Terpene

The actinide or actinoid (IUPAC nomenclature) series encompasses the 15 chemical elements with atomic numbers from 89 to 103, actinium to lawrencium.[2] The actinide series derives its name from the group 3 element actinium; although actinoid (actinide) means "actinium-like" and therefore should exclude actinium, it is usually included in the series for the purpose of comparison.[3] Only thorium and uranium occur in more than trace quantities in nature; the other actinides are synthetic elements.[4] The actinides are usually considered to be f-block elements; they show much more variable valence than the lanthanides. All actinides are radioactive and release energy upon radioactive decay; uranium, thorium, and plutonium, the most abundant actinides on Earth, are used in nuclear reactors and nuclear weapons.

89Ac 90Th 91Pa 92U 93Np 94Pu 95Am 96Cm 97Bk 98Cf 99Es 100Fm 101Md 102No 103Lr

Contents

[edit] Synthesis

Synthesis of transuranium elements[5][6]
Element Year Method
Neptunium 1940 Bombarding 238U by neutrons
Plutonium 1941 Bombarding 238U by deuterons
Americium 1944 Bombarding 239Pu by neutrons
Curium 1944 Bombarding 239Pu by α-particles
Berkelium 1949 Bombarding 241Am by α-particles
Californium 1950 Bombarding 242Cm by α-particles
Einsteinium 1952 As a product of nuclear explosion
Fermium 1952 As a product of nuclear explosion
Mendelevium 1955 Bombarding 253Es by α-particles
Nobelium 1965 Bombarding 243Am by 15N
or 238U with α-particles
Lawrencium 1961–1971 Bombarding 252Cf by 10B or 11B
and of 243Am with 18O

Like lanthanides, actinides form a family of elements with similar properties. Within 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 and thorium, followed by plutonium, americium, actinium, protactinium and neptunium.

At present, there are two major methods of producing isotopes of transplutonium elements: irradiation of the lighter elements with neutrons or accelerated charged particles. The first method is most important for applications, as only neutron irradiation using nuclear reactors allows to produce sizeable amounts of synthetic actinides; however, it is limited to relatively light elements. The advantage of the second method is that it allows obtaining elements heavier than plutonium, as well as neutron-deficient isotopes, which are not formed during neutron irradiation.[7]

In 1962–1966, there were attempts in the United States to produce transplutonium isotopes using a series of 6 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 and nuclear fission.[7]

The existence of transuranium elements was suggested by Enrico Fermi based on his experiments in 1934,[8][9] and the terms "actinides" (and "lanthanides"; from Greek éidos meaning "type") were proposed in 1948 by Sergey Shchukarev.[10]

[edit] From actinium to neptunium

Enrico Fermi suggested the existence of transuranium elements in 1934.

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,[4] 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.[11] He also isolated the first sample of uranium metal by heating uranium tetrachloride with potassium.[12] 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 experimentally in 1882 by K. Zimmerman.[13][14]

Thorium oxide was discovered by Friedrich Wöhler in the mineral, which was found in Norway (1827).[15] Jöns Jacob Berzelius characterized this material in more detail by in 1828. By reduction of thorium tetrachloride with potassium, he isolated the metal and named it thorium after the Norse god of thunder and lightning Thor.[16][17] The same isolation method was later used by Péligot for uranium.[4]

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[18] and (in 1900) as similar to thorium.[19] The discovery of actinium by Debierne was however questioned in 1971[20] and 2000,[21] arguing that Debierne's publications in 1904 contradict to his earlier work of 1899–1900. The name for word actinium comes from the Greek aktis, aktinos (ακτίς, ακτίνος), meaning beam or ray. This metal was discovered not by its own radiation but by the radiation of the daughter products.[15][22] Owing to the close similarity of actinium and lanthanum and low abundance, pure actinium could only be produced in 1950.

Protactinium was possibly isolated in 1900 by William Crookes.[23] It was first identified in 1913, when Kasimir Fajans and Oswald Helmuth Göhring encountered the short-lived isotope 234mPa (half-life 1.17 minutes) during their studies of the 238U decay. They named the new element brevium (from Latin brevis meaning brief);[24][25] the name was changed to protoactinium (from Greek πρῶτος + ἀκτίς meaning "first beam element") in 1918 when two groups of scientists, led by Otto Hahn and Lise Meitner of Germany and Frederick Soddy and John Cranston of Great Britain, independently discovered 231Pa. 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 from 60 tonnes of waste left after extraction of uranium from its ore.[26]

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.[27] They produced the 239Np isotope (half-life 2.4 days) by bombarding uranium with slow neutrons.[26] It was the first transuranium element produced synthetically.[28]

[edit] Plutonium and above

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 developed more than 100 atomic actinide isotopes.

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:

\mathrm{ {}^{238}_{92}U + {}^{1}_{0}n\ \xrightarrow\ \ {}^{239}_{92}U\ \xrightarrow[23.5\ min]{\beta^-}\ {}^{239}_{93}Np\ \xrightarrow[2.3\ days]{\beta^-}\ {}^{239}_{94}Pu\ \xrightarrow[2.4\cdot 10^4\ years]{\alpha} }

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.[29]

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

~\mathrm{{}_{~92}^{238}{}U+{}_{10}^{22}{}Ne \longrightarrow {}_{102}^{256}{}No+4{}_0^1{}n}.

First isotopes of transplutonium elements, americium-241 and curium-242, were synthesized in 1944 by Glenn T. Seaborg, Ralph A. James and Albert Ghiorso.[30] Curium-242 was obtained by bombarding plutonium-239 with 32-MeV α-particles

~\mathrm{{}_{~94}^{239}{}Pu+{}_2^4{}He \longrightarrow {}_{~96}^{242}{}Cm+{}_0^1{}n}.

The americium-241 and curium-242 isotopes also were produced by irradiating plutonium in a nuclear reactor.[7] The latter element was named after Marie Curie and her husband Pierre who are noted for discovering radium and for their work in radioactivity.

Bombarding curium-242 with α-particles resulted in an isotope of californium 245Cf (1950), and a similar procedure yielded in 1949 berkelium-243 from americium-241.[31] 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.[32]

Americium hydroxide

In 1945, B. B. Cunningham obtained the first bulk chemical compound of a transplutonium element, namely americium hydroxide.[33] Over the next 3–4 years, milligram quantities of americium and microgram amounts of curium were accumulated that allowed production of isotopes of berkelium (Thomson, 1949)[34][35] and californium (Thomson, 1950).[36][37][38] Sizeable amounts of these elements were produced only in 1958 (Burris B. Cunningham and Stanley G. Thomson),[39] and the first californium compound (0.3 µg of CfOCl) was obtained only in 1960 by B. B. Cunningham and J. C. Wallmann).[40]

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.[41][42] 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.[43][44] The "Ivy Mike" studies were declassified and published in 1955.[41] The first significant (submicrograms) amounts of einsteinium were produced in 1961 by Cunningham and colleagues, but this has not been done for fermium yet.[45]

The first isotope of mendelevium, 256Md (half-life 87 min), was synthesized by Albert Ghiorso, Glenn T. Seaborg, Gregory R. Choppin, Bernard G. Harvey and Stanley G. Thompson when they bombarded an 253Es target with alpha particles in the Berkeley Radiation Laboratory's 60-inch cyclotron; this was the first isotope of any element to be synthesized one atom at a time.[46]

There were several attempts to obtain isotopes of nobelium by Swedish (1957) and American (1958) groups, but the first reliable results was the synthesis of 256No by the Russian group (Georgy Flyorov et al.) in 1965, as acknowledges by the IUPAC in 1992. In their experiments, Flyorov et al. bombarded uranium-238 with neon-22.[5]

In 1961, Ghiorso et al. obtained the first isotope of lawrencium by irradiating californium (mostly californium-252) with boron-10 and boron-11 ions.[5] The mass number of this isotope was not clearly established (possibly 258 or 259) at the time. In 1965, 256Lr were synthesized by Flyorov et al. from 243Am and 18O. Thus IUPAC recognized the nuclear physics teams at Dubna and Berkeley as the co-discoverers of lawrencium.

[edit] Isotopes

Most actinides have 89 to 103 protons and 117 to 159 neutrons.

Thirty one isotopes of actinium and 8 excited isomeric states of some of its nuclides were identified by 2010.[47] Three isotopes, 225Ac, 227Ac and 228Ac, 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 neptunium series;[50] 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.[22] 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 2×10−10 gram of 227Ac.[22][47] Actinium-228 is a member of radioactive thorium series formed by the decay of 228Ra;[50] it is a β emitter with a half-life of 6.15 hours. In one tonne of thorium there is 5×10−8 gram of 228Ac. It was discovered by Otto Hahn in 1906.[22]

Twenty nine isotopes of protactinium are known with mass numbers 212–240[47] as well as three excited isomeric states. Only 231Pa and 234Pa have been found in nature. All the isotopes have short lifetime, except for protactinium-231 (half-life 32,760 years). The most important isotopes are 231Pa and 233Pa, which is an intermediate product in obtaining uranium-233 and is the most affordable among artificial isotopes of protactinium. 233Pa 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.[47][51]

Uranium has the highest number (25) of both natural and synthetic isotopes. They have mass numbers of 217–242,[48] and three of them, 234U, 235U and 238U, are present in appreciable quantities in nature. Among others, the most important is 233U, which is a final product of transformations of 232Th irradiated by slow neutrons. 233U has a very higher fission efficiency by low-energy (thermal) neutrons, compared e.g. with 235U. Most uranium chemistry studies were carried out on uranium-238 owing to its long half-life of 4.4×109 years.[52]

There are 19 isotopes of neptunium with mass numbers from 225 to 244;[48] they are all highly radioactive. The most popular among scientists are long-lived 237Np (t½ = 2.20×106 years) and short-lived 239Np, 238Np (t½ ~ 2 days).[28]

Sixteen isotopes of americium are known with mass numbers from 232 to 248.[48] The most important are 241Am and 243Am, which are alpha-emitters and also emit soft, but intense γ-rays; both of them can be obtained in an isotopically pure form. Chemical properties of americium were first studied with 241Am, but later shifted to 243Am, which is almost 20 times less radioactive. The disadvantage of 243Am is production of the short-lived daughter isotope 239Np, which has to be considered in the data analysis.[7]

Among 19 isotopes of curium,[48] the most accessible are 242Cm and 244Cm; they are α-emitters, but with much shorter lifetime than the americium isotopes. These isotopes emit almost no γ-radiation, but undergo spontaneous fission with the associated emission of neutrons. More long-lived isotopes of curium (245–248Cm, all α-emitters) are formed as a mixture during neutron irradiation of plutonium or americium. Upon short irradiation, this mixture is dominated by curium-246, and then curium-248 begins to accumulate. Both of these isotopes, especially 248Cm, have a longer half-life (3.48×105 years) and are much more convenient for carrying out chemical research than 242Cm and 244Cm, but they also have a rather high rate of spontaneous fission. 247Cm has the longest lifetime among isotopes of curium (1.56×107 years), but is not formed in large quantities because of the strong fission induced by thermal neutrons.

Fourteen isotopes of berkelium were identified with mass numbers 238–252.[48] Only 249Bk 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 detection. Its alpha radiation is rather weak (1.45×10−3% with respect to β-radiation), but is sometimes used to detect this isotope. 247Bk 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.[7]

Isotopes of californium with mass numbers 237–256 are formed in nuclear reactors;[48] californium-253 is a β-emitter and the rest are α-emitters. The isotopes with even mass numbers (250Cf, 252Cf and 254Cf) have a high rate of spontaneous fission, especially 254Cf 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. 249Cf is not formed in large quantities in a nuclear reactor because of the slow β-decay of the parent isotope 249Bk and a large cross section of interaction with neutrons, but it can be accumulated in the isotopically pure form as the β-decay product of (pre-selected) 249Bk. Californium produced by reactor-irradiation of plutonium mostly consists of 250Cf and 252Cf, the latter being predominant for large neutron fluences, and its study is hindered by the strong neutron radiation.[7]

Properties of some transplutonium isotope pairs[7]
Parent
isotope
t½ Daughter
isotope
t½ Time to establish
radioactive equilibrium
243Am 7370 years 239Np 2.35 days 47.3 days
245Cm 8265 years 241Pu 14 years 129 years
247Cm 1.64×107 years 243Pu 4.95 hours 7.2 days
254Es 270 days 250Bk 3.2 hours 35.2 hours
255Es 39.8 days 255Fm 22 hours 5 days
257Fm 79 days 253Cf 17.6 days 49 days

Among the 16 known isotopes of einsteinium with mass numbers from 241 to 257[48] the most affordable is 253Es. It is an α-emitter with a half-life of 20.47 days, a relatively weak γ-emission and small spontaneous fission rate as compared with the isotopes of californium. Prolonged neutron irradiation also produces a long-lived isotope 254Es (t½ = 275.5 days).[7]

Nineteen isotopes of fermium are known with mass numbers of 242–260. 254Fm, 255Fm and 256Fm are α-emitters with a short half-life (hours), which can be isolated in significant amounts. 257Fm (t½ = 100 days) can accumulate upon prolonged and strong irradiation. All these isotopes are characterized by high rates of spontaneous fission.[7][53]

Among the 15 known isotopes of mendelevium (mass numbers from 245 to 260),[48] the most studied is 256Md, which mainly decays through the electron capture (α-radiation is ≈10%) with the half-life of 77 minutes. Another alpha emitter, 258Md, has a half-life of 53 days. Both these isotopes are produced from rare einsteinium (253Es and 255Es respectively), that limits their so their availability.[47]

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 nobelium and lawrencium were studied with 255No (t½ = 3 min) and 256Lr (t½ = 35 s). The longest-lived nobelium isotope 259No has a half-life of 1.5 hours.[47]

[edit] Distribution in nature

Unprocessed uranium ore

Thorium and uranium are the most abundant actinides in nature with the respective mass concentrations of 1.6×10−3% and 4×10−4%.[54] 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 (KUO2VO4·3H2O) and autunite (Ca(UO2)2(PO4)2·nH2O). The isotopic composition of natural uranium is 238U (relative abundance 99.2742%), 235U (0.7204%) and 234U (0.0054%); of these 238U has the largest half-life of 4.51×109 years.[55][56] The worldwide production of uranium in 2009 amounted to 50,572 tonnes, of which 27.3% was mined in Kazakhstan. Other important uranium mining countries are Canada (20.1%), Australia (15.7%), Namibia (9.1%), Russia (7.0%), and Niger (6.4%).[57]

The most abundant thorium minerals are thorianite (ThO2), thorite (ThSiO4) and monazite, ((Th, Ca, Ce)PO4). Most 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) and Canada (~100,000 tonnes).[58]

Content of plutonium in uranium and thorium ores[59]
Ore Location Uranium
content, %
Mass ratio
239Pu/ore
Ratio
239Pu/U (×1012)
Uraninite Canada 13.5 9.1×10−12 7.1
Uraninite Congo 38 4.8×10−12 12
Uraninite Colorado, US 50 3.8×10−12 7.7
Monazite Brazil 0.24 2.1×10−14 8.3
Monazite North Carolina, US 1.64 5.9×10−14 3.6
Fergusonite - 0.25 <1×10−14 <4
Carnotite - 10 <4×10−14 <0.4

The abundance of actinium in the Earth's crust is only about 5×10−15%.[51] Actinium is mostly present in uranium-containing, but also in other minerals, though in much smaller quantities. The content of actinium in most natural objects corresponds to the isotopic equilibrium of parent isotope 235U, and it is not affected by the weak Ac migration.[22] Protactinium is more abundant (10−12%) in the Earth's crust than actinium. It was discovered in the uranium ore in 1913 by Fajans and Göhring.[24] As actinium, the distribution of protactinium follows that of 235U.[51]

The half-life of the longest-lived isotope of neptunium, 237Np, 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.[28] Traces of plutonium in uranium minerals were first found in 1942, and the more systematic results on 239Pu are summarized in the table (no other plutonium isotopes could be detected in those samples). The upper limit of abundance of the longest-living isotope of plutonium, 244Pu, is 3×10−20%. Plutonium could not be detected in samples of lunar soil. Owing to its scarcity in nature, most plutonium is produced synthetically.[59]

[edit] Extraction

Monazite—a major thorium mineral

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. Fluorides are reduced with calcium, magnesium or barium:[15]

~\mathrm{2AmF_3+3Ba \ \xrightarrow{1150-1350^\circ C} \ 3BaF_2+2Am}
~\mathrm{PuF_4+2Ba \ \xrightarrow{1200^\circ C} \ 2BaF_2+Pu}
~\mathrm{UF_4+2Mg \ \xrightarrow{>500^\circ C} \ U+2MgF_2}

Among the actinides, thorium and uranium are the easiest to isolate. Thorium is extracted mostly from monazite: thorium diphosphate (Th(PO4)2) is reacted with nitric acid, and the produced thorium nitrate treated with tributyl phosphate. Rare-earth impurities are separated by increasing the pH in sulfate solution.

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)4) contaminated with ~3% of rare-earth hydroxides; the rest of rare-earth hydroxides 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:[15]

Separation of uranium and plutonium from nuclear fuel[60]

Th(OH)4 + 4 HNO3 → Th(NO3)4 + 4 H2O

Metallic thorium is separated from the anhydrous oxide, chloride or fluoride by reacting it with calcium in an inert atmosphere:[61]

ThO2 + 2 Ca → 2 CaO + Th

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.[62]

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 into an organic form UO2(NCS)2·2TBP. The insoluble impurities are filtered and the uranium is extracted by reaction with hydroxides as (NH4)2U2O7 or with hydrogen peroxide as UO4·2H2O.[15]

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 carbonate, which converts uranium into a complex [UO2(CO3)3]4–, which is stable in aqueous solutions at low 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 CO32– → 2 [UO2(CO3)3]4–

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.[15]

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, UO2(NO3)2·6H2O. When heated, it turns into UO3, which is converted to UO2 with hydrogen:

UO3 + H2 → UO2 + H2O

Reacting uranium dioxide with fluoric acid changes it to uranium tetrafluoride, which yields uranium metal upon reaction with magnesium metal:[61]

4 HF + UO2 → UF4 + 2 H2O

To extract plutonium, neutron-irradiated uranium is dissolved in nitric acid, and a reducing agent (FeSO4, or H2O2) 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 (UO2(NO3)2). The solution is treated with a reducing agent and neutralized with ammonium carbonate to pH = 8 that results in precipitation of Pu4+ compounds.[15]

In another method, Pu4+and UO22+ are first extracted with tributyl phosphate, then reacted with hydrazine washing out the recovered plutonium.[15]

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.[22]

[edit] Properties

Actinides have similar properties to lanthanides. The 6d and 7s electronic shells are completed 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.[63] As in lanthanides (see lanthanide contraction), the ionic radius of actinides monotonically decreases with atomic number (see also Aufbau principle).[15]

Properties of actinides (the mass of the most long-lived isotope is in square brackets) [56][64]
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] 232.038 [231] 238.029 [237] [244] [243] [247] [247] [251] [252] [257] [258] [259] [262]
Number of natural isotopes 3 1 2 3
Longest-lived isotope 227 232 231 238 237 244 243 247 247 251 252 257 258 259 262
Half-life of the longest-lived isotope 21.8 years 14 billion years 32,500 years 4.47 billion years 2.14 million years 8.2 million years 7,370 years 15.6 million years 1,400 years 900 years 1.29 years 100.5 days 52 days 58 min 261 min
Electronic configuration in the ground state 6d17s2 6d27s2 5f26d17s2or 5f16d27s2 5f36d17s2 5f46d17s2or 5f57s2 5f67s2 5f77s2 5f76d17s2 5f97s2or 5f86d17s2 5f107s2 5f117s2 5f127s2 5f137s2 5f147s2 5f147s27p1
Oxidation state 3 3, 4 3, 4, 5 3, 4, 5, 6 3, 4, 5, 6, 7 3, 4, 5, 6, 7 2, 3, 4 3, 4 3, 4 2, 3 2, 3 2, 3 2, 3 2, 3 3
Metallic radius, nm 0.203 0.180 0.162 0.153 0.150 0.162 0.173 0.174 0.170 0.186 0.186
Ionic radius, nm:
An4+
An3+


0.126

0.114

0.104
0.118

0.103
0.118

0.101
0.116

0.100
0.115

0.099
0.114

0.099
0.112

0.097
0.110

0.096
0.109

0.085
0.098

0.084
0.091

0.084
0.090

0.084
0.095

0.083
0.088
Temperature, °C:
melting
boiling

1050
3300

1750
4800

1572
4400

1130
3800

640
3900

640
3230

1176
2610

1340

1050

900

860

1530

830

830

1630
Density, g/cm3 10.07 11.78 15.37 19.06 20.25 1.84 11.7 7.0 14.78
Standard electrode potential, V:
E° (An4+/An0)
E° (An3+/An0)


−2.13

−1.83

−1.47

−1.38
−1.66

−1.30
−1.79

−1.25
−2.00

−0.90
−2.07

−0.75
−2.06

−0.55
−1.96

−0.59
−1.97

−0.36
−1.98

−0.29
−1.96


−1.74


−1.20

-
−2.10
Color
[M(H2O)n]4+
[M(H2O)n]3+


Colorless

Colorless
Blue

Yellow
Dark blue

Green
Purple

Yellow-green
Purple

Brown
Violet

Red
Rose

Yellow
Colorless

Beige
Yellow-green

Green
Green


Pink








[edit] Physical properties

ActinidesLattice.png ACTIION.PNG
Major crystal structures of some actinides vs. temperature Metallic and ionic radii of actinides[64]
A pellet of 238PuO2 to be used in a radioisotope thermoelectric generator for either the Cassini or Galileo mission. The pellet produces 62 watts of heat and glows because of the heat generated by the radioactive decay (primarily α). Photo is taken after insulating the pellet under a graphite blanket for minutes and removing the blanket.

Actinides are typical metals. All of them are soft and have a silvery color (but tarnish in air),[67] relatively high density and plasticity. Some of them can be cut with a knife. Their electrical resistivity varies between 15 and 150 µOhm·cm.[64] 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, neptunium and californium three. Crystal structures of protactinium, uranium, neptunium and plutonium do not have clear analogs among the lanthanide and are more similar to those of the 3d-transition metals.[56]

All actinides are pyrophoric, especially when finely divided, that is, they spontaneously ignite upon reaction with air.[67] The melting point of 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.[56]

[edit] Chemical properties

Like the lanthanides, all actinides are highly reactive with halogens and chalcogens; however, 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.[15] 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 plutonium and +3 for actinium and other actinides.[15]

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 properties. Among other trivalent actinides Ac3+ is least acidic, i.e. has the weakest tendency to hydrolyze in aqueous solutions.[22][56]

Thorium is rather active chemically. Owing to lack of electrons on 6d-and 4f-orbitals, the tetravalent thorium compounds are colorless. At pH < 3, the solutions of thorium salts are dominated by the cations [Th(H2O)8]4+. The Th4+ 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 a hydrolysis. The distinctive ability of thorium salts is their high solubility, not only in water, but also in polar organic solvents.[56]

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 UIV and thorium. Fluorides, phosphates, hypophosphate, iodate and phenylarsonates of PaIV 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.[51]

Uranium has a valence from 3 to 6, the last being most stable. In the hexavalent state, uranium is very similar to the sixth group elements. Many compounds of uranium(IV) and uranium(VI) are nonstoichiometric, i.e. have variable composition. For example, the actual chemical formula of uranium dioxide is UO2+x, where x varies between −0.4 and 0.32. Uranium(VI) compounds are weak oxidants. Most of them contain the linear "uranyl" group, UO22+. 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. NpO22+ and PuO22+ 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 as UIII(C5H5)3 and UIV(C5H5)4.[56][68]

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.[28]

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.[69][70]

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 to be erroneous. The most stable valence of americium is 3 in the aqueous solutions and 3 or 4 in solid compounds.[7]

Valence 3 is dominant in all subsequent elements up to lawrencium (maybe except for 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 of Bk4+ in aqueous solution is close to that of Ce4+. Only valence 3 was observed for californium, einsteinium and fermium. Bivalent 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. Rutherfordium chloride behaves similar to hafnium chloride, i.e. might contain tetravalent rutherfordium.[7]

The redox potential ~E_\frac{M^{4+}}{AnO^{2+}_{2}} increases from −0.32 V in uranium, through 0.34 V (Np) and 1.04 V (Pu) to 1.34 V in americium revealing the increasing reduction ability of the An4+ ion from americium to uranium. All actinides form AnH3 hydrides of black color with salt-like properties. Actinides also produce carbides with the general formula of AnC or AnC2 (U2C3 for uranium) as well as sulfides An2S3 and AnS2.[15]

[edit] Compounds

[edit] Oxides and hydroxides

An – actinide
**Depending on the isotopes

Some actinides can exists in various oxide forms: An2O3, AnO2, An2O5 and AnO3. For all actinides, oxides AnO3 are amphoteric and An2O3, AnO2 and An2O5 are basic, they easily react with water, forming bases:[15]

An2O3 + 3 H2O → 2 An(OH)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 (Ac2S3).[15] Dioxides of tetravalent actinides crystallize in the cubic system, same as in calcium fluoride.

Thorium reacting with oxygen exclusively forms dioxide:

~\mathrm{Th+O_2 \ \xrightarrow{1000^\circ C} \ ThO_2}

Thorium dioxide is a refractory material with the highest melting point among any known oxide (3390 °C).[74] Adding 0.8–1% ThO2to tungsten stabilizes its structure, so the doped filaments have better mechanical stability to vibrations. To dissolve ThO2acids, it is heated to 500–600 °C; heating above 600 °C produces a very resistant to acids and other reagents form of ThO2. Small addition of fluoride ions catalyses dissolution of thorium dioxide in acids.

Two protactinium oxides were obtained: PaO2 (black) and Pa2O5(white); the former is isomorphic with ThO2 and the latter is easier to obtain. Both oxides are basic, and Pa(OH)5 is a weak, poorly soluble base.[15]

Decomposition of certain salts of uranium, for example UO2(NO3)·6H2O in air at 400 °C, yields orange or yellow UO3.[74] This oxide is amphoteric and forms several hydroxides, the most stable being UO2(OH)2. Reaction of uranium(VI) oxide with hydrogen results in uranium dioxide, which is similar in its properties with ThO2. This oxide is also basic and corresponds to the uranium hydroxide (U(OH)4).[15]

Plutonium, neptunium and americium form two basic oxides: An2O3 and AnO2. Neptunium trioxide is unstable, thus only Np3O8 could be obtained so far. However, the oxides of plutonium and neptunium with the chemical formula AnO2 and An2O3 are well characterized.[15]

[edit] Salts

*An – actinide
**Depending on the isotopes
Einsteinium triiodide glowing in the dark

Actinides easily react with halogens forming salts with the formulas MX3 and MX4 (X = halogen). So the first berkelium compound, BkCl3, was synthesized in 1962 with an amount of 3 nanogram. 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 AnO2F2.[78] Pentachloride and black hexachloride of uranium were synthesized, but they are both unstable.[15]

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 + 2 H2SO4 → U (SO4)2 + 2 H2
2 Pu + 6 HCl → 2 PuCl3 + 3 H2

However, in these reactions the regenerating hydrogen can react with the metal, forming the corresponding hydride. Uranium reacts with acids and water much easier than thorium.[15]

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(NO3)4·6H2O, Th(SO4)2·9H2O and Pu2(SO4)3·7H2O. Salts of high-valence actinides easily hydrolyze. So, colorless sulfate, chloride, perchlorate and nitrate of thorium transform into basic salts with formulas Th(OH)2SOsub>4 and Th(OH)3NO3. The solubility and insolubility of trivalent and tetravalent actinides is like that of lanthanide salts. So phosphates, fluorides, oxalates, iodates and carbonates of actinides are weakly soluble in water; they precipitate as hydrates, such as ThF4·3H2O and Th(CrO4)2·3H2O.[15]

Actinides with oxidation state +6, except for the AnO22+-type cations, form [AnO4]2–, [An2O7]2– and other complex anions. For example, uranium, neptunium and plutonium form salts of the Na2UO4 (uranate) and (NH4)2U2O7 (diuranate) types. In comparison with lanthanides, actinides easier form coordination compounds, and this ability increases with the actinide valence. Trivalent actinides do not form fluoride coordination compounds, whereas tetravalent thorium forms K2ThF6, KThF5, and even K5ThF9 complexes. Thorium also forms the corresponding sulfates (for example Na2SO4·Th (SO4)2·5H2O), nitrates and thiocyanates. Salts with the general formula An2Th(NO3)6·nH2O are of coordination nature, with the coordination number of thorium equal to 12. Even 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.[15]

[edit] Applications

Inside of a smoke detector containing americium-241.

While actinides have some established daily-life applications, such as in smoke detectors (americium)[79][80] and gas mantles (thorium),[61] they are mostly used in nuclear weapons and use as a fuel in nuclear reactors.[61] 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.

Self-illumination of a nuclear reactor by the Cherenkov radiation.

The most important isotope for nuclear energy applications is uranium-235. It is used in the most popular thermal reactor, and its concentration in natural uranium does not exceed 0.72%. This isotope strongly absorbs thermal neutrons releasing much energy. One fission act of 1 gram of 235U converts into about 1 MW·day. Important is that 235U emits more neutrons than it absorbs.[15] Upon reaching the critical mass, 235U enters into a self-sustaining chain reaction.[56] Typically, uranium nucleus is divided into two fragments with the release of 2–3 neutrons, for example:

~\mathrm{{}_{~}^{235}U+ \ ^1_0n \longrightarrow {}_{~}^{115}Rh+{}_{~}^{118}Ag + 3^1_0n}

Other promising actinide isotopes for nuclear energy are thorium-232 and its fission product, formed uranium-233.

Nuclear reactor [56][81][82]
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.
Nuclear reactor scheme

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.[61] 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.[15] The relative content of thorium and uranium isotopes is widely used to estimate the age of various objects, including stars (see radiometric dating).[83]

Plutonium-238 is potentially more efficient isotope for nuclear reactors, as it has smaller critical mass than uranium-235, but releases much more thermal energy (0.56 W/g).[80][84] However, 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 spacecrafts (e.g. Apollo 14[85]) had heaters powered by kilogram quantities of plutonium-238 oxide; this heat is also transformed into electricity with thermopiles. The decay of plutonium-238 is produced relatively harmless alpha particles and is not accompanied by gamma-irradiation. Therefore and this isotope (~160 mg) is used as the energy source in heart pacemakers where it lasts about 5 times longer than conventional batteries.[80]

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 for remote use. 228Ac is used as an indicator of radioactivity in chemical research, as it emits high-energy electrons (2.18 MeV) that can be easily detected. 228Ac-228Ra mixtures are widely used as an intense gamma-source in industry and medicine.[22]

[edit] Toxicity

Schematic illustration of penetration of radiation through sheets of paper, aluminium and lead brick
Periodic table with elements colored according to the half-life of their most stable isotope.
  Stable elements;
  Radioactive elements with very long-lived isotopes. Their half-lives of over four million years confer them very small, if not negligible radioactivities;
  Radioactive elements that may present low health hazards. Their most stable isotopes have half-lives between 800 and 34,000 years. Because of this, they usually have some commercial applications;
  Radioactive elements that are known to pose high safety risks. Their most stable isotopes have half-lives between one day and 103 years. Their radioactivities confers them little potential for commercial uses;
  Highly radioactive elements. Their most stable isotopes have half-lives between one day and several minutes. They pose severe health risks. Few of them receive uses outside basic research;
  Extremely radioactive elements. Very little is known about these elements due to their extreme instability and radioactivity.

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 and γ-radiation. Together with radium and transuranium elements, actinium is one of the most dangerous radioactive 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.[22]

Protactinium in the body tends to accumulate in the kidneys and bones. The maximum safe dose of Pa in the human body is 0.03 µCi that corresponds to 0.5 micrograms of 231Pa. This isotope, which might be present in the air as aerosol, is 2.5×108 times more toxic than hydrocyanic acid.[51]

Plutonium, when entering the body through air, food or blood (e.g. a 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 (LD50) for 30 days in dogs after 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 4 times greater. Another estimate assumes that plutonium is 50 times less toxic than radium, and thus permissible 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 Pu 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, Pu 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.[69]

[edit] See also

[edit] References and notes

  1. ^ The Manhattan Project
  2. ^ Actinide element, Encyclopedia Britannica on-line
  3. ^ Connelly, Neil G.; et al. (2005). "Elements". Nomenclature of Inorganic Chemistry. London: Royal Society of Chemistry. p. 52. ISBN 0854044388. http://books.google.com/books?id=w1Kf1CakyZIC&pg=PA52. 
  4. ^ a b c Greenwood, p. 1250
  5. ^ a b c Greenwood, p. 1252
  6. ^ Nobelium and lawrencium were almost simultaneously discovered by Soviet and American scientists
  7. ^ a b c d e f g h i j k l m B. Myasoedov et al. (1972). Analytical chemistry of transplutonium elements. Moscow: Nauka. ISBN 0470627158. 
  8. ^ "О возможном нахождении элемента с атомным номером выше 92 (On the possible finding of the element with atomic number greater than 92)". Uspekhi Fizicheskikh Nauk 14 (7): 829. 1934. http://ufn.ru/ru/articles/1934/7/b/. 
  9. ^ P. Kudryavtsev (1982). "Опыты Ферми". The course of the history of physics. Moscow. p. 73. http://nplit.ru/books/item/f00/s00/z0000004/st070.shtml. 
  10. ^ "О периодическом законе и таблице" (in Russian). http://www.xenoid.ru/materials/materials_chem/history/pertable.php. Retrieved 2010-07-16. 
  11. ^ M. H. Klaproth (1789). "Chemische Untersuchung des Uranits, einer neuentdeckten metallischen Substanz". Chemische Annalen 2: 387–403. 
  12. ^ E.-M. Péligot (1842). "Recherches Sur L'Uranium". Annales de chimie et de physique 5 (5): 5–47. http://gallica.bnf.fr/ark:/12148/bpt6k34746s/f4.table. 
  13. ^ Ingmar Grenthe et al. (2006). "Uranium". The Chemistry of the Actinide and Transactinide Elements. doi:10.1007/1-4020-3598-5_5. 
  14. ^ Zimmerman, Ann., 213, 290 (1882); 216, 1 (1883); Ber. 15 (1882) 849
  15. ^ a b c d e f g h i j k l m n o p q r s t u v w x A. M. Golub (1971). Общая и неорганическая химия (General and Inorganic Chemistry). 2. 
  16. ^ Berzelius, J. J. (1829). "Untersuchung eines neues Minerals und einer darin erhalten zuvor unbekannten Erde (Investigation of a new mineral and of a previously unknown earth contained therein)". Annalen der Physik und Chemie 16: 385–415. http://gallica.bnf.fr/ark:/12148/bpt6k151010.pleinepage.r=Annalen+der+Physic.f395.langFR.  (modern citation: Annalen der Physik, vol. 92, no. 7, pages 385–415)
  17. ^ Berzelius, J. J. (1829). "Undersökning af ett nytt mineral (Thorit), som innehåller en förut obekant jord" (Investigation of a new mineral (thorite), as contained in a previously unknown earth)". Kungliga Svenska Vetenskaps Akademiens Handlingar (Transactions of the Royal Swedish Science Academy): 1–30. 
  18. ^ Debierne, André-Louis (1899). "Sur un nouvelle matière radio-active" (in French). Comptes rendus 129: 593–595. http://gallica.bnf.fr/ark:/12148/bpt6k3085b/f593.table. 
  19. ^ Debierne, André-Louis (1900–1901). "Sur un nouvelle matière radio-actif – l'actinium" (in French). Comptes rendus 130: 906–908. http://gallica.bnf.fr/ark:/12148/bpt6k3086n/f906.table. 
  20. ^ Kirby, H. W. (1971). "The Discovery of Actinium". Isis 62 (3): 290–308. doi:10.1086/350760. http://www.jstor.org/stable/view/229943?seq=1. 
  21. ^ Adloff, J. P. (2000). "The centenary of a controversial discovery: actinium". Radiochim. Acta, 88: 123–128. doi:10.1524/ract.2000.88.3-4.123. 
  22. ^ a b c d e f g h i j Z. K. Karalova, B. Myasoedov (1982). Actinium. Analytical chemistry items. Moscow: Nauka. 
  23. ^ Emsley, John (2003-08-11). "Protactinium". Nature's Building Blocks: An A-Z Guide to the Elements. Oxford, England, UK: Oxford University Press. pp. 347–349. ISBN 0198503407. http://books.google.com/?id=j-Xu07p3cKwC&pg=PA348. 
  24. ^ a b K. Fajans and 0. Gohring, (1913). "Über die komplexe Natur des Ur X". Naturwissenschaften 14: 339. doi:10.1007/BF01495360. http://www.digizeitschriften.de/no_cache/home/jkdigitools/loader/?tx_jkDigiTools_pi1%5BIDDOC%5D=201162&tx_jkDigiTools_pi1%5Bpp%5D=425. 
  25. ^ K. Fajans and 0. Gohring, (1913). "Über das Uran X2-das neue Element der Uranreihe". Physikalische Zeitschrift 14: 877–84. 
  26. ^ a b Greenwood, p. 1251
  27. ^ Mcmillan, Edwin; Abelson, Philip (1940). "Radioactive Element 93". Physical Review 57: 1185. doi:10.1103/PhysRev.57.1185.2. 
  28. ^ a b c d e f V.A. Mikhailov, ed (1971). Analytical chemistry of neptunium. Moscow: Nauka. 
  29. ^ E. Fermi (1946). "The Development of the first chain reaction pile". Proceedings of the American Philosophical Society 90 (1): 20–24. http://www.jstor.org/pss/3301034. 
  30. ^ Hall, Nina (2000). The New Chemistry: A Showcase for Modern Chemistry and Its Applications. Cambridge University Press. pp. 8–9. ISBN 9780521452243. http://books.google.com/?id=U4rnzH9QbT4C. 
  31. ^ Thompson, S. G.; Ghiorso, A.; Seaborg, G. T. (1950). "Element 97". Phys. Rev. 77 (6): 838–39. doi:10.1103/PhysRev.77.838.2. 
  32. ^ Thompson, S. G.; Ghiorso, A.; Seaborg, G. T. (1950). "The New Element Berkelium (Atomic Number 97)". Phys. Rev. 80: 781–89. doi:10.1103/PhysRev.80.781. 
  33. ^ Wallace W. Schulz (1976) The Chemistry of Americium, U. S. Department of Commerce, p. 1
  34. ^ Thompson, S.; Ghiorso, A.; Seaborg, G. (1950). "Element 97". Physical Review 77: 838. doi:10.1103/PhysRev.77.838.2. 
  35. ^ Thompson, S.; Ghiorso, A.; Seaborg, G. (1950). "The New Element Berkelium (Atomic Number 97)". Physical Review 80: 781. doi:10.1103/PhysRev.80.781. 
  36. ^ S. G. Thompson, K. Street, Jr., A. Ghiorso, G. T. Seaborg (1950). "Element 98". Physical Review 78: 298. doi:10.1103/PhysRev.78.298.2. http://repositories.cdlib.org/cgi/viewcontent.cgi?article=7072&context=lbnl. 
  37. ^ S. G. Thompson, K. Street, Jr., A. Ghiorso, G. T. Seaborg (1950). "The New Element Californium (Atomic Number 98)". Physical Review 80: 790. doi:10.1103/PhysRev.80.790. http://www.osti.gov/accomplishments/documents/fullText/ACC0050.pdf. 
  38. ^ K. Street, Jr., S. G. Thompson, G. T. Seaborg (1950). "Chemical Properties of Californium". J. Am. Chem. Soc. 72: 4832. doi:10.1021/ja01166a528. http://handle.dtic.mil/100.2/ADA319899. 
  39. ^ 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
  40. ^ Darleane C. Hoffman, Albert Ghiorso, Glenn Theodore Seaborg The transuranium people: the inside story, Imperial College Press, 2000 ISBN 1860940870, pp. 141–142
  41. ^ a b A. Ghiorso, S. G. Thompson, G. H. Higgins, G. T. Seaborg, M. H. Studier, P. R. Fields, S. M. Fried, H. Diamond, J. F. Mech, G. L. Pyle, J. R. Huizenga, A. Hirsch, W. M. Manning, C. I. Browne, H. L. Smith, R. W. Spence (1955). "New Elements Einsteinium and Fermium, Atomic Numbers 99 and 100". Phys. Rev. 99 (3): 1048–1049. doi:10.1103/PhysRev.99.1048. 
  42. ^ Fields, P.; Studier, M.; Diamond, H.; Mech, J.; Inghram, M.; Pyle, G.; Stevens, C.; Fried, S. et al. (1956). "Transplutonium Elements in Thermonuclear Test Debris". Physical Review 102: 180. doi:10.1103/PhysRev.102.180. 
  43. ^ Thompson, S. G. and Ghiorso, A.; Harvey, B. G.; Choppin, G. R. (1954). "Transcurium Isotopes Produced in the Neutron Irradiation of Plutonium". Physical Review 93: 908. doi:10.1103/PhysRev.93.908. 
  44. ^ Choppin, G. R.; Thompson, S. G.; Ghiorso, A.; Harvey, B. G. (1954). "Nuclear Properties of Some Isotopes of Californium, Elements 99 and 100". Physical Review 94 (4): 1080–1081. doi:10.1103/PhysRev.94.1080. 
  45. ^ Ghiorso, Albert (2003). "Einsteinium and Fermium". Chemical and Engineering News 81 (36). http://pubs.acs.org/cen/80th/einsteiniumfermium.html. 
  46. ^ Ghiorso, A.; Harvey, B.; Choppin, G.; Thompson, S.; Seaborg, G. (1955). "New Element Mendelevium, Atomic Number 101". Physical Review 98: 1518. doi:10.1103/PhysRev.98.1518. ISBN 9789810214401. http://books.google.com/books?id=e53sNAOXrdMC&pg=PA101. 
  47. ^ a b c d e f g h G. Audi, O. Bersillon, J. Blachot, A.H. Wapstra (2003). "The Nubase evaluation of nuclear and decay properties". Nuclear Physics 624: 3–128. doi:10.1016/S0375-9474(97)00482-X. http://amdc.in2p3.fr/nubase/Nubase2003.pdf. 
  48. ^ a b c d e f g h i "Table of nuclides, IAEA". http://www-nds.iaea.org/relnsd/vcharthtml/VChartHTML.html. Retrieved 2010-07-07. 
  49. ^ Specific activity is calculated by given in the table half-lives and the probability of spontaneous fission
  50. ^ a b Greenwood, p. 1254
  51. ^ a b c d e f g E.S. Palshin et al. (1968). Analytical chemistry of protactinium. Moscow: Nauka. 
  52. ^ I.P. Alimarin et al. (1962). A.P. Vinogradov. ed. Analytical chemistry of uranium. Moscow: Publisher USSR Academy of Sciences. 
  53. ^ "Table of elements, compounds, isotopes" (in Russian). http://elm.e-science.ru/. Retrieved 2010-07-07. 
  54. ^ Jay H. Lehr, Janet K. Lehr (2000). Standard handbook of environmental science, health, and technology. McGraw-Hill Professional. pp. 2–38. ISBN 007038309X. http://books.google.com/books?id=w0wa4b9CGkcC&pg=SA2-PA38. 
  55. ^ Lide, D. R., ed. (2005), CRC Handbook of Chemistry and Physics (86th ed.), Boca Raton (FL): CRC Press, ISBN 0-8493-0486-5 
  56. ^ a b c d e f g h i Yu.D. Tretyakov, ed (2007). Non-organic chemistry in three volumes. Chemistry of transition elements. 3. Moscow: Academy. ISBN 5-7695-2533-9. 
  57. ^ "World Uranium Mining". World Nuclear Association. http://www.world-nuclear.org/info/inf23.html. Retrieved 2010-06-11. 
  58. ^ Thorium, USGS Mineral Commodities
  59. ^ a b c F. Weigel, J. Katz, G. Seaborg et al. (1997). The Chemistry of the Actinide Elements. 2. Moscow: Mir. ISBN 5-03-001885-9. 
  60. ^ Greenwood, pp. 1255, 1261
  61. ^ a b c d e Greenwood, p. 1255
  62. ^ van Arkel, A. E.; de Boer, J. H. (1925). "Darstellung von reinem Titanium-, Zirkonium-, Hafnium- und Thoriummetall" (in German). Zeitschrift für anorganische und allgemeine Chemie 148 (1): 345–350. doi:10.1002/zaac.19251480133. 
  63. ^ I.L. Knunyants et al. (1961). Short Chemical Encyclopedia. 1. Moscow: Soviet Encyclopedia. 
  64. ^ a b c Greenwood, p. 1263
  65. ^ Greenwood, p. 1265
  66. ^ dtv-Atlas zur Chemie 1981, Teil 1, S. 224
  67. ^ a b Greenwood, p. 1264
  68. ^ Greenwood, p. 1278
  69. ^ a b B.A. Nadykto and L.F. Timofeeva, ed (2003). Plutonium. Fundamental Problems. 1. Sarov: VNIIEF. ISBN 5-9515-00-24-9. 
  70. ^ M. S. Milyukova et al. (1965). Analytical chemistry of plutonium. Moscow: Nauka. ISBN 0250399180. 
  71. ^ a b "Таблица Inorganic and Coordination compounds" (in Russian). http://chemanalytica.com/book/novyy_spravochnik_khimika_i_tekhnologa/01_osnovnye_svoystva_neorganicheskikh_organicheskikh_i_elementoorganicheskikh_soedineniy. Retrieved 2010-07-11. 
  72. ^ According to other sources, cubic sesquioxide of curium is olive-green. See "Соединения curium site XuMuK.ru" (in Russian). http://www.xumuk.ru/encyklopedia/2248.html. Retrieved 2010-07-11. 
  73. ^ 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).
  74. ^ a b c Greenwood, p. 1268
  75. ^ a b "Information from webelements.com". http://www.webelements.com/. 
  76. ^ a b Greenwood, p. 1270
  77. ^ Nave, S.; Haire, R.; Huray, Paul (1983). "Magnetic properties of actinide elements having the 5f^{6} and 5f^{7} electronic configurations". Physical Review B 28: 2317. doi:10.1103/PhysRevB.28.2317. 
  78. ^ Greenwood, p.1269
  79. ^ Smoke Detectors and Americium, Nuclear Issues Briefing Paper 35, May 2002
  80. ^ a b c Greenwood, p. 1262
  81. ^ GG Bartolomei, Baybakov VD, Alkhutov MS, Bach G. A. (1982). Basic theories and methods of calculation of nuclear reactors. Moscow: Energoatomizdat. 
  82. ^ Greenwood, pp. 1256–1261
  83. ^ Sergey Popov, Alexander Sergeev (2008). "Universal Alchemy" (in Russian). Vokrug Sveta 2811 (4). http://www.vokrugsveta.ru/vs/article/6214/. 
  84. ^ Nuclear Weapons Design & Materials
  85. ^ Apollo 14 Press Kit – 01/11/71, NASA, pp. 38–39

[edit] Bibliography

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

[edit] Further reading

  • Tamer Andrea and Moris S. Eisen (2008). "Recent advances in organothorium and organouranium catalysis". Chem. Soc. Rev. 37: 550–567. doi:10.1039/b614969n. 
  • Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean, eds (2006). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer. ISBN 13978-1-4020-3555-5. 

[edit] External links

Personal tools
  • Log in / create account
Namespaces
Variants
Actions

Leave a Reply