actinium trichloride is obtained by reacting actinium hydroxide or oxalate with carbon tetrachloride vapors at
temperatures above 960 °C. Similar to oxyfluoride, actinium oxychloride can be prepared by hydrolyzing actinium
trichloride with ammonium hydroxide at 1000 °C. However, in contrast to the oxyfluoride, the oxychloride could
well be synthesized by igniting a solution of actinium trichloride in hydrochloric acid with ammonia.
Reaction of aluminium bromide and actinium oxide yields actinium
tribromide:
Ac O + 2 AlBr → 2 AcBr + Al O
2 3 3 2 3
3
and treating it with ammonium hydroxide at 500 °C results in the oxybromide AcOBr.
Other compounds
Actinium hydride was obtained by reduction of actinium trichloride with potassium at 300 °C, and its structure was
deduced by analogy with the corresponding LaH [5]
2 hydride. The source of hydrogen in the reaction was uncertain.
Mixing monosodium phosphate (NaH PO ) with a solution of actinium in hydrochloric acid yields white-colored
2 4
actinium phosphate hemihydrate (AcPO ·0.5H O), and heating actinium oxalate with hydrogen sulfide vapors at
4 2
1400 °C for a few minutes results in a black actinium sulfide Ac S . It may possibly be produced by acting with a
2 3
mixture of hydrogen sulfide and carbon disulfide on actinium oxide at 1000 °C.

Isotopes
Main article: Isotopes of actinium
Naturally occurring actinium is composed of one radioactive isotope; 227
Ac. Thirty-six radioisotopes have been identified, the most stable being 227
Ac with a half-life of 21.772 years, 225
Ac with a half-life of 10.0 days and 226
Ac with a half-life of 29.37 hours. All remaining radioactive isotopes have half-lives that are less than 10 hours and
the majority of them have half-lives shorter than one minute. The shortest-lived known isotope of actinium is 217
Ac (half-life of 69 nanoseconds) which decays through alpha decay and electron capture. Actinium also has two
meta states.
Purified 227
Ac comes into equilibrium with its decay products at the end of 185 days. It decays according to its 21.772-year
[]
half-life emitting mostly beta (98.8%) and some alpha particles (1.2%); the successive decay products are part of
the actinium series. Owing to the low available amounts, low energy of its beta particles (46 keV) and low intensity
of alpha radiation, 227
Ac is difficult to detect directly by its emission and it is therefore traced via its decay products. The isotopes of
actinium range in atomic weight from 206 u (206
Ac) to 236 u (236
Ac).

Isotope Production Decay Half-life

221 232 225 221 α 52 ms


Ac Th(d,9n) Pa(α)→ Ac

222 232 226 222 α 5.0 s


Ac Th(d,8n) Pa(α)→ Ac

223 232 227 223 α 2.1 min


Ac Th(d,7n) Pa(α)→ Ac

224 232 228 224 α 2.78 hours


Ac Th(d,6n) Pa(α)→ Ac

225 232 233 − 233 − 233 229 225 − 225 α 10 days


Ac Th(n,γ) Th(β )→ Pa(β )→ U(α)→ Th(α)→ Ra(β ) Ac

226 226 226 − 29.37 hours


Ac Ra(d,2n) Ac α, β
electron
capture
227 235 231 − 231 227 − 21.77 years
Ac U(α)→ Th(β )→ Pa(α)→ Ac α, β

228 232 228 − 228 − 6.15 hours


Ac Th(α)→ Ra(β )→ Ac β

229 228 229 − 229 − 62.7 min


Ac Ra(n,γ) Ra(β )→ Ac β

230 232 230 − 122 s


Ac Th(d,α) Ac β

231 232 231 − 7.5 min


Ac Th(γ,p) Ac β
232 232 232 − 119 s
Ac Th(n,p) Ac β

Occurrence and synthesis


227
Actinium is found only in traces in uranium ores as Ac – one tonne of ore
227
contains about 0.2 milligrams of actinium. The actinium isotope Ac is a
transient member of the actinium series decay chain, which begins with the
235 239 207
parent isotope U (or Pu) and ends with the stable lead isotope Pb.
225
Another actinium isotope ( Ac) is transiently present in the neptunium
237 233
series decay chain, beginning with Np (or U) and ending with thallium
205 209
( Tl) and near-stable bismuth ( Bi).

The low natural concentration, and the close similarity of physical and
chemical properties to those of lanthanum and other lanthanides, which are
always abundant in actinium-bearing ores, render separation of actinium from
the ore impractical, and complete separation was never achieved. Instead,
actinium is prepared, in milligram amounts, by the neutron irradiation of
226
Ra in a nuclear reactor.

Uraninite ores have


227 elevated concentrations of
The reaction yield is about 2% of the radium weight. Ac can further
228 actinium.
capture neutrons resulting in small amounts of Ac. After the synthesis,
actinium is separated from radium and from the products of decay and nuclear fusion, such as thorium, polonium,
lead and bismuth. The extraction can be performed with thenoyltrifluoroacetone-benzene solution from an aqueous
solution of the radiation products, and the selectivity to a certain element is achieved by adjusting the pH (to about
6.0 for actinium). An alternative procedure is anion exchange with an appropriate resin in nitric acid, which can
result in a separation factor of 1,000,000 for radium and actinium vs. thorium in a two-stage process. Actinium can
then be separated from radium, with a ratio of about 100, using a low cross-linking cation exchange resin and nitric
acid as eluant.
225
Ac was first produced artificially at the Institute for Transuranium Elements (ITU) in Germany using a cyclotron
and at St George Hospital in Sydney using a linac in 2000. This rare isotope has potential applications in radiation
therapy and is most efficiently produced by bombarding a radium-226 target with 20–30 MeV deuterium ions. This
226 225 [6]
reaction also yields Ac which however decays with a half-life of 29 hours and thus does not contaminate Ac.
Actinium metal has been prepared by the reduction of actinium fluoride with lithium vapor in vacuum at a
temperature between 1100 and 1300 °C. Higher temperatures resulted in evaporation of the product and lower ones
lead to an incomplete transformation. Lithium was chosen among other alkali metals because its fluoride is most
[]
volatile.

Applications
Owing to its scarcity, high price and radioactivity, actinium currently has no significant industrial
use.
227
Ac is highly radioactive and was therefore studied for use as an active element of radioisotope
227
thermoelectric generators, for example in spacecraft. The oxide of Ac pressed with beryllium is also an
efficient neutron source with the activity exceeding that of the standard americium-beryllium and radium-
[7] 227
beryllium pairs. In all those applications, Ac (a beta source) is merely a progenitor which generates alpha-
emitting isotopes upon its decay. Beryllium captures alpha particles and emits neutrons owing to its large cross-
section for the (α,n) nuclear reaction:
227
The AcBe neutron sources can be applied in a neutron probe – a standard device for measuring the quantity of
[8][9]
water present in soil, as well as moisture/density for quality control in highway construction. Such probes are
also used in well logging applications, in neutron radiography, tomography and other radiochemical investigations.
225 213
Ac is applied in medicine to produce Bi in a reusable generator or
can be used alone as an agent for radiation therapy, in particular targeted
alpha therapy (TAT). This isotope has a half-life of 10 days that makes it
213
much more suitable for radiation therapy than Bi (half-life 46 minutes).
225
Not only Ac itself, but also its decay products emit alpha particles which
kill cancer
Chemical structure of the DOTA carrier
cells in the body. The major difficulty with application of 225Ac was that 225
for Ac in radiation therapy.
intravenous injection of simple actinium complexes resulted in their
accumulation in the bones and liver for a period of tens of years. As a result, after the cancer cells were quickly
225
killed by alpha particles from Ac, the radiation from the actinium and its decay products might induce new
225
mutations. To solve this problem, Ac was bound to a chelating agent, such as citrate, ethylenediaminetetraacetic
acid (EDTA) or diethylene triamine pentaacetic acid (DTPA). This reduced actinium accumulation in the bones, but
the excretion from the body remained slow. Much better results were obtained with such chelating agents as
HEHA(1,4,7,10,13,16-hexaazacyclohexadecane-N,N',N'`,N'``,N'``',N'``'`-hexaacetic acid) or DOTA (1,4,7,10-
tetraazacyclododecane-1,4,7,10-tetraacetic acid) coupled to trastuzumab, a monoclonal antibody that interferes with
the HER2/neu receptor. The latter delivery combination was tested on mice and proved to be effective against
leukemia, lymphoma, breast, ovarian, neuroblastoma and prostate cancers.
227
The medium half-life of Ac (21.77 years) makes it very convenient radioactive isotope in modeling the slow
vertical mixing of oceanic waters. The associated processes cannot be studied with the required accuracy by direct
measurements of current velocities (of the order 50 meters per year). However, evaluation of the concentration
depth-profiles for different isotopes allows estimating the mixing rates. The physics behind this method is as follows:
235 231
oceanic waters contain homogeneously dispersed U. Its decay product, Pa, gradually precipitates to the bottom,
231 227
so that its concentration first increases with depth and then stays nearly constant. Pa decays to Ac; however,
231
the concentration of the latter isotope does not follow the Pa depth profile, but instead increases toward the sea
227
bottom. This occurs because of the mixing processes which raise some additional Ac from the sea bottom. Thus
231 227
analysis of both Pa and Ac depth profiles allows to model the mixing behavior.

Precautions
227
Ac is highly radioactive and experiments with it are carried out in a specially designed laboratory equipped with a
glove box. When actinium trichloride is administered intravenously to rats, about 33% of actinium is deposited into
the bones and 50% into the liver. Its toxicity is comparable to, but slightly lower than that of americium and
plutonium.

References
[1] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_actinium&action=edit
[2] Actinium, in Encyclopædia Britannica, 15th edition, 1995, p. 70
[3] Frederick Seitz, David Turnbull Solid state physics: advances in research and applications (http:// books. google. com/ books?id=F9V3a-
0V3r8C&pg=PA289), Academic Press, 1964 ISBN 0-12-607716-9 pp. 289–291
[4] Meyer, p. 71
[5] Meyer, p. 43
[6] Russell, Pamela J.; Jackson, Paul and Kingsley, Elizabeth Anne Prostate cancer methods and protocols (http:// books. google. com/ books?
id=K1y6k5bdlWkC&pg=PA336), Humana Press, 2003, ISBN 0-89603-978-1, p. 336
[7] Russell, Alan M. and Lee, Kok Loong Structure-property relations in nonferrous metals (http:/ /books.google. com/ books?id=fIu58uZTE-
gC&pg=PA470), Wiley, 2005, ISBN 0-471-64952-X, pp. 470–471
[8] Majumdar, D. K. Irrigation Water Management: Principles and Practice (http://books. google.com/books?id=hf1j9v4v3OEC&pg=PA108),
2004 ISBN 81-203-1729-7 p. 108
[9] Chandrasekharan, H. and Gupta, Navindu Fundamentals of Nuclear Science – Application in Agriculture (http:/ /books.google. com/
books?id=45IDh4Lt8xsC&pg=PA203), 2006 ISBN 81-7211-200-9 pp. 202 ff

Bibliography
• Meyer, Gerd and Morss, Lester R. Synthesis of lanthanide and actinide compounds (http://books.google.com/
books?id=bnS5elHL2w8C&pg=PA87), Springer, 1991, ISBN 0-7923-1018-7

External links
• Actinium (http://www.periodicvideos.com/videos/089.htm) at The Periodic Table of Videos (University
of Nottingham)
• NLM Hazardous Substances Databank – Actinium, Radioactive (http://toxnet.nlm.nih.gov/cgi-bin/sis/search/
r?dbs+hsdb:@term+@na+@rel+actinium,+radioactive)
• Actinium (http://radchem.nevada.edu/classes/rdch710/files/actinium.pdf) in Haire, Richard G. (2006).
Morss; Edelstein, Norman M.; Fuger, Jean, eds. The Chemistry of the Actinide and Transactinide Elements (3rd
ed.). Dordrecht, The Netherlands: Springer. ISBN 1-4020-3555-1.
90 Thorium 1

90 – Thorium
Not to be confused with Thallium or Thulium.

Thorium
Th
90

actinium ← thorium → protactiniumCe



Th

(Uqb)

Thorium in the periodic table


Appearance

silvery, often with black tarnish

General properties

Name, symbol, number thorium, Th, 90

Pronunciation /ˈθɔəriəm/
THAWR-ee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight 232.0377(4)

Electron configuration 2 2
[Rn] 6d 7s
2, 8, 18, 32, 18, 10, 2

Physical properties

Phase solid

Density (near r.t.) −3


11.7 g·cm

Melting point 2115 K, 1842 °C, 3348 °F

Boiling point 5061 K, 4788 °C, 8650 °F


Heat of fusion −1
13.81 kJ·mol

Heat of vaporization −1
514 kJ·mol

Molar heat capacity −1 −1


26.230 J·mol ·K

Vapor pressure

P (Pa) 1 10 100 1k 10 k 100 k

at T (K) 2633 2907 3248 3683 4259 5055

Atomic properties

Oxidation states 4, 3, 2, 1
(weakly basic oxide)

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 587 kJ·mol

−1
2nd: 1110 kJ·mol

−1
3rd: 1930 kJ·mol

Atomic radius 179 pm

Covalent radius 206±6 pm

Miscellanea

Crystal structure face-centered cubic

Magnetic ordering [1]


paramagnetic

Electrical resistivity (0 °C) 147 nΩ·m

Thermal conductivity −1 −1
54.0 W·m ·K

Thermal expansion −1 −1
(25 °C) 11.0 µm·m ·K

Speed of sound (thin rod) −1


(20 °C) 2490 m·s

Young's modulus 79 GPa

Shear modulus 31 GPa

Bulk modulus 54 GPa

Poisson ratio 0.27

Mohs hardness 3.0

Vickers hardness 350 MPa

Brinell hardness 400 MPa

CAS registry number 7440-29-1

History

Discovery Jöns Jakob Berzelius (1829)

Most stable isotopes

Main article: Isotopes of thorium


iso NA half-life DM DE (MeV) DP

228 trace 1.9116 y α 5.520 224


Th Ra

229 trace 7340 y α 5.168 225


Th Ra

230 trace 75380 y α 4.770 226


Th Ra

231 trace 25.5 h − 0.39 231


Th β Pa

232 100% 10 α 4.083 228


Th 1.405×10 y Ra

234 trace 24.1 d − 0.27 234


Th β Pa

• v
• t
[2]
• e

Thorium is a naturally occurring radioactive chemical element with the symbol Th and atomic number 90. It was
discovered in 1828 by the Norwegian mineralogist Morten Thrane Esmark and identified by the Swedish chemist
Jöns Jakob Berzelius, who named it after Thor, the Norse god of thunder.
Thorium produces a radioactive gas, radon-220, as one of its decay products. Secondary decay products of thorium
include radium and actinium. In nature, virtually all thorium is found as thorium-232, which undergoes alpha decay
with a half-life of about 14.05 billion years. Other isotopes of thorium are short-lived intermediates in the decay
chains of higher elements, and only found in trace amounts. Thorium is estimated to be about three to four times
more abundant than uranium in the Earth's crust, and is chiefly refined from monazite sands as a by-product of
extracting rare earth metals.
Thorium was once commonly used as the light source in gas mantles and as an alloying material, but these
applications have declined due to concerns about its radioactivity. Thorium is also used as an alloying element in
nonconsumable TIG welding electrodes. It remains popular as a material in high-end optics and scientific
instrumentation; thorium and uranium are the only radioactive elements with major commercial applications that do
not rely on their radioactivity.
Canada, China, Germany, India, the Netherlands, the United Kingdom and the United States have experimented with
using thorium as a substitute nuclear fuel in nuclear reactors. When compared to uranium, there is a growing interest
in thorium-based nuclear power due to its greater safety benefits, absence of non-fertile isotopes and its higher
occurrence and availability. India's three stage nuclear power programme is possibly the most well known and well
funded of such efforts.

Characteristics

Physical properties
Pure thorium is a silvery-white metal that is air-stable and retains its luster for several months. When contaminated
with the oxide, thorium slowly tarnishes in air, becoming gray and finally black. The degree of oxide contamination
[3]
greatly influences thorium's physical properties.
The purest specimens often contain several tenths of a percent of the oxide. Pure thorium is soft, very ductile, and
can be cold-rolled, swaged, and drawn. Thorium is dimorphic, changing at 1360 °C from a face-centered cubic to a
body-centered cubic structure; a body-centered tetragonal lattice form exists at high pressure with impurities driving
[3]
the exact transition temperatures and pressures.
Powdered thorium metal is often pyrophoric and requires careful handling. When heated in air, thorium metal
turnings ignite and burn brilliantly with a white light. Thorium has one of the largest liquid temperature ranges of
any element, with 2946 °C between the melting point and boiling point. Thorium metal is paramagnetic with a
2 2 [3]
ground state of 6d 7s .

Chemical properties
Thorium is slowly attacked by water, but does not dissolve readily in most common acids, except hydrochloric
acid. It dissolves in concentrated nitric acid containing a small amount of catalytic fluoride ion.
Thorium's oxide is ThO . Thorium's most common oxidation state is +4, as in ThF , but thorium also has an
2 4
oxidation state of +3, as in ThI . Thorium has been shown to activate carbon-hydrogen bonds, forming unusual
3
compounds. Thorium atoms can also bond to more atoms than any other element. For instance, in the compound
[4]
thorium aminodiboranate, thorium has a coordination number of fifteen.

Compounds
Thorium compounds are stable in the +4 oxidation state.
Thorium dioxide has the highest melting point (3300 °C) of all oxides.
Thorium(IV) nitrate and thorium(IV) fluoride are known in their hydrated forms: Th(NO
3)

4·4H

2O and ThF

4·4H

2O, respectively. Thorium(IV) carbonate, Th(CO

3)

2, is also known.
4+
When treated with potassium fluoride and hydrofluoric acid, Th forms the complex anion ThF2−
6, which precipitates as an insoluble salt, K

2ThF

6.

Thorium(IV) hydroxide, Th(OH)


4, is highly insoluble in water, and is not amphoteric. The peroxide of thorium, ThO or Th(O ) , is rare in being an
4 22
insoluble solid. This property can be used to separate thorium from other ions in solution.
4+
In the presence of phosphate anions, Th forms precipitates of various compositions, which are insoluble in water
and acid solutions.
Thorium monoxide has recently been produced through laser ablation of thorium in the presence of oxygen. This
[5]
highly polar molecule has the largest known internal electric field.
Isotopes
Main article: Isotopes of
thorium
210
Twenty-seven radioisotopes have been characterized, with a range in atomic weight from 210 u ( Th) to 236 u
236
( Th). The most stable isotopes are:
232
• Th with a half-life of 14.05 billion years, it represents all but a trace of naturally occurring
thorium.
230 238
• Th with a half-life of 75,380 years. Occurs as the daughter product of U decay.
229
• Th with a half-life of 7340 years. It has a nuclear isomer (or metastable state) with a remarkably low excitation
energy of 7.6 eV.
228
• Th with a half-life of 1.92 years.
All of the remaining radioactive isotopes have half-lives that are less than thirty days and the majority of these have
half-lives that are less than ten minutes.

Applications
Thorium is a component of the magnesium alloy series, called Mag-Thor, used in aircraft engines and rockets and
imparting high strength and creep resistance at elevated temperatures. Thoriated magnesium was used to build the
CIM-10 Bomarc missile, although concerns about radioactivity have resulted in several missiles being removed from
public display.
Thorium is also used in its oxide form (thoria) in gas tungsten arc welding (GTAW) to increase the high-temperature
[6]
strength of tungsten electrodes and improve arc stability. The electrodes labeled EWTH-1 contain 1% thoria, while
the EWTH-2 contain 2%. In electronic equipment, thorium coating of tungsten wire improves the electron emission
of heated cathodes.
Thorium is a very effective radiation shield, although it has not been used for this purpose as much as lead or
depleted uranium. Uranium-thorium age dating has been used to date hominid fossils, seabeds, and mountain
[7]
ranges.
Environmental concerns related to radioactivity led to a sharp decrease in demand for nonnuclear uses of
[7]
thorium in the 2000s.

Thorium compounds
Thorium dioxide (ThO ) and thorium nitrate (Th(NO
2
3)

4) are used in mantles of portable gas lights, including natural gas


lamps, oil lamps and camping lights. These mantles glow with an
intense white light (unrelated to radioactivity) when heated in a gas
flame, and its color could be shifted to yellow by addition of cerium.

Thorium dioxide is a material for heat-resistant ceramics, e.g., for high-


[6]
temperature laboratory crucibles. When added to glass, it helps Examples of thoriated lenses.
increase refractive index and decrease dispersion. Such glass finds
application in high-quality lenses for cameras and scientific instruments. The radiation from these lenses can self-
[8]
darken (yellow) them over a period of years and degrade film, but the health risks are minimal. Yellowed
lenses may be restored to their original colorless state with lengthy exposure to intense UV light.

Thorium dioxide was used to control the grain size of tungsten metal used for spirals of electric lamps. Thoriated
tungsten elements are found in the filaments of magnetron tubes. Thorium is added because of its ability to emit
electrons at relatively low temperatures when heated in vacuum. Those tubes generate microwave frequencies and
are applied in microwave ovens and radars.
[6]
Thorium dioxide has been used as a catalyst in the conversion of ammonia to nitric acid, in petroleum cracking and
in producing sulfuric acid. It is the active ingredient of Thorotrast, which was used as radiocontrast agent for X-ray
diagnostics. This use has been abandoned due to its carcinogenic nature.
Despite its radioactivity, thorium fluoride (ThF ) is used as an antireflection material in multilayered optical
4
coatings. It has excellent optical transparency in the range 0.35–12 µm, and its radiation is primarily due to alpha
particles, which can be easily stopped by a thin cover layer of another material. Thorium fluoride was also used in
manufacturing carbon arc lamps, which provided high-intensity illumination for movie projectors and search lights.

Thorium as a nuclear fuel


See also: Thorium-based nuclear
power

Benefits and challenges


The naturally occurring isotope thorium-232 is a fertile material, and with a suitable neutron source can be used as
nuclear fuel in nuclear reactors, including breeder reactors. In 1997, the U.S. Energy Department underwrote
research into thorium fuel, and research also was begun in 1996 by the International Atomic Energy Agency (IAEA),
to study the use of thorium reactors. Nuclear scientist Alvin Radkowsky of Tel Aviv University in Israel founded a
consortium to develop thorium reactors, which included other companies: Raytheon Nuclear Inc., Brookhaven
[9]
National Laboratory and the Kurchatov Institute in Moscow.
Radkowsky was chief scientist in the U.S. nuclear submarine program directed by Admiral Hyman Rickover and
later headed the design team that built the USA's first civilian nuclear power plant at Shippingport, Pennsylvania,
which was a scaled-up version of the first naval reactor. The third Shippingport core, initiated in 1977, bred thorium.
Even earlier examples of reactors using fuel with thorium exist, including the first core at the Indian Point Energy
Center in 1962.
Some countries, including India, are now investing in research to build thorium-based nuclear reactors. A 2005
report by the International Atomic Energy Agency discusses potential benefits along with the challenges of
[10]
thorium reactors. India has also made thorium-based nuclear reactors a priority with its focus on developing
[11][12]
fast breeder technology.
Some benefits of thorium fuel when compared with uranium were summarized as follows:
• Weapons-grade fissionable material (233U) is harder to retrieve safely and clandestinely from a thorium
reactor;
• Thorium mining produces a single pure isotope, whereas the mixture of natural uranium isotopes must be
enriched to function in most common reactor designs. The same cycle could also use the fissionable U-238
component of the natural uranium, and also contained in the depleted reactor fuel;
[13]
• Thorium cannot sustain a nuclear chain reaction without priming, so fission stops by default in an
accelerator driven reactor.
When used in a breeder-like reactor, however, unlike uranium-based light water reactors, thorium requires
irradiation and reprocessing before the above-noted advantages of thorium-232 can be realized, which initially
[7]
makes solid thorium fuels more expensive than uranium fuels. But experts note that "the second thorium
reactor may activate a third thorium reactor. This could continue in a chain of reactors for a millennium if we so
choose." They add that because of thorium's abundance, it will not be exhausted in 1,000 years.
The Thorium Energy Alliance (TEA), an educational advocacy organization, emphasizes that "there is enough
[14]
thorium in the United States alone to power the country at its current energy level for over 10,000 years."
90 Thorium 1

Thorium energy fuel cycle


Main article: Thorium fuel cycle
"Thorium is like wet wood […it] needs to be turned into fissile uranium just as wet wood needs to be dried in a furnace."

— Ratan Kumar Sinha, current Chairman of the Atomic Energy Commission of India.
238 232 233
Like U, Th is not fissile itself, but it is fertile: it absorbs slow neutrons to produce, after two beta decays, U,
[7]
which is fissile. The preparation of thorium fuel does not require isotopic separation, unlike the preparation of
uranium fuels.
233
The thorium fuel cycle creates U, which, if separated from the reactor's fuel, could with some difficulty be
used for making nuclear weapons. This is one reason why a liquid-fuel cycle (e.g., molten salt reactor or
233
MSR) is preferred—only a limited amount of U ever exists in the reactor and its heat-transfer systems,
preventing access to weapons material. However, the neutrons produced by the reactor can be absorbed by a
233 239 233
thorium or uranium blanket to produce fissile U or Pu. Also, the U could be continuously extracted from
the molten fuel as the reactor runs. Neutrons from the decay of uranium-233 can be fed back into the fuel cycle
[7]
to start the cycle again.
233 233
The neutron flux from spontaneous fission of U is negligible. U can thus be used easily in a simple gun-type
nuclear bomb design. In 1977, a light-water reactor at the Shippingport Atomic Power Station was used to establish a
232 233
Th- U fuel cycle. The reactor worked until its decommissioning in 1982. Thorium can be and has been used to
power nuclear energy plants using both the modified traditional Generation III reactor design and prototype
Generation IV reactor designs. The use of thorium as an alternative fuel is one innovation being explored by the
[15]
International Project on Innovative Nuclear Reactors and Fuel Cycles (INPRO), conducted by the International
Atomic Energy Agency (IAEA).
Unlike its use in Molten salt reactors, when using solid thorium in modified light water reactor (LWR) problems
include: the undeveloped technology for fuel fabrication; in traditional, once-through LWR designs potential
228
problems in recycling thorium due to highly radioactive Th; some weapons proliferation risk due to production of
233
U; and the technical problems (not yet satisfactorily solved) in reprocessing. Much development work is still
required before the thorium fuel cycle can be commercialized for use in LWR. The effort required has not seemed
worth it while abundant uranium is available.

Commercial nuclear power station


India's Kakrapar-1 reactor is the world's first reactor that uses thorium rather than depleted uranium for power
flattening across the reactor core. India, which has about 25% of the world's thorium reserves, is developing a 300
MW prototype of a thorium-based Advanced Heavy Water Reactor (AHWR). The prototype is expected to be fully
operational by 2016, after which they plan to construct five more reactors. The reactor is a fast breeder reactor and
uses a plutonium core rather than an accelerator to produce neutrons. As accelerator-based systems can operate at
sub-criticality they could be developed too, but that would require more research. India currently envisages meeting
30% of its electricity demand through thorium-based reactors by 2050.

Existing thorium energy projects


Main article: Thorium-based nuclear power § Current thorium projects
While research is under way in many countries, only India is building utility-scale plants, mostly planned to be
thorium-fueled. In 2012, the first commercial fast reactor capable of using thorium was allegedly nearing
[16]
completion, according to Srikumar Banerjee, former Chairman of the Indian Atomic Energy Commission.
Cadillac produced a concept vehicle in 2009 using thorium as its propulsion. The vehicle was called the Cadillac
[17]
World Thorium Fuel Concept. The vehicle affectionately became known as "the WTF".
Projects combining uranium and thorium
Fort St. Vrain Generating Station, a demo HTGR in Colorado, USA, operating from 1977 until 1992, employed
enriched uranium fuel that also contained thorium. This resulted in high fuel efficiency because the thorium was
converted to uranium and then fissioned.

History
Morten Thrane Esmark found a black mineral on Løvøya island,
Norway, and gave a sample to his father, Jens Esmark, a noted
mineralogist. The elder Esmark was not able to identify it and sent a
sample to the Swedish chemist Jöns Jakob Berzelius for examination in
1828. Berzelius determined that it contained a new element, which he
[6]
named thorium after Thor, the Norse god of thunder. He published
[18]
his findings in 1829. Berzelius reused the name of a previous
element discovery from a mineral from the Falun, which later proved
to be a yttrium mineral. The metal had no practical uses until Carl Auer
[6]
von Welsbach invented the gas mantle in 1885. The Earth's thorium originated when ancient stars
went supernova.
Thorium was first observed to be radioactive in 1898, independently,
by the Polish-French physicist Marie Curie and the German chemist
[19]
Gerhard Carl Schmidt. Between 1900 and 1903, Ernest Rutherford and Frederick Soddy showed how thorium
decayed at a fixed rate over time into a series of other elements. This observation led to the identification of half-life
as one of the outcomes of the alpha particle experiments that led to their disintegration theory of radioactivity.
The crystal bar process (or "iodide process") was discovered by Anton Eduard van Arkel and Jan Hendrik de Boer in
1925 to produce high-purity metallic thorium.
230
The name ionium was given early in the study of radioactive elements to the Th isotope produced in the decay
238
chain of U before it was realized that ionium and thorium were chemically identical. The symbol Io was used for
this supposed element.
Thorium-232 is a primordial nuclide, having existed in its current form
for over 4.5 billion years, predating the formation of the Earth; it was
forged in the cores of dying stars through the r-process and scattered
[20]
across the galaxy by supernovas. Its radioactive decay produces a
significant amount of the Earth's internal heat.

232
The radiogenic heat from the decay of Th is a
major contributor to the earth's internal heat
budget.
Occurrence
Thorium is found in small amounts in most rocks and soils; it is three
times more abundant than tin in the Earth's crust and is about as
[21]
common as lead. Soil commonly contains an average of around 6
[22]
parts per million (ppm) of thorium. Thorium occurs in several
minerals including thorite (ThSiO ), thorianite (ThO + UO ) and
4 2 2
monazite. Thorianite is a rare mineral and may contain up to about
12% thorium oxide. Monazite contains 2.5% thorium, allanite has
[23]
0.1 to 2% thorium and zircon can have up to 0.4% thorium.
Thorium-containing minerals occur on all continents. Thorium is
Partial North American map of thorium
several times more abundant in Earth's crust than all isotopes of
concentrations from the United States Geological
uranium combined and thorium-232 is several hundred times more Survey.
[7]
abundant than uranium-235.

Thorium concentrations near the surface of the earth can be mapped


using gamma spectroscopy. The same technique has been used to
detect concentrations on the surface of the moon; the near side has
high abundances of relatively Thorium-rich KREEP, while the
Compton–Belkovich Thorium Anomaly was detected on the far side.
[24]
Martian thorium has also been mapped by 2001 Mars Odyssey.
232
Th decays very slowly (its half-life is comparable to the age of the Thorium concentrations on the moon, as
universe) but other thorium isotopes occur in the thorium and uranium mapped by Lunar Prospector.

decay chains. Most of these are short-lived and hence much more
232
radioactive than Th, though on a mass basis they are negligible.

Extraction
Main article:
Monazite
Thorium has been extracted chiefly from monazite through a complex
multi-stage process. The monazite sand is dissolved in hot
concentrated sulfuric acid (H SO ). Thorium is extracted as an
2 4
insoluble residue into an organic phase containing an amine. Next it is
separated or stripped using an ion such as nitrate, chloride, hydroxide,
or carbonate, returning the thorium to an aqueous phase. Finally, the
thorium is precipitated and collected.

Several methods are available for producing thorium metal: it can be


obtained by reducing thorium oxide with calcium, by electrolysis of Monazite, a rare earth and thorium phosphate
anhydrous thorium chloride in a fused mixture of sodium and mineral, is the primary source of the world's
potassium chlorides, by calcium reduction of thorium tetrachloride thorium.

mixed with anhydrous zinc chloride, and by reduction of thorium


tetrachloride with an alkali metal.

Reserve estimates
Present knowledge of the distribution of thorium resources is poor
because of the relatively low-key exploration efforts arising out of
insignificant demand. There are two sets of estimates that define world
thorium reserves, one set by the US Geological Survey (USGS) and the
other supported by reports from the OECD and the International
Atomic Energy Agency (the IAEA). Under the USGS estimate, USA,
Australia, and India have particularly large reserves of thorium.

India and Australia are believed to possess about 300,000 tonnes each;
i.e., each has 25% of the world's thorium reserves. In the OECD
reports, however, estimates of Australia's Reasonably Assured
Reserves (RAR) of thorium indicate only 19,000 tonnes and not
300,000 tonnes as indicated by USGS. The two sources vary wildly for
countries such as Brazil, Turkey, and Australia, however, both reports
appear to show some consistency with respect to India's thorium India's thorium is mostly found in a contiguous
belt formed by its eastern coastal states. 2012
reserve figures, with 290,000 tonnes (USGS) and 319,000 tonnes
reserve estimates: 35% (Andhra
(OECD/IAEA). Pradesh) 15-20% (Tamil Nadu, Odisha) 10-15%
(Kerala, West Bengal) 0-5% (Bihar)
Both the IAEA and OECD appear to conclude that India may possess
the lion's share of world's thorium deposits.
The IAEA's 2005 report estimates India's reasonably assured reserves of thorium at 319,000 tonnes, but mentions
recent reports of India's reserves at 650,000 tonnes. A government of India estimate, shared in the country's
Parliament in August 2011, puts the recoverable reserve at 846,477 tonnes. The Indian Minister of State V.
Narayanasamy stated that as of May 2013, the country's thorium reserves were 11.93 million tonnes (monazite,
having 9-10% ThO ), with a significant majority (8.59 Mt; 72%) found in the three eastern coastal states of Andhra
2
Pradesh (3.72 Mt; 31%), Tamil Nadu (2.46 Mt; 21%) and Odisha (2.41 Mt; 20%).
The prevailing estimate of the economically available thorium reserves comes from the U.S. Geological Survey,
Mineral Commodity Summaries (1996–2010):
USGS Estimates in tonnes (1999)
Country Reserves

Australia 300,000

India 290,000

Norway 170,000

United States 160,000

Canada 100,000

South Africa 35,000

Brazil 16,000

Other Countries 95,000

World Total 1,200,000

USGS Estimates in tonnes (2011)


Country Reserves

India 963,000

United States 440,000

Australia 300,000

Canada 100,000

South Africa 35,000

Brazil 16,000

Malaysia 4,500

Other Countries 90,000

World Total 1,913,000

Note: The OECD/NEA report notes that the estimates (that the Australian figures are based on) are subjective, due to
the variability in the quality of the data, a lot of which is old and incomplete. Adding to the confusion are subjective
claims made by the Australian government (in 2009, through its Geoscience department) that combine the
reasonably assured reserves (RAR) estimates with "inferred" data (i.e., subjective guesses). This strange combined
figure of RAR and "guessed" reserves yields a figure, published by the Australian government, of 489,000 tonnes,
however, using the same criteria for Brazil or India would yield reserve figures of between 600,000 to 1,300,000
tonnes for Brazil and between 300,000 to 600,000 tonnes for India. Irrespective of isolated claims by the Australian
government, the most credible third-party and multi-lateral reports, those of the OECD/IAEA and the USGS,
consistently report high thorium reserves for India while not doing the same for Australia.
Another estimate of reasonably assured reserves (RAR) and estimated additional reserves (EAR) of thorium comes
from OECD/NEA, Nuclear Energy, "Trends in Nuclear Fuel Cycle", Paris, France (2001):
IAEA Estimates in tonnes (2005)
Country RAR Th EAR Th

India 519,000 21%

Australia 489,000 19%

USA 400,000 13%

Turkey 344,000 11%

Venezuela 302,000 10%

Brazil 302,000 10%

Norway 132,000 4%

Egypt 100,000 3%

Russia 75,000 2%

Greenland 54,000 2%

Canada 44,000 2%

South Africa 18,000 1%

"Other countries" 33,000 2%

"World total" 2,810,000

The preceding reserve figures refer to the amount of thorium in high-concentration deposits inventoried so far and
19
estimated to be extractable at current market prices; millions of times more total exist in Earth's 3×10 tonne crust,
[25]
around 120 trillion tons of thorium, and lesser but vast quantities of thorium exist at intermediate concentrations.
[26][27]
Proved reserves are "a poor indicator of the total future supply of a mineral resource."
The Lemhi Pass, along the Idaho-Montana border, has one of the world's largest known high quality thorium
deposits. Thorium Energy, Inc. has the mineral rights to approximately 1360 acres (5.5 sq km) of it and states that
they have proven thorium oxide reserves of 600 thousand tons and probable reserves of an additional 1.8 million tons
[28]
within their claim.
In event of a thorium fuel cycle, Conway granite with 56 (±6) parts per million thorium could provide a major
low-grade resource; a 307 sq mile (795 sq km) "main mass" in New Hampshire is estimated to contain over three
million metric tons per 100 feet (30 m) of depth (i.e. 1 kg thorium in eight cubic metres of rock), of which two-thirds
is "readily leachable". Even common granite rock with 13 PPM thorium concentration (just twice the crustal average,
[29]
along with 4 ppm uranium) contains potential nuclear energy equivalent to 50 times the entire rock's mass in coal,
although there is no incentive to resort to such very low-grade deposits so long as much higher-grade deposits
[30]
remain available and cheaper to extract. Thorium has been produced in excess of demand from the refining of
[31]
rare earth elements.
Dangers and biological roles
Powdered thorium metal is pyrophoric and often ignites spontaneously
in air. Natural thorium decays very slowly compared to many other
radioactive materials, and the alpha radiation emitted cannot penetrate
human skin meaning owning and handling small amounts of thorium,
such as a gas mantle, is considered safe. Exposure to an aerosol of
thorium, however, can lead to increased risk of cancers of the lung,
[citation needed]
pancreas, and blood, as lungs and other internal organs
can be penetrated by alpha radiation. Exposure to thorium internally
leads to increased risk of liver diseases. Thorium is radioactive and Experiment on the effect of radiation (from an
unburned thorium gas mantle) on the germination
produces a radioactive gas, radon-220, as one of its decay products.
and growth of timothy-grass seed; from Popular
Secondary decay products of thorium include radium and actinium. Science, 1909.
Because of this, there are concerns about the safety of thorium
mantles.
[32]
Some nuclear safety agencies make recommendations about their use. Production of gas mantles has led to some
safety concerns during manufacture.

The element has no known biological role. Humans typically consume three micrograms per day of thorium. Of this,
99.98% does not remain in the body. Out of the thorium that does remain in the body, three quarters of it
accumulates in the skeleton. A number of thorium compounds are chemically moderately toxic. People who work
with thorium compounds are at a risk of dermatitis. It can take as much as thirty years after the ingestion of thorium
for symptoms to manifest themselves.

References
[1] Magnetic susceptibility of the elements and inorganic compounds (http://www-d0.fnal. gov/ hardware/cal/ lvps_info/engineering/
elementmagn. pdf), in Handbook of Chemistry and Physics 81st edition, CRC press.
[2] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_thorium&action=edit
[3] Wickleder 2006, p.
61.
[4] Scott R. Daly et al., "Synthesis and Properties of a Fifteen-Coordinate Complex: The Thorium Aminodiboranate [Th(H BNMe BH )
]",
Angewandte Chemie International Edition 49:3379-3381 (2010) 3 2 3 4

[5] "The ACME EDM Experiment." (http:/ /www.electronedm.org/ )


electronedm.org [6] Wickleder 2006, p. 52.
[7] Wickleder 2006, p. 53.
[8] Thoriated Camera Lens (ca. 1970s) (http://www.orau.org/ptp/collection/consumer products/ cameralens. htm)
[9] Bulletin of the Atomic Scientists (http:/ /books.google. com/ books?id=2wsAAAAAMBAJ&pg=PA19). September/October 1997 pp. 19–20
[10] "Thorium fuel cycle — Potential benefits and challenge" (http:/ /www-pub.iaea.org/ MTCD/publications/PDF/ TE_1450_web.pdf),
IAEA, May 2005
[11] Progress on India's Thorium Nuclear Reactor and South Africa's Pebble Bed (http:// nextbigfuture.com/2008/08/ indias-thorium-nuclear-
reactor-and.html). Nextbigfuture.com (22 August 2008). Retrieved on 2011-05-01.
[12] Nuclear Power in India|Indian Nuclear Energy (http://www.world-nuclear.org/ info/inf53.html). World-nuclear.org. Retrieved on 1 May
2011.
[13] "Thorium: Is It the Better Nuclear Fuel?" (http:/ /www.cavendishscience.org/ bks/ nuc/ thrupdat.htm), Cavendish Press, Dec
2008 [14] What Is Thorium? (http:/ /thoriumenergyalliance. com/ ThoriumSite/ portal.html) Thorium Energy Alliance
[15] Sollychin, Ray. (3 September 2009) Exploring Fuel Alternatives (http://www.iaea.org/ Publications/Magazines/Bulletin/Bull511/
51104894344. html). Iaea.org. Retrieved on 2011-05-01.
[16] "First commercial fast reactor nearly ready" (http://www.thehindu.com/todays-paper/tp-national/article3582922.ece), The Hindu, June
29, 2012
[17] "Not in Detroit: Cadillac World Thorium Fuel Concept" (http://www.autoblog.com/2009/ 01/ 13/ not-in-detroit-cadillac-world-thorium-fuel-
concept), AutoBlog, January 13, 2009
[18] (modern citation: Annalen der Physik, vol. 92, no. 7, pp. 385–415)
[19] (modern citation: Annalen der Physik, vol. 301, pages 141–151 (1898)).
[20] Synthesis of heavy elements (http://www.gsi.de/forschung/kp/kp2/nuc-astro/ HeavyElements_e. html)
[21] Wickleder 2006, p. 55.
[22] THORIUM (http://www.atsdr.cdc. gov/tfacts147. pdf) Agency for Toxic Substances and Disease Registry. July 1999.
[23] Wickleder 2006, p. 56.
[24] Lunar & Planetary Lab at The University of Arizona (http:// grs.lpl.arizona.edu/latestresults.jsp?lrid=32) January 2008: Thorium Map
[25] Ragheb, M. (12 August 2011) Thorium Resources In Rare Earth Elements (http:// www.scribd.com/doc/105448071/
Thorium-Resources-in-Rare-Earth-Elements-Ragheb-M-Aug-2011). scribd.com
[26] American Geophysical Union, Fall Meeting 2007, abstract #V33A-1161. Mass and Composition of the Continental Crust (http:// adsabs.
harvard. edu/ abs/2007AGUFM.V33A1161P)
[27] James D. Gwartney, Richard L. Stroup, Russell S. Sobel, David MacPherson. Economics: Private and Public Choice, 12th Edition.
South-Western Cengage Learning, p. 730 (http:// books.google.com/books?id=yIbH4R77OtMC&pg=PA730)
[28] Lemhi Pass Thorium (http:/ /www.thoriumenergy.com/index.php?option=com_content&task=view&id=17&Itemid=33).
thoriumenergy.com
[29] Hubbert, M. King Nuclear Energy and the Fossil Fuels (http:// www.energybulletin.net/ node/ 13630). American Petroleum Institute
Conference, 8 March 1956. Republished on 8 March 2006, by the Energy Bulletin.
[30] Brown, Harrison (1954). The Challenge of Man's Future. New York: Viking Press.
[31] Hedrick, James B. (1997) Thorium (http://minerals.usgs. gov/minerals/pubs/commodity/thorium/690497.pdf). US Geological Survey.
[32] Radioactivity in Lantern Mantles (http://web.archive.org/web/20071014211034/http://arpansa. gov.au/RadiationProtection/
Factsheets/is_lantern.cfm). Australian Radiation Protection and Nuclear Safety Agency

Bibliography
• Wickleder, Mathias S.; Fourest, Blandine; Dorhourt, Peter K. (2006). "Thorium". In Morss, Lester R.; Edelstein,
Norman M.; Fuger, Jean. The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Springer
Science+Business Media. ISBN 1-4020-3555-1.

Further reading
• Martin, Richard (2012). Super Fuel: Thorium, the Green Energy Source for the Future (1st ed.). Palgrave
MacMillan. p. 240. ISBN 978-0-230-11647-4.
• Hargraves, Robert (25 July 2012). Thorium: Energy Cheaper than Coal. CreateSpace Independent Publishing
Platform. p. 482. ISBN 978-1478161295.

External links
• International Thorium Energy Committee – iThEC (http://www.ithec.org)
• International Thorium Energy Organisation – IThEO.org (http://www.itheo.org)
• European Nuclear Society – Natural Decay Chains (http://www.euronuclear.org/info/encyclopedia/d/
decaybasinnatural.htm)
• ATSDR CDC ToxFAQs: health questions about thorium (http://www.atsdr.cdc. gov/tfacts147. html)
• FactSheet on Thorium (http://www.world-nuclear.org/info/inf62.html), World Nuclear Association
• Thorium TV – A review of the element (http://www.thorium.tv/en/index.php)
• EnergyFromThorium.com – Content-rich site on Thorium as a future energy source, and its extraction
technology (http://energyfromthorium.com/)
• TED talk by former NASA engineer Kirk Sorensen about Thorium energy production (video) (http://www.ted.
com/talks/kirk_sorensen_thorium_an_alternative_nuclear_fuel.html)
• India's experimental Thorium Fuel Cycle Nuclear Reactor (NDTV Report)
(http://www.youtube.com/ watch?v=Nl5DiTPw3dk)
• Thorium Remix 2011 – 120 minute Creative Commons Share-Alike documentary on Thorium as an energy
source (http://www.youtube. com/watch?v=P9M yYbsZ4)
• Newspaper article about thorium power in India (http://www.guardian.co.uk/environment/2011/nov/01/
india-thorium-nuclear-plant)
• China Blazes Trail for Clean Nuclear Power from Thorium (http://www.telegraph.co.uk/finance/comment/
ambroseevans_pritchard/9784044/China-blazes-trail-for-clean-nuclear-power-from-thorium.html)
• Thorium (http://www.periodicvideos.com/videos/090.htm) at The Periodic Table of Videos (University
of Nottingham)
• Thorium Deposits of the United States—Energy Resources for the Future? (http://pubs.usgs.gov/circ/1336/)
(USGS, 2009)

91 – Protactinium
Protactinium
Pa
91

thorium ← protactinium → uraniumPr



Pa

(Uqt)

Protactinium in the periodic table


Appearance

bright, silvery metallic luster

General properties

Name, symbol, number protactinium, Pa, 91

Pronunciation /ˌproʊtækˈtɪniəm/
PROH-tak-TIN-ee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight 231.03588

Electron configuration 2 1 2
[Rn] 5f 6d 7s
2, 8, 18, 32, 20, 9, 2

Physical properties

Phase solid

Density (near r.t.) −3


15.37 g·cm

Melting point 1841 K, 1568 °C, 2854 °F

Boiling point ? 4300 K, ? 4027 °C, ? 7280 °F

Heat of fusion −1
12.34 kJ·mol
91 Protactinium 1

Heat of vaporization −1
481 kJ·mol

Atomic properties

Oxidation states 2, 3, 4, 5
(weakly basic oxide)

Electronegativity 1.5 (Pauling scale)

Ionization energies −1
1st: 568 kJ·mol

Atomic radius 163 pm

Covalent radius 200 pm

Miscellanea

Crystal structure tetragonal

Magnetic ordering [1]


paramagnetic

Electrical resistivity (0 °C) 177 nΩ·m

Thermal conductivity −1 −1
47 W·m ·K

CAS registry number 7440-13-3

History

Prediction Dmitri Mendeleev (1869)

Discovery William Crookes (1900)

First isolation William Crookes (1900)

Named by Otto Hahn and Lise Meitner (1917–8)

Most stable isotopes

Main article: Isotopes of protactinium

iso NA half-life DM DE (MeV) DP

229 syn 1.5 d ε 0.311 229


Pa Th

230 syn 17.4 d ε 1.310 230


Pa Th

231 ~100% 4 α 5.150 227


Pa 3.276×10 y Ac

232 syn 1.31 d − 1.337 232


Pa β U

233 trace 26.967 d − 0.5701 233


Pa β U

234m trace 1.17 min − 2.29 234


Pa β U

234 trace 6.75 h − 2.195 234


Pa β U

• v
• t
[2]
• e
Protactinium is a chemical element with the symbol Pa and atomic number 91. It is a dense, silvery-gray metal
which readily reacts with oxygen, water vapor and inorganic acids. It forms various chemical compounds where
protactinium is usually present in the oxidation state +5, but can also assume +4 and even +2 or +3 states. The
average concentrations of protactinium in the Earth's crust is typically on the order of a few parts per trillion, but
may reach up to a few parts per million in some uraninite ore deposits. Because of its scarcity, high radioactivity and
high toxicity, there are currently no uses for protactinium outside of scientific research, and for this purpose,
protactinium is mostly extracted from spent nuclear fuel.
Protactinium was first identified in 1913 by Kasimir Fajans and Oswald Helmuth Göhring and named brevium
because of the short half-life of the specific isotope studied, namely protactinium-234. A more stable isotope
231
( Pa) of protactinium was discovered in 1917/18 by Otto Hahn and Lise Meitner, and they chose the name
proto-actinium, but then the IUPAC named it finally protactinium in 1949 and confirmed Hahn and Meitner as
discoverers. The new name meant "parent of actinium" and reflected the fact that actinium is a product of
radioactive decay of protactinium.
The longest-lived and most abundant (nearly 100%) naturally occurring isotope of protactinium, protactinium-
231, has a half-life of 32,760 years and is a decay product of uranium-235. Much smaller trace amounts of the
short-lived nuclear isomer protactinium-234m occur in the decay chain of uranium-238. Protactinium-233 results
from the decay of thorium-233 as part of the chain of events used to produce uranium-233 by neutron
irradiation of thorium-232. It is an undesired intermediate product in thorium-based nuclear reactors and is
therefore removed from the active zone of the reactor during the breeding process. Analysis of the relative
concentrations of various uranium, thorium and protactinium isotopes in water and minerals is used in
radiometric dating of sediments which are up to 175,000 years old and in modeling of various geological
processes.

History
In 1871, Dmitri Mendeleev predicted the existence of an element between thorium and uranium. The actinide
element group was unknown at the time. Therefore, uranium was positioned below tungsten, and thorium below
zirconium, leaving the space below tantalum empty and, until the 1950s, periodic tables were published with this
structure. For a long time chemists searched for eka-tantalum as an element with similar chemical properties to
tantalum, making a discovery of protactinium nearly impossible.
In 1900, William Crookes isolated protactinium as an intensely radioactive material from uranium; however, he
[3]
could not characterize it as a new chemical element and thus named it uranium-X (UX). Crookes dissolved
uranium nitrate in ether, the residual aqueous phase contains most of the 234
90Th and 234
91Pa. His method was still used in the 1950s to isolate 234
90Th and 234
91Pa from uranium compounds. Protactinium was first identified in 1913, when Kasimir Fajans and Oswald Helmuth
234
Göhring encountered the isotope Pa during their studies of the decay chains of uranium-238: 238
92U →
234
90Th →
234
91Pa → 234
92U. They named the new element brevium (from the Latin word, brevis, meaning brief or short) because of its short
half-life, 6.7 hours for 234
[4][5]
91Pa. In 1917/18, two groups of scientists, Otto Hahn and Lise Meitner of Germany and Frederick Soddy
and
231
John Cranston of Great Britain, independently discovered another isotope of protactinium, Pa having much longer
half-life of about 32,000 years. Thus the name brevium was changed to protoactinium as the new element was part of
the decay chain of uranium-235 before the actinium (from Greek: πρῶτος = protos meaning first, before). For ease
[6]
of pronunciation, the name was shortened to protactinium by the IUPAC in 1949. The discovery of protactinium
completed the last gap in the early versions of the periodic table, proposed by Mendeleev in 1869, and it brought to
[7]
fame the involved scientists.
Aristid von Grosse produced 2 milligrams of Pa in 1927, and in 1934 first isolated elemental protactinium from
O
2 5
0.1 milligrams of Pa O . He used two different procedures: in the first one, protactinium oxide was irradiated by
2 5
35 keV electrons in vacuum. In another method, called the van Arkel–de Boer process, the oxide was chemically
converted to a halide (chloride, bromide or iodide) and then reduced in a vacuum with an electrically heated
metallic filament:
2 PaI → 2 Pa + 5 I
5 2
In 1961, the United Kingdom Atomic Energy Authority (UKAEA) produced 125 grams of 99.9% pure
protactinium by processing 60 tonnes of waste material in a 12-stage process, at a cost of about 500,000 USD.
For many years, this was the world's only significant supply of protactinium, which was provided to various
laboratories for scientific studies. Oak Ridge National Laboratory in the US is currently providing
protactinium at a cost of about
[8]
280 USD/gram.

Occurrence
Protactinium is one of the rarest and
most expensive naturally occurring
elements. It is found in the form of
231 234
two isotopes – Pa and Pa,
234
with the isotope Pa occurring in
two different energy states. Nearly
all natural protactinium is
protactinium-231. It is an alpha
emitter and is formed by the decay of
Mendeleev's 1869 periodic table with a gap for protactinium on the bottom row of
uranium-235, whereas the beta
the chart, between thorium and uranium
radiating protactinium-234 is
produced
234m [9]
as a result of uranium-238 decay. Nearly all uranium-238 (99.8%) decays first to the Pa isomer.
231
Protactinium occurs in uraninite (pitchblende) at concentrations of about 0.3-3 parts Pa per million parts (ppm) of
ore. Whereas the usual content is closer to 0.3 ppm (e.g. in Jáchymov, Czech Republic), some ores from the
Democratic Republic of the Congo have about 3 ppm. Protactinium is homogeneously dispersed in most natural
materials and in water, but at much lower concentrations on the order of one part per trillion, that corresponds to the
radioactivity of 0.1 picocuries (pCi)/g. There is about 500 times more protactinium in sandy soil particles than in
water, even the water present in the same sample of soil. Much higher ratios of 2,000 and above are measured in
[10]
loam soils and clays, such as bentonite.
In nuclear reactors
231 233
Two major protactinium isotopes, Pa and Pa, are produced from thorium in nuclear reactors; both are
undesirable and are usually removed, thereby adding complexity to the reactor design and operation. In particular,
232 231 231
Th via (n,2n) reactions produces Th which quickly (half-life 25.5 hours) decays to Pa. The last isotope,
while not a transuranic waste, has a long half-life of 32,760 years and is a major contributor to the long term
radiotoxicity of spent nuclear fuel.
232
Protactinium-233 is formed upon neutron capture by Th. It further either decays to uranium-233 or captures
233
another neutron and converts into the non-fissile uranium-234. Pa has a relatively long half-life of 27 days and
high cross section for neutron capture (the so-called "neutron poison"). Thus instead of rapidly decaying to the useful
233 233
U, a significant fraction of Pa converts to non-fissile isotopes and consumes neutrons, degrading the reactor
233
efficiency. To avoid this, Pa is extracted from the active zone of thorium molten salt reactors, during their
233
operation, so that it only decays to U. This is achieved using several meters tall columns of molten bismuth with
lithium dissolved in it. In a simplified scenario, lithium selectively reduces protactinium salts to protactinium metal
which is then extracted from the molten-salt cycle, and bismuth is merely a carrier. It is chosen because of its low
melting point (271 °C), low vapor pressure, good solubility for lithium and actinides, and immiscibility with molten
[7]
halides.

Preparation
Before the advent of nuclear reactors, protactinium was separated for scientific
experiments from uranium ores. Nowadays, it is mostly produced as an
intermediate product of nuclear fission in thorium high-temperature reactors:

Protactinium occurs in uraninite ores.

The times are half-lives.

Protactinium metal can be prepared by reduction of its fluoride with calcium fluoride, lithium or barium at a
temperature of 1300–1400 °C.
Physical and chemical
properties
Protactinium is an actinide which is positioned in the periodic table to the left of uranium and to the right of
thorium, and many of its physical properties are intermediate between those two actinides. So, protactinium is
more dense and rigid than thorium but is lighter than uranium, and its melting point is lower than that of thorium
and higher than that of uranium. The thermal expansion, electrical and thermal conductivities of these three
elements are comparable and are typical of "poor metals". The estimated shear modulus of protactinium is
[11]
similar to that of titanium. Protactinium is a metal with silvery-gray luster that is preserved for some time in
air. Protactinium easily reacts with oxygen, water vapor and acids, but not with alkali metals.
At room temperature, protactinium crystallizes in body-centered tetragonal structure which can be regarded as
distorted body-centered cubic lattice; this structure does not change upon compression up to 53 GPa. The structure
[12]
changes to face-centered cubic (fcc) upon cooling from high temperature, at about 1200 °C. The thermal
−6
expansion coefficient of the tetragonal phase between room temperature and 700 °C is 9.9×10 /°C.
[13]
Protactinium is paramagnetic and no magnetic transitions are known for it at any temperature. It becomes
superconductive at temperatures below 1.4 K. Protactinium tetrachloride is paramagnetic at room temperature but
turns ferromagnetic upon cooling to 182 K.
Protactinium exists in two major oxidation states, +4 and +5, both in solids and solutions, and the +3 and +2 states
2 1 2
were observed in some solid phases. As the electron configuration of the neutral atom is [Rn]7s 6d 5f , the +5
0
oxidation state corresponds to the low-energy (and thus favored) 5f configuration. Both +4 and +5 states easily form
3+
hydroxides in water with the predominant ions being Pa(OH) , Pa(OH)2+
2, Pa(OH)+
[14]
3 and Pa(OH) , all colorless. Other known protactinium ions include PaCl2+
4
2, PaSO2+
3+
4, PaF , PaF2+
2, PaF−

6, PaF2−

7 and PaF3−
[15]
8.

Chemical compounds
Formula color symmetry space group No Pearson symbol a (pm) b (pm) c (pm) Z 3
density, g/cm

Pa silvery-gray tetragonal I4/mmm 139 tI2 392.5 392.5 323.8 2 15.37

PaO rocksalt Fm3m 225 cF8 496.1 4 13.44

PaO black fcc Fm3m 225 cF12 550.5 4 10.47


2

Pa O white Fm3m 225 cF16 547.6 547.6 547.6 4 10.96


2 5

Pa O white orthorhombic 692 402 418


2 5

PaH black cubic Pm3n 223 cP32 664.8 664.8 664.8 8 10.58
3

PaF brown-red monoclinic C2/c 15 mS60 2


4

PaCl green-yellow tetragonal I4 /amd 141 tI20 837.7 837.7 748.1 4 4.72
4 1

PaBr brown tetragonal I4 /amd 141 tI20 882.4 882.4 795.7


4 1

PaCl yellow monoclinic C2/c 15 mS24 797 1135 836 4 3.74


5

PaBr red monoclinic P2 /c 14 mP24 838.5 1120.5 1214.6 4 4.98


5 1

PaOBr monoclinic C2 1691.1 387.1 933.4


3
Pa(PO ) orthorhombic 696.9 895.9 1500.9
3 4

Pa P O cubic Pa3 865 865 865


2 2 7

Pa(C H ) golden-yellow monoclinic 709 875 1062


8 8 2

Here a, b and c are lattice constants in picometers, No is space group number and Z is the number of formula units
per unit cell; fcc stands for the face-centered cubic symmetry. Density was not measured directly but calculated from
the lattice parameters.

Oxides and oxygen-containing salts


Protactinium oxides are known for the metal oxidation states +2, +4 and +5. The most stable is white pentoxide
[16]
Pa O , which can be produced by igniting protactinium(V) hydroxide in air at a temperature of 500 °C. Its
crystal
2 5
structure is cubic, and the chemical composition is often non-stoichiometric, described as PaO . Another phase of
2.25
this oxide with orthorhombic symmetry has also been reported. The black dioxide PaO is obtained from the
2
pentoxide by reducing it at 1550 °C with hydrogen. It is not readily soluble in either dilute or concentrated nitric,
hydrochloric or sulfuric acids, but easily dissolves in hydrofluoric acid. The dioxide can be converted back to
pentoxide by heating in oxygen-containing atmosphere to 1100 °C. The monoxide PaO has only been observed as a
thin coating on protactinium metal, but not in an isolated bulk form.
Protactinium forms mixed binary oxides with various metals. With alkali metals A, the crystals have a chemical
formula APaO and perovskite structure, or A PaO and distorted rock-salt structure, or A PaO where oxygen
3 3 4 7 6
[17]
atoms for a hexagonal close-packed lattice. In all these materials, protactinium ions are octahedrally coordinated.
The pentoxide Pa O combines with rare-earth metal oxides R O to form various nonstoichiometric mixed-oxides,
2 5 2 3
also of perovskite structure.
Protactinium oxides are basic; they easily convert to hydroxides and can form various salts, such as sulfates,
phosphates, nitrates, etc. The nitrate is usually white but can be brown due to radiolytic decomposition. Heating the
nitrate in air at 400 °C converts it to the white protactinium pentoxide. The polytrioxophosphate Pa(PO ) can be
34
produced by reacting difluoride sulfate PaF SO with phosphoric acid (H PO ) under inert gas atmosphere. Heating
2 4 3 4
the product to about 900 °C eliminates the reaction by-products such as hydrofluoric acid, sulfur trioxide and
phosphoric anhydride. Heating to higher temperatures in an inert atmosphere decomposes Pa(PO ) into the
34
diphosphate PaP O , which is analogous to diphosphates of other actinides. In the diphosphate, the PO groups form
2 7 3
pyramids of symmetry. Heating PaP O in air to 1400 °C decomposes it into the pentoxides of phosphorus and
C
2v 2 7
protactinium.

Halides
Protactinium(V) fluoride forms white crystals where protactinium ions are arranged in pentagonal bipyramids and
coordinated by 7 other ions. The coordination is the same in protactinium(V) chloride, but the color is yellow. The
coordination changes to octahedral in the brown protactinium(V) bromide and is unknown for protactinium(V)
iodide. The protactinium coordination in all its tetrahalides is 8, but the arrangement is square antiprismatic in
protactinium(IV) fluoride and dodecahedral in the chloride and bromide. Brown-colored protactinium(III) iodide has
[18]
been reported where protactinium ions are 8-coordinated in a bicapped trigonal prismatic arrangement.
Protactinium(V) fluoride and protactinium(V) chloride have a polymeric
structure of monoclinic symmetry. There, within one polymeric chain,
all the halide atoms lie in one graphite-like plane and form planar
pentagons around the protactinium ions. The coordination 7 of
protactinium originates from the 5 halide atoms and two bonds to
Coordination of protactinium (solid
protactinium atoms belonging to the nearby chains. These
circles) and halogen atoms (open circles)
compounds easily hydrolyze in water. The pentachloride melts at 300 in protactinium(V) fluoride or chloride.
°C and sublimates at even lower temperatures.

Protactinium(V) fluoride can be prepared by reacting protactinium oxide with either bromine pentafluoride or
bromine trifluoride at about 600 °C, and protactinium(IV) fluoride is obtained from the oxide and a mixture of
hydrogen and hydrogen fluoride at 600 °C; a large excess of hydrogen is required to remove atmospheric
oxygen leaks into the reaction.
Protactinium(V) chloride is prepared by reacting protactinium oxide with carbon tetrachloride at temperature of
200–300 °C. The by-products (such as PaOCl ) are removed by fractional sublimation. Reduction of
3
protactinium(V) chloride with hydrogen at about 800 °C yields protactinium(IV) chloride – a yellow-green solid
which sublimes in vacuum at 400 °C; it can also be obtained directly from protactinium dioxide by treating it with
carbon tetrachloride at 400 °C.
Protactinium bromides are produced by the action of aluminum bromide, hydrogen bromide, carbon
tetrabromide or a mixture of hydrogen bromide and thionyl bromide on protactinium oxide. An alternative
reaction is between protactinium pentachloride and hydrogen bromide or thionyl bromide. Protactinium(V)
bromide has two similar monoclinic forms, one is obtained by sublimation at 400–410 °C and another by
sublimation at slightly lower temperature of 390–400 °C.
Protactinium iodides result from the oxides and aluminum iodide or ammonium iodide heated to 600 °C.
Protactinium(III) iodide was obtained by heating protactinium(V) iodide in vacuum. As with oxides, protactinium
forms mixed halides with alkali metals. Among those, most remarkable is Na PaF where protactinium ion is
3 8

symmetrically surrounded by 8 F ions which form a nearly perfect cube.
[]
More complex protactinium fluorides are also known such as Pa F and ternary fluorides of the types MPaF (M =
2 9 6
Li, Na, K, Rb, Cs or NH ), M PaF (M = K, Rb, Cs or NH ) and M PaF (M = Li, Na, Rb, Cs), all being white
4 2 7 4 3 8
crystalline solids. The MPaF formula can be represented as a combination of MF and PaF . These compounds can
6 5
be obtained by evaporating a hydrofluoric acid solution containing these both complexes. For the small alkali cations
+ + +
like Na, the crystal structure is tetragonal, whereas it lowers to orthorphombic for larger cations K , Rb , Cs or
+
NH . A similar variation was observed for the M PaF fluorides, namely the crystal symmetry was dependent on
4 2 7
the cation and differed for Cs PaF and M PaF (M = K, Rb or NH ).
2 7 2 7 4

Other inorganic compounds


Oxyhalides and oxysulfides of protactinium are known. PaOBr
3 has a monoclinic structure composed of
double-chain units where protactinium has coordination 7 and is arranged into pentagonal bipyramids. The chains
are interconnected through oxygen and bromine atoms, and each oxygen atom is related to three protactinium
atoms. PaOS is a light-yellow non-volatile solid with a cubic crystal lattice isostructural to that of other actinide
oxysulfides. It is obtained by reacting protactinium(V) chloride with a mixture of hydrogen sulfide and carbon
disulfide at 900
°C.
In hydrides and nitrides, protactinium has a low oxidation state of about +3. The hydride is obtained by direct action
of hydrogen on the metal at 250 °C, and the nitride is a product of ammonia and protactinium tetrachloride or
pentachloride. This bright yellow solid is stable to heating to 800 °C in vacuum. Protactinium carbide PaC is formed
by reduction of protactinium tetrafluoride with barium in a carbon crucible at a temperature of about 1400 °C.
Protactinium forms borohydrides which include Pa(BH ) . It has an unusual polymeric structure with helical chains
44 – [19]
where the protactinium atom has coordination number of 12 and is surrounded by six BH ions.
4

Organometallic compounds
Protactinium(IV) forms a tetrahedral complex
tetrakis(cyclopentadienyl)protactinium(IV) (or Pa(C H ) ) with four
5 5 4
cyclopentadienyl rings, which can be synthesized by reacting protactinium(IV)
chloride with molten Be(C H ) . One ring can be substituted with a halide
5 5 2
[20]
atom. Another organometallic complex is golden-yellow
bis(π-cyclooctatetraene) protactinium, Pa(C H ) , which is analogous in
8 8 2
structure to uranocene. There, the metal atom is sandwiched between two
cyclooctatetraene ligands. Similar to uranocene, it can be prepared by
reacting protactinium tetrachloride with dipotassium cyclooctatetraenide, K
C H , in
2 8 8
tetrahydrofuran. The proposed structure of
the Pa(C H ) molecule
8 8 2

Isotopes
Main article: Isotopes of
protactinium
231
Twenty-nine radioisotopes of protactinium have been discovered, the most stable being Pa with a half-life of
233 230
32,760 years, Pa with a half-life of 27 days, and Pa with a half-life of 17.4 days. All of the remaining isotopes
have half-lives shorter than 1.6 days, and the majority of these have half-lives less than 1.8 seconds. Protactinium
217m 234m
also has two nuclear isomers, Pa (half-life 1.2 milliseconds) and Pa (half-life 1.17 minutes).
231
The primary decay mode for isotopes of protactinium lighter than (and including) the most stable isotope Pa (i.e.,
212 231 232 240
Pa to Pa) is alpha decay and the primary mode for the heavier isotopes (i.e., Pa to Pa) is beta decay. The
231
primary decay products of isotopes of protactinium lighter than (and including) Pa are actinium isotopes and the
primary decay products for the heavier isotopes of protactinium are uranium isotopes.

Applications
Although protactinium is located in the periodic table between uranium and thorium, which both have numerous
applications, owing to its scarcity, high radioactivity and high toxicity, there are currently no uses for protactinium
outside of scientific research.
Protactinium-231 arises from the decay of uranium-235 formed in nuclear reactors, and by the reaction
232 231
Th + n → Th + 2n and subsequent beta decay. It may support a nuclear chain reaction, which could in principle
be used to build nuclear weapons. The physicist Walter Seifritz once estimated the associated critical mass as
[21] [22]
750±180 kg, but this possibility (of a chain reaction) has been ruled out by other nuclear physicists since then.
231
With the advent of highly sensitive mass spectrometers, an application of Pa as a tracer in geology and
paleoceanography has become possible. So, the ratio of protactinium-231 to thorium-230 is used for radiometric
dating of sediments which are up to 175,000 years old and in modeling of the formation of minerals. In particular, its
evaluation in oceanic sediments allowed to reconstruct the movements of North Atlantic water bodies during the last
melting of Ice Age glaciers. Some of the protactinium-related dating variations rely on the analysis of the relative
concentrations for several long-living members of the uranium decay chain – uranium, thorium and protactinium, for
example. These elements have 6, 5 and 4 f-electrons in the outer shell and thus favor +6, +5 and +4 oxidation states,
respectively, and show different physical and chemical properties. So, thorium and protactinium, but not uranium
compounds are poorly soluble in aqueous solutions, and precipitate into sediments; the precipitation rate is faster for
thorium than for protactinium. Besides, the concentration analysis for both protactinium-231 (half-life 32,750 years)
and thorium-230 (half-life 75,380 years) allows to improve the accuracy compared to when only one isotope is
measured; this double-isotope method is also weakly sensitive to inhomogeneities in the spatial distribution of the
[2]
isotopes and to variations in their precipitation rate.

Precautions
Protactinium is both toxic and highly radioactive and thus all manipulations with it are performed in a sealed
231
glove box. Its major isotope Pa has a specific activity of 0.048 curies (1.8 GBq) per gram and primarily
emits alpha-particles with an energy of 5 MeV, which can be stopped by a thin layer of any material. However, it
227
slowly decays, with a half-life of 32,760 years, into Ac, which has a specific activity of 74 curies (2,700 GBq)
227
per gram, emits both alpha and beta radiation, and has a much shorter half-life of 22 years. Ac, in turn,
decays into lighter isotopes with even shorter half-lives and much greater specific activities (SA), as
summarized in the table below showing the decay chain of protactinium-231.

Isotope 231 227 227 223 219 215 211 211 207
Pa Ac Th Ra Rn Po Pb Bi Tl
SA (Ci/g) 0.048 73 4 4 10 13 7 8 8
3.1×10 5.2×10 1.3×10 3×10 2.5×10 4.2×10 1.9×10

Decay α α, β α α α α β α, β β

Half-life 33 ka 22 a 19 days 11 days 4s 1.8 ms 36 min 2.1 min 4.8 min

As protactinium is present in small amounts in most natural products and materials, it is ingested with food or
water and inhaled with air. Only about 0.05% of ingested protactinium is absorbed into the blood and the
remainder is excreted. From the blood, about 40% of the protactinium deposits in the bones, about 15% goes to
the liver, 2% to the kidneys, and the rest leaves the body. The biological half-life of protactinium is about 50
years in the bones, whereas in other organs the kinetics has a fast and slow component. So in the liver 70% of
protactinium have a half-life of 10 days and 30% remain for 60 days. The corresponding values for kidneys are
20% (10 days) and 80% (60 days). In all these organs, protactinium promotes cancer via its radioactivity. The
maximum safe dose of Pa in the human body is 0.03 µCi (1.1 kBq), which corresponds to 0.5 micrograms of
231 8
Pa. This isotope is 2.5×10 times more toxic than hydrocyanic acid. The maximum allowed
231 −4 3
concentrationsWikipedia:WikiProject Countering systemic bias of Pa in the air is 3×10 Bq/m .

References
[1] Magnetic susceptibility of the elements and inorganic compounds (http://www-d0.fnal. gov/ hardware/cal/ lvps_info/engineering/
elementmagn. pdf), in Handbook of Chemistry and Physics 81st edition, CRC press.
[2] http://en.wikipedia.org/w/index.php?title=Template:Infobox_protactinium&action=edit
[3] http://books. google.co.uk/books?id=-zgrAAAAYAAJ&pg=PA180
[4] Greenwood, p. 1250
[5] Greenwood, p. 1254
[6] Greenwood, p. 1251
[7] Shea, William R. (1983) Otto Hahn and the rise of nuclear physics (http://books. google.com/books?id=W7xyvXc-hgEC&pg=PA213),
Springer, p. 213, ISBN 90-277-1584-X.
[8] Protactinium (http:/ /periodic.lanl.gov/91.shtml), Los Alamos Laboratory
[9] Protactinium (http://www.ead. anl. gov/ pub/doc/protactinium. pdf), Argonne National Laboratory, Human Health Fact Sheet, August
2005
[10] Cornelis, Rita (2005) Handbook of elemental speciation II: species in the environment, food, medicine & occupational health, Vol. 2
(http:// books. google. com/books?id=1PmjurlE6KkC&pg=PA520), John Wiley and Sons, pp. 520–521, ISBN 0-470-85598-3.
[11] Seitz, Frederick and Turnbull, David (1964) Solid state physics: advances in research and applications (http:// books. google. com/ books?
id=F9V3a-0V3r8C&pg=PA289), Academic Press, pp. 289–291, ISBN 0-12-607716-9.
[12] Young, David A. (1991) Phase diagrams of the elements (http:/ /books.google. com/ books?id=F2HVYh6wLBcC&pg=PA222),
University of California Press, p. 222, ISBN 0-520-07483-1.
[13] Buschow, K. H. J. (2005) Concise encyclopedia of magnetic and superconducting materials (http:/ /books.google. com/ books?
id=N9mvytGEBtwC&pg=PA129), Elsevier, pp. 129–130, ISBN 0-08-044586-1.
[14] Greenwood, p. 1265
[15] Greenwood, p. 1275
[16] Greenwood, p. 1268
[17] Greenwood, p. 1269
[18] Greenwood, p. 1270
[19] Greenwood, p. 1277
[20] Greenwood, pp. 1278–1279
[21] Seifritz, Walter Nukleare Sprengkörper – Bedrohung oder Energieversorgung für die Menschheit, Thiemig-Verlag (1984), ISBN
3-521-06143-4.
231
[22] Ganesan, S. et al. A Re-calculation of Criticality Property of Pa Using New Nuclear Data (http:/ /www.iisc.ernet. in/currsci/sept10/
researcharticle. pdf), Current Science, 1999, 77 (5) 667–677.

Bibliography
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth–Heinemann.
ISBN 0080379419.

External links
• Protactinium (http://www.periodicvideos.com/videos/091.htm) at The Periodic Table of Videos
(University of Nottingham)
92 Uranium 1

92 – Uranium
This article is about the chemical element. For other uses, see Uranium (disambiguation).

Uranium
U
92

protactinium ← uranium → neptuniumNd



U

(Uqq)

Uranium in the periodic table


Appearance

silvery gray metallic; corrodes to a spalling black oxide coat in air

General properties

Name, symbol, number uranium, U, 92

Pronunciation /jʊˈreɪniəm/
ew-RAY-nee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight 238.02891(3)

Electron configuration 3 1 2
[Rn] 5f 6d 7s
2, 8, 18, 32, 21, 9, 2

Physical properties

Phase solid

Density (near r.t.) −3


19.1 g·cm
Liquid density at m.p. −3
17.3 g·cm

Melting point 1405.3 K, 1132.2 °C, 2070 °F

Boiling point 4404 K, 4131 °C, 7468 °F

Heat of fusion −1
9.14 kJ·mol

Heat of vaporization −1
417.1 kJ·mol

Molar heat capacity −1 −1


27.665 J·mol ·K

Vapor pressure

P (Pa) 1 10 100 1k 10 k 100 k

at T (K) 2325 2564 2859 3234 3727 4402

Atomic properties

Oxidation states 6, 5, 4, 3, 2, 1
(weakly basic oxide)

Electronegativity 1.38 (Pauling scale)

Ionization energies −1
1st: 597.6 kJ·mol

−1
2nd: 1420 kJ·mol

Atomic radius 156 pm

Covalent radius 196±7 pm

Van der Waals radius 186 pm

Miscellanea

Crystal structure orthorhombic

Magnetic ordering paramagnetic

Electrical resistivity (0 °C) 0.280 µΩ·m

Thermal conductivity −1 −1
27.5 W·m ·K

Thermal expansion −1 −1
(25 °C) 13.9 µm·m ·K

Speed of sound (thin rod) −1


(20 °C) 3155 m·s

Young's modulus 208 GPa

Shear modulus 111 GPa

Bulk modulus 100 GPa

Poisson ratio 0.23

CAS registry number 7440-61-1

History

Naming after planet Uranus, itself named after Greek god of the sky Uranus

Discovery Martin Heinrich Klaproth (1789)

First isolation Eugène-Melchior Péligot (1841)


Most stable isotopes

Main article: Isotopes of uranium

iso NA half-life DM DE (MeV) DP

232 trace 68.9 y SF - -


U
α 5.414 228
Th

233 trace 5 SF 197.93 -


U 1.592×10 y
α 4.909 229
Th

234 0.005% 5 SF 197.78 -


U 2.455×10 y
α 4.859 230
Th

235 0.720% 8 SF 202.48 -


U 7.04×10 y
α 4.679 231
Th

236 trace 7 SF 201.82 -


U 2.342×10 y
α 4.572 232
Th

238 99.274% 9 α 4.270 234


U 4.468×10 y Th
SF 205.87 -

− − - 238
β β Pu

• v
• t
[1]
• e

Uranium is a silvery-white metallic chemical element in the actinide series of the periodic table, with symbol U and
atomic number 92. A uranium atom has 92 protons and 92 electrons, of which 6 are valence electrons. Uranium is
weakly radioactive because all its isotopes are unstable (with half-lives of the 6 naturally known isotopes, U-233 -
U-238, varying between 69 years and 4½ billion years). The most common isotopes of uranium are uranium-238
(which has 146 neutrons and accounts for almost 99.3% of the uranium found in nature) and uranium-235 (which has
143 neutrons, accounting for 0.7% of the element found naturally). Uranium has the second highest atomic weight of
the primordially occurring elements, lighter only than plutonium. Its density is about 70% higher than that of lead,
but not as dense as gold or tungsten. It occurs naturally in low concentrations of a few parts per million in soil, rock
and water, and is commercially extracted from uranium-bearing minerals such as uraninite.
In nature, uranium is found as uranium-238 (99.2739–99.2752%), uranium-235 (0.7198–0.7202%), and a very small
amount of uranium-234 (0.0050–0.0059%). Uranium decays slowly by emitting an alpha particle. The half-life of
uranium-238 is about 4.47 billion years and that of uranium-235 is 704 million years, making them useful in dating
the age of the Earth.
Many contemporary uses of uranium exploit its unique nuclear properties. Uranium-235 has the distinction of being
the only naturally occurring fissile isotope. Uranium-238 is fissionable by fast neutrons, and is fertile, meaning it can
be transmuted to fissile plutonium-239 in a nuclear reactor. Another fissile isotope, uranium-233, can be produced
from natural thorium and is also important in nuclear technology. While uranium-238 has a small probability for
spontaneous fission or even induced fission with fast neutrons, uranium-235 and to a lesser degree uranium-233 have
a much higher fission cross-section for slow neutrons. In sufficient concentration, these isotopes maintain a sustained
nuclear chain reaction. This generates the heat in nuclear power reactors, and produces the fissile material for
238
nuclear weapons. Depleted uranium ( U) is used in kinetic energy penetrators and armor plating.
Uranium is used as a colorant in uranium glass, producing orange-red to lemon yellow hues. It was also used for
tinting and shading in early photography. The 1789 discovery of uranium in the mineral pitchblende is credited to
Martin Heinrich Klaproth, who named the new element after the planet Uranus. Eugène-Melchior Péligot was the
first person to isolate the metal and its radioactive properties were discovered in 1896 by Henri Becquerel. Research
by Enrico Fermi and others starting in 1934 led to its use as a fuel in the nuclear power industry and in Little Boy, the
first nuclear weapon used in war. An ensuing arms race during the Cold War between the United States and the
Soviet Union produced tens of thousands of nuclear weapons that used uranium metal and uranium-derived
plutonium-239. The security of those weapons and their fissile material following the breakup of the Soviet Union in
[2]
1991 is an ongoing concern for public health and safety. See Nuclear proliferation.

Characteristics
When refined, uranium is a silvery white, weakly radioactive metal, which is
harder than most elements. It is malleable, ductile, slightly paramagnetic,
strongly electropositive and is a poor electrical conductor. Uranium metal has
very high density, being approximately 70% denser than lead, but slightly less
dense than gold.
Uranium metal reacts with almost all non-metal elements(with an exception
of the group 18 elements) and their compounds, with reactivity increasing
with temperature. Hydrochloric and nitric acids dissolve uranium, but non-
oxidizing acids other than hydrochloric acid attack the element very slowly.
When finely divided, it can react with cold water; in air, uranium metal
becomes coated with a dark layer of uranium oxide. Uranium in ores is
extracted chemically and converted into uranium dioxide or other chemical
forms usable in industry.

Uranium-235 was the first isotope that was found to be fissile. Other
A neutron-induced nuclear fission event
naturally occurring isotopes are fissionable, but not fissile. Upon
involving uranium-235
bombardment with slow neutrons, its uranium-235 isotope will most of the
time divide into two smaller nuclei, releasing nuclear binding energy and
more neutrons. If too
many of these neutrons are absorbed by other uranium-235 nuclei, a nuclear chain reaction occurs that results in a
burst of heat or (in special circumstances) an explosion. In a nuclear reactor, such a chain reaction is slowed and
controlled by a neutron poison, absorbing some of the free neutrons. Such neutron absorbent materials are often part
of reactor control rods (see nuclear reactor physics for a description of this process of reactor control).

As little as 15 lb (7 kg) of uranium-235 can be used to make an atomic bomb. The first nuclear bomb used in war,
Little Boy, relied on uranium fission, while the very first nuclear explosive (The gadget) and the bomb that destroyed
Nagasaki (Fat Man) were plutonium bombs.
Uranium metal has three allotropic
forms:
• α (orthorhombic) stable up to 660
°C
• β (tetragonal) stable from 660 °C to 760 °C
• γ (body-centered cubic) from 760 °C to melting point—this is the most malleable and ductile state.
Applications

Military
The major application of uranium in the military sector is in high-
density penetrators. This ammunition consists of depleted uranium
(DU) alloyed with 1–2% other elements. At high impact speed,
the density, hardness, and pyrophoricity of the projectile enable
destruction of heavily armored targets. Tank armor and other
removable vehicle armor are also hardened with depleted uranium
plates. The use of depleted uranium became politically and
environmentally contentious after the use of depleted uranium
munitions by the US, UK and other countries during wars in the
Depleted uranium is used by various militaries as Persian Gulf and the Balkans raised questions of uranium
high-density penetrators. compounds left in the soil (see Gulf War Syndrome).

Depleted uranium is also used as a shielding material in some


containers used to store and transport radioactive materials. While the metal itself is radioactive, its high density
makes it more effective than lead in halting radiation from strong sources such as radium. Other uses of depleted
uranium include counterweights for aircraft control surfaces, as ballast for missile re-entry vehicles and as a
shielding material. Due to its high density, this material is found in inertial guidance systems and in gyroscopic
compasses. Depleted uranium is preferred over similarly dense metals due to its ability to be easily machined and
cast as well as its relatively low cost. The main risk of exposure to depleted uranium is chemical poisoning by
uranium oxide rather than radioactivity (uranium being only a weak alpha emitter).

During the later stages of World War II, the entire Cold War, and to a lesser extent afterwards, uranium-235 has been
used as the fissile explosive material to produce nuclear weapons. Initially, two major types of fission bombs were
built: a relatively simple device that uses uranium-235 and a more complicated mechanism that uses plutonium-239
derived from uranium-238. Later, a much more complicated and far more powerful type of fission/fusion bomb
(thermonuclear weapon) was built, that uses a plutonium-based device to cause a mixture of tritium and deuterium to
undergo nuclear fusion. Such bombs are jacketed in a non-fissile (unenriched) uranium case, and they derive more
than half their power from the fission of this material by fast neutrons from the nuclear fusion process.

Civilian
The main use of uranium in the civilian sector is to fuel nuclear power
plants. One kilogram of uranium-235 can theoretically produce about
13
20 terajoules of energy (2×10 joules), assuming complete fission; as
much energy as 1500 tonnes of coal.
Commercial nuclear power plants use fuel that is typically enriched to
around 3% uranium-235. The CANDU and Magnox designs are the
only commercial reactors capable of using unenriched uranium fuel.
Fuel used for United States Navy reactors is typically highly enriched
in uranium-235 (the exact values are classified). In a breeder reactor,
The most visible civilian use of uranium is as the
uranium-238 can also be converted into plutonium through the thermal power source used in nuclear power
238 239 239
following reaction: U (n, gamma) → U -(beta) → Np -(beta) plants.
239
→ Pu.
[citation needed]
Before (and, occasionally, after) the discovery of
radioactivity, uranium was primarily used in small amounts for yellow
glass and pottery glazes, such as uranium glass and in Fiestaware.
The discovery and isolation of radium in uranium ore (pitchblende)
by Marie Curie sparked the development of uranium mining to
extract the radium, which was used to make glow-in-the-dark paints
for clock and aircraft dials. This left a prodigious quantity of
uranium as a waste product, since it takes three tonnes of uranium
to extract one gram of radium. This waste product was diverted
Uranium glass glowing under UV
light to the glazing industry, making uranium glazes very inexpensive
and abundant. Besides the pottery glazes, uranium tile glazes
accounted for the bulk of the use,
including common bathroom and kitchen tiles which can be produced in green, yellow, mauve, black, blue, red and
other colors.

Uranium was also used in photographic chemicals (especially


[citation needed]
uranium nitrate as a toner), in lamp filaments, to
improve the appearance of dentures, and in the leather and wood
industries for stains and dyes. Uranium salts are mordants of silk
or wool. Uranyl acetate and uranyl formate are used as electron-
dense "stains" in transmission electron microscopy, to increase the
contrast of biological specimens in ultrathin sections and in
negative staining of viruses, isolated cell organelles and
macromolecules.
Uranium glass used as lead-in seals in a vacuum
The discovery of the radioactivity of uranium ushered in additional capacitor
scientific and practical uses of the element. The long half-life of the
9
isotope uranium-238 (4.51×10 years) makes it well-suited for use
in
estimating the age of the earliest igneous rocks and for other types of radiometric dating, including uranium-thorium
dating, uranium-lead dating and uranium-uranium dating. Uranium metal is used for X-ray targets in the making of
high-energy X-rays.

History

Prehistoric naturally occurring fission


Main article: Natural nuclear fission
reactor
In 1972 the French physicist Francis Perrin discovered fifteen ancient and no longer active natural nuclear fission
reactors in three separate ore deposits at the Oklo mine in Gabon, West Africa, collectively known as the Oklo Fossil
Reactors. The ore deposit is 1.7 billion years old; then, uranium-235 constituted about 3% of the total uranium on
Earth. This is high enough to permit a sustained nuclear fission chain reaction to occur, provided other supporting
conditions exist. The capacity of the surrounding sediment to contain the nuclear waste products has been cited by
the U.S. federal government as supporting evidence for the feasibility to store spent nuclear fuel at the Yucca
Mountain nuclear waste repository.
Pre-discovery use
The use of uranium in its natural oxide form dates back to at least the year 79 CE, when it was used to add a yellow
color to ceramic glazes. Yellow glass with 1% uranium oxide was found in a Roman villa on Cape Posillipo in the
Bay of Naples, Italy, by R. T. Gunther of the University of Oxford in 1912. Starting in the late Middle Ages,
pitchblende was extracted from the Habsburg silver mines in Joachimsthal, Bohemia (now Jáchymov in the Czech
Republic), and was used as a coloring agent in the local glassmaking industry. In the early 19th century, the world's
only known sources of uranium ore were these mines.

Discovery
The discovery of the element is credited to the German chemist Martin
Heinrich Klaproth. While he was working in his experimental
laboratory in Berlin in 1789, Klaproth was able to precipitate a yellow
compound (likely sodium diuranate) by dissolving pitchblende in nitric
acid and neutralizing the solution with sodium hydroxide. Klaproth
assumed the yellow substance was the oxide of a yet-undiscovered
element and heated it with charcoal to obtain a black powder, which he
thought was the newly discovered metal itself (in fact, that powder was
an oxide of uranium). He named the newly discovered element after
the planet Uranus, (named after the primordial Greek god of the sky),
Antoine Henri Becquerel discovered the
which had been discovered eight years earlier by William Herschel. phenomenon of radioactivity by exposing a
photographic plate to uranium in 1896.
In 1841, Eugène-Melchior Péligot, Professor of Analytical Chemistry
at the Conservatoire National des Arts et Métiers (Central School of
Arts and Manufactures) in Paris, isolated the first sample of uranium metal by heating uranium tetrachloride with
potassium. Uranium was not seen as being particularly dangerous during much of the 19th century, leading to the
development of various uses for the element. One such use for the oxide was the aforementioned but no longer
secret coloring of pottery and glass.

Henri Becquerel discovered radioactivity by using uranium in 1896. Becquerel made the discovery in Paris by
leaving a sample of a uranium salt, K UO (SO ) (potassium uranyl sulfate), on top of an unexposed photographic
2 2 42
plate in a drawer and noting that the plate had become "fogged". He determined that a form of invisible light or rays
emitted by uranium had exposed the plate.

Fission research
A team led by Enrico Fermi in 1934 observed that bombarding
uranium with neutrons produces the emission of beta rays (electrons or
positrons from the elements produced; see beta particle). The fission
products were at first mistaken for new elements of atomic numbers 93
and 94, which the Dean of the Faculty of Rome, Orso Mario Corbino,
christened ausonium and hesperium, respectively. The experiments
leading to the discovery of uranium's ability to fission (break apart)
into lighter elements and release binding energy were conducted by
Otto Hahn and Fritz Strassmann in Hahn's laboratory in Berlin. Lise
Cubes and cuboids of uranium produced during
Meitner and her nephew, the physicist Otto Robert Frisch, published
the Manhattan project
the physical explanation in February 1939 and named the process
"nuclear fission". Soon after, Fermi hypothesized that the fission of uranium might release enough neutrons to
sustain a fission reaction. Confirmation of this hypothesis came in 1939, and later work found that on average about
2.5 neutrons are released by each fission of the rare uranium isotope uranium-235. Further work found that the far
more common uranium-238 isotope can be transmuted into plutonium, which, like uranium-235, is also fissionable
by thermal neutrons. These discoveries led numerous countries to begin working on the development of nuclear
weapons and nuclear power.
On 2 December 1942, as part of the Manhattan Project, another team led by Enrico Fermi was able to initiate the first
artificial self-sustained nuclear chain reaction, Chicago Pile-1. Working in a lab below the stands of Stagg Field at
the University of Chicago, the team created the conditions needed for such a reaction by piling together 400 short
tons (360 metric tons) of graphite, 58 short tons (53 metric tons) of uranium oxide, and six short tons (5.5 metric
tons) of uranium metal, a majority of which was supplied by Westinghouse Lamp Plant in a makeshift production
process.

Nuclear weaponry
Two major types of atomic bombs were developed by the United States
during World War II: a uranium-based device (codenamed "Little
Boy") whose fissile material was highly enriched uranium, and a
plutonium-based device (see Trinity test and "Fat Man") whose
plutonium was derived from uranium-238. The uranium-based Little
Boy device became the first nuclear weapon used in war when it was
detonated over the Japanese city of Hiroshima on 6 August 1945.
Exploding with a yield equivalent to 12,500 tonnes of TNT, the blast
and thermal wave of the bomb destroyed nearly 50,000 buildings and
killed approximately 75,000 people (see Atomic bombings of
Hiroshima and Nagasaki). Initially it was believed that uranium was
relatively rare, and that nuclear proliferation could be avoided by
simply buying up all known uranium stocks, but within a decade large
[3]
deposits of it were discovered in many places around the world. The mushroom cloud over Hiroshima after the
dropping of the uranium-based atomic bomb
nicknamed 'Little Boy'

Reactors

The X-10 Graphite Reactor at Oak Ridge National Laboratory (ORNL)


in Oak Ridge, Tennessee, formerly known as the Clinton Pile and X-10
Pile, was the world's second artificial nuclear reactor (after Enrico
Fermi's Chicago Pile) and was the first reactor designed and built for
continuous operation. Argonne National Laboratory's Experimental
Breeder Reactor I, located at the Atomic Energy Commission's
National Reactor Testing Station near Arco, Idaho, became the first
nuclear reactor to create electricity on 20 December 1951. Initially,
four 150-watt light bulbs were lit by the reactor, but improvements
eventually enabled it to power the whole facility (later, the town of Four light bulbs lit with electricity generated
from the first artificial electricity-producing
Arco became the first in the world to have all its electricity come from
nuclear reactor, EBR-I (1951)
nuclear power generated by BORAX-III, another reactor designed and operated by Argonne National Laboratory).
The world's first commercial scale nuclear power station, Obninsk in the Soviet Union, began generation with its
reactor AM-1 on 27 June 1954. Other early nuclear power plants were Calder Hall in England which began
generation on 17 October 1956 and the Shippingport Atomic Power Station in Pennsylvania which began on 26 May
1958. Nuclear power was used for the first time for propulsion by a submarine, the USS Nautilus, in 1954.

Contamination and the Cold War legacy


Above-ground nuclear tests by the Soviet Union and the United States
in the 1950s and early 1960s and by France into the 1970s and 1980s
spread a significant amount of fallout from uranium daughter isotopes
around the world. Additional fallout and pollution occurred from
[4]
several nuclear accidents.
Uranium miners have a higher incidence of cancer. An excess risk of
lung cancer among Navajo uranium miners, for example, has been
documented and linked to their occupation. The Radiation Exposure
Compensation Act, a 1990 law in the USA, required $100,000 in U.S. and USSR/Russian nuclear
"compassion payments" to uranium miners diagnosed with cancer or weapons stockpiles, 1945–
2005
other respiratory ailments.
During the Cold War between the Soviet Union and the United States, huge stockpiles of uranium were amassed and
tens of thousands of nuclear weapons were created using enriched uranium and plutonium made from uranium. Since
the break-up of the Soviet Union in 1991, an estimated 600 short tons (540 metric tons) of highly enriched weapons
grade uranium (enough to make 40,000 nuclear warheads) have been stored in often inadequately guarded facilities
in the Russian Federation and several other former Soviet states. Police in Asia, Europe, and South America on at
least 16 occasions from 1993 to 2005 have intercepted shipments of smuggled bomb-grade uranium or plutonium,
most of which was from ex-Soviet sources. From 1993 to 2005 the Material Protection, Control, and Accounting
Program, operated by the federal government of the United States, spent approximately US $550 million to help
safeguard uranium and plutonium stockpiles in Russia. This money was used for improvements and security
enhancements at research and storage facilities. Scientific American reported in February 2006 that in some of the
facilities security consisted of chain link fences which were in severe states of disrepair. According to an interview
from the article, one facility had been storing samples of enriched (weapons grade) uranium in a broom closet before
the improvement project; another had been keeping track of its stock of nuclear warheads using index cards kept in a
shoe box.
Occurrence

Biotic and abiotic


Main article: Uranium in the environment
Uranium is a naturally occurring element that can be found in low
levels within all rock, soil, and water. Uranium is the 51st element in
order of abundance in the Earth's crust. Uranium is also the
highest-numbered element to be found naturally in significant
quantities on Earth and is almost always found combined with other
elements. Along with all elements having atomic weights higher than
that of iron, it is only naturally formed in supernovae. The decay of
uranium, thorium, and potassium-40 in the Earth's mantle is thought to
be the main source of heat that keeps the outer core liquid and drives
mantle convection, which in turn drives plate tectonics.
Uraninite, also known as pitchblende, is the most
Uranium's average concentration in the Earth's crust is (depending on
common ore mined to extract uranium.
the reference) 2 to 4 parts per million, or about 40 times as abundant as
silver. The Earth's crust from the surface to 25 km (15 mi) down is
17 17
calculated to contain 10 kg (2×10 lb) of uranium while the oceans
13 13
may contain 10 kg (2×10 lb). The concentration of uranium in soil
ranges from 0.7 to 11 parts per million (up to 15 parts per million in
farmland soil due to use of phosphate fertilizers), and its concentration
in sea water is 3 parts per billion.

Uranium is more plentiful than antimony, tin, cadmium, mercury, or


silver, and it is about as abundant as arsenic or molybdenum. Uranium
is found in hundreds of minerals including uraninite (the most common
uranium ore), carnotite, autunite, uranophane, torbernite, and coffinite. The evolution of Earth's radiogenic heat
235
Significant concentrations of uranium occur in some substances such flow over time: contribution from U in
238
as phosphate rock deposits, and minerals such as lignite, and monazite pink and from U in light blue.

sands in uranium-rich ores (it is recovered commercially from sources


with as little as 0.1% uranium).
Some bacteria such as S. putrefaciens and G. metallireducens have
been shown to reduce U(VI) to U(IV). Some organisms, such as the
lichen Trapelia involuta or microorganisms such as the bacterium
Citrobacter, can absorb concentrations of uranium that are up to 300
times higher than in their environment. Citrobacter species absorb
uranyl ions when given glycerol phosphate (or other similar organic
phosphates). After one day, one gram of bacteria can encrust
themselves with nine grams of uranyl phosphate crystals; this creates
Citrobacter species can have concentrations of
the possibility that these organisms could be used in bioremediation to
uranium in their bodies 300 times higher than in
decontaminate uranium-polluted water. The proteobacterium the surrounding environment.
Geobacter has also been shown to bioremediate uranium in ground
water. The mycorrhizal fungus Glomus intraradices increases uranium content in the roots of its symbiotic plant.

In nature, uranium(VI) forms highly soluble carbonate complexes at alkaline pH. This leads to an increase in
mobility and availability of uranium to groundwater and soil from nuclear wastes which leads to health hazards.
However, it is difficult to precipitate uranium as phosphate in the presence of excess carbonate at alkaline pH. A
Sphingomonas sp. strain BSAR-1 has been found to express a high activity alkaline phosphatase (PhoK) that
has been applied for bioprecipitation of uranium as uranyl phosphate species from alkaline solutions. The
precipitation ability was enhanced by overexpressing PhoK protein in E. coli.
Plants absorb some uranium from soil. Dry weight concentrations of uranium in plants range from 5 to 60 parts per
billion, and ash from burnt wood can have concentrations up to 4 parts per million. Dry weight concentrations of
uranium in food plants are typically lower with one to two micrograms per day ingested through the food people eat.

Production and mining


Main article: Uranium
mining
The worldwide production of uranium in 2010 amounted to 53,663
tonnes, of which 17,803 t (33.2%) was mined in Kazakhstan. Other
important uranium mining countries are Canada (9,783 t), Australia
(5,900 t), Namibia (4,496 t), Niger (4,198 t) and Russia (3,562 t).
Uranium ore is mined in several ways: by open pit, underground, in-
situ leaching, and borehole mining (see uranium mining). Low-grade
uranium ore mined typically contains 0.01 to 0.25% uranium oxides.
Extensive measures must be employed to extract the metal from its World uranium production (mines) and demand

ore. High-grade ores found in Athabasca Basin deposits in


Saskatchewan, Canada can contain up to 23% uranium oxides on
average. Uranium ore is crushed and rendered into a fine powder and
then leached with either an acid or alkali. The leachate is subjected to
one of several sequences of precipitation, solvent extraction, and ion
exchange. The resulting mixture, called yellowcake, contains at least
75% uranium oxides U O . Yellowcake is then calcined to remove
3 8
impurities from the milling process before refining and conversion.

Commercial-grade uranium can be produced through the reduction of Yellowcake is a concentrated mixture of uranium
uranium halides with alkali or alkaline earth metals. Uranium metal oxides that is further refined to extract pure
can also be prepared through electrolysis of KUF uranium.

5 or UF

4, dissolved in molten calcium chloride


(CaCl
2) and sodium chloride (NaCl) solution. Very pure uranium is produced through the thermal decomposition of
uranium halides on a hot filament.

Resources and reserves


It is estimated that 5.5 million tonnes of uranium exists in ore reserves that are economically viable at US$59
per lb of uranium, while 35 million tonnes are classed as mineral resources (reasonable prospects for eventual
economic extraction). Prices went from about $10/lb in May 2003 to $138/lb in July 2007. This has caused a big
increase in spending on exploration, with US$200 million being spent world wide in 2005, a 54% increase on
the previous year. This trend continued through 2006, when expenditure on exploration rocketed to over $774
million, an increase of over 250% compared to 2004. The OECD Nuclear Energy Agency said exploration
figures for 2007 would likely match those for 2006.
Australia has 31% of the world's known uranium ore reserves and the world's largest single uranium deposit, located
at the Olympic Dam Mine in South Australia. There is a significant reserve of uranium in Bakouma a sub-prefecture
in the prefecture of Mbomou in Central African Republic.
Some nuclear fuel comes from nuclear weapons being dismantled, such as from the Megatons to Megawatts
Program.
An additional 4.6 billion tonnes of uranium are estimated to be in sea water (Japanese scientists in the 1980s showed
that extraction of uranium from sea water using ion exchangers was technically feasible). There have been
experiments to extract uranium from sea water, but the yield has been low due to the carbonate present in the water.
In 2012, ORNL researchers announced the successful development of a new absorbent material dubbed HiCap
which performs surface retention of solid or gas molecules, atoms or ions and also effectively removes toxic metals
from water, according to results verified by researchers at Pacific Northwest National Laboratory.

Supplies
Main article: Uranium market
In 2005, seventeen countries produced concentrated uranium oxides,
with Canada (27.9% of world production) and Australia (22.8%) being
the largest producers and Kazakhstan (10.5%), Russia (8.0%), Namibia
(7.5%), Niger (7.4%), Uzbekistan (5.5%), the United States (2.5%),
Argentina (2.1%), Ukraine (1.9%) and China (1.7%) also producing
significant amounts. Kazakhstan continues to increase production and
may have become the world's largest producer of uranium by 2009
with an expected production of 12,826 tonnes, compared to Canada
with 11,100 t and Australia with 9,430 t. The ultimate available Monthly uranium spot price in US$ per pound.
The 2007 price peak is clearly visible.
uranium is believed to be sufficient for at least the next 85 years,
although some studies indicate underinvestment in the late twentieth
century may produce supply problems in the 21st century. Uranium deposits seem to be log-normal distributed.
There is a 300-fold increase in the amount of uranium recoverable for each tenfold decrease in ore grade. In other
words, there is little high grade ore and proportionately much more low grade ore available.

Compounds

Oxidation states and


oxides

Oxides

Triuranium octoxide (left) and uranium dioxide (right) are the two most common uranium
oxides.
Calcined uranium yellowcake as produced in many large mills contains a distribution of uranium oxidation species
in various forms ranging from most oxidized to least oxidized. Particles with short residence times in a calciner will
generally be less oxidized than those with long retention times or particles recovered in the stack scrubber. Uranium
content is usually referenced to U
3O

8,which dates to the days of the Manhattan project when U


3O

8 was used as an analytical chemistry reporting standard.

Phase relationships in the uranium-oxygen system are complex. The most important oxidation states of uranium are
uranium(IV) and uranium(VI), and their two corresponding oxides are, respectively, uranium dioxide (UO
2) and uranium trioxide
(UO
3). Other uranium oxides such as uranium monoxide (UO), diuranium pentoxide (U

2O

5), and uranium peroxide (UO

4·2H

2O) also exist.

The most common forms of uranium oxide are triuranium octoxide


(U
3O

8) and
UO
2. Both oxide forms are solids that have low solubility in water and are relatively stable over a wide range of

environmental conditions. Triuranium octaoxide is (depending on conditions) the most stable compound of uranium
and is the form most commonly found in nature. Uranium dioxide is the form in which uranium is most commonly
used as a nuclear reactor fuel. At ambient temperatures, UO
2 will gradually convert to
U
3O

8. Because of their stability, uranium oxides are generally considered the preferred chemical form for storage or

disposal.

Aqueous chemistry

Salts of many oxidation states of uranium are water-soluble and may be


studied in aqueous solutions. The most common ionic forms are U3+
(brown-red),
U4+ (green),
UO+
2 (unstable),and UO2+
2 (yellow), for U(III), U(IV), U(V), and U(VI), respectively. A few solid and

semi-metallic compounds such as UO and US exist for the formal


oxidation state uranium(II), but no simple ions are known to exist in
Uranium in its oxidation states III, IV,
solution for that state. Ions of U3+
V, VI
liberate hydrogen from water and are therefore considered to be highly
unstable. The UO2+
2 ion represents the uranium(VI) state and is known to form compounds such

as uranyl carbonate, uranyl chloride and uranyl sulfate. UO2+


2 also forms complexes with various organic chelating agents, the most commonly encountered of which is uranyl

acetate.

Unlike the uranyl salts of uranium and polyatomic ion uranium-oxide cationic forms, the uranates, salts containing a
polyatomic uranium-oxide anion, are generally not water-soluble.
Carbonates
The interactions of carbonate anions with uranium(VI) cause the Pourbaix diagram to change greatly when the
medium is changed from water to a carbonate containing solution. While the vast majority of carbonates are
insoluble in water (students are often taught that all carbonates other than those of alkali metals are insoluble in
water), uranium carbonates are often soluble in water. This is because a U(VI) cation is able to bind two terminal
oxides and three or more carbonates to form anionic complexes.

Pourbaix diagrams[5]

Uranium in a non-complexing aqueous medium (e.g. perchloric Uranium in carbonate solution


acid/sodium hydroxide).

Relative concentrations of the different chemical forms of uranium in a non- Relative concentrations of the different chemical forms of uranium in an
complexing aqueous medium (e.g. perchloric acid/sodium hydroxide). aqueous carbonate solution.
Effects of pH
The uranium fraction diagrams in the presence of carbonate illustrate this further: when the pH of a uranium(VI)
solution increases, the uranium is converted to a hydrated uranium oxide hydroxide and at high pHs it becomes an
anionic hydroxide complex.
When carbonate is added, uranium is converted to a series of carbonate complexes if the pH is increased. One
effect of these reactions is increased solubility of uranium in the pH range 6 to 8, a fact which has a direct
bearing on the long term stability of spent uranium dioxide nuclear fuels.

Hydrides, carbides and nitrides


Uranium metal heated to 250 to 300 °C (482 to 572 °F) reacts with hydrogen to form uranium hydride. Even
higher temperatures will reversibly remove the hydrogen. This property makes uranium hydrides convenient
starting materials to create reactive uranium powder along with various uranium carbide, nitride, and halide
compounds. Two crystal modifications of uranium hydride exist: an α form that is obtained at low temperatures
and a β form that is created when the formation temperature is above 250 °C.
Uranium carbides and uranium nitrides are both relatively inert semimetallic compounds that are minimally soluble
in acids, react with water, and can ignite in air to form U
3O

8. Carbides of uranium include uranium monocarbide (UC), uranium dicarbide (UC


2), and diuranium tricarbide
(U
2C

3). Both UC and


UC
2 are formed by adding carbon to molten uranium or by exposing the metal to carbon monoxide at high
temperatures. Stable below 1800 °C, U
2C

3 is prepared by subjecting a heated mixture of UC and


UC
2 to mechanical stress. Uranium nitrides obtained by direct exposure of the metal to nitrogen include uranium

mononitride (UN), uranium dinitride (UN


2), and diuranium trinitride
(U
2N

3).

Halides
All uranium fluorides are created using uranium tetrafluoride (UF 17, and UF
4); UF 5.

4 itself is prepared by hydrofluorination of uranium dioxide.

Reduction of UF
4 with hydrogen at 1000 °C produces uranium trifluoride (UF

3). Under the right conditions of temperature and pressure, the reaction

of solid UF
4 with gaseous uranium hexafluoride (UF

6) can form the intermediate fluorides of U

2F

9, U

4F
Uranium hexafluoride is the feedstock used to
separate uranium-235 from natural uranium.
At room temperatures,
UF
6 has a high vapor pressure, making it useful in the gaseous diffusion process to separate the rare uranium-235 from

the common uranium-238 isotope. This compound can be prepared from uranium dioxide and uranium hydride by
the following process:
UO
2 + 4 HF → UF

4+ 2 H

2O (500 °C, endothermic)

UF
4+ F
2 → UF

6 (350 °C, endothermic)

The resulting UF
6,a white solid, is highly reactive (by fluorination), easily sublimes (emitting a vapor that behaves as a nearly
ideal
gas), and is the most volatile compound of uranium known to exist.
One method of preparing uranium tetrachloride (UCl
4) is to directly combine chlorine with either uranium metal or uranium hydride. The reduction of UCl
4 by hydrogen produces uranium trichloride
(UCl
3) while the higher chlorides of uranium are prepared by reaction with additional chlorine. All uranium chlorides

react with water and air.


Bromides and iodides of uranium are formed by direct reaction of, respectively, bromine and iodine with uranium or
by adding UH
3 tothose element's acids. Known examples include: UBr
3, UBr

4,
UI
3, and
UI
4. Uranium oxyhalides are water-soluble and include UO

2F

2, UOCl

2,
UO
2Cl

2, and UO

2Br

2. Stability of the oxyhalides decrease as the atomic weight of the component halide increases.

Isotopes

Natural concentrations
Main article: Isotopes of
uranium
Natural uranium consists of three major isotopes: uranium-238 (99.28% natural abundance), uranium-235 (0.71%),
and uranium-234 (0.0054%). All three are radioactive, emitting alpha particles, with the exception that all three of
these isotopes have small probabilities of undergoing spontaneous fission, rather than alpha emission.
9
Uranium-238 is the most stable isotope of uranium, with a half-life of about 4.468×10 years, roughly the age of the
8 5
Earth. Uranium-235 has a half-life of about 7.13×10 years, and uranium-234 has a half-life of about 2.48×10 years.
238 234
For natural uranium, about 49% of its alpha rays are emitted by each of U atom, and also 49% by U (since the
92 Uranium 76

235
latter is formed from the former) and about 2.0% of them by the U. When the Earth was young, probably about
234
one-fifth of its uranium was uranium-235, but the percentage of U was probably much lower than this.
Uranium-238 is usually an α emitter (occasionally, it undergoes spontaneous fission), decaying through the
"Uranium Series" of nuclear decay, which has 18 members, all of which eventually decay into lead-206, by a variety
of different decay paths.
235
The decay series of U, which is called the actinium series has 15 members, all of which eventually decay into
lead-207. The constant rates of decay in these decay series makes the comparison of the ratios of parent to daughter
elements useful in radiometric dating.
Uranium-234 is a member of the "Uranium Series", and it decays to lead-206 through a series of relatively short-
lived isotopes.
233
Uranium-233 is made from thorium-232 by neutron bombardment, usually in a nuclear reactor, and U is also
fissile. Its decay series ends with thallium-205.
Uranium-235 is important for both nuclear reactors and nuclear weapons, because it is the only uranium isotope
existing in nature on Earth in any significant amount that is fissile. This means it can be split into two or three
fragments (fission products) by thermal neutrons.
Uranium-238 is not fissile, but is a fertile isotope, because after neutron activation it can produce plutonium-239,
238
another fissile isotope. Indeed, the U nucleus can absorb one neutron to produce the radioactive isotope uranium-
239
239. U decays by beta emission to neptunium-239, also a beta-emitter, that decays in its turn, within a few
239
days into plutonium-239. Pu was used as fissile material in the first atomic bomb detonated in the "Trinity test"
on 15 July 1945 in New Mexico.

Enrichment
Main article: Enriched
uranium
In nature, uranium is found as uranium-238 (99.2742%) and uranium-
235 (0.7204%). Isotope separation concentrates (enriches) the
fissionable uranium-235 for nuclear weapons and most nuclear power
plants, except for gas cooled reactors and pressurised heavy water
reactors. Most neutrons released by a fissioning atom of uranium-235
must impact other uranium-235 atoms to sustain the nuclear chain
reaction. The concentration and amount of uranium-235 needed to
achieve this is called a 'critical mass'.

To be considered 'enriched', the uranium-235 fraction should be


between 3% and 5%. This process produces huge quantities of Cascades of gas centrifuges are used to
enrich uranium ore to concentrate its
uranium that is depleted of uranium-235 and with a correspondingly
fissionable isotopes.
increased fraction of uranium-238, called depleted uranium or 'DU'.
To be considered 'depleted', the uranium-235 isotope concentration
should be
no more than 0.3%. The price of uranium has risen since 2001, so enrichment tailings containing more than 0.35%
uranium-235 are being considered for re-enrichment, driving the price of depleted uranium hexafluoride above $130
per kilogram in July 2007 from $5 in 2001.
The gas centrifuge process, where gaseous uranium hexafluoride
(UF
235 238
6) is separated by the difference in molecular weight between UF and UF using high-speed centrifuges, is the
6 6
cheapest and leading enrichment process. The gaseous diffusion process had been the leading method for
enrichment
and was used in the Manhattan Project. In this process, uranium hexafluoride is repeatedly diffused through a
silver-zinc membrane, and the different isotopes of uranium are separated by diffusion rate (since uranium 238 is
heavier it diffuses slightly slower than uranium-235). The molecular laser isotope separation method employs a laser
92 Uranium 1

beam of precise energy to sever the bond between uranium-235 and fluorine. This leaves uranium-238 bonded to
fluorine and allows uranium-235 metal to precipitate from the solution. An alternative laser method of enrichment is
known as atomic vapor laser isotope separation (AVLIS) and employs visible tunable lasers such as dye lasers.
Another method used is liquid thermal diffusion.

Human exposure
A person can be exposed to uranium (or its radioactive daughters such as radon) by inhaling dust in air or by
ingesting contaminated water and food. The amount of uranium in air is usually very small; however, people who
work in factories that process phosphate fertilizers, live near government facilities that made or tested nuclear
weapons, live or work near a modern battlefield where depleted uranium weapons have been used, or live or work
near a coal-fired power plant, facilities that mine or process uranium ore, or enrich uranium for reactor fuel, may
have increased exposure to uranium. Houses or structures that are over uranium deposits (either natural or man-
made slag deposits) may have an increased incidence of exposure to radon gas.
Most ingested uranium is excreted during digestion. Only 0.5% is absorbed when insoluble forms of uranium, such
as its oxide, are ingested, whereas absorption of the more soluble uranyl ion can be up to 5%. However, soluble
uranium compounds tend to quickly pass through the body whereas insoluble uranium compounds, especially when
inhaled by way of dust into the lungs, pose a more serious exposure hazard. After entering the bloodstream, the
absorbed uranium tends to bioaccumulate and stay for many years in bone tissue because of uranium's affinity for
phosphates. Uranium is not absorbed through the skin, and alpha particles released by uranium cannot penetrate the
skin.
Incorporated uranium becomes uranyl ions, which accumulate in bone, liver, kidney, and reproductive tissues.
Uranium can be decontaminated from steel surfaces and aquifers.

Effects and
precautions
Normal functioning of the kidney, brain, liver, heart, and other systems can be affected by uranium exposure,
because, besides being weakly radioactive, uranium is a toxic metal. Uranium is also a reproductive toxicant.
238
Radiological effects are generally local because alpha radiation, the primary form of U decay, has a very short
range, and will not penetrate skin. Uranyl (UO2+
2) ions, such as from uranium trioxide or uranyl nitrate and other hexavalent uranium compounds, have been shown

to cause birth defects and immune system damage in laboratory animals. While the CDC has published one study
that no human cancer has been seen as a result of exposure to natural or depleted uranium, exposure to uranium and
its decay products, especially radon, are widely known and significant health threats. Exposure to strontium-90,
iodine-131, and other fission products is unrelated to uranium exposure, but may result from medical procedures or
[6]
exposure to spent reactor fuel or fallout from nuclear weapons. Although accidental inhalation exposure to a high
concentration of uranium hexafluoride has resulted in human fatalities, those deaths were associated with generation
of highly toxic hydrofluoric acid and uranyl fluoride rather than with uranium itself. Finely divided uranium metal
presents a fire hazard because uranium is pyrophoric; small grains will ignite spontaneously in air at room
temperature.
Uranium metal is commonly handled with gloves as a sufficient precaution. Uranium concentrate is handled and
contained so as to ensure that people do not inhale or ingest it.
Body system Human studies Animal studies In vitro

Renal Elevated levels of protein excretion, Damage to proximal convoluted tubules, No studies
urinary catalase and diuresis necrotic cells cast from tubular epithelium,
glomerular changes (renal failure)

Brain/CNS Decreased performance Acute cholinergic toxicity; No studies


on neurocognitive tests Dose-dependent accumulation in cortex,
midbrain, and vermis;
Electrophysiological changes in
hippocampus

DNA Increased reports of cancers Increased mutagenicity (in mice) Binucleated cells with
and induction of tumors micronuclei, Inhibition of cell
cycle kinetics and proliferation;
Sister chromatid induction,
tumorigenic phenotype
Bone/muscle No studies Inhibition of periodontal bone formation; No studies
and alveolar wound healing

Reproductive Uranium miners have more first Moderate to severe focal tubular atrophy; No studies
born female children vacuolization of Leydig cells

Lungs/respiratory No adverse health effects reported Severe nasal congestion and hemorrhage, No studies
lung lesions and fibrosis, edema and
swelling, lung cancer

Gastrointestinal Vomiting, diarrhea, albuminuria No studies No studies

Liver No effects seen at exposure dose Fatty livers, focal necrosis No studies

Skin No exposure assessment data available Swollen vacuolated epidermal cells, No studies
damage to hair follicles and sebaceous
glands

Tissues surrounding Elevated uranium urine concentrations Elevated uranium urine concentrations, No studies
embedded DU fragments perturbations in biochemical and
neuropsychological testing

Immune system Chronic fatigue, rash, ear and eye No studies No studies
infections, hair and weight loss, cough.
May be due to combined chemical
exposure rather than DU alone

Eyes No studies Conjunctivitis, irritation inflammation, No studies


edema, ulceration of conjunctival sacs

Blood No studies Decrease in RBC count and hemoglobin No studies


concentration

Cardiovascular Myocarditis resulting from the No effects No studies


uranium ingestion, which ended 6
months after ingestion

|+Compilation of 2004 review on uranium toxicity


Notes
[1] http://en. wikipedia.org/w/index.php?title=Template:Infobox_uranium&action=edit
[2] "U.S. to pump money into nuke stockpile, increase security," (http:// en.rian.ru/world/20100218/157925732. html) RIA Novosti 18
February 2010
[3] Helmreich, J.E. Gathering Rare Ores: The Diplomacy of Uranium Acquisition, 1943–1954, Princeton UP, 1986: ch. 10 ISBN 0-7837-9349-9
[4] Newtan, Samuel Upton (2007). Nuclear War I and Other Major Nuclear Disasters of the 20th Century, AuthorHouse ISBN 1-4259-8512-2.
[5] Puigdomenech, Ignasi Hydra/Medusa Chemical Equilibrium Database and Plotting Software (2004) KTH Royal Institute of Technology,
freely downloadable software at (http:// www.kemi.kth.se/ medusa/ )
[6] Chart of the Nuclides, US Atomic Energy Commission 1968

References
Full reference information for multi-page works cited
• Emsley, John (2001). "Uranium" (http://books.google.com/?id=j-Xu07p3cKwC&printsec=frontcover).
Nature's Building Blocks: An A to Z Guide to the Elements. Oxford: Oxford University Press. pp. 476–482.
ISBN 0-19-850340-7.
• Seaborg, Glenn T. (1968). "Uranium". The Encyclopedia of the Chemical Elements. Skokie, Illinois: Reinhold
Book Corporation. pp. 773–786. LCCCN 68-29938.

External links
• Uranium Resources and Nuclear Energy (http://www.lbst.de/publications/studies e/2006/
EWG-paper_1-06_Uranium-Resources-Nuclear-Energy_03DEC2006 .pdf)
• U.S. EPA: Radiation Information for Uranium (http://www.epa.gov/radiation/radionuclides/uranium.html)
• "What is Uranium?" from World Nuclear Association (http://world-nuclear.org/education/uran.htm)
• Nuclear fuel data and analysis from the U.S. Energy Information Administration (http://www.eia.gov/nuclear/
)
• Current market price of uranium (http://www.uxc.com)
• World Uranium deposit maps (http://www.wise-uranium.org/umaps.html)
• Annotated bibliography for uranium from the Alsos Digital Library (http://alsos.wlu.edu/qsearch.
aspx?browse=science/Uranium)
• NLM Hazardous Substances Databank—Uranium, Radioactive (http://toxnet.nlm.nih.gov/cgi-bin/sis/search/
r?dbs+hsdb:@term+@na+@rel+uranium,+radioactive)
• CDC - NIOSH Pocket Guide to Chemical Hazards (http://www.cdc.gov/niosh/npg/npgd0650.html)
• Mining Uranium at Namibia's Langer Heinrich Mine (http://viewer.zmags.com/showmag.php?mid=pfgsh#/
page34/)
• World Nuclear News (http://www.world-nuclear-news.org/)
• ATSDR Case Studies in Environmental Medicine: Uranium Toxicity (http://www.atsdr.cdc.gov/csem/
uranium/) U.S. Department of Health and Human Services
• Real Time Uranium Prices (http://www.bloomberg.com/quote/MFURMDUR:IND)
• Uranium (http://www.periodicvideos.com/videos/092.htm) at The Periodic Table of Videos (University
of Nottingham)
93 Neptunium 1

93 – Neptunium
Neptunium
Np
93

uranium ← neptunium → plutoniumPm



Np

(Uqp)

Neptunium in the periodic table


Appearance

silvery metallic

General properties

Name, symbol, number neptunium, Np, 93

Pronunciation UK /nɛpˈtjuːniəm/
nep-TEW-nee-əm
US /nɛpˈtuːniəm/
nep-TOO-nee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (237)

Electron configuration 4 1 2
[Rn] 5f 6d 7s
2, 8, 18, 32, 22, 9, 2

Physical properties

Phase solid
[1] −3
Density (near r.t.) (alpha) g·cm
20.45
Density (near r.t.) −3
(accepted standard value) 19.38 g·cm

Melting point 912 K, 639 °C, 1182 °F

Boiling point 4447 K, 4174 °C, 7545 (extrapolated) °F

Heat of fusion −1
5.19 kJ·mol

Heat of vaporization −1
336 kJ·mol

Molar heat capacity −1 −1


29.46 J·mol ·K

Vapor pressure

P (Pa) 1 10 100 1 k 10 k 100 k

at T (K) 2194 2437

Atomic properties

Oxidation states 7, 6, 5, 4, 3
(amphoteric oxide)

Electronegativity 1.36 (Pauling scale)

Ionization energies −1
1st: 604.5 kJ·mol

Atomic radius 155 pm

Covalent radius 190±1 pm

Miscellanea

Crystal structure orthorhombic

Magnetic ordering [2]


paramagnetic

Electrical resistivity (22 °C) 1.220 µΩ·m

Thermal conductivity −1 −1
6.3 W·m ·K

CAS registry number 7439-99-8

History

Naming after planet Neptune, itself named after Roman god of the sea Neptune

Discovery Edwin McMillan and Philip H. Abelson (1940)

Most stable isotopes

Main article: Isotopes of neptunium


iso NA half-life DM DE (MeV) DP

235 syn 396.1 d α 5.192 231


Np Pa
ε 0.124 235
U

236 syn 5 ε 0.940 236


Np 1.54×10 y U

− 0.940 236
β Pu
α 5.020 232
Pa

237 trace 6 α 4.959 233


Np 2.144×10 y Pa

239 trace 2.356 d − 0.218 239


Np β Pu

• v
• t
[3]
• e

Neptunium is a chemical element with the symbol Np and atomic number 93. A radioactive actinide metal,
neptunium is the first transuranic element. Its position in the periodic table just after uranium, named after the planet
Uranus, led to its being named after Neptune, the next planet beyond Uranus. A neptunium atom has 93 protons and
93 electrons, of which seven are valence electrons. Neptunium metal is silvery and tarnishes when exposed to air. It
occurs in three allotropic forms. The element normally exhibits five oxidation states, ranging from +3 to +7.
In the 1870s, Dmitri Mendeleev first predicted the existence of neptunium, and many false claims of its
discovery were made over the years. The element was first synthesized by Edwin McMillan and Philip H.
Abelson at the Berkeley Radiation Laboratory in 1940. Most neptunium is produced by bombarding uranium
with neutrons in nuclear reactors; neptunium is also generated as a by-product in conventional nuclear
reactors. Though neptunium has no commercial uses at present, it is widely used as a precursor for the
formation of plutonium-238, used in radioisotope thermal generators which are used to power some spacecraft.
Neptunium itself can be used in detectors of high-energy neutrons.
The most stable isotope of neptunium, neptunium-237, is a by-product of nuclear reactors and plutonium production,
and it can be used as a component in neutron detection equipment. The isotopes neptunium-237 and neptunium-239
are also found in trace amounts in uranium ores due to neutron capture reactions and beta decay.

History

Pre-discovery
Even before its discovery, the existence of neptunium had been predicted several times. The periodic table of Dmitri
Mendeleev published in the 1870s showed a " — " in place after uranium similar to several other places for at that
point undiscovered elements. Also, a 1913 publication of the known radioactive isotopes by Kasimir Fajans shows
the empty place after uranium.
The search for element 93 in minerals was encumbered by the fact that the predictions on the chemical properties of
element 93 were based on a periodic table which lacked the actinide series, and therefore placed thorium below
hafnium, protactinium below tantalum, and uranium below tungsten. This periodic table suggested that element 93,
at that point often named eka-rhenium, should be similar to manganese or rhenium. With this misconception it was
impossible to isolate element 93 from minerals, although neptunium was later found in uranium ore, in 1952.
In 1934, Odolen Koblic extracted a small amount of material from the wash water of roasted pitchblende. He
assumed the sample was element 93, and called it bohemium, but after being analyzed, it turned out that the sample
was a mixture of tungsten and vanadium.
In 1938, Horia Hulubei, a Romanian physicist; and Yvette Cauchois, a French chemist; claimed to have discovered
element 93 via spectroscopy in minerals. They named their element sequanium, but the claim was opposed at the
time because neptunium was thought to occur exclusively artificially. As neptunium does occur in nature, it is
possible that Hulubei and Cauchois did discover neptunium.
Enrico Fermi believed that bombarding uranium with neutrons and subsequent beta decay would lead to the
formation of element 93. Chemical separation of the new formed elements from the uranium yielded material
with low half-life, and, therefore, Fermi announced the discovery of a new element in 1934, though this was
soon found to be mistaken. Soon it was speculated and later proven that most of the material is created by
nuclear fission of uranium by neutrons. Small quantities of neptunium had to be produced in Otto Hahn's
239
experiments in late 1930s as a result of decay of U. Hahn and his colleagues experimentally confirmed
239
production and chemical properties of U, but were unsuccessful at isolating and detecting neptunium.

Discovery
In 1939, Edwin McMillan did some preliminary work leading to the eventual discovery of element 93. Uranium
trioxide was placed on paper together with aluminium or paper foils and the result was irradiated with neutrons from
a cyclotron. Examination revealed the presence of two components, both beta decaying: one turned out to be
uranium-239 (half-life 23 minutes), while the other could not be characterized firmly. Emilio Segrè contradictorily
labelled the second component as having atomic number 93 but not being a transuranic element, also noting that it
behaved similarly to the rare earth elements. This lack of confidence in labelling the product as transuranic was from
[4]
the absence of any observed activity from the beta decay product of element 93.
Neptunium (named for the planet Neptune, the next planet out from Uranus, after which uranium was named)
was finally convincingly discovered by Edwin McMillan and Philip H. Abelson in 1940 at the Berkeley
Radiation Laboratory of the University of California, Berkeley. The team produced the neptunium isotope
239
Np (2.4 day half-life) by bombarding uranium with slow moving neutrons. It was the first transuranium
element produced synthetically and the first actinide series transuranium element discovered.

Occurrence
237
The most stable isotope of neptunium is Np, with a half-life of two million years. Thus, all primordial neptunium
should have decayed by now. Trace amounts of the neptunium isotopes neptunium-237 through neptunium-240, are
found naturally as decay products from transmutation reactions in uranium ores.
237 237
Artificial Np is produced through a reaction of NpF with liquid barium or lithium at around 1200 °C and is
3
most often extracted from spent nuclear fuel rods in kilogram amounts as a by-product in plutonium production.
2 NpF
3 + 3 Ba → 2 Np + 3 BaF

By weight, neptunium-237 discharges are about 5% as great as plutonium discharges and about 0.05% of spent
[5]
nuclear fuel discharges. This amounts to more than fifty tons per year.
Characteristics
Silvery in appearance, neptunium metal is chemically fairly reactive and is found in at least three allotropes:
• α-neptunium, orthorhombic, density 20.45 g/cm3
• β-neptunium (above 280 °C), tetragonal, density (313 °C) 19.36 g/cm3
• γ-neptunium (above 577 °C), cubic, density (600 °C) 18 g/cm3
Neptunium has the largest liquid range of any element, 3363 K, between the melting point and boiling point. It is the
[6]
densest of all the actinides and the fifth-densest of all naturally occurring elements. Neptunium has no biological
role. It is not absorbed by the digestive tract. When injected into the body, it accumulates in bones, from which it is
slowly released.

Isotopes
Main article: Isotopes of neptunium
237
19 neptunium radioisotopes have been characterized, with the most stable being Np with a half-life of 2.14
236 235
million years, Np with a half-life of 154,000 years, and Np with a half-life of 396.1 days. All of the remaining
radioactive isotopes have half-lives that are less than 4.5 days, and the majority of these have half-lives that are less
236m
than 50 minutes. This element also has 4 meta states, with the most stable being Np (t 22.5 hours).
½
225 244
The isotopes of neptunium range in atomic weight from 225.0339 u ( Np) to 244.068 u ( Np). The primary
237
decay mode before the most stable isotope, Np, is electron capture (with a good deal of alpha emission), and the
237
primary mode after is beta emission. The primary decay products before Np are element 92 (uranium) isotopes
(alpha emission produces element 91, protactinium, however) and the primary products after are element 94
(plutonium) isotopes.
237 237
Np is fissionable. Np eventually decays to form bismuth-209 and thallium-205, unlike most other common
heavy nuclei which decay to make isotopes of lead. This decay chain is known as the neptunium series.

Synthesis
Chemically, neptunium is prepared by the reduction of NpF with barium or lithium vapor at about 1200 °C. Most
3
Np is produced in nuclear reactions:
• When an 235U atom captures a neutron, it is converted to an excited state of 236U. About 81% of the excited 236U
236
nuclei undergo fission, but the remainder decay to the ground state of U by emitting gamma radiation. Further
237 237
neutron capture creates U which has a half-life of 7 days and thus quickly decays to Np through beta decay.
237
During beta decay, the excited U emits an electron, while the atomic weak interaction converts a neutron to a
237
proton, thus creating Np.

237 238
• U is also produced via an (n,2n) reaction with U. This only happens with very energetic neutrons.
237 241
• Np is the product of alpha decay of Am.
Heavier isotopes of neptunium decay quickly, and lighter isotopes of neptunium cannot be produced by neutron
237
capture, so chemical separation of neptunium from cooled spent nuclear fuel gives nearly pure Np.
Chemistry
This element has five ionic oxidation states while in solution:
• Np3+ (pale purple), analogous to the rare earth ion Pm3+
• Np4+ (yellow-green)
• NpO+
2 (green-blue)

• NpO2+
2 (pale pink)
• NpO3− Neptunium ions in solution.
5 (green). This may better be labelled as a hydroxo species [NpO
4(OH)

2]3−

.
Neptunium(III) hydroxide is not soluble in water and does not dissolve in excess alkali. Neptunium(III) is
susceptible to oxidation in contact to air forming neptunium(IV).
Neptunium forms tri- and tetrahalides such as NpF , NpF , NpCl , NpBr , NpI , and oxides of the various
3 4 4 3 3
compositions such as are found in the uranium-oxygen system, including Np O and NpO .
3 8 2
Neptunium hexafluoride, NpF , is volatile like uranium hexafluoride.
6
Further information: fluoride volatility and uranium enrichment
Neptunium, like protactinium, uranium, plutonium, and americium readily forms a linear dioxo neptunyl core
n+ –
(NpO ), in its 5+ and 6+ oxidation states, which readily complexes with hard O-donor ligands such as OH ,

NO ,2 2
– 2–
NO , and SO to form soluble anionic complexes which tend to be readily mobile with low affinities to soil.
3 4
Neptunium is very reactive when in contact with oxygen, steam, or acid. However it is not attacked by alkalis.
• NpO (OH) –
2 2
• NpO2(CO3)–
• NpO2(CO3)23–
• NpO (CO ) 5–
2 33
See also: Actinides in the environment

Applications

Precursor in plutonium-238 production


237 238
Np is irradiated with neutrons to create Pu, an alpha emitter for radioisotope thermal generators for spacecraft
237 238
and military applications. Np will capture a neutron to form Np and beta decay with a half-life of two days to
238
Pu.

238
Pu also exists in sizable quantities in spent nuclear fuel but would have to be separated from other isotopes of
plutonium.
Weapons applications
Neptunium is fissionable, and could theoretically be used as fuel in a fast neutron reactor or a nuclear weapon, with a
critical mass of around 60 kilograms. In 1992, the U.S. Department of Energy declassified the statement that
[7]
neptunium-237 "can be used for a nuclear explosive device". It is not believed that an actual weapon has ever been
constructed using neptunium. As of 2009, the world production of neptunium-237 by commercial power reactors
was over 1000 critical masses a year, but to extract the isotope from irradiated fuel elements would be a major
industrial undertaking.
In September 2002, researchers at the Los Alamos National Laboratory briefly created the first known nuclear
critical mass using neptunium in combination with shells of enriched uranium (U-235), discovering that the critical
mass of a bare sphere of neptunium-237 "ranges from kilogram weights in the high fifties to low sixties," showing
that it "is about as good a bomb material as U-235." The United States Federal government made plans in March
2004 to move America's supply of separated neptunium to a nuclear-waste disposal site in Nevada.

Physics applications
237
Np is used in devices for detecting high-energy (MeV) neutrons.

Role in nuclear
waste
Neptunium-237 is the most mobile actinide in the deep geological repository environment. This makes it and its
predecessors such as americium-241 candidates of interest for destruction by nuclear transmutation. Neptunium
accumulates in commercial household ionization-chamber smoke detectors from decay of the (typically) 0.2
microgram of americium-241 initially present as a source of ionizing radiation. With a half-life of 432 years, the
americium-241 in a smoke detector includes about 3% neptunium after 20 years, and about 15% after 100 years.
Due to its long half-life, neptunium becomes the major contributor of the total radiation in 10,000 years. As it is
unclear what happens to the containment in that long time span, an extraction of the neptunium would minimize
the contamination of the environment if the nuclear waste could be mobilized after several thousand years.

Biological role and


precautions
Neptunium does not have any biological role. It is absorbed via the digestive tract. When injected it
concentrates in the bones, from which it is slowly released.

References
237
[1] Criticality of a Np Sphere (http:// typhoon.jaea.go. jp/icnc2003/Proceeding/paper/2. 14_107. pdf)
[2] Magnetic susceptibility of the elements and inorganic compounds (http://www-d0.fnal. gov/ hardware/cal/ lvps_info/engineering/
elementmagn. pdf), in Handbook of Chemistry and Physics 81st edition, CRC press.
[3] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_neptunium&action=edit
[4] Yoshida et al., pp. 699–700
[5] http://www.rsc.org/chemistryworld/podcast/ interactive_periodic_table_transcripts/neptunium.asp
[6] Theodore Gray. The Elements. Page 215
[7] "Restricted Data Declassification Decisions from 1946 until Present" (http:// www.fas.org/sgp/ othergov/doe/rdd-7.html), accessed Sept
23, 2006
Bibliography
• Yoshida, Zenko; Johnson, Stephen G.; Kimura, Takaumi; Krsul, John R. (2006). "Neptunium" (http://radchem.
nevada. edu/ classes/rdch710/files/neptunium.pdf). In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean.
The Chemistry of the Actinide and Transactinide Elements 3 (3rd ed.). Dordrecht, the Netherlands: Springer.
pp. 699–812. doi: 10.1007/1-4020-3598-5_6 (http://dx.doi.org/10. 1007/1-4020-3598-5_6).

Literature
• Guide to the Elements – Revised Edition, Albert Stwertka, (Oxford University Press; 1998) ISBN 0-19-508083-1
• Lester R. Morss, Norman M. Edelstein, Jean Fuger (Hrsg.): The Chemistry of the Actinide and Transactinide
Elements, Springer-Verlag, Dordrecht 2006, ISBN 1-4020-3555-1.
• Ida Noddack (1934). "Über das Element 93" (http://www.chemteam.info/Chem-History/Noddack-1934.html).
Zeitschrift für Angewandte Chemie 47 (37): 653. doi: 10.1002/ange.19340473707 (http://dx.doi.org/10.1002/
ange. 19340473707).
• Eric Scerri, A Very Short Introduction to the Periodic Table, Oxford University Press, Oxford, 2011, ISBN
978-0-19-958249-5.

External links
• Neptunium (http://www.periodicvideos.com/videos/093.htm) at The Periodic Table of Videos (University
of Nottingham)
• Lab builds world's first neptunium sphere (http://www.eurekalert.org/features/doe/2001-08/danl-lbw060502.
php), U.S. Department of Energy Research News
• NLM Hazardous Substances Databank – Neptunium, Radioactive (http://toxnet.nlm.nih.gov/cgi-bin/sis/
search/ r?dbs+hsdb:@term+@na+@rel+neptunium,+radioactive)
• Neptunium: Human Health Fact Sheet (http://www.ead.anl.gov/pub/doc/neptunium.pdf)
• C&EN: It's Elemental: The Periodic Table – Neptunium (http://pubs.acs.org/cen/80th/neptunium.html)
94 Plutonium 1

94 – Plutonium
This article is about the radioactive element. For other uses, see Plutonium (disambiguation).

Plutonium
Pu
94

neptunium ← plutonium → americiumSm



Pu

(Uqh)

Plutonium in the periodic table


Appearance

silvery white, tarnishing to dark gray in air

General properties

Name, symbol, number plutonium, Pu, 94

Pronunciation /pluːˈtoʊniəm/
ploo-TOH-nee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (244)

Electron configuration 6 2
[Rn] 5f 7s
2, 8, 18, 32, 24, 8, 2

Physical properties

Phase solid

Density (near r.t.) −3


19.816 g·cm
Liquid density at m.p. −3
16.63 g·cm

Melting point 912.5 K, 639.4 °C, 1182.9 °F

Boiling point 3505 K, 3228 °C, 5842 °F

Heat of fusion −1
2.82 kJ·mol

Heat of vaporization −1
333.5 kJ·mol

Molar heat capacity −1 −1


35.5 J·mol ·K

Vapor pressure

P (Pa) 1 10 100 1k 10 k 100 k

at T (K) 1756 1953 2198 2511 2926 3499

Atomic properties

Oxidation states 8, 7, 6, 5, 4, 3, 2, 1
(amphoteric oxide)

Electronegativity 1.28 (Pauling scale)

Ionization energies −1
1st: 584.7 kJ·mol

Atomic radius 159 pm

Covalent radius 187±1 pm

Miscellanea

Crystal structure monoclinic

Magnetic ordering [1]


paramagnetic

Electrical resistivity (0 °C) 1.460 µΩ·m

Thermal conductivity −1 −1
6.74 W·m ·K

Thermal expansion −1 −1
(25 °C) 46.7 µm·m ·K

Speed of sound −1
2260 m·s

Young's modulus 96 GPa

Shear modulus 43 GPa

Poisson ratio 0.21

CAS registry number 7440-07-5

History

Naming after minor planet Pluto, itself a recently chosen name

Discovery Glenn T. Seaborg, Arthur Wahl, Joseph W. Kennedy, Edwin McMillan (1940–
1)
Most stable isotopes

Main article: Isotopes of plutonium


iso NA half-life DM DE (MeV) DP

trace 87.74 y SF [2] —


238 204.66
Pu
α 5.5 234
U

239 100% 4 SF 207.06 —


Pu 2.41 × 10 y
α 5.157 235
U

240 trace 3 SF 205.66 —


Pu 6.5 × 10 y
α 5.256 236
U

241 syn 14 y − 0.02078 241


Pu β Am
SF 210.83 —

242 trace 5 SF 209.47 —


Pu 3.73 × 10 y
α 4.984 238
U

244 trace 7 α 4.666 240


Pu 8.08 × 10 y U
SF —

• v
• t
[3]
• e

Plutonium is a transuranic radioactive chemical element with the symbol Pu and atomic number 94. It is an actinide
metal of silvery-gray appearance that tarnishes when exposed to air, and forms a dull coating when oxidized. The
element normally exhibits six allotropes and four oxidation states. It reacts with carbon, halogens, nitrogen, silicon
and hydrogen. When exposed to moist air, it forms oxides and hydrides that expand the sample up to 70% in volume,
which in turn flake off as a powder that can spontaneously ignite. It is radioactive and can accumulate in the bones.
These properties make the handling of plutonium dangerous.
Plutonium is the heaviest primordial element by virtue of its most stable isotope, plutonium-244, whose half-life of
about 80 million years is just long enough for the element to be found in trace quantities in nature. Plutonium is
mostly a byproduct of nuclear reactions in reactors where some of the neutrons released by the fission process
[4]
convert uranium-238 nuclei into plutonium.
Both plutonium-239 and plutonium-241 are fissile, meaning that they can sustain a nuclear chain reaction, leading to
applications in nuclear weapons and nuclear reactors. Plutonium-240 exhibits a high rate of spontaneous fission,
raising the neutron flux of any sample containing it. The presence of plutonium-240 limits a plutonium sample's
usability for weapons or its quality as reactor fuel, and the percentage of plutonium-240 determines its grade
(weapons grade, fuel grade, or reactor grade).
Plutonium-238 has a half-life of 88 years and emits alpha particles. It is a heat source in radioisotope thermoelectric
generators, which are used to power some spacecraft. Plutonium isotopes are expensive and inconvenient to
separate, so particular isotopes are usually manufactured in specialized reactors.
A team led by Glenn T. Seaborg and Edwin McMillan at the University of California, Berkeley, first synthesized
plutonium in 1940 by bombarding uranium-238 with deuterons. Trace amounts of plutonium were subsequently
discovered in nature. Producing plutonium in useful quantities for the first time was a major part of the
Manhattan Project during World War II, which developed the first atomic bombs. The first nuclear test, "Trinity"
(July 1945), and the second atomic bomb used to destroy a city (Nagasaki, Japan, in August 1945), "Fat Man",
both had cores of plutonium-239. Human radiation experiments studying plutonium were conducted without
informed consent, and
several criticality accidents, some lethal, occurred during and after the war. Disposal of plutonium waste from
nuclear power plants and dismantled nuclear weapons built during the Cold War is a nuclear-proliferation and
environmental concern. Other sources of plutonium in the environment are fallout from numerous above-ground
nuclear tests (now banned).

Characteristics

Physical properties
Plutonium, like most metals, has a bright silvery appearance at first, much like nickel, but it oxidizes very quickly to
[5]
a dull gray, although yellow and olive green are also reported. At room temperature plutonium is in its α form
(alpha). This, the most common structural form of the element (allotrope), is about as hard and brittle as grey cast
iron unless it is alloyed with other metals to make it soft and ductile. Unlike most metals, it is not a good conductor
of heat or electricity. It has a low melting point (640 °C) and an unusually high boiling point (3,228 °C).
Alpha decay, the release of a high-energy helium nucleus, is the most common form of radioactive decay for
239 24 12
plutonium. A 5 kg mass of Pu contains about 12.5×10 atoms. With a half-life of 24,100 years, about 11.5×10
of its atoms decay each second by emitting a 5.157 MeV alpha particle. This amounts to 9.68 watts of power. Heat
[6]
produced by the deceleration of these alpha particles makes it warm to the touch.
Resistivity is a measure of how strongly a material opposes the flow of electric current. The resistivity of plutonium
at room temperature is very high for a metal, and it gets even higher with lower temperatures, which is unusual for
metals. This trend continues down to 100 K, below which resistivity rapidly decreases for fresh samples. Resistivity
then begins to increase with time at around 20 K due to radiation damage, with the rate dictated by the isotopic
composition of the sample.
Because of self-irradiation, a sample of plutonium fatigues throughout its crystal structure, meaning the ordered
arrangement of its atoms becomes disrupted by radiation with time. Self-irradiation can also lead to annealing which
counteracts some of the fatigue effects as temperature increases above 100 K.
Unlike most materials, plutonium increases in density when it melts, by 2.5%, but the liquid metal exhibits a linear
decrease in density with temperature. Near the melting point, the liquid plutonium has also very high viscosity and
surface tension as compared to other metals.
Allotrope
s
Main article: Allotropes of plutonium
Plutonium normally has six allotropes and
forms a seventh (zeta, ζ) at high temperature
within a limited pressure range. These
allotropes, which are different structural
modifications or forms of an element, have
very similar internal energies but significantly
varying densities and crystal structures. This
makes plutonium very sensitive to changes in
temperature, pressure, or chemistry, and
allows for dramatic volume changes following
phase transitions from one allotropic form to
another. The densities of the different
3 Plutonium has six allotropes at ambient pressure: alpha (α), beta (β),
allotropes vary from 16.00 g/cm to
3 gamma (γ), delta (δ), delta prime (δ'), & epsilon (ε)
19.86 g/cm .

The presence of these many allotropes makes


machining plutonium very difficult, as it changes
state very readily. For example, the α form exists at room temperature in unalloyed plutonium. It has machining
characteristics similar to cast iron but changes to the plastic and malleable β form (beta) at slightly higher
temperatures. The reasons for the complicated phase diagram are not entirely understood. The α form has a
low-symmetry monoclinic structure, hence its brittleness, strength, compressibility, and poor thermal
conductivity.

Plutonium in the δ form normally exists in the 310 °C to 452 °C range but is stable at room temperature when
alloyed with a small percentage of gallium, aluminium, or cerium, enhancing workability and allowing it to be
welded. The delta form has more typical metallic character, and is roughly as strong and malleable as
aluminium. In fission weapons, the explosive shock waves used to compress a plutonium core will also cause a
transition from the usual delta phase plutonium to the denser alpha form, significantly helping to achieve
supercriticality. The ε phase, the highest temperature solid allotrope, exhibits anomalously high atomic self-
diffusion compared to other elements.

Nuclear
fission
Plutonium is a radioactive actinide metal whose isotope, plutonium-239, is
one of the three primary fissile isotopes (uranium-233 and uranium-235 are
the other two); plutonium-241 is also highly fissile. To be considered fissile,
an isotope's atomic nucleus must be able to break apart or fission when struck
by a slow moving neutron and to release enough additional neutrons to
sustain the nuclear chain reaction by splitting further nuclei.
Pure plutonium-239 may have a multiplication factor (k ) larger than one,
eff tons of TNT (88,000 GJ). It is this
which means that if the metal is present in sufficient quantity and with an
energy that makes plutonium-239
appropriate geometry (e.g., a sphere of sufficient size), it can form a critical
useful in nuclear weapons and
mass. During fission, a fraction of the binding energy, which holds a nucleus
reactors.
together, is released as a large amount of electromagnetic and kinetic energy
(much of the latter being quickly converted to thermal energy). Fission of a
kilogram of plutonium-239 can produce an explosion equivalent to 21,000
core. The ring weighs 5.3 kg, is ca. 11
cm in diameter and its shape helps with
criticality safety.
A ring of weapons-grade 99.96% pure electrorefined plutonium, enough for one bomb
The presence of the isotope plutonium-240 in a sample limits its nuclear bomb potential, as plutonium-240 has a
relatively high spontaneous fission rate (~440 fissions per second per gram—over 1,000 neutrons per second per
[7]
gram), raising the background neutron levels and thus increasing the risk of predetonation. Plutonium is identified
as either weapons-grade, fuel grade, or power reactor grade based on the percentage of plutonium-240 that it
contains. Weapons-grade plutonium contains less than 7% plutonium-240. Fuel grade plutonium contains from 7%
to less than 19%, and power reactor grade contains 19% or more plutonium-240. Supergrade plutonium, with less
than 4% of plutonium-240, is used in U.S. Navy weapons stored in proximity to ship and submarine crews, due to its
[8]
lower radioactivity. The isotope plutonium-238 is not fissile but can undergo nuclear fission easily with fast
neutrons as well as alpha decay.

Isotopes and synthesis


Main article: Isotopes of plutonium
Twenty radioactive isotopes of
plutonium have been characterized.
The longest-lived are plutonium-244,
with a half-life of 80.8 million years,
plutonium-242, with a half-life of
373,300 years, and plutonium-239,
with a half-life of 24,110 years. All of
the remaining radioactive isotopes
have half-lives that are less than
7,000 years. This element also has
eight metastable states, though all
have half-lives less than one second.

The isotopes of plutonium range in


mass number from 228 to 247. The
primary decay modes of isotopes Uranium-plutonium and thorium-uranium chains

with mass numbers lower than the


most
stable isotope, plutonium-244, are spontaneous fission and α emission, mostly forming uranium (92 protons) and
neptunium (93 protons) isotopes as decay products (neglecting the wide range of daughter nuclei created by fission
processes). The primary decay mode for isotopes with mass numbers higher than plutonium-244 is β emission,
mostly forming americium (95 protons) isotopes as decay products. Plutonium-241 is the parent isotope of the
neptunium decay series, decaying to americium-241 via β or electron emission.

Plutonium-238 and 239 are the most widely synthesized isotopes. Plutonium-239 is synthesized via the following

reaction using uranium (U) and neutrons (n) via beta decay (β ) with neptunium (Np) as an intermediate:

Neutrons from the fission of uranium-235 are captured by uranium-238 nuclei to form uranium-239; a beta decay
converts a neutron into a proton to form Np-239 (half-life 2.36 days) and another beta decay forms plutonium-239.
Workers on the Tube Alloys project had predicted this reaction theoretically in 1940.
Plutonium-238 is synthesized by bombarding uranium-238 with deuterons (D, the nuclei of heavy hydrogen) in
the following reaction:

In this process, a deuteron hitting uranium-238 produces two neutrons and neptunium-238, which spontaneously
decays by emitting negative beta particles to form plutonium-238.
Decay heat and fission
properties
Plutonium isotopes undergo radioactive decay, which produces decay heat. Different isotopes produce different
amounts of heat per mass. The decay heat is usually listed as watt/kilogram, or milliwatt/gram. In case of larger
pieces of plutonium (e.g. a weapon pit) and inadequate heat removal the resulting self-heating may be significant.
All isotopes produce weak gamma on decay.

Decay heat of plutonium isotopes


Isotope Decay mode Half-life Decay Spontaneous fission Comment
(years) heat neutrons (1/(g·s))
(W/kg)

238 234 87.74 560 2600 Very high decay heat. Even in small amounts can cause significant
Pu alpha to U
self-heating. Used on its own in radioisotope thermoelectric
generators.

239 235 24100 1.9 0.022 The principal fissile isotope in use.
Pu alpha to U

240 236 6560 6.8 910 239


Pu alpha to U, The principal impurity in samples of the Pu isotope. The
spontaneous 240
plutonium grade is usually listed as percentage of Pu. High
fission spontaneous fission hinders use in nuclear weapons.

241 14.4 4.2 0.049 Decays to americium-241; its buildup presents a radiation hazard
Pu beta-minus, to
241 in older samples.
Am

242 238 376000 0.1 1700


Pu alpha to U

Americium-241, the decay product of plutonium-241, has half-life of 430 years, 1.2 spontaneous fissions per gram
per second, and decay heat of 114 watts per kilogram. As its decay produces highly penetrative gamma rays, its
presence in plutonium, determined by the original concentration of plutonium-241 and the sample age, increases the
radiation exposure of surrounding structures and personnel.

Compounds and
chemistry
At room temperature, pure plutonium is
silvery in color but gains a tarnish when
oxidized. The element displays four
common ionic oxidation states in aqueous
solution and one rare one:

• Pu(III), as Pu3+ (blue lavender)


• Pu(IV), as Pu4+ (yellow brown)
• Pu(V), as PuO+
[9]
2 (light pink)
• Pu(VI), as PuO2+
2 (pink orange)
• Pu(VII), as PuO3−
5 (green)–the heptavalent ion is rare
The color shown by plutonium solutions
Various oxidation states of plutonium in solution
depends on both the oxidation state and the
nature of the acid anion. It is the acid anion
that influences the degree of complexing—how atoms connect to a central atom—of the plutonium species.
Metallic plutonium is produced by reacting plutonium tetrafluoride with barium, calcium or lithium at 1200
°C. It is attacked by acids, oxygen, and steam but not by alkalis and dissolves easily in concentrated
hydrochloric, hydroiodic and perchloric acids. Molten metal must be kept in a vacuum or an inert
atmosphere to avoid reaction with air. At 135 °C the metal will ignite in air and will explode if placed in
carbon tetrachloride.
Plutonium is a reactive metal. In moist air or moist argon, the metal oxidizes
rapidly, producing a mixture of oxides and hydrides. If the metal is exposed long
enough to a limited amount of water vapor, a powdery surface coating of PuO is
2
formed. Also formed is plutonium hydride but an excess of water vapor forms
only PuO .
2
Plutonium shows enormous, and reversible, reaction rates with pure
Plutonium pyrophoricity can cause
hydrogen, forming plutonium hydride. It also reacts readily with oxygen,
it to look like a glowing ember
under forming PuO and
certain conditions. PuO as well as intermediate oxides; plutonium oxide fills 40% more volume
2
than plutonium metal. The metal reacts with the halogens, giving rise to
compounds with the general formula PuX where X can be F, Cl, Br or I and
3
PuF is also seen. The following oxyhalides are observed: PuOCl, PuOBr and
4
PuOI. It will react with carbon to form PuC, nitrogen to form PuN and silicon to
form PuSi .
2
Powders of plutonium, its hydrides and certain oxides like Pu are pyrophoric,
O
2 3
meaning they can ignite spontaneously at ambient temperature and are therefore
handled in an inert, dry atmosphere of nitrogen or argon. Bulk plutonium ignites
only when heated above 400 °C. Pu O spontaneously heats up and transforms
2 3
into PuO , which is stable in dry air, but reacts with water vapor when heated.
2
Crucibles used to contain plutonium need to be able to withstand its strongly
reducing properties. Refractory metals such as tantalum and tungsten along with
Twenty micrograms of pure
plutoniu hydroxide.
[10] the more stable oxides, borides, carbides, nitrides and silicides can tolerate this.
m Melting in an electric arc furnace can be used to produce small ingots of the
metal without the need for a crucible.
Cerium is used as a chemical simulant of plutonium for development of containment, extraction, and other
technologies.

Electronic structure
Plutonium is an element in which the 5f electrons are the transition border between delocalized and localized; it is
therefore considered one of the most complex elements. The anomalous behavior of plutonium is caused by its
electronic structure. The energy difference between the 6d and 5f subshells is very low. The size of the 5f shell is just
enough to allow the electrons to form bonds within the lattice, on the very boundary between localized and bonding
behavior. The proximity of energy levels leads to multiple low-energy electron configurations with near equal
n 2 n-1 2 1
energy levels. This leads to competing 5f 7s and 5f 7s 6d configurations, which causes the complexity of its
chemical behavior. The highly directional nature of 5f orbitals is responsible for directional covalent bonds in
molecules and complexes of plutonium.
Alloys
Plutonium can form alloys and intermediate compounds with most other metals. Exceptions include lithium, sodium,
potassium, rubidium and caesium of the alkali metals; and magnesium, calcium, strontium, and barium of the
alkaline earth metals; and europium and ytterbium of the rare earth metals. Partial exceptions include the refractory
metals chromium, molybdenum, niobium, tantalum, and tungsten, which are soluble in liquid plutonium, but
insoluble or only slightly soluble in solid plutonium. Gallium, aluminium, americium, scandium and cerium can
stabilize the δ phase of plutonium for room temperature. Silicon, indium, zinc and zirconium allow formation of
metastable δ state when rapidly cooled. High amounts of hafnium, holmium and thallium also allows retaining some
of the δ phase at room temperature. Neptunium is the only element that can stabilize the α phase at higher
temperatures.
Plutonium alloys can be produced by adding a metal to molten plutonium. If the alloying metal is sufficiently
reductive, plutonium can be added in the form of oxides or halides. The δ phase plutonium-gallium and
plutonium-aluminium alloys are produced by adding plutonium(III) fluoride to molten gallium or aluminium,
which has the advantage of avoiding dealing directly with the highly reactive plutonium metal.
• Plutonium-gallium is used for stabilizing the δ phase of plutonium, avoiding the α-phase and α-δ related
issues.
Its main use is in pits of implosion nuclear weapons.
• Plutonium-aluminium is an alternative to the Pu-Ga alloy. It was the original element considered for δ phase
stabilization, but its tendency to react with the alpha particles and release neutrons reduces its usability for nuclear
weapon pits. Plutonium-aluminium alloy can be also used as a component of nuclear fuel.
• Plutonium-gallium-cobalt alloy (PuCoGa ) is an unconventional superconductor, showing superconductivity
5
below 18.5 kelvin, an order of magnitude higher than the highest between heavy fermion systems, and has large
critical current.
• Plutonium-zirconium alloy can be used as nuclear fuel.[11]
• Plutonium-cerium and plutonium-cerium-cobalt alloys are used as nuclear fuels.
• Plutonium-uranium, with about 15–30 mol.% plutonium, can be used as a nuclear fuel for fast breeder reactors.
Its pyrophoric nature and high susceptibility to corrosion to the point of self-igniting or disintegrating after
exposure to air require alloying with other components. Addition of aluminium, carbon or copper did not improve
disintegration rates markedly, zirconium and iron alloys have better corrosion resistance but they disintegrate in
several months in air as well. Addition of titanium and/or zirconium significantly increases the melting point of
the alloy.
• Plutonium-uranium-titanium and plutonium-uranium-zirconium were investigated for use as nuclear fuels.
The addition of the third element increases corrosion resistance, reduces flammability, and improves ductility,
fabricability, strength, and thermal expansion. Plutonium-uranium-molybdenum has the best corrosion
resistance, forming a protective film of oxides, but titanium and zirconium are preferred for physics reasons.
• Thorium-uranium-plutonium was investigated as a nuclear fuel for fast breeder reactors.

Occurrence
Trace amounts of at least three plutonium isotopes (plutonium-238, 239, and 244) can be found in nature. Small
traces of plutonium-239, a few parts per trillion, and its decay products are naturally found in some
concentrated ores of uranium, such as the natural nuclear fission reactor in Oklo, Gabon. The ratio of
−12 −12
plutonium-239 to uranium at the Cigar Lake Mine uranium deposit ranges from 2.4×10 to 44×10 .
Even smaller amounts of primordial plutonium-244 occur naturally due to its relatively long half-life of about 80
239 238
million years. These trace amounts of Pu originate in the following fashion: On rare occasions, U
undergoes spontaneous fission, and in the process, the nucleus emits one or two free neutrons with some
238
kinetic energy. When one of these neutrons strikes the nucleus of another U atom, it is absorbed by the
239 239
atom, which becomes U. With a relatively short half-life, U-239 decays to neptunium-239 ( Np), and then
239 239
Np decays into Pu.
Since the relatively long-lived isotope plutonium-240 occurs in the decay chain of plutonium-244 it should also
be present, albeit 10,000 times rarer still. Finally, exceedingly small amounts of plutonium-238, attributed to
the extremely rare double beta decay of uranium-238, have been found in natural uranium samples.
Minute traces of plutonium are usually found in the human body due to the 550 atmospheric and underwater
nuclear tests that have been carried out, and to a small number of major nuclear accidents. Most atmospheric
and underwater nuclear testing was stopped by the Limited Test Ban Treaty in 1963, which was signed and
ratified by the United States, the United Kingdom, the Soviet Union, and other nations. Continued atmospheric
nuclear weapons testing since 1963 by non-treaty nations included those by China (atomic bomb test above
the Gobi Desert in 1964, hydrogen bomb test in 1967, and follow-on tests), and France (tests as recently as
the 1990s). Because it is deliberately manufactured for nuclear weapons and nuclear reactors, plutonium-239
is the most abundant isotope of plutonium by far.

History

Discovery
Enrico Fermi and a team of scientists at the University of Rome reported that they had discovered element 94 in
1934. Fermi called the element hesperium and mentioned it in his Nobel Lecture in 1938. The sample was actually a
mixture of barium, krypton, and other elements, but this was not known at the time because nuclear fission had not
been discovered yet.
The breakthrough with plutonium was at the Cavendish Laboratory, Cambridge by Egon Bretscher and Norman
Feather. They realized that a slow neutron reactor fuelled with uranium would theoretically produce substantial
amounts of plutonium-239 as a by-product. This is because U-238 absorbs slow neutrons and forms a new isotope
U-239. The new isotope's nucleus rapidly emits an electron through beta decay producing a new element with a mass
of 239 and an atomic number of 93. This element's nucleus then also emits an electron and becomes a new element
of mass 239 but with an atomic number 94 and a much greater half-life. Bretscher and Feather showed theoretically
feasible grounds that element 94 would be readily 'fissionable' by both slow and fast neutrons, and had the added
advantage of being chemically different from uranium, and could easily be separated from it.
This new development was also confirmed in independent work by Edwin M. McMillan and Philip Abelson at
Berkeley Radiation Laboratory also in 1940. Nicholas Kemmer of the Cambridge team proposed the names
neptunium for the new element 93 and plutonium for 94 by analogy with the outer planets Neptune and Pluto beyond
Uranus (uranium being element 92). The Americans fortuitously suggested the same names.
Plutonium (specifically, plutonium-238) was first produced and
isolated on December 14, 1940, and chemically identified on February
23, 1941, by Dr. Glenn T. Seaborg, Edwin M. McMillan, J. W.
Kennedy, and A. C. Wahl by deuteron bombardment of uranium in the
60-inch (150 cm) cyclotron at the University of California, Berkeley.
In the 1940 experiment, neptunium-238 was created directly by the
bombardment but decayed by beta emission with a half-life of a little
over two days, which indicated the formation of element 94.

A paper documenting the discovery was prepared by the team and


sent to the journal Physical Review in March 1941. The paper was
withdrawn before publication after the discovery that an isotope of
the new element (plutonium-239) could undergo nuclear fission in a
way that might be useful in an atomic bomb. Publication was
delayed until a year after the end of World War II due to security
Glenn T. Seaborg and his team at Berkeley
concerns. were the first to produce plutonium.

Edwin McMillan had recently named the first transuranium element


after the planet Neptune and suggested that element 94, being the next
[12]
element in the series, be named for what was then considered the next planet, Pluto. Seaborg originally
considered the name "plutium", but later thought that it did not sound as good as "plutonium." He chose the letters
[13]
"Pu" as a joke, which passed without notice into the periodic table. Alternative names considered by Seaborg and
others were "ultimium" or "extremium" because of the erroneous belief that they had found the last possible element
on the periodic table.

Early research
The basic chemistry of plutonium was found to resemble uranium after a few months of initial study. Early research
was continued at the secret Metallurgical Laboratory of the University of Chicago. On August 20, 1942, a trace
quantity of this element was isolated and measured for the first time. About 50 micrograms of plutonium-239
combined with uranium and fission products was produced and only about 1 microgram was isolated. This procedure
[14]
enabled chemists to determine the new element's atomic weight.
In November 1943 some plutonium trifluoride was reduced to create the first sample of plutonium metal: a few
micrograms of metallic beads. Enough plutonium was produced to make it the first synthetically made element
to be visible with the unaided eye.
The nuclear properties of plutonium-239 were also studied; researchers found that when it is hit by a neutron it
breaks apart (fissions) by releasing more neutrons and energy. These neutrons can hit other atoms of
plutonium-239 and so on in an exponentially fast chain reaction. This can result in an explosion large enough to
destroy a city if enough of the isotope is concentrated to form a critical mass.
Production during the Manhattan Project
During World War II the U.S. government established the Manhattan Project, which was tasked with developing an
atomic bomb. The three primary research and production sites of the project were the plutonium production facility
at what is now the Hanford Site, the uranium enrichment facilities at Oak Ridge, Tennessee, and the weapons
research and design laboratory, now known as Los Alamos National Laboratory.
The first production reactor that made plutonium-239 was the X-10
Graphite Reactor. It went online in 1943 and was built at a facility in
[15]
Oak Ridge that later became the Oak Ridge National Laboratory.
On April 5, 1944, Emilio Segrè at Los Alamos received the first
sample of reactor-produced plutonium from Oak Ridge. Within ten
days, he discovered that reactor-bred plutonium had a higher
concentration of the isotope plutonium-240 than cyclotron-produced
plutonium. Plutonium-240 has a high spontaneous fission rate, raising
the overall background neutron level of the plutonium sample. The
original gun-type plutonium weapon, code-named "Thin Man", had to The Hanford B Reactor face under
construction—the first plutonium-
be abandoned as a result—the increased number of spontaneous production reactor.
neutrons meant that nuclear pre-detonation (a fizzle) would be likely.

The entire plutonium weapon design effort at Los Alamos was soon
changed to the more complicated implosion device, code-named "Fat
Man." With an implosion weapon, a solid (or, in later designs, hollow)
sphere of plutonium is compressed to a high density with explosive
lenses—a technically more daunting task than the simple gun-type
design, but necessary to use plutonium for weapons purposes.
(Enriched uranium, by contrast, can be used with either method.)

Construction of the Hanford B Reactor, the first industrial-sized


nuclear reactor for the purposes of material production, was
The Hanford site represents two-thirds of the
completed in March 1945. B Reactor produced the fissile material
[16] nation's high-level radioactive waste by volume.
for the plutonium weapons used during World War II. B, D and F Nuclear reactors line the riverbank at the Hanford
were the initial reactors built at Hanford, and six additional Site along the Columbia River in January 1960.
plutonium-producing reactors were built later at the site.

In the 2013 book, Plutopia: Nuclear Families, Atomic Cities, and the Great Soviet and American Plutonium
Disasters (Oxford), Kate Brown explores the health of affected citizens in the United States, and the “slow-motion
disasters” that still threaten the environments where the plutonium production plants are located. According to
Brown, the plants at Hanford (and Mayak in the USSR), over a period of four decades, “both released more than 200
million curies of radioactive isotopes into the surrounding environment -- twice the amount expelled in the
Chernobyl disaster in each instance”. Most of this radioactive contamination over the years were part of normal
operations, but unforeseen accidents did occur and plant management kept this secret, as the pollution continued
unabated. Even today, as pollution threats to health and the environment persist, the government keeps knowledge
about the associated risks from the public.

In 2004, a safe was discovered during excavations of a burial trench at the Hanford nuclear site. Inside the safe were
various items, including a large glass bottle containing a whitish slurry which was subsequently identified as the
oldest sample of weapons-grade plutonium known to exist. Isotope analysis by Pacific Northwest National
Laboratory indicated that the plutonium in the bottle was manufactured in the X-10 reactor at Oak Ridge during
1944.
94 Plutonium 100

Trinity and Fat Man atomic bombs


The first atomic bomb test, codenamed "Trinity" and detonated on July
16, 1945, near Alamogordo, New Mexico, used plutonium as its fissile
material. The implosion design of "the gadget", as the Trinity device
was code-named, used conventional explosive lenses to compress a
sphere of plutonium into a supercritical mass, which was
simultaneously showered with neutrons from the "Urchin", an initiator
made of polonium and beryllium (neutron source: (α, n) reaction).
Together, these ensured a runaway chain reaction and explosion. The
overall weapon weighed over 4 tonnes, although it used just 6.2 kg of
plutonium in its core. About 20% of the plutonium used in the Trinity
weapon underwent fission, resulting in an explosion with an energy
[17]
equivalent to approximately 20,000 tons of TNT.

An identical design was used in the "Fat Man" atomic bomb dropped
on Nagasaki, Japan, on August 9, 1945, killing 70,000 people and Because of the presence of plutonium-240 in
reactor-bred plutonium, the implosion design was
wounding another 100,000. The "Little Boy" bomb dropped on
developed for the "Fat Man" and "Trinity"
Hiroshima three days earlier used uranium-235, not plutonium. Japan weapons
capitulated on August 15 to General Douglas MacArthur. Only after
the announcement of the first atomic bombs was the existence of plutonium made public.

Cold War use and waste


Large stockpiles of weapons-grade plutonium were built up by both the Soviet Union and the United States during
the Cold War. The U.S. reactors at Hanford and the Savannah River Site in South Carolina produced 103 tonnes, and
[18]
an estimated 170 tonnes of military-grade plutonium was produced in USSR. Each year about 20 tonnes of the
element is still produced as a by-product of the nuclear power industry. As much as 1000 tonnes of plutonium may
be in storage with more than 200 tonnes of that either inside or extracted from nuclear weapons. SIPRI estimated the
world plutonium stockpile in 2007 as about 500 tons, divided equally between weapon and civilian stocks.
Since the end of the Cold War these stockpiles have become a focus of
nuclear proliferation concerns. In the U.S., some plutonium extracted
from dismantled nuclear weapons is melted to form glass logs of
plutonium oxide that weigh two tonnes. The glass is made of
[19]
borosilicates mixed with cadmium and gadolinium. These logs are
planned to be encased in stainless steel and stored as much as 4 km
(2 mi) underground in bore holes that will be back-filled with concrete.
As of 2008, the only facility in the U.S. that was scheduled to store
plutonium in this way was the Yucca Mountain nuclear waste Proposed waste storage tunnel design for
repository, which is about 100 miles (160 km) north-east of Las Vegas, the Yucca Mountain nuclear waste
repository
Nevada. Local and state opposition to this plan delayed efforts to store
nuclear waste at Yucca Mountain. In March 2010, the Department of Energy withdrew its license application for the
Yucca Mountain repository "with prejudice" and eliminated funding for the Office of Civilian Radioactive Waste
Management, which had managed the Yucca Mountain site for 25 years, canceling the program.
94 Plutonium 1

Medical experimentation
See also: Human radiation experiments and Albert
Stevens
During and after the end of World War II, scientists working on the Manhattan Project and other nuclear weapons
research projects conducted studies of the effects of plutonium on laboratory animals and human subjects. Animal
studies found that a few milligrams of plutonium per kilogram of tissue is a lethal dose.
In the case of human subjects, this involved injecting solutions containing (typically) five micrograms of
plutonium into hospital patients thought to be either terminally ill, or to have a life expectancy of less than ten
years either due to age or chronic disease condition. This was reduced to one microgram in July 1945 after
animal studies found that the way plutonium distributed itself in bones was more dangerous than radium. Many
[citation needed]
of these experiments resulted in strong mutation. Most of the subjects, Eileen Welsome says, were
[20]
poor, powerless, and sick.
From 1945 to 1947, eighteen human test subjects were injected with plutonium without informed consent. The tests
were used to create diagnostic tools to determine the uptake of plutonium in the body in order to develop safety
standards for working with plutonium. Other experiments directed by the United States Atomic Energy Commission
and the Manhattan Project continued into the 1970s. The Plutonium Files chronicles the lives of the subjects of the
secret program by naming each person involved and discussing the ethical and medical research conducted in secret
by the scientists and doctors. The episode is now considered to be a serious breach of medical ethics and of the
Hippocratic Oath.
The government covered up most of these radiation mishaps until 1993, when President Bill Clinton ordered a
change of policy and federal agencies then made available relevant records. The resulting investigation was
undertaken by the president’s Advisory Committee on Human Radiation Experiments, and it uncovered much of the
material about plutonium research on humans. The committee issued a controversial 1995 report which said that
"wrongs were committed" but it did not condemn those who perpetrated them.

Applications

Explosives
The isotope plutonium-239 is a key fissile component in nuclear weapons, due to
its ease of fission and availability. Encasing the bomb's plutonium pit in a tamper
(an optional layer of dense material) decreases the amount of plutonium needed
to reach critical mass by reflecting escaping neutrons back into the plutonium
core. This reduces the amount of plutonium needed to reach criticality from
16 kg to 10 kg, which is a sphere with a diameter of about 10 centimeters (4 in).
This critical mass is about a third of that for uranium-235.

The "Fat Man"–type plutonium bombs produced during the Manhattan Project
used explosive compression of plutonium to obtain significantly higher
densities than normal, combined with a central neutron source to begin the
reaction and increase efficiency. Thus only 6.2 kg of plutonium was needed The atomic bomb dropped on
Nagasaki, Japan in 1945 had
for an explosive yield equivalent to 20 kilotons of TNT. (See also nuclear
a plutonium core.
weapon design.) Hypothetically, as little as 4 kg of plutonium—and maybe
even less—could be used to make a single atomic bomb using very
sophisticated assembly designs.
Mixed oxide fuel
Main article: Nuclear reprocessing
Spent nuclear fuel from normal light water reactors contains plutonium, but it is a mixture of plutonium-242, 240,
239 and 238. The mixture is not sufficiently enriched for efficient nuclear weapons, but can be used once as
MOX fuel. Accidental neutron capture causes the amount of plutonium-242 and 240 to grow each time the
plutonium is irradiated in a reactor with low-speed "thermal" neutrons, so that after the second cycle, the
plutonium can only be consumed by fast neutron reactors. If fast neutron reactors are not available (the normal
case), excess plutonium is usually discarded, and forms the longest-lived component of nuclear waste. The
desire to consume this plutonium and other transuranic fuels and reduce the radiotoxicity of the waste is the
usual reason nuclear engineers give to make fast neutron reactors.
The most common chemical process, PUREX (Plutonium–URanium EXtraction) reprocesses spent nuclear fuel
to extract plutonium and uranium which can be used to form a mixed oxide "MOX fuel" for reuse in nuclear
reactors. Weapons grade plutonium can be added to the fuel mix. MOX fuel is used in light water reactors and
consists of 60 kg of plutonium per tonne of fuel; after four years, three-quarters of the plutonium is burned
(turned into other elements). Breeder reactors are specifically designed to create more fissionable material than they
consume.
MOX fuel has been in use since the 1980s and is widely used in Europe. In September 2000, the United States and
the Russian Federation signed a Plutonium Management and Disposition Agreement by which each agreed to
[21]
dispose of 34 tonnes of weapon grade plutonium. The U.S. Department of Energy plans to dispose of 34 tonnes of
weapon grade plutonium in the United States before the end of 2019 by converting the plutonium to a MOX fuel to
be used in commercial nuclear power reactors.
MOX fuel improves total burnup. A fuel rod is reprocessed after three years of use to remove waste products,
which by then account for 3% of the total weight of the rods. Any uranium or plutonium isotopes produced
[22]
during those three years are left and the rod goes back into production. The presence of up to 1% gallium
per mass in weapon grade plutonium alloy has the potential to interfere with long-term operation of a light water
reactor.
Plutonium recovered from spent reactor fuel poses a less significant proliferation hazard, because of excessive
contamination with non-fissile plutonium-240 and plutonium-242. Separation of the isotopes is not feasible. A
dedicated reactor operating on very low burnup (hence minimal exposure of newly formed Pu-239 to additional
neutrons which causes it to be transformed to heavier isotopes of plutonium) is generally required to produce
material suitable for use in efficient nuclear weapons. While 'weapons-grade' plutonium is defined to contain at
least 92% plutonium-239 (of the total plutonium), the United States have managed to detonate an under-20Kt
device using plutonium believed to contain only about 85% plutonium-239, so called 'fuel-grade' plutonium. The
'reactor grade' plutonium produced by a regular LWR burnup cycle typically contains less than 60% Pu-239,
with up to 30% parasitic Pu-240/Pu-242, and 10–15% fissile Pu-241. It's unknown if a device using
plutonium obtained from reprocessed civil nuclear waste can be detonated, however such a device could
hypothetically fizzle and spread radioactive materials over a large urban area. The IAEA conservatively
classifies plutonium of all isotopic vectors as "direct-use" material, that is, "nuclear material that can be used
for the manufacture of nuclear explosives components without transmutation or further enrichment".
241
Am has recently been suggested for use as a denaturing agent in plutonium reactor fuel rods to further limit its
proliferation potential.
Power and heat source
The isotope plutonium-238 has a half-life of 87.74 years. It emits a
large amount of thermal energy with low levels of both gamma
rays/particles and spontaneous neutron rays/particles. Being an alpha
emitter, it combines high energy radiation with low penetration and
thereby requires minimal shielding. A sheet of paper can be used to
shield against the alpha particles emitted by plutonium-238. One
kilogram of the isotope can generate about 570 watts of heat.

These characteristics make it well-suited for electrical power


generation for devices which must function without direct maintenance
238
A glowing cylinder of PuO
for timescales approximating a human lifetime. It is therefore used in 2

radioisotope thermoelectric generators and radioisotope heater units


such as those in the Cassini, Voyager, and New Horizons space
probes.
The twin Voyager spacecraft were launched in 1977, each containing a 500 watt plutonium power source. Over 30
[23]
years later, each source is still producing about 300 watts which allows limited operation of each spacecraft. An
earlier version of the same technology powered five Apollo Lunar Surface Experiment Packages, starting with
Apollo 12 in 1969.
Plutonium-238 has also been used successfully to power artificial heart pacemakers, to reduce the risk of repeated
[24]
surgery. It has been largely replaced by lithium-based primary cells, but as of 2003[25] there were somewhere
between 50 and 100 plutonium-powered pacemakers still implanted and functioning in living patients. Plutonium-
238 was studied as a way to provide supplemental heat to scuba diving. Plutonium-238 mixed with beryllium
is used to generate neutrons for research purposes.

Precautions
See also: Plutonium in the environment

Toxicity
There are two aspects to the harmful effects of plutonium: the radioactivity and the heavy metal poison effects.
These are by no means different from many other heavy materials, such as uranium (natural, depleted and enriched),
or natural thorium, for example. The two scariest aspects of plutonium to the layman, that it is a primary fuel for
nuclear weapons and that it did not exist in any sensible amount before 1944, are not relevant.
Isotopes and compounds of plutonium are radioactive and accumulate in bone marrow. Contamination by plutonium
oxide has resulted from nuclear disasters and radioactive incidents, including military nuclear accidents where
nuclear weapons have burned. Studies of the effects of these smaller releases, as well as of the widespread radiation
poisoning sickness and death following the atomic bombings of Hiroshima and Nagasaki, have provided
considerable information regarding the dangers, symptoms and prognosis of radiation poisoning, which in the case of
the Japanese Hibakusha/survivors was largely unrelated to direct plutonium exposure.
During the decay of plutonium, three types of radiation are released—alpha, beta, and gamma. Alpha radiation can
travel only a short distance and cannot travel through the outer, dead layer of human skin. Beta radiation can
penetrate human skin, but cannot go all the way through the body. Gamma radiation can go all the way through the
[26]
body. Alpha, beta, and gamma radiation are all forms of ionizing radiation. Either acute or longer-term exposure
carries a danger of serious health outcomes including radiation sickness, genetic damage, cancer, and death. The
danger increases with the amount of exposure.
Even though alpha radiation cannot penetrate the skin, ingested or inhaled plutonium does irradiate internal
organs. The skeleton, where plutonium accumulates, and the liver, where it collects and becomes concentrated,
are at risk. Plutonium is not absorbed into the body efficiently when ingested; only 0.04% of plutonium oxide is
absorbed after ingestion. Plutonium absorbed by the body is excreted very slowly, with a biological half-life
of 200 years. Plutonium passes only slowly through cell membranes and intestinal boundaries, so absorption
by ingestion and incorporation into bone structure proceeds very slowly.
Plutonium is more dangerous when inhaled than when ingested. The risk of lung cancer increases once the total
radiation dose equivalent of inhaled plutonium exceeds 400 mSv. The U.S. Department of Energy estimates that the
lifetime cancer risk from inhaling 5,000 plutonium particles, each about 3 µm wide, to be 1% over the background
U.S. average. Ingestion or inhalation of large amounts may cause acute radiation poisoning and death; no human is
known to have died because of inhaling or ingesting plutonium, and many people have measurable amounts of
plutonium in their bodies.
The "hot particle" theory in which a particle of plutonium dust radiates a localized spot of lung tissue is not
supported by mainstream research — such particles are more mobile than originally thought and toxicity is not
measurably increased due to particulate form.
When inhaled, plutonium can pass into the bloodstream. Once in the bloodstream, plutonium moves throughout the
body and into the bones, liver, or other body organs. Plutonium that reaches body organs generally stays in the body
for decades and continues to expose the surrounding tissue to radiation and thus may cause cancer.
A commonly cited quote by Ralph Nader states that a pound of plutonium dust spread into the atmosphere
would be enough to kill 8 billion people. However, calculations show that one pound of plutonium could kill no
more than 2 million people by inhalation. This makes the toxicity of plutonium roughly equivalent with that of
[27]
nerve gas. Nader's views were challenged in 1976 by Bernard Cohen, as described in the book Nuclear
Power, Both Sides: The Best Arguments for and Against the Most Controversial Technology. Cohen's own
estimate is that a dose of 200 micrograms would likely be necessary to cause cancer.
Several populations of people who have been exposed to plutonium dust (e.g. people living down-wind of
Nevada test sites, Nagasaki survivors, nuclear facility workers, and "terminally ill" patients injected with Pu in
1945–46 to study Pu metabolism) have been carefully followed and analyzed. These studies generally do not
show especially high plutonium toxicity or plutonium-induced cancer results, such as Albert Stevens who
survived into old age after being injected with plutonium. "There were about 25 workers from Los Alamos
National Laboratory who inhaled a considerable amount of plutonium dust during 1940s; according to the hot-
particle theory, each of them has a 99.5% chance of being dead from lung cancer by now, but there has not
been a single lung cancer among them."
Plutonium has a metallic taste.
Criticality potential
Toxicity issues aside, care must be taken to avoid the accumulation of
amounts of plutonium which approach critical mass, particularly
because plutonium's critical mass is only a third of that of uranium-
235. A critical mass of plutonium emits lethal amounts of neutrons and
gamma rays. Plutonium in solution is more likely to form a critical
mass than the solid form due to moderation by the hydrogen in water.

Criticality accidents have occurred in the past, some of them with


lethal consequences. Careless handling of tungsten carbide bricks
A sphere of simulated plutonium surrounded by
around a 6.2 kg plutonium sphere resulted in a fatal dose of radiation at neutron-reflecting tungsten carbide blocks in a
Los Alamos on August 21, 1945, when scientist Harry K. Daghlian, Jr. re-enactment of Harry Daghlian's 1945
received a dose estimated to be 5.1 Sievert (510 rems) and died experiment

25 days later. Nine months later, another Los Alamos scientist, Louis
Slotin, died from a similar accident involving a beryllium reflector and the same plutonium core (the so-called
"demon core") that had previously claimed the life of Daghlian. These incidents were fictionalized in the 1989
film Fat Man and Little Boy.

In December 1958, during a process of purifying plutonium at Los Alamos, a critical mass was formed in a mixing
vessel, which resulted in the death of a chemical operator named Cecil Kelley. Other nuclear accidents have occurred
in the Soviet Union, Japan, the United States, and many other countries.

Flammability
Metallic plutonium is a fire hazard, especially if the material is finely divided. In a moist environment, plutonium
forms hydrides on its surface, which are pyrophoric and may ignite in air at room temperature. Plutonium
expands up to 70% in volume as it oxidizes and thus may break its container. The radioactivity of the burning
material is an additional hazard. Magnesium oxide sand is probably the most effective material for extinguishing
a plutonium fire. It cools the burning material, acting as a heat sink, and also blocks off oxygen. Special
precautions are necessary to store or handle plutonium in any form; generally a dry inert gas atmosphere is
[28]
required.

Transportation

Air
The U.S. Government air transport regulations permit the transport of plutonium by air, subject to restrictions on
other dangerous materials carried on the same flight, packaging requirements, and stowage in the rearmost part of
the aircraft.
In 2012 media revealed that plutonium has been flown out of Norway on commercial passenger airlines—around
every other year—including one time in 2011. Regulations permit an airplane to transport 15 grams of fissionable
material. Such plutonium transportation is without problems, according to a Senior Advisor (seniorrådgiver) at
Statens strålevern.
Notes

Footnotes
[1] Magnetic susceptibility of the elements and inorganic compounds (http://www-d0.fnal. gov/ hardware/cal/ lvps_info/engineering/
elementmagn. pdf), in
[2] Magurno, B.A.; Pearlstein, S. (eds.) Workshop on nuclear data evaluation methods and procedures
(http://www.osti.gov/bridge//product. biblio.jsp?query_id=0&page=0&osti_id=5972980), Upton, NY, USA, 22 Septmber 1980, vol. II
(1981), pp. 835 ff
[3] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_plutonium&action=edit
[4] "Contaminated Water Escaping Nuclear Plant, Japanese Regulator Warns" (http:// www.nytimes.com/2011/03/29/world/asia/ 29japan.
html). The New York Times.
[5] (public domain text)
[6] Leona Marshall: "When you hold a lump of it in your hand, it feels warm, like a live rabbit"
[7] Samuel Glasstone and Leslie M. Redman, An Introduction to Nuclear Weapons (http://www.doeal. gov/opa/ docs/RR00171.pdf) (Atomic
Energy Commission Division of Military Applications Report WASH-1038, June 1972), p. 12.
[8] (public domain text)
4+ 4+
[9] The ion is unstable in solution and will disproportionate into Pu and ; the Pu will then oxidize the remaining to , being reduced in turn to
3+ 3+
Pu . Thus, aqueous solutions of tend over time towards a mixture of Pu and . Uranium#Aqueous chemistry is unstable for the same reason.
[10] Pure plutonium hydroxide in capillary tube (http:/ /imglib.lbl.gov/ImgLib/COLLECTIONS/BERKELEY-LAB/
RESEARCH-1930-1990/NUCLEAR-PHYSICS/TRANSURANIUM-ELEMENTS /index/ 96602765.html), LBNL Image Library
[11] McCuaig, Franklin D. "Pu-Zr alloy for high-temperature foil-type fuel" , Issued on November 22, 1977
[12] This was not the first time somebody suggested that an element be named "plutonium." A decade after barium was discovered, a Cambridge
University professor suggested it be renamed to "plutonium" because the element was not (as suggested by the Greek root, barys, it was
named for) heavy. He reasoned that, since it was produced by the relatively new technique of electrolysis, its name should refer to fire. Thus
he suggested it be named for the Roman god of the underworld, Pluto.
[13] As one article puts it, referring to information Seaborg gave in a talk: "The obvious choice for the symbol would have been Pl, but
facetiously, Seaborg suggested Pu, like the words a child would exclaim, 'Pee-yoo!' when smelling something bad. Seaborg thought that he
would receive a great deal of flak over that suggestion, but the naming committee accepted the symbol without a word."
[14] Room 405 of the George Herbert Jones Laboratory, where the first isolation of plutonium took place, was named a National Historic
Landmark in May 1967.
[15] During the Manhattan Project, plutonium was also often referred to as simply "49": the number 4 was for the last digit in 94 (atomic
number of plutonium), and 9 was for the last digit in plutonium-239, the weapon-grade fissile isotope used in nuclear bombs.
[16] The American Society of Mechanical Engineers (ASME) established B Reactor as a National Historic Mechanical Engineering Landmark in
September 1976. In August 2008, B Reactor was designated a U.S. National Historic Landmark.
[17] The efficiency calculation is based on the fact that 1 kg of plutonium-239 (or uranium-235) fissioning results in an energy release of
approximately 17 kt, leading to a rounded estimate of 1.2 kg plutonium actually fissioned to produce the 20 kt yield. On the figure of 1 kg =
17 kt,
[18] Much of this plutonium was used to make the fissionable cores of a type of thermonuclear weapon employing the Teller–Ulam design.
These so-called 'hydrogen bombs' are a variety of nuclear weapon that use a fission bomb to trigger the nuclear fusion of heavy hydrogen
isotopes. Their destructive yield is commonly in the millions of tons of TNT equivalent compared with the thousands of tons of TNT
equivalent of fission-only devices.
[19] Gadolinium zirconium oxide () has been studied because it could hold plutonium for up to 30 million years.
[20] R.C. Longworth. Injected! Book review:The Plutonium Files: America's Secret Medical Experiments in the Cold War (http:// intl-bos.
sagepub. com/content/ 55/6/58.full.pdf+html), The Bulletin of the Atomic Scientists, Nov/Dec 1999, 55(6): 58-61.
[21] (public domain text)
[22] Breakdown of plutonium in a spent nuclear fuel rod: plutonium-239 (~58%), 240 (24%), 241 (11%), 242 (5%), and 238 (2%).
[23] Voyager-Spacecraft Lifetime (http:// voyager.jpl.nasa.gov/spacecraft/spacecraftlife.html)
[24] Defunct pacemakers with Pu power source (http:// www.orau. org/ptp/collection/Miscellaneous/pacemaker. htm)
[25] http://en.wikipedia.org/w/index.php?title=Plutonium&action=edit
[26] Plutonium (http://www.atsdr.cdc. gov/substances/ toxsubstance. asp?toxid=119), CAS ID #: 7440-07-5, Centers for Disease Control and
Prevention (CDC) Agency for Toxic Substances and Disease Registry
[27] (Online version of Cohen's book The Nuclear Energy Option (Plenum Press, 1990) ISBN 0-306-
43567-5). [28] There was a major plutonium-initiated fire at the Rocky Flats Plant near Boulder, Colorado
in 1969.
Citations

References
• CRC contributors (2006). David R. Lide, ed. Handbook of Chemistry and Physics (87th ed.). Boca Raton (FL):
CRC Press, Taylor & Francis Group. ISBN 0-8493-0487-3.
• Emsley, John (2001). "Plutonium". Nature's Building Blocks: An A–Z Guide to the Elements. Oxford (UK):
Oxford University Press. pp. 324–329. ISBN 0-19-850340-7.
• Greenwood, N. N.; Earnshaw, A. (1997). Chemistry of the Elements (2nd ed.). Oxford (UK):
Butterworth-Heinemann. ISBN 0-7506-3365-4.
• Heiserman, David L. (1992). "Element 94: Plutonium". Exploring Chemical Elements and their Compounds.
New York (NY): TAB Books. pp. 337–340. ISBN 0-8306-3018-X.
• Miner, William N.; Schonfeld, Fred W. (1968). "Plutonium". In Clifford A. Hampel (editor). The Encyclopedia
of the Chemical Elements. New York (NY): Reinhold Book Corporation. pp. 540–546. LCCN 68-29938
(http:// lccn.loc.gov/68-29938) Check |lccn= value (help).
• Stwertka, Albert (1998). "Plutonium". Guide to the Elements (Revised ed.). Oxford (UK): Oxford University
Press. ISBN 0-19-508083-1.

External links
• Sutcliffe, W.G.; et al. (1995). "A Perspective on the Dangers of Plutonium"
(http://web.archive.org/web/ 20060929015050/http://www.llnl.gov/csts/publications/sutcliffe/). Lawrence
Livermore National Laboratory. Archived from the original
(http://www.llnl.gov/csts/publications/sutcliffe/) on September 29, 2006.
• Johnson, C.M.; Davis, Z.S. (1997). "Nuclear Weapons: Disposal Options for Surplus Weapons-Usable
Plutonium" (http://www.globalsecurity.org/wmd/library/report/crs/97-564.htm). CRS Report for Congress #
97-564 ENR. Retrieved February 15, 2009.
• IEER contributors (2005). "Physical, Nuclear, and Chemical, Properties of Plutonium" (http://www.ieer.org/
fctsheet/ pu-props. html). IEER. Retrieved February 15, 2009.
• Bhadeshia, H. "Plutonium crystallography" (http://www.msm.cam.ac.uk/phase-trans/2006/Plutonium/
Plutonium.html).
• Samuels, D. (2005). "End of the Plutonium Age" (http://discovermagazine.com/2005/nov/end-of-plutonium).
Discover Magazine 26 (11).
• Pike, J.; Sherman, R. (2000). "Plutonium production" (http://www.fas.org/nuke/intro/nuke/plutonium.htm).
Federation of American Scientists. Retrieved February 15, 2009.
• Nuclear Weapon Archive contributors. "Plutonium Manufacture and Fabrication" (http://nuclearweaponarchive.
org/Library/Plutonium/). Nuclearweaponarchive.org.
• Ong, C. (1999). "World Plutonium Inventories" (http://www.nuclearfiles.org/menu/key-issues/
nuclear-energy/ issues/world-plutonium-inventories-ong.htm). Nuclear Files.org. Retrieved February 15, 2009.
• LANL contributors (2000). "Challenges in Plutonium Science"
(http://www.fas.org/sgp/othergov/doe/lanl/ pubs/number26.htm). Los Alamos Science. I & II (26).
Retrieved February 15, 2009.
• NLM contributors. "Plutonium, Radioactive" (http://toxnet.nlm.nih.gov/cgi-bin/sis/search/r?dbs+
hsdb:@term+@na+@rel+plutonium,+radioactive). NLM Hazardous Substances Databank. Retrieved February
15, 2009.
• Alsos contributors. "Annotated Bibliography on plutonium" (http://alsos.wlu.edu/qsearch. aspx?
browse=science/Plutonium). Alsos Digital Library for Nuclear Issues. Retrieved February 15, 2009.
• Chemistry in its element podcast (http://www.rsc.org/chemistryworld/podcast/element.asp) (MP3) from the
Royal Society of Chemistry's Chemistry World: Plutonium (http://www.rsc.org/images/
CIIE_plutonium_48kbps_tcm18-121120 .MP3)
• Plutonium (http://www.periodicvideos.com/videos/094.htm) at The Periodic Table of Videos (University
of Nottingham)

95 – Americium
Americium
Am
95

plutonium ← americium → curiumEu



Am

(Uqs)

Americium in the periodic table


Appearance

silvery white

General properties

Name, symbol, number americium, Am, 95

Pronunciation /ˌæməˈrɪsiəm/
AM-ə-RIS-ee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (243)

Electron configuration 7 2
[Rn] 5f 7s
2, 8, 18, 32, 25, 8, 2

Physical properties
95 Americium 1

Phase solid

Density (near r.t.) −3


12 g·cm

Melting point 1449 K, 1176 °C, 2149 °F

Boiling point 2880 K, 2607 °C, 4725 °F

Heat of fusion −1
14.39 kJ·mol

Molar heat capacity −1 −1


62.7 J·mol ·K

Vapor pressure

P (Pa) 1 10 100 1 k 10 k 100 k

at T (K) 1239 1356

Atomic properties

Oxidation states 7, 6, 5, 4, 3, 2
(amphoteric oxide)

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 578 kJ·mol

Atomic radius 173 pm

Covalent radius 180±6 pm

Miscellanea

Crystal structure hexagonal

Magnetic ordering paramagnetic

Electrical resistivity 0.69 µΩ·m

Thermal conductivity −1 −1
10 W·m ·K

CAS registry number 7440-35-9

History

Naming after the Americas

Discovery Glenn T. Seaborg, Ralph A. James, Leon O. Morgan, Albert Ghiorso (1944)

Most stable isotopes

Main article: Isotopes of americium


iso NA half-life DM DE (MeV) DP

241 trace 432.2 y SF - -


Am
α 5.486 237
Np

242m trace 141 y IT 0.049 242


Am Am
α 5.637 238
Np
SF - -

243 trace 7370 y SF - -


Am
α 5.275 239
Np

• v
• t
[1]
• e

Americium (pronounced AM-ə-RISH-ee-əm) is a transuranic radioactive chemical element that has the symbol Am
and atomic number 95. This transuranic element of the actinide series is located in the periodic table below the
lanthanide element europium, and thus by analogy was named after another continent, America.
Americium was first produced in 1944 by the group of Glenn T. Seaborg at the University of California, Berkeley.
Although it is the third element in the transuranic series, it was discovered fourth, after the heavier curium. The
discovery was kept secret and only released to the public in November 1945. Most americium is produced by
bombarding uranium or plutonium with neutrons in nuclear reactors – one tonne of spent nuclear fuel contains about
100 grams of americium. It is widely used in commercial ionization chamber smoke detectors, as well as in neutron
sources and industrial gauges. Several unusual applications, such as a nuclear battery or fuel for space ships with
242m
nuclear propulsion, have been proposed for the isotope Am, but they are as yet hindered by the scarcity and high
price of this nuclear isomer.
241
Americium is a relatively soft radioactive metal with silvery appearance. Its most common isotopes are Am
243
and Am. In chemical compounds, they usually assume the oxidation state +3, especially in solutions. Several
other oxidation states are known, which range from +2 to +7 and can be identified by their characteristic optical
absorption spectra. The crystal lattice of solid americium and its compounds contains intrinsic defects, which
are induced by self-irradiation with alpha particles and accumulate with time; this results in a drift of some
material properties.
History
Although americium was likely produced in previous nuclear
experiments, it was first intentionally synthesized, isolated and
identified in late autumn 1944, at the University of California,
Berkeley, by Glenn T. Seaborg, Leon O. Morgan, Ralph A. James, and
Albert Ghiorso. They used a 60-inch cyclotron at the University of
[2]
California, Berkeley. The element was chemically identified at the
Metallurgical Laboratory (now Argonne National Laboratory) of the
University of Chicago. Following the lighter neptunium, plutonium,
and heavier curium, americium was the fourth transuranium element to
The 60-inch cyclotron at the Lawrence Radiation be discovered. At the time, the periodic table had been restructured by
Laboratory, University of California, Berkeley, in Seaborg to its present layout, containing the actinide row below the
August 1939. lanthanide one. This led to americium being located right below its
twin lanthanide element europium; it was thus by analogy named after
another continent, America: "The name americium (after the Americas) and the symbol Am are suggested for
the element on the basis of its position as the sixth member of the actinide rare-earth series, analogous to
[3][4]
europium, Eu, of the lanthanide series."

The new element was isolated from its oxides in a complex, multi-step process. First plutonium-239 nitrate
239 2
( PuNO ) solution was coated on a platinum foil of about 0.5 cm area, the solution was evaporated and the
3
residue was converted into plutonium dioxide (PuO ) by annealing. After cyclotron irradiation, the coating was
2
dissolved with nitric acid, and then precipitated as the hydroxide using concentrated aqueous ammonia solution. The
residue was dissolved in perchloric acid. Further separation was carried out by ion exchange, yielding a certain
isotope of curium. The separation of curium and americium was so painstaking that those elements were initially
called by the Berkeley group as pandemonium (from Greek for all demons or hell) and delirium (from Latin for
madness).
241 242 239 238
Initial experiments yielded four americium isotopes: Am, Am, Am and Am. Americium-241 was
237
directly obtained from plutonium upon absorption of one neutron. It decays by emission of a α-particle to Np;
the half-life of this decay was first determined as 510 ± 20 years but then corrected to 432.2 years.

The times are half-lives


242 241
The second isotope Am was produced upon neutron bombardment of the already-created Am. Upon rapid
242 242
β-decay, Am converts into the isotope of curium Cm (which had been discovered previously). The half-life of
this decay was initially determined at 17 hours, which was close to the presently accepted value of 16.02 h.

The discovery of americium and curium in 1944 was closely related to the Manhattan Project; the results were
confidential and declassified only in 1945. Seaborg leaked the synthesis of the elements 95 and 96 on the U.S. radio
show for children, the Quiz Kids, five days before the official presentation at an American Chemical Society meeting
on 11 November 1945, when one of the listeners asked whether any new transuranium element beside plutonium and
241 242
neptunium had been discovered during the war. After the discovery of americium isotopes Am and Am, their
[5]
production and compounds were patented listing only Seaborg as the inventor. The initial americium samples
weighed a few micrograms; they were barely visible and were identified by their radioactivity. The first substantial
amounts of metallic americium weighing 40–200 micrograms were not prepared until 1951 by reduction of
americium(III) fluoride with barium metal in high vacuum at 1100 °C.
Occurrence
See also: Nuclear
reprocessing
241
The longest-lived and most common isotopes of americium, Am
243
and Am, have half-lives of 432.2 and 7,370 years, respectively.
Therefore, all primordial americium (americium that was present on
Earth during its formation) should have decayed by now.
Existing americium is concentrated in the areas used for the
atmospheric nuclear weapons tests conducted between 1945 and
1980, as well as at the sites of nuclear incidents, such as the
Chernobyl disaster. For example, the analysis of the debris at the
testing site of the first U.S. hydrogen bomb, Ivy Mike, (1 November Americium was detected in the fallout from the
1952, Enewetak Atoll), revealed high concentrations of various Ivy Mike nuclear test.
actinides including americium; due to military secrecy, this result
was published only in 1956. Trinitite, the glassy residue left on
the desert floor near
Alamogordo, New Mexico, after the plutonium-based Trinity nuclear bomb test on 16 July 1945, contains traces of
americium-241. Elevated levels of americium were also detected at the crash site of a US B-52 bomber, which
carried four hydrogen bombs, in 1968 in Greenland.

In other regions, the average radioactivity of surface soil due to residual americium is only about 0.01 picocuries/g
(0.37 mBq/g). Atmospheric americium compounds are poorly soluble in common solvents and mostly adhere to soil
particles. Soil analysis revealed about 1,900 times higher concentration of americium inside sandy soil particles than
[6]
in the water present in the soil pores; an even higher ratio was measured in loam soils.
Americium is produced mostly artificially in small quantities, for research purposes. A tonne of spent nuclear fuel
241 243 [7]
contains about 100 grams of various americium isotopes, mostly Am and Am. Their prolonged radioactivity
is undesirable for the disposal, and therefore americium, together with other long-lived actinides, have to be
neutralized. The associated procedure may involve several steps, where americium is first separated and then
converted by neutron bombardment in special reactors to short-lived nuclides. This procedure is well known as
[8][9]
nuclear transmutation, but it is still being developed for americium.
A few atoms of americium can be produced by neutron capture reactions and beta decay in very highly concentrated
uranium-bearing deposits.
Synthesis and extraction

Isotope nucleosyntheses
Americium has been produced in small quantities in nuclear reactors
241 243
for decades, and kilograms of its Am and Am isotopes have been
[10]
accumulated by now. Nevertheless, since it was first offered for sale
241
in 1962, its price, about 1,500 USD per gram of Am, remains almost
[11]
unchanged owing to the very complex separation procedure. The
243
heavier isotope Am is produced in much smaller amounts; it is thus
more difficult to separate, resulting in a higher cost of the order
[12]
100,000–160,000 USD/g.

Americium is not synthesized directly from uranium – the most


239
common reactor material – but from the plutonium isotope Pu. The
latter needs to be produced first, according to the following nuclear
process:

Chromatographic elution curves revealing the


similarity between the lanthanides Tb, Gd, and
Eu and the corresponding actinides Bk, Cm, and
Am.

239 241
The capture of two neutrons by Pu (a so-called (n,γ) reaction), followed by a β-decay, results in Am:

241
The plutonium present in spent nuclear fuel contains about 12% of Pu. Because it spontaneously converts to
241 241 241
Am, Pu can be extracted and may be used to generate further Am. However, this process is rather slow: half
241 241 241
of the original amount of Pu decays to Am after about 15 years, and the Am amount reaches a maximum
[13]
after 70 years.
241
The obtained Am can be used for generating heavier americium isotopes by further neutron capture inside a
241 242
nuclear reactor. In a light water reactor (LWR), 79% of Am converts to Am and 10% to its nuclear isomer
242m [14][15]
Am:

79%:

10%:
243
Americium-242 has a half-life of only 16 hours, which makes its further up-conversion to Am, extremely
239
inefficient. The latter isotope is produced instead in a process where Pu captures four neutrons under high neutron
flux:
Metal generation
Most synthesis routines yield a mixture of different actinide isotopes in oxide forms, from which isotopes of
americium need to be separated. In a typical procedure, the spent reactor fuel (e.g. MOX fuel) is dissolved in nitric
acid, and the bulk of uranium and plutonium is removed using a PUREX-type extraction (Plutonium –URanium
EXtraction) with tributyl phosphate in a hydrocarbon. The lanthanides and remaining actinides are then separated
from the aqueous residue (raffinate) by a diamide-based extraction, to give, after stripping, a mixture of trivalent
actinides and lanthanides. Americium compounds are then selectively extracted using multi-step chromatographic
[16]
and centrifugation techniques with an appropriate reagent. A large amount of work has been done on the solvent
extraction of americium. For example, a recent EU funded project codenamed "EUROPART" studied triazines and
other compounds as potential extraction agents. Bis-triazinyl bipyridine complex has been recently proposed as such
reagent as highly selective to americium (and curium). Separation of americium from the highly similar curium can
be achieved by treating a slurry of their hydroxides in aqueous sodium bicarbonate with ozone, at elevated
temperatures. Both Am and Cm are mostly present in solutions in the +3 valence state; whereas curium remains
[17]
unchanged, americium oxidizes to soluble Am(IV) complexes which can be washed away.
Metallic americium is obtained by reduction from its compounds. Americium(III) fluoride was first used for this
purpose. The reaction was conducted using elemental barium as reducing agent in a water- and oxygen-free
[18][19]
environment inside an apparatus made of tantalum and tungsten.

An alternative is the reduction of americium dioxide by metallic lanthanum or thorium:

Physical properties
In the periodic table, americium is located to the right of plutonium, to
the left of curium, and below the lanthanide europium, with which it
shares many similarities in physical and chemical properties.
Americium is a highly radioactive element. When freshly prepared, it
has a silvery-white metallic lustre, but then slowly tarnishes in air.
3
With a density of 12 g/cm , americium is less dense than both curium
3 3
(13.52 g/cm ) and plutonium (19.8 g/cm ); but has a higher density
3
than europium (5.264 g/cm )—mostly because of its higher atomic
mass. Americium is relatively soft and easily deformable and has a
significantly lower bulk modulus than the actinides before it: Th, Pa,
U, Np and Pu. Its melting point of 1173 °C is significantly higher
Double-hexagonal close packing with the layer
than that of plutonium (639 °C) and europium (826 °C), but lower sequence ABAC in the crystal structure of
than for curium (1340 °C). α-americium (A: green, B: blue, C: red).

At ambient conditions, americium is present in its most stable α form


which has a hexagonal crystal symmetry, and a space group P6 /mmc with lattice parameters a = 346.8 pm and c =
3
1124 pm, and four atoms per unit cell. The crystal consists of a double-hexagonal close packing with the layer
sequence ABAC and so is isotypic with α-lanthanum and several actinides such as α-curium. The crystal structure of
americium changes with pressure and temperature. When compressed at room temperature to 5 GPa, α-Am
transforms to the β modification, which has a face-centered cubic (fcc) symmetry, space group Fm3m and lattice
constant a = 489 pm. This fcc structure is equivalent to the closest packing with the sequence ABC. Upon further
compression to 23 GPa, americium transforms to an orthorhombic γ-Am structure similar to that of α-uranium.
There are no further transitions observed up to 52 GPa, except for an appearance of a monoclinic phase at pressures
between 10 and 15 GPa. There is no consistency on the status of this phase in the literature, which also sometimes
lists the α, β and γ phases as I, II and III. The β-γ transition is accompanied by a 6% decrease in the crystal volume;
although theory also predicts a significant volume change for the α-β transition, it is not observed experimentally.
The pressure of the α-β transition decreases with increasing temperature, and when α-americium is heated at ambient
pressure, at 770 °C it changes into an fcc phase which is different from β-Am, and at 1075 °C it converts to a
body-centered cubic structure. The pressure-temperature phase diagram of americium is thus rather similar to those
of lanthanum, praseodymium and neodymium.
As with many other actinides, self-damage of the crystal lattice due to alpha-particle irradiation is intrinsic to
americium. It is especially noticeable at low temperatures, where the mobility of the produced lattice defects is
relatively low, by broadening of X-ray diffraction peaks. This effect makes somewhat uncertain the temperature of
americium and some of its properties, such as electrical resistivity. So for americium-241, the resistivity at 4.2 K
increases with time from about 2 µOhm·cm to 10 µOhm·cm after 40 hours, and saturates at about 16 µOhm·cm after
140 hours. This effect is less pronounced at room temperature, due to annihilation of radiation defects; also heating
to room temperature the sample which was kept for hours at low temperatures restores its resistivity. In fresh
samples, the resistivity gradually increases with temperature from about 2 µOhm·cm at liquid helium to 69 µOhm·cm
at room temperature; this behavior is similar to that of neptunium, uranium, thorium and protactinium, but is
different from plutonium and curium which show a rapid rise up to 60 K followed by saturation. The room
temperature value for americium is lower than that of neptunium, plutonium and curium, but higher than for
uranium, thorium and protactinium.
Americium is paramagnetic in a wide temperature range, from that of liquid helium, to room temperature, and above.
This behavior is markedly different from that of its neighbor curium which exhibit antiferromagnetic transition at
−6
52 K. The thermal expansion coefficient of americium is slightly anisotropic and amounts to (7.5 ± 0.2)×10 /°C
−6
along the shorter a axis and (6.2 ± 0.4)×10 /°C for the longer c hexagonal axis. The enthalpy of dissolution of
americium metal in hydrochloric acid at standard conditions is −620.6 ± 1.3 kJ/mol, from which the standard
3+ −1
enthalpy change of formation (Δ fH°) of aqueous Am ion is −621.2 ± 2.0 kJ/mol . The standard potential
3+ 0
Am /Am is 2.08 ± 0.01 V.

Chemical properties
Americium readily reacts with oxygen and dissolves well in acids. The most common oxidation state for americium
[20]
is +3, in which americium compounds are rather stable against oxidation and reduction. In this sense, americium
is chemically similar to most lanthanides. The trivalent americium forms insoluble fluoride, oxalate, iodate,
hydroxide, phosphate and other salts. Other oxidation states have been observed between +2 and +7, which is the
3+
widest range among the actinide elements. Their color in aqueous solutions varies as follows: Am (colorless to
4+ V
yellow-reddish), Am (yellow-reddish), Am O+
VI
2; (yellow), Am O2+
VII
2 (brown) and Am O5−
[21][22]
6 (dark green). All oxidation states have their characteristic optical absorption spectra, with a few sharp peaks
in the visible and mid-infrared regions, and the position and intensity of these peaks can be converted into the
[23]
concentrations of the corresponding oxidation states. For example, Am(III) has two sharp peaks at 504 and
811 nm, Am(V) at 514 and 715 nm, and Am(VI) at 666 and 992 nm.
Americium compounds with oxidation state +4 and higher are strong oxidizing agents, comparable in strength to the
permanganate ion (MnO−
[24] 4+ 3+
4) in acidic solutions. Whereas the Am ions are unstable in solutions and readily convert to Am , the +4
oxidation state occurs well in solids, such as americium dioxide (AmO ) and americium(IV) fluoride (AmF
).
2 4
All pentavalent and hexavalent americium compounds are complex salts such as KAmO F , Li AmO and
2 2 3 4
Li AmO , Ba AmO , AmO F . These high oxidation states Am(IV), Am(V) and Am(VI) can be prepared from
6 6 3 6 2 2
Am(III) by oxidation with ammonium persulfate in dilute nitric acid, with silver(I) oxide in perchloric acid, or with
ozone or sodium persulfate in sodium carbonate solutions. The pentavalent oxidation state of americium was
first observed in 1951. It is present in aqueous solution in the form of AmO+
2 ions (acidic) or AmO−
3 ions (alkaline) which are however unstable and subject to several rapid disproportionation reactions:[25]

Chemical compounds

Oxygen compounds
Two americium oxides are known, with the oxidation states +3 (Am O ) and +4 (AmO ). Americium(III) oxide is a
2 3 2
[26]
red-brown solid with a melting point of 2205 °C. Americium(IV) oxide is the main form of solid americium
which is used in nearly all its applications. As most other actinide dioxides, it is a black solid with a cubic (fluorite)
[27]
crystal structure.
The oxalate of americium(III), vacuum dried at room temperature, has the chemical formula Am (C O ) ·7H O.
2 2 4 3 2
Upon heating in vacuum, it loses water at 240 °C and starts decomposing into AmO2 at 300 °C, the decomposition
[28]
completes at about 470 °C. The initial oxalate dissolves in nitric acid with the maximum solubility of 0.25 g/L.

Halides
[29]
Halides of americium are known for the oxidation states +2, +3 and +4, where the +3 is most stable, especially in
solutions.

Oxidation state F Cl Br I

+4 Americium(IV)
fluoride
AmF
4
pale pink

+3 Americium(III) Americium(III) Americium(III) Americium(III)


fluoride chloride bromide iodide
AmF AmCl AmBr AmI
3 3 3 3
pink pink light yellow light yellow

+2 Americium(II) chloride Americium(II) bromide Americium(II) iodide


AmCl AmBr AmI
2 2 2
black black black

Reduction of Am(III) compounds with sodium amalgam yields Am(II) salts – the black halides AmCl , AmBr and
2 2
AmI . They are very sensitive to oxygen and oxidize in water, releasing hydrogen and converting back to the
2
Am(III) state. Specific lattice constants are:
• Orthorhombic AmCl : a = 896.3 ± 0.8 pm, b = 757.3 ± 0.8 pm and c = 453.2 ± 0.6 pm
2
• Tetragonal AmBr : a = 1159.2 ± 0.4 and c = 712.1 ± 0.3 pm.
2
They can also be prepared by reacting metallic americium with an appropriate mercury halide HgX 2, where X = Cl,
[30]
Br or I:
3+
Americium(III) fluoride (AmF ) is poorly soluble and precipitates upon reaction of Am and fluoride ions in weak
3
acidic solutions:

The tetravalent americium(IV) fluoride (AmF4) is obtained by reacting solid americium(III) fluoride with molecular
[31]
fluorine:

[32]
Another known form of solid tetravalent americium chloride is KAmF . Tetravalent americium has also been
5
observed in the aqueous phase. For this purpose, black Am(OH) was dissolved in 15-M NH F with the americium
4 3
concentration of 0.01 M. The resulting reddish solution had a characteristic optical absorption spectrum which is
similar to that of AmF but differed from other oxidation states of americium. Heating the Am(IV) solution to 90 °C
4
did not result in its disproportionation or reduction, however a slow reduction was observed to Am(III) and assigned
to self-irradiation of americium by alpha particles.
Most americium(III) halides form hexagonal crystals with slight variation of the color and exact structure between
the halogens. So, chloride (AmCl ) is reddish and has a structure isotypic to uranium(III) chloride (space group
3
P6 /m) and the melting point of 715 °C. The fluoride is isotypic to LaF (space group P6 /mmc) and the iodide to
3 3 3
BiI (space group R3). The bromide is an exception with the orthorhombic PuBr -type structure and space group
3 3
Cmcm. Crystals of americium hexahydrate (AmCl ·6H O) can be prepared by dissolving americium dioxide in
3 2
hydrochloric acid and evaporating the liquid. Those crystals are hygroscopic and have yellow-reddish color and a
monoclinic crystal structure.
VI V III
Oxyhalides of americium in the form Am O X , Am O X, and Am OX can be obtained by reacting
IV
Am OX
2 2 2 2
the corresponding americium halide with oxygen or Sb O , and AmOCl can also be produced by vapor phase
2 3
hydrolysis:

Chalcogenides and pnictides


The known chalcogenides of americium include the sulfide AmS , selenides AmSe and Am Se , and tellurides
2 2 3 4
Am Te and AmTe . The pnictides of americium (243Am) of the AmX type are known for the elements phosphorus,
2 3 2
arsenic, antimony and bismuth. They crystallize in the rock-salt lattice.

Silicides and borides


Americium monosilicide (AmSi) and "disilicide" (nominally AmSi with: 1.87 < x < 2.0) were obtained by reduction
x
of americium(III) fluoride with elementary silicon in vacuum at 1050 °C (AmSi) and 1150−1200 °C (AmSi ). AmSi
x
is a black solid isomorphic with LaSi, it has an orthorhombic crystal symmetry. AmSi has a bright silvery lustre and
x
a tetragonal crystal lattice (space group I4 /amd), it is isomorphic with PuSi and ThSi . Borides of americium
1 2 2
include AmB 4 and AmB 6. The tetraboride can be obtained by heating an oxide or halide of americium with
[33]
magnesium diboride in vacuum or inert atmosphere.
Organoamericium compounds
Analogous to uranocene, americium forms an organometallic compound with two
8
cyclooctatetraene ligands, that is (η -C8H 8) 2Am. It also makes trigonal
5
(η -C H ) Am complexes with three cyclopentadienyl rings.
5 5 3
Formation of the complexes of the type Am(n-C H -BTP) , where BTP stands for
3 7 3 3+
2,6-di(1,2,4-triazin-3-yl)pyridine, in solutions containing n-C H -BTP and Am
3 7
ions has been confirmed by EXAFS. Some of these BTP-type complexes selectively
interact with americium and therefore are useful in its selective separation from
lanthanides and another actinides.

8
(η -C H ) Am structure
8 8 2

Biological aspects
[34]
Americium is an artificial element, and thus a biological function involving the element would be impossible. It
has been proposed to use bacteria for removal of americium and other heavy metals from rivers and streams. Thus,
Enterobacteriaceae of the genus Citrobacter precipitate americium ions from aqueous solutions, binding them into a
metal-phosphate complex at their cell walls. Several studies have been reported on the biosorption and
bioaccumulation of americium by bacteria and fungi.

Fission
242m1
The isotope Am (half-life 141 years) has the largest cross sections for absorption of thermal neutrons (5,700
[35]
barns), that results in a small critical mass for a sustained nuclear chain reaction. The critical mass for a bare
242m1
Am sphere is about 9–14 kg (the uncertainty results from insufficient knowledge of its material properties). It
[36]
can be lowered to 3–5 kg with a metal reflector and should become even smaller with a water reflector. Such
242m1
small critical mass is favorable for portable nuclear weapons, but those based on Am are not known yet,
241
probably because of its scarcity and high price. The critical masses of two other readily available isotopes, Am
243 241 243 [37]
and Am, are relatively high – 57.6 to 75.6 kg for Am and 209 kg for Am. Scarcity and high price yet
hinder application of americium as a nuclear fuel in nuclear reactors.
242m1
There are proposals of very compact 10-kW high-flux reactors using as little as 20 grams of Am. Such
low-power reactors would be relatively safe to use as neutron sources for radiation therapy in hospitals.

Isotopes
See also: Isotopes of
americium
241
About 19 isotopes and 8 nuclear isomers are known for americium. There are two long-lived alpha-emitters, Am
243 242m1
and Am with half-lives of 432.2 and 7,370 years, respectively, and the nuclear isomer Am has a long
245m1
half-life of 141 years. The half-lives of other isotopes and isomers range from 0.64 microseconds for Am to
240
50.8 hours for Am. As with most other actinides, the isotopes of americium with odd number of neutrons have
relatively high rate of nuclear fission and low critical mass.
237
Americium-241 decays to Np emitting alpha particles of 5 different energies, mostly at 5.486 MeV (85.2%)
and
5.443 MeV (12.8%). Because many of the resulting states are metastable, they also emit gamma rays with the
discrete energies between 26.3 and 158.5 keV.
242
Americium-242 is a short-lived isotope with a half-life of 16.02 h. It mostly (82.7%) converts by β-decay to Cm,
242 242 242
but also by electron capture to Pu (17.3%). Both Cm and Pu transform via nearly the same decay chain
238 234
through Pu down to U.
242m1 242
Nearly all (99.541%) of Am decays by internal conversion to Am and the remaining 0.459% by α-decay to
238 238 234
Np. The latter breaks down to Pu and then to U.
239 239 239
Americium-243 transforms by α-emission into Np, which converts by β-decay to Pu, and the Pu
235
changes into U by emitting an α-particle.

Applications

Outside and inside view of an americium-based smoke detector

Ionization
detectors
Main article: Americium smoke
detector
Americium is the only synthetic element to have found its way into the household, where one common type of
241 [38]
smoke detector uses Am in the form of americium dioxide as its source of ionizing radiation. This isotope is
226
preferred over Ra because it emits 5 times more alpha particles and relatively little harmful γ-radiation. The
amount of americium in a typical new smoke detector is 1 microcurie (37 kBq) or 0.28 microgram. This amount
declines slowly as the americium decays into neptunium-237, a different transuranic element with a much longer
half-life (about 2.14 million years). With its half-life of 432.2 years, the americium in a smoke detector includes
about 3% neptunium after 19 years, and about 5% after 32 years. The radiation passes through an ionization
chamber, an air-filled space between two electrodes, and permits a small, constant current between the electrodes.
Any smoke that enters the chamber absorbs the alpha particles, which reduces the ionization and affects this current,
triggering the alarm. Compared to the alternative optical smoke detector, the ionization smoke detector is cheaper
and can detect particles which are too small to produce significant light scattering; however, it is more prone to false
[39][40]
alarms.
Radionuclide
241 238
As Am has a significantly longer half-life than Pu (432.2 years vs. 87 years), it has been proposed as an active
[41]
element of radioisotope thermoelectric generators, for example in spacecraft. Although americium produces less
241 243 238
heat and electricity – the power yield is 114.7 mW/g for Am and 6.31 mW/g for Am (cf. 390 mW/g for Pu)
– and its radiation poses more threat to humans owing to neutron emission, the European Space Agency is planning
[42]
to use americium for its space probes.
Another proposed space-related application of americium is a fuel for space ships with nuclear propulsion. It relies
242m
on the very high rate of nuclear fission of Am, which can be maintained even in a micrometer-thick foil. Small
thickness avoids the problem of self-absorption of emitted radiation. This problem is pertinent to uranium or
242m
plutonium rods, in which only surface layers provide alpha-particles. The fission products of Am can either
directly propel the spaceship or they can heat up a thrusting gas; they can also transfer their energy to a fluid and
generate electricity through a magnetohydrodynamic generator.
242m
One more proposal which utilizes the high nuclear fission rate of Am is a nuclear battery. Its design relies not on
the energy of the emitted by americium alpha particles, but on their charge, that is the americium acts as the
242m
self-sustaining "cathode". A single 3.2 kg Am charge of such battery could provide about 140 kW of power over
[43] 242m
a period of 80 days. With all the potential benefits, the current applications of Am are as yet hindered by the
scarcity and high price of this nuclear isomer.

Neutron
source
241
The oxide of Am pressed with beryllium is an efficient neutron source. Here americium acts as the alpha source,
and beryllium produces neutrons owing to its large cross-section for the (α,n) nuclear reaction:

241
The most widespread use of AmBe neutron sources is a neutron probe – a device used to measure the quantity of
241
water present in soil, as well as moisture/density for quality control in highway construction. Am neutron sources
are also used in well logging applications, as well as in neutron radiography, tomography and other radiochemical
investigations.

Production of other
elements
Americium is a starting material for the production of other transuranic elements and transactinides – for example,
242 242 242 242
82.7% of Am decays to Cm and 17.3% to Pu. In the nuclear reactor, Am is also up-converted by neutron
243 244 244
capture to Am and Am, which transforms by β-decay to Cm:

241 12 22 247 260


Irradiation of Am by C or Ne ions yields the isotopes Es (einsteinium) or Db (dubnium), respectively.
243
Furthermore, the element berkelium ( Bk isotope) had been first intentionally produced and identified by
241
bombarding Am with alpha particles, in 1949, by the same Berkeley group, using the same 60-inch cyclotron.
Similarly, nobelium was produced at the Joint Institute for Nuclear Research, Dubna, Russia, in 1965 in several
243 15
reactions, one of which included irradiation of Am with N ions. Besides, one of the synthesis reactions for
243 18
lawrencium, discovered by scientists at Berkeley and Dubna, included bombardment of Am with O.
Spectrometer
Americium-241 has been used as a portable source of both gamma rays and alpha particles for a number of medical
241
and industrial uses. The 60-keV gamma ray emissions from Am in such sources can be used for indirect analysis
of materials in radiography and X-ray fluorescence spectroscopy, as well as for quality control in fixed nuclear
density gauges and nuclear densometers. For example, the element has been employed to gauge glass thickness to
help create flat glass. Americium-241 is also suitable for calibration of gamma-ray spectrometers in the low-energy
range, since its spectrum consists of nearly a single peak and negligible Compton continuum (at least three orders of
[44]
magnitude lower intensity). Americium-241 gamma rays were also used to provide passive diagnosis of thyroid
function. This medical application is however obsolete.

Health concerns
As a highly radioactive element, americium and its compounds must be handled only in an appropriate laboratory
under special arrangements. Although most americium isotopes predominantly emit alpha particles which can be
blocked by thin layers of common materials, many of the daughter products emit gamma-rays and neutrons which
[45]
have a long penetration depth.
If consumed, americium is excreted within a few days and only 0.05% is absorbed in the blood. From there,
roughly 45% of it goes to the liver and 45% to the bones, and the remaining 10% is excreted. The uptake to the
liver depends on the individual and increases with age. In the bones, americium is first deposited over cortical
241
and trabecular surfaces and slowly redistributes over the bone with time. The biological half-life of Am is 50
years in the bones and 20 years in the liver, whereas in the gonads (testicles and ovaries) it remains
[46]
permanently; in all these organs, americium promotes formation of cancer cells as a result of its radioactivity.
Americium often enters landfills from discarded smoke detectors. The rules associated with the disposal of smoke
detectors are relaxed in most jurisdictions. In the U.S., the "Radioactive Boy Scout" David Hahn was able to
concentrate americium from smoke detectors after managing to buy a hundred of them at remainder prices and also
[47]
stealing a few. There have been cases of humans being contaminated with americium, the worst case being that of
Harold McCluskey, who at the age of 64 was exposed to 500 times the occupational standard for americium-241 as a
result of an explosion in his lab. McCluskey died at the age of 75, not as a result of exposure, but of a heart disease
which he had before the accident.

Notes
[1] http://en.wikipedia.org/w/index.php?title=Template:Infobox_americium&action=edit
[2] Obituary of Dr. Leon Owen (Tom) Morgan (1919–2002) (http:/ /www.utexas.edu/ faculty/council/2002-2003/memorials/Morgan/
morgan. html), Retrieved 28 November 2010
[3] Seaborg, G. T.; James, R.A. and Morgan, L. O.: "The New Element Americium (Atomic Number 95)", THIN PPR (National Nuclear
Energy Series, Plutonium Project Record), Vol 14 B The Transuranium Elements: Research Papers, Paper No. 22.1, McGraw-Hill Book
Co., Inc., New York, 1949. Abstract (http:// www.osti.gov/cgi-bin/rd_accomplishments/ display_biblio.cgi?
id=ACC0046&numPages=43&fp=N);
Full text (http:// www.osti.gov/accomplishments/ documents/ fullText/ACC0046.pdf) (January 1948), Retrieved 28 November 2010
[4] Greenwood, p. 1252
[5] Seaborg, Glenn T. "Element", Filing date: 23 August 1946, Issue date: 10 November 1964
[6] Human Health Fact Sheet on Americium (http:// www.ead. anl.gov/pub/doc/americium.pdf), Los Alamos National Laboratory, Retrieved
28 November 2010
[7] Hoffmann, Klaus Kann man Gold machen? Gauner, Gaukler und Gelehrte. Aus der Geschichte der chemischen Elemente (Can you make
gold? Crooks, clowns and scholars. From the history of the chemical elements), Urania-Verlag, Leipzig, Jena, Berlin 1979, no ISBN, p. 233
[8] Baetslé, L. Application of Partitioning/Transmutation of Radioactive Materials in Radioactive Waste Management (http://www.ictp.trieste.
it/~pub_off/lectures/ lns012/Baetsle.pdf), Nuclear Research Centre of Belgium Sck/Cen, Mol, Belgium, September 2001, Retrieved 28
November 2010
[9] Fioni, Gabriele; Cribier, Michel and Marie, Frédéric Can the minor actinide, americium-241, be transmuted by thermal neutrons? (http://
www.cea.fr/var/cea/ storage/ static/gb/library/Clefs46/pagesg/clefs46_30.html), Department of Astrophysics, CEA/Saclay, Retrieved 28
November 2010
[10] Greenwood, p. 1262
[11] Smoke detectors and americium (http://www.world-nuclear.org/ info/inf57.html), World Nuclear Association, January 2009,
Retrieved 28 November 2010
[12] Hammond C. R. "The elements" in
[13] BREDL Southern Anti-Plutonium Campaign (http://www.bredl.org/sapc/Pu_ReportI.htm), Blue Ridge Environmental Defense League,
Retrieved 28 November 2010
[14] The "metastable" state is marked by the letter m.
[15] article/200410/000020041004A0333355.php Abstract (http://sciencelinks. jp/j-east/)
[16] Penneman, pp. 34–48
[17] Penneman, p. 25
[18] Gmelin Handbook of Inorganic Chemistry, System No. 71, transuranics, Part B 1, pp. 57–67.
[19] Penneman, p. 3
[20] Penneman, p. 4
[21] Americium (http://www.chemie-master.de/FrameHandler.php?loc=http://www.chemie-master.de/pse/pse. php?modul=Am), Das
Periodensystem der Elemente für den Schulgebrauch (The periodic table of elements for schools) chemie-master.de (in German), Retrieved
28 November 2010
[22] Greenwood, p. 1265
[23] Penneman, pp. 10–14
[24] Holleman, p. 1956
[25] Greenwood, p. 1275
[26] Holleman, p. 1972
[27] Greenwood, p. 1267
[28] Penneman, p. 5
[29] Holleman, p. 1969
[30] Greenwood, p. 1272
[31] Greenwood, p. 1271
[32] Penneman, p. 6
[33] Lupinetti, A. J. et al. "Low-temperature synthesis of actinide tetraborides by solid-state metathesis reactions", Filed 4 Apr 2002, Issued 14
Dec 2004
[34] Toeniskoetter, Steve; Dommer, Jennifer and Dodge, Tony The Biochemical Periodic Tables – Americium (http://umbbd.ethz. ch/periodic/
elements/ am.html), University of Minnesota, Retrieved 28 November 2010
[35] Pfennig, G.; Klewe-Nebenius, H and Seelmann Eggebert, W. (Eds.): Karlsruhe nuclide, 7 Edition 2006.
[36] Abstract (http://sciencelinks. jp/j-east/article/200403/000020040303A0828431.php)
[37] Institut de Radioprotection et de Sûreté Nucléaire, "Evaluation of nuclear criticality safety data and limits for actinides in transport" (http://
ec.europa.eu/ energy/ nuclear/ transport/ doc/irsn_sect03_146.pdf), p. 16.
[38] Smoke Detectors and Americium (http://web.archive.org/web/19960101-re_/http://www.uic.com.au/ nip35.htm), Nuclear Issues
Briefing Paper 35, May 2002. (Internet Archive), Retrieved 28 November 2010
[39] Residential Smoke Alarm Performance, Thomas Cleary. Building and Fire Research Laboratory, National Institute of Standards and
Technology; UL Smoke and Fire Dynamics Seminar. November 2007
[40] Bukowski, R. W. et al. (2007) Performance of Home Smoke Alarms Analysis of the Response of Several Available Technologies in
Residential Fire Settings (http:/ /www.fire.nist.gov/bfrlpubs/fire07/ art063. html), NIST Technical Note 1455-1
[41] Basic elements of static RTGs (http://fti.neep.wisc.edu/neep602/SPRING00/lecture5.pdf), G.L. Kulcinski, NEEP 602 Course Notes
(Spring 2000), Nuclear Power in Space, University of Wisconsin Fusion Technology Institute (see last page)
[42] Space agencies tackle waning plutonium stockpiles (http:// www.spaceflightnow.com/news/ n1007/ 09rtg/), Spaceflight now, 9 July 2010
[43] Genuth, Iddo Americium Power Source (http://thefutureofthings.com/articles.php?itemId=26/64/), The Future of Things, 3 October
2006, Retrieved 28 November 2010
[44] Nuclear Data Viewer 2.4 (http:/ /www.nndc.bnl.gov/nudat2/indx_dec. jsp), NNDC
[45] Public Health Statement for Americium (http:// www.atsdr.cdc. gov/phs/phs. asp?id=809&tid=158) Section 1.5., Agency for Toxic
Substances and Disease Registry, April 2004, Retrieved 28 November 2010
[46] Frisch, Franz Crystal Clear, 100 x energy, Bibliographisches Institut AG, Mannheim 1977, ISBN 3-411-01704-X, p. 184
[47] Ken Silverstein, The Radioactive Boy Scout: When a teenager attempts to build a breeder reactor (http://www.harpers. org/archive/1998/
11/0059750). Harper's Magazine, November 1998
References

Bibliography
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
ISBN 0080379419.
• Wiberg, Nils (2007). Lehrbuch Der Anorganischen Chemie. De Gruyter. ISBN 978-3-11-017770-1.
• Penneman, R. A. and Keenan T. K. The radiochemistry of americium and curium
(http://www.osti.gov/bridge/ purl.cover.jsp?purl=/4187189-IKQUwY/), University of California, Los
Alamos, California, 1960

Further reading
• Nuclides and Isotopes – 14th Edition, GE Nuclear Energy, 1989.
• Fioni, Gabriele; Cribier, Michel and Marie, Frédéric. "Can the minor actinide, americium-241, be transmuted by
thermal neutrons?" (http://www.cea.fr/var/cea/ storage/ static/gb/library/Clefs46/pagesg/clefs46_30.html).
Commissariat à l'énergie atomique.
• Stwertka, Albert (1999). A Guide to the Elements. Oxford University Press, USA. ISBN 0-19-508083-1.

External links
• Americium (http://www.periodicvideos.com/videos/095.htm) at The Periodic Table of Videos (University
of Nottingham)
• ATSDR – Public Health Statement: Americium (http://www.atsdr.cdc.gov/toxprofiles/phs156.html)
• World Nuclear Association – Smoke Detectors and Americium (http://world-nuclear.org/info/inf57.html)
96 Curium 1

96 – Curium
This article is about the chemical element. For the ancient city located in Cyprus, see Kourion.

Curium
Cm
96

americium ← curium → berkeliumGd



Cm

(Uqo)

Curium in the periodic table


Appearance

silvery

General properties

Name, symbol, number curium, Cm, 96

Pronunciation /ˈkjʊəriəm/
KEWR-ee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (247)

Electron configuration 7 1 2
[Rn] 5f 6d 7s
2, 8, 18, 32, 25, 9, 2

Physical properties

Phase solid

Density (near r.t.) −3


13.51 g·cm

Melting point 1613 K, 1340 °C, 2444 °F

Boiling point 3383 K, 3110 °C, 5630 °F

Heat of fusion −1
? 15 kJ·mol

Vapor pressure

P (Pa) 1 10 100 1 k 10 k 100 k

at T (K) 1788 1982


Atomic properties

Oxidation states 8, 6, 4, 3, 2 (amphoteric oxide)

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 581 kJ·mol

Atomic radius 174 pm

Covalent radius 169±3 pm

Miscellanea

Crystal structure hexagonal close-packed

Magnetic ordering antiferromagnetic-paramagnetic transition at 52 K

Electrical resistivity 1.25 µΩ·m

CAS registry number 7440-51-9

History

Naming named after Marie Skłodowska-Curie and Pierre Curie

Discovery Glenn T. Seaborg, Ralph A. James, Albert Ghiorso (1944)

Most stable isotopes

Main article: Isotopes of curium

iso NA half-life DM DE (MeV) DP

242 trace 160 d SF -


Cm
α 6.1 238
Pu

243 trace 29.1 y α 6.169 239


Cm Pu
ε 0.009 243
Am
SF -

244 trace 18.1 y SF -


Cm
α 5.8048 240
Pu

245 trace 8500 y SF -


Cm
α 5.623 241
Pu

246 trace 4730 y α 5.475 242


Cm Pu
SF -

247 trace 7 α 5.353 243


Cm 1.56×10 y Pu

248 trace 5 α 5.162 244


Cm 3.40×10 y Pu
SF -

250 syn 9000 y SF -


Cm
α 5.169 246
Pu

− 0.037 250
β Bk
• v
• t
[1]
• e

Curium is a transuranic radioactive chemical element with the symbol Cm and atomic number 96. This element
of the actinide series was named after Marie and Pierre Curie – both were known for their research on
radioactivity. Curium was first intentionally produced and identified in July 1944 by the group of Glenn T.
Seaborg at the University of California, Berkeley. The discovery was kept secret and only released to the public
in November 1945. Most curium is produced by bombarding uranium or plutonium with neutrons in nuclear
reactors – one tonne of spent nuclear fuel contains about 20 grams of curium.
Curium is a hard, dense, silvery metal with a relatively high melting point and boiling point for an actinide. Whereas
it is paramagnetic at ambient conditions, it becomes antiferromagnetic upon cooling, and other magnetic transitions
are also observed for many curium compounds. In compounds, curium usually exhibits valence +3 and sometimes
+4, and the +3 valence is predominant in solutions. Curium readily oxidizes, and its oxides are a dominant form
of this element. It forms strongly fluorescent complexes with various organic compounds, but there is no
evidence of its incorporation into bacteria and archaea. When introduced into the human body, curium
accumulates in the bones, lungs and liver, where it promotes cancer.
All known isotopes of curium are radioactive and have a small critical mass for a sustained nuclear chain reaction.
They predominantly emit α-particles, and the heat released in this process can potentially produce electricity in
radioisotope thermoelectric generators. This application is hindered by the scarcity, high cost and radioactivity of
238
curium isotopes. Curium is used in production of heavier actinides and of the Pu radionuclide for power sources
in artificial pacemakers. It served as the α-source in the alpha particle X-ray spectrometers installed on the
Sojourner, Mars, Mars 96, Athena, Spirit and Opportunity rovers as well as the Mars Science Laboratory to analyze
the composition and structure of the rocks on the surface of Mars and the Moon. Such a spectrometer will also be
used by the Philae lander of the Rosetta spacecraft to probe the surface of the 67P/Churyumov-Gerasimenko comet.

History
Although curium had likely been produced in previous nuclear experiments, it
was first intentionally synthesized, isolated and identified in 1944, at the
University of California, Berkeley, by Glenn T. Seaborg, Ralph A. James, and
Albert Ghiorso. In their experiments, they used a 60-inch (150 cm) cyclotron.
Curium was chemically identified at the Metallurgical Laboratory (now Argonne
National Laboratory) at the University of Chicago. It was the third transuranium
element to be discovered even though it is the fourth in the series – the lighter
[2]
element americium was unknown at the time.
The sample was prepared as follows: first plutonium nitrate solution was coated
2
on a platinum foil of about 0.5 cm area, the solution was evaporated and the
Glenn T. Seaborg residue was converted into plutonium dioxide (PuO ) by annealing. Following
2
cyclotron irradiation of the oxide, the coating was dissolved with nitric acid
and
then precipitated as the hydroxide using concentrated aqueous ammonia solution. The residue was dissolved in
perchloric acid, and further separation was carried out by ion exchange to yield a certain isotope of curium. The
separation of curium and americium was so painstaking that the Berkeley group initially called those elements
[3]
pandemonium (from Greek for all demons or hell) and delirium (from Latin for madness).
The curium-242 isotope was produced in July–August 1944 by bombarding
239
Pu with α-particles to produce curium with the release of a neutron:

Curium-242 was unambiguously identified by the characteristic energy of the α-


particles emitted during the decay:

The 60-inch (150 cm) cyclotron at


the Lawrence Radiation Laboratory,
University of California, Berkeley, in
August 1939.

The half-life of this alpha decay was first measured as 150 days and then corrected to 162.8 days.
240
Another isotope Cm was produced in a similar reaction in March 1945:

240
The half-life of the Cm α-decay was correctly determined as 26.7 days.
The discovery of curium, as well as americium, in 1944 was closely related to the Manhattan Project, the results
were confidential and declassified only in 1945. Seaborg leaked the synthesis of the elements 95 and 96 on the U.S.
radio show for children, the Quiz Kids, five days before the official presentation at an American Chemical Society
meeting on November 11, 1945, when one of the listeners asked whether any new transuranium element beside
242 240
plutonium and neptunium had been discovered during the war. The discovery of curium ( Cm and Cm), their
[4]
production and compounds were later patented listing only Seaborg as the inventor.

Marie and Pierre Curie


The new element was named after Marie Skłodowska-Curie and her husband Pierre Curie who are noted for
discovering radium and for their work in radioactivity. It followed the example of gadolinium, a lanthanide element
[5]
above curium in the periodic table, which was named after the explorer of the rare earth elements Johan Gadolin:
"As the name for the element of atomic number 96 we should like to propose "curium", with symbol Cm.
The evidence indicates that element 96 contains seven 5f electrons and is thus analogous to the element
gadolinium with its seven 4f electrons in the regular rare earth series. On this base element 96 is named
after the Curies in a manner analogous to the naming of gadolinium, in which the chemist Gadolin was
honored."
The first curium samples were barely visible, and were identified by their radioactivity. Louis Werner and Isadore
Perlman created the first substantial sample of 30 µg curium-242 hydroxide at the University of California in 1947
[6][7]
by bombarding americium-241 with neutrons. Macroscopic amounts of curium fluoride were obtained in 1950
by W. W. T. Crane, J. C. Wallmann and B. B. Cunningham. Its magnetic susceptibility was very close to that of
GdF providing the first experimental evidence for the +3 valence of curium in its compounds. Curium metal was
3
produced only in 1951 by reduction of curium fluoride with barium.

Characteristics

Physical
A synthetic, radioactive element, curium is a hard dense metal with
silvery-white appearance and physical and chemical properties
resembling those of gadolinium. Its melting point of 1340 °C is
significantly higher than that of the previous transuranic elements
neptunium (637 °C), plutonium (639 °C) and americium (1173 °C). In
comparison, gadolinium melts at 1312 °C. The boiling point of curium
3
is 3110 °C. With a density of 13.52 g/cm , curium is significantly
3 3
lighter than neptunium (20.45 g/cm ) and plutonium (19.8 g/cm ), but
is heavier than most other metals. Between two crystalline forms of
curium, the α-Cm is more stable at ambient conditions. It has a
hexagonal symmetry, space group P6 /mmc, lattice parameters a = 365
3 Double-hexagonal close packing with the layer
pm and c = 1182 pm, and four formula units per unit cell. The crystal sequence ABAC in the crystal structure of
consists of a double-hexagonal close packing with the layer sequence α-curium (A: green, B: blue, C: red)
ABAC and so is isotypic with α-lanthanum. At pressures above 23
GPa, at room temperature, α-Cm transforms into β-Cm, which has a
face-centered cubic symmetry, space group Fm3m and the lattice
constant a = 493 pm. Upon further compression to 43 GPa, curium
transforms to an orthorhombic γ-Cm structure similar to that of α-
uranium, with no further transitions observed up to 52 GPa. These
[8]
three curium phases are also referred to as Cm I, II and III.

Curium has peculiar magnetic properties. Whereas its neighbor


element americium shows no deviation from Curie-Weiss
paramagnetism in the entire temperature range, α-Cm transforms to an 3+
Orange fluorescence of Cm ions in a solution
antiferromagnetic state upon cooling to 65–52 K, and β-Cm exhibits a of tris(hydrotris)pyrazolylborato-Cm(III)
ferrimagnetic transition at about 205 K. Meanwhile, curium pnictides complex, excited at 396.6 nm.
244 244
show ferromagnetic transitions upon cooling: CmN and CmAs at
248 248
109 K, CmP at 73 K and CmSb at 162 K. Similarly, the lanthanide analogue of curium, gadolinium, as well as
its pnictides also show magnetic transitions upon cooling, but the transition character is somewhat different: Gd and
[9]
GdN become ferromagnetic, and GdP, GdAs and GdSb show antiferromagnetic ordering.
In accordance with magnetic data, electrical resistivity of curium increases with temperature – about twice between
4 and 60 K – and then remains nearly constant up to room temperature. There is a significant increase in resistvity
over time (about 10 µΩ·cm/h) due to self-damage of the crystal lattice by alpha radiation. This makes uncertain the
absolute resistivity value for curium (about 125 µΩ·cm). The resistivity of curium is similar to that of gadolinium
and of the actinides plutonium and neptunium, but is significantly higher than that of americium, uranium, polonium
and thorium.
Under ultraviolet illumination, curium(III) ions exhibit strong and stable yellow-orange fluorescence with a
maximum in the range about 590–640 nm depending on their environment. The fluorescence originates from the
6 8
transitions from the first excited state D and the ground state S . Analysis of this fluorescence allows
7/2 7/2
[10]
monitoring interactions between Cm(III) ions in organic and inorganic complexes.

Chemical
Curium ions in solution almost exclusively assume the oxidation state of +3, which is the most stable oxidation state
[11]
for curium. The +4 oxidation state is observed mainly in a few solid phases, such as CmO and CmF . Aqueous
2 4
curium(IV) is only known in the presence of strong oxidizers such as potassium persulfate, and is easily reduced to
curium(III) by radiolysis and even by water. The chemical behavior of curium is different from the actinides thorium
3+
and uranium, and is similar to that of americium and many lanthanides. In aqueous solution, the Cm ion is
[12] 4+ [13] 3+
colorless to pale green, and Cm ion is pale yellow. The optical absorption of Cm ions contains three sharp
peaks at 375.4, 381.2 and 396.5 nanometers and their strength can be directly converted into the concentration of the
[14]
ions. The +6 oxidation state has only been reported once in solution in 1978, as the curyl ion (CmO2+
2): this was prepared from the beta decay of americium-242 in the americium(V) ion 242

AmO+
4+ 3+
2.Failure to obtain Cm(VI) from oxidation of Cm(III) and Cm(IV) may be due to the high Cm /Cm ionization
potential and the instability of Cm(V).
Curium ions are hard Lewis acids and thus form most stable complexes with hard bases. The bonding is mostly
ionic, with a small covalent component. Curium in its complexes commonly exhibits a 9-fold coordination
[15]
environment, within a tricapped trigonal prismatic geometry.

Isotopes
[16]
Thermal neutron cross sections (barns
)

242 243 244 245 246 247


Cm Cm Cm Cm Cm Cm
Fission 5 617 1.04 2145 0.14 81.90

Capture 16 130 15.20 369 1.22 57

C/F ratio 3.20 0.21 14.62 0.17 8.71 0.70

LEU spent fuel 20 years after 53 MWd/kg burnup

3 common isotopes 51 3700 390

Fast reactor MOX fuel (avg 5 samples, burnup 66-120GWd/t)

−3 27.64% 70.16% 2.166% 0.0376% 0.000928%


Total curium 3.09×10 %

Isotope 242 243 244 245 246 247 248 250


Cm Cm Cm Cm Cm Cm Cm Cm
Critical mass, kg 25 7.5 33 6.8 39 7 40.4 23.5

See also: Isotopes of curium


96 Curium 1

233 252
About 20 radioisotopes and 7 nuclear isomers between Cm and Cm are known for curium, and no stable
247 248
isotopes. The longest half-lives have been reported for Cm (15.6 million years) and Cm (348,000 years). Other
245 250 246
long-lived isotopes are Cm (half-life 8500 years), Cm (8,300 years) and Cm (4,760 years). Curium-250 is
unusual in that it predominantly (about 86%) decays via spontaneous fission. The most commonly used curium
242 244
isotopes are Cm and Cm with the half-lives of 162.8 days and 18.1 years, respectively.
242
All isotopes between Cm and
248 250
Cm, as well as Cm, undergo a
self-sustaining nuclear chain reaction
and thus in principle can act as a
nuclear fuel in a reactor. As in most
transuranic elements, the nuclear
fission cross section is especially high
for the odd-mass curium
243 245 247
isotopes Cm, Cm and Cm.
These can be used in thermal-neutron
reactors, whereas a mixture of curium
isotopes is only suitable for fast
breeder reactors since the even-mass
isotopes are not fissile in a thermal
reactor and accumulate as burn-up
[17]
increases. The mixed-oxide (MOX)
fuel, which is to be used in power
reactors, should contain little or no
curium because the neutron activation
248
of Cm will create californium. This
is strong neutron emitter, and would 238 244
Transmutation flow between Pu and Cm in LWR.
pollute the back end of the fuel cycle Fission percentage is 100 minus shown percentages.
Total rate of transmutation varies greatly by nuclide.
and increase the dose to reactor 245 248
Cm– Cm are long-lived with negligible decay.
personnel. Hence, if the minor
actinides are to be used as fuel in a thermal neutron reactor, the curium should be excluded from the fuel or placed in
special fuel rods where it is the only actinide present.

The table to the right lists the critical masses for curium isotopes for a sphere, without a moderator and reflector.
With a metal reflector (30 cm of steel), the critical masses of the odd isotopes are about 3–4 kg. When using water
245 243
(thickness ~20–30 cm) as the reflector, the critical mass can be as small as 59 gram for Cm, 155 gram for Cm
247
and 1550 gram for Cm. There is a significant uncertainty in these critical mass values. Whereas it is usually of the
242 246
order 20%, the values for Cm and Cm were listed as large as 371 kg and 70.1 kg, respectively, by some
research groups.
[18] 245 247
Currently, curium is not used as a nuclear fuel owing to its low availability and high price. Cm and Cm
have a very small critical mass and therefore could be used in portable nuclear weapons, but none have been
reported thus far. Curium-243 is not suitable for this purpose because of its short half-life and strong α emission
which would result in excessive heat. Curium-247 would be highly suitable, having a half-life 647 times that of
plutonium-239.
Occurrence
247
The longest-lived isotope of curium, Cm, has a half-life of 15.6
million years. Therefore, all primordial curium, that is curium present
on the Earth during its formation, should have decayed by now.
Curium is produced artificially, in small quantities for research
purposes. Furthermore, it occurs in spent nuclear fuel. Curium is
present in nature in certain areas used for the atmospheric nuclear
[19]
weapons tests, which were conducted between 1945 and 1980.
So the analysis of the debris at the testing site of the first U.S.
hydrogen bomb, Ivy Mike, (1 November 1952, Enewetak Atoll),
beside einsteinium, fermium, plutonium and americium also Several isotopes of curium were detected in the
revealed isotopes of berkelium, californium and curium, in particular fallout from the Ivy Mike nuclear test.
245 246 247 248 249
Cm, Cm and smaller quantities of Cm, Cm and Cm.
For reasons of military secrecy, this result was published only in
1956.

Atmospheric curium compounds are poorly soluble in common solvents and mostly adhere to soil particles. Soil
analysis revealed about 4,000 times higher concentration of curium at the sandy soil particles than in water present in
the soil pores. An even higher ratio of about 18,000 was measured in loam soils.
A few atoms of curium can be produced by neutron capture reactions and beta decay in very highly concentrated
uranium-bearing deposits.

Synthesis

Isotope preparation
Curium is produced in small quantities in nuclear reactors, and by now only kilograms of it have been accumulated
242 244
for the Cm and Cm and grams or even milligrams for heavier isotopes. This explains the high price of curium,
242
which has been be quoted at 160–185 USD per milligram, with a more recent estimate at 2,000 USD/g for Cm
244 238
and 170 USD/g for Cm. In nuclear reactors, curium is formed from U in a series of nuclear reactions. In the
238 239 – 239 239
first chain, U captures a neutron and converts into U, which via β decay transforms into Np and Pu.

(the times are half-lives).


– 241
Further neutron capture followed by β -decay produces the Am isotope of americium which further converts into
242
Cm:

For research purposes, curium is obtained by irradiating not uranium but plutonium, which is available in large
amounts from spent nuclear fuel. Much higher neutron flux is used for the irradiation that results in a different
244 []
reaction chain and formation of Cm:

240
Curium-244 decays into Pu by emission of alpha particle, but it also absorbs neutrons resulting in a small amount
247 248
of heavier curium isotopes. Among those, Cm and Cm are popular in scientific research because of their long
247
half-lives. However, the production rate of Cm in thermal neutron reactors is relatively low because of it is prone
250
to undergo fission induced by thermal neutrons. Synthesis of Cm via neutron absorption is also rather unlikely
249 –
because of the short half-life of the intermediate product Cm (64 min), which converts by β decay to the
249
berkelium isotope Bk.

(for A = 244–248)
The above cascade of (n,γ) reactions produces a mixture of different curium isotopes. Their post-synthesis separation
is cumbersome, and therefore a selective synthesis is desired. Curium-248 is favored for research purposes because
of its long half-life. The most efficient preparation method of this isotope is via α-decay of the californium isotope
252
Cf, which is available in relatively large quantities due to its long half-life (2.65 years). About 35–50 mg of
248 248
Cm is being produced by this method every year. The associated reaction produces Cm with isotopic purity of
97%.

245 249
Another interesting for research isotope Cm can be obtained from the α-decay of Cf, and the latter isotope is
– 249
produced in minute quantities from the β -decay of the berkelium isotope Bk.

Metal
preparation
Most synthesis routines yield a mixture of different actinide isotopes
as oxides, from which a certain isotope of curium needs to be
separated. An example procedure could be to dissolve spent
reactor fuel (e.g. MOX fuel) in nitric acid, and remove the bulk of
the uranium and plutonium using a PUREX (Plutonium –
URanium EXtraction) type extraction with tributyl phosphate in a
hydrocarbon. The lanthanides and the remaining actinides are
then separated from the aqueous residue (raffinate) by a diamide-
based extraction to give, after stripping, a mixture of trivalent
actinides and lanthanides. A curium compound is then selectively
extracted using multi-step chromatographic and centrifugation
[20]
techniques with an appropriate reagent. Bis-triazinyl bipyridine
complex has been recently proposed as such reagent which is
highly selective to curium. Separation of curium from a very
similar americium can also be achieved by treating a slurry of
their hydroxides in aqueous sodium bicarbonate with ozone at
Chromatographic elution curves revealing the
elevated temperature. Both americium and curium are present in similarity between Tb, Gd, Eu lanthanides and
solutions mostly in the +3 valence state; whereas americium corresponding Bk, Cm, Am actinides.
oxidizes to soluble Am(IV) complexes, curium remains unchanged
[21]
and can thus be isolated by repeated centrifugation.

Metallic curium is obtained by reduction of its compounds. Initially, curium(III) fluoride was used for this purpose.
The reaction was conducted in the environment free from water and oxygen, in the apparatus made of tantalum and
[22]
tungsten, using elemental barium or lithium as reducing agents.

Another possibility is the reduction of curium(IV) oxide using a magnesium-zinc alloy in a melt of magnesium
chloride and magnesium fluoride.
Compounds and reactions

Oxides
Curium readily reacts with oxygen forming mostly Cm O and CmO oxides, but the divalent oxide CmO is also
[23] 2 3 2
known. Black CmO can be obtained by burning curium oxalate (Cm (C O ) ), nitrate (Cm(NO ) ) or hydroxide
[24] 2 2 2 4 3 33
in pure oxygen. Upon heating to 600–650 °C in vacuum (about 0.01 Pa), it transforms into the whitish Cm O :
2 3
.
Alternatively, Cm O can be obtained by reducing CmO with molecular hydrogen:
2 3 2

Furthermore, a number of ternary oxides of the type M(II)CmO


3 are known, where M stands for a divalent metal,
such as barium.
Thermal oxidation of trace quantities of curium hydride (CmH ) has been reported to produce a volatile form of
2–3
CmO and the volatile trioxide CmO , one of the two known examples of the very rare +6 state for curium. Another
2 3
observed species was reported to behave similarly to plutonium tetroxide and was tentatively characterized as CmO ,
4
with curium in the extremely rare +8 state only known in this compound.

Halides
The colorless curium(III) fluoride (CmF ) can be produced by introducing fluoride ions into curium(III)-containing
3
solutions. The brown tetravalent curium(IV) fluoride (CmF ) on the other hand is only obtained by reacting
4
curium(III) fluoride with molecular fluorine:

A series of ternary fluorides are known of the form A Cm , where A stands for alkali metal.
F
7 6 31
The colorless curium(III) chloride (CmCl ) is produced in the reaction of curium(III) hydroxide (Cm(OH) ) with
3 3
anhydrous hydrogen chloride gas. It can further be converted into other halides, such as curium(III) bromide
(colorless to light green) and curium(III) iodide (colorless), by reacting it with the ammonia salt of the corresponding
halide at elevated temperature of about 400–450 °C:

An alternative procedure is heating curium oxide to about 600 °C with the corresponding acid (such as hydrobromic
for curium bromide). Vapor phase hydrolysis of curium(III) chloride results in curium oxychloride:
Chalcogenides and pnictides
Sulfides, selenides and tellurides of curium have been obtained by treating curium with gaseous sulfur, selenium or
[25]
tellurium in vacuum at elevated temperature. The pnictides of curium of the type CmX are known for the
elements nitrogen, phosphorus, arsenic and antimony. They can be prepared by reacting either curium(III) hydride
[26]
(CmH ) or metallic curium with these elements at elevated temperatures.
3

Organocurium compounds and biological aspects


Organometallic complexes analogous to uranocene are known also for other
actinides, such as thorium, protactinium, neptunium, plutonium and americium.
8
Molecular orbital theory predicts a stable "curocene" complex (η -C H ) Cm, but it
[27] 8 8 2
has not been reported experimentally yet.
Formation of the complexes of the type Cm(n-C H -BTP) , where BTP stands for
3 7 3 3+
2,6-di(1,2,4-triazin-3-yl)pyridine, in solutions containing n-C H -BTP and Cm
3 7
ions has been confirmed by EXAFS. Some of these BTP-type complexes selectively
interact with curium and therefore are useful in its selective separation from
3+
lanthanides and another actinides. Dissolved Cm ions bind with many organic
compounds, such as hydroxamic acid, urea, fluorescein and adenosine triphosphate. Predicted curocene structure
Many of these compounds are related to biological activity of various
microorganisms. The resulting complexes exhibit strong yellow-orange emission under UV light excitation, which is
3+
convenient not only for their detection, but also for studying the interactions between the Cm ion and the ligands
via changes in the half-life (of the order ~0.1 ms) and spectrum of the fluorescence.
3+
Curium has no biological significance. There are a few reports on biosorption of Cm by bacteria and archaea,
however no evidence for incorporation of curium into them.

Applications

Radionuclides
242 244
Curium is one of the most radioactive isolable elements. Its two most common isotopes Cm and Cm are
strong alpha emitters (energy 6 MeV); they have relatively short half-lives of 162.8 days and 18.1 years, and
[28][29]
produce as much as 120 W/g and 3 W/g of thermal energy, respectively. Therefore, curium can be used in
its common oxide form in radioisotope thermoelectric generators like those in spacecraft. This application has
244 242
been studied for the Cm isotope, while Cm was abandoned due to its prohibitive price of around 2000
USD/g. Curium-243 with a ~30 year half-life and good energy yield of ~1.6 W/g could make for a suitable fuel,
but it produces significant amounts of harmful gamma and beta radiation from radioactive decay products.
244
Though as an α-emitter, Cm requires a much thinner radiation protection shielding, it has a high
spontaneous fission rate, and thus the neutron and gamma radiation rate are relatively strong. As compared to
238 244
a competing thermoelectric generator isotope such as Pu, Cm emits a 500 time greater fluence of
neutrons, and its higher gamma emission requires a shield that is 20 times thicker — about 2 inches of lead for
238
a 1 kW source, as compared to 0.1 in for Pu. Therefore this application of curium is currently considered
[]
impractical.
242 238
A more promising application of Cm is to produce Pu, a more suitable radioisotope for thermoelectric
238 237
generators such as in cardiac pacemakers. The alternative routes to Pu use the (n,γ) reaction of Np, or the
236
deuteron bombardment of uranium, which both always produce Pu as an undesired by-product — since the latter
208 [30]
decays to Tl with strong gamma emission.
Curium is also a common starting material for the production of higher transuranic elements and transactinides.
248 18 26
Thus, bombardment of Cm with oxygen ( O) or magnesium ( Mg) yielded certain isotopes of seaborgium
265 269 270 [31]
( Sg) and hassium ( Hs and Hs). Californium was discovered when a microgram-sized target of
curium-242 was irradiated with 35 MeV alpha particles using the 60-inch (150 cm) cyclotron at Berkeley:
242
96Cm + 4
2He → 245
98Cf + 1
0n

Only about 5,000 atoms of californium were produced in this experiment.

X-ray spectrometer
244
The most practical application of Cm — though rather limited in
total volume — is as α-particle source in the alpha particle X-ray
spectrometers (APXS). These instruments were installed on the
Sojourner, Mars, Mars 96, Spirit, Athena and Opportunity rovers, as
well as the Mars Science Laboratory to analyze the composition and
structure of the rocks on the surface of planet Mars. APXS was also
242 [][32]
used in the Surveyor 5–7 moon probes but with a Cm source.

An elaborated APXS setup is equipped with a sensor head containing


six curium sources having the total radioactive decay rate of several
Alpha-particle X-ray spectrometer of a Mars
tens of millicuries (roughly a gigabecquerel). The sources are exploration rover
collimated on the sample, and the energy spectra of the alpha particles
and protons scattered from the sample are analyzed (the proton analysis is implemented only in some
spectrometers). These spectra contain quantitative information on all major elements in the samples except for
hydrogen, helium and
[33]
lithium. An APXS will also be used by the Philae lander of the Rosetta spacecraft to probe the surface of the
67P/Churyumov-Gerasimenko comet.

Safety
Owing to its high radioactivity, curium and its compounds must be handled in appropriate laboratories under special
arrangements. Whereas curium itself mostly emits α-particles which are absorbed by thin layers of common
materials, some of its decay products emit significant fractions of beta and gamma radiation, which require a more
elaborate protection. If consumed, curium is excreted within a few days and only 0.05% is absorbed in the blood.
From there, about 45% goes to the liver, 45% to the bones, and the remaining 10% is excreted. In the bone, curium
accumulates on the inside of the interfaces to the bone marrow and does not significantly redistribute with time; its
radiation destroys bone marrow and thus stops red blood cell creation. The biological half-life of curium is about 20
years in the liver and 50 years in the bones. Curium is absorbed in the body much more strongly via inhalation, and
244 242 244
the allowed total dose of Cm in soluble form is 0.3 μC. Intravenous injection of Cm and Cm containing
solutions to rats increased the incidence of bone tumor, and inhalation promoted pulmonary and liver cancer.
[34]
Curium isotopes are inevitably present in spent nuclear fuel with a concentration of about 20 g/tonne. Among
245 248
them, the Cm– Cm isotopes have decay times of thousands of years and need to be removed to neutralize the
[35]
fuel for disposal. The associated procedure involves several steps, where curium is first separated and then
converted by neutron bombardment in special reactors to short-lived nuclides. This procedure, nuclear transmutation,
while well documented for other elements, is still being developed for curium.
References
[1] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_curium&action=edit
[2] Seaborg, G. T.; James, R. A. and Ghiorso, A.: "The New Element Curium (Atomic Number 96)", NNES PPR (National Nuclear
Energy Series, Plutonium Project Record), Vol. 14 B, The Transuranium Elements: Research Papers, Paper No. 22.2, McGraw-Hill
Book Co., Inc., New York, 1949; Abstract (http:// www.osti.gov/cgi-bin/rd_accomplishments/ display_biblio.cgi?
id=ACC0049&numPages=13&fp=N); Full text (January 1948) (http://www.osti.gov/accomplishments/ documents/ fullText/ACC0049.pdf).
[3] Krebs, Robert E. The history and use of our earth's chemical elements: a reference guide (http:/ /books.google. com/ books?
id=yb9xTj72vNAC&pg=PA322), Greenwood Publishing Group, 2006, ISBN 0-313-33438-2 p. 322
[4] Seaborg, G. T. "Element", Filing date: 7 February 1949, Issue date: December 1964
[5] Greenwood, p. 1252
[6] Hammond C. R. "The elements" in
[7] L. B. Werner, I. Perlman: "Isolation of Curium", NNES PPR (National Nuclear Energy Series, Plutonium Project Record), Vol. 14 B, The
Transuranium Elements: Research Papers, Paper No. 22.5, McGraw-Hill Book Co., Inc., New York, 1949.
[8] Young, D. A. Phase diagrams of the elements (http:/ /books.google. com/ books?id=F2HVYh6wLBcC&pg=PA227), University of
California Press, 1991, ISBN 0-520-07483-1p. 227
[9] Nave, S. E.; Huray, P. G.; Peterson, J. R. and Damien, D. A. Magnetic susceptibility of curium pnictides (http://www.osti.gov/bridge/purl.
cover.jsp;jsessionid=ECF73C70531D64E8B663048ECE8C10F9?purl=/6263633-jkoGGI/), Oak Ridge National Laboratory
[10] Bünzli, J.-C. G. and Choppin, G. R. Lanthanide probes in life, chemical, and earth sciences: theory and practice, Elsevier, Amsterdam,
1989 ISBN 0-444-88199-9
[11] Penneman, p. 24
[12] Greenwood, p. 1265
[13] Holleman, p. 1956
[14] Penneman, pp. 25–26
[15] Greenwood, p. 1267
[16] Pfennig, G.; Klewe-Nebenius, H. and Seelmann Eggebert, W. (Eds.): Karlsruhe nuclide, 6th Ed. 1998
[17] Institut de Radioprotection et de Sûreté Nucléaire: "Evaluation of nuclear criticality safety. data and limits for actinides in transport" (http://
ec.europa.eu/ energy/ nuclear/ transport/ doc/irsn_sect03_146.pdf), p. 16
[18] § 2 Begriffsbestimmungen (Atomic Energy Act) (http:/ /bundesrecht.juris.de/ atg/ 2.html) (in German)
[19] Curium (http:/ /www.lenntech. de/ pse/ pse. htm) (in German)
[20] Penneman, pp. 34–48
[21] Penneman, p. 25
[22] Gmelin Handbook of Inorganic Chemistry, System No. 71, Volume 7 a, transuranics, Part B 1, pp. 67–68.
[23] Holleman, p. 1972
[24] Greenwood, p. 1268
[25] Troc, R. Actinide Monochalcogenides, Volume 27 (http://books. google.com/books?id=vkzx_t3zLR0C&pg=PA4), Springer, 2009 ISBN 3-540-
29177-6, p. 4
[26] Lumetta, G. J.; Thompson, M. C.; Penneman, R. A.; Eller, P. G. Curium (http://radchem.nevada.edu/classes/ rdch710/files/curium.pdf),
Chapter Nine in Radioanalytical Chemistry, Springer, 2004, pp. 1420-1421. ISBN 0387341226, ISBN 978-0387 341224
[27] Elschenbroich, Ch. Organometallic Chemistry, 6th edition, Wiesbaden 2008, ISBN 978-3-8351-0167-8, p. 589
[28] Binder, Harry H.: Lexikon der chemischen Elemente, S. Hirzel Verlag, Stuttgart 1999, ISBN 3-7776-0736-3, pp. 174–178.
[29] Gmelin Handbook of Inorganic Chemistry, System No. 71, Volume 7a, transuranics, Part A2, p. 289
[30] Kronenberg, Andreas (http:// www.kronenberg. kernchemie. de/), Plutonium-Batterien (http://www.kernenergie-wissen.de/ pu-batterien.
html) (in German)
[31] Holleman, pp. 1980–1981.
[32] Leitenberger, Bernd Die Surveyor Raumsonden (http:// www.bernd-leitenberger.de/ surveyor.shtml) (in German)
[33] Alpha Particle X-Ray Spectrometer (APXS) (http://web.archive.org/web/20060302040531/http://athena. cornell.edu/pdf/tb_apxs.
pdf), Cornell University
[34] Hoffmann, K. Kann man Gold machen? Gauner, Gaukler und Gelehrte. Aus der Geschichte der chemischen Elemente (Can you make gold?
Crooks, clowns and scholars. From the history of the chemical elements), Urania-Verlag, Leipzig, Jena, Berlin 1979, no ISBN, p. 233
[35] Baetslé, L. H. Application of Partitioning/Transmutation of Radioactive Materials in Radioactive Waste Management (http:// www.ictp.
trieste.it/~pub_off/lectures/lns012/Baetsle. pdf), Nuclear Research Centre of Belgium Sck/Cen, Mol, Belgium, September 2001.
Bibliography
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
ISBN 0080379419.
• Holleman, Arnold F. and Wiberg, Nils Textbook of Inorganic Chemistry, 102 Edition, de Gruyter, Berlin 2007,
ISBN 978-3-11-017770-1.
• Penneman, R. A. and Keenan T. K. The radiochemistry of americium and curium
(http://www.osti.gov/bridge/ purl.cover.jsp?purl=/4187189-IKQUwY/), University of California, Los
Alamos, California, 1960

External links
• Curium (http://www.periodicvideos.com/videos/096.htm) at The Periodic Table of Videos (University
of Nottingham)
• NLM Hazardous Substances Databank – Curium, Radioactive (http://toxnet.nlm.nih.gov/cgi-bin/sis/search/
r?dbs+hsdb:@term+@na+@rel+curium,+radioactive)
97 Berkelium 1

97 – Berkelium
Berkelium
Bk
97

curium ← berkelium → californiumTb



Bk

(Uqe)

Berkelium in the periodic table


Appearance

silvery

General properties

Name, symbol, number berkelium, Bk, 97

Pronunciation /bərˈkiːli.əm/
bər-KEE-lee-
əm less
commonly:
/ˈbɜrkli.əm/
BERK-lee-əm
Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (247)

Electron configuration 9 2
[Rn] 5f 7s
2, 8, 18, 32, 27, 8, 2

Physical properties

Phase solid

Density (near r.t.) −3


(alpha) 14.78 g·cm
Density (near r.t.) −3
(beta) 13.25 g·cm

Melting point (beta) 1259 K, 986 °C, 1807 °F

Boiling point (beta) 2900 K, 2627 °C, 4760 °F

Atomic properties

Oxidation states 3, 4

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 601 kJ·mol

Atomic radius 170 pm

Miscellanea

Crystal structure hexagonal close-packed

Magnetic ordering paramagnetic

Thermal conductivity −1 −1
10 W·m ·K

CAS registry number 7440-40-6

History

Naming after Berkeley, California, where it was


discovered
Discovery Lawrence Berkeley National Laboratory (1949)

Most stable isotopes

Main article: Isotopes of berkelium

iso NA half-life DM DE (MeV) DP

245 syn 4.94 d ε 0.810 245


Bk Cm
α 6.455 241
Am

246 syn 1.8 d α 6.070 242


Bk Am
ε 1.350 246
Cm

247 syn 1380 y α 5.889 243


Bk Am

248 syn >9 y α 5.803 244


Bk Am

249 trace 330 d α 5.526 245


Bk Am
SF - -

− 0.125 249
β Cf

• v
• t
[1]
• e

Berkelium is a transuranic radioactive chemical element with the symbol Bk and atomic number 97, a member of
the actinide and transuranium element series. It is named after the city of Berkeley, California, the location of the
University of California Radiation Laboratory where it was discovered in December 1949. This was the fifth
transuranium element discovered after neptunium, plutonium, curium and americium.
The major isotope of berkelium, berkelium-249, is synthesized in minute quantities in dedicated high-flux nuclear
reactors, mainly at the Oak Ridge National Laboratory in Tennessee, USA, and at the Research Institute of Atomic
Reactors in Dimitrovgrad, Russia. The production of the second-important isotope berkelium-247 involves the
irradiation of the rare isotope curium-244 with high-energy alpha particles.
Just over one gram of berkelium has been produced in the United States since 1967. There is no practical application
of berkelium outside of scientific research which is mostly directed at the synthesis of heavier transuranic elements
and transactinides. A 22 milligram batch of berkelium-249 was prepared during a 250-day irradiation period and
then purified for a further 90 days at Oak Ridge in 2009. This sample was used to synthesize the element
ununseptium for the first time in 2009 at the Joint Institute for Nuclear Research, Russia, after it was bombarded
with calcium-48 ions for 150 days. This was a culmination of the Russia–US collaboration on the synthesis of
elements 113 to 118.
Berkelium is a soft, silvery-white, radioactive metal. The berkelium-249 isotope emits low-energy electrons and
thus is relatively safe to handle. It decays with a half-life of 330 days to californium-249, which is a strong
emitter of ionizing alpha particles. This gradual transformation is an important consideration when studying the
properties of elemental berkelium and its chemical compounds, since the formation of californium brings not
only chemical contamination, but also free-radical effects and self-heating from the emitted helium nuclei.

Characteristics

Physical
Berkelium is a soft, silvery-white, radioactive actinide metal. In the
periodic table, it is located to the right of the actinide curium, to the
left of the actinide californium and below the lanthanide terbium
with which it shares many similarities in physical and chemical
3
properties. Its density of 14.78 g/cm lies between those of curium
3 3
(13.52 g/cm ) and californium (15.1 g/cm ), as does its melting
point of 986 °C, below that of curium (1340 °C) but higher than
that of californium (900 °C). Berkelium is relatively soft and has one
of the lowest bulk moduli among the actinides, at about 20 GPa
10
(2×10 Pa).

Berkelium(III) ions shows two sharp fluorescence peaks at


652 nanometers (red light) and 742 nanometers (deep red – near Double-hexagonal close packing with the layer
infrared) due to internal transitions at the f-electron shell. The relative sequence ABAC in the crystal structure of
intensity of these peaks depends on the excitation power and α-berkelium (A: green, B: blue, C: red)

temperature of the sample. This emission can be observed, for


example, after dispersing berkelium ions in a silicate glass, by melting the glass in presence of berkelium oxide or
[2]
halide.

Between 70 K and room temperature, berkelium behaves as a Curie–Weiss paramagnetic material with an effective
magnetic moment of 9.69 Bohr magnetons (µ ) and a Curie temperature of 101 K. This magnetic moment is almost
B
equal to the theoretical value of 9.72 µ B calculated within the simple atomic L-S coupling model. Upon cooling to
[3]
about 34 K, berkelium undergoes a transition to an antiferromagnetic state. Enthalpy of dissolution in hydrochloric
−1
acid at standard conditions is −600 kJ/mol , from which the standard enthalpy change of formation (Δ H°) of
f
3+ −1 3+ 0
aqueous Bk ions is obtained as −601 kJ/mol . The standard potential Bk /Bk is −2.01 V. The ionization
[4]
potential of a neutral berkelium atom is 6.23 eV.
Allotrope
s
At ambient conditions, berkelium assumes its most stable α form which has a hexagonal symmetry, space group
P6 /mmc, lattice parameters of 341 pm and 1107 pm. The crystal has a double-hexagonal close packing structure
3
with the layer sequence ABAC and so is isotypic (having a similar structure) with α-lanthanum and α-forms of
actinides beyond curium. This crystal structure changes with pressure and temperature. When compressed at room
temperature to 7 GPa, α-berkelium transforms to the beta modification, which has a face-centered cubic (fcc)
symmetry and space group Fm3m. This transition occurs without change in volume, but the enthalpy increases by
[5]
3.66 kJ/mol. Upon further compression to 25 GPa, berkelium transforms to an orthorhombic γ-berkelium structure
similar to that of α-uranium. This transition is accompanied by a 12% volume decrease and delocalization of the
[6]
electrons at the 5f electron shell. No further phase transitions are observed up to 57 GPa.
Upon heating, α-berkelium transforms into another phase with an fcc lattice (but slightly different from β-
berkelium), space group Fm3m and the lattice constant of 500 pm; this fcc structure is equivalent to the closest
packing with the sequence ABC. This phase is metastable and will gradually revert to the original α-berkelium
phase at room temperature. The temperature of the phase transition is believed to be quite close to the melting
point.

Chemical
Like all actinides, berkelium dissolves in various aqueous inorganic acids, liberating gaseous hydrogen and
converting into the berkelium(III) state. This trivalent oxidation state (+3) is the most stable, especially in
aqueous solutions, but tetravalent (+4) and possibly divalent (+2) berkelium compounds are also known. The
existence of divalent berkelium salts is uncertain and has only been reported in mixed lanthanum chloride-
[7]
strontium chloride melts. A similar behavior is observed for the lanthanide analogue of berkelium, terbium.
3+ 4+
Aqueous solutions of Bk ions are green in most acids. The color of Bk ions is yellow in hydrochloric acid
[7][8][9]
and orange-yellow in sulfuric acid. Berkelium does not react rapidly with oxygen at room
temperature, possibly due to the formation of a protective oxide layer surface. However, it reacts with molten
[3]
metals, hydrogen, halogens, chalcogens and pnictogens to form various binary compounds.

Isotopes
Main article: Isotopes of
berkelium
About twenty isotopes and six nuclear isomers (excited states of an isotope) of berkelium have been characterized
with the mass numbers ranging from 235 to 254. All of them are radioactive. The longest half-lives are observed for
247 248 249
Bk (1,380 years), Bk (9 years) and Bk (330 days); the half-lives of the other isotopes range from
microseconds to several days. The isotope which is the easiest to synthesize is berkelium-249. This emits mostly soft
−3
β-particles which are inconvenient for detection. Its alpha radiation is rather weak – 1.45×10 % with respect to the
β-radiation – but is sometimes used to detect this isotope. The second important berkelium isotope, berkelium-247,
is an alpha-emitter, as are most actinide isotopes.

Occurrence
All berkelium isotopes have a half-life far too short to be primordial. Therefore, all primordial berkelium, that is,
berkelium present on the Earth during its formation, has decayed by now.
On Earth, berkelium is mostly concentrated in certain areas, which were used for the atmospheric nuclear weapons
tests between 1945 and 1980, as well as at the sites of nuclear incidents, such as the Chernobyl disaster, Three Mile
Island accident and 1968 Thule Air Base B-52 crash. Analysis of the debris at the testing site of the first U.S.
hydrogen bomb, Ivy Mike, (1 November 1952, Enewetak Atoll), revealed high concentrations of various actinides,
including berkelium. For reasons of military secrecy, this result was published only in 1956.
Nuclear reactors produce mostly, among the berkelium isotopes, berkelium-249. During the storage and before the
fuel disposal, most of it beta decays to californium-249. The latter has a half-life of 351 years, which is relatively
long when compared to the other isotopes produced in the reactor, and is therefore undesirable in the disposal
products.
A few atoms of berkelium can be produced by neutron capture reactions and beta decay in very highly concentrated
uranium-bearing deposits, thus making it the rarest naturally occurring element.

History
Although very small amounts of berkelium were possibly produced in
previous nuclear experiments, it was first intentionally synthesized, isolated
and identified in December 1949 by Glenn T. Seaborg, Albert Ghiorso
and Stanley G. Thompson. They used the 60-inch cyclotron at the
University of California, Berkeley. Similar to the nearly simultaneous
discovery of americium (element
95) and curium (element 96) in 1944, the new elements berkelium and
[10]
californium (element 98) were both produced in 1949–1950.

The name choice for element 97 followed the previous tradition of the
Californian group to draw an analogy between the newly discovered actinide and
the lanthanide element positioned above it in the periodic table. Previously,
Glenn T. Seaborg americium was named after a continent as its analogue europium, and curium
honored scientists Marie and Pierre Curie as the lanthanide above it, gadolinium,
was named after the explorer of the rare earth elements Johan Gadolin. Thus the
discovery report by the Berkeley group reads: "It is suggested that element 97 be
given the name berkelium (symbol Bk) after the city of Berkeley in a manner
similar to that used in naming its chemical homologue terbium (atomic number
65) whose name was derived from the town of Ytterby, Sweden, where the rare
earth minerals were first found." This tradition ended on berkelium, though, as
the naming of the next discovered actinide, californium, was not related to its
lanthanide analogue dysprosium, but after the discovery place.
The 60-inch cyclotron at the
Lawrence Radiation Laboratory, The most difficult steps in the synthesis of berkelium were its separation from
University of California, Berkeley, in
the final products and the production of sufficient quantities of americium for the
August 1939 241
target material. First, americium ( Am) nitrate solution was coated on a
platinum foil, the solution was evaporated and the residue converted by
annealing to americium dioxide (AmO ). This target was irradiated with 35 MeV
2
alpha particles for 6 hours in the 60-inch cyclotron at the Lawrence Radiation
Laboratory, University of California, Berkeley. The (α,2n) reaction induced by
243
the irradiation yielded the Bk isotope and two free neutrons:

After the irradiation, the coating was dissolved with nitric acid and then
precipitated as the hydroxide using concentrated aqueous ammonia solution. The
product was centrifugated and re-dissolved in nitric acid. To separate berkelium
Berkelium is named after from the unreacted americium, this solution was added to a mixture of
UC Berkeley
ammonium and ammonium sulfate and heated to convert all the dissolved
americium into the oxidation state +6. Unoxidized residual americium was
precipitated by the addition of hydrofluoric acid as americium(III) fluoride
(AmF
3). This step yielded a mixture of the accompanying product curium and the expected element 97 in form of

trifluorides. The mixture was converted to the corresponding hydroxides by treating it with potassium hydroxide,
and after centrifugation, was dissolved in perchloric acid.
Further separation was carried out in the presence of a citric
acid/ammonium buffer solution in a weakly acidic medium (pH≈3.5),
using ion exchange at elevated temperature. The chromatographic
separation behavior was then unknown for the element 97, but was
anticipated by analogy with terbium (see elution curves). First results
were disappointing as no alpha-particle emission signature could be
detected from the elution product. Only the further search for
characteristic X-rays and conversion electron signals resulted in the
identification of a berkelium isotope. Its mass number was uncertain
between 243 and 244 in the initial report, but was later established as
243.

Synthesis and
extraction

Preparation of
isotopes Chromatographic elution curves revealing the
similarity between the lanthanides terbium (Tb),
238
Berkelium is produced by bombarding lighter actinides uranium ( U) gadolinium (Gd), and europium (Eu) and their
239 corresponding actinides berkelium (Bk), curium
or plutonium ( Pu) with neutrons in a nuclear reactor. In a more
(Cm), and americium (Am)
common case of uranium fuel, plutonium is produced first by neutron
capture (the so-called (n,γ) reaction or neutron fusion) followed by beta-
decay:

(the times are half-lives)

Plutonium-239 is further irradiated by a source that has a high neutron flux, several times higher than a conventional
nuclear reactor, such as the 85-megawatt High Flux Isotope Reactor (HFIR) at the Oak Ridge National Laboratory in
239
Tennessee, USA. The higher flux promotes fusion reactions involving not one but several neutrons, converting Pu
244 249
to Cm and then to Cm:

250
Curium-249 has a short half-life of 64 minutes, and thus its further conversion to Cm has a low probability.
249
Instead, it transforms by beta-decay into Bk:

249
The thus-produced Bk has a long half-life of 330 days and thus can capture another neutron. However, the
250
product, Bk, again has a relatively short half-life of 3.212 hours and thus, does not yield any heavier berkelium
250
isotopes. Instead decays to the californium isotope Cf:

247
Although Bk is the most stable isotope of berkelium, its production in nuclear reactors is very inefficient due to
the long half-life of its potential progenitor curium-247, which does not allow it sufficient time to beta decay before
249
capturing another neutron. Thus, Bk is the most accessible isotope of berkelium, which still, is available only in
[11]
small quantities (only 0.66 grams have been produced in the US over the period 1967–1983 ) at a high price of the
[6]
order 185 USD per microgram.
248
The isotope Bk was first obtained in 1956 by bombarding a mixture of curium isotopes with 25 MeV α-particles.
245
Although its direct detection was hindered by strong signal interference with Bk, the existence of a new isotope
248 248
was proven by the growth of the decay product Cf which had been previously characterized. The half-life of Cf
was estimated as 23 ± 5 hours and a more reliable value still is not known. Berkelium-247 was produced during the
244
same year by irradiating Cm with alpha-particles:

235 11 238 10 232 14 232


Berkelium-242 was synthesized in 1979 by bombarding U with B, U with B, Th with N or Th
15 242
with N. It converts by electron capture to Cm with a half-life of 7.0 ± 1.3 minutes. A search for an initially
241 241
suspected isotope Bk was then unsuccessful; Bk has since been synthesized.

Separation
The fact that berkelium readily assumes oxidation state +4 in solids, and is relatively stable in this state in liquids
greatly assists separation of berkelium away from many other actinides. These are inevitably produced in relatively
large amounts during the nuclear synthesis and often favor the +3 state. This fact was not yet known in the initial
experiments, which used a more complex separation procedure. Various oxidation agents can be applied to the
berkelium(III) solutions to convert it to the +4 state, such as bromates (BrO−
3), bismuthates (BiO−

3 ), chromates
(CrO2−
4 and
Cr
2O2−
7), silver(I) thiolate (Ag
2S

2O

8), lead(IV) oxide


(PbO
2 ), ozone
(O
3 ), or photochemical oxidation procedures. Berkelium(IV) is then extracted with ion exchange, extraction
chromatography or liquid-liquid extraction using HDEHP (bis-(2-ethylhexyl) phosphoric scid), amines, tributyl
phosphate or various other reagents. These procedures separate berkelium from most trivalent actinides and
lanthanides, except for the lanthanide cerium (lanthanides are absent in the irradiation target but are created in
[12]
various nuclear fission decay chains).
A more detailed procedure adopted at the Oak Ridge National Laboratory was as follows: the initial mixture of
actinides is processed with ion exchange using lithium chloride reagent, then precipitated as hydroxides, filtered and
dissolved in nitric acid. It is then treated with high-pressure elution from cation exchange resins, and the berkelium
[12]
phase is oxidized and extracted using one of the procedures described above. Reduction of the thus-obtained
berkelium(IV) to the +3 oxidation state yields a solution, which is nearly free from other actinides (but contains
[13]
cerium). Berkelium and cerium are then separated with another round of ion-exchange treatment.
Bulk metal preparation
In order to characterize chemical and physical properties of solid berkelium and its compounds, a program was
initiated in 1952 at the Material Testing Reactor, Arco, Idaho, US. It resulted in preparation of an eight-gram
plutonium-239 target and in the first production of macroscopic quantities (0.6 micrograms) of berkelium by
Burris
B. Cunningham and Stanley G. Thompson in 1958, after a continuous reactor irradiation of this target for six
[11][14]
years. This irradiation method was and still is the only way of producing weighable amounts of the element,
[15]
and most solid-state studies of berkelium have been conducted on microgram or submicrogram-sized samples.
The world's major irradiation sources are the 85-megawatt High Flux Isotope Reactor at the Oak Ridge National
Laboratory in Tennessee, USA, and the SM-2 loop reactor at the Research Institute of Atomic Reactors (NIIAR) in
Dimitrovgrad, Russia, which are both dedicated to the production of transcurium elements (atomic number greater
than 96). These facilities have similar power and flux levels, and are expected to have comparable production
capacities for transcurium elements, although the quantities produced at NIIAR are not publicly reported. In a
"typical processing campaign" at Oak Ridge, tens of grams of curium are irradiated to produce decigram quantities
[16]
of californium, milligram quantities of berkelium-249 and einsteinium, and picogram quantities of fermium. In
total, just over one gram of berkelium-249 has been produced at Oak Ridge since 1967.
The first berkelium metal sample weighing 1.7 micrograms was prepared in 1971 by the reduction of berkelium(III)
fluoride with lithium vapor at 1000 °C; the fluoride was suspended on a tungsten wire above a tantalum crucible
[17]
containing molten lithium. Later, metal samples weighting up to 0.5 milligrams were obtained with this method.

Similar results are obtained with berkelium(IV) fluoride. Berkelium metal can also be produced by the reduction of
[17]
berkelium(IV) oxide with thorium or lanthanum.

Compounds
Main article: Compounds of berkelium

Oxides
Two oxides of berkelium are known, with the berkelium oxidation state of +3 (Bk O ) and +4 (BkO ).
2 3 2
Berkelium(IV) oxide is a brown solid, while berkelium(III) oxide is a yellow-green solid with a melting point of
[18]
1920 °C and is formed from BkO by reduction with molecular hydrogen:
2

Upon heating to 1200 °C, the oxide Bk O undergoes a phase change; it undergoes another phase change at 1750 °C.
2 3
Such three-phase behavior is typical for the actinide sesquioxides. Berkelium(II) oxide, BkO, has been reported as a
[19]
brittle gray solid but its exact chemical composition remains uncertain.

Halides
[20]
In halides, berkelium assumes the oxidation states +3 and +4. The +3 state is the most stable, especially in
solutions, while the tetravalent halides BkF and Cs BkCl are only known in the solid phase.[21] The coordination of
4 2 6
berkelium atom in its trivalent fluoride and chloride is tricapped trigonal prismatic, with the coordination number of
9. In trivalent bromide, it is bicapped trigonal prismatic (coordination 8) or octahedral (coordination 6), and in the
[22]
iodide it is octahedral.
Oxidation number F Cl Br I

+3 BkF BkCl BkBr BkI


(yellow ) (green ) (yellow-green ) (yellow )
[23]
Cs NaBkCl

+4 BkF Cs BkCl
(yellow ) (orange )

Berkelium(IV) fluoride (BkF ) is a yellow-green ionic solid and is isotypic with uranium tetrafluoride or
[23] 4
zirconium(IV) fluoride. Berkelium(III) fluoride (BkF ) is also a yellow-green solid, but it has two crystalline
3
structures. The most stable phase at low temperatures is isotypic with yttrium(III) fluoride, while upon heating to
[23]
between 350 and 600 °C, it transforms to the structure found in lanthanum(III) fluoride.
Visible amounts of berkelium(III) chloride (BkCl ) were first isolated and characterized in 1962, and weighed
only 3
3
billionths of a gram. It can be prepared by introducing hydrogen chloride vapors into an evacuated quartz tube
[20]
containing berkelium oxide at a temperature about 500 °C. This green solid has a melting point of 600 °C, and is
isotypic with uranium(III) chloride. Upon heating to nearly melting point, BkCl3 converts into an orthorhombic
[24]
phase.
Two forms of berkelium(III) bromide are known: one with berkelium having coordination 6, and one with
[15]
coordination 8. The latter is less stable and transforms to the former phase upon heating to about 350 °C. An
important phenomenon for radioactive solids has been studied on these two crystal forms: the structure of fresh and
249
aged BkBr samples was probed by X-ray diffraction over a period longer than 3 years, so that various fractions
3
of berkelium-249 had beta decayed to californium-249. No change in structure was observed upon the
249 249 249 249
BkBr — CfBr transformation. However, other differences were noted for BkBr and CfBr . For
3 3 3 3
249
example, the latter could be reduced with hydrogen to CfBr , but the former could not – this result was
2
249 249
reproduced on individual BkBr and CfBr samples, as well on the samples containing both bromides. The
3 3
intergrowth of californium in berkelium occurs at a rate of 0.22% per day and is an intrinsic obstacle in studying
249
berkelium properties. Beside a chemical contamination, Cf, being an alpha emitter, brings undesirable self-
damage of the crystal lattice and the resulting self-heating. The chemical effect however can be avoided by
[21]
performing measurements as a function of time and extrapolating the obtained results.

Other inorganic
compounds
The pnictides of berkelium-249 of the type BkX are known for the elements nitrogen, phosphorus, arsenic and
antimony. They crystallize in the rock-salt structure and are prepared by the reaction of either berkelium(III) hydride
(BkH ) or metallic berkelium with these elements at elevated temperature (about 600 °C) under high vacuum.
3
Berkelium(III) sulfide, Bk S , is prepared by either treating berkelium oxide with a mixture of hydrogen sulfide and
2 3
carbon disulfide vapors at 1130 °C, or by directly reacting metallic berkelium with elemental sulfur. These
[25]
procedures yield brownish-black crystals.
Berkelium(III) and berkelium(IV) hydroxides are both stable in 1 molar solutions of sodium hydroxide.
Berkelium(III) phosphate (BkPO )4 has been prepared as a solid, which shows strong fluorescence under excitation
[26]
with a green light. Berkelium hydrides are produced by reacting metal with hydrogen gas at temperatures about
250 °C. They are non-stoichiometric with the nominal formula BkH [25]
2+x (0 < x < 1). Several other salts of
berkelium are known, including an oxysulfide (Bk O S), and hydrated nitrate (Bk(NO
2 2
3)

3·4H

2O), chloride (BkCl


3·6H
2O), sulfate (Bk
2(SO

4)

3·12H

2O) and oxalate (Bk

2(C

2O

4)

3·4H
[21]
2O). Thermal decomposition at about 600 °C in an argon atmosphere (to avoid oxidation to BkO
2) of Bk

2(SO

4)

3·12H

2O yields the crystals of berkelium(III) oxysulfate (Bk

2O

2SO
[27]
4). This compound is thermally stable to at least 1000 °C in inert atmosphere.

Organoberkelium compounds
5
Berkelium forms a trigonal (η –C H ) Bk metallocene complex with three cyclopentadienyl rings, which can be
5 5 3
synthesized by reacting berkelium(III) chloride with the molten beryllocene (Be(C H ) ) at about 70 °C. It has an
5 5 2
3
amber color and a density of 2.47 g/cm . The complex is stable to heating to at least 250 °C, and sublimates without
melting at about 350 °C. The high radioactivity of berkelium gradually destroys the compound (within a period of
[28] 5
weeks). One cyclopentadienyl ring in (η –C 5H 5) 3Bk can be substituted by chlorine to yield [Bk(C 5H 5) 2Cl] 2. The
5 [17][27]
optical absorption spectra of this compound are very similar to those of (η –C H ) Bk.
5 5 3

Applications
There is currently no use for any isotope of berkelium outside of basic
scientific research. Berkelium-249 is a common target nuclide to
prepare still heavier transuranic elements and transactinides, such as
lawrencium, rutherfordium and bohrium. It is also useful as a source of
the isotope californium-249, which is used for studies on the chemistry
of californium in preference to the more radioactive californium-252
that is produced in neutron bombardment facilities such as the HFIR.

A 22 milligram batch of berkelium-249 was prepared in a 250-day


irradiation and then purified for 90 days at Oak Ridge in 2009. This
target yielded the first 6 atoms of ununseptium at the Joint Institute for
Nuclear Research (JINR), Dubna, Russia, after bombarding it with The berkelium target used for the synthesis of
[29]
calcium ions in the U400 cyclotron for 150 days. This synthesis was a ununseptium (in dissolved state)
culmination of the Russia—US collaboration between JINR and
Lawrence Livermore National Laboratory on the synthesis of elements 113 to 118 which was initiated in
[30][31]
1989.
Nuclear fuel
cycle
The nuclear fission properties of berkelium are different from those of the neighboring actinides curium and
californium, and they suggest berkelium to perform poorly as a fuel in a nuclear reactor. Specifically, berkelium-249
has a moderately large neutron capture cross section of 710 barns for thermal neutrons, 1200 barns resonance
integral, but very low fission cross section for thermal neutrons. In a thermal reactor, much of it will therefore be
[32]
converted to berkelium-250 which quickly decays to californium-250. In principle, berkelium-249 can sustain a
nuclear chain reaction in a fast breeder reactor. Its critical mass is relatively high at 192 kg; it can be reduced with a
[33]
water or steel reflector but would still exceed the world production of this isotope.
Berkelium-247 can maintain chain reaction both in a thermal-neutron and in a fast-neutron reactor, however, its
production is rather complex and thus the availability is much lower than its critical mass, which is about 75.7 kg for
a bare sphere, 41.2 kg with a water reflector and 35.2 kg with a steel reflector (30 cm thickness).

Health issues
Little is known about the effects of berkelium on human body, and analogies with other elements may not be
drawn because of different radiation products (electrons for berkelium and alpha particles, neutrons, or both for
most other actinides). The low energy of electrons emitted from berkelium-249 (less than 126 keV) hinders its
detection, due to signal interference with other decay processes, but also makes this isotope relatively
harmless to humans as compared to other actinides. However, berkelium-249 transforms with a half-life of
only 330 days to the strong alpha-emitter californium-249, which is rather dangerous and has to be handled in
[34]
a glove box in a dedicated laboratory.
Most available berkelium toxicity data originate from research on animals. Upon ingestion by rats, only about 0.01%
berkelium ends in the blood stream. From there, about 65% goes to the bones, where it remains for about 50 years,
25% to the lungs (biological half-life about 20 years), 0.035% to the testicles or 0.01% to the ovaries where
[35]
berkelium stays indefinitely. The balance of about 10% is excreted. In all these organs berkelium might promote
cancer, and in the skeletal system its radiation can damage red blood cells. The maximum permissible amount of
[36]
berkelium-249 in the human skeleton is 0.4 nanograms.

References
[1] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_berkelium&action=edit
[2] Rita Cornelis, Joe Caruso, Helen Crews, Klaus Heumann Handbook of elemental speciation II: species in the environment, food,
medicine & occupational health. Volume 2 of Handbook of Elemental Speciation (http://books. google. com/books?
id=1PmjurlE6KkC&pg=PA553),
John Wiley and Sons, 2005, ISBN 0-470-85598-3 p. 553
[3] Peterson 1984, p. 45.
[4] Peterson 1984, p. 34.
[5] Peterson 1984, p. 44.
[6] Young, David A. Phase diagrams of the elements (http:/ /books.google. com/ books?id=F2HVYh6wLBcC&pg=PA228), University of
California Press, 1991, ISBN 0-520-07483-1 p. 228
[7] Peterson 1984, p. 55.
[8] Holleman 2007, p.
1956.
[9] Greenwood 1997, p. 1265.
[10] Abstract (http://www.osti.gov/cgi-bin/rd_accomplishments/ display_biblio.cgi?id=ACC0045&numPages=38&fp=N)
[11] Peterson 1984, p. 30.
[12] Peterson 1984, p. 32.
[13] Peterson 1984, pp. 33–34.
[14] S. G. Thompson, BB Cunningham: "First Macroscopic Observations of the Chemical Properties of Berkelium and californium," supplement
to Paper P/825 presented at the Second International Conference on Peaceful Uses of Atomic Energy, Geneva, 1958
[15] Peterson 1984, p. 38.
[16] Greenwood 1997, p. 1262.
[17] Peterson 1984, p. 41.
[18] Holleman 2007, p. 1972.
[19] Peterson 1984, p. 51.
[20] Holleman 2007, p. 1969.
[21] Peterson 1984, p. 47.
[22] Greenwood 1997, p. 1270.
[23] Peterson 1984, p. 48.
[24] Peterson 1984, p. 52.
[25] Peterson 1984, p. 53.
[26] Peterson 1984, pp. 39–40.
[27] Peterson 1984, p. 54.
[28] Christoph Elschenbroich Organometallic Chemistry, 6th Edition, Wiesbaden 2008, ISBN 978-3-8351-0167-8, pp. 583–584
[29] Finally, Element 117 Is Here! (http://news. sciencemag. org/sciencenow/ 2010/04/finally-element-117-is-here.html), Science Now, 7
April 2010
[30] Collaboration Expands the Periodic Table, One Element at a Time (https:// str.llnl.gov/OctNov10/shaughnessy. html), Science and
Technology Review, Lawrence Livermore National Laboratory, October/November 2010
[31] Nuclear Missing Link Created at Last: Superheavy Element 117 (http:// www.sciencedaily.com/releases/2010/04/100406181611. htm),
Science daily, 7 April 2010
[32] G. Pfennig, H. Klewe-Nebenius, W. Seelmann Eggebert (Eds.): Karlsruhe nuclide, 7 Edition, 2006
[33] Institut de Radioprotection et de Sûreté Nucléaire: "Evaluation of nuclear criticality safety. data and limits for actinides in transport" (http://
ec.europa.eu/ energy/ nuclear/ transport/ doc/irsn_sect03_146.pdf), p. 16
[34] Emeleus, H. J. Advances in inorganic chemistry (http:// books. google. com/books?id=K5_LSQqeZ_IC&pg=PA32), Academic Press,
1987, ISBN 0-12-023631-1 p. 32
[35] International Commission on Radiological Protection Limits for intakes of radionuclides by workers, Part 4, Volume 19, Issue 4 (http:/ /
books. google. com/books?id=WTxcCV4w0VEC&pg=PA14), Elsevier Health Sciences, ISBN, 0080368867 p. 14
[36] Pradyot Patnaik. Handbook of Inorganic Chemicals McGraw-Hill, 2002, ISBN 0-07-049439-8

Bibliography
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2 ed.). Oxford:
Butterworth-Heinemann. ISBN 0-08-037941-9.
• Holleman, Arnold F.; Wiberg, Nils (2007). Textbook of Inorganic Chemistry (102 ed.). Berlin: de Gruyter.
ISBN 978-3-11-017770-1.
• Peterson, J. R.; Hobart, D. E. (1984). "The Chemistry of Berkelium" (http://books.google.com/
books?id=U-YOlLVuV1YC&pg=PA29). In Emeléus, Harry Julius. Advances in inorganic chemistry
and
radiochemistry 28. Academic Press. pp. 29–64. doi: 10.1016/S0898-8838(08)60204-4 (http://dx.doi.org/10.
1016/S0898-8838(08)60204-4). ISBN 0-12-023628-1.

External links
• Berkelium (http://www.periodicvideos.com/videos/097.htm) at The Periodic Table of Videos (University
of Nottingham)
98 Californium 1

98 – Californium
Californium
Cf
98

berkelium ← californium → einsteiniumDy



Cf

(Upn)

Californium in the periodic table


Appearance

silvery

General properties

Name, symbol, number californium, Cf, 98

Pronunciation /ˌkælɨˈfɔrniəm/
KAL-i-FOR-nee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight [1]


(251)

Electron configuration 10 2 [2]


[Rn] 5f 7s
2, 8, 18, 32, 28, 8, 2

Physical properties

Phase solid

Density (near r.t.) [1] −3


15.1 g·cm
[1]
Melting point 1173 K, °C, 1652 °F
900

Boiling point (estimation) 1743 K, 1470 °C, 2678 °F


Atomic properties
[3]
Oxidation states 2, 3,
4

Electronegativity [4]
1.3 (Pauling scale)
[5] −1
Ionization energies 1st: kJ·mol
608

Miscellanea

Crystal structure simple hexagonal

Mohs hardness [6]


3–4

CAS registry number [1]


7440-71-3

History

Naming after California, where it was discovered

Discovery Lawrence Berkeley National Laboratory (1950)

Most stable isotopes

Main article: Isotopes of californium

iso NA half-life DM DE (MeV) DP

248 syn 333.5 d α (100%) 6.369 244


Cf Cm
−3 0.0029 –
SF (2.9×10 %)

249 trace 351 y α (100%) 6.295 245


Cf Cm
−7 −7 –
SF (5.0×10 %) 4.4×10

250 trace 13.08 y α (99.92%) 6.129 246


Cf Cm
SF (0.08%) 0.077 –

251 trace 898 y α 6.172 247


Cf Cm

252 trace 2.645 y α (96.91%) 6.217 248


Cf Cm
SF (3.09%) – –

253 trace 17.81 d − 0.29 253


Cf β (99.69%) Es
α (0.31%) 6.126 249
Cm

254 syn 60.5 d SF (99.69%) – –


Cf
α (0.31%) 5.930 250
Cm
[7]
Isotope
references:

• v
• t
[8]
• e

Californium is a radioactive metallic chemical element with the symbol Cf and atomic number 98. The element
was first made at the University of California, Berkeley, in 1950 by bombarding curium with alpha particles
(helium-4
ions). It is an actinide element, the sixth transuranium element to be synthesized, and has the second-highest atomic
mass of all the elements that have been produced in amounts large enough to see with the unaided eye (after
einsteinium). The element was named after California and the University of California. It is the heaviest element to
occur naturally on Earth; heavier elements can only be produced by synthesis.
Two crystalline forms exist for californium under normal pressure: one above and one below 900 °C. A third
form exists at high pressure. Californium slowly tarnishes in air at room temperature. Compounds of
californium are dominated by a chemical form of the element, designated californium(III), that can participate in
three chemical bonds. The most stable of californium's twenty known isotopes is californium-251, which
has a half-life of 898 years. This short half-life means the element is not found in significant quantities in
the Earth's crust. Californium-252, with a half-life of about 2.64 years, is the most common isotope used and is
produced at the Oak Ridge National Laboratory in the United States and the Research Institute of Atomic
Reactors in Russia.
Californium is one of the few transuranium elements that have practical applications. Most of these applications
exploit the property of certain isotopes of californium to emit neutrons. For example, californium can be used to help
start up nuclear reactors, and it is employed as a source of neutrons when studying materials with neutron diffraction
and neutron spectroscopy. Californium can also be used in nuclear synthesis of higher mass elements; ununoctium
(element 118) was synthesized by bombarding californium-249 atoms with calcium-48 ions. Use of californium must
take into account radiological concerns and the element's ability to disrupt the formation of red blood cells by
bioaccumulating in skeletal tissue.

Characteristics

Physical properties
[9]
Californium is a silvery white actinide metal with a melting point of 900 ± 30 °C and an estimated boiling point of
[10]
1745 K. The pure metal is malleable and is easily cut with a razor blade. Californium metal starts to vaporize
[11]
above 300 °C when exposed to a vacuum. Below 51 K (−220 °C) californium metal is either ferromagnetic or
ferrimagnetic (it acts like a magnet), between 48 and 66 K it is antiferromagnetic (an intermediate state), and above
[12]
160 K (−110 °C) it is paramagnetic (external magnetic fields can make it magnetic). It forms alloys with
[11]
lanthanide metals but little is known about them.
The element has two crystalline forms under 1 standard atmosphere of pressure: A double-hexagonal close-packed
[13]
form dubbed alpha (α) and a face-centered cubic form designated beta (β). </ref> The α form exists below
3 3 [14]
900 °C with a density of 15.10 g/cm and the β form exists above 900 °C with a density of 8.74 g/cm . At 48 GPa
of pressure the β form changes into an orthorhombic crystal system due to de-localization of the atom's 5f electrons,
[15][16]
which frees them to bond. </ref>
The bulk modulus of a material is a measure of its resistance to uniform pressure. Californium's bulk modulus is
50 ± 5 GPa, which is similar to trivalent lanthanide metals but smaller than more familiar metals, such as
[15]
aluminium (70 GPa).
Chemical properties and compounds
Further information: Compounds of californium

Representative californium compounds[9][17] Compounds in the +4 oxidation state are


strong oxidizing agents and those in the +2 state are strong reducing agents.[9] </ref>
state compound formula color

+2 californium(II) bromide CfBr yellow


2

+2 californium(II) iodide CfI dark violet


2

+3 californium(III) oxide Cf O yellow-green


2 3

+3 californium(III) fluoride CfF bright green


3

+3 californium(III) chloride CfCl emerald green


3

+3 californium(III) iodide CfI lemon yellow


3

+4 californium(IV) oxide CfO black brown


2

+4 californium(IV) fluoride CfF green


4

Californium exhibits valences of 4, 3, or 2; indicating the number of chemical bonds one atom of this element can
[14] [18]
form. Its chemical properties are predicted to be similar to other primarily 3+ valence actinide elements and
[19]
the element dysprosium, which is the lanthanide above californium in the periodic table. The element slowly
[14]
tarnishes in air at room temperature, with the rate increasing when moisture is added. Californium reacts when
heated with hydrogen, nitrogen, or a chalcogen (oxygen family element); reactions with dry hydrogen and aqueous
[14]
mineral acids are rapid.
Californium is only water soluble as the californium(III) cation. Attempts to reduce or oxidize the +3 ion in solution
[19]
have failed. The element forms a water-soluble chloride, nitrate, perchlorate, and sulfate and is precipitated as a
[18]
fluoride, oxalate, or hydroxide.

Isotopes
Main article: Isotopes of
californium
Twenty radioisotopes of californium have been characterized, the most stable being californium-251 with a half-life
of 898 years, californium-249 with a half-life of 351 years, californium-250 with a half-life of 13.08 years, and
californium-252 with a half-life of 2.645 years. All the remaining isotopes have half-lives shorter than a year, and the
majority of these have half-lives shorter than 20 minutes. The isotopes of californium range in mass number from
237 to 256.
Californium-249 is formed from the beta decay of berkelium-249, and most other californium isotopes are made by
[19]
subjecting berkelium to intense neutron radiation in a nuclear reactor. Although californium-251 has the longest
half-life, its production yield is only 10% due to its tendency to collect neutrons (high neutron capture) and its
[20]
tendency to interact with other particles (high neutron cross-section).
Californium-252 is a very strong neutron emitter, which makes it extremely radioactive and harmful.
Californium-252 undergoes alpha decay (the loss of two protons and two neutrons) 96.9% of the time to form
curium-248 while the remaining 3.1% of decays are spontaneous fission. One microgram (µg) of californium-
252 emits 2.3 million neutrons per second, an average of 3.7 neutrons per spontaneous fission. Most of the
other isotopes of californium decay to isotopes of curium (atomic number 96) via alpha decay.
History
Californium was first synthesized at the University of California,
Berkeley, by the physics researchers Stanley G. Thompson, Kenneth
Street, Jr., Albert Ghiorso, and Glenn T. Seaborg on or about February
[21]
9, 1950. It was the sixth transuranium element to be discovered; the
team announced its discovery on March 17, 1950.
To produce californium, a microgram-sized target of curium-242 (242
96Cm) was bombarded with 35 MeV-alpha particles (4
2He) in the 60-inch-diameter (1,500 mm) cyclotron at Berkeley,
California, which produced californium-245 (245
[21]
The 60-inch-diameter (1,500 mm) cyclotron used 98Cf) plus one free neutron (n).
to first synthesize californium
242
96Cm + 4
2He → 245
98Cf + 1
0n
[22]
Only about 5,000 atoms of californium were produced in this experiment, and these atoms had a half-life of
[21]
44 minutes.
The discoverers named the new element after California and the University of California. This was a break from
the convention used for elements 95 to 97, which drew inspiration from how the elements directly above them in
[23][24]
the periodic table were named. </ref> However, the element directly above element 98 in the periodic
table, dysprosium, has a name that simply means "hard to get at" so the researchers decided to set aside the
[25]
informal naming convention. They added that "the best we can do is to point out [that] ... searchers a century
[26]
ago found it difficult to get to California."
Weighable quantities of californium were first produced by the irradiation of plutonium targets at the Materials
Testing Reactor at the Idaho National Laboratory; and these findings were reported in 1954. The high
spontaneous fission rate of californium-252 was observed in these samples. The first experiment with
[21]
californium in concentrated form occurred in 1958. The isotopes californium-249 to californium-252 were
isolated that same year from a sample of plutonium-239 that had been irradiated with neutrons in a nuclear
[9]
reactor for five years. Two years later, in 1960, Burris Cunningham and James Wallman of the Lawrence
Radiation Laboratory of the University of California created the first californium compounds—californium
trichloride, californium oxychloride, and californium oxide—by treating californium with steam and hydrochloric
acid.
The High Flux Isotope Reactor (HFIR) at the Oak Ridge National Laboratory (ORNL) in Oak Ridge, Tennessee,
started producing small batches of californium in the 1960s. By 1995, the HFIR nominally produced 500
[27]
milligrams of californium annually. Plutonium supplied by the United Kingdom to the United States under the
1958 US-UK Mutual Defence Agreement was used for californium production.
The Atomic Energy Commission sold californium-252 to industrial and academic customers in the early 1970s
for
[28][29]
$10 per microgram and an average of 150 mg of californium-252 were shipped each year from 1970 to 1990.
Californium metal was first prepared in 1974 by Haire and Baybarz who reduced californium(III) oxide with
[30][31]
lanthanum metal to obtain microgram amounts of sub-micrometer thick films. The 1974 work was confirmed
[30]
in 1976 and work on californium metal continued. </ref>
Occurrence
Very minute amounts of californium have been found to exist on Earth
due to neutron capture reactions and beta decay in very highly-
concentrated uranium-bearing deposits. Traces of californium can be
found near facilities that use the element in mineral prospecting and in
[32]
medical treatments. The element is fairly insoluble in water, but it
adheres well to ordinary soil; and concentrations of it in the soil can
be 500 times higher than in the water surrounding the soil particles. Nuclear testing has contaminated the
environment with traces of californium.
Fallout from atmospheric nuclear testing prior to 1980 contributed a
small amount of californium to the environment. Californium isotopes
with mass numbers 249, 252, 253, and 254 have been observed in the radioactive dust collected from the air after a
nuclear explosion. Californium is not a major radionuclide at United States Department of Energy legacy sites since
it was not produced in large quantities.
254
Californium was once believed to be produced in supernovas, as their decay matches the 60 day half-life of Cf.
However, subsequent studies failed to demonstrate any californium spectra, and supernova light curves are now
[33]
thought to follow the decay of nickel-56.

Production
See also: Nuclear fuel cycle
[34]
Californium is produced in nuclear reactors and particle accelerators. Californium-250 is made by bombarding
berkelium-249 (249
97Bk) with neutrons, forming berkelium-250
(250

97Bk) via neutron capture (n,γ) which, in turn, quickly beta decays (β ) to californium-250 (250
[35]
98Cf) in the following reaction:
249
97Bk(n,γ)250
97Bk → 250

98Cf + β
Bombardment of californium-250 with neutrons produces californium-251 and californium-252.
[35]

Prolonged irradiation of americium, curium, and plutonium with neutrons produces milligram amounts of
[36]
californium-252 and microgram amounts of californium-249. As of 2006, curium isotopes 244 to 248 are
irradiated by neutrons in special reactors to produce primarily californium-252 with lesser amounts of isotopes
[37]
249 to 255.
Microgram quantities of californium-252 are available for commercial use through the U.S. Nuclear Regulatory
[34]
Commission. Only two sites produce californium-252 – the Oak Ridge National Laboratory in the United States,
and the Research Institute of Atomic Reactors in Dimitrovgrad, Russia. As of 2003, the two sites produce 0.25 grams
[38]
and 0.025 grams of californium-252 per year, respectively.
Three californium isotopes with significant half-lives are produced, requiring a total of 15 neutron captures by
[38]
uranium-238 without nuclear fission or alpha decay occurring during the process. Californium-253 is at the
end of a production chain that starts with uranium-238, includes several isotopes of plutonium, americium,
curium, berkelium, and the californium isotopes 249 to 253 (see diagram).
Scheme of the production of californium-252 from uranium-238 by neutron irradiation

Applications
Californium-252 has a number of specialized applications as a strong
neutron emitter, and each microgram of fresh californium produces
139 million neutrons per minute. This property makes californium
[14]
useful as a neutron startup source for some nuclear reactors and as
a
portable (non-reactor based) neutron source for neutron activation
[40]
analysis to detect trace amounts of elements in samples. </ref>
Neutrons from californium are employed as a treatment of certain
cervical and brain cancers where other radiation therapy is
[14]
ineffective. It has been used in educational applications since 1969
when the Georgia Institute of Technology received a loan of 119 µg of
[41]
californium-252 from the Savannah River Plant. It is also used with
online elemental coal analyzers and bulk material analyzers in the coal
and cement industries.

Neutron penetration into materials makes californium useful in


[14]
detection instruments such as fuel rod scanners; neutron Fifty-ton shipping cask built at Oak Ridge
radiography of aircraft and weapons components to detect corrosion, National Laboratory
[39] which can transport up to 1
[42] gram of Large and heavily shielded
bad welds, cracks and trapped moisture; and in portable metal 252
Cf.
transport containers are needed to prevent the
detectors. Neutron moisture gauges use californium-252 to find water
release of highly radioactive material in case of
and petroleum layers in oil wells, as a portable neutron source for gold normal and hypothetical accidents.
[19]
and silver prospecting for on-the-spot analysis, and to detect ground
water movement. The major uses of californium-252 in 1982 were, in order of use, reactor start-up (48.3%), fuel
[43]
rod scanning (25.3%), and activation analysis (19.4%). By 1994 most californium-252 was used in
neutron radiography (77.4%), with fuel rod scanning (12.1%) and reactor start-up (6.9%) as important but
[43]
distant secondary uses.
Californium-251 has a very small calculated critical mass (about 5 kg), high lethality, and a relatively short period of
toxic environmental irradiation. The low critical mass of californium led to some exaggerated claims about possible
[44]
uses for the element. </ref>
In October 2006, researchers announced that three atoms of ununoctium (element 118) had been identified at the
Joint Institute for Nuclear Research in Dubna, Russia, as the product of bombardment of californium-249 with
calcium-48, making it the heaviest element ever synthesized. The target for this experiment contained about 10 mg
2
of californium-249 deposited on a titanium foil of 32 cm area. Californium has also been used to produce other
transuranium elements; for example, element 103 (later named lawrencium) was first synthesized in 1961 by
bombarding californium with boron nuclei.

Precautions
Californium that bioaccumulates in skeletal tissue releases radiation that disrupts the body's ability to form red blood
[45]
cells. The element plays no natural biological role in any organism due to its intense radioactivity and low
[32]
concentration in the environment.
Californium can enter the body from ingesting contaminated food or drinks or by breathing air with suspended
particles of the element. Once in the body, only 0.05% of the californium will reach the bloodstream. About 65% of
that californium will be deposited in the skeleton, 25% in the liver, and the rest in other organs, or excreted, mainly
in urine. Half of the californium deposited in the skeleton and liver are gone in 50 and 20 years, respectively.
Californium in the skeleton adheres to bone surfaces before slowly migrating throughout the bone.
The element is most dangerous if taken into the body. In addition, californium-249 and californium-251 can
cause tissue damage externally, through gamma ray emission. Ionizing radiation emitted by californium on bone
and in the liver can cause cancer.

Notes
[1] CRC 2006, p.
4.56.
[2] CRC 2006, p.
1.14.
[3] Greenwood 1997, p. 1265.
[4] Emsley 1998, p. 50.
[5] CRC 2006, p. 10.204.
[6] CRC 1991, p.
254.
[7] CRC 2006, p. 11.196.
[8] http://en.wikipedia.org/ w/index. php?title=Template:Infobox_californium&action=edit
[9] Jakubke 1994, p. 166.
[10] Haire 2006, pp. 1522–1523.
[11] Haire 2006, p. 1526.
[12] Haire 2006, p. 1525.
[13] A double hexagonal close-packed (dhcp) unit cell consists of two hexagonal close-packed structures that share a common hexagonal
plane, giving dhcp an ABACABAC sequence.<ref name="FOOTNOTESzwacki201080">Szwacki 2010, p. 80.
[14] O'Neil 2006, p.
276.
[15] Haire 2006, p. 1522.
[16] The three lower-mass transplutonium elements—americium, curium, and berkelium—require much less pressure to delocalize their 5f
electrons.<ref name="FOOTNOTEHaire20061522">Haire 2006, p. 1522.
[17] Other +3 oxidation states include the sulfide and metallocene.<ref name="FOOTNOTECotton19991163">Cotton 1999, p. 1163.
[18] Seaborg 2004.
[19] CRC 2006, p. 4.8.
[20] Haire 2006, p. 1504.
[21] Cunningham 1968, p. 103.
[22] Seaborg 1996, p. 82.
[23] Weeks 1968, p. 849.
[24] Europium, in the sixth period directly above element 95, was named for the continent it was discovered on, so element 95 was named
americium. Element 96 was named for Marie Curie and Pierre Curie as an analog to the naming of gadolinium, which was named for the
scientist and engineer Johan Gadolin. Terbium was named for the city it was discovered in, so element 97 was named berkelium.<ref
name="FOOTNOTEWeeks1968848">Weeks 1968, p. 848.
[25] Heiserman 1992, p. 347.
[26] Weeks 1968, p. 848.
[27] Osborne-Lee 1995, p. 11.
[28] Osborne-Lee 1995, p. 6.
[29] The Nuclear Regulatory Commission replaced the Atomic Energy Commission when the Energy Reorganization Act of 1974 was
implemented. The price of californium-252 was increased by the NRC several times and was $60 per microgram by 1999; this price does not
include the cost of encapsulation and transportation.CRC 2006, p. 4.56.
[30] Haire 2006, p. 1519.
[31] In 1975, another paper stated that the californium metal prepared the year before was the hexagonal compound Cf O S and face-centered
2 2
cubic compound CfS.<ref>
[32] Emsley 2001, p. 90.
[33] Ruiz-Lapuente1996, p. 274.
[34] Krebs 2006, pp. 327–328.
[35] Heiserman 1992, p. 348.
[36] Cunningham 1968, p. 105.
[37] Haire 2006, p. 1503.
[38] NRC 2008, p. 33.
[39] Seaborg 1994, p. 245.
[40] By 1990, californium-252 had replaced plutonium-beryllium neutron sources due to its smaller size and lower heat and gas generation.<ref
name="FOOTNOTESeaborg1990318">Seaborg 1990, p. 318.
[41] Osborne-Lee 1995, p. 33.
[42] Osborne-Lee 1995, pp. 26–27.
[43] Osborne-Lee 1995, p. 12.
[44] An article entitled "Facts and Fallacies of World War III" in the July 1961, edition of Popular Science magazine read "A californium
atomic bomb need be no bigger than a pistol bullet. You could build a hand-held six-shooter to fire bullets that would explode on contact
with the force of 10 tons of TNT."<ref>"force of 10 tons of TNT" on page 180.
[45] Cunningham 1968, p. 106.

References

Bibliography
• Cotton, F. Albert; Wilkinson, Geoffrey; Murillo, Carlos A.; Bochmann, Manfred (1999). Advanced Inorganic
Chemistry (6th ed.). John Wiley & Sons. ISBN 978-0-471-19957-1.
• CRC contributors (1991). Walker, Perrin; Tarn, William H., eds. Handbook of Metal Etchants. CRC Press.
ISBN 978-0-8493-3623-2.
• CRC contributors (2006). Lide, David R., ed. Handbook of Chemistry and Physics (87th ed.). CRC Press, Taylor
& Francis Group. ISBN 978-0-8493-0487-3.
• Cunningham, B. B. (1968). "Californium". In Hampel, Clifford A. The Encyclopedia of the Chemical Elements.
Reinhold Book Corporation. LCCN 68-29938 (http://lccn.loc.gov/68-29938) Check |lccn= value (help).
• Emsley, John (1998). The Elements (http://books.google.com/books?id=qgYpAAAAYAAJ). Oxford
University Press. ISBN 978-0-19-855818-7.
• Emsley, John (2001). "Californium" (http://books.google.com/books?id=Yhi5X7OwuGkC&
printsec=frontcover#v=onepage&q&f=false). Nature's Building Blocks: An A-Z Guide to the Elements.
Oxford University Press. ISBN 978-0-19-850340-8.
• Greenwood, N. N.; Earnshaw, A. (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
ISBN 978-0-7506-3365-9.
• Haire, Richard G. (2006). "Californium". In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean. The Chemistry
of the Actinide and Transactinide Elements (3rd ed.). Springer Science+Business Media.
ISBN 978-1-4020-3555-5.
• Heiserman, David L. (1992). "Element 98: Californium" (http://books.google.com/books?id=24l-Cpal9oIC).
Exploring Chemical Elements and their Compounds. TAB Books. ISBN 978-0-8306-3018-9.
• Jakubke, Hans-Dieter; Jeschkeit, Hans, eds. (1994). Concise Encyclopedia Chemistry (http://books.google.com/
books?id=Owuv-c9L_IMC&printsec=frontcover). trans. rev. Eagleson, Mary. Walter de Gruyter.
ISBN 978-3-11-011451-5.
• Krebs, Robert (2006). The History and Use of our Earth's Chemical Elements: A Reference Guide. Greenwood
Publishing Group. ISBN 978-0-313-33438-2.
• National Research Council (U.S.). Committee on Radiation Source Use and Replacement (2008). Radiation
Source Use and Replacement: Abbreviated Version (http://books. google. com/books?id=3cT2REdXJ98C&
printsec=frontcover). National Academies Press. ISBN 978-0-309-11014-3.
• O'Neil, Marydale J.; Heckelman, Patricia E.; Roman, Cherie B., eds. (2006). The Merck Index: An
Encyclopedia of Chemicals, Drugs, and Biologicals (http://books.google.com/books?id=kEYfRAAACAAJ)
(14th ed.). Merck Research Laboratories, Merck & Co. ISBN 978-0-911910-00-1.
• Osborne-Lee, I. W.; Alexander, C. W. (1995). "Californium-252: A Remarkable Versatile Radioisotope"
(http:// www.osti.gov/bridge/product.biblio.jsp?query_id=1&page=0&osti_id=205871). Oak Ridge Technical
Report ORNL/TM-12706. doi: 10.2172/205871 (http://dx.doi.org/10. 2172/205871).
• Ruiz-Lapuente, P.; Canal, R.; Isern, J. (1996). Thermonuclear Supernovae (http://books.google.com/
books?id=I6Rl1VAAX3QC). Springer Science+Business Media. ISBN 978-0-7923-4359-2.
• Seaborg, Glenn T.; Loveland, Walter D. (1990). The Elements Beyond Uranium
(http://books.google.com/ books?id=QFhRAAAAMAAJ). John Wiley & Sons, Inc. ISBN 978-0-471-
89062-1.
• Seaborg, G. T. (1994). Modern alchemy: selected papers of Glenn T. Seaborg (http://books.google.com/
books?id=e53sNAOXrdMC&pg=PA245). World Scientific. ISBN 978-981-02-1440-1.
• Seaborg, G. T. (1996). Adloff, J. P., ed. One Hundred Years after the Discovery of Radioactivity
(http://books. google. com/books?id=whGiCQywLi8C&printsec=frontcover#v=onepage&q&f=false).
Oldenbourg Wissenschaftsverlag. ISBN 978-3-486-64252-0.
• Seaborg, Glenn T. (2004). "Californium" (http://books.google.com/books?id=Owuv-c9L_IMC&
printsec=frontcover#v=onepage&q&f=false). In Geller, Elizabeth. Concise Encyclopedia of Chemistry.
McGraw-Hill. p. 94. ISBN 978-0-07-143953-4.
• Szwacki, Nevill Gonzalez; Szwacka, Teresa (2010). Basic Elements of Crystallography (http://www.scribd.
com/doc/39588027/ Basic-Elements-of-Crystallography). Pan Stanford. ISBN 978-981-4241-59-5.
• Weeks, Mary Elvira; Leichester, Henry M. (1968). "21: Modern Alchemy". Discovery of the Elements.
Journal of Chemical Education. pp. 848–850. ISBN 978-0-7661-3872-8. LCCN 68-15217
(http://lccn.loc.gov/68-15217) Check |lccn= value (help).

External links
• Californium (http://www.periodicvideos.com/videos/098.htm) at The Periodic Table of Videos (University
of Nottingham)
• NuclearWeaponArchive.org – Californium (http://www.nuclearweaponarchive.org/Nwfaq/Nfaq6.
html#nfaq6.2)
• Hazardous Substances Databank – Californium, Radioactive (http://toxnet.nlm.nih.gov/cgi-bin/sis/search/
r?dbs+hsdb:@term+@rel+@na+californium,radioactive)
99 Einsteinium 1

99 – Einsteinium
"Athenium" redirects here. For other uses, see Athenaeum.

Einsteinium
Es
99

californium ← einsteinium → fermiumHo



Es

(Upu)

Einsteinium in the periodic table


Appearance

silver-colored

General properties

Name, symbol, number einsteinium, Es, 99

Pronunciation /aɪnˈstaɪniəm/
eyen-STY-nee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (252)

Electron configuration 11 2
[Rn] 5f 7s
2, 8, 18, 32, 29, 8, 2

Physical properties

Phase solid

Density (near r.t.) −3


8.84 g·cm
Melting point 1133 K, 860 °C, 1580 °F

Boiling point (estimated) 1269 K, 996 °C, 1825 °F

Atomic properties

Oxidation states 2, 3, 4

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 619 kJ·mol

Miscellanea

Crystal structure face-centered cubic

Magnetic ordering paramagnetic

CAS registry number 7429-92-7

History

Naming after Albert Einstein

Discovery Lawrence Berkeley National Laboratory (1952)

Most stable isotopes

Main article: Isotopes of einsteinium

iso NA half-life DM DE (MeV) DP

252 syn 471.7 d α 6.760 248


Es Bk
ε 1.260 252
Cf

− 0.480 252
β Fm

253 syn 20.47 d SF - -


Es
α 6.739 249
Bk

254 syn 275.7 d ε 0.654 254


Es Cf

− 1.090 254
β Fm
α 6.628 250
Bk

255 syn 39.8 d − 0.288 255


Es β Fm
α 6.436 251
Bk
SF - -

• v
• t
[1]
• e

Einsteinium is a synthetic element with the symbol Es and atomic number 99. It is the seventh transuranic element,
and an actinide.
Einsteinium was discovered as a component of the debris of the first hydrogen bomb explosion in 1952, and named
after Albert Einstein. Its most common isotope einsteinium-253 (half life 20.47 days) is produced artificially from
decay of californium-253 in a few dedicated high-power nuclear reactors with a total yield on the order of one
milligram per year. The reactor synthesis is followed by a complex procedure of separating einsteinium-253 from
other actinides and products of their decay. Other isotopes are synthesized in various laboratories, but at much
smaller amounts, by bombarding heavy actinide elements with light ions. Owing to the small amounts of produced
einsteinium and the short half-life of its most easily produced isotope, there are currently almost no practical
applications for it outside of basic scientific research. In particular, einsteinium was used to synthesize, for the first
time, 17 atoms of the new element mendelevium in 1955.
Einsteinium is a soft, silvery, paramagnetic metal. Its chemistry is typical of the late actinides, with a preponderance
of the +3 oxidation state; the +2 oxidation state is also accessible, especially in solids. The high radioactivity of
einsteinium-253 produces a visible glow and rapidly damages its crystalline metal lattice, with released heat of about
1000 watts per gram. Difficulty in studying its properties is due to einsteinium-253's conversion to berkelium and
then californium at a rate of about 3% per day. The isotope of einsteinium with the longest half life, einsteinium-252
(half life 471.7 days) would be more suitable for investigation of physical properties, but it has proven far more
[2]
difficult to produce and is available only in minute quantities, and not in bulk. Einsteinium is the element with the
highest atomic number which has been observed in macroscopic quantities in its pure form, and this was the
common short-lived isotope einsteinium-253.
Like all synthetic transuranic elements, isotopes of einsteinium are extremely radioactive and are considered highly
dangerous to health on ingestion.

History
Einsteinium was first identified in December 1952 by Albert Ghiorso
and co-workers at the University of California, Berkeley in
collaboration with the Argonne and Los Alamos National
Laboratories, in the fallout from the Ivy Mike nuclear test. The test
was carried out on November 1, 1952 at Enewetak Atoll in the
Pacific Ocean and was the first successful test of a hydrogen
bomb. Initial examination of the debris from the explosion had
shown the production of a new isotope of plutonium, 244
94Pu, which could only have formed by the absorption of six neutrons

Einsteinium was first observed in the fallout by a uranium-238 nucleus followed by two beta decays.
from the Ivy Mike nuclear test.

At the time, the multiple neutron absorption was thought to be an extremely rare process, but the identification of 244
94Pu indicated that still more neutrons could have been captured by the uranium nuclei, thereby producing new
elements heavier than californium.
Ghiorso and co-workers analyzed filter papers which had been flown through the explosion cloud on airplanes (the
same sampling technique that had been used to discover 244
[3]
94Pu). Larger amounts of radioactive material were later isolated from coral debris of the atoll, which were
delivered to the U.S. The separation of suspected new elements was carried out in the presence
of a citric acid/ammonium buffer solution in a weakly acidic medium
(pH ≈ 3.5), using ion exchange at elevated temperatures; fewer than
[4]
200 atoms of einsteinium were recovered in the end. Nevertheless,
253
element 99 (einsteinium), namely its Es isotope, could be detected
via its characteristic high-energy alpha decay at 6.6 MeV. It was
produced by the capture of 15 neutrons by uranium-238 nuclei
followed by seven beta-decays, and had a half-life of 20.5 days. Such
multiple neutron absorption was made possible by the high neutron
flux density during the detonation, so that newly generated heavy
isotopes had plenty of available neutrons to absorb before they could
disintegrate into lighter elements. Neutron capture initially raised the
mass number without changing the atomic number of the nuclide, and
the concomitant beta-decays resulted in a gradual increase in the
atomic number:
The element was named after Albert Einstein.

238 255
Some U atoms, however, could absorb another two neutrons (for a total of 17), resulting in Es, as well as in the
255 [5]
Fm isotope of another new element, fermium. The discovery of the new elements and the associated new
data on multiple neutron capture were initially kept secret on the orders of the U.S. military until 1955 due to
[6][7]
Cold War tensions and competition with Soviet Union in nuclear technologies. However, the rapid capture
of so many neutrons would provide needed direct experimental confirmation of the so-called r-process
multiple neutron absorption needed to explain the cosmic nucleosynthesis (production) of certain heavy
chemical elements (heavier than nickel) in supernova explosions, before beta decay. Such a process is needed
[8]
to explain the existence of many stable elements in the universe.
Meanwhile, isotopes of element 99 (as well as of new element 100, fermium) were produced in the Berkeley and
Argonne laboratories, in a nuclear reaction between nitrogen-14 and uranium-238, and later by intense neutron
irradiation of plutonium or californium:

These results were published in several articles in 1954 with the disclaimer that these were not the first studies that
had been carried out on the elements. The Berkeley team also reported some results on the chemical properties of
[9]
einsteinium and fermium. The Ivy Mike results were declassified and published in 1955.
In their discovery of the elements 99 and 100, the American teams had competed with a group at the Nobel Institute
for Physics, Stockholm, Sweden. In late 1953 – early 1954, the Swedish group succeeded in the synthesis of light
250
isotopes of element 100, in particular Fm, by bombarding uranium with oxygen nuclei. These results were also
published in 1954. Nevertheless, the priority of the Berkeley team was generally recognized, as its publications
preceded the Swedish article, and they were based on the previously undisclosed results of the 1952 thermonuclear
explosion; thus the Berkeley team was given the privilege to name the new elements. As the effort which had led to
the design of Ivy Mike was codenamed Project PANDA, element 99 had been jokingly nicknamed "Pandamonium"
but the official names suggested by the Berkeley group derived from two prominent scientists, Albert Einstein and
Enrico Fermi: "We suggest for the name for the element with the atomic number 99, einsteinium (symbol E) after
Albert Einstein and for the name for the element with atomic number 100, fermium (symbol Fm), after Enrico
[10]
Fermi." Both Einstein and Fermi died before the names were announced. The discovery of these new elements
was announced by Albert Ghiorso at the first Geneva Atomic Conference held on 8–20 August 1955. The symbol for
[11][12]
einsteinium was first given as "E" and later changed to "Es" by IUPAC.

Characteristics

Physical
Einsteinium is a synthetic, silvery-white, radioactive metal. In the periodic
table, it is located to the right of the actinide californium, to the left of the
actinide fermium and below the lanthanide holmium with which it shares
many similarities in physical and chemical properties. Its density of 8.84
3 3
g/cm is lower than that of californium (15.1 g/cm ) and is nearly the same as
that of holmium
3
(8.79 g/cm ), despite atomic einsteinium being much heavier than holmium.
The melting point of einsteinium (860 °C) is also relatively low – below
[6][14]
californium (900 °C), fermium (1527 °C) and holmium (1461 °C).
Einsteinium is a soft metal, with the bulk modulus of only 15 GPa, which
[15]
value is one of the lowest among non-alkali metals.

Contrary to the lighter actinides californium, berkelium, curium and americium


which crystallize in a double hexagonal structure at ambient conditions,
einsteinium is believed to have a face-centered cubic (fcc) symmetry with the Glow due to the intense radiation
253 [13]
space group Fm3m and the lattice constant a = 575 pm. However, there is a from ~300 µg of Es.
report of room-temperature hexagonal einsteinium metal with a = 398 pm and c
= 650 pm, which converted to the fcc phase upon heating to 300 °C.

The self-damage induced by the radioactivity of einsteinium is so strong that it rapidly destroys the crystal lattice,
253 []
and the energy release during this process, 1000 watts per gram of Es, induces a visible glow. These processes
[16]
may contribute to the relatively low density and melting point of einsteinium. Further, owing to the small size of
the available samples, the melting point of einsteinium was often deduced by observing the sample being heated
[17]
inside an electron microscope. Thus the surface effects in small samples could reduce the melting point value.
The metal is divalent and has a noticeably high volatility. In order to reduce the self-radiation damage, most
[18]
measurements of solid einsteinium and its compounds are performed right after thermal annealing. Also, some
compounds are studied under the atmosphere of the reductant gas, for example H O+HCl for EsOCl so that the
2
sample is partly regrown during its decomposition.
Apart from the self-destruction of solid einsteinium and its compounds, other intrinsic difficulties in studying this
253
element include scarcity – the most common Es isotope is available only once or twice a year in sub-milligram
amounts – and self-contamination due to rapid conversion of einsteinium to berkelium and then to californium at a
[19]
rate of about 3.3% per day:

Thus, most einsteinium samples are contaminated, and their intrinsic properties are often deduced by extrapolating
back experimental data accumulated over time. Other experimental techniques to circumvent the contamination
problem include selective optical excitation of einsteinium ions by a tunable laser, such as in studying its
[20]
luminescence properties.
Magnetic properties have been studied for einsteinium metal, its oxide and fluoride. All three materials showed
Curie–Weiss paramagnetic behavior from liquid helium to room temperature. The effective magnetic moments were
deduced as 10.4 ± 0.3 µ for Es O and 11.4 ± 0.3 µ for the EsF , which are the highest values among actinides, and
B 2 3 B 3
the corresponding Curie temperatures are 53 and 37 K.
Chemical
Like all actinides, einsteinium is rather reactive. Its trivalent oxidation state is most stable in solids and aqueous
[21]
solution where it induced pale pink color. The existence of divalent einsteinium is firmly established,
especially in solid phase; such +2 state is not observed in many other actinides, including protactinium,
uranium, neptunium, plutonium, curium and berkelium. Einsteinium(II) compounds can be obtained, for
[22]
example, by reducing einsteinium(III) with samarium(II) chloride. The oxidation state +4 was postulated from
[23]
vapor studies and is yet uncertain.

Isotopes
Main article: Isotopes of
einsteinium
Nineteen nuclides and three nuclear isomers are known for einsteinium with atomic weights ranging from 240 to
252
258. All are radioactive and the most stable nuclide, Es, has a half-life of 471.7 days. Next most stable isotopes
254 255 253
are Es (half-life 275.7 days), Es (39.8 days) and Es (20.47 days). All of the remaining isotopes have
half-lives shorter than 40 hours, and most of them decay within less than 30 minutes. Of the three nuclear isomers,
254m
the most stable is Es with half-life of 39.3 hours.

Nuclear
fission
Einsteinium has a high rate of nuclear fission that results in a low critical mass for a sustained nuclear chain reaction.
254
This mass is 9.89 kilograms for a bare sphere of Es isotope, and can be lowered to 2.9 or even 2.26 kilograms,
respectively, by adding a 30 centimeter thick steel or water reflector. However, even this small critical mass greatly
254 [24]
exceeds the total amount of einsteinium isolated thus far, especially of the rare Es isotope.

Natural occurrence
Because of the short half-life of all isotopes of einsteinium, all primordial einsteinium, that is einsteinium that could
possibly be present on the Earth during its formation, has decayed by now. Synthesis of einsteinium from naturally
occurring actinides uranium and thorium in the Earth crust requires multiple neutron capture, which is an extremely
unlikely event. Therefore, most einsteinium is produced on Earth in scientific laboratories, high-power nuclear
reactors, or in nuclear weapons tests, and is present only within a few years from the time of the synthesis.
Einsteinium and fermium did occur naturally in the natural nuclear fission reactor at Oklo, but no longer do so.
[25]
Einsteinium was observed in a very specular Przybylski's Star 2008
Synthesis and extraction
Einsteinium is produced in minute quantities
by bombarding lighter actinides with
neutrons in dedicated high-flux nuclear
reactors. The world's major irradiation
sources are the 85-megawatt High Flux
Isotope Reactor (HFIR) at the Oak Ridge
National Laboratory in Tennessee, U.S., and
the SM-2 loop reactor at the Research
Institute of Atomic Reactors (NIIAR) in
Dimitrovgrad, Russia, which are both
dedicated to the production of transcurium
(Z > 96) elements. These facilities have
similar power and flux levels, and are
expected to have comparable production Early evolution of einsteinium production in theU.S.
[26]
[27]
capacities for transcurium elements,
although the quantities produced at NIIAR are not widely reported. In a "typical processing campaign" at Oak Ridge,
tens of grams of curium are irradiated to produce decigram quantities of californium, milligram quantities of
249 [28]
berkelium ( Bk) and einsteinium and picogram quantities of fermium.
253
The first microscopic sample of Es sample weighing about 10 nanograms was prepared in 1961 at HFIR. A
[29]
special magnetic balance was designed to estimate its weight. Larger batches were produced later starting from
253
several kilograms of plutonium with the einsteinium yields (mostly Es) of 0.48 milligrams in 1967–1970, 3.2
[30]
milligrams in 1971–1973, followed by steady production of about 3 milligrams per year between 1974 and 1978.
These quantities however refer to the integral amount in the target right after irradiation. Subsequent separation
procedures reduced the amount of isotopically pure einsteinium roughly tenfold.

Laboratory
synthesis
253
Heavy neutron irradiation of plutonium results in four major isotopes of einsteinium: Es (α-emitter with half-
5 254m
life of 20.03 days and with a spontaneous fission half-life of 7×10 years); Es (β-emitter with half-life of
254 255
38.5 hours), Es (α-emitter with half-life of about 276 days) and Es (β-emitter with half-life of 24
days). An alternative route involves bombardment of uranium-238 with high-intensity nitrogen or oxygen ion
beams.
Einsteinium-247 (half-life 4.55 minutes) was produced by irradiating americium-241 with carbon or uranium-238
[31]
with nitrogen ions. The latter reaction was first realized in 1967 in Dubna, Russia, and the involved scientists
[32]
were awarded the Lenin Komsomol Prize.
248 249
The isotope Es was produced by irradiating Cf with deuterium ions. It mainly decays by emission of electrons
248
to Cf with a half-life of 25 (±5) minutes, but also releases α-particles of 6.87 MeV energy, with the ratio of
electrons to α-particles of about 400.

249 250 251 252 249


The heavier isotopes Es, Es, Es and Es were obtained by bombarding Bk with α-particles. One to four
neutrons are liberated in this process making possible the formation of four different isotopes in one reaction.

252
Einsteinium-253 was produced by irradiating a 0.1–0.2 milligram Cf target with a thermal neutron flux of (2–
14 −2 −1
5)×10 neutrons·cm ·s for 500–900 hours:
Synthesis in nuclear explosions
The analysis of the debris at the 10-megaton
Ivy Mike nuclear test was a part of long-term
project. One of the goals of which was
studying the efficiency of production of
transuranium elements in high-power
nuclear explosions. The motivation for these
experiments was that synthesis of such
elements from uranium requires multiple
neutron capture. The probability of such
events increases with the neutron flux, and
nuclear explosions are the most powerful
man-made neutron sources, providing
23
densities of the order 10 neutrons/cm²
29
within a microsecond, or about 10
neutrons/(cm²·s). In comparison, the flux of
15
Estimated yield of transuranium elements in the U.S. nuclear tests Hutch and the HFIR reactor is 5×10 neutrons/(cm²·s).
Cyclamen. A dedicated laboratory was set up right at
Enewetak Atoll for preliminary analysis of
debris, as some isotopes could have decayed by the time the debris samples reached the mainland U.S. The
laboratory was receiving samples for analysis as soon as possible, from airplanes equipped with paper filters which
flew over the atoll after the tests. Whereas it was hoped to discover new chemical elements heavier than fermium,
none of these were found even after a series of megaton explosions conducted between 1954 and 1956 at the atoll.

The atmospheric results were supplemented by the underground test data accumulated in the 1960s at the Nevada
Test Site, as it was hoped that powerful explosions conducted in confined space might result in improved yields and
heavier isotopes. Apart from traditional uranium charges, combinations of uranium with americium and thorium
have been tried, as well as a mixed plutonium-neptunium charge, but they were less successful in terms of yield and
was attributed to stronger losses of heavy isotopes due to enhanced fission rates in heavy-element charges. Product
isolation was problematic as the explosions were spreading debris through melting and vaporizing the surrounding
rocks at depths of 300–600 meters. Drilling to such depths to extract the products was both slow and inefficient in
[]
terms of collected volumes.
[33][34]
Among the nine underground tests that were carried between 1962 and 1969, the last one was the most
powerful and had the highest yield of transuranium elements. Milligrams of einsteinium that would normally take a
year of irradiation in a high-power reactor, were produced within a microsecond. However, the major practical
problem of the entire proposal was collecting the radioactive debris dispersed by the powerful blast. Aircraft filters
−14
adsorbed only about 4×10 of the total amount, and collection of tons of corals at Enewetak Atoll increased this
fraction by only two orders of magnitude. Extraction of about 500 kilograms of underground rocks 60 days after the
−7
Hutch explosion recovered only about 1×10 of the total charge. The amount of transuranium elements in this
500-kg batch was only 30 times higher than in a 0.4 kg rock picked up 7 days after the test which demonstrated the
[35]
highly non-linear dependence of the transuranium elements yield on the amount of retrieved radioactive rock.
Shafts were drilled at the site before the test in order to accelerate sample collection after explosion, so that
explosion would expel radioactive material from the epicenter through the shafts and to collecting volumes near the
surface. This method was tried in two tests and instantly provided hundreds kilograms of material, but with actinide
concentration 3 times lower than in samples obtained after drilling. Whereas such method could have been efficient
in scientific studies of short-lived isotopes, it could not improve the overall collection efficiency of the produced
[36]
actinides.
Although no new elements (apart from einsteinium and fermium) could be detected in the nuclear test debris, and the
total yields of transuranium elements were disappointingly low, these tests did provide significantly higher amounts
[37]
of rare heavy isotopes than previously available in laboratories.

Separation
Separation procedure of einsteinium depends on the synthesis method.
In the case of light-ion bombardment inside a cyclotron, the heavy ion
target is attached to a thin foil, and the generated einsteinium is simply
washed off the foil after the irradiation. However, the produced
[38]
amounts in such experiments are relatively low. The yields are
much higher for reactor irradiation, but there, the product is a mixture
of various actinide isotopes, as well as lanthanides produced in the
nuclear fission decays. In this case, isolation of einsteinium is a tedious
procedure which involves several repeating steps of cation exchange,
at elevated temperature and pressure, and chromatography. Separation
from berkelium is important, because the most common einsteinium
253
isotope produced in nuclear reactors, Es, decays with a half-life of
249
only 20 days to Bk, which is fast on the timescale of most
experiments. Such separation relies on the fact that berkelium easily
oxidizes to the solid +4 state and precipitates, whereas other actinides, Elution curves: chromatographic separation of
[39] Fm(100), Es(99), Cf, Bk, Cm and Am
including einsteinium, remain in their +3 state in solutions.

Separation of trivalent actinides from lanthanide fission products can be done by a cation-exchange resin
column using a 90% water/10% ethanol solution saturated with hydrochloric acid (HCl) as eluant. It is usually
followed by anion-exchange chromatography using 6 molar HCl as eluant. A cation-exchange resin column
(Dowex-50 exchange column) treated with ammonium salts is then used to separate fractions containing
elements 99, 100 and 101. These elements can be then identified simply based on their elution position/time,
using α-hydroxyisobutyrate solution (α-HIB), for example, as eluant.

Separation of the 3+ actinides can also be achieved by solvent extraction chromatography, using bis-(2-ethylhexyl)
phosphoric acid (abbreviated as HDEHP) as the stationary organic phase, and nitric acid as the mobile aqueous
phase. The actinide elution sequence is reversed from that of the cation-exchange resin column. The einsteinium
separated by this method has the advantage to be free of organic complexing agent, as compared to the separation
using a resin column.

Preparation of the metal


Einsteinium is highly reactive and therefore strong reducing agents are required to obtain the pure metal from its
[40]
compounds. This can be achieved by reduction of einsteinium(III) fluoride with metallic lithium:
EsF + 3 Li → Es + 3 LiF
3
However, owing to its low melting point and high rate of self-radiation damage, einsteinium has high vapor
pressure, which is higher than that of lithium fluoride. This makes this reduction reaction rather inefficient. It was
tried in the early preparation attempts and quickly abandoned in favor of reduction of einsteinium(III) oxide with
lanthanum
[41]
metal:
Es O + 2 La → 2 Es + La O
2 3 2 3
Chemical compounds

Crystal structure and lattice constants of some Es compounds


Compound Color Symmetry Space group No Pearson symbol a (pm) b (pm) c (pm)

Es O Colorless Cubic Ia3 206 cI80 1076.6


2 3

Es O Colorless Monoclinic C2/m 12 mS30 1411 359 880


2 3

Es O Colorless Hexagonal P3m1 164 hP5 370 600


2 3

EsF Hexagonal
3

EsF Monoclinic C2/c 15 mS60


4

EsCl Orange Hexagonal C6 /m hP8 727 410


3 3

EsBr Yellow Monoclinic C2/m 12 mS16 727 1259 681


3

EsI Amber [42] R3 148 hR24 753 2084


3 Hexagonal

EsOCl [] P4/nmm 394.8 670.2


Tetragonal

Oxides
Einsteinium(III) oxide (Es O ) was obtained by burning einsteinium(III) nitrate. It forms colorless cubic crystals,
2 3
[24]
which were first characterized from microgram samples sized about 30 nanometers. Two other phases,
monoclinic and hexagonal, are known for this oxide. The formation of a certain Es O phase depends on the
2 3
preparation technique and sample history, and there is no clear phase diagram. Interconversions between the three
[43]
phases can occur spontaneously, as a result of self-irradiation or self-heating. The hexagonal phase is
3+ 2–
isotypic with lanthanum(III) oxide where the Es ion is surrounded by a 6-coordinated group of O ions.

Halides
[44]
Einsteinium halides are known for the oxidation states +2 and +3.
The most stable state is +3 for all halides from fluoride to iodide.
Einsteinium(III) fluoride (EsF ) can be precipitated from
3
einsteinium(III) chloride solutions upon reaction with fluoride ions. An
alternative preparation procedure is to exposure einsteinium(III) oxide
to chlorine trifluoride (ClF ) or F gas at a pressure of 1–2 atmospheres
3 2
and a temperature between 300 and 400 °C. The EsF3 crystal structure
3+
is hexagonal, as in californium(III) fluoride (CfF ) where the Es ions
3
are 8-fold coordinated by fluorine ions in a bicapped trigonal prism
[45]
arrangement.
Einsteinium(III) chloride (EsCl ) can be prepared by annealing
3
einsteinium(III) oxide in the atmosphere of dry hydrogen chloride
vapors at about 500 °C for some 20 minutes. It crystallizes upon Einsteinium(III) iodide glowing in the dark
cooling at about 425 °C into an orange solid with a hexagonal structure
of UCl3 type, where einsteinium atoms are 9-fold coordinated by chlorine atoms in a tricapped trigonal prism
[] type, where
geometry. Einsteinium(III) bromide (EsBr ) is a pale-yellow solid with a monoclinic structure of
AlCl
3 3
the einsteinium atoms are octahedrally coordinated by bromine (coordination number 6).
[46]
The divalent compounds of einsteinium are obtained by reducing the trivalent halides with hydrogen:
2 EsX + H → 2 EsX + 2 HX, X = F, Cl, Br, I
3 2 2 [47] [48]
Einsteinium(II) chloride (EsCl ), einsteinium(II) bromide (EsBr ), and einsteinium(II) iodide (EsI ) have been
2 2 2
produced and characterized by optical absorption, with no structural information available yet.
Known oxyhalides of einsteinium include EsOCl, EsOBr and EsOI. They are synthesized by treating a trihalide with
a vapor mixture of water and the corresponding hydrogen halide: for example, EsCl + H O/HCl to obtain EsOCl.
3 []
2

Organoeinsteinium compounds
The high radioactivity of einsteinium has a potential use in radiation therapy, and organometallic complexes have
been synthesized in order to deliver einsteinium atoms to an appropriate organ in the body. Experiments have been
performed on injecting einsteinium citrate (as well as fermium compounds) to dogs. Einsteinium(III) was also
incorporated into beta-diketone chelate complexes, since analogous complexes with lanthanides previously showed
strongest UV-excited luminescence among metallorganic compounds. When preparing einsteinium complexes, the
3+ 3+
Es ions were 1000 times diluted with Gd ions. This allowed reducing the radiation damage so that the
compounds did not disintegrate during the period of 20 minutes required for the measurements. The resulting
3+
luminescence from Es was much too weak to be detected. This was explained by the unfavorable relative energies
3+
of the individual constituents of the compound that hindered efficient energy transfer from the chelate matrix to Es
ions. Similar conclusion was drawn for other actinides americium, berkelium and fermium.
3+
Luminescence of Es ions was however observed in inorganic hydrochloric acid solutions as well as in organic
solution with di(2-ethylhexyl)orthophosphoric acid. It shows a broad peak at about 1064 nanometers (half-width
about 100 nm) which can be resonantly excited by green light (ca. 495 nm wavelength). The luminescence has
a lifetime of several microseconds and the quantum yield below 0.1%. The relatively high, compared to
3+
lanthanides, non-radiative decay rates in Es were associated with the stronger interaction of f-electrons with
3+
the inner Es electrons.

Applications
There is almost no use for any isotope of einsteinium outside of basic scientific research aiming at production of
[49]
higher transuranic elements and transactinides.
9 253
In 1955, mendelevium was synthesized by irradiating a target consisting of about 10 atoms of Es in the 60-inch
253 256
cyclotron at Berkeley Laboratory. The resulting Es(α,n) Md reaction yielded 17 atoms of the new element with
the atomic number of 101.
The rare isotope einsteinium-254 is favored for production of ultraheavy elements because of its large mass,
relatively long half-life of 270 days, and availability in significant amounts of several micrograms. Hence
einsteinium-254 was used as a target in the attempted synthesis of ununennium (element 119) in 1985 by
bombarding it with calcium-48 ions at the superHILAC linear accelerator at Berkeley, California. No atoms were
identified, setting an upper limit for the cross section of this reaction at 300 nanobarns.

Einsteinium-254 was used as the calibration marker in the chemical analysis spectrometer ("alpha-scattering surface
analyzer") of the Surveyor 5 lunar probe. The large mass of this isotope reduced the spectral overlap between signals
from the marker and the studied lighter elements of the lunar surface.
Safety
Most of the available einsteinium toxicity data originates from research on animals. Upon ingestion by rats, only
about 0.01% einsteinium ends in the blood stream. From there, about 65% goes to the bones, where it remains
for about 50 years, 25% to the lungs (biological half-life about 20 years, although this is rendered irrelevant by
the short half-lives of einsteinium isotopes), 0.035% to the testicles or 0.01% to the ovaries – where
einsteinium stays indefinitely. About 10% of the ingested amount is excreted. The distribution of einsteinium
over the bone surfaces is uniform and is similar to that of plutonium.

References
[1] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_einsteinium&action=edit
[2] Einsteinium (http:/ /periodic.lanl.gov/99.shtml). periodic.lanl.gov
[3] Seaborg, p. 39
[4] John Emsley Nature's building blocks: an A-Z guide to the elements (http://books. google.com/books?id=j-Xu07p3cKwC&pg=PA133),
Oxford University Press, 2003, ISBN 0-19-850340-7 pp. 133–135
254 254 253 254
[5] Es, Fm and Fm would not be produced because of lack of beta decay in Cf and
253
Es [6] Google Books (http://books. google.com/books?id=e53sNAOXrdMC&pg=PA91)
[7] Google Books (http:/ /books.google. com/ books?id=e53sNAOXrdMC&pg=PA93)
[8] Byrne, J. Neutrons, Nuclei, and Matter, Dover Publications, Mineola, NY, 2011, ISBN 978-0-486-48238-5 (pbk.) pp. 267.
[9] Seaborg, G. T.; Thompson, S.G.; Harvey, B.G. and Choppin, G.R. (July 23, 1954) "Chemical Properties of Elements 99 and 100" (http://
www.osti. gov/ accomplishments/documents/fullText/ACC0047. pdf), Radiation Laboratory, University of California, Berkeley,
UCRL-2591
[10] Both died between the time the names were originally proposed and when they were announced.
[11] Haire, p. 1577
[12] Seaborg, G.T. (1994) Modern alchemy: selected papers of Glenn T. Seaborg (http:// books. google. com/books?id=e53sNAOXrdMC&
pg=PA6), World Scientific, p. 6, ISBN 981-02-1440-5.
[13] Haire, p. 1580
[14] Haire, R. G. (1990) "Properties of the Transplutonium Metals (Am-Fm)", in: Metals Handbook, Vol. 2, 10th edition, (ASM International,
Materials Park, Ohio), pp. 1198–1201.
[15] Haire, p. 1591
[16] draft manuscript (http:// www.osti.gov/bridge/servlets/purl/6582609-SrTVod/6582609.pdf)
[17] Seaborg, p. 61
[18] Seaborg, p. 52
[19] Seaborg, p. 55
[20] Seaborg, p. 76
[21] Holleman, p. 1956
[22] Seaborg, p. 53
[23] Haire, p. 1578
[24] Institut de Radioprotection et de Sûreté Nucléaire, "Evaluation of nuclear criticality safety data and limits for actinides in transport" (http://
ec.europa.eu/ energy/ nuclear/ transport/ doc/irsn_sect03_146.pdf), p. 16.
[25] http:// link.springer. com/article/10.3103%2FS0884591308020049
[26] Seaborg, p. 51
[27] Haire, p. 1582
[28] Greenwood, p. 1262
[29] Hoffman, Darleane C.; Ghiorso, Albert and Seaborg, Glenn Theodore (2000) The Transuranium People: The Inside Story, Imperial College
Press, pp. 190–191, ISBN 978-1-86094-087-3.
[30] Seaborg, pp. 36–37
[31] Harry H. Binder: Lexikon der chemischen Elemente, S. Hirzel Verlag, Stuttgart 1999, ISBN 3-7776-0736-3, pp. 18–
23. [32] Эйнштейний (http://n-t.ru/ri/ps/ pb099. htm) (in Russian, a popular article by one of the involved scientists)
[33] These were codenamed: "Anacostia" (5.2 kilotons, 1962), "Kennebec" (<5 kilotons, 1963), "Par" (38 kilotons, 1964), "Barbel" (<20
kilotons, 1964), "Tweed" (<20 kilotons, 1965), "Cyclamen" (13 kilotons, 1966), "Kankakee" (20-200 kilotons, 1966), "Vulcan" (25 kilotons,
1966) and "Hutch" (20-200 kilotons, 1969)
[34] United States Nuclear Tests July 1945 through September 1992 (http://www.nv.doe. gov/ library/publications/ historical/
DOENV_209_REV15. pdf), DOE/NV--209-REV 15, December 2000.
[35] Seaborg, p. 43
[36] Seaborg, p. 44
[37] Seaborg, p. 47
[38] Haire, p. 1583
[39] Haire, pp. 1584–1585
[40] Haire, p. 1588
[41] Haire, p. 1590
[42] Seaborg, p. 62
[43] Haire, p. 1598
[44] Holleman, p. 1969
[45] Greenwood, p. 1270
[46] manuscript draft (http://www.osti.gov/energycitations/product.biblio.jsp?osti_id=6593662)
[47] Fellows, R.L.; Young, J.P.; Haire, R.G. and Peterson J.R. (1977) in: GJ McCarthy and JJ Rhyne (eds) The Rare Earths in Modern Science
and Technology, Plenum Press, New York, pp. 493–499.
[48] Young, J.P.; Haire R.G., Fellows, R.L.; Noe, M. and Peterson, J.R. (1976) "Spectroscopic and X-Ray Diffraction Studies of the
Bromides of Californium-249 and Einsteinium-253", in: W. Müller and R. Lindner (eds.) Plutonium 1975, North Holland, Amsterdam, pp.
227–234.
[49] It's Elemental – The Element Einsteinium (http://education. jlab.org/itselemental/ele099.html). Retrieved 2 December 2007.

Bibliography
• Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth–Heinemann.
ISBN 0080379419.
• Haire, Richard G. (2006). "Einsteinium" (http://radchem.nevada.edu/classes/rdch710/files/einsteinium.pdf).
In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean. The Chemistry of the Actinide and Transactinide
Elements 3 (3rd ed.). Dordrecht, the Netherlands: Springer. pp. 1577–1620. doi: 10.1007/1-4020-3598-5_12
(http://dx.doi.org/10.1007/1-4020-3598-5_12).
• Holleman, Arnold F. and Wiberg, Nils (2007). Textbook of Inorganic Chemistry (102 ed. ed.). Berlin: de Gruyter.
ISBN 978-3-11-017770-1.
• Seaborg, G.T., ed. (23 January 1978). Proceedings of the Symposium Commemorating the 25th Anniversary
of Elements 99 and 100 (http://www.escholarship.org/uc/item/92g2p7cd.pdf). Report LBL-7701.

External links
• Einsteinium (http://www.periodicvideos.com/videos/099.htm) at The Periodic Table of Videos (University
of Nottingham)
• Age-related factors in radionuclide metabolism and dosimetry: Proceedings (http://books.google.com/
books?id=cgqNoNWLGBMC&pg=PA311) – contains several health related studies of einsteinium
100 Fermium 1

100 – Fermium
Not to be confused with ferrum.

Fermium
Fm
100

einsteinium ← fermium → mendeleviumEr



Fm

(Upb)

Fermium in the periodic table


Appearance

unknown

General properties

Name, symbol, number fermium, Fm, 100

Pronunciation /ˈfɜrmiəm/
FER-mee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (257)

Electron configuration 12 2
[Rn] 5f 7s
2, 8, 18, 32, 30, 8, 2

Physical properties

Phase solid (predicted)

Melting point 1125 K, 852 °C, 1565 (predicted) °F

Atomic properties

Oxidation states 2, 3

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 627 kJ·mol

Miscellanea

CAS registry number 7440-72-4

History
Naming after Enrico Fermi

Discovery Lawrence Berkeley National Laboratory (1952)

Most stable isotopes

Main article: Isotopes of fermium

iso NA half-life DM DE (MeV) DP

252 syn 25.39 h SF - -


Fm
α 7.153 248
Cf

253 syn 3d ε 0.333 253


Fm Es
α 7.197 249
Cf

255 syn 20.07 h SF - -


Fm
α 7.241 251
Cf

257 syn 100.5 d α 6.864 253


Fm Cf
SF - -

• v
• t
[1]
• e

Fermium is a synthetic element with symbol Fm and atomic number 100. It is a member of the actinide series. It is
the heaviest element that can be formed by neutron bombardment of lighter elements, and hence the last element
that can be prepared in macroscopic quantities, although pure fermium metal has not yet been prepared. A total
257
of 19 isotopes are known, with Fm being the longest-lived with a half-life of 100.5 days.
It was discovered in the debris of the first hydrogen bomb explosion in 1952, and named after Enrico Fermi, one of
the pioneers of nuclear physics. Its chemistry is typical for the late actinides, with a preponderance of the +3
oxidation state but also an accessible +2 oxidation state. Owing to the small amounts of produced fermium and all of
its isotopes having relatively short half-lives, there are currently no uses for it outside of basic scientific research.

Discovery
Fermium was first discovered in the fallout from the 'Ivy Mike' nuclear
[2]
test (1 November 1952), the first successful test of a hydrogen bomb.
Initial examination of the debris from the explosion had shown the
production of a new isotope of plutonium, 244
94Pu: this could only have formed by the absorption of six neutrons

by a uranium-238 nucleus followed by two β decays. At the time,
the absorption of neutrons by a heavy nucleus was thought to be a
rare process, but the identification of 244
94Pu raised the possibility that still more neutrons could have been

Fermium was first observed in the fallout from absorbed by the uranium nuclei, leading to new elements.
the Ivy Mike nuclear test.
Element 99 (einsteinium) was quickly discovered on filter papers
which had been flown through the cloud from the explosion (the
same
sampling technique that had been used to discover
244
94Pu). It was then identified in December 1952 by Albert Ghiorso and
[2]
co-workers at the University of California at Berkeley. They
253
discovered the isotope Es (half-life 20.5 days) that was made by the
capture of 15 neutrons by uranium-238 nuclei – which then underwent
seven successive beta decays:

238
Some U atoms, however, could capture another amount of neutrons
(most likely, 16 or
17).
The discovery of fermium (Z = 100) required more material, as the
yield was expected to be at least an order of magnitude lower than that
of element 99, and so contaminated coral from the Enewetak atoll
The element was named after Enrico Fermi. (where the test had taken place) was shipped to the University of
California Radiation Laboratory in Berkeley, California, for processing
and analysis. About two months after the test, a new component was
isolated emitting high-energy α-particles (7.1 MeV) with a half-life of about a day. With such a short half-life, it

could only arise from the β decay of an isotope of einsteinium, and so had to be an isotope of the new element 100:
255
it was quickly identified as Fm (t = 20.07(7) hours).
½
The discovery of the new elements, and the new data on neutron capture, was initially kept secret on the orders of the
[3]
U.S. military until 1955 due to Cold War tensions. Nevertheless, the Berkeley team were able to prepare
elements 99 and 100 by civilian means, through the neutron bombardment of plutonium-239, and published this
work in 1954 with the disclaimer that it was not the first studies that had been carried out on the elements. The
'Ivy Mike' studies were declassified and published in 1955.
The Berkeley team had been worried that another group might discover lighter isotopes of element 100 through ion
bombardment techniques before they could publish their classified research, and this proved to be the case. A group
at the Nobel Institute for Physics in Stockholm independently discovered the element, producing an isotope later
250
confirmed to be Fm (t = 30 minutes) by bombarding a 238
½
92U target with oxygen-16 ions, and published their work in May 1954. Nevertheless, the priority of the Berkeley
team was generally recognized, and with it the prerogative to name the new element in honour of the recently
deceased Enrico Fermi, the developer of the first artificial self-sustained nuclear reactor.
Isotopes
Main article: Isotopes of fermium

Decay pathway of fermium-257

[4] 257
There are 19 isotopes of fermium listed in NUBASE 2003, with atomic weights of 242 to 260, of which Fm is the
253 251 252
longest-lived with a half-life of 100.5 days. Fm has a half-life of 3 days, while Fm of 5.3 h, Fm of 25.4 h,
254 255 256
Fm of 3.2 h, Fm of 20.1 h, and Fm of 2.6 hours. All the remaining ones have half-lives ranging from 30
258
minutes to less than a millisecond. The neutron-capture product of fermium-257, Fm, undergoes spontaneous
259 260
fission with a half-life of just 370(14) microseconds; Fm and Fm are also unstable with respect to spontaneous
fission (t ½ = 1.5(3) s and 4 ms respectively). This means that neutron capture cannot be used to create nuclides with
257
a mass number greater than 257, unless carried out in a nuclear explosion. As Fm is an α-emitter, decaying to
253
Cf, and no fermium isotopes undergo beta minus decay (which would produce isotopes of the next element,
mendelevium), fermium is also the last element that can be prepared by a neutron-capture process.

Production
Fermium is produced by the bombardment of lighter actinides with
neutrons in a nuclear reactor. Fermium-257 is the heaviest isotope that
is obtained via neutron capture, and can only be produced in nanogram
[5]
quantities. The major source is the 85 MW High Flux Isotope
Reactor (HFIR) at the Oak Ridge National Laboratory in Tennessee,
USA, which is dedicated to the production of transcurium (Z > 96)
elements. In a "typical processing campaign" at Oak Ridge, tens of
grams of curium are irradiated to produce decigram quantities of
californium, milligram quantities of berkelium and einsteinium and
picogram quantities of fermium. However, nanogram and microgram
quantities of fermium can be prepared for specific experiments. The
quantities of fermium produced in 20–200 kiloton thermonuclear
explosions is believed to be of the order of milligrams, although it is
257
mixed in with a huge quantity of debris; 40 picograms of Fm was
recovered from 10 kilograms of debris from the 'Hutch' test (16 July Elution: chromatographic separation of Fm(100),
1969). Es(99), Cf, Bk, Cm and Am.

After production, the fermium must be separated from other actinides and from lanthanoid fission products. This
is usually achieved by ion exchange chromatography, with the standard process using a cation exchanger such
as Dowex 50 or TEVA eluted with a solution of ammonium α-hydroxyisobutyrate. Smaller cations form more
stable complexes with the α-hydroxyisobutyrate anion, and so are preferentially eluted from the column. A rapid
fractional crystallization method has also been described.
257
Although the most stable isotope of fermium is Fm, with a half-life of 100.5 days, most studies are conducted on
255 255
Fm (t = 20.07(7) hours) as this isotope can be easily isolated as required as the decay product of Es (t =
½ ½
39.8(12) days).

Synthesis in nuclear explosions


The analysis of the debris at the 10-megaton Ivy Mike nuclear test was a part of long-term project, one of the goals of
which was studying the efficiency of production of transuranium elements in high-power nuclear explosions. The
motivation for these experiments was as follows: synthesis of such elements from uranium requires multiple neutron
capture. The probability of such events increases with the neutron flux, and nuclear explosions are the most powerful
23 29
neutron sources, providing densities of the order 10 neutrons/cm² within a microsecond, i.e. about 10
15
neutrons/(cm²·s). In comparison, the flux of the HFIR reactor is 5×10 neutrons/(cm²·s). A dedicated laboratory was
set up right at Enewetak Atoll for preliminary analysis of debris, as some isotopes could have decayed by the time
the debris samples reached the U.S. The laboratory was receiving samples for analysis, as soon as possible, from
airplanes equipped with paper filters which flew over the atoll after the tests. Whereas it was hoped to discover new
chemical elements heavier than fermium, those were not found after a series of megaton explosions conducted
[6]
between 1954 and 1956 at the atoll.
The atmospheric results were supplemented
by the underground test data accumulated in
the 1960s at the Nevada Test Site, as it was
hoped that powerful explosions conducted
in confined space might result in improved
yields and heavier isotopes. Apart from
traditional uranium charges, combinations
of uranium with americium and thorium
have been tried, as well as a mixed
plutonium-neptunium charge. They were
less successful in terms of yield that was
attributed to stronger losses of heavy
isotopes due to enhanced fission rates in
heavy-element charges. Isolation of the
products was found to be rather problematic,
Estimated yield of transuranium elements in the U.S. nuclear tests Hutch and as the explosions were spreading debris
Cyclamen. through melting and vaporizing rocks under
the great depth of 300–600 meters, and
[]
drilling to such depth in order to extract the products was both slow and inefficient in terms of collected volumes.

Among the nine underground tests, which were carried between 1962 and 1969 and codenamed Anacostia (5.2
kilotons, 1962), Kennebec (<5 kilotons, 1963), Par (38, kilotons, 1964), Barbel (<20 kilotons, 1964), Tweed (<20
kilotons, 1965), Cyclamen (13 kilotons, 1966), Kankakee (20-200 kilotons, 1966), Vulcan (25 kilotons, 1966) and
[7]
Hutch (20-200 kilotons, 1969), the last one was most powerful and had the highest yield of transuranium elements.
In the dependence on the atomic mass number, the yield showed a saw-tooth behavior with the lower values for odd
isotopes, due to their higher fission rates. The major practical problem of the entire proposal was however collecting
−14
the radioactive debris dispersed by the powerful blast. Aircraft filters adsorbed only about 4×10 of the total
amount and collection of tons of corals at Enewetak Atoll increased this fraction by only two orders of magnitude.
Extraction of about 500 kilograms of underground rocks 60 days after the Hutch explosion recovered only about
−7
10 of the total charge. The amount of transuranium elements in this 500-kg batch was only 30 times higher than in
a 0.4 kg rock picked up 7 days after the test. This observation demonstrated the highly nonlinear dependence of the
[8]
transuranium elements yield on the amount of retrieved radioactive rock. In order to accelerate sample collection
after explosion, shafts were drilled at the site not after but before the test, so that explosion would expel radioactive
material from the epicenter, through the shafts, to collecting volumes near the surface. This method was tried in the
Anacostia and Kennebec tests and instantly provided hundreds kilograms of material, but with actinide concentration
3 times lower than in samples obtained after drilling; whereas such method could have been efficient in scientific
[9]
studies of short-lived isotopes, it could not improve the overall collection efficiency of the produced actinides.
Although no new elements (apart from einsteinium and fermium) could be detected in the nuclear test debris, and the
total yields of transuranium elements were disappointingly low, these tests did provide significantly higher amounts
9 257
of rare heavy isotopes than previously available in laboratories. So 6×10 atoms of Fm could be recovered after
257
the Hutch detonation. They were then used in the studies of thermal-neutron induced fission of Fm and in
258 250
discovery of a new fermium isotope Fm. Also the rare Cm isotope was synthesized in large quantities, which is
249 249
very difficult to produce in nuclear reactors from its progenitor Cm – the half-life of Cm (64 minutes) is much
[10]
too short for months-long reactor irradiations, but is very "long" on the explosion timescale.
Natural occurrence
Because of the short half-life of all isotopes of fermium, all primordial fermium, that is fermium that could possibly
be present on the Earth during its formation, has decayed by now. Synthesis of fermium from naturally occurring
actinides uranium and thorium in the Earth crust requires multiple neutron capture, which is an extremely unlikely
event. Therefore, most fermium is produced on Earth in scientific laboratories, high-power nuclear reactors, or in
nuclear weapons tests, and is present only within a few months from the time of the synthesis. Einsteinium and
fermium did occur naturally in the natural nuclear fission reactor at Oklo, but no longer do so.

Chemistry
The chemistry of fermium has only been studied in solution using
tracer techniques, and no solid compounds have been prepared. Under
3+
normal conditions, fermium exists in solution as the Fm ion, which
has a hydration number of 16.9 and an acid dissociation constant of
−4
1.6×10 (pK = 3.8). Fm3+ forms complexes with a wide variety of
a
organic ligands with hard donor atoms such as oxygen, and these
complexes are usually more stable than those of the preceding
actinides. It also forms anionic complexes with ligands such as
chloride or nitrate and, again, these complexes appear to be more
stable than those formed by einsteinium or californium. It is believed
that the bonding in the complexes of the later actinides is mostly
A fermium-ytterbium alloy used for measuring
ionic in
3+
character: the Fm ion is expected to be smaller than the preceding the enthalpy of vaporization of fermium metal.
3+
An ions because of the higher effective nuclear charge of
fermium,
and hence fermium would be expected to form shorter and stronger
metal–ligand bonds.

Fermium(III) can be fairly easily reduced to fermium(II), for example with samarium(II) chloride, with which
fermium coprecipitates. The electrode potential has been estimated to be similar to that of the ytterbium(III)/(II)
couple, or about −1.15 V with respect to the standard hydrogen electrode, a value which agrees with theoretical
2+ 0
calculations. The Fm /Fm couple has an electrode potential of −2.37(10) V based on polarographic measurements.

Toxicity
Although few people come in contact with fermium, the International Commission on Radiological Protection has
7
set annual exposure limits for the two most stable isotopes. For fermium-253, the ingestion limit was set at 10
5 5
Becquerels (1 Bq is equivalent to one decay per second), and the inhalation limit at 10 Bq; for fermium-257, at 10
Bq and 4000 Bq respectively.

Notes and references

Notes
[1] http:// en. wikipedia.org/w/index.php?title=Template:Infobox_fermium&action=edit
[2] Fermium – National Research Council Canada (http://www.nrc-cnrc.gc. ca/eng/education/elements/el/fm.html). Retrieved 2 December
2007
[3] Fields, P. R.; Studier, M. H.; Diamond, H.; Mech, J. F.; Inghram, M. G. Pyle, G. L.; Stevens, C. M.; Fried, S.; Manning, W. M.
(Argonne National Laboratory, Lemont, Illinois); Ghiorso, A.; Thompson, S. G.; Higgins, G. H.; Seaborg, G. T. (University of
California, Berkeley, California): "Transplutonium Elements in Thermonuclear Test Debris", in:
260
[4] The discovery of Fm is considered "unproven" in NUBASE
2003.
[5] All isotopes of elements Z > 100 can only be produced by accelerator-based nuclear reactions with charged particles and can be obtained only
in tracer quantities (e.g., 1 million atoms for Md (Z = 101) per hour of irradiation (see reference 1 below)).
[6] Seaborg, p. 39
[7] United States Nuclear Tests July 1945 through September 1992 (http://www.nv.doe. gov/ library/publications/ historical/
DOENV_209_REV15. pdf), DOE/NV--209-REV 15, December 2000
[8] Seaborg, p. 43
[9] Seaborg, p. 44
[10] Seaborg, p. 47

References

Further reading
• Robert J. Silva: Fermium, Mendelevium, Nobelium, and Lawrencium (http://radchem.nevada.edu/classes/
rdch710/files/Fm to Lr.pdf), in: Lester R. Morss, Norman M. Edelstein, Jean Fuger (Hrsg.): The Chemistry of
the Actinide and Transactinide Elements, Springer, Dordrecht 2006; ISBN 1-4020-3555-1, p. 1621–1651; doi:
10.1007/1-4020-3598-5_13 (http://dx.doi.org/10. 1007/1-4020-3598-5_13).
• Seaborg, G.T. (ed.) (1978) Proceedings of the Symposium Commemorating the 25th Anniversary of Elements 99
and 100 (http://www.escholarship. org/uc/item/92g2p7cd. pdf), 23 January 1978, Report LBL-7701
• Gmelins Handbuch der anorganischen Chemie, System Nr. 71, Transurane: Teil A 1 II, p. 19–20; Teil A 2, p.
47; Teil B 1, p. 84.

External links
• Fermium (http://www.periodicvideos.com/videos/100.htm) at The Periodic Table of Videos (University
of Nottingham)
101 Mendelevium 1

101 – Mendelevium
Mendelevium
Md
101

fermium ← mendelevium → nobeliumTm



Md

(Upt)

Mendelevium in the periodic table


Appearance

unknown

General properties

Name, symbol, number mendelevium, Md, 101

Pronunciation /ˌmɛndəˈlɛviəm/
or /ˌmɛndəˈliːviəm/

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight (258)

Electron configuration 13 2
[Rn] 5f 7s
2, 8, 18, 32, 31, 8, 2

Physical properties

Phase solid (predicted)

Melting point 1100 K, 827 °C, 1521 (predicted) °F

Atomic properties

Oxidation states 2, 3

Electronegativity 1.3 (Pauling scale)

Ionization energies −1
1st: 635 kJ·mol

Miscellanea

CAS registry number 7440-11-1

History

Naming after Dmitri Mendeleev


Discovery Lawrence Berkeley National Laboratory (1955)

Most stable isotopes

Main article: Isotopes of mendelevium

iso NA half-life DM DE (MeV) DP

257 syn 5.52 h ε 0.406 257


Md Fm
α 7.558 253
Es
SF - -

258 syn 51.5 d ε 1.230 258


Md Fm

260 syn 31.8 d SF - -


Md
α 7.000 256
Es
ε - 260
Fm

− 1.000 260
β No

• v
• t
[1]
• e

Mendelevium is a synthetic element with the symbol Md (formerly Mv) and the atomic number 101. A metallic
radioactive transuranic element in the actinide series, mendelevium is usually synthesized by bombarding
einsteinium with alpha particles. It was named after Dmitri Ivanovich Mendeleev, who created the periodic table, the
standard way to classify all the chemical elements.

History
Mendelevium (for Dimitri Ivanovich Mendeleev, surname commonly transliterated into Latin script as Mendeleev,
Mendeleyev, Mendeléef, or even Mendelejeff, and first name sometimes transliterated as Dmitry or Dmitriy) was
first synthesized by Albert Ghiorso, Glenn T. Seaborg, Gregory R. Choppin, Bernard G. Harvey, and Stanley G.
256
Thompson (team leader) in early 1955 at the University of California, Berkeley. The team produced Md (half-life
253
of 87 minutes) when they bombarded an Es target with alpha particles (helium nuclei) in the Berkeley Radiation
256
Laboratory's 60-inch cyclotron ( Md was the first isotope of any element to be synthesized one atom at a time).
Mendelevium was the ninth transuranic element synthesized. The first 17 atoms of this element were created and
analyzed using the ion-exchange adsorption-elution method. During the process, mendelevium behaved very much
like thulium, its naturally occurring homologue.

Discovery
253 256
The discovery was based on a grand total of only 17 atoms. It is synthesized via the Es (α,n) 101 reaction
in the 60-Inch-Cyclotron (= 152 cm) at Berkeley (California). The target can be produced by irradiation of
lighter isotopes as plutonium in the Materials Testing Reactor at the Arco Reactor Station in Idaho. Remarkable
9 253
is that this target consisted of only 10 atoms of highly radioactive Es (with a half-life of 20.5 days). By
elution through a calibrated cation exchange resin column, mendelevium was separated and chemically identified.
Determining feasibility
To predict if this method would be possible, they made use of a rough calculation. The number of atoms that
would be produced would be approximately equal to the number of atoms of target material times its cross
section times the ion beam intensity times the time of bombardment related to the half-life of the product when
bombarding for a time of the order of its half-life. This gave one atom per experiment. Thus under optimum
conditions, the preparation of only one atom of element 101 per experiment could be expected. This calculation
demonstrated that it was feasible to go ahead with the experiment.

Recoil
technique
The actual synthesis was done by a recoil technique, introduced by Albert Ghiorso. In this technique, the target
element was placed on the opposite side of the target from the beam and caught the recoiling atoms on a catcher foil.
This recoil target was made by an electroplating technique, developed by Alfred Chetham-Strode. This technique
gave a very high yield, which is absolutely necessary when working with such a rare product as the einsteinium
target material.
9 253
The recoil target consisted of 10 atoms of Es which were deposited electrolytically on a thin gold foil (also Be,
Al and Pt can be used). It was bombarded by 41 eV α-particles in the Berkeley cyclotron with a very high beam
13 2
density of 6∙10 particles per second over an area of 0.05 cm . The target was cooled by water or liquid helium.
The use of helium, in a gaseous atmosphere, slowed down the recoil atoms. This gas could be pumped out of the
reaction chamber through a small orifice to form a ‘gas-jet’. Some fraction of the nonvolatile product atoms carried
along with the gas, were deposited permanently on the foil surface. The foil could be removed periodically and a
253 4
new foil could be installed. The next reaction was used for the mendelevium discovery experiment: Es + He →
256 1
Md + n.

Purification and isolation


The removal of the mendelevium atoms from the collector foil was done by acid etching or total dissolution of
the thin gold foil. They can be purified and isolated from other product activities by several techniques.
Separation of trivalent actinides from lanthanide fission products and La carrier can be done by a cation-
exchange resin column using a 90% water/10% ethanol solution saturated with HCl as eluant. To separate Md
rapidly from the catcher foil, an anion-exchange chromatography using 6M HCl as eluant can be used. The gold
3+
remained on the column while the Md and other actinides passed through. A final isolation of Md from other
trivalent actinides was also required. To separate fractions containing elements 99, 100 and 101, a cation-
exchange resin column (Dowex-50 exchange column) treated with ammonium salts was used. A chemical
identification was made on the basis of its elution position just before fermium. In series of repetitive
experiments, they made use of the eluant: α-hydroxyisobutyrate solution (α-HIB). Using the ‘gas-jet’ method,
the first two steps can be eliminated. There was shown that in this method it is possible to transport and collect
individual product atoms in a fraction of a second some tens of meters away from the target area. Effective
transport over long distances requires the presence of large clusters (KCl aerosols) in the ‘carrier’ gas. It is
used frequently in the production and isolation of transeinsteinium elements.
Another possible way to separate the 3+ actinides can be achieved by solvent extraction chromatography using
bis-(2-ethylhexyl) phosphoric acid (abbreviated as HDEHP) as the stationary organic phase and HNO as the mobile
3
aqueous phase. The actinide elution sequence is reversed from that of the cation-exchange resin column. The Md
separated by this method has the advantage to be free of organic complexing agent compared to the resin column.
The disadvantage of this method is that Md elutes after Fm late in the sequence.
The first "Hooray!"
There was no direct detection, but by observation of spontaneous fission events arising from its electron-capture
256
daughter Fm. These events were recorded during the night of February 19, 1955. The first one was identified with
a "hooray" followed by a "double hooray" and a "triple hooray". The fourth one eventually officially proved the
chemical identification of the 101st element, mendelevium. Additional analysis and further experimentation, showed
the isotope to have mass 256 and to decay by electron capture with a half-life of 1.5 h.

Naming
The name "mendelevium" was accepted by the International Union of Pure and Applied Chemistry (IUPAC) in
1955 with symbol "Mv", which was changed to "Md" in the next IUPAC General Assembly (Paris, 1957).

Characteristics
Researchers have shown that mendelevium has a moderately stable dipositive (II) oxidation state in addition to the
more characteristic (for actinide elements) tripositive (III) oxidation state, the latter being the more dominantly
256
exhibited state in an aqueous solution (chromatography being the process used). Md has been used to find out
some of the chemical properties of this element while in an aqueous solution. There are no other known uses of
mendelevium and only trace amounts of the element have ever been produced. Other isotopes of mendelevium, all
258
radioactive, have been discovered, with Md being the most stable with a two-month half-life (about 55 days).
Other isotopes range from 248 to 258 mass numbers and half-lives from a few seconds to about 51 days. The original
256
Md had a half-life of 87 minutes.

Radioactivity
253 255 255
The trivalent element is radioactive. It was expected that the reaction would be Es (α,n) Md, where Md was
α-active with a t of 5 minutes and the corresponding α-energy. No such α-activity was observed, but the 101
½
fraction showed spontaneous fission representing a t ½ less than 3 hours. Because spontaneous fission was also
253 256
observed in the fraction containing fermium, the α-bombardment of Es produced Md. The latter underwent
256 256
electron capture to become Fm, which then decayed by spontaneous fission. So Fm was produced by the
decays of cyclotron-synthesized mendelevium.

Metallic state
Johansson and Rosengren predicted in 1975 that Md would prefer a divalent metallic state, similar to europium (Eu)
and ytterbium (Yb), rather than a trivalent one. Thermochromatographic studies conducted with trace amounts of Md
concluded that Md forms a divalent metal. With the aid of empirical correlation method, a divalent metallic radius of
(0.194 ± 0.010) nm has been estimated. The estimated enthalpy of sublimation is in the range of 134-142
kJ/mol.

Solution
chemistry
Before the actual discovery of mendelevium, the trivalent state was predicted to be the most stable one in aqueous
solution. Accordingly, a similar chemical behavior to the other 3+ actinides and lanthanides was expected. The
elution of Md just before Fm in the elution sequence of the trivalent actinides from the cation-exchange resin
column, confirmed this prediction. Afterwards, Md in the form of insoluble hydroxides and fluorides that are
quantitatively coprecipitated with trivalent lanthanides was found. The cation-exchange resin column as well as the
HDEHP solvent extraction column elution date is consistent with a trivalent state for Md and an ionic radius smaller
3+
than Fm. An ionic radius of 0.0192 nm and a coordination number of 6 for Md was predicted using empirical
correlations. Using the known ionic radii for the trivalent rare earths and the linear correlation of log distribution
3+
coefficient with ionic radius, an average ionic radius of 0.089 nm was estimated for Md and a heat of hydration of
–(3654 ± 12) kJ/mol calculated using empirical models and the Born-Haber cycle. In reducing conditions, an
anomalous chemical behavior of Md was found. Coprecipitation with BaSO4 and solvent extraction chromatography
3+
experiments using HDEHP were carried out in different reducing agents. These showed that Md could easily be
2+
reduced to a stable Md in aqueous solution. Mendelevium can also be reduced to the monovalent state in
+
water-ethanol solutions. The cocrystallization of Md with salts of divalent ions is due to the formation of mixed
+ 3+ 4+
crystals. For Md , an ionic radius of 0.117 nm was found. The oxidation of Md to Md was rather
unsuccessful.

Isotopes
Main article: Isotopes of
mendelevium
258
Sixteen isotopes of mendelevium from mass 245 to 260 have been characterized, with the most stable being Md
260 257
with a half-life of 51.5 days, Md with a half-life of 31.8 days, and Md with a half-life of 5.52 hours. All of the
remaining radioactive isotopes have half-lives that are less than 97 minutes, and the majority of these have half-lives
258m
that are less than 5 minutes. This element also has 5 meta states, with the longest-lived being Md (t = 58
½
245 260
minutes). The isotopes of mendelevium range in atomic weight from 245.091 u ( Md) to 260.104 u ( Md).

References
[1] http:// en.wikipedia.org/w/index.php?title=Template:Infobox_mendelevium&action=edit

Further reading
• Hoffman, D.C., Ghiorso, A., Seaborg, G. T. The transuranium people: the inside story, (2000), 201–229
• Morss, L. R., Edelstein, N. M., Fuger, J., The chemistry of the actinide and transactinide element, 3, (2006),
1630–1636
• Seaborg, G. T., Les elements tranuraniens artificiels, (1967), 39–45
• Gol’danskii, V. I., Polikanov, S. M., The transuranium elements, (1973), 101–103
• Seaborg, G.T., The transcalifornium elements. Journal of Chemical Education, (1959), 36, 38–44
• Guide to the Elements – Revised Edition, Albert Stwertka, (Oxford University Press; 1998) ISBN 0-19-508083-1

External links
• Los Alamos National Laboratory – Mendelevium (http://periodic.lanl.gov/elements/101.html)
• It's Elemental – Mendelevium (http://education.jlab.org/itselemental/ele101.html)
• Mendelevium (http://www.periodicvideos.com/videos/101.htm) at The Periodic Table of Videos
(University of Nottingham)
• Environmental Chemistry – Md info (http://environmentalchemistry.com/yogi/periodic/Md.html)
102 Nobelium 1

102 – Nobelium
Nobelium
No
102

mendelevium ← nobelium → lawrenciumYb



No

(Upq)

Nobelium in the periodic table


Appearance

unknown

General properties

Name, symbol, number nobelium, No, 102

Pronunciation i
/noʊˈbɛliəm/ noh-BEL-ee-əm
or /noʊˈbiːliəm/ noh-BEE-lee-əm

Element category actinide

Group, period, block n/a, 7, f

Standard atomic weight [259]

Electron configuration 14 2
[Rn] 5f 7s
2, 8, 18, 32, 32, 8, 2

Physical properties

Phase solid (predicted)

Melting point 1100 K, 827 °C, 1521 (predicted) °F

Atomic properties

Oxidation states 2, 3
[1]
Electronegativity 1.3 (Pauling scale)
(predicted)

Ionization energies −1
1st: 641.6 kJ·mol

−1
2nd: 1254.3 kJ·mol

−1
3rd: 2605.1 kJ·mol

Miscellanea
CAS registry number 10028-14-5

History

Naming after Alfred Nobel

Discovery Joint Institute for Nuclear Research (1966)

Most stable isotopes

Main article: Isotopes of nobelium

iso NA half-life DM DE (MeV) DP

253 syn 1.62 min 80% α 8.14, 8.06, 8.04, 8.01 249
No Fm
+ 253
20% β Md

254 syn 51 s 90% α 250


No Fm
+ 254
10% β Md

255 syn 3.1 min 61% α 8.12, 8.08, 7.93 251


No Fm
+ 2.012 255
39% β Md

257 syn 25 s 99% α 8.32, 8.22 253


No Fm
+ 257
1% β Md

259 syn 58 min 75% α 7.69, 7.61, 7.53.... 255


No Fm
25% ε 259
Md
10% SF

only isotopes with half-lives over 5 seconds are included


here

• v
• t
[2]
• e

Nobelium is a synthetic element with the symbol No and atomic number 102. It was first correctly identified in 1966
by scientists at the Flerov Laboratory of Nuclear Reactions in Dubna, Soviet Union. Little is known about the
element but limited chemical experiments have shown that it forms a stable divalent ion in solution as well as the
predicted trivalent ion that is associated with its presence as one of the actinides.

Discovery profile
The discovery of element 102 was first announced by physicists at the Nobel Institute in Sweden in 1957. The team
reported that they created an isotope with a half-life of 10 minutes, decaying by emission of an 8.5 MeV alpha
244 13 251 253
particle, after bombarding Cm with C nuclei. The activity was assigned to No or No. The scientists
[3]
proposed the name nobelium (No) for the new element. Later they retracted their claim and associated the activity
to background effects.
The synthesis of element 102 was then claimed in April 1958 at the University of California, Berkeley, by Albert
Ghiorso, Glenn T. Seaborg, John R. Walton and Torbjørn Sikkeland. The team used the new heavy-ion linear
244 246 13 12
accelerator (HILAC) to bombard a curium target (95% Cm and 5% Cm) with C and C ions. They were
250
unable to confirm the 8.5 MeV activity claimed by the Swedes but were instead able to detect decays from Fm,
254
supposedly the daughter of 102, which had an apparent half-life of ~3 s. In 1959 the team continued their studies
and claimed that they were able to produce an isotope that decayed predominantly by emission of an 8.3 MeV alpha
particle, with a half-life of 3 s with an associated 30% spontaneous fission branch. The activity was initially assigned
254 252
to No but later changed to No. The Berkeley team decided to adopt the name nobelium for the element.
244
96Cm + 12
6C → 256
102No*
→ 252
102No + 4 1
0n

Further work in 1961 on the attempted synthesis of element 103 produced evidence for a Z=102 alpha activity
255
decaying by emission of an 8.2 MeV particle with a half-life of 15 s, and assigned to No.
Following initial work between 1958–1964, in 1966, a team at the Flerov Laboratory of Nuclear Reactions (FLNR)
250 254
reported that they had been able to detect Fm from the decay of a parent nucleus ( No) with a half-life of ~50s,
in contradiction to the Berkeley claim. Furthermore, they were able to show that the parent decayed by emission of
8.1 MeV alpha particles with a half-life of ~35 s.
238
92U + 22
10Ne → 260
102No*
→ 254
102No + 6 1
0n

In 1969, the Dubna team carried out chemical experiments on element 102 and concluded that it behaved as the
heavier homologue of ytterbium. The Russian scientists proposed the name joliotium (Jo) for the new element.
Later work in 1967 at Berkeley and 1971 at Oak Ridge fully confirmed the discovery of element 102 and clarified
earlier observations.
In 1992, the IUPAC-IUPAP Transfermium Working Group (TWG) assessed the claims of discovery and concluded
that only the Dubna work from 1966 correctly detected and assigned decays to Z=102 nuclei at the time. The Dubna
team are therefore officially recognised as the discoverers of nobelium although it is possible that it was detected at
Berkeley in 1959.

Naming
Element 102 was first named nobelium (No) by its claimed discoverers in 1957 by scientists at the Nobel Institute in
Sweden. The name was later adopted by Berkeley scientists who claimed its discovery in 1959.
The International Union of Pure and Applied Chemistry (IUPAC) officially recognised the name nobelium following
the Berkeley results.
In 1994, and subsequently in 1997, the IUPAC ratified the name nobelium (No) for the element on the basis that it
had become entrenched in the literature over the course of 30 years and that Alfred Nobel should be commemorated
in this fashion.
Physical properties
The appearance of this element is unknown, however it is most likely silvery-white or gray and metallic. If sufficient
amounts of nobelium were produced, it would pose a radiation hazard. Some sources quote a melting point of 827 °C
for nobelium but this cannot be substantiated from an official source and seems implausible regarding the
[citation
requirements of such a measurement. However, the 1st, 2nd and 3rd ionization energies have been measured
needed]
. In addition, an electronegativity value of 1.3 is also sometimes quoted.

Experimental chemistry

Aqueous phase chemistry


First experiments involving nobelium assumed that it predominantly formed a +III state like earlier actinoids.
However, it was later found that nobelium forms a stable +II state in solution, although it can be oxidized to the
2+ 3+ [4]
+III state. The ionic radii were estimated at 110 pm for No and 90 pm for No .

Summary of compounds and (complex)


ions

Formula Names(s)

3+ hexaaquanobelium(III)
[No(H O) ]
2 6
2+ hexaaquanobelium(II)
[No(H O) ]
2 6

Nucleosynthesis

By cold
fusion
208 48 256-x
Pb( Ca,xn) No (x=1,2,3,4)
This cold fusion reaction was first studied in 1979 at the FLNR. Further work in 1988 at the GSI measured EC and
254
SF branchings in No. In 1989, the FLNR used the reaction to measure SF decay characteristics for the two
254
isomers of No. The measurement of the 2n excitation function was reported in 2001 by Yuri Oganessian at the
FLNR.
255-252
Patin et al. at the LBNL reported in 2002 the synthesis of No in the 1-4n exit channels and measured further
decay data for these isotopes.
The reaction has recently been used at the Jyvaskylan Yliopisto Fysiikan Laitos (JYFL) using the RITU set-up to
254
study K-isomerism in No. The scientists were able to measure two K-isomers with half-lives of 275 ms and
198
- +
µs, respectively. They were assigned to 8 and 16 K-isomeric levels.
255-253
The reaction was used in 2004–5 at the FLNR to study the spectroscopy of No. The team were able to confirm
253
an isomeric level in No with a half-life of 43.5 µs.
208 44 252-x
Pb( Ca,xn) No (x=2)
250
This reaction was studied in 2003 at the FLNR in a study of the spectroscopy of No.
207 48 255-x
Pb( Ca,xn) No (x=2)
The measurement of the 2n excitation function for this reaction was reported in 2001 by Yuri Oganessian and
253
co-workers at the FLNR. The reaction was used in 2004-5 to study the spectroscopy of No.
206 48 254-x
Pb( Ca,xn) No (x=1,2,3,4)
The measurement of the 1-4n excitation functions for this reaction were reported in 2001 by Yuri Oganessian and
co-workers at the FLNR. The 2n channel was further studied by the GSI to provide a spectroscopic determination of
252 -
K-isomerism in No. A K-isomer with spin and parity 8 was detected with a half-life of 110
ms.
204 48 252-x
Pb( Ca,xn) No (x=2)
The measurement of the 2n excitation function for this reaction was reported in 2001 by Yuri Oganessian at the
250
FLNR. They reported a new isotope No with a half-life of 36µs. The reaction was used in 2003 to study the
250
spectroscopy of No.They were able to observe two spontaneous fission activities with half-lives of 5.6µs and 54µs
250 249 250
and assigned to No and No, respectively. The latter activity was later assigned to a K-isomer in No. The
250
reaction was reported in 2006 by Peterson et al. at the Argonne National Laboratory (ANL) in a study of SF in No.
250
They detected two activities with half-lives of 3.7 µs and 43 µs and both assigned to No, the latter associated with
a K-isomer.

By hot
fusion
232 26 258-x
Th( Mg,xn) No (x=4,5,6)
The cross sections for the 4-6n exit channels have been measured for this reaction at the
FLNR.
238 22 260-x
U( Ne,xn) No (x=4,5,6)
252 250
This reaction was first studied in 1964 at the FLNR. The team were able to detect decays from Fm and Fm.
252 256
The Fm activity was associated with an ~8 s half-life and assigned to 102 from the 4n channel, with a yield of
256
45 nb. They were also able to detect a 10 s spontaneous fission activity also tentatively assigned to 102. Further
250
work in 1966 on the reaction examined the detection of Fm decay using chemical separation and a parent activity
254
with a half-life of ~50 s was reported and correctly assigned to 102. They also detected a 10 s spontaneous fission
256
activity tentatively assigned to 102. The reaction was used in 1969 to study some initial chemistry of nobelium at
the FLNR. They determined eka-ytterbium properties, consistent with nobelium as the heavier homologue. In 1970,
256 256
they were able to study the SF properties of No. In 2002, Patin et al. reported the synthesis of No from the 4n
257
channel but were unable to detect No.
The cross section values for the 4-6n channels have also been studied at the
FLNR.
238 20 258-x
U( Ne,xn) No
This reaction was studied in 1964 at the FLNR. No spontaneous fission activities were observed.
236 22 258-x
U( Ne,xn) No (x=4,5,6)
The cross sections for the 4-6n exit channels have been measured for this reaction at the
FLNR.
235 22 257-x
U( Ne,xn) No
(x=5)
252
This reaction was studied in 1970 at the FLNR. It was used to study the SF decay properties of No.
233 22 255-x
U( Ne,xn) No
The synthesis of neutron deficient nobelium isotopes was studied in 1975 at the FLNR. In their experiments
250 [5]
they observed a 250 µs SF activity which they tentatively assigned to No in the 5n exit channel. Later
results have not been able to confirm this activity and it is currently unidentified.
242 18 260-x
Pu( O,xn) No (x=4?)
This reaction was studied in 1966 at the FLNR. The team identified an 8.2 s SF activity tentatively assigned to
256
102.
241 16 257-x
Pu( O,xn) No
This reaction was first studied in 1958 at the FLNR. The team measured ~8.8 MeV alpha particles with a half-life of
253,252,251
30 s and assigned to 102. A repeat in 1960 produced 8.9 MeV alpha particles with a half-life of 2–40 s and
253
assigned to 102 from the 4n channel. Confidence in these results was later diminished.
239 18 257-x
Pu( O,xn) No (x=5)
252
This reaction was studied in 1970 at the FLNR in an effort to study the SF decay properties of No.
239 16 255-x
Pu( O,xn) No
This reaction was first studied in 1958 at the FLNR. The team were able to measure ~8.8 MeV alpha particles with a
253,252,251
half-life of 30 s and assigned to 102. A repeat in 1960 was unsuccessful and it was concluded the first
results were probably associated with background effects.
243 15 258-x
Am( N,xn) No (x=4)
250
This reaction was studied in 1966 at the FLNR. The team were able to detect Fm using chemical techniques and
determined an associated half-lifesignificantly higher than the reported 3 s by Berkeley for the supposed parent
254
No. Further work later the same year measured 8.1 MeV alpha particles with a half-life of 30–40 s.
243 14 257-x
Am( N,xn) No
This reaction was studied in 1966 at the FLNR. They were unable to detect the 8.1 MeV alpha particles detected
when using a N-15 beam.
241 15 256-x
Am( N,xn) No (x=4)
252
The decay properties of No were examined in 1977 at Oak Ridge. The team calculated a half-life of 2.3 s and
measured a 27% SF branching.
248 18 262-x
Cm( O,αxn) No (x=3)
259
The synthesis of the new isotope No was reported in 1973 from the LBNL using this reaction.
248 13 261-x
Cm( C,xn) No (x=3?,4,5)
258 257 256
This reaction was first studied in 1967 at the LBNL. The new isotopes No, No and No were detected in the
257
3-5n channels. The reaction was repeated in 1970 to provide further decay data for No.
248 12 260-x
Cm( C,xn) No (4,5?)
This reaction was studied in 1967 at the LBNL in their seminal study of nobelium isotopes. The reaction was used in
256
1990 at the LBNL to study the SF of No.
246 13 259-x
Cm( C,xn) No (4?,5?)
This reaction was studied in 1967 at the LBNL in their seminal study of nobelium
isotopes.
246 12 258-x
Cm( C,xn) No (4,5)
246
This reaction was studied in 1958 by scientists at the LBNL using a 5% Cm curium target. They were able to
250 254
measure 7.43 MeV decays from Fm, associated with a 3 s No parent activity, resulting from the 4n channel.
252 244
The 3 s activity was later reassigned to No, resulting from reaction with the predominant Cm component in the
250m
target. It could however not be proved that it was not due to the contaminant Fm, unknown at the time. Later
work in 1959 produced 8.3 MeV alpha particles with a half-life of 3 s and a 30% SF branch. This was initially
254 252 244
assigned to No and later reassigned to No, resulting from reaction with the Cm component in the target. The
254 253
reaction was restudied in 1967 and activities assigned to No and No were detected.
244 13 257-x
Cm( C,xn) No (x=4)
This reaction was first studied in 1957 at the Nobel Institute in Stockholm. The scientists detected 8.5 MeV alpha
251 253
particles with a half-life of 10 minutes. The activity was assigned to No or No. The results were later dismissed
as background. The reaction was repeated by scientists at the LBNL in 1958 but they were unable to confirm the 8.5
253
MeV alpha particles. The reaction was further studied in 1967 at the LBNL and an activity assigned to No was
measured.
244 12 256-x
Cm( C,xn) No(x=4,5)
244
This reaction was studied in 1958 by scientists at the LBNL using a 95% Cm curium target. They were able to
250 254
measure 7.43 MeV decays from Fm, associated with a 3 s No parent activity, resulting from the reaction
246 252 244
( Cm,4n). The 3 s activity was later reassigned to No, resulting from reaction ( Cm,4n). It could however not
250m
be proved that it was not due to the contaminant Fm, unknown at the time. Later work in 1959 produced 8.3
254
MeV alpha particles with a half-life of 3 s and a 30% SF branch. This was initially assigned to No and later
252 244
reassigned to No, resulting from reaction with the Cm component in the target. The reaction was restudied in
251
1967 at the LBNL and a new activity assigned to No was measured.
252 12 260-x
Cf( C,αxn) No
(x=3?)
This reaction was studied at the LBNL in 1961 as part of their search for element 104. They detected 8.2 MeV alpha
particles with a half-life of 15 s. This activity was assigned to a Z=102 isotope. Later work suggests an assignment to
257 252
No, resulting most likely from the α3n channel with the Cf component of the californium target.
252 11 262-x
Cf( B,pxn) No
(x=5?)
This reaction was studied at the LBNL in 1961 as part of their search for element 103. They detected 8.2 MeV alpha
particles with a half-life of 15 s. This activity was assigned to a Z=102 isotope. Later work suggests an assignment to
257 252
No, resulting most likely from the p5n channel with the Cf component of the californium target.
249 12 257-x
Cf( C,αxn) No
(x=2)
This reaction was studied in the early 1970s at the Oak Ridge Laboratory. Scientists were able to measure coincident
255
Z=100 K X-rays from No, confirming the discovery of the element.

As decay products
Isotopes of nobelium have also been identified in the decay of heavier elements. Observations to date are
summarised in the table below:

Evaporation Residue Observed No isotope

262 262
Lr No

269 265 261 257


Hs, Sg, Rf No

267 263 259 255


Hs, Sg, Rf No

254 254
Lr No

261 257 253


Sg, Rf No

264 260 256 252


Hs, Sg, Rf No

255 251
Rf No

Isotopes
Main article: Isotopes of
nobelium
259
Twelve radioisotopes of nobelium have been characterized, with the most stable being No with a half-life of 58
261 263
minutes. Longer half-lives are expected for the as-yet-unknown No and No. An isomeric level has been found
253 250 252 254
in No and K-isomers have been found in No, No and No to date.
Chronology of isotope
discovery
Isotope Year discovered Discovery reaction

250 m 2001 204 48


No Pb( Ca,2n)

250 g 2006 204 48


No Pb( Ca,2n)

251 1967 244 12


No Cm( C,5n)

252 g 1959 244 12


No Cm( C,4n)

252 m ~2002 206 48


No Pb( Ca,2n)

253 g 1967 242 16 239 18


No Pu( O,5n), Pu( O,4n)

253 m 1971 249 12


No Cf( C,4n)

254 1966 243 15


Nog Am( N,4n)

254 1967? 246 13 246 12


No m1
Cm( C,5n), Cm( C,4n)

254 ~2003 208 48


No m2
Pb( Ca,2n)

255 1967 246 13 248 12


No Cm( C,4n), Cm( C,5n)

256 1967 248 12 248 13


No Cm( C,4n), Cm( C,5n)

257 1961?, 1967 248 13


No Cm( C,4n)

258 1967 248 13


No Cm( C,3n)

259 1973 248 18


No Cm( O,α3n)

260 ? 254 22 18 13
No Es + Ne, O, C – transfer

261 unknown
No

262 1988 254 22 262


No Es + Ne – transfer (EC of Lr)

Nuclear
isomerism
254
No
The study of K-isomerism was recently studied by physicists at the University of Jyväskylä physics laboratory
(JYFL). They were able to confirm a previously reported K-isomer and detected a second K-isomer. They
- +
assigned spins and parities of 8 and 16 to the two K-isomers.
253
No
In the early 1970s, Bemis et al. determined an isomeric level decaying with a half-life of 31 µs from the decay of
257 257
Rf. This was confirmed in 2003 at the GSI by also studying the decay of Rf. Further support in the same year
from the FLNR appeared with a slightly higher half-life of 43.5 µs, decaying by M2 gamma emission to the ground
state.
252
No
In a recent study by the GSI into K-isomerism in even-even isotopes, a K-isomer with a half-life of 110 ms was
252 -
detected for No. A spin and parity of 8 was assigned to the isomer.
250
No
249
In 2003, scientists at the FLNR reported that they had been able to synthesise No which decayed by SF with a
half-life of 54µs. Further work in 2006 by scientists at the ANL showed that the activity was actually due to a
250
K-isomer in No. The ground state isomer was also detected with a very short half-life of 3.7 µs.

Retracted isotopes
In 2003, scientists at the FLNR claimed to have discovered the lightest known isotope of nobelium. However,
250 249
subsequent work showed that the 54 µs activity was actually due to No and the isotope No was retracted.

References
[1] J.A. Dean (ed), Lange's Handbook of Chemistry (15th Edition), McGraw-Hill, 1999; Section 4; Table 4.5, Electronegativities of the
Elements. [2] http:// en. wikipedia.org/w/index.php?title=Template:Infobox_nobelium&action=edit
[3] Silva, p. 1637.
[4] Silva, p. 1640.
[5] Silva, p. 1638.

Bibliography
• Silva, Robert J. (2011). "Chapter 13. Fermium, Mendelevium, Nobelium, and Lawrencium". In Morss, Lester R.;
Edelstein, Norman M. and Fuger, Jean. The Chemistry of the Actinide and Transactinide Elements. Netherlands:
Springer. doi: 10.1007/978-94-007-0211-0_13 (http://dx.doi.org/10.1007/978-94-007-0211-0_13).
ISBN 978-94-007-0210-3.

External links
• Chart of Nuclides (http://www.nndc.bnl.gov/chart/). nndc.bnl.gov
• Los Alamos National Laboratory – Nobelium (http://periodic.lanl.gov/102.shtml)
• Nobelium (http://www.periodicvideos.com/videos/102.htm) at The Periodic Table of Videos (University
of Nottingham)
103 Lawrencium 1

103 – Lawrencium
Lawrencium
Lr
103

nobelium ← lawrencium → rutherfordiumLu



Lr

(Upp)

Lawrencium in the periodic table


Appearance

unknown

General properties

Name, symbol, number lawrencium, Lr, 103

Pronunciation i
/ləˈrɛnsiəm/
lə-REN-see-əm

Element category actinide


sometimes considered a transition metal

Group, period, block n/a, 7, d

Standard atomic weight [262]

Electron configuration 2 14 1
[Rn] 7s 5f 7p
2, 8, 18, 32, 32, 8, 3

Physical properties

Phase solid (predicted)

Melting point 1900 K, 1627 °C, 2961 (predicted) °F

Atomic properties

Oxidation states 3

Ionization energies −1
1st: 443.8 kJ·mol

−1
2nd: 1428.0 kJ·mol

−1
3rd: 2219.1 kJ·mol

Miscellanea
Crystal structure hexagonal close-packed (predicted)

CAS registry number 22537-19-5

History

Naming after Ernest Lawrence

Discovery Lawrence Berkeley National Laboratory (1961)

Most stable isotopes

Main article: Isotopes of lawrencium

iso NA half-life DM DE (MeV) DP

262 syn 3.6 h ε 262


Lr No

261 syn 44 min SF/ε?


Lr

260 syn 2.7 min α 8.04 256


Lr Md

259 syn 6.2 s 78% α 8.44 255


Lr Md
22% SF

256 syn 27 s α 8.62,8.52,8.32... 252


Lr Md

255 syn 21.5 s α 8.43,8.37 251


Lr Md

254 syn 13 s 78% α 8.46,8.41 250


Lr Md
22% ε 254
No
only isotopes with half-lives over 5 seconds are included here

• v
• t
[1]
• e

Lawrencium is a radioactive synthetic chemical element with the symbol Lr (formerly Lw) and atomic number 103.
In the periodic table of the elements, it is a period 7 d-block element and the last element of the actinide series.
Chemistry experiments have confirmed that lawrencium behaves as the heavier homologue to lutetium and is
chemically similar to other actinides.
Lawrencium was first synthesized by the nuclear-physics team led by Albert Ghiorso on February 14, 1961, at the
Lawrence Berkeley National Laboratory of the University of California. The first atoms of lawrencium were
produced by bombarding a three-milligram target consisting of three isotopes of the element californium with boron-
10 and boron-11 nuclei from the Heavy Ion Linear Accelerator. The team suggested the name lawrencium (after
[2]
Ernest Lawrence), and the symbol "Lw", but IUPAC changed the symbol to "Lr" in 1963. It was the final and
the heaviest element of the actinide series to be synthesized.
All isotopes of lawrencium are radioactive; its most stable known isotope is lawrencium-262, with a half-life of
approximately 3.6 hours. All its isotopes except for lawrencium-260, -261 and -262 decay with a half-life of less
than a minute.
History

Discovery
An attempt to synthesize lawrencium was performed in 1958 by the Lawrence Berkeley Laboratory. This attempt
244 14
involved bombarding Cm with N. A follow-up on this experiment was not performed, as the target was
destroyed. Later, in 1960, the Lawrence Berkeley Laboratory attempted to synthesize the element by bombarding
252 10 11
Cf with B and B. The results of this experiment were not conclusive.
Lawrencium was first synthesized by the nuclear-physics team of Albert Ghiorso, Torbjørn Sikkeland, Almon Larsh,
Robert M. Latimer, and their co-workers on February 14, 1961, at the Lawrence Radiation Laboratory (now called
the Lawrence Berkeley National Laboratory) at the University of California. The first atoms of lawrencium were
produced by bombarding a three-milligram target consisting of three isotopes of the element californium with boron-
10 and boron-11 nuclei from the Heavy Ion Linear Accelerator (HILAC). The Berkeley team reported that the
257
isotope Lr was detected in this manner, and that it decayed by emitting an 8.6 MeV alpha particle with a half-life
258
of about eight seconds. This identification was later corrected to be Lr.
252
98Cf + 11
5B → 263–x
103Lr → 258
103Lr + 5 1
0n

In 1967, nuclear-physics researchers in Dubna, Russia, reported that they were not able to confirm assignment of an
257 258
alpha emitter with a half-life of eight seconds to Lr. This isotope was later deduced to be Lr. Instead, the
256
Dubna team reported an isotope with a half-life of about 45 seconds as Lr.
243
95Am + 18
8O → 261–x
103Lr → 256
103Lr + 5 1
0n
243 18
The Russian Joint Institute for Nuclear Research created lawrencium-256 in 1965 by bombarding Am with O to
form lawrencium-256. The scientists at the Joint Institute for Nuclear Research also discovered that the Lawrence
Berkeley Laboratory's claims for the isotope they synthesized, its energy, and its half-life were mistaken. The
Russians proposed the name "rutherfordium" for the new element in 1967. They also changed their data to fit the
claims made by the Joint Institute for Nuclear Research.
Further experiments have demonstrated an actinide chemistry for the new element, so by 1970 it was known that
lawrencium is the last actinide. In 1971, the nuclear physics team at the University of California at Berkeley
successfully performed a whole series of experiments aimed at measuring the nuclear decay properties of the
lawrencium isotopes with mass numbers from 255 through 260.
In 1971, the IUPAC granted the discovery of lawrencium to the Lawrence Berkeley Laboratory, even though they
did not have ideal data for the element's existence. However, in 1992, the IUPAC Trans-fermium Working Group
(TWG) officially recognized the nuclear physics teams at Dubna and Berkeley as the co-discoverers of lawrencium.
Because the name "lawrencium" had been in use for a long time by this point, it was made official by the Trans-
fermium Working Group.
Naming
The origin of the name, ratified by the American Chemical Society, is in reference to the nuclear-physicist Ernest O.
Lawrence, of the University of California, who invented the cyclotron particle accelerator. The symbol Lw was used
originally, but the element was assigned the Lr symbol. In August 1997, the International Union of Pure and Applied
Chemistry (IUPAC) ratified the name lawrencium and the symbol Lr during a meeting in Geneva.

Characteristics

Electronic structure
Lawrencium is element 103 in the periodic table. It is the first member of the 6d-block; in accordance with the
2 14 1
Madelung rule, its electronic configuration should be [Rn]7s 5f 6d . However, results from quantum mechanical
2 14 1
research have suggested that this configuration is incorrect, and is in fact [Rn]7s 5f 7p . A direct measurement of
this is not possible. Though early calculations gave conflicting results, more recent studies and calculations confirm
the suggestion.
A strict correlation between the periodic table blocks and the orbital-shell configurations for neutral atoms would
classify lawrencium as a transition metal because it could be classed as a d-block element. However, lawrencium is
[3]
classified as an actinide element according to the IUPAC recommendations.
Lawrencium is expected to be a solid under normal conditions and assume a hexagonal close-packed crystal
c
structure ( / = 1.58), similar to its lighter congener lutetium, though this is not yet known experimentally.
a

Experimental chemical properties

Summary of all lawrencium compounds known

Formula Names(s)
LrCl lawrencium trichloride ; lawrencium(III) chloride
3

Gaseous
phase
The first gaseous-phase studies of lawrencium were reported in 1969 by a nuclear physics team at the Flerov
243 18
Laboratory of Nuclear Reactions (FLNR) in the Soviet Union. They used the nuclear reaction Am+ O to produce
lawrencium nuclei, which they then exposed to a stream of chlorine gas, and a volatile chloride product was formed.
256
This product was deduced to be LrCl , and this confirmed that lawrencium is a typical actinide element.
3

Aqueous phase
The first aqueous-phase studies of lawrencium were reported in 1970 by a nuclear physics team at the
249 11
Lawrence Berkeley National Laboratory in California. This team used the nuclear reaction Cf+ B to produce
lawrencium nuclei. They were able to show that lawrencium forms a trivalent ion, similar to those of the other
actinide elements, but in contrast with that of nobelium. Further experiments in 1988 confirmed the
formation of a trivalent lawrencium(III) ion using anion-exchange chromatography using α-
hydroxyisobutyrate (α-HIB) complex. Comparison of the elution time with other actinides allowed a
3+ [4]
determination of 88.6 ± 0.3 picometers for the ionic radius for Lr . Attempts to reduce lawrencium in the
lawrencium(III) ionization state to lawrencium(I) using the potent reducing agent hydroxylamine hydrochloride
[5]
were unsuccessful.
Nucleosynthesis

Fusion
205 50 255-x
Tl( Ti,xn) Lr (x=2?)
This reaction was studied in a series of experiments in 1976 by Yuri Oganessian and his team at the FLNR. Evidence
253
was provided for the formation of Lr in the 2n exit channel.
203 50 253-x
Tl( Ti,xn) Lr
This reaction was studied in a series of experiments in 1976 by Yuri Oganessian and his team at the FLNR.
208 48 255-x
Pb( Ti,pxn) Lr
(x=1?)
246
This reaction was reported in 1984 by Yuri Oganessian at the FLNR. The team was able to detect decays of Cf, a
254
descendant of Lr.
208 45 253-x
Pb( Sc,xn) Lr
This reaction was studied in a series of experiments in 1976 by Yuri Oganessian and his team at the FLNR. Results
are not readily available.
209 48 257-x
Bi( Ca,xn) Lr
(x=2)
255
This reaction has been used to study the spectroscopic properties of Lr. The team at GANIL used the reaction in
2003 and the team at the FLNR used it between 2004 and 2006 to provide further information for the decay scheme
255 255
of Lr. The work provided evidence for an isomeric level in Lr.

Hot fusion
243 18 261-x
Am( O,xn) Lr (x=5)
This reaction was first studied in 1965 by the team at the FLNR. They were able to detect activity with a
256 257
characteristic decay of 45 seconds, which was assigned to Lr or Lr. Later work suggests an assignment to
256
Lr. Further studies in 1968 produced an 8.35–8.60 MeV alpha activity with a half-life of 35 seconds. This activity
256 257 256
was also initially assigned to Lr or Lr and later to solely Lr.
243 16 259-x
Am( O,xn) Lr (x=4)
This reaction was studied in 1970 by the team at the FLNR. They were able to detect an 8.37 MeV alpha activity
255
with a half-life of 22s. This was assigned to Lr.
248 15 263-x
Cm( N,xn) Lr
(x=3,4,5)
This reaction was studied in 1971 by the team at the LBNL in their large study of lawrencium isotopes. They were
260 259 258
able to assign alpha activities to Lr, Lr and Lr from the 3-5n exit channels.
248 18 265-x
Cm( O,pxn) Lr (x=3,4)
262 261
This reaction was studied in 1988 at the LBNL in order to assess the possibility of producing Lr and Lr
254
without using the exotic Es target. It was also used to attempt to measure an electron capture (EC) branch in
261m
Rf from the 5n exit channel. After extraction of the Lr(III) component, they were able to measure the
261
spontaneous fission of Lr with an improved half-life of 44 minutes. The production cross-section was 700
pb. On this basis, a 14% electron capture branch was calculated if this isotope was produced via the 5n
channel rather than the p4n channel. A lower bombarding energy (93 MeV c.f. 97 MeV) was then used to
262
measure the production of Lr in the p3n channel. The isotope was successfully detected and a yield of 240
pb was measured. The yield was lower than expected compared to the p4n channel. However, the results were
261
judged to indicate that the Lr was most likely produced by a p3n channel and an upper limit of 14% for the
261m
electron capture branch of Rf was therefore suggested.
246 14 260-x
Cm( N,xn) Lr
(x=3?)
103 Lawrencium 200

244 246
This reaction was studied briefly in 1958 at the LBNL using an enriched Cm target (5% Cm). They observed a
257
~9 MeV alpha activity with a half-life of ~0.25 seconds. Later results suggest a tentative assignment to Lr from
the 3n channel
244 14 258-x
Cm( N,xn) Lr
244 246
This reaction was studied briefly in 1958 at the LBNL using an enriched Cm target (5% Cm). They observed a
257
~9 MeV alpha activity with a half-life of ~0.25s. Later results suggest a tentative assignment to Lr from the 3n
246 244
channel with the Cm component. No activities assigned to reaction with the Cm component have been
reported.
249 18 263-x
Bk( O,αxn) Lr (x=3)
This reaction was studied in 1971 by the team at the LBNL in their large study of lawrencium isotopes. They were
260
able to detect an activity assigned to Lr. The reaction was repeated in 1988 to study the aqueous chemistry of
260
lawrencium. A total of 23 alpha decays were measured for Lr, with a mean energy of 8.03 MeV and an improved
half-life of 2.7 minutes. The calculated cross-section was 8.7 nb.
252 11 263-x
Cf( B,xn) Lr (x=5,7??)
This reaction was first studied in 1961 at the University of California by Albert Ghiorso by using a californium target
252
(52% Cf). They observed three alpha activities of 8.6, 8.4 and 8.2 MeV, with half-lives of about 8 and 15 seconds,
257 258
respectively. The 8.6 MeV activity was tentatively assigned to Lr. Later results suggest a reassignment to Lr,
257
resulting from the 5n exit channel. The 8.4 MeV activity was also assigned to Lr. Later results suggest a
256 250
reassignment to Lr. This is most likely from the 33% Cf component in the target rather than from the 7n
channel. The 8.2 MeV was subsequently associated with nobelium.
252 10 262-x
Cf( B,xn) Lr (x=4,6)
This reaction was first studied in 1961 at the University of California by Albert Ghiorso by using a californium target
252
(52% Cf). They observed three alpha activities of 8.6, 8.4 and 8.2 MeV, with half-lives of about 8 and 15 seconds,
257 258
respectively. The 8.6 MeV activity was tentatively assigned to Lr. Later results suggest a reassignment to Lr.
257 256
The 8.4 MeV activity was also assigned to Lr. Later results suggest a reassignment to Lr. The 8.2 MeV was
subsequently associated with nobelium.
250 15 260-x
Cf( N,αxn) Lr (x=3)
This reaction was studied in 1971 at the LBNL. They were able to identify a 0.7 s alpha activity with two alpha lines
257
at 8.87 and 8.82 MeV. This was assigned to Lr.
249 11 260-x
Cf( B,xn) Lr (x=4)
This reaction was first studied in 1970 at the LBNL in an attempt to study the aqueous chemistry of lawrencium.
3+
They were able to measure a Lr activity. The reaction was repeated in 1976 at Oak Ridge and the 26s lifetime of
256
Lr was confirmed by measurement of coincident X-rays.
249 12 260-x
Cf( C,pxn) Lr (x=2)
258
This reaction was studied in 1971 by the team at the LBNL. They were able to detect an activity assigned to Lr
from the p2n channel.
249 15 260-x
Cf( N,αxn) Lr (x=2,3)
258
This reaction was studied in 1971 by the team at the LBNL. They were able to detect an activities assigned to Lr
257
and Lr from the α2n and α3n and channels. The reaction was repeated in 1976 at Oak Ridge and the synthesis of
258
Lr was confirmed.
254 22
Es + Ne – transfer
This reaction was studied in 1987 at the LLNL. They were able to detect new spontaneous fission (SF) activities
261 262 22 254
assigned to Lr and Lr, resulting from transfer from the Ne nuclei to the Es target. In addition, a 5 ms SF
262
activity was detected in delayed coincidence with nobelium K-shell X-rays and was assigned to No, resulting
103 Lawrencium 1

262
from the electron capture of Lr.

Decay products
Isotopes of lawrencium have also been identified in the decay of heavier elements. Observations to date are
summarised in the table below:

List of lawrencium isotopes produced as other nuclei decay products


Parent nuclide Observed lawrencium isotope

267 263 259


Bh, Db Lr

278 274 270 266 262 258


Uut, Rg, Mt, Bh, Db Lr

261 257
Db Lr

272 268 264 260 256


Rg, Mt, Bh, Db Lr

259 255
Db Lr

266 262 258 254


Mt, Bh, Db Lr

261 257 g,m 253 g,m


Bh, Db Lr

260 256 252


Bh, Db Lr

Isotopes
Main article: Isotopes of lawrencium

Summary of all lawrencium isotopes known

Isotope Year discovered discovery reaction

252 2001 209 50


Lr Bi( Ti,3n)

253 g 1985 209 50


Lr Bi( Ti,2n)

253 m 2001 209 50


Lr Bi( Ti,2n)

254 1985 209 50


Lr Bi( Ti,n)

255 1970 243 16


Lr Am( O,4n)

256 1961? 1965? 1968? 1971 252 10


Lr Cf( B,6n)

257 1958? 1971 249 15


Lr Cf( N,α3n)

258 1961? 1971 249 15


Lr Cf( N,α2n)

259 1971 248 15


Lr Cm( N,4n)

260 1971 248 15


Lr Cm( N,3n)

261 1987 254 22


Lr Es + Ne

262 1987 254 22


Lr Es + Ne
262
Eleven isotopes of lawrencium plus one isomer have been synthesized with Lr being the longest-lived and the
252 [6]
heaviest, with a half-life of 216 minutes. Lr is the lightest isotope of lawrencium to be produced to date.

Nuclear isomerism
257
A study of the decay properties of Db (see dubnium) in 2001 by Hessberger et al. at the GSI provided some data
253 253
for the decay of Lr. Analysis of the data indicated the population of two isomeric levels in Lr from the decay of
257
the corresponding isomers in Db. The ground state was assigned spin and parity of 7/2-, decaying by emission of
an 8794 keV alpha particle with a half-life of 0.57s. The isomeric level was assigned spin and parity of 1/2-,
decaying by emission of an 8722 keV alpha particle with a half-life of 1.49 s.
255 209 48 255
Recent work on the spectroscopy of Lr formed in the reaction Bi( Ca,2n) Lr has provided evidence for an
isomeric level.

References
[1] http://en.wikipedia.org/w/index.php?title=Template:Infobox_lawrencium&action=edit
[2] Silva, p. 1641.
[3] Provisional Recommendations for the Nomenclature of Inorganic Chemistry (2004) (http://www.iupac. org/reports/ provisional/abstract04/
connelly_310804.html). IUPAC.org (Broken link: please see this archived version (http://web.archive.org/web/20080219122415/http://
www.iupac. org/reports/ provisional/abstract04/connelly_310804.html) instead).
[4] Silva, p. 1645.
[5] Silva, p. 1646.
[6] Silva, p. 1642.

Bibliography
• Silva, Robert J. (2011). "Chapter 13. Fermium, Mendelevium, Nobelium, and Lawrencium". In Morss, Lester R.;
Edelstein, Norman M. and Fuger, Jean. The Chemistry of the Actinide and Transactinide Elements. Netherlands:
Springer. doi: 10.1007/978-94-007-0211-0_13 (http://dx.doi.org/10.1007/978-94-007-0211-0_13).
ISBN 978-94-007-0210-3.

External links
• Chart of Nuclides (http://www.nndc.bnl.gov/chart/). nndc.bnl.gov
• Los Alamos National Laboratory's Chemistry Division: Periodic Table – Lawrencium
(http://periodic.lanl.gov/ 103.shtml)
• Lawrencium (http://www.periodicvideos.com/videos/103.htm) at The Periodic Table of Videos (University
of Nottingham)
1

Miscellany

Actinides in the environment


Actinides in the environment refer to the sources, environmental behaviour and effects of actinides in Earth's
environment. Environmental radioactivity is not limited solely to actinides; non-actinides such as radon and radium
are of note.

Inhalation versus
ingestion
Generally, ingested insoluble actinide compounds such as high-fired uranium dioxide and mixed oxide (MOX) fuel
will pass through the digestive system with little effect since they cannot dissolve and be absorbed by the body.
Inhaled actinide compounds, however, will be more damaging as they remain in the lungs and irradiate the lung
tissue.
Ingested Low-fired oxides and soluble salts such as nitrate can be absorbed into the blood stream. If they are
inhaled then it is possible for the solid to dissolve and leave the lungs. Hence the dose to the lungs will be
lower for the soluble form.

Radon and radium in the environment


Radon and radium are not actinides—they are both radioactive daughters from the decay of uranium. Aspects of
their biology and environmental behaviour is discussed at radium in the environment.

Thorium in the environment


In India, a large amount of thorium ore can be found in the form of
monazite in placer deposits of the Western and Eastern coastal dune
sands, particularly in the Tamil Nadu coastal areas. The residents of
this area are exposed to a naturally occurring radiation dose ten times
higher than the worldwide average.

Occurrence
Thorium is found at low levels in most rocks and soils, where it is
about three times more abundant than uranium, and is about as Monazite, a rare-earth-and-thorium-phosphate
common as lead. Soil commonly contains an average of around 6 parts mineral is the primary source of the world's
[1] thorium
per million (ppm) of thorium. Thorium occurs in several minerals,
the most common being the rare earth-thorium-phosphate mineral,
232
monazite, which contains up to about 12% thorium oxide. There are substantial deposits in several countries. Th
decays very slowly (its half-life is about three times the age of the earth) but other thorium isotopes occur in the
232
thorium and uranium decay chains. Most of these are short-lived and hence much more radioactive than Th,
though on a mass basis they are negligible.
Actinides in the environment 1

Effects in
humans
Thorium has been linked to liver cancer. In the past thoria (thorium dioxide) was used as a contrast agent for medical
X-ray radiography but its use has been discontinued. It was sold under the name Thorotrast.

Uranium in the environment


Main article: Uranium in the environment
Further information: Uranium § Human exposure
Uranium is a natural metal which is widely found. It is present in almost all soils and it is more plentiful than
antimony, beryllium, cadmium, gold, mercury, silver, or tungsten and is about as abundant as arsenic or
molybdenum. Significant concentrations of uranium occur in some substances such as phosphate rock deposits, and
minerals such as lignite, and monazite sands in uranium-rich ores (it is recovered commercially from these sources).
Seawater contains about 3.3 parts per billion of uranium by weight as uranium(VI) forms soluble carbonate
complexes. The extraction of uranium from seawater has been considered as a means of obtaining the element.
Because of the very low specific activity of uranium the chemical effects of it upon living things can often outweigh
the effects of its radioactivity. Additional uranium has been added to the environment in some locations as a result of
the nuclear fuel cycle and the use of depleted uranium in munitions.

Neptunium in the environment


Like plutonium, neptunium has a high affinity for soil. However, it is relatively mobile over the long term, and
diffusion of neptunium-237 in groundwater is a major issue in designing a deep geological repository for permanent
237
storage of spent nuclear fuel. Np has a halflife of 2.144 million years, so it is a long-term problem; but its halflife
237
is still much shorter than those of uranium-238, uranium-235, or uranium-236, and Np therefore has higher
specific activity than those nuclides.

Plutonium in the environment


Main article: Plutonium in the environment

Sources
Plutonium in the environment has several sources. These
include:
• Atomic batteries
• In space
• In pacemakers
• Bomb detonations
• Bomb safety trials
• Nuclear accidents (such as
Chernobyl)
• Nuclear crime
• Nuclear fuel
cycle
Environmental chemistry
Plutonium, like other actinides, readily forms a plutonium dioxide (plutonyl) core (PuO ). In the environment, this
- 2 - - 2−
plutonyl core readily complexes with carbonate as well as other oxygen moieties (OH , NO , NO , and SO ) to
2 3 4
form charged complexes which can be readily mobile with low affinities to
soil.
• PuO CO 2−
2 3
• PuO2 (CO3)24−
• PuO (CO ) 6−
2 33
PuO formed from neutralizing highly acidic nitric acid solutions tends to form polymeric which is resistant to
2
PuO
2
complexation. Plutonium also readily shifts valences between the +3, +4, +5 and +6 states. It is common for some
fraction of plutonium in solution to exist in all of these states in
equilibrium.
Plutonium is known to bind to soil particles very strongly, see above for a X-ray spectroscopic study of plutonium in
soil and concrete. While caesium has very different chemistry to the actinides, it is well known that both caesium and
134
many of the actinides bind strongly to the minerals in soil. Hence it has been possible to use Cs labeled soil to
study the migration of Pu and Cs is soils. It has been shown that colloidal transport processes control the migration
of Cs (and will control the migration of Pu) in the soil at the Waste Isolation Pilot Plant.

Americium in the environment


Americium often enters landfills from discarded smoke detectors. The rules associated with the disposal of
smoke detectors are very relaxed in most municipalities. For instance in the UK it is permissible to dispose of
an americium containing smoke detector by placing it in the dustbin with normal household rubbish, but each
dustbin worth of rubbish is limited to only containing one smoke detector.
In France a truck transporting 900 smoke detectors had been reported to have caught fire, it is claimed that this led to
a release of americium into the environment. In the U.S., the "Radioactive Boy Scout" David Hahn was able to buy
thousands of smoke detectors at remainder prices and concentrate the americium from them.
There have been cases of humans being contaminated with americium, the worst case being that of Harold
McCluskey. It is interesting to note that Harold McCluskey did not die of cancer but of heart disease (which he
had before the accident). It is likely that the medical care which he was given saved his life; because of the
difference in the chemistry of americium (the +3 oxidation state is very stable) to plutonium (where the +4 state
[citation needed]
can form in the human body) the americium has very different biochemistry to plutonium.
The most common isotope americium-241 decays (halflife 431 years) to neptunium-237 which has a much longer
[citation needed]
halflife, so in the long term, the issues discussed above for neptunium apply.

References
B.E. Burakov, M.I Ojovan,W.E. Lee. Crystalline Materials for Actinide Immobilisation, Imperial College Press,
London, 198 pp. (2010). http://www.icpress.co. uk/engineering/ p652. html
[1] THORIUM (http:// www.atsdr. cdc.gov/tfacts147.pdf) Agency for Toxic Substances and Disease Registry. July 1999.

Further
reading
• Hala, Jiri, and James D. Navratil. Radioactivity, Ionizing Radiation and Nuclear Energy. Konvoj: Brno,
Czech Republic, 2003. ISBN 80-7302-053-X.

External links
• Royal Society for Chemistry (http://www.rsc.org/images/Livens_tcm18-47506.pdf) - Why do mechanisms
matter in radioactive waste management?
• Federation of American Scientists (http://www.fas.org/sgp/othergov/doe/lanl/pubs/00818041.pdf) -
Spectroscopies for Environmental Studies of Actinide Species

Major actinide
Major actinides is a term used in the nuclear power industry that refers to the plutonium and uranium present in
used nuclear fuel, as opposed to the minor actinides neptunium, americium, curium, berkelium, and californium.

Minor actinide
The minor actinides are the actinide
elements in used nuclear fuel other
than uranium and plutonium, which are
termed the major actinides. The minor
actinides include neptunium,
americium, curium, berkelium,
californium, einsteinium, and fermium.
The most important isotopes in spent
nuclear fuel are neptunium-237,
americium-241, americium-243,
curium-242 through -248, and
californium-249 through -252.

Plutonium and the minor actinides


will be responsible for the bulk of the
radiotoxicity and heat generation of
used nuclear fuel in the medium term
(300 to 20,000 years in the future).
There are no fission products with
halflife in this range.
238 244
The plutonium from a power reactor Transmutation flow between Pu and Cm in LWR.
tends to have a greater amount of Fission percentage is 100 minus shown percentages.
Total rate of transmutation varies greatly by nuclide.
Pu-241 than the plutonium generated 245 248
Cm– Cm are long-lived with negligible decay.
by the lower burnup operations
designed to create weapons-grade
plutonium. Because the reactor-grade plutonium contains so much Pu-241 the presence of americium-241 makes the
plutonium less suitable for making a nuclear weapon. The ingrowth of americium in plutonium is one of the methods
for identifying the origin of an unknown sample of plutonium and the time since it was last separated chemically
from the americium.

Americium is commonly used in industry as both an alpha particle and low photon energy gamma radiation
source. For instance it is used in many smoke detectors. Americium can be formed by neutron capture of Pu-
239 and Pu-240 forming Pu-241 which then decays by beta decay to Am-241. In general, as the energy of the
neutrons increases, the ratio of the fission cross section to the neutron capture cross section changes in favour
of fission. Hence if MOX is used in a thermal reactor such as a boiling water reactor (BWR) or pressurized
water reactor (PWR) then more americium can be expected in the used fuel than that from a fast neutron
reactor.
Minor actinide 1

Some of them have been found in fallout from bomb tests. See Actinides in the environment for details of the
actinides in the environment.

Transuranics in LWR spent fuel (burnup 55 GWd /T) and mean neutron consumption to
th
fission
Isotope Fraction D D D
LWR fast superthermal

Np-237 0.0539 1.12 -0.59 -0.46

Pu-238 0.0364 0.17 -1.36 -0.13

Pu-239 0.451 -0.67 -1.46 -1.07

Pu-240 0.206 0.44 -0.96 0.14

Pu-241 0.121 -0.56 -1.24 -0.86

Pu-242 0.0813 1.76 -0.44 1.12

Am-241 0.0242 1.12 -0.62 -0.54

Am-242m 0.000088 0.15 -1.36 -1.53

Am-243 0.0179 0.82 -0.60 0.21

Cm-243 0.00011 -1.90 -2.13 -1.63

Cm-244 0.00765 -0.15 -1.39 -0.48

Cm-245 0.000638 -1.48 -2.51 -1.37

Weighted sum -0.03 -1.16 -0.51

|+ Negative numbers mean net neutron producer


Article Sources and Contributors 1

Article Sources and Contributors


Actinide Source: http://en.wikipedia.org/w/index.php?oldid=604745607 Contributors: 28bytes, 4, A876, AXRL, Aelffin, Ahoerstemeier, Akadruid, Alexf, Andre Engels, Aniten21, Anoop.m,
Antandrus, BJ Axel, Badocter, Barticus88, Beetstra, Beland, Bgwhite, Bob0the0mighty, Burubuz, Camw, Can't sleep, clown will eat me, ChongDae, Chris the speller, ChrisGualtieri,
ChrisHodgesUK, Chrishmt0423, Circeus, Cjthellama, CommonsDelinker, Conversion script, Cyrius, David Gerard, Davidryan168, DePiep, Deglr6328, Den nis1607, Derek Ross, Dicklyon, Dmn,
Donner60, Double sharp, DragonflySixtyseven, DuncanHill, Ebe123, Eddideigel, Edgar181, Ehrenkater, Ekren, Ericoides, Etincelles, Excirial, Fartherred, Feline1, Femto, Finalius, Flying Jazz,
FourBlades, Fred Bauder, Fredrik, Furrykef, Gap9551, Gentgeen, Geraki, Glenn4pr, Hairy Dude, Hammer1980, Hans Dunkelberg, HarDNox, Hashar, Hede2000, Herbee, Hugger666,
Hydrargyrum, Icairns, Intrr, Itinerant1, JWB, JaGa, Jamespoke, Jasper Deng, Jet57, Jet66, Jhd, Joan-of-arc, Johnleemk, Jrockley, Julesd, Keenan Pepper, Koavf, Kvng, LOLthulu,
Lanthanum-138, LarryMorseDCOhio, Levil, Limulus, Little Mountain 5, Looxix, LorenzoB, Lstanley1979, Lucent, MWHAHAHARAWRRRRRR, MagnaGraecia, Malleus Fatuorum,
Mandolinface, MarsJenkar, Materialscientist, Mfearby, Myshare, Nasir8891, Nergaal, Nickkid5, Njaard, Ohconfucius, Ojovan, Old Moonraker, Oldmanyes, Ospalh, Petergans, Peyre,
Piledhigheranddeeper, Pjstewart, Polyamorph, Pstudier, Pyrotec, Remember, Rfc1394, Rich Farmbrough, Ricknightcrawler, Rifleman 82, Rjwilmsi, Rotten, Rursus, Salix alba, ScAvenger,
Shalom Yechiel, Sl, Some1 forget, Speciman00, SpookyMulder, Squids and Chips, Stephen C. Carlson, StephenBuxton, Stone, Supremevo01, SvenAERTS, Szaszicska, TAnthony, TarmoK,
Tarquin, TheVault, TrygveFlathen, Tygrrr, Ugog Nizdast, V1adis1av, Vicki Rosenzweig, Vsmith, Whoop whoop pull up, Wikipelli, Wimt, Wknight94, Woohookitty, Wwoods, Zereshk, Zfr,

Ztobor, आशशीष भटननागर, 154 anonymous edits


89 – Actinium Source: http://en.wikipedia.org/w/index.php?oldid=603480307 Contributors: 28bytes, 3.14159265358pi, 65.68.87.xxx, ABF, Abhishek Jacob, AbigailAbernathy,
Aceygg, Adabow, Ahoerstemeier, Alansohn, Amatulic, Andre Engels, Anypodetos, Arkuat, Astatine-210, AuroraReloaded, Beetstra, Benbest, Benjiboi, Bryan Derksen, Bubby Scubblyfupt,
CRGreathouse, Carnildo, Casliber, Catgut, Ch'marr, Chill doubt, ChrisCork, ChrisGualtieri, Conversion script, Cybercobra, Danny, Dave Foley, David Latapie, DePiep, Deepak harshal nagle,
Deflective, Deviator13, Donarreiskoffer, Double sharp, Dtgriscom, Dureo, Dwmyers, Edgar181, Emperorbma, FMSrtTM, Femto, Fibonacci, GeeIsUs., George Burgess, Glenn, Gmel,
Goldenphoenix2007, Graham87, Greatpatton, Gypsypkd, Gyrobo, Hallpriest9, HazyM, Headbomb, HenryLi, Icairns, Ideyal, Igodard, JWB, Jaan513, Jaeger Lotno, Jaraalbe, Jasper Deng,
Jauhienij, Jimp, Jorfer, Kalamkaar, Karelj, Karlhahn, Keilana, Kurykh, Kwamikagami, Lamro, LarryMorseDCOhio, Lethalgeek, LordLuneth, Magioladitis, Marc Venot, Matani2005,
Materialscientist, Mattman723, Mav, Mdf, Michael Devore, Micothemagic, Mike Dill, Mikespedia, Muke, Naddy, Nergaal, Nick Y., Nihiltres, No1lakersfan, Ohconfucius, Orzetto, Parcly Taxel,
Paris moo, PeepP, Piledhigheranddeeper, PlanetStar, Polonium, Poolkris, Poor Yorick, Ranveig, Redux, Remember, Reyk, Reza kalani, Rich Farmbrough, Rjwilmsi, Romanm, Rwflammang,
SamSim, Saperaud, Schneelocke, Sfgiants1995, Sinrh0816, Sl, Snickerdo, SoCalSuperEagle, SpeedyGonsales, Srich32977, Stifynsemons, Stone, Tagishsimon, Tbhotch, Tetracube, Thisisbossi,
Thumperward, Trevor MacInnis, Tsogo3, Urhixidur, Vanished user vjhsduheuiui4t5hjri, Vsmith, Vuerqex, Warut, WereSpielChequers, WikiPuppies, Yekrats, Ynhockey, Youssefsan, Yyy,
Zachlipton, 147 anonymous edits

90 – Thorium Source: http://en.wikipedia.org/w/index.php?oldid=606831290 Contributors: 2mcm, 65.68.87.xxx, 78.26, 83d40m, Aarchiba, Acalamari, Ace ofgabriel, Acid88,
Adapter, Ahoerstemeier, Ajithrshenoy, Akaluzniacki, Alansohn, Albin Norlin, Aleksander.adamowski, AlexW, Alterrabe, Anaxial, Ancient Ginkgo, Andage01, Andonic, Andres, Andrewa,
Anoop.m, Antandrus, Anythingyouwant, Archimerged, Arg, Arkuat, Arru, Ashyanbhog, Attinio, Audunv, Avegab, Awatral, Azivegu, Bank top, Barneyg, BartonM, Bayou Banjo, Beetstra,
BenAlbahari, Benbest, Benjah-bmm27, Bentogoa, Bmcgann, Bobo192, Bobrayner, Borislav Dopudja, BoulderDuck, Brandon5485, Brian Huffman, Brisvegas, Bryan Derksen, C1010, CAPS
LOCK, CJGB, CMG, Cadmium, Calabe1992, Calmer Waters, Cameronw22, Canthusus, Carnildo, Ceropegius, Charminarin, ChemNerd, Chill doubt, Chris 73, Chris Roy, Chris the speller, Chriscf,
Christine T, Chuckiesdad, Classaccount073, Clovis Sangrail, Cmacd123, CommonsDelinker, Conversion script, Courcelles, Cryptic, Crzrussian, Cwkmail, Cxbrx, Cybercobra, Cyberevil, Czkraal,
DMahalko, Danpat, Darrien, David Gerard, David Latapie, David R. Ingham, DavidMIA, DePiep, DeXXus, Deepak D'Souza, Deflective, Denisarona, Deogratias5, Deor, Der Golem, Derek
Andrews, Devout Atheist, Dewan357, Diwas, Donarreiskoffer, Double sharp, DrHonzik, Dwayne, Dwmyers, E090, Ederiel, Edgar181, Edgy01, El C, Elium2, Elmozo, Emperorbma, EoGuy,
Epbr123, Excirial, Facts707, Falcon8765, Femto, Finefellow, Fionaclee, Foo1942, FordGT90Concept, Fraggle81, Fresheneesz, Fundamental metric tensor, Furrykef, Fæ, G-Man, Gaius Cornelius,
Gary, GenQuest, GeorgeTSLC, Gizziiusa, GoingBatty, GoonerDP, Gordonmcdowell, Gracophilus, Graeme Bartlett, Graham87, Graibeard, Greatpatton, Grey Shadow, Gsmcolect, Gthb, Guthrie,
Gypsypkd,
Hak-kâ-ngìn, Hallpriest9, Hcantero, Hcobb, Headbomb, Helge Skjeveland, Hellbus, Hersfold, Hesacon, Hinakana, Hitssquad, Holy Ganga, Icairns, Ideyal, Idleguy, Inglok, InverseHypercube,
Itub, J.delanoy, JWB, Jakec, Janke, Jaraalbe, Jauhienij, Jayshuler, JdH, Jdurg, Jim62sch, Jimp, Joanjoc, John, JonathanCobb, Jonathanischoice, Jons63, Jorge Iani, Julian Mendez, JunCTionS,
KFSorensen, Kaihsu, Karelj, Kathzerion, Kayleigheliz, Kbrose, Keenan Pepper, Keine entschuldigung, Kelovy, King Zebu, King jakob c, Kinu, Kjramesh, Kmarinas86, Kniwor, Kolbasz,
KoshVorlon, Ktsquare, Kudz75, Kurykh, Kwamikagami, Kwilbur, LA2, Lambiam, Lamlott, Lamro, Lanthanum-138, LarryMorseDCOhio, LeBron Michael Bryant, Leftbrainstuff, Lfstevens,
Light show, Limulus, Lockesdonkey, LorenzoB, Lothar von Richthofen, M11101, MJBurrage, Macrakis, Madstat, Magioladitis, Malcolm Farmer, Malleus4, Marc Venot, Mark.murphy,
Materialscientist, Matt Gies, Mattisse, Maury Markowitz, Mav, McMillen1, Megan1967, Member, Michelet, Mindmatrix, MinorFixes, Mio222, Mion, MisterDub, Mithridates, Mkweise,
Mnyaseen, Mogism, Monster eagle, Moogsi, MooreJethro, Mortense, Mosa Dawas, Mr0t1633, Mrintel, N2e, NPguy, Naddy, Nailedtooth, Nakkiel, Neil916, Nergaal, Nevadaone,
Neverquick, Niceguyedc, Nick Y., Nicola.Manini, Nihiltres, Northumbrian, O.Koslowski, Oberono, OhanaUnited, Ohnoitsjamie, OlEnglish, Ondrejch, Onepebble, Pakaran, Paradigm,
Parcly Taxel, PeepDussy, PeepP, Peter Greenwell, Peter.thejackos, Peter8212991, PeterJeremy, Peyre, PhasmatisNox, Philip Trueman, PiersOffshore, Piperh, Pjedicke, PlanetStar, Plunkey,
Pmurfey, Polonium, Polyamorph, Poolkris, Portnadler, Prosopon, Pstudier, Q43, REQC, Racerx11, RamanVirk, Rbakker99, Rcnet, Reatlas, Redrose64, Remember, ResearchRave, RexNL, Reyk,
Rgoodermote, Rich Farmbrough, Richard W.M. Jones, Rickterp, Riick, Rjwilmsi, Rmhermen, Robert Merkel, Robert1947, Robertiki, Robertrade, Roentgenium111, Rogerzilla, Romanm,
Royk, Ryan Lonswell, SDC, SEWilco, Santăr, Saperaud, Schlice, Schneelocke, Scott Ritchie, Scs, Sfuerst, Sgokoluk, Shaddack, Shadowjams, Sholt, ShotmanMaslo, Shyam, Sigsmund,
Sillybilly, Sl, Sleigh, Slidersv, Smusser, Soliloquial, Spangineer, Srich32977, Srikeit, Sstidman, Stephenacooke, Stifynsemons, Stone, StringTheory11, Svante, Sveinbjornpalsson, TDGood,
Tabletop, Tagishsimon, Tempodivalse, Terrencemorgan, Terribleidea, Tetracube, Thayts, The Haunted Angel, The Thing That Should Not Be, TheCaroline!, TheDestitutionOfOrganizedReligion,
TheoClarke, Thingg, Thoreaulylazy, Thorml, Thumperward, Tide rolls, Tkreh, Tobias Hoevekamp, Toms2866, Trevorloflin, Triforce of Power, Trojancowboy, UnitedStatesian,
Unliquidated, Vanished User 1004, Vegaswikian, Velella, Versus22, Volland, Vsmith, Wachholder0, Walkerma, Warut, Wayne Slam, Wcwarren, WhiteDragon, Whkoh, Whosasking,
Wikieditor06, Wikitiki89, William Avery, Worstnightmare99, Wwoods, Wyn.junior, XP1, Yekrats, Yogirox234, YordanGeorgiev, Yyy, Zerpent, Zoomzoom316, 525 anonymous edits

91 – Protactinium Source: http://en.wikipedia.org/w/index.php?oldid=605171191 Contributors: 28bytes, 65.68.87.xxx, Ahoerstemeier, AlimanRuna, AmericanXplorer13,


Andres, Archimerged, ArglebargleIV, Arkuat, Armando-Martin, Bayou Banjo, Benbest, Berkay0652, Bkell, Bobblewik, Brammers, Bryan Derksen, Bushmillsmccallan, CalicoCatLover,
Captain-n00dle, Chrislk02, Christopher Parham, Circeus, Conversion script, Cwkmail, Cybercobra, David Gerard, David Latapie, DePiep, Deflective, DerHexer, Diwas, Dmn, Donarreiskoffer,
Double sharp, DragonflySixtyseven, Dthomsen8, Dwmyers, Edgar181, Egil, Emperorbma, Exert, Favonian, Femto, Fibonacci, Fresheneesz, Geo Swan, Graham87, Greatpatton, Gypsypkd, Hak-
kâ-ngìn, Hallpriest9, Hammer1980, Hede2000, Heron, Hu, Icairns, Igny, InfoHaunter, Itub, J.delanoy, JWB, Jake Spooky, Jaraalbe, Jasper Deng, Jauhienij, Joanjoc, John, Karelj, Keilana, Kelovy,
Kolbasz, Kukini, Kwamikagami, LA2, Lamro, Lanthanum-138, LarryMorseDCOhio, Leonard G., Lightscamera808, Lugia2453, M11101, MagnaGraecia, Mandarax, Marc Venot, Marek69,
Materialscientist, Mattman723, Mav, Mgiganteus1, Mgobrien14, Mike Rosoft, Mikespedia, Mimihitam, Monkey Bounce, Montgomery '39, Muqman 52, NAHID, Naddy, Nergaal, Nyttend, Olof,
PCock, Peyre, Polonium, Polyamorph, Poolkris, Pras, PurpleChez, RJHall, Remember, Res2216firestar, Reyk, Rich Farmbrough, Rjwilmsi, Romanm, Runt00001, Sam Hocevar, Santăr,
Saperaud, Schneelocke, Semperf, Shalom Yechiel, Shoy, Sl, Smartse, Some jerk on the Internet, Speciman00, Squids and Chips, Srich32977, Stifynsemons, Stone, StringTheory11, Svante,
TJ-Grite, Tagishsimon, Tetracube, Tex, The High Fin Sperm Whale, Thumperward, Torsmo, Trovatore, Ulm, Urhixidur, Usinsk, Venona, Vsmith, Vuerqex, Walkerma, Warut,
Wavelength, Wayne Slam, WhiteDragon, Widefox, Xeworlebi, Yartrebo, Yekrats, Yintan, Yyy, 161 anonymous edits

92 – Uranium Source: http://en.wikipedia.org/w/index.php?oldid=606802487 Contributors: .:Ajvol:., .:CoReHaCk:., 10strein, 13afuse, 144.132.70.xxx, 200.191.188.xxx, 2D,
3sXk3, 56, 8ty3hree, A little insignificant, A.Ou, ABF, ASDFGHJKL, AVand, Aaron of Mpls, Abitslow, Abopardikar, Abrech, Aces lead, Acroterion, Adashiel, AdjustShift, Adssadasdasdasdas,
Aff123a, Ahoerstemeier, Airbreather, Aitias, Alansohn, Alarbus, AlexiusHoratius, Alexparent, Alhutch, AlmostReadytoFly, Alpha Kim, Alphachimp, Altus Prosator, Amarkov, Amonkey
person123456, Amplitude101, Anaxial, Andesite, Andonic, Andrea105, Andres, Andrewa, Andrewkim99, Angela, Angusmclellan, Anlace, Anomalocaris, Anonymousous, Antandrus, Antatnsu,
Antonio Lopez, Anwar saadat, Apollo2011, Apotheosis247, Arakunem, Archimerged, Arenhaus, Arjun01, Arnaugir, ArnoldReinhold, Art10, Artaxiad, Arthena, Artichoker, Artivist, Ashley
Pomeroy, Aspects, Asshairbutt, Atarr, Atomicskier, Atoyotaisatoyota, Austinfidel, Avenged Eightfold, Axl, AySz88, B.d.mills, BCV, Babajobu, Badboies, Badkick, Barneca, Barneyg, Barticus88,
Baum, Bayou Banjo, Bbatsell, Bduke, Beagle17, Beast of traal, BeefRendang, Beetstra, BenB4, Benbest, Bender235, Benjah-bmm27, Benjiboi, Bhadani, BiT, Big Brother 1984, Bigbumjellybelly,
Bigmak, BillC, BinaryTed, Biochem67, BlaiseFEgan, Bluerasberry, Bmicomp, Bobo The Ninja, Bobo192, Bobtbuilder, Boing! said Zebedee, Boneyard90, Bongwarrior, BorgQueen, Borislav
Dopudja, Borne nocker, Br77rino, BradBeattie, Branden, BrianY, Brianga, Bridgecross, BrightStarSky, Brighterorange, Bryan Derksen, Bucephalus, Buck Mulligan, Burntsauce, Burtonpe, C'est
moi, CMBJ, Cabhan, Cacahueten, Cactus.man, Cadmium, Calabe1992, Calbaer, Caleson, Calvin 1998, CambridgeBayWeather, Can't sleep, clown will eat me, CanadianLinuxUser, Capricorn42,
Carbon-16, Carcharoth, CardinalDan, Carel.jonkhout, Carlosp420, Carlwillis, Cbdorsett, Cbrodersen, Ceaser, Celuca, Cenarium, CensoredScribe, Cflm001, Chachebo, Charles Matthews,
CharlotteWebb, CharonX, Chem-awb, ChemNerd, Cherhillsnow, Cherry1000, Chowbok, Chris 73, Chris Roy, ChrisCork, ChrisHodgesUK, ChrisO, Chtito, CiTrusD, Cicik, Cirus206,
Citizensunshine, Cmprince, Coffee, CoffeeAddictUK, Cologinal, Connorbourke, Conversion script, Corelachew, Corrigendas, Corvus cornix, Courcelles, Craig.lz, Crazymonkey1123, Crisco
1492, Cryptic C62, Crystal whacker, Crystallina, Ctbolt, Curps, Cybercobra, Cyberevil, Cyberpower678, CyclePat, DMacks, DRE, DV8 2XL, DVD R W, DVdm, Da J Myster, Dallas Noyes,
Damirgraffiti, Danny, Darrien, Darsie42, Darth Krayt, Darth Mike, Datrollmaster, Davewild, David Gerard, David Latapie, Daybike, Dbastro, DePiep, Deacon of Pndapetzim, Deanos, Decltype,
Deflective, Deglr6328, Dekisugi, Delldot, Deltasam, DemocraticLuntz, DerHexer, Dialectric, Diberri, Digitalmind, Dina, Discospinster, Diwas, Djcam, Djdrake, Dlohcierekim, Dlohcierekim's
sock, Doc0000, Donarreiskoffer, Double sharp, Dougher, Doulos Christos, Dr U, Dr Zak, DrGatsby1962, Dragonjk96, Drex, Drgnaw, Drmies, Dwmyers, Dzlinker, EJF, ESkog, Ebe123, Ec5618,
Echuck215, Eclipse777xp, Edgar181, EdgeOfEpsilon, Edivorce, Eeekster, Eequor, Efried2, Eggishorn, Egomaniac, Einsi22, El C, ElTyrant, Ellywa, Emilfarb, Emperorbma, Epbr123, Eregli bob,
Eric Kvaalen, Eric-Wester, Erik9, Erkcan, Eubulides, EugeneZelenko, EuroCarGT, Everyking, Excirial, Expatkiwi, Extransit, FF2010, Fairyfreak1423, Fangfufu, Farseer, Fastfission, Feezo,
Article Sources and Contributors 2
Feinoha, Femto, Feydey, Fhempen, Fibonacci, Fionaclee, FisherQueen, Fligher, Floatingloopingtrees, Flyer22, Fraggle81, Fresheneesz, Friedorange33, Friginator, Fritzpoll, Furrykef, Fuzheado,
Fvasconcellos, Fyyer, Gadig, Galloramenu, Gamer007, Gareth Wyn, Garrett.mitchener, Garyzx, Gatortpk, Gavino uno, Gbleem, Gegnome, Gekedo, Gene Nygaard, George The Dragon,
George08, GeorgeBarnick, GetLinkPrimitiveParams, Gfoley4, GhostPirate, Giftlite, Gil johnson, Gilliam, Gimboid13, Ginsengbomb, Ginsuloft, Gioto, Give Peace A Chance, Glane23,
Glenn, Gobonobo, Gogo Dodo, Gopher78, Gortu, Graeme Bartlett, Graeme11, Graham87, GreatWhiteNortherner, Greatrobo76, Greenhorn1, Greg L, Gsmaster, Gtstricky, Gulbenk, Gurch,
Guttyut, Gwernol, Gypsypkd, Hadal, Hak-kâ-ngìn, HalfShadow, Hanacy, Happysailor, Hardcorespeedbump, Hardmoose, Harryboyles, Hawkeye7, Hawkzn, HazyM, Hdt83, Headbomb, Hellbus,
Herbee, Hipocrite, Hotshot977, Hulten, Hunter524, Husond, Hut 8.5, I am One of Many, I suck balloons, IForgotToEatBreakFast, IHateXx-sleepy-star-xxShe'sABitch, IMiiTH, IW.HG,
Iacobus, Ic3lvi4g3, Icairns, Ieat3hchildren, Iloveiraq, Ilyanep, Imsocoolithink, Inglok, Inkan1969, Intelligentsium, Ionas, Isis Stafford, Itai, Itangalo, Ixfd64, J Di, J.delanoy, JDX, JJGD,
JSquish, JTN, JZNIOSH, Jackelfive, Jackmcbarn, Jacob1207, Jaraalbe, Jarodalien, Jasper Deng, Jav13rg, Jaxl, Jay32183, JayHenry, Jc3s5h, JerrySteal, Jetnerd, Jim Douglas, Jim1138,
Jim77742, Jimfbleak, Jimp, Jm727, Jmajeremy, Jni, Joefromrandb, John, John Harland, JohnOwens, JohnSRoberts99, Johnfos, JohnnyB256, JohnnyRush10, Jojit fb, Jonesey95, Joren, Joriki,
Jose77, Joseph Solis in Australia, Josh, Josh Cherry, JoshHolloway, Jossi, Joyous!, Jrcoote, Jredmond, Jsavit, Ju66l3r, Juliancolton, Juliawashere, Julietkiri, Junglecat, Junvfr, Jusdafax, Justin
Eiler, Jóna Þórunn, Kaihsu, Karlhahn, Karn, Katieh5584, Kbdank71, Kbk, Kbrose, Keilana, KeithB, Kelapstick, Kenfreeman12, Kerotan, Ketsuekigata, Kevmitch, Kgrr, Killiondude, Kilo-
Lima, Kimon, Kingaustin42, Kingsimz, Kinneyboy90, Kinu, Knettles1, KnightLago, KnowledgeOfSelf, Knowlesadam, Knulclunk, Kolbasz, Korath, Koyaanis Qatsi, Krispos42, Kstinch,
Ktlabe, Ktsquare, Kudz75, Kurykh, Kwamikagami, Kyp4fish, KyuubiSeal, L Kensington, L1A1 FAL, Lamro, LargeBacon, LarryMorseDCOhio, Lcolson, LeaveSleaves, Lee J Haywood,
Leithp, Leuko, Liface, Liftarn, Light current, Lightmouse, LilHelpa, Limulus, Little Mountain 5, LogicDictates, Loltowne, Louisrsnblm, Lugia2453, LuigiManiac, Luisivanmezaparrales,
Lukobe, Lumos3, Luna Santin, M Holland, M-le-mot-dit, MER-C, MGTom, MK8, MZMcBride, Mac, MadGuy7023, MagnaGraecia, Maharashtraexpress, Majorclanger, Mako098765,
Malaret, Malbi, Malcolm Farmer, Malo, Mani1, Marc Venot, Marek69, Mark Arsten, MarkForeman, MarkRose, MarlinMr, Marminnetje, Martin451, Martpol, Mass09, MastCell,
Masterjamie, Materialscientist, Matthew Yeager, Matthewali55, Matthuxtable, Mav, MaxMahem, Maxamegalon2000, Maximus Rex, Mboverload, Mdf, Mechoption, Meekywiki, Mejor Los
Indios, Melaen, Mentifisto, Mgway, Michaelo80, Micromaster,
MightyWarrior, Mike Rosoft, Mikespedia, Mikiemike, Mild Bill Hiccup, MindstormsKid, Minesweeper, Minitech.me, Miranda, MisfitToys, Mithaca, Mma whaley, Mo0, Mohawkjohn, Moiz789,
Mokopila, Monkeyman024, Moondyne, Moreschi, Mormegil, Mouse980, Mr Stephen, Mr. Wheely Guy, Mr.Yim, MrFish, Mstblue, Muon, Muqman 52, Murcielago, MusikAnimal,
Mwhockeydude, Mwtoews, Myanw, Myasuda, NAHID, NIMSoffice, Naddy, Nakkiel, Nakon, NameIsRon, Natgel, Naught101, NawlinWiki, Nedim Ardoğa, Neel2210, Neifion, NellieBly, Nemo
bis, NerdyNSK, Nergaal, Nestoristhebest, Neurolysis, NewEnglandYankee, NewebNL, Nfreader, Nick Connolly, Nigash nagul, Nihiltres, Nirmos, Njerseyguy, Nn123645, Nono64, Nopetro,
NormalGoddess, NormanEinstein, NotAnonymous0, NotWorkin, Novacio, Nrcprm2026, Nsaa, Nthiery, NuclearWarfare, Numbafieve, Nydas, Nyttend, Oldnoah, Oleg Alexandrov, Olin, OllieFury,
Omcnew, Omegatron, Omg boy, Omg farts are cool, Oncemarried, Onebravemonkey, Opelio, Oroso, Orzetto, Osip7315, Othercjosen1, Otto ter Haar, Overthinkingly, Oxymoron83, PAntoni,
PFHLai, PIrish, Pakaran, ParticleMan, Patrick, Perey, Perry Tyler, Persian Poet Gal, Pgk, Pharaoh of the Wizards, Phileas, Philip Trueman, PierreAbbat, Pinethicket, Pip2andahalf, Piperh,
PlanetStar, Plantsurfer, Plasma Twa 2, Plasticup, Plazak, Poindexter Propellerhead, Polonium, Polyamorph, Polyparadigm, Poolkris, Porsche997SBS, Potatoswatter, Prannoymathew, Pratyya
Ghosh, Prewrite1, Prodego, Proofreader77, Pseudoanonymous, Pstudier, Public Kanonkas, Puchiko, Pyro Peanut, Pyro-stuntman, Qoose, Quark1005, Queenmomcat, Quintote, R6144, RG2,
RJC, RJHall, RSido, RTC, RadTek, Raeky, RandomKidMan, RandomLittleHelper, Raul654, Readeraml86, Reaper779, Reatlas, RedWolf, Reflynn, Remember, Res2216firestar, Retired username,
Rettetast, RexNL, Reyk, Riana, Rich Farmbrough, Rich257, Richard Arthur Norton (1958- ), Richard0612, Rirving, Risk one, Riverpilot, Rjwilmsi, Rlevse, Rmhermen, Roadrunner, Robert Foley,
RobertG, Rock2e, RockMFR, Roentgenium111, Roger-lc, Romanm, Rorschach, Rottenpeach, RoyBoy, Rrburke, Rtcoles, Rupert105, Rursus, RussianDeathToUSA, Ryan vicks, RyanB88,
Ryansilke, Samuria9, Santăr, Saperaud, Sarepr91, Sarregouset, Satellizer, Savi89, SaxicolousOne, Sbharris, Scarian, Scavenger348, Scetoaux, Schwern, Scientizzle, Sciurinæ, Scorpion451,
ScottyBerg, Secretfomrula, Seddon, Semila76, Sengkang, SenorToenails, Sfuerst, Shaddack, Shadow Tower, Shadowjams, ShakingSpirit, Shanes, ShaunMacPherson, Shawn in Montreal,
Shinkolobwe, Shirik, Shirulashem, Shivait, Shmoop, ShotmanMaslo, Sikkema, SimonP, SineWave, Sintaku, Sir Vicious, Sirsanjuro, Skarebo, Skaterdude7732, Skysmith, Sl, Slaggart, Slash,
Snigbrook, Snowolf, Soakologist, Some standardized rigour, Someguy1221, Sora22cloud, Souperman, SpK, Sparrowman980, Spencer, SpikeToronto, Splargo, SpookyMulder, Sprocket 2008,
Squirepants101, Srgreene, Srich32977, Stan Ison, Starcare, Starranger00, Staxringold, Stemonitis, Stenchrsbanna, Steorra, Stephenacooke, Stephenb, SteveMcCluskey, Stifynsemons, Stone,
StringTheory11, Suruena, Svante, Svick, TDC, THEN WHO WAS PHONE?, THF, TUF-KAT, TYelliot, Taggard, Tagishsimon, Tango, Tariqabjotu, Tatrgel, Tawker, Taxman, Tay tay charlie
charlsie, Tbhotch, Tbonnie, Tcncv, Techauthor, Techman224, Technical Foul, Teentje, Teh C b0x, Teixant, Tempodivalse, Terrell Larson, Tetracube, Texashomeboy09, Texture, Tgundry, Tharikrish,
The High Fin Sperm Whale, The Rambling Man, The Thing That Should Not Be, TheEditor12942, TheGrappler, Theda, Themanfromtheplace, Theo F, Therealnotfunnyguy1, Thingg, Thomaslisi,
Thomyris, Thumperward, Thunderwing, Tide rolls, TimothyHorrigan, Titoxd, Tkircher, Tmobsterdon, Tnxman307, Toagac, Togo, Tohd8BohaithuGh1, Tom harrison, TommorowIsNotFriday,
Tommy2010, Tony Fox, Tow, Tpbradbury, Traxs7, Trelvis, Tresiden, Trevor MacInnis, Trojancowboy, Turgan, Tweenk, Ulf Hermjakob, Uli, Ultraman11, Umapathy, Uranium 4 Your Cranium,
Urhixidur, Username237, Utcursch, Vanished User 1004, Vanished user 39948282, Vargenau, Vector Potential, Vegaswikian, Versus22, Vianello, Vimalkalyan, Vina, Violetbonmua, Viriditas,
Vmelkon, Vrenator, Vsmith, Vyznev Xnebara, WJBscribe, Wabernat, Wachholder0, WadeSimMiser, Wadewitz, Wafulz, WahreJakob, Walkerma, Warut, Watch37264, Wavelength, Wayward,
Wbrameld, Wdanwatts, Webrats, Wereon, Weyes, WhatamIdoing, Whitlock, Wiccan Quagga, Wiki alf, Wiki ian, Wiki13, WikiTed 2000, Wikibofh, Wikidudeman, Wikimedes, Wikipelli,
WillKemp, William Avery, Wimt, Wimvandorst, Wisdom89, Witoman, Wknight94, Wormow, Woseph, Wowlookitsjoe, Wtmitchell, WulfTheSaxon, Wyn.junior, X201, Xiahou, Xxpor, Yann,
Yapete, Yekrats, Yellowdesk, Yggdræsil, Yonatan, Ytterbium2, Yyy, Z22, Zapvet, Zbeganovic, Zeno Panthakree, Zeszezu, Zscout370, Zzuuzz, Zzyzx11,

2082 ,‫ﻞﯧﺨﻤﻴﻫﺍﺮﺑﺍ‬-‫ﻲﻧﺎﺑﺎﻴﺑ‬-‫ﺐﻴﺠﻧ‬-‫ ﺪﻤﺣﺍ‬anonymous edits


93 – Neptunium Source: http://en.wikipedia.org/w/index.php?oldid=600560577 Contributors: 28bytes, Addshore, Ahoerstemeier, AlimanRuna, Andonic, Andres, Andrewa,
Anoop.m, Arkuat, Avm1, Barticus88, Bayou Banjo, Beetstra, Benbest, Bird, Borislav Dopudja, Breno, Bryan Derksen, Cadmium, Cheeseeeeeeee, ChemNerd, Cholmes75, Chris the speller,
Conversion script, Cryvksr, Cybercobra, Daggerstab, Darrien, David Latapie, DePiep, DeadEyeArrow, Deanos, Deconstructhis, Deflective, Discospinster, Diwas, DocWatson42, Donarreiskoffer,
Double sharp, DragonflySixtyseven, Edgar181, Edward J. Picardy, Egil, El C, Emperorbma, Falcanary, Farseer, Father McKenzie, Femto, Fennec, Furrykef, Fuzzform, Gene Nygaard, Gilliam,
Give Peace A Chance, Goldenfool, GraemeL, Grimwald13, Gypsypkd, Hak-kâ-ngìn, Hallpriest9, HarDNox, HazyM, Helge Skjeveland, Hqb, Icairns, Inter, Itinerant1, IvanLanin, JWB, JWBE,
Jakec, Jaraalbe, Jauhienij, Jhinman, Jiang, Jneg, Joanjoc, JoanneB, John, Jovianeye, Kazvorpal, Kbrose, Keegan, Kelovy, Kimse, King jakob c, Kobrabones, Kuratowski's Ghost, Kwamikagami,
LA2, Lamro, LarryMorseDCOhio, Limulus, Marc Venot, Materialscientist, Mav, Mboverload, Mdf, Morwen, Mr.98, Muqman 52, Myasuda, Naddy, Nedim Ardoğa, Neil916, Nergaal, NightBear,
Nolimits5017, Novangelis, Nunquam Dormio, Nyroc1234567892468, OlEnglish, PeepP, Piano non troppo, Pinethicket, Polonium, Polyamorph, Poolkris, Pras, PsychoCola, Queenmomcat, Raeky,
Razorflame, Reatlas, RexNL, Reza kalani, Riana, Rich Farmbrough, Rjwilmsi, Roberta F., Roentgenium111, Runningonbrains, Rursus, Rwendland, Rwflammang, Ryukichiro, Sanogo700, Santăr,
Saperaud, Schneelocke, Sengkang, Sfuerst, Shanedidona, SimonP, Skizzik, Sl, Sobieski Wanda, Specs112, Spiff, Squids and Chips, Srich32977, Stephenb, Stone, Stratocracy, Szaszicska,
Tagishsimon, Tetracube, Thingg, Thumperward, Tide rolls, Trogdor57, Uncle Dick, VASANTH S.N., Velella, Versus22, Vipulksoni, Virekleatherwong, Volland, Vsmith, Vuerqex, Warut, Whoop
whoop pull up, Yekrats, Yyy, Zalgo, 215 anonymous edits

94 – Plutonium Source: http://en.wikipedia.org/w/index.php?oldid=606399210 Contributors: .:CoReHaCk:., 130.94.122.xxx, 1stlegionarmy, 28bytes, 2D, 5 albert square, A. di
M., A876, ACSE, ARTE, AWeishaupt, Aarchiba, Abalacha, Abcdefgy2, Acroterion, Action Jackson IV, Adzze, Afernand74, Ahoerstemeier, Aitias, Alan Peakall, Alansohn, Alarbus, Algebraist,
AlimanRuna, Alistair1978, Allen Moore, Alpha Quadrant (alt), AnakngAraw, Anclation, Andres, Andrewa, Andrewman327, Andy M. Wang, Angelofdeath275, Ankitbhatt, Anonymous
Dissident, Antandrus, Antatnsu, Anthony Appleyard, Anyeverybody, April Arcus, Apyule, Ardonik, ArglebargleIV, Arkuat, Armistej, ArnoldReinhold, Art LaPella, ArthurBorges, Artichoker,
Aschwole, Ash, Asro940, Atif.t2, Attilios, Audin, Aunva6, Avenue, AwesomeEvilGenius, Axl, Bananaman2222, Barneyg, Basti1basti, Bayerischermann, Bayou Banjo, Bbartlog, Bcorr, Bdweiler,
BeefRendang, Beeline23, Beetstra, Benbest, Bender235, Benjiboi, BerserkerBen, Big cbutty, BigDukeSix, Bigmikeaziz, Bill-on-the-Hill, Black lupin, Bladiebla, Blethering Scot, Bobathon71,
Bobo192, Bongwarrior, Booboobear7, BoomerAB, Borgx, Borislav Dopudja, Boundarylayer, Brat32, Brianga, Brothejr, Bryan Derksen, Bth, Buijs, Bulbeck, C.Fred, CWenger, CWii, CaSJer,
Cadmium, Calmer Waters, Camembert, Can't sleep, clown will eat me, Cannolis, Carlsotr, Catbar, Ccmdav, Centrx, Cgingold, Chealer, ChemGardener, Cherkash, Chris the speller, Christhi,
Christian75, Chuck Carroll, Chuckwatson, Chuunen Baka, Citizensunshine, Clarkcj12, Clawson, Closedmouth, CommonsDelinker, Conversion script, Cosmic Latte, Courcelles, CredoFromStart,
Criticality, Crystal whacker, Cs302b, Cuhlik, Cybercobra, DARTH SIDIOUS 2, DV8 2XL, Dagonking123, Dajwilkinson, DaneCrazy, Daners, Darrien, Darth Krayt, David Latapie, David from
Downunder, Dawn Bard, Dcandeto, DePiep, DeadEyeArrow, Deaxman, Deflective, Defordj, Deglr6328, Deltabeignet, DexDor, Diagonalfish, Discospinster, Dispenser, Djmackie, Dkozza, Dnwq,
DocWatson42, Dolphin51, Domthedude001, Donarreiskoffer, Donreed, Double sharp, Dougofborg, Doulos Christos, Download, DrKiernan, DragonflySixtyseven, Dragons flight, Drj,
Dszerxftcghjk, Dtobin123, Durin, Dysmorodrepanis, E Wing, EHRice, Eatabug1234, Ec5618, EchetusXe, Edgar181, EdgeOfEpsilon, Edivorce, Edward J. Picardy, Einsteinmc2300, El C, El
Suizo, ElationAviation, Emperorbma, Endicott65, Eoghain0708, Epbr123, Eras-mus, ErkinBatu, Euku, Evil Monkey, Evildeathmath, Excirial, Eyreland, FF2010, FRS, Fan-1967, Fang Aili,
Farosdaughter, Farseer, Fasettle, Fastfission, FellGleaming, Femto, Finlay McWalter, FirstPrinciples, Flbribri8788, Forai, FormerNukeSubmariner, Frankenpuppy, Fred Bauder, Fredrik,
Fredrock800, Frehley, Frog is God, Froshirt, Fullerene, Furrykef, Fuzzform, GCarty, Gail, GainLine, Galoubet, Gbleem, Gcalis, Gderbysh, Geni, Geoffrey.landis, Gerhardvalentin, Getaway,
Ghewgill, Giftlite, Gigemag76, Gillis, Give Peace A Chance, Glane23, GoneAwayNowAndRetired, Gordonrox24, Graeme Bartlett, Graham87, GregorB, Ground Zero, Grunkhead, Guest9999,
Gzornenplatz, Gökhan, H2eddsf3, H2g2bob, HPaul, HVS, Hahnium, Hak-kâ-ngìn, HarDNox, Hawkeye7, HazyM, Headbomb, Heimstern, Heron, HexaChord, Hinto, Honeycake, Hosain54,
Howcheng, Howdyfolk, Hqb, Hugo999, Hurricane Devon, Husond, Hydrargyrum, I am Super Ryan, IIXII, IRP, Icairns, Igodard, Igoldste, Iliev, Infeh, Inglok, Inver471ness, Iridescent,
IronGargoyle, Itub, J.delanoy, JForget, JHunterJ, JLKrause, JWB, JWBE, JaGa, Jaerik, Jake Wartenberg, Janiberry, Jaraalbe, Jauerback, Jauhienij, Jay L09, Jeff G., Jeffq, Jhfjdhfjhsdfkd, Jiang,
Jmcc150, Joanjoc, Joe Friendly, Joema, Joesmith989, John, John254, JohnOwens, JohnSRoberts99, Johnfos, Joriki, Jossi, Julesd, KDS4444, Kalamkaar, Karlhahn, Karn, Kazvorpal, Kbrose,
Keegan, Keeplookin, Keeves, Kendrick7, Kenny1678, Kenyon, Kimchi.sg, Kimdino, Kirk Hilliard, Kjkolb, KnowledgeOfSelf, Kolbasz, Korderrius, Kouhoutek, Kurykh, Kwamikagami, LA2,
LOL, Lamro, Lawrence King, Lcolson, Leafyplant, LeaveSleaves, Lfh, LilHelpa, Limulus, Little green rosetta, LittleOldMe, Lomn, LonelyMarble, Lop7685, LovesMacs, Lucas0425, Ludde23,
Lugia2453, Luk, Lupo, MBK004, Madtrolls, Magioladitis, Magister Mathematicae, Mako098765, Man with two legs, Marc Venot, Marco Krohn, Marek69, Margareta, Marhault, Markjoseph125,
Martin Hogbin, Martylunsford, Matalovita, Materialscientist, Mav, Maxim, Mboverload, Mcshadypl, Me1423, MeegsC, Megansmith18, MelbourneStar, Melchoir, Mentifisto, Mgiganteus1,
Michael Devore, Midgley, Mike Rosoft, Milkbreath, Mirithing, Misiu mp, Miss Madeline, Misterballs1337, MistyMorn, Mkweise, Modemac, Mojo Hand, Moletrouser, Monkey Bounce, Mono,
Mr Stephen, Mr.98, Mr0t1633, MrWhipple, Muqman 52, MyFaceBeAFunnyFace, Myanw, Mygerardromance, NHRHS2010, Nabokov, Naddy, Nakon, NapoliRoma, Narson, NavyPunk426,
NawlinWiki, Nedim Ardoğa, Nelhowt4, Neoshero5, Nergaal, Nesstopher, Neurolysis, NeuronExMachina, Neverquick, NewbieDoo, NewebNL, Newyorkadam, Nihiltres, Ninjamarmot, Ninly,
Nitrobutane, Nosebud, NuclearWarfare, Ocdnctx, Old Moonraker, Omega 13, Omegatron, Omicronpersei8, Onebravemonkey, Oneiros, Open4D, Orenburg1, Orlady, Ouishoebean, Oxymoron83,
Pagw, Pajz, Pakaran, PaleCloudedWhite, Parcly Taxel, Patrick, Paul-L, Pediddle, Pen1234567, Petri Krohn, Pgk, Philip Trueman, Phixt, Phmer, Physchim62, Pi zero, Piano non troppo,
Piledhigheranddeeper, Pinethicket, Pinkadelica, Pinktulip, Piperh, PlanetStar, Plantsurfer, Plasticspork, Plutonium12345, Pol098, Polarscribe, Pollinator, Polonium, Polyamorph, Polymerbringer,
Poolkris, Positron, Pstanton, Pstudier, Psy guy, Pyfan, Quadalpha, Quartic, Quatro dose, Qwe, R8R Gtrs, RJHall, RTC, Railgun, Rama's Arrow, RandomP, RaptorHunter, Ratemonth, Ray Van De
Walker, Rccoms, Rcnet, Rdsmith4, Reaper Eternal, Reatlas, Redfarmer, Remember, RexNL, Reyk, Rich Farmbrough, Riffsyphon1024,
Rjwilmsi, Rmhermen, Roadrunner, Robert Foley, RobertG, Rockvee, Roentgenium111, Rogermw, Romanm, Ron Magic, Rreagan007, Ruby Murray, Ruffin' writer, Rune.welsh, Rursus,
Rwendland, Rwflammang, Ryanrs, SEWilco, Saikiri, Salamancer42, Sam Hocevar, Samcz, SandyGeorgia, Saperaud, Sasper, Savant13, Sbharris, SchfiftyThree, Schneelocke,
SchnitzelMannGreek, Science Focus, Sdaaw4g, Sean Goodwin, Sebastian barnes, Sengkang, Sfan00 IMG, Sfuerst, Shaddack, Shadowjams, Shanedidona, Shanes, Shenme, Shinkolobwe,
Shoemaker's Holiday, Shoy, Sia15998, Simesa, Skier lad, Skoch3, SkoreKeep, Sl, Sladen, Smalljim, Smsarmad, Snowolf, Snugggy, Soarhead77, Soralin, Speciman00, Spellmaster, Spencer,
Spidaman23, Spiff, SpookyMulder, Squids and Chips, Srich32977, Starkrm, StaticGull, Stebbins, Steel, Stellar-TO, StephenHart, Stevertigo, Stone, Stratocracy, Strike Eagle, StringTheory11,
Stux, Sugarbat, Super8Guy, Supremeknowledge, Svante, Svick, Sweet smell of salami, Syrthiss, THEJUDGE24, THEN WHO WAS PHONE?, Tagishsimon, Tanaats, Tbhotch, Tdent,
Techman224, TedE, Tedernst, Tempodivalse, TenOfAllTrades, TerraFrost, Tetracube, Texboy, Tgeairn, That Guy, From That Show!, The Cake is a Lie, The High Fin Sperm Whale, The Jacobin,
The Thing That Should Not Be, Thedigitalabe, Thedjatclubrock, Theseeker4, Thincat, Thingg, Thom.fynn, Thue, Thumperward, Tobias Bergemann, ToddFincannon, Tpth, Tresiden,
Trojancowboy, Trollepedia troll, Trusilver, Trần Nam Hạ 2001, Tstrobaugh, Twang, Tweenk, TwoWildnCrazyKids, Ugog Nizdast, Ulric1313, Ultramince, Uncle Dick, Ur mom sux, Uriber,
Valentinian, VampWillow, Vanischenu from public computers, Veemonkamiya, Versus22, Vsmith, WadeSimMiser, Walkinglikeahuricane, Wapcaplet, Warbeck, Warut, Watch37264, Wavelength,
Wdfarmer, Wearyalchemist21, WereSpielChequers, Wertuose, Wervo, WhickityWhite, Whiteskin420, Whitneyjones, Whoamiswagger, Whoop whoop pull up, Whosasking,
Whydidthatpagechange, Widr, Wiki alf, Wiki fanatic, WikiDao, Wikipelli, William Avery, Wimt, WolfmanSF, Woohookitty, Wuhwuzdat, Wwoods, Xiaphias, XinaNicole, Xiong Chiamiov,
Yalbik, Yann, Yath, Yaush, Yekrats, YouRang?, Yyy, Zenswashbuckler, Zidonuke, ZimZalaBim, Zizzybaluba, ZooFari, Zotel, சசசசச , 1212 anonymous edits

95 – Americium Source: http://en.wikipedia.org/w/index.php?oldid=606325390 Contributors: A. Carty, A876, A8UDI, AWeishaupt, Abeg92, Ahoerstemeier, AlimanRuna,
Andres, Angela, Avidallred, Ayengar, B, Baccyak4H, Bahahs, Ban Bridges, Banana04131, Bayou Banjo, Bcorr, Beetstra, Ben Ben, Beorhtwulf, Billingd, Blanchardb, Bobo The Ninja, Bogey97,
Bovineone, Bradbaka, Brianski, Bryan Derksen, Burtonpe, Cadmium, Camw, Canthusus, Capricorn42, CaptainVindaloo, ChemGardener, ChemNerd, Chemkid1, Cherkash, Chris the speller,
ChrisGualtieri, Chuunen Baka, Cockus, Coinmanj, CommonsDelinker, Conversion script, CorinneSD, Courcelles, Crazytales, Cybercobra, DV8 2XL, Dachshund, Darrien, David Latapie,
David.Mestel, Davidprior, Dbrodbeck, DePiep, Deflective, Delphwhite, Deor, Dispenser, DocWatson42, Donarreiskoffer, Double sharp, Drmies, Dspradau, Edgar181, Ekren, Emperorbma, Espi,
Farseer, Femto, Flehmen, Fluffernutter, ForestAngel, Forteblast, Frosty, Gibson43, Glc9144, Glenn, Gluck 123, GoingBatty, Goldenband, Graham87, Gtstricky, Gypsypkd, Hak-kâ-ngìn,
Hamiltonstone, HazyM, Headbomb, Hemmingsen, Herbee, Hhhippo, Hqb, I dream of horses, ICAPTCHA, Icairns, Ideyal, Inglok, Ishikawa Minoru, Itub, Iviney, J miester25, J.delanoy, JWB,
JWBE, Jamesilames, Jaraalbe, Jauhienij, Jiang, Jimfbleak, Joanjoc, JoaoRicardo, John, Julesd, Kalamkaar, Kay Dekker, Kbrose, Keilana, Kelisi, Kelovy, Koavf, Kolbasz, Kurykh, Kwamikagami,
Kwardle33, LA2, Lamro, LarryMorseDCOhio, Likeitsmyjob, Limulus, LindsayH, Lord Voldemort, Lottiotta, Luna Santin, Lysithix, Magnus Manske, Marc Venot, Martin451, Materialscientist,
Mav, Mayur, Mdf, Meaghan, Mgiganteus1, Mikenorton, Milkunderwood, Muqman 52, Murtasa, Naddy, Nedim Ardoğa, Nergaal, Nick Y., NickCT, Nihiltres, No1lakersfan, Nofutureuk, NotWith,
Omnipaedista, Ospalh, Panu, Parcly Taxel, Perlmonger42, Pharaoh of the Wizards, Polonium, Polyamorph, Poolkris, Potatoswatter, Pras, Pseudomonas, R'n'B, RA0808, Rcnet, Recognizance,
Redux, Remember, Reza kalani, Rifleman 82, Rjwilmsi, Robert K S, Roberta F., Robyrockets, Roentgenium111, Romanm, Samw, Saperaud, Sargentzan, Schneelocke, Semperf, Sengkang,
Shadeslayer036, Sheila Rogers, Sionus, Skepicalcynic, Sl, Smalleditor, SmartyPants321, Socrates2008, Some jerk on the Internet, Srich32977, Stifynsemons, Stone, Stratocracy, StringTheory11,
StuartH, Supremeknowledge, Surya Prakash.S.A., TJRC, Tagishsimon, Tetracube, The Firewall, TheWildColonialBoy, Thingg, Thumperward, Tide rolls, Tim Q. Wells, Titoxd, Tlusťa, Tmopkisn,
Trannylover3, Trappist the monk, Tsogo3, Ttiotsw, Ttony21, Tweenk, Uruiamme, Uwe W., VASANTH S.N., Vary, Vsmith, WODUP, WRK, Warut, We hope, WeniWidiWiki, Werieth, Whoop
whoop pull up, Wikipelli, William Avery, Wingman4l7, Wmahan, Woohookitty, Wrynne, Xenophon777, Xeworlebi, XinaNicole, Xxis, Ytrottier, Yyy, Zelmerszoetrop, 318 anonymous edits

96 – Curium Source: http://en.wikipedia.org/w/index.php?oldid=606325165 Contributors: A876, APerson, AXRL, Ahoerstemeier, AlimanRuna, Andres, Arkuat, Arthena, Auric,
B.d.mills, Beetstra, Benbest, BerserkerBen, Borislav Dopudja, Bryan Derksen, Cadmium, Carlossuarez46, Carnildo, Casliber, CommonsDelinker, Conversion script, Cybercobra, Daniel Case,
Danielle dk, Darrien, David Latapie, DePiep, Deflective, Deor, Double sharp, El C, Emerson7, Emperorbma, Enok Walker, Epbr123, Falcanary, Femto, Fibonacci, Finalius, Flewis, Gaius Cornelius,
GeeIsUs., Ginsuloft, Glenn4pr, Go-in, GregorB, Gypsypkd, Gökhan, Hak-kâ-ngìn, Hashar, HazyM, Headbomb, Hqb, I love chemistry, IForgotToEatBreakFast, Icairns, Ideyal, Inglok, J.delanoy,
JTN, JWB, JWBE, JaGa, Jaraalbe, Jauhienij, JellyFishes, Jennavecia, Jiang, Joanjoc, JoshuaZ, K6ka, Karelj, Kbrose, Kelovy, Kipala, Knowledgeum, Kpalion, Krenair, Kurykh, Kwamikagami,
Kwardle33, LA2, Lamro, Lanthanum-138, Lightmouse, Lugia2453, M-le-mot-dit, Marc Venot, Materialscientist, Mav, Mblormand, Mentifisto, Mild Bill Hiccup, Miss Madeline, Muqman 52,
Naddy, Nedim Ardoğa, Nergaal, New4325, Nihiltres, Nosebud, Novangelis, Omicronpersei8, Oo64eva, Ospalh, Parcly Taxel, Piperh, PlanetStar, Polonium, Polverone, Poolkris, Pras, Pretzelpaws,
Reedy, Res2216firestar, Reyk, Rjwilmsi, Roberta F., Romanm, Sakus, Saperaud, Sbharris, Schneelocke, Sengkang, Sfuerst, Sl, Smartse, Sonicology, Squids and Chips, Srich32977, Stephenb,
Steve Hart, Stifynsemons, Stone, Stratocracy, Suisui, Tagishsimon, Tetracube, Thingg, Thinghy, Thumperward, Trackteur, Uannis, UkPaolo, Unara, Volland, Vsmith, Warut, Watch37264, Werieth,
Whoop whoop pull up, Widr, William Avery, Xeworlebi, Yekrats, Yyy, சசசசச , 107 anonymous edits

97 – Berkelium Source: http://en.wikipedia.org/w/index.php?oldid=607301003 Contributors: A. Carty, A2Kafir, Ahoerstemeier, Alansohn, AlimanRuna, Andres, Angela,
Anoop.m, Arcandam, BRW, Bart133, Beetstra, Benbest, Brian Huffman, Bryan Derksen, CMW275, Carnildo, Casecrer, Chowbok, ChrisGualtieri, CommonsDelinker, Conversion script, Cowbert,
Cryptoid, Cybercobra, Darrien, Dave6, David Latapie, DePiep, Deflective, Doenut1793, Double sharp, Doulos Christos, Edgar181, El C, Eleuther, Emperorbma, Epicgenius, FF2010, Falcanary,
Femto, FruitMart07, Gilliam, GrahamN, Gypsypkd, Hahaha2006, Hashar, Headbomb, Hede2000, Hoops gza, Hqb, I am One of Many, Ideyal, InfoCan, Ioeth, Irishguy, IvanLanin, JWB, JWBE,
JaGa, Jack Greenmaven, Jackson Peebles, Jerem43, Jiang, Joanjoc, Johnlp, Kalamkaar, Kaobear, Karelj, Kbrose, Kelovy, Khazar2, Kurykh, Kwamikagami, Kyoko, LA2, Lamro, Lotje,
M0th3rT3r3asa, Marc Venot, Materialscientist, Mav, Mdf, Muqman 52, Mygerardromance, Naddy, Nedim Ardoğa, Nergaal, Nick Y., Nima1024, No1lakersfan, Oo64eva, PCHS-NJROTC, PP
Jewel, PamD, Parcly Taxel, Pepspirit, Petrb, Physchim62, PlanetStar, Polonium, Polyamorph, Poolkris, R8R Gtrs, Recognizance, Remember, Rettetast, Reza kalani, Rjwilmsi, Roberta F.,
Roentgenium111, Romanm, Saperaud, Schneelocke, Scwlong, Serpentuous, Sl, Sprinkler21, Srich32977, Stifynsemons, Stone, Stratocracy, StringTheory11, Tagishsimon, Tetracube, Tholme,
Thricecube, Thumperward, Tmangray, UpstateNYer, VASANTH S.N., Vsmith, Watch37264, Whoop whoop pull up, Wimt, Xeworlebi, Yekrats, Yyy, Zzuuzz, 132 anonymous edits

98 – Californium Source: http://en.wikipedia.org/w/index.php?oldid=607337526 Contributors: !Darkfire!6'28'14, 1exec1, 2D, 2help, AXRL, Aeroknight, Ahoerstemeier, Akkida,
Alansohn, Alarbus, AlimanRuna, Andres, Anonymous101, Anoop.m, Antandrus, Anttie987, Appasionata, Atarr, Aurocker49, B07, BD2412, BJ Axel, Bart133, Beetstra, Benbest, Bender235,
BillFlis, Blockhouse, Bobo192, Bongwarrior, Brad101, Bryan Derksen, C.A.T.S. CEO, CWenger, Cadmium, Capricorn42, Carcharoth, Carnildo, Casliber, Ccmcguire, Chemicalinterest, Chowbok,
Chris the speller, ChrisGualtieri, ChrisHodgesUK, Christacool, ChromiumCurium, Cimon Avaro, ClamDip, Cliff Dugal, Cmprince, Colonies Chris, Connormah, Conversion script, Coolaidan6,
Courcelles, Cryptic C62, Cybercobra, DMacks, Dabomb87, Dan100, Darrien, David Latapie, Davidwr, Dbloemers612, DePiep, Deflective, Deli nk, Demiurge, DocWatson42, Double sharp,
Dragons flight, Dudley Miles, Duhhhhh, Duhhhhhh, Eddie tejeda, Edgar181, Emperorbma, Encyclopedia77, Epbr123, Eso es todo folks, Evan Robidoux, Eyu100, Femto, Finalius, Finetooth,
Flower Priest, Flyguy649, Flying Jazz, Frosty, GB fan, Gaius Cornelius, George8211, Gilliam, Glenn4pr, GoingBatty, Greenhousegeorge, GregorB, Gyrobo, Gökhan, Hameryko, HazyM,
Headbomb, Henrybauer, Hqb, Hurricane111, I80and, Iamthedeus, Infinite.08, Inglok, Integrity84, Integrity86, Iohannes Animosus, J.delanoy, JNW, JWB, JWBE, JaGa, James Balsom, Jaraalbe,
Jauhienij, Jbrennen, Jeff G., Jiang, Joanjoc, Josh Parris, Jrdioko, Kalamkaar, Karelj, Karlhahn, Katieh5584, Kbh3rd, Kbrose, KeepItClean, Kelovy, Kelseymh, Kevin McE, Kingdon, Knoxville
Physics, KoenigseggCCXR, Kralizec!, Kshieh, Kuru, Kurykh, Kwamikagami, LA2, Lanthanum-138, LarryMorseDCOhio, Lightdarkness, LionMans Account, Lotje, Luna Santin, Lupin, MK8,
Madhero88, Marc Venot, Martin Hogbin, Materialscientist, Mathonius, Mav, Mayur, Mdf, Meteor sandwich yum, Mikeo, Mouse20080706, Mouser, Mumiemonstret, Myanw, Naddy, Nedim
Ardoğa, Nergaal, Nick Y., Nihiltres, Nonsequiturmine, Oliphaunt, Oo64eva, Orez119, Oxymoron83, PDCP, Parcly Taxel, Paul August, Paulistano, PeepP, Peruvianllama, Physchim62, Piano non
troppo, Pinethicket, PlanetStar, Poepkop, Polonium, Poolkris, Pras, Pt, Q43, RJHall, RUL3R, RadiantRay, Randommelon, Ratemonth, Remember, Reyk, Rich Farmbrough, Rjwilmsi, Robert
Keiden, Roberta F., Roentgenium111, Romanm, Rsrikanth05, Rwendland, Rwflammang, Ryan Vesey, Saperaud, Sbharris, Schneelocke, Shaddack, Shawn R. Peasley, Shawnsta42, Shoaler,
Sin-man, Sir Nils, Skarebo, Sl, Smitbritt2014, Squids and Chips, Srich32977, Statue2, Stedder, Steven Zhang, Stifynsemons, Stone, Stratocracy, StringTheory11, Swtpc6800, TCO, TEB728,
TFNorman, Tagishsimon, Tbhotch, Tbranch0507, Tcncv, Tedernst, Tenorcnj, Tetracube, That Guy, From That Show!, Thenormalyears, Thumperward, Tmangray, ToBeFree, Tolly4bolly, Tony1,
Trappist the monk, Trevor MacInnis, Uannis, UnitedStatesian, Vary, Velvetron, Versus22, Vsmith, Vuo, Wackywace, Walkerma, Warut, Watch37264, Weeeeeeeeeeeeee, What!?Why?Who?,

Whosasking, Wiki alf, Will Beback Auto, William Avery, Willking1979, Wknight94, WolfmanSF, Yekrats, Ytterbium2, Yyy, Zantolak, 392 , ‫ יבצ לאינד‬anonymous edits
99 – Einsteinium Source: http://en.wikipedia.org/w/index.php?oldid=604518091 Contributors: 28bytes, Addshore, AdjustShift, AdultSwim, Ahoerstemeier, Alansohn, Alba,
Aleenf1, Alex Lin, AlimanRuna, AlphaEta, Arctic Night, Armbrust, Bachrach44, Bbb2007, Beetstra, Beland, Bender235, Bgwhite, BlueDevil, Bogey97, Brian Huffman, Bryan Derksen, Bth,
CASportsFan, Caltas, CardinalDan, Chowbok, Chris 73, CiTrusD, CommonsDelinker, Conversion script, Cyan, Cybercobra, Cyclopia, DVdm, DanielCD, Darrien, David Latapie, DePiep,
Deflective, Deor, Diehard4.0, Difu Wu, Dirac66, Discospinster, DocWatson42, Double sharp, DragonflySixtyseven, Edgar181, Egomaniac, El C, El aprendelenguas, Element16, Emperorbma,
Encyclopedia77, Epbr123, Esowteric, Everyking, Eweisser, Faeriesaire, Femto, Finlay McWalter, Forteblast, Fraggle81, Fritzpoll, Func, GeeJo, GeorgeMoney, Gholam, Gianluigi, Giftlite,
Ginsuloft, Good Olfactory, Gypsypkd, Gzornenplatz, Haham hanuka, Hak-kâ-ngìn, Hardysinep, Headbomb, Hqb, Hydra Rider, II MusLiM HyBRiD II, Iamkanadian, Igodard, Igoldste, Ipatrol,
IvanLanin, J.meija, JDP90, JEH, JWBE, Jaan513, Jaked122, Jaraalbe, Jasper Deng, Jauhienij, Jeff G., Jiang, Joanjoc, John, JohnCD, Jose Ramos, Josve05a, Jschnur, Jsmith86, KGasso,
Kalamkaar, Kenrick95, Kenyon, Kilo-Lima, King of Hearts, Kingpin13, Koolade343, Kwamikagami, LA2, Lanthanum-138, Logan, LonelyBeacon, Lugia2453, Luna Santin, MJDietrich, MK8,
Magog the Ogre, Makeemlighter, Marc Venot, Materialscientist, Matty4123, Mav, Maxis ftw, Mdf, Mgiganteus1, Michael Devore, Mikael V, Mike.lifeguard, Mikeo, Milonica, Milosci, Mormegil,
Mtpaley, Naddy, Nakon, NawlinWiki, Nedim Ardoğa, Nergaal, NewEnglandYankee, Nick Y., Nickst, Nihiltres, Noctibus, NrDg, Ntsimp, Opelio, Parawiki, Parcly Taxel, Pharaoh of the Wizards,
Philip Trueman, Physicistjedi, Polonium, Polyamorph, Poolkris, Prodego, Puchiko, Pushkar613, R8R Gtrs, Rdsmith4, Reddi, Redsox345678, Reywas92, Reza kalani, Rjwilmsi, Roberta F.,
Roentgenium111, Romanm, Rumkneebeard the Pyrate, Santăr, Saperaud, Sbharris, Schneelocke, Shawn81, Silver starfish, Skunkboy74, Sl, Snezzy, Solomonfromfinland, Spiesr, Squids and
Chips, Srich32977, Stickee, Stifynsemons, Storm Rider, Stratocracy, SymlynX, Tagishsimon, Tempodivalse, Tetracube, The Real Jean-Luc, The Thing That Should Not Be, Thecheesykid, Tholme,
ThomasK, Thumperward, Tide rolls, Uannis, Ucbearcatfan007, Uncle Dick, Vchorozopoulos, Versus22, Victuallers, Vieque, Violinchick1995, Vroo, Vsmith, WODUP, Warut, Wayne Slam,
Wer900, Whoop whoop pull up, Whosasking, Widr, William Avery, Xeworlebi, Xiahou, XinaNicole, Yamakiri, Yekrats, Yyy, Zagalejo, சசசசச , 406 anonymous edits

100 – Fermium Source: http://en.wikipedia.org/w/index.php?oldid=600801295 Contributors: AXRL, Achaemenes, Ahoerstemeier, Airsoftgi88, Ale jrb, AlimanRuna, Arkuat,
Beetstra, Bentley4, Bkell, Brian Huffman, Bryan Derksen, Carnildo, Casliber, ChemNerd, Chowbok, Chris Roy, CommonsDelinker, Conversion script, Cybercobra, Darrien, David Latapie,
DavidLeighEllis, Ddama, DePiep, Deflective, Dina, Double sharp, Edgar181, Egomaniac, Emperorbma, Femto, Finalius, Gijoend81, Glenn4pr, Graham87, Hak-kâ-ngìn, Headbomb, Hqb, Ideyal,
JWBE, Jaan513, Jamo58, Jaraalbe, Jasper Deng, Jiang, John, Justfred, Karelj, Kbrose, Kelovy, Kjramesh, Kwamikagami, LA2, Lainex, Lanthanum-138, Lfh, Ling.Nut, Looxix,
Materialscientist, Mav, Mdf, Mikespedia, Mimihitam, Naddy, Nb07wiki, Nedim Ardoğa, Nergaal, Nick Y., Physchim62, PierceG, Pl1324, PlanetStar, Polonium, Polyamorph, Poolkris, Pras,
Remember, Rjwilmsi, Roberta F., Roentgenium111, Sandman, Saperaud, Schneelocke, Sl, Solomonfromfinland, Stifynsemons, Stone, Stratocracy, Sue Rangell, Tagishsimon, Tetracube, Tholme,
Thumperward, Tide rolls, Uannis, Vsmith, Whoop whoop pull up, Whosasking, William Avery, Willking1979, Wilson44691, Wknight94, Xeworlebi, XinaNicole, Yekrats, Yyy, 99 anonymous
edits

101 – Mendelevium Source: http://en.wikipedia.org/w/index.php?oldid=605201651 Contributors: 1exec1, Ahoerstemeier, AlimanRuna, Arkuat, B3 sii tZ, Beetstra, Benbest, Bill,
Bonzostar, Brian Huffman, Bryan Derksen, Carnildo, Charlottevranken, ClanCC, Clovis Sangrail, Colonies Chris, Conversion script, Darrien, David Latapie, Deflective, Delawaresky, Double
sharp, Edgar181, El C, Emperorbma, Excirial, Femto, Gdommett, Glenn4pr, Hak-kâ-ngìn, Hede2000, Hqb, IanOsgood, Ideyal, Infinite.08, J.meija, JJJiiimmmyyy, Jaan513, Jaraalbe, Jasper Deng,
Jiang, JimVC3, Joanjoc, John beaker, Karl-Henner, Kelovy, Kwamikagami, LairepoNite, Lanthanum-138, LcawteHuggle, LittleOldMe, Log1c bg, Malcolm Farmer, Mark Musante,
Materialscientist, Mav, Mdf, Michel32Nl, Mikespedia, Mimihitam, Mitch Ames, Naddy, Narval, Nb07wiki, Nedim Ardoğa, Nergaal, Nick Y., Nihiltres, OverMyHead, PeepP, PleaseStand,
Polonium, Poolkris, Pras, Pstanton, Remember, Reza kalani, Rjwilmsi, Roberta F., Roentgenium111, Romanm, Salix alba, Saperaud, Schneelocke, Sgr927, Shirik, Shogunzhu, Shynash, SimonP,
Sl, Squids and Chips, Srich32977, Stifynsemons, Stratocracy, Tagishsimon, Tasc, Tetracube, ThaddeusB, Tholme, Thumperward, Uannis, VASANTH S.N., Varlaam, Vcelloho, Videogamer344,
Vsmith, Wavelength, Widr, Wikid77, William Avery, Wiwaxia, Writtenright, Xanchester, XinaNicole, Yekrats, Ynhockey, 122 anonymous edits

102 – Nobelium Source: http://en.wikipedia.org/w/index.php?oldid=598765420 Contributors: Ahoerstemeier, Alfio, AlimanRuna, Andres, AndyVolykhov, Atjesse, Bender235,
Black-Velvet, Bobo192, BostonMA, Brian Huffman, Bryan Derksen, Bubblegumski, CO, Carnildo, ChemNerd, Church, Conversion script, Curb Chain, Cybercobra, Dac04, Darrien, David Latapie,
DePiep, Deflective, DesertAngel, Discospinster, Double sharp, Drjezza, Edgar181, Emperorbma, Eog1916, Everyking, Fadesga, Fangfufu, Favonian, Femto, Glenn4pr, Glinness, Gwankney, Hak-
kâ-ngìn, Headbomb, Hede2000, IRP, Ideyal, Inglok, J.delanoy, J04n, Jaraalbe, Jauhienij, Jiang, Joanjoc, John, Karelj, Karl-Henner, Khazar, Kwamikagami, LA2, LeaveSleaves, LilHelpa, Limulus,
LizardWizard, Marc9510000, Materialscientist, Mav, Mdf, Melonkelon, Mimihitam, Miquonranger03, Miss Madeline, Mmxx, Montrealais, Naddy, Nedim Ardoğa, Nergaal, Nibbleadams,
Nihiltres, Ohconfucius, Polonium, Poolkris, Pras, PrimeCupEevee, Rambam rashi, Remember, Renrenren, Reza kalani, Roberta F., Roentgenium111, Rursus, Santăr, Saperaud, Schneelocke,
Seegoon, Sl, Slazenger, Snigbrook, Srich32977, Steve2011, Stifynsemons, Stone, Swalot, T42N24T, Tagishsimon, Tetracube, Thinking of England, Thumperward, Tinss, Tom harrison,
Tommy2010, Uannis, Undead warrior, VASANTH S.N., Van helsing, Vsmith, Vuerqex, Vuo, Warut, Watch37264, Widr, William Avery, XinaNicole, Yekrats, Yyy, 130 anonymous edits

103 – Lawrencium Source: http://en.wikipedia.org/w/index.php?oldid=599645212 Contributors: 03jmgibbens, 777sms, 78.26, Adashiel, Adventurer, Ahoerstemeier,
AlimanRuna, ApsbaMd2, Arctic Kangaroo, Ascidian, Avicennasis, Beetstra, Bigtimepeace, Bkell, Bryan Derksen, Burubuz, CASportsFan, CJ, Cannolis, Carnildo, Chimneyfreak, Christian List,
Conversion script, Cybercobra, D6, Dajwilkinson, Darrien, Dave6, David Shay, David1010, Deadbeef, Deflective, Digital infinity, Dirt palace, Dom497, Double sharp, Drjezza, Drmies, Edgar181,
ElCubano09, Eltomo11, Emperorbma, Enchanter, Enviroboy, Eog1916, Ericamick, Femto, Flying Jazz, Francs2000, GB fan, Gogo Dodo, Goudzovski, Graham87, Granf, Gypsypkd, Gyrobo, Hak-
kâ-ngìn, Headbomb, Howarjos000, Hqb, Icairns, Imdwalrus, J.delanoy, J04n, JLaTondre, Jaan513, JackSchmidt, Jakec, Jamesontai, Jan1nad, Jaraalbe, Jauhienij, Jiang, John, Karl-Henner,
Kbscvvin, Keyboard mouse, Kgf0, Kingdon, Kingpin13, Kolbasz, Kumorifox, Kwamikagami, LA2, Lad+ve, Larryincinci, Lelan doodie, Lowe4091, Materialscientist, Mav, Mdf, Metal Militia,
Mikespedia, Mimihitam, Naddy, Nedim Ardoğa, Nergaal, Nihiltres, Nikai, PeterJeremy, Philip Trueman, Physchim62, Polonium, Polyamorph, Poolkris, Pras, Psyche825, Qtsc, R8R Gtrs,
Remember, Rich Farmbrough, Rjwilmsi, Roberta F., Roentgenium111, Romanskolduns, Santăr, Saperaud, Schneelocke, Serinoah, Simon12, Sinnyg, Sl, Smelliegibbs, Snclucas, Squids and Chips,
Srich32977, Srnec, Steinbach, Stifynsemons, Stone, Stratocracy, Syrthiss, Tagishsimon, Tetracube, The Thing That Should Not Be, Thinking of England, Thumperward, Tide rolls, Uannis, Vsmith,
Werdan7, William Avery, XinaNicole, Yekrats, Zephyrus67, Zerabat, Zoicon5, 220 anonymous edits

Actinides in the environment Source: http://en.wikipedia.org/w/index.php?oldid=597007453 Contributors: A876, Alan Liefting, Amarande, Barkeep, BeefRendang, Beetstra, Cadmium,
Carstensen, Cedders, Cgingold, Chris83, Clicketyclack, Colonies Chris, D.h, DV8 2XL, Del245, DocWatson42, Double sharp, Dr Zak, Ettrig, Everyking, Flyguy649, Furrykef, GangofOne,
Gilderien, Give Peace A Chance, Grumpyyoungman01, Iothiania, JWB, Jaganath, Jcabraham, John Walker (fourmilab.ch), Johnherrick, Joseph Solis in Australia, Karmos, Kelly Martin,
Keoniphoenix, Kimdino, Kolbasz, Kwertii, Limulus, LizardWizard, Malik8U, Materialscientist, Nono64, Ojovan, Optigan13, PaperTruths, Pearle, Polyamorph, Pstudier, Puzl bustr, Scientific29,
Sinus, Sj, Templationist, Thomas Connor, Touch Of Light, Trappist the monk, Travelbird, Velella, Vuo, Warut, Wavelength, 29 anonymous edits

Major actinide Source: http://en.wikipedia.org/w/index.php?oldid=535766581 Contributors: Cadmium, Garion96, Headbomb, JWB, Kjkolb, Skapur

Minor actinide Source: http://en.wikipedia.org/w/index.php?oldid=596674545 Contributors: Aaron Brenneman, Cadmium, Cdecoro, DePiep, DexDor, Dr Zak, Elminster Aumar, GregorB, Itub,
JWB, Kjkolb, Midgley, Molinari, Nergaal, Nikai, PranksterTurtle, Reyk, Rigadoun, Rjwilmsi, Thomas Connor, Vuo, Whoop whoop pull up, Xuanmingzi, 4 anonymous edits
Image Sources, Licenses and 1
Contributors

Image Sources, Licenses and Contributors


File:Transparent.gif Source: http://en.wikipedia.org/w/index.php?title=File:Transparent.gif License: Public Domain Contributors: Edokter
File:Nagasakibomb.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Nagasakibomb.jpg License: Public Domain Contributors: The picture was taken by Charles Levy from one of
the B-29 Superfortresses used in the attack.
File:Enrico Fermi 1943-49.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Enrico_Fermi_1943-49.jpg License: Public Domain Contributors: Department of Energy. Office
of Public Affairs
File:Glenn Seaborg - 1964.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Glenn_Seaborg_-_1964.jpg License: Public Domain Contributors: Atomic Energy Commission. (1946 -
01/19/1975)
File:Isotopes and half-life.svg Source: http://en.wikipedia.org/w/index.php?title=File:Isotopes_and_half-life.svg License: Public Domain Contributors: BenRG
File:Uranium ore square.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_ore_square.jpg License: Public Domain Contributors: Hawkeye7, JWBE, Nolanus, Tamorlan,
Tungsten, Zscout370, 1 anonymous edits
File:MonaziteUSGOV.jpg Source: http://en.wikipedia.org/w/index.php?title=File:MonaziteUSGOV.jpg License: Public Domain Contributors: Saperaud
File:Plutonium and uranium extraction from nuclear fuel-eng.svg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_and_uranium_extraction_from_nuclear_fuel-eng.svg
License: Creative Commons Attribution-Sharealike 3.0 Contributors: Plutonium_and_uranium_extraction_from_nuclear_fuel-rus.svg: HarDNox derivative work: Materialscientist (talk)
File:ActinidesLattice.png Source: http://en.wikipedia.org/w/index.php?title=File:ActinidesLattice.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Materialscientist
(talk)
File:ACTIION.PNG Source: http://en.wikipedia.org/w/index.php?title=File:ACTIION.PNG License: Creative Commons Attribution-Sharealike 3.0 Contributors: Materialscientist
File:Radioisotope thermoelectric generator plutonium pellet.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Radioisotope_thermoelectric_generator_plutonium_pellet.jpg
License: Public Domain Contributors: Original uploader was Deglr6328 at en.wikipedia. Later version(s) were uploaded by Raeky, Johnny--Bravo, Aarchiba at en.wikipedia.
File:Californium.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Californium.jpg License: Public Domain Contributors: United States Department of Energy (see
File:Einsteinium.jpg)
File:Uranyl Nitrate.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Uranyl_Nitrate.jpg License: Creative Commons Attribution 3.0 Contributors: Minerallad at en.wikipedia
File:U Oxstufen.jpg Source: http://en.wikipedia.org/w/index.php?title=File:U_Oxstufen.jpg License: Public Domain Contributors: Los Alamos National Lab.
File:Np ox st .jpg Source: http://en.wikipedia.org/w/index.php?title=File:Np_ox_st_.jpg License: Public Domain Contributors: Los Alamos National Lab.
File:Plutonium in solution.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_in_solution.jpg License: Public Domain Contributors: Los Alamos National Laboratory
File:UCl4.jpg Source: http://en.wikipedia.org/w/index.php?title=File:UCl4.jpg License: Creative Commons Attribution 3.0 Contributors: Original uploader was Minerallad at en.wikipedia
File:UF6.jpg Source: http://en.wikipedia.org/w/index.php?title=File:UF6.jpg License: Public Domain Contributors: Benjah-bmm27, Fallschirmjäger, Tungsten
File:Yellowcake.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Yellowcake.jpg License: Public Domain Contributors: Bomazi, JWBE, Mattes, Pfctdayelise, Tungsten, 1
anonymous edits
File:CaF2 polyhedra.png Source: http://en.wikipedia.org/w/index.php?title=File:CaF2_polyhedra.png License: Public Domain Contributors: Solid State
File:UCl3 without caption.png Source: http://en.wikipedia.org/w/index.php?title=File:UCl3_without_caption.png License: Creative Commons Attribution-Sharealike 3.0,2.5,2.0,1.0
Contributors: UCl3.png: Solid State derivative work: Leyo
File:Einsteinium triiodide by transmitted light.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Einsteinium_triiodide_by_transmitted_light.jpg License: Public Domain
Contributors: Los Alamos National Lab.
File:InsideSmokeDetector.jpg Source: http://en.wikipedia.org/w/index.php?title=File:InsideSmokeDetector.jpg License: Creative Commons Attribution-Sharealike 2.0 Contributors: MD111
File:Cerenkov Effect.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Cerenkov_Effect.jpg License: Public Domain Contributors: United States Nuclear Regulatory Commission
File:Heterogeneous reactor scheme.png Source: http://en.wikipedia.org/w/index.php?title=File:Heterogeneous_reactor_scheme.png License: Creative Commons Attribution-ShareAlike 3.0
Unported Contributors: Panther
File:Alfa beta gamma radiation penetration.svg Source: http://en.wikipedia.org/w/index.php?title=File:Alfa_beta_gamma_radiation_penetration.svg License: Creative Commons
Attribution-ShareAlike 3.0 Unported Contributors: Alfa_beta_gamma_radiation.svg: User:Stannered derivative work: Ehamberg (talk)
File:Periodic Table Radioactivity.svg Source: http://en.wikipedia.org/w/index.php?title=File:Periodic_Table_Radioactivity.svg License: Creative Commons Attribution-Sharealike 2.5
Contributors: Periodic_Table_Armtuk3.svg: Armtuk (talk) derivative work: Alessio Rolleri (talk) derivative work: Gringer (talk)
File:Cubic-face-centered.svg Source: http://en.wikipedia.org/w/index.php?title=File:Cubic-face-centered.svg License: GNU Free Documentation License Contributors: Original PNGs by
Daniel Mayer and DrBob, traced in Inkscape by User:Stannered
File:Uraninite-39029.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Uraninite-39029.jpg License: unknown Contributors: -
File:DOTA polyaminocarboxylic acid.png Source: http://en.wikipedia.org/w/index.php?title=File:DOTA_polyaminocarboxylic_acid.png License: Public Domain Contributors: Petergans at
en.wikipedia
file:Thorium sample 0.1g.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Thorium_sample_0.1g.jpg License: Free Art License Contributors: Alchemist-hp (talk) (
www.pse-mendelejew.de)
File:Radioactive Lenses (group shot).jpg Source: http://en.wikipedia.org/w/index.php?title=File:Radioactive_Lenses_(group_shot).jpg License: Creative Commons Attribution 2.0
Contributors: s58y
File:Keplers supernova.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Keplers_supernova.jpg License: Public Domain Contributors: NASA/ESA/JHU/R.Sankrit & W.Blair
File:Evolution of Earth's radiogenic heat.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Evolution_of_Earth's_radiogenic_heat.jpg License: Creative Commons Attribution-
Sharealike 3.0 Contributors: User:Bkilli1
File:NAMrad Th let.gif Source: http://en.wikipedia.org/w/index.php?title=File:NAMrad_Th_let.gif License: Public Domain Contributors: Limulus
File:Lunar KREEP concentrations.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Lunar_KREEP_concentrations.jpg License: Public Domain Contributors: NASA
Image:Monazit opening acid.gif Source: http://en.wikipedia.org/w/index.php?title=File:Monazit_opening_acid.gif License: Creative Commons Attribution-Sharealike 3.0 Contributors:
Hermann Luyken
File:India-locator-map-thorium2012.svg Source: http://en.wikipedia.org/w/index.php?title=File:India-locator-map-thorium2012.svg License: Creative Commons Attribution-Sharealike
3.0,2.5,2.0,1.0 Contributors: User:Limulus
File:PSM V74 D233 Thorium radioactive incandescent gas mantle placed above plant seeds.png Source: http://en.wikipedia.org/w/index.php?
title=File:PSM_V74_D233_Thorium_radioactive_incandescent_gas_mantle_placed_above_plant_seeds.png License: Public Domain Contributors: Ineuw, Limulus
File:Tetragonal.svg Source: http://en.wikipedia.org/w/index.php?title=File:Tetragonal.svg License: GNU Free Documentation License Contributors: Original PNGs by Daniel Mayer, traced in
Inkscape by User:Stannered
File:Mendelejevs periodiska system 1871.png Source: http://en.wikipedia.org/w/index.php?title=File:Mendelejevs_periodiska_system_1871.png License: Public Domain Contributors:
Original uploader was Den fjättrade ankan at sv.wikipedia
File:PaF5geometry.PNG Source: http://en.wikipedia.org/w/index.php?title=File:PaF5geometry.PNG License: Creative Commons Attribution-Sharealike 3.0 Contributors: Materialscientist
File:Uranocene-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Uranocene-3D-balls.png License: Public Domain Contributors: Benjah-bmm27
file:HEUraniumC.jpg Source: http://en.wikipedia.org/w/index.php?title=File:HEUraniumC.jpg License: Public Domain Contributors: Original uploader was Zxctypo at en.wikipedia
File:Orthorhombic.svg Source: http://en.wikipedia.org/w/index.php?title=File:Orthorhombic.svg License: GNU Free Documentation License Contributors: Original PNGs by Daniel Mayer,
traced in Inkscape by User:Stannered
File:Nuclear fission.svg Source: http://en.wikipedia.org/w/index.php?title=File:Nuclear_fission.svg License: Public Domain Contributors: User:Fastfission
File:30mm DU slug.jpg Source: http://en.wikipedia.org/w/index.php?title=File:30mm_DU_slug.jpg License: Public Domain Contributors: Original uploader was Nrcprm2026 at en.wikipedia
File:Nuclear Power Plant 2.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Nuclear_Power_Plant_2.jpg License: Copyrighted free use Contributors: User Alvinrune on
en.wikipedia
File:U glass with black light.jpg Source: http://en.wikipedia.org/w/index.php?title=File:U_glass_with_black_light.jpg License: Creative Commons Attribution-Sharealike 2.5 Contributors: JJ
Harrison, Nolanus, Túrelio, WikipediaMaster, Zvesoulis
File:Vacuum capacitor with uranium glass.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Vacuum_capacitor_with_uranium_glass.jpg License: Creative Commons
Attribution-Sharealike 3.0,2.5,2.0,1.0 Contributors: Warut Roonguthai
File:Becquerel plate.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Becquerel_plate.jpg License: Public Domain Contributors: Ranveig, Sumanch, Svajcr
File:UraniumCubesLarge.jpg Source: http://en.wikipedia.org/w/index.php?title=File:UraniumCubesLarge.jpg License: Public Domain Contributors: U.S. Department of Energy
File:Atomic cloud over Hiroshima.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Atomic_cloud_over_Hiroshima.jpg License: Public Domain Contributors: Enola Gay Tail
Gunner S/Sgt. George R. (Bob) Caron
File:First four nuclear lit bulbs.jpeg Source: http://en.wikipedia.org/w/index.php?title=File:First_four_nuclear_lit_bulbs.jpeg License: Public Domain Contributors: Howcheng,
LilHelpa, NewebNL, Thuresson, Tungsten, Túrelio, Werewombat, 3 anonymous edits
File:US and USSR nuclear stockpiles.svg Source: http://en.wikipedia.org/w/index.php?title=File:US_and_USSR_nuclear_stockpiles.svg License: Public Domain Contributors: Created by
User:Fastfission first by mapping the lines using OpenOffice.org's Calc program, then exporting a graph to SVG, and the performing substantial aesthetic modifications in Inkscape.
File:Pichblende.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Pichblende.jpg License: Creative Commons Attribution-Sharealike 2.5 Contributors: Original uploader was Kgrr at
en.wikipedia
File:Citrobacter freundii.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Citrobacter_freundii.jpg License: Public Domain Contributors: Copydays, Georgeryp, Kookaburra, NEON
ja, Ninjatacoshell, Romary, 2 anonymous edits
File:U production-demand.png Source: http://en.wikipedia.org/w/index.php?title=File:U_production-demand.png License: Creative Commons Attribution-Sharealike 3.0 Contributors:
Materialscientist
File:MonthlyUraniumSpot.png Source: http://en.wikipedia.org/w/index.php?title=File:MonthlyUraniumSpot.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Celuca
file:U3O8lattice.jpg Source: http://en.wikipedia.org/w/index.php?title=File:U3O8lattice.jpg License: Public Domain Contributors: Original uploader was Cadmium at en.wikipedia
file:UO2lattice.jpg Source: http://en.wikipedia.org/w/index.php?title=File:UO2lattice.jpg License: Public Domain Contributors: Original uploader was Cadmium at en.wikipedia
File:Uranium pourdaix diagram in water.png Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_pourdaix_diagram_in_water.png License: Public Domain Contributors:
Original uploader was Cadmium at en.wikipedia
File:Uranium pourdiax diagram in carbonate media.png Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_pourdiax_diagram_in_carbonate_media.png License: Public
Domain Contributors: Original uploader was Cadmium at en.wikipedia
File:Uranium fraction diagram with no carbonate.png Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_fraction_diagram_with_no_carbonate.png License: Public Domain
Contributors: Original uploader was Cadmium at en.wikipedia
File:Uranium fraction diagram with carbonate present.png Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_fraction_diagram_with_carbonate_present.png License: Public
Domain Contributors: Original uploader was Cadmium at en.wikipedia
File:Uranium hexafluoride crystals sealed in an ampoule.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Uranium_hexafluoride_crystals_sealed_in_an_ampoule.jpg License:
Public Domain Contributors: Original uploader was Deglr6328 at en.wikipedia
File:Gas centrifuge cascade.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Gas_centrifuge_cascade.jpg License: Public Domain Contributors: U.S. DOE
file:neptunium2.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Neptunium2.jpg License: Public Domain Contributors: Los Alamos National Laboratory,
file:Plutonium3.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium3.jpg License: Public Domain Contributors: U.S. Department of Energy,
File:Monoclinic.svg Source: http://en.wikipedia.org/w/index.php?title=File:Monoclinic.svg License: GNU Free Documentation License Contributors: Original PNGs by Daniel Mayer,
traced in Inkscape by User:Stannered
File:Plutonium density-eng.svg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_density-eng.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors:
Plutonium_density.svg: HarDNox derivative work: Materialscientist (talk)
File:Plutonium ring.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_ring.jpg License: Public Domain Contributors: Los Alamos National Laboratory
File:PuIsotopes.png Source: http://en.wikipedia.org/w/index.php?title=File:PuIsotopes.png License: Creative Commons Attribution 3.0 Contributors: Изотопы.svg: HarDNox derivative work:
Materialscientist (talk)
File:Plutonium pyrophoricity.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_pyrophoricity.jpg License: Public Domain Contributors: Los Alamos National Laboratory
File:96602765.lowres.jpeg Source: http://en.wikipedia.org/w/index.php?title=File:96602765.lowres.jpeg License: Public Domain Contributors: Berkeley-Laboratory
File:Hanford B Reactor.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Hanford_B_Reactor.jpg License: Public Domain Contributors: Original uploader was Northwest-historian
at en.wikipedia
Image:Hanford N Reactor adjusted.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Hanford_N_Reactor_adjusted.jpg License: Public Domain Contributors: United States
Department of Energy
File:Fission bomb assembly methods.svg Source: http://en.wikipedia.org/w/index.php?title=File:Fission_bomb_assembly_methods.svg License: Public Domain Contributors: Fastfission
File:Yucca Mountain emplacement drifts.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Yucca_Mountain_emplacement_drifts.jpg License: Public Domain Contributors: U.S.
Department of Energy
File:Plutonium pellet.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Plutonium_pellet.jpg License: Public Domain Contributors: Adam Cuerden, Bomazi, Craigboy, D-Kuru,
Fastfission, Mav, Rlevse, Uwe W.
File:Partially-reflected-plutonium-sphere.jpeg Source: http://en.wikipedia.org/w/index.php?title=File:Partially-reflected-plutonium-sphere.jpeg License: Public Domain Contributors: Los
Alamos National Laboratory
file:Americium microscope.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Americium_microscope.jpg License: Creative Commons Attribution 3.0 Contributors: Bionerd
File:Hexagonal.svg Source: http://en.wikipedia.org/w/index.php?title=File:Hexagonal.svg License: BSD Contributors: Original uploader was Danieljamesscott at en.wikipedia
File:Berkeley 60-inch cyclotron.gif Source: http://en.wikipedia.org/w/index.php?title=File:Berkeley_60-inch_cyclotron.gif License: unknown Contributors: -
File:Ivy Mike - mushroom cloud.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ivy_Mike_-_mushroom_cloud.jpg License: Public Domain Contributors: Avron, Bomazi,
Chetvorno, Fastfission, 2 anonymous edits
File:Elutionskurven Tb Gd Eu und Bk Cm Am.png Source: http://en.wikipedia.org/w/index.php?title=File:Elutionskurven_Tb_Gd_Eu_und_Bk_Cm_Am.png License: Public Domain
Contributors: S. G. Thompson, A. Ghiorso, and G. T. Seaborg () - originally published in US-gov classified documents , thus public domain
File:Closest packing ABAC.png Source: http://en.wikipedia.org/w/index.php?title=File:Closest_packing_ABAC.png License: Public Domain Contributors: Solid State
file:Residential smoke detector.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Residential_smoke_detector.jpg License: Public Domain Contributors: Oleg Alexandrov
file:InsideSmokeDetector.jpg Source: http://en.wikipedia.org/w/index.php?title=File:InsideSmokeDetector.jpg License: Creative Commons Attribution-Sharealike 2.0 Contributors: MD111
File:Hexagonal close packed.svg Source: http://en.wikipedia.org/w/index.php?title=File:Hexagonal_close_packed.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors:
DePiep, Torsch
file:Marie Curie (Nobel-Chem).png Source: http://en.wikipedia.org/w/index.php?title=File:Marie_Curie_(Nobel-Chem).png License: Public Domain Contributors: Afrank99, Bohème,
DEDB, Fastfission, Kjunix, Martin H., Materialscientist, Staszek Lem, 4 anonymous edits
file:Curie-pierre.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Curie-pierre.jpg License: Public Domain Contributors: Original uploader was Matthias Bock at de.wikipedia
File:Cm-Fluoreszenz.GIF Source: http://en.wikipedia.org/w/index.php?title=File:Cm-Fluoreszenz.GIF License: Public Domain Contributors: Eschenmoser + Kollegen
File:Sasahara.svg Source: http://en.wikipedia.org/w/index.php?title=File:Sasahara.svg License: GNU Free Documentation License Contributors: Original uploader was JWB at en.wikipedia
File:MER APXS PIA05113.jpg Source: http://en.wikipedia.org/w/index.php?title=File:MER_APXS_PIA05113.jpg License: Public Domain Contributors: 84user, Bricktop, D-Kuru,
Huntster file:Berkelium metal.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Berkelium_metal.jpg License: Public Domain Contributors: Oak Ridge National Laboratory, US
Department of
Energy
File:The University of California Berkeley 1868.svg Source: http://en.wikipedia.org/w/index.php?title=File:The_University_of_California_Berkeley_1868.svg License: Public Domain
Contributors: Original University of California seal: probably Tiffany & Co,; This SVG file: User:Casecrer
File:Berkelium.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Berkelium.jpg License: Public Domain Contributors: ORNL, Department of Energy
file:Californium.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Californium.jpg License: Public Domain Contributors: United States Department of Energy (see
File:Einsteinium.jpg)
File:Operation Crossroads Baker Edit.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Operation_Crossroads_Baker_Edit.jpg License: Public Domain Contributors:
Operation_Crossroads_Baker_(wide).jpg: United States Department of Defense (either the U.S. Army or the U.S. Navy) derivative work: Victorrocha (talk)
File:Cf 252 Produktion.png Source: http://en.wikipedia.org/w/index.php?title=File:Cf_252_Produktion.png License: Public Domain Contributors: JWBE
File:CfShield.JPG Source: http://en.wikipedia.org/w/index.php?title=File:CfShield.JPG License: Public Domain Contributors: Unknown, US Department of Energy
file:Einsteinium.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Einsteinium.jpg License: Public Domain Contributors: Haire, R. G., US Department of Energy. Touched up
by Materialscientist at en.wikipedia.
File:Einstein1921 by F Schmutzer 2.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Einstein1921_by_F_Schmutzer_2.jpg License: Public Domain Contributors: Diego pmc,
Emijrp, Frank C. Müller, Hemulen, Lobo, Quibik, Vonvon, Андрей Романенко, 6 anonymous edits
File:EinsteiniumGlow.JPG Source: http://en.wikipedia.org/w/index.php?title=File:EinsteiniumGlow.JPG License: Public Domain Contributors: Haire, R. G., Department of Energy
File:EsProduction.png Source: http://en.wikipedia.org/w/index.php?title=File:EsProduction.png License: Public Domain Contributors: Lawrence Berkeley National Laboratory, Dep of Energy
File:ActinideExplosionSynthesis.png Source: http://en.wikipedia.org/w/index.php?title=File:ActinideExplosionSynthesis.png License: Public Domain Contributors: Lawrence Berkeley
National Laboratory, Dep of Energy
File:Elutionskurven Fm Es Cf Bk Cm Am.png Source: http://en.wikipedia.org/w/index.php?title=File:Elutionskurven_Fm_Es_Cf_Bk_Cm_Am.png License: Public Domain Contributors:
UCRL
File:Decay of Fermium-257.PNG Source: http://en.wikipedia.org/w/index.php?title=File:Decay_of_Fermium-257.PNG License: Creative Commons Attribution-Sharealike 2.5
Contributors: Decay_of_Fermium-257_to_the_Neptunium_series.gif: Uwe W. derivative work: Materialscientist (talk)
File:Fermium-Ytterbium Alloy.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Fermium-Ytterbium_Alloy.jpg License: Public Domain Contributors: Ben E. Lewis.
file:speakerlink-new.svg Source: http://en.wikipedia.org/w/index.php?title=File:Speakerlink-new.svg License: Creative Commons Zero Contributors: User:Kelvinsong
Image:Sasahara.svg Source: http://en.wikipedia.org/w/index.php?title=File:Sasahara.svg License: GNU Free Documentation License Contributors: Original uploader was JWB at en.wikipedia
License 1

License
Creative Commons Attribution-Share Alike 3.0
//creativecommons.org/licenses/by-sa/3.0/

Vous aimerez peut-être aussi