Astatine

Last updated

Astatine, 85At
Astatine
Pronunciation /ˈæstətn,-tɪn/ (ASS-tə-teen, -tin)
Appearanceunknown, probably metallic
Mass number [210]
Astatine in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson
 I 

At

Ts
poloniumastatineradon
Atomic number (Z)85
Group group 17 (halogens)
Period period 6
Block   p-block
Electron configuration [ Xe ] 4f14 5d10 6s2 6p5
Electrons per shell2, 8, 18, 32, 18, 7
Physical properties
Phase at  STP solid (predicted)
Density (near  r.t.)8.91–8.95 g/cm3(estimated) [1]
Molar volume 23.6 cm3/mol(estimated) [1]
Atomic properties
Oxidation states −1, +1, +3, +5, +7 [2]
Ionization energies
  • 1st: 899.003 kJ/mol [3]
Other properties
Natural occurrence from decay
Crystal structure face-centered cubic (fcc)
Cubic-face-centered.svg

(predicted) [4]
CAS Number 7440-68-8
History
Namingfrom Ancient Greek ἄστατος (ástatos) 'unstable'
Discovery Dale R. Corson, Kenneth Ross MacKenzie, Emilio Segrè (1940)
Isotopes of astatine
Main isotopes [5] Decay
abun­dance half-life (t1/2) mode pro­duct
209At synth 5.41 h β+ 209Po
α 205Bi
210Atsynth8.1 hβ+ 210Po
α 206Bi
211Atsynth7.21 h ε 211Po
α 207Bi
Symbol category class.svg  Category: Astatine
| references

Astatine is a chemical element; it has symbol At and atomic number 85. It is the rarest naturally occurring element in the Earth's crust, occurring only as the decay product of various heavier elements. All of astatine's isotopes are short-lived; the most stable is astatine-210, with a half-life of 8.1 hours. Consequently, a solid sample of the element has never been seen, because any macroscopic specimen would be immediately vaporized by the heat of its radioactivity.

Contents

The bulk properties of astatine are not known with certainty. Many of them have been estimated from its position on the periodic table as a heavier analog of fluorine, chlorine, bromine, and iodine, the four stable halogens. However, astatine also falls roughly along the dividing line between metals and nonmetals, and some metallic behavior has also been observed and predicted for it. Astatine is likely to have a dark or lustrous appearance and may be a semiconductor or possibly a metal. Chemically, several anionic species of astatine are known and most of its compounds resemble those of iodine, but it also sometimes displays metallic characteristics and shows some similarities to silver.

The first synthesis of astatine was in 1940 by Dale R. Corson, Kenneth Ross MacKenzie, and Emilio G. Segrè at the University of California, Berkeley. They named it from the Ancient Greek ἄστατος (astatos) 'unstable'. [6] Four isotopes of astatine were subsequently found to be naturally occurring, although much less than one gram is present at any given time in the Earth's crust. Neither the most stable isotope, astatine-210, nor the medically useful astatine-211 occur naturally; they are usually produced by bombarding bismuth-209 with alpha particles.

Characteristics

Astatine is an extremely radioactive element; all its isotopes have half-lives of 8.1 hours or less, decaying into other astatine isotopes, bismuth, polonium, or radon. Most of its isotopes are very unstable, with half-lives of seconds or less. Of the first 101 elements in the periodic table, only francium is less stable, and all the astatine isotopes more stable than the longest-lived francium isotopes (205–211At) are in any case synthetic and do not occur in nature. [7]

The bulk properties of astatine are not known with any certainty. [8] Research is limited by its short half-life, which prevents the creation of weighable quantities. [9] A visible piece of astatine would immediately vaporize itself because of the heat generated by its intense radioactivity. [10] It remains to be seen if, with sufficient cooling, a macroscopic quantity of astatine could be deposited as a thin film. [4] Astatine is usually classified as either a nonmetal or a metalloid; [11] [12] metal formation has also been predicted. [4] [13]

Physical

Most of the physical properties of astatine have been estimated (by interpolation or extrapolation), using theoretically or empirically derived methods. [14] For example, halogens get darker with increasing atomic weight fluorine is nearly colorless, chlorine is yellow-green, bromine is red-brown, and iodine is dark gray/violet. Astatine is sometimes described as probably being a black solid (assuming it follows this trend), or as having a metallic appearance (if it is a metalloid or a metal). [15] [16] [17]

Astatine sublimes less readily than iodine, having a lower vapor pressure. [9] Even so, half of a given quantity of astatine will vaporize in approximately an hour if put on a clean glass surface at room temperature. [lower-alpha 1] The absorption spectrum of astatine in the middle ultraviolet region has lines at 224.401 and 216.225 nm, suggestive of 6p to 7s transitions. [19] [20]

The structure of solid astatine is unknown. [21] As an analog of iodine it may have an orthorhombic crystalline structure composed of diatomic astatine molecules, and be a semiconductor (with a band gap of 0.7  eV). [22] [23] Alternatively, if condensed astatine forms a metallic phase, as has been predicted, it may have a monatomic face-centered cubic structure; in this structure, it may well be a superconductor, like the similar high-pressure phase of iodine. [4] Metallic astatine is expected to have a density of 8.91–8.95 g/cm3. [1]

Evidence for (or against) the existence of diatomic astatine (At2) is sparse and inconclusive. [24] [25] [26] [27] [28] Some sources state that it does not exist, or at least has never been observed, [29] [30] while other sources assert or imply its existence. [31] [32] [33] Despite this controversy, many properties of diatomic astatine have been predicted; [34] for example, its bond length would be 300±10  pm , dissociation energy 83.7±12.5 kJ/mol, [35] and heat of vaporization (∆Hvap) 54.39 kJ/mol. [36] Many values have been predicted for the melting and boiling points of astatine, but only for At2. [37]

Chemical

The chemistry of astatine is "clouded by the extremely low concentrations at which astatine experiments have been conducted, and the possibility of reactions with impurities, walls and filters, or radioactivity by-products, and other unwanted nano-scale interactions". [22] Many of its apparent chemical properties have been observed using tracer studies on extremely dilute astatine solutions, [33] [38] typically less than 10−10 mol·L−1. [39] Some properties, such as anion formation, align with other halogens. [9] Astatine has some metallic characteristics as well, such as plating onto a cathode, [lower-alpha 2] and coprecipitating with metal sulfides in hydrochloric acid. [41] It forms complexes with EDTA, a metal chelating agent, [42] and is capable of acting as a metal in antibody radiolabeling; in some respects, astatine in the +1 state is akin to silver in the same state. Most of the organic chemistry of astatine is, however, analogous to that of iodine. [43] It has been suggested that astatine can form a stable monatomic cation in aqueous solution. [41] [44]

Astatine has an electronegativity of 2.2 on the revised Pauling scale  – lower than that of iodine (2.66) and the same as hydrogen. In hydrogen astatide (HAt), the negative charge is predicted to be on the hydrogen atom, implying that this compound could be referred to as astatine hydride according to certain nomenclatures. [45] [46] [47] [48] That would be consistent with the electronegativity of astatine on the Allred–Rochow scale (1.9) being less than that of hydrogen (2.2). [49] [lower-alpha 3] However, official IUPAC stoichiometric nomenclature is based on an idealized convention of determining the relative electronegativities of the elements by the mere virtue of their position within the periodic table. According to this convention, astatine is handled as though it is more electronegative than hydrogen, irrespective of its true electronegativity. The electron affinity of astatine, at 233 kJ mol−1, is 21% less than that of iodine. [51] In comparison, the value of Cl (349) is 6.4% higher than F (328); Br (325) is 6.9% less than Cl; and I (295) is 9.2% less than Br. The marked reduction for At was predicted as being due to spin–orbit interactions. [39] The first ionization energy of astatine is about 899 kJ mol−1, which continues the trend of decreasing first ionization energies down the halogen group (fluorine, 1681; chlorine, 1251; bromine, 1140; iodine, 1008). [3]

Compounds

Less reactive than iodine, astatine is the least reactive of the halogens; [52] the chemical properties of tennessine, the next-heavier group 17 element, have not yet been investigated, however. [53] Astatine compounds have been synthesized in nano-scale amounts and studied as intensively as possible before their radioactive disintegration. The reactions involved have been typically tested with dilute solutions of astatine mixed with larger amounts of iodine. Acting as a carrier, the iodine ensures there is sufficient material for laboratory techniques (such as filtration and precipitation) to work. [54] [55] [lower-alpha 4] Like iodine, astatine has been shown to adopt odd-numbered oxidation states ranging from −1 to +7. [58]

Only a few compounds with metals have been reported, in the form of astatides of sodium, [10] palladium, silver, thallium, and lead. [59] Some characteristic properties of silver and sodium astatide, and the other hypothetical alkali and alkaline earth astatides, have been estimated by extrapolation from other metal halides. [60]

Hydrogen astatide space-filling model Hydrogen-astatide-calculated-3D-sf.svg
Hydrogen astatide space-filling model

The formation of an astatine compound with hydrogen – usually referred to as hydrogen astatide  – was noted by the pioneers of astatine chemistry. [61] As mentioned, there are grounds for instead referring to this compound as astatine hydride. It is easily oxidized; acidification by dilute nitric acid gives the At0 or At+ forms, and the subsequent addition of silver(I) may only partially, at best, precipitate astatine as silver(I) astatide (AgAt). Iodine, in contrast, is not oxidized, and precipitates readily as silver(I) iodide. [9] [62]

Astatine is known to bind to boron, [63] carbon, and nitrogen. [64] Various boron cage compounds have been prepared with At–B bonds, these being more stable than At–C bonds. [65] Astatine can replace a hydrogen atom in benzene to form astatobenzene C6H5At; this may be oxidized to C6H5AtCl2 by chlorine. By treating this compound with an alkaline solution of hypochlorite, C6H5AtO2 can be produced. [66] The dipyridine-astatine(I) cation, [At(C5H5N)2]+, forms ionic compounds with perchlorate [64] (a non-coordinating anion [67] ) and with nitrate, [At(C5H5N)2]NO3. [64] This cation exists as a coordination complex in which two dative covalent bonds separately link the astatine(I) centre with each of the pyridine rings via their nitrogen atoms. [64]

With oxygen, there is evidence of the species AtO and AtO+ in aqueous solution, formed by the reaction of astatine with an oxidant such as elemental bromine or (in the last case) by sodium persulfate in a solution of perchloric acid. [9] [68] The species previously thought to be AtO2 has since been determined to be AtO(OH)2, a hydrolysis product of AtO+ (another such hydrolysis product being AtOOH). [69] The well characterized AtO3 anion can be obtained by, for example, the oxidation of astatine with potassium hypochlorite in a solution of potassium hydroxide. [66] [70] Preparation of lanthanum triastatate La(AtO3)3, following the oxidation of astatine by a hot Na2S2O8 solution, has been reported. [71] Further oxidation of AtO3, such as by xenon difluoride (in a hot alkaline solution) or periodate (in a neutral or alkaline solution), yields the perastatate ion AtO4; this is only stable in neutral or alkaline solutions. [72] Astatine is also thought to be capable of forming cations in salts with oxyanions such as iodate or dichromate; this is based on the observation that, in acidic solutions, monovalent or intermediate positive states of astatine coprecipitate with the insoluble salts of metal cations such as silver(I) iodate or thallium(I) dichromate. [66] [73]

Astatine may form bonds to the other chalcogens; these include S7At+ and At(CSN)2 with sulfur, a coordination selenourea compound with selenium, and an astatine–tellurium colloid with tellurium. [74]

Structure of astatine monoiodide, one of the astatine interhalogens and the heaviest known diatomic interhalogen Astatine-iodide-3D-vdW.svg
Structure of astatine monoiodide, one of the astatine interhalogens and the heaviest known diatomic interhalogen

Astatine is known to react with its lighter homologs iodine, bromine, and chlorine in the vapor state; these reactions produce diatomic interhalogen compounds with formulas AtI, AtBr, and AtCl. [56] The first two compounds may also be produced in water – astatine reacts with iodine/iodide solution to form AtI, whereas AtBr requires (aside from astatine) an iodine/iodine monobromide/bromide solution. The excess of iodides or bromides may lead to AtBr2 and AtI2 ions, [56] or in a chloride solution, they may produce species like AtCl2 or AtBrCl via equilibrium reactions with the chlorides. [57] Oxidation of the element with dichromate (in nitric acid solution) showed that adding chloride turned the astatine into a molecule likely to be either AtCl or AtOCl. Similarly, AtOCl2 or AtCl2 may be produced. [56] The polyhalides PdAtI2, CsAtI2, TlAtI2, [75] [76] [77] and PbAtI [78] are known or presumed to have been precipitated. In a plasma ion source mass spectrometer, the ions [AtI]+, [AtBr]+, and [AtCl]+ have been formed by introducing lighter halogen vapors into a helium-filled cell containing astatine, supporting the existence of stable neutral molecules in the plasma ion state. [56] No astatine fluorides have been discovered yet. Their absence has been speculatively attributed to the extreme reactivity of such compounds, including the reaction of an initially formed fluoride with the walls of the glass container to form a non-volatile product. [lower-alpha 5] Thus, although the synthesis of an astatine fluoride is thought to be possible, it may require a liquid halogen fluoride solvent, as has already been used for the characterization of radon fluoride. [56] [72]

History

Cercle rouge 100%25.svg
Periodic table by Mendeleev, 1871.svg
Dmitri Mendeleev's table of 1871, with an empty space at the eka-iodine position

In 1869, when Dmitri Mendeleev published his periodic table, the space under iodine was empty; after Niels Bohr established the physical basis of the classification of chemical elements, it was suggested that the fifth halogen belonged there. Before its officially recognized discovery, it was called "eka-iodine" (from Sanskrit eka – "one") to imply it was one space under iodine (in the same manner as eka-silicon, eka-boron, and others). [82] Scientists tried to find it in nature; given its extreme rarity, these attempts resulted in several false discoveries. [83]

The first claimed discovery of eka-iodine was made by Fred Allison and his associates at the Alabama Polytechnic Institute (now Auburn University) in 1931. The discoverers named element 85 "alabamine", and assigned it the symbol Ab, designations that were used for a few years. [84] [85] [86] In 1934, H. G. MacPherson of University of California, Berkeley disproved Allison's method and the validity of his discovery. [87] There was another claim in 1937, by the chemist Rajendralal De. Working in Dacca in British India (now Dhaka in Bangladesh), he chose the name "dakin" for element 85, which he claimed to have isolated as the thorium series equivalent of radium F (polonium-210) in the radium series. [88] The properties he reported for dakin do not correspond to those of astatine, [88] and astatine's radioactivity would have prevented him from handling it in the quantities he claimed. [89] Moreover, astatine is not found in the thorium series, and the true identity of dakin is not known. [88]

In 1936, the team of Romanian physicist Horia Hulubei and French physicist Yvette Cauchois claimed to have discovered element 85 by observing its X-ray emission lines. In 1939, they published another paper which supported and extended previous data. In 1944, Hulubei published a summary of data he had obtained up to that time, claiming it was supported by the work of other researchers. He chose the name "dor", presumably from the Romanian for "longing" [for peace], as World War II had started five years earlier. As Hulubei was writing in French, a language which does not accommodate the "ine" suffix, dor would likely have been rendered in English as "dorine", had it been adopted. In 1947, Hulubei's claim was effectively rejected by the Austrian chemist Friedrich Paneth, who would later chair the IUPAC committee responsible for recognition of new elements. Even though Hulubei's samples did contain astatine-218, his means to detect it were too weak, by current standards, to enable correct identification; moreover, he could not perform chemical tests on the element. [89] He had also been involved in an earlier false claim as to the discovery of element 87 (francium) and this is thought to have caused other researchers to downplay his work. [90]

Emilio Segre, one of the discoverers of the main-group element astatine Segre.jpg
Emilio Segrè, one of the discoverers of the main-group element astatine

In 1940, the Swiss chemist Walter Minder announced the discovery of element 85 as the beta decay product of radium A (polonium-218), choosing the name "helvetium" (from Helvetia , the Latin name of Switzerland). Berta Karlik and Traude Bernert were unsuccessful in reproducing his experiments, and subsequently attributed Minder's results to contamination of his radon stream (radon-222 is the parent isotope of polonium-218). [91] [lower-alpha 6] In 1942, Minder, in collaboration with the English scientist Alice Leigh-Smith, announced the discovery of another isotope of element 85, presumed to be the product of thorium A (polonium-216) beta decay. They named this substance "anglo-helvetium", [92] but Karlik and Bernert were again unable to reproduce these results. [54]

Later in 1940, Dale R. Corson, Kenneth Ross MacKenzie, and Emilio Segrè isolated the element at the University of California, Berkeley. Instead of searching for the element in nature, the scientists created it by bombarding bismuth-209 with alpha particles in a cyclotron (particle accelerator) to produce, after emission of two neutrons, astatine-211. [93] The discoverers, however, did not immediately suggest a name for the element. The reason for this was that at the time, an element created synthetically in "invisible quantities" that had not yet been discovered in nature was not seen as a completely valid one; in addition, chemists were reluctant to recognize radioactive isotopes as legitimately as stable ones. [94] In 1943, astatine was found as a product of two naturally occurring decay chains by Berta Karlik and Traude Bernert, first in the so-called uranium series, and then in the actinium series. [95] [96] (Since then, astatine was also found in a third decay chain, the neptunium series. [97] ) Friedrich Paneth in 1946 called to finally recognize synthetic elements, quoting, among other reasons, recent confirmation of their natural occurrence, and proposed that the discoverers of the newly discovered unnamed elements name these elements. In early 1947, Nature published the discoverers' suggestions; a letter from Corson, MacKenzie, and Segrè suggested the name "astatine" [94] coming from the Ancient Greek αστατος (astatos) meaning 'unstable', because of its propensity for radioactive decay, with the ending "-ine", found in the names of the four previously discovered halogens. The name was also chosen to continue the tradition of the four stable halogens, where the name referred to a property of the element. [98]

Corson and his colleagues classified astatine as a metal on the basis of its analytical chemistry. [99] Subsequent investigators reported iodine-like, [100] [101] cationic, [102] [103] or amphoteric behavior. [104] [105] In a 2003 retrospective, Corson wrote that "some of the properties [of astatine] are similar to iodine ... it also exhibits metallic properties, more like its metallic neighbors Po and Bi." [98]

Isotopes

Alpha decay characteristics for sample astatine isotopes [lower-alpha 7]
Mass
number
Half-life [7] Probability
of alpha
decay [7]
Alpha
decay
half-life
2071.80 h8.6%20.9 h
2081.63 h0.55%12.3 d
2095.41 h4.1%5.5 d
2108.1 h0.175%193 d
2117.21 h41.8%17.2 h
2120.31 s0.31 s
213125 ns100%125 ns
214558 ns100%558 ns
21956 s97%58 s
2203.71 min8%46.4 min
2212.3 min

There are 41 known isotopes of astatine, with mass numbers of 188 and 190–229. [106] [107] Theoretical modeling suggests that about 37 more isotopes could exist. [106] No stable or long-lived astatine isotope has been observed, nor is one expected to exist. [108]

Astatine's alpha decay energies follow the same trend as for other heavy elements. [108] Lighter astatine isotopes have quite high energies of alpha decay, which become lower as the nuclei become heavier. Astatine-211 has a significantly higher energy than the previous isotope, because it has a nucleus with 126 neutrons, and 126 is a magic number corresponding to a filled neutron shell. Despite having a similar half-life to the previous isotope (8.1 hours for astatine-210 and 7.2 hours for astatine-211), the alpha decay probability is much higher for the latter: 41.81% against only 0.18%. [7] [lower-alpha 8] The two following isotopes release even more energy, with astatine-213 releasing the most energy. For this reason, it is the shortest-lived astatine isotope. [108] Even though heavier astatine isotopes release less energy, no long-lived astatine isotope exists, because of the increasing role of beta decay (electron emission). [108] This decay mode is especially important for astatine; as early as 1950 it was postulated that all isotopes of the element undergo beta decay, [109] though nuclear mass measurements indicate that 215At is in fact beta-stable, as it has the lowest mass of all isobars with A = 215. [7] Astatine-210 and most of the lighter isotopes exhibit beta plus decay (positron emission), astatine-217 and heavier isotopes except astatine-218 exhibit beta minus decay, while astatine-211 undergoes electron capture. [5]

The most stable isotope is astatine-210, which has a half-life of 8.1 hours. The primary decay mode is beta plus, to the relatively long-lived (in comparison to astatine isotopes) alpha emitter polonium-210. In total, only five isotopes have half-lives exceeding one hour (astatine-207 to -211). The least stable ground state isotope is astatine-213, with a half-life of 125 nanoseconds. It undergoes alpha decay to the extremely long-lived bismuth-209. [7]

Astatine has 24 known nuclear isomers, which are nuclei with one or more nucleons (protons or neutrons) in an excited state. A nuclear isomer may also be called a "meta-state", meaning the system has more internal energy than the "ground state" (the state with the lowest possible internal energy), making the former likely to decay into the latter. There may be more than one isomer for each isotope. The most stable of these nuclear isomers is astatine-202m1, [lower-alpha 9] which has a half-life of about 3 minutes, longer than those of all the ground states bar those of isotopes 203–211 and 220. The least stable is astatine-213m1; its half-life of 110 nanoseconds is shorter than 125 nanoseconds for astatine-213, the shortest-lived ground state. [5]

Natural occurrence

Neptunium series, showing the decay products, including astatine-217, formed from neptunium-237 Decay Chain(4n+1, Neptunium Series).svg
Neptunium series, showing the decay products, including astatine-217, formed from neptunium-237

Astatine is the rarest naturally occurring element. [lower-alpha 10] The total amount of astatine in the Earth's crust (quoted mass 2.36 × 1025 grams) [110] is estimated by some to be less than one gram at any given time. [9] Other sources estimate the amount of ephemeral astatine, present on earth at any given moment, to be up to one ounce [111] (about 28 grams).

Any astatine present at the formation of the Earth has long since disappeared; the four naturally occurring isotopes (astatine-215, -217, -218 and -219) [112] are instead continuously produced as a result of the decay of radioactive thorium and uranium ores, and trace quantities of neptunium-237. The landmass of North and South America combined, to a depth of 16 kilometers (10 miles), contains only about one trillion astatine-215 atoms at any given time (around 3.5 × 10−10 grams). [113] Astatine-217 is produced via the radioactive decay of neptunium-237. Primordial remnants of the latter isotope—due to its relatively short half-life of 2.14 million years—are no longer present on Earth. However, trace amounts occur naturally as a product of transmutation reactions in uranium ores. [114] Astatine-218 was the first astatine isotope discovered in nature. [115] Astatine-219, with a half-life of 56 seconds, is the longest lived of the naturally occurring isotopes. [7]

Isotopes of astatine are sometimes not listed as naturally occurring because of misconceptions [104] that there are no such isotopes, [116] or discrepancies in the literature. Astatine-216 has been counted as a naturally occurring isotope but reports of its observation [117] (which were described as doubtful) have not been confirmed. [118]

Synthesis

Formation

Possible reactions after bombarding bismuth-209 with alpha particles
Reaction [lower-alpha 11] Energy of alpha particle
209
83
Bi
+ 4
2
He
211
85
At
+ 2 1
0
n
26 MeV [54]
209
83
Bi
+ 4
2
He
210
85
At
+ 3 1
0
n
40 MeV [54]
209
83
Bi
+ 4
2
He
209
85
At
+ 4 1
0
n
60 MeV [119]

Astatine was first produced by bombarding bismuth-209 with energetic alpha particles, and this is still the major route used to create the relatively long-lived isotopes astatine-209 through astatine-211. Astatine is only produced in minuscule quantities, with modern techniques allowing production runs of up to 6.6  gigabecquerels [120] (about 86  nanograms or 2.47×1014 atoms). Synthesis of greater quantities of astatine using this method is constrained by the limited availability of suitable cyclotrons and the prospect of melting the target. [120] [121] [lower-alpha 12] Solvent radiolysis due to the cumulative effect of astatine decay [123] is a related problem. With cryogenic technology, microgram quantities of astatine might be able to be generated via proton irradiation of thorium or uranium to yield radon-211, in turn decaying to astatine-211. Contamination with astatine-210 is expected to be a drawback of this method. [124]

The most important isotope is astatine-211, the only one in commercial use. To produce the bismuth target, the metal is sputtered onto a gold, copper, or aluminium surface at 50 to 100 milligrams per square centimeter. Bismuth oxide can be used instead; this is forcibly fused with a copper plate. [125] The target is kept under a chemically neutral nitrogen atmosphere, [126] and is cooled with water to prevent premature astatine vaporization. [125] In a particle accelerator, such as a cyclotron, [127] alpha particles are collided with the bismuth. Even though only one bismuth isotope is used (bismuth-209), the reaction may occur in three possible ways, producing astatine-209, astatine-210, or astatine-211. In order to eliminate undesired nuclides, the maximum energy of the particle accelerator is set to a value (optimally 29.17 MeV) [128] above that for the reaction producing astatine-211 (to produce the desired isotope) and below the one producing astatine-210 (to avoid producing other astatine isotopes). [125]

Separation methods

Since astatine is the main product of the synthesis, after its formation it must only be separated from the target and any significant contaminants. Several methods are available, "but they generally follow one of two approaches—dry distillation or [wet] acid treatment of the target followed by solvent extraction." The methods summarized below are modern adaptations of older procedures, as reviewed by Kugler and Keller. [129] [lower-alpha 13] Pre-1985 techniques more often addressed the elimination of co-produced toxic polonium; this requirement is now mitigated by capping the energy of the cyclotron irradiation beam. [120]

Dry

The astatine-containing cyclotron target is heated to a temperature of around 650 °C. The astatine volatilizes and is condensed in (typically) a cold trap. Higher temperatures of up to around 850 °C may increase the yield, at the risk of bismuth contamination from concurrent volatilization. Redistilling the condensate may be required to minimize the presence of bismuth [131] (as bismuth can interfere with astatine labeling reactions). The astatine is recovered from the trap using one or more low concentration solvents such as sodium hydroxide, methanol or chloroform. Astatine yields of up to around 80% may be achieved. Dry separation is the method most commonly used to produce a chemically useful form of astatine. [121] [132]

Wet

The irradiated bismuth (or sometimes bismuth trioxide) target is first dissolved in, for example, concentrated nitric or perchloric acid. Following this first step, the acid can be distilled away to leave behind a white residue that contains both bismuth and the desired astatine product. This residue is then dissolved in a concentrated acid, such as hydrochloric acid. Astatine is extracted from this acid using an organic solvent such as dibutyl ether, diisopropyl ether (DIPE), or thiosemicarbazide. Using liquid-liquid extraction, the astatine product can be repeatedly washed with an acid, such as HCl, and extracted into the organic solvent layer. A separation yield of 93% using nitric acid has been reported, falling to 72% by the time purification procedures were completed (distillation of nitric acid, purging residual nitrogen oxides, and redissolving bismuth nitrate to enable liquid–liquid extraction). [133] [134] Wet methods involve "multiple radioactivity handling steps" and have not been considered well suited for isolating larger quantities of astatine. However, wet extraction methods are being examined for use in production of larger quantities of astatine-211, as it is thought that wet extraction methods can provide more consistency. [134] They can enable the production of astatine in a specific oxidation state and may have greater applicability in experimental radiochemistry. [120]

Uses and precautions

Several 211At-containing molecules and their experimental uses [135]
AgentApplications
[211At]astatine-tellurium colloidsCompartmental tumors
6-[211At]astato-2-methyl-1,4-naphtaquinol diphosphate Adenocarcinomas
211At-labeled methylene blue Melanomas
Meta-[211At]astatobenzyl guanidineNeuroendocrine tumors
5-[211At]astato-2'-deoxyuridineVarious
211At-labeled biotin conjugatesVarious pretargeting
211At-labeled octreotide Somatostatin receptor
211At-labeled monoclonal antibodies and fragmentsVarious
211At-labeled bisphosphonates Bone metastases

Newly formed astatine-211 is the subject of ongoing research in nuclear medicine. [135] It must be used quickly as it decays with a half-life of 7.2 hours; this is long enough to permit multistep labeling strategies. Astatine-211 has potential for targeted alpha-particle therapy, since it decays either via emission of an alpha particle (to bismuth-207), [136] or via electron capture (to an extremely short-lived nuclide, polonium-211, which undergoes further alpha decay), very quickly reaching its stable granddaughter lead-207. Polonium X-rays emitted as a result of the electron capture branch, in the range of 77–92  keV, enable the tracking of astatine in animals and patients. [135] Although astatine-210 has a slightly longer half-life, it is wholly unsuitable because it usually undergoes beta plus decay to the extremely toxic polonium-210. [137]

The principal medicinal difference between astatine-211 and iodine-131 (a radioactive iodine isotope also used in medicine) is that iodine-131 emits high-energy beta particles, and astatine does not. Beta particles have much greater penetrating power through tissues than do the much heavier alpha particles. An average alpha particle released by astatine-211 can travel up to 70 μm through surrounding tissues; an average-energy beta particle emitted by iodine-131 can travel nearly 30 times as far, to about 2 mm. [125] The short half-life and limited penetrating power of alpha radiation through tissues offers advantages in situations where the "tumor burden is low and/or malignant cell populations are located in close proximity to essential normal tissues." [120] Significant morbidity in cell culture models of human cancers has been achieved with from one to ten astatine-211 atoms bound per cell. [138]

Astatine ... [is] miserable to make and hell to work with. [139]

P Durbin, Human Radiation Studies: Remembering the Early Years, 1995

Several obstacles have been encountered in the development of astatine-based radiopharmaceuticals for cancer treatment. World War II delayed research for close to a decade. Results of early experiments indicated that a cancer-selective carrier would need to be developed and it was not until the 1970s that monoclonal antibodies became available for this purpose. Unlike iodine, astatine shows a tendency to dehalogenate from molecular carriers such as these, particularly at sp3 carbon sites [lower-alpha 14] (less so from sp2 sites). Given the toxicity of astatine accumulated and retained in the body, this emphasized the need to ensure it remained attached to its host molecule. While astatine carriers that are slowly metabolized can be assessed for their efficacy, more rapidly metabolized carriers remain a significant obstacle to the evaluation of astatine in nuclear medicine. Mitigating the effects of astatine-induced radiolysis of labeling chemistry and carrier molecules is another area requiring further development. A practical application for astatine as a cancer treatment would potentially be suitable for a "staggering" number of patients; production of astatine in the quantities that would be required remains an issue. [124] [140] [lower-alpha 15]

Animal studies show that astatine, similarly to iodine—although to a lesser extent, perhaps because of its slightly more metallic nature [111] —is preferentially (and dangerously) concentrated in the thyroid gland. Unlike iodine, astatine also shows a tendency to be taken up by the lungs and spleen, possibly because of in-body oxidation of At to At+. [43] If administered in the form of a radiocolloid it tends to concentrate in the liver. Experiments in rats and monkeys suggest that astatine-211 causes much greater damage to the thyroid gland than does iodine-131, with repetitive injection of the nuclide resulting in necrosis and cell dysplasia within the gland. [141] Early research suggested that injection of astatine into female rodents caused morphological changes in breast tissue; [142] this conclusion remained controversial for many years. General agreement was later reached that this was likely caused by the effect of breast tissue irradiation combined with hormonal changes due to irradiation of the ovaries. [139] Trace amounts of astatine can be handled safely in fume hoods if they are well-aerated; biological uptake of the element must be avoided. [143]

See also

Notes

  1. This half-vaporization period grows to 16 hours if it is instead put on a gold or platinum surface; this may be caused by poorly understood interactions between astatine and these noble metals. [18]
  2. It is also possible that this is sorption on a cathode. [40]
  3. The algorithm used to generate the Allred-Rochow scale fails in the case of hydrogen, providing a value that is close to that of oxygen (3.5). Hydrogen is instead assigned a value of 2.2. Despite this shortcoming, the Allred-Rochow scale has achieved a relatively high degree of acceptance. [50]
  4. Iodine can act as a carrier despite it reacting with astatine in water because these reactions require iodide (I), not (only) I2. [56] [57]
  5. An initial attempt to fluoridate astatine using chlorine trifluoride resulted in formation of a product which became stuck to the glass. Chlorine monofluoride, chlorine, and tetrafluorosilane were formed. The authors called the effect "puzzling", admitting they had expected formation of a volatile fluoride. [79] Ten years later, the compound was predicted to be non-volatile, out of line with the lighter halogens but similar to radon fluoride; [80] by this time, the latter had been shown to be ionic. [81]
  6. In other words, some other substance was undergoing beta decay (to a different end element), not polonium-218.
  7. In the table, "alpha decay half-life" refers to the half-life if decay modes other than alpha are omitted.
  8. This means that, if decay modes other than alpha are omitted, then astatine-210 has an alpha decay half-life of 4,628.6 hours (128.9 days) and astatine-211 has one of only 17.2 hours (0.7 days). Therefore, astatine-211 is very much less stable toward alpha decay than astatine-210.
  9. "m1" means that this state of the isotope is the next possible one above – with an energy greater than – the ground state. "m2" and similar designations refer to further higher energy states. The number may be dropped if there is only one well-established meta state, such as astatine-216m. Other designation techniques are sometimes used.
  10. Emsley [10] states that this title has been lost to berkelium, "a few atoms of which can be produced in very-highly concentrated uranium-bearing deposits"; however, his assertion is not corroborated by any primary source.
  11. A nuclide is commonly denoted by a symbol of the chemical element this nuclide belongs to, preceded by a non-spaced superscript mass number and a subscript atomic number of the nuclide located directly under the mass number. (Neutrons may be considered as nuclei with the atomic mass of 1 and the atomic charge of 0, with the symbol being n.) With the atomic number omitted, it is also sometimes used as a designation of an isotope of an element in isotope-related chemistry.
  12. See however Nagatsu et al. [122] who encapsulate the bismuth target in a thin aluminium foil and place it in a niobium holder capable of holding molten bismuth.
  13. See also Lavrukhina and Pozdnyakov. [130]
  14. In other words, where carbon's one s atomic orbital and three p orbitals hybridize to give four new orbitals shaped as intermediates between the original s and p orbitals.
  15. "Unfortunately, the conundrum confronting the ... field is that commercial supply of 211At awaits the demonstration of clinical efficacy; however, the demonstration of clinical efficacy requires a reliable supply of 211At." [120]

Related Research Articles

<span class="mw-page-title-main">Actinium</span> Chemical element, symbol Ac and atomic number 89

Actinium is a chemical element; it has symbol Ac and atomic number 89. It was first isolated by Friedrich Oskar Giesel in 1902, who gave it the name emanium; the element got its name by being wrongly identified with a substance André-Louis Debierne found in 1899 and called actinium. Actinium gave the name to the actinide series, a set of 15 elements between actinium and lawrencium in the periodic table. Together with polonium, radium, and radon, actinium was one of the first non-primordial radioactive elements to be isolated.

<span class="mw-page-title-main">Bohrium</span> Chemical element, symbol Bh and atomic number 107

Bohrium is a synthetic chemical element; it has symbol Bh and atomic number 107. It is named after Danish physicist Niels Bohr. As a synthetic element, it can be created in particle accelerators but is not found in nature. All known isotopes of bohrium are highly radioactive; the most stable known isotope is 270Bh with a half-life of approximately 2.4 minutes, though the unconfirmed 278Bh may have a longer half-life of about 11.5 minutes.

A chemical element is a chemical substance that cannot be broken down into other substances by chemical reactions. The basic particle that constitutes a chemical element is the atom. Chemical elements are identified by the number of protons in the nuclei of their atoms, known as the element's atomic number. For example, oxygen has an atomic number of 8, meaning that each oxygen atom has 8 protons in its nucleus. Two or more atoms of the same element can combine to form molecules, in contrast to chemical compounds or mixtures, which contain atoms of different elements. Atoms can be transformed into different elements in nuclear reactions, which change an atom's atomic number.

<span class="mw-page-title-main">Chalcogen</span> Group of chemical elements

The chalcogens are the chemical elements in group 16 of the periodic table. This group is also known as the oxygen family. Group 16 consists of the elements oxygen (O), sulfur (S), selenium (Se), tellurium (Te), and the radioactive elements polonium (Po) and livermorium (Lv). Often, oxygen is treated separately from the other chalcogens, sometimes even excluded from the scope of the term "chalcogen" altogether, due to its very different chemical behavior from sulfur, selenium, tellurium, and polonium. The word "chalcogen" is derived from a combination of the Greek word khalkόs (χαλκός) principally meaning copper, and the Latinized Greek word genēs, meaning born or produced.

<span class="mw-page-title-main">Francium</span> Chemical element, symbol Fr and atomic number 87

Francium is a chemical element; it has symbol Fr and atomic number 87. It is extremely radioactive; its most stable isotope, francium-223, has a half-life of only 22 minutes. It is the second-most electropositive element, behind only caesium, and is the second rarest naturally occurring element. Francium's isotopes decay quickly into astatine, radium, and radon. The electronic structure of a francium atom is [Rn] 7s1; thus, the element is classed as an alkali metal.

<span class="mw-page-title-main">Halogen</span> Group of chemical elements

The halogens are a group in the periodic table consisting of six chemically related elements: fluorine (F), chlorine (Cl), bromine (Br), iodine (I), and the radioactive elements astatine (At) and tennessine (Ts), though some authors would exclude tennessine as its chemistry is unknown and is theoretically expected to be more like that of gallium. In the modern IUPAC nomenclature, this group is known as group 17.

<span class="mw-page-title-main">Promethium</span> Chemical element, symbol Pm and atomic number 61

Promethium is a chemical element; it has symbol Pm and atomic number 61. All of its isotopes are radioactive; it is extremely rare, with only about 500–600 grams naturally occurring in Earth's crust at any given time. Promethium is one of only two radioactive elements that are followed in the periodic table by elements with stable forms, the other being technetium. Chemically, promethium is a lanthanide. Promethium shows only one stable oxidation state of +3.

<span class="mw-page-title-main">Polonium</span> Chemical element, symbol Po and atomic number 84

Polonium is a chemical element; it has symbol Po and atomic number 84. A rare and highly radioactive metal with no stable isotopes, polonium is a chalcogen and chemically similar to selenium and tellurium, though its metallic character resembles that of its horizontal neighbors in the periodic table: thallium, lead, and bismuth. Due to the short half-life of all its isotopes, its natural occurrence is limited to tiny traces of the fleeting polonium-210 in uranium ores, as it is the penultimate daughter of natural uranium-238. Though longer-lived isotopes exist, such as the 124 years half-life of polonium-209, they are much more difficult to produce. Today, polonium is usually produced in milligram quantities by the neutron irradiation of bismuth. Due to its intense radioactivity, which results in the radiolysis of chemical bonds and radioactive self-heating, its chemistry has mostly been investigated on the trace scale only.

A radionuclide (radioactive nuclide, radioisotope or radioactive isotope) is a nuclide that has excess numbers of either neutrons or protons, giving it excess nuclear energy, and making it unstable. This excess energy can be used in one of three ways: emitted from the nucleus as gamma radiation; transferred to one of its electrons to release it as a conversion electron; or used to create and emit a new particle (alpha particle or beta particle) from the nucleus. During those processes, the radionuclide is said to undergo radioactive decay. These emissions are considered ionizing radiation because they are energetic enough to liberate an electron from another atom. The radioactive decay can produce a stable nuclide or will sometimes produce a new unstable radionuclide which may undergo further decay. Radioactive decay is a random process at the level of single atoms: it is impossible to predict when one particular atom will decay. However, for a collection of atoms of a single nuclide the decay rate, and thus the half-life (t1/2) for that collection, can be calculated from their measured decay constants. The range of the half-lives of radioactive atoms has no known limits and spans a time range of over 55 orders of magnitude.

Livermorium is a synthetic chemical element; it has symbol Lv and atomic number 116. It is an extremely radioactive element that has only been created in a laboratory setting and has not been observed in nature. The element is named after the Lawrence Livermore National Laboratory in the United States, which collaborated with the Joint Institute for Nuclear Research (JINR) in Dubna, Russia, to discover livermorium during experiments conducted between 2000 and 2006. The name of the laboratory refers to the city of Livermore, California, where it is located, which in turn was named after the rancher and landowner Robert Livermore. The name was adopted by IUPAC on May 30, 2012. Five isotopes of livermorium are known, with mass numbers of 288 and 290–293 inclusive; the longest-lived among them is livermorium-293 with a half-life of about 60 milliseconds. A sixth possible isotope with mass number 294 has been reported but not yet confirmed.

Moscovium is a synthetic chemical element; it has symbol Mc and atomic number 115. It was first synthesized in 2003 by a joint team of Russian and American scientists at the Joint Institute for Nuclear Research (JINR) in Dubna, Russia. In December 2015, it was recognized as one of four new elements by the Joint Working Party of international scientific bodies IUPAC and IUPAP. On 28 November 2016, it was officially named after the Moscow Oblast, in which the JINR is situated.

Tennessine is a synthetic chemical element; it has symbol Ts and atomic number 117. It has the second-highest atomic number and joint-highest atomic mass of all known elements, and is the penultimate element of the 7th period of the periodic table.

A period 6 element is one of the chemical elements in the sixth row (or period) of the periodic table of the chemical elements, including the lanthanides. The periodic table is laid out in rows to illustrate recurring (periodic) trends in the chemical behaviour of the elements as their atomic number increases: a new row is begun when chemical behaviour begins to repeat, meaning that elements with similar behaviour fall into the same vertical columns. The sixth period contains 32 elements, tied for the most with period 7, beginning with caesium and ending with radon. Lead is currently the last stable element; all subsequent elements are radioactive. For bismuth, however, its only primordial isotope, 209Bi, has a half-life of more than 1019 years, over a billion times longer than the current age of the universe. As a rule, period 6 elements fill their 6s shells first, then their 4f, 5d, and 6p shells, in that order; however, there are exceptions, such as gold.

<span class="mw-page-title-main">Decay chain</span> Series of radioactive decays

In nuclear science, the decay chain refers to a series of radioactive decays of different radioactive decay products as a sequential series of transformations. It is also known as a "radioactive cascade". The typical radioisotope does not decay directly to a stable state, but rather it decays to another radioisotope. Thus there is usually a series of decays until the atom has become a stable isotope, meaning that the nucleus of the atom has reached a stable state.

In chemistry, an interhalogen compound is a molecule which contains two or more different halogen atoms and no atoms of elements from any other group.

Astatine (85At) has 41 known isotopes, all of which are radioactive; their mass numbers range from 188 to 229. There are also 24 known metastable excited states. The longest-lived isotope is 210At, which has a half-life of 8.1 hours; the longest-lived isotope existing in naturally occurring decay chains is 219At with a half-life of 56 seconds.

Bismuth-209 (209Bi) is an isotope of bismuth; with the longest known half-life of any radioisotope that undergoes α-decay. It has 83 protons and a magic number of 126 neutrons, and an atomic mass of 208.9803987 amu. Primordial bismuth consists entirely of this isotope.

<span class="mw-page-title-main">Hydrogen astatide</span> Chemical compound

Hydrogen astatide, also known as astatine hydride, astatane, astatidohydrogen or hydroastatic acid, is a chemical compound with the chemical formula HAt, consisting of an astatine atom covalently bonded to a hydrogen atom. It thus is a hydrogen halide.

<span class="mw-page-title-main">Astatine compounds</span>

Astatine compounds are compounds that contain the element astatine (At). As this element is very radioactive, few compounds have been studied. Less reactive than iodine, astatine is the least reactive of the halogens. Its compounds have been synthesized in nano-scale amounts and studied as intensively as possible before their radioactive disintegration. The reactions involved have been typically tested with dilute solutions of astatine mixed with larger amounts of iodine. Acting as a carrier, the iodine ensures there is sufficient material for laboratory techniques to work. Like iodine, astatine has been shown to adopt odd-numbered oxidation states ranging from −1 to +7.

Organoastatine chemistry describes the synthesis and properties of organoastatine compounds, chemical compounds containing a carbon to astatine chemical bond.

References

  1. 1 2 3 Arblaster, JW, ed. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. p. 604. ISBN   978-1-62708-154-2.
  2. Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. p. 28. ISBN   978-0-08-037941-8.
  3. 1 2 Rothe, S.; Andreyev, A. N.; Antalic, S.; Borschevsky, A.; Capponi, L.; Cocolios, T. E.; De Witte, H.; Eliav, E.; et al. (2013). "Measurement of the First Ionization Potential of Astatine by Laser Ionization Spectroscopy". Nature Communications. 4: 1–6. Bibcode:2013NatCo...4E1835R. doi:10.1038/ncomms2819. PMC   3674244 . PMID   23673620.
  4. 1 2 3 4 Hermann, A.; Hoffmann, R.; Ashcroft, N. W. (2013). "Condensed Astatine: Monatomic and Metallic". Physical Review Letters. 111 (11): 116404-1–116404-5. Bibcode:2013PhRvL.111k6404H. doi:10.1103/PhysRevLett.111.116404. PMID   24074111.
  5. 1 2 3 Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  6. "Astatine (At) | AMERICAN ELEMENTS". American Elements: The Materials Science Company. Retrieved 4 April 2024.
  7. 1 2 3 4 5 6 7 Audi, Georges; Bersillon, Olivier; Blachot, Jean; Wapstra, Aaldert Hendrik (2003). "The NUBASE evaluation of nuclear and decay properties". Nuclear Physics A. 729: 3–128. Bibcode:2003NuPhA.729....3A. doi:10.1016/j.nuclphysa.2003.11.001.
  8. Greenwood & Earnshaw 2002, p. 795.
  9. 1 2 3 4 5 6 Wiberg, N., ed. (2001). Holleman-Wiberg: Inorganic Chemistry. Translation of 101st German edition by M. Eagleson and W. D. Brewer, English language editor B. J. Aylett. Academic Press. p. 423. ISBN   978-0-12-352651-9.
  10. 1 2 3 Emsley, J. (2011). Nature's Building Blocks: An A-Z Guide to the Elements (New ed.). Oxford University Press. pp. 57–58. ISBN   978-0-19-960563-7.
  11. Kotz, J. C.; Treichel, P. M.; Townsend, J. (2011). Chemistry & Chemical Reactivity (8th ed.). Cengage Learning. p. 65. ISBN   978-0-8400-4828-8.
  12. Jahn, T. P. (2010). MIPS and Their Role in the Exchange of Metalloids. Vol. 679. Springer. p. 41. ISBN   978-1-4419-6314-7.
  13. Siekierski, S.; Burgess, J. (2002). Concise Chemistry of the Elements. Horwood. pp. 65, 122. ISBN   978-1-898563-71-6.
  14. Maddock, A. G. (1956). "Astatine". Supplement to Mellor's Comprehensive Treatise on Inorganic and Theoretical Chemistry, Supplement II, Part 1, (F, Cl, Br, I, At). Longmans, Green & Co. (Ltd.). pp. 1064–1079.
  15. Garrett, A. B.; Richardson, J. B.; Kiefer, A. S. (1961). Chemistry: A First Course in Modern Chemistry. Ginn. p. 313.
  16. Seaborg, G. T. (2015). "Transuranium element". Encyclopædia Britannica. Retrieved 24 February 2015.
  17. Oon, H. L. (2007). Chemistry Expression: An Inquiry Approach. John Wiley and Sons. p. 300. ISBN   978-981-271-162-5.
  18. Lavrukhina & Pozdnyakov 1970, p. 251.
  19. McLaughlin, R. (1964). "Absorption Spectrum of Astatine". Journal of the Optical Society of America. 54 (8): 965–967. Bibcode:1964JOSA...54..965M. doi:10.1364/JOSA.54.000965.
  20. Lavrukhina & Pozdnyakov 1970, p. 235.
  21. Donohue, J. (1982). The Structures of the Elements. Robert E. Krieger. p. 400. ISBN   978-0-89874-230-5.
  22. 1 2 Vernon, R. (2013). "Which Elements are Metalloids?". Journal of Chemical Education. 90 (12): 1703–1707 (1704). Bibcode:2013JChEd..90.1703V. doi:10.1021/ed3008457.
  23. Batsanov, S. S. (1972). "Quantitative characteristics of bond metallicity in crystals". Journal of Structural Chemistry. 12 (5): 809–813. doi:10.1007/BF00743349. ISSN   0022-4766. S2CID   96816296.
  24. Merinis, J.; Legoux, G.; Bouissières, G. (1972). "Etude de la formation en phase gazeuse de composés interhalogénés d'astate par thermochromatographie" [Study of the gas-phase formation of interhalogen compounds of astatine by thermochromatography ]. Radiochemical and Radioanalytical Letters (in French). 11 (1): 59–64.
  25. Takahashi, N.; Otozai, K. (1986). "The Mechanism of the Reaction of Elementary Astatine with Organic Solvents". Journal of Radioanalytical and Nuclear Chemistry. 103: 1–9. doi:10.1007/BF02165358. S2CID   93572282.
  26. Takahashi, N.; Yano, D.; Baba, H. (1992). "Chemical Behavior of Astatine Molecules". Proceedings of the International Conference on Evolution in Beam Applications, Takasaki, Japan, 5–8 November 1991. pp. 536–539.
  27. Zuckerman & Hagen 1989, p. 21.
  28. Kugler & Keller 1985, pp. 110, 116, 210–211, 224.
  29. Meyers, R. A. (2001). "Halogen Chemistry". Encyclopedia of Physical Science and Technology (3rd ed.). Academic Press. pp. 197–222 (202). ISBN   978-0-12-227410-7.
  30. Keller, C.; Wolf, W.; Shani, J. (2011). "Radionuclides, 2. Radioactive Elements and Artificial Radionuclides". Ullmann's Encyclopedia of Industrial Chemistry. Vol. 31. pp. 89–117 (96). doi:10.1002/14356007.o22_o15. ISBN   978-3-527-30673-2.
  31. Otozai, K.; Takahashi, N. (1982). "Estimation Chemical Form Boiling Point Elementary Astatine by Radio Gas Chromatography". Radiochimica Acta. 31 (3–4): 201–203. doi:10.1524/ract.1982.31.34.201. S2CID   100363889.
  32. Zumdahl, S. S.; Zumdahl, S. A. (2008). Chemistry (8th ed.). Cengage Learning. p. 56. ISBN   978-0-547-12532-9.
  33. 1 2 Housecroft, C. E.; Sharpe, A. G. (2008). Inorganic chemistry (3rd ed.). Pearson Education. p. 533. ISBN   978-0-13-175553-6.
  34. Kugler & Keller 1985, p. 116.
  35. Visscher, L.; Dyall, K. G. (1996). "Relativistic and Correlation Effects on Molecular properties. I. The Dihalogens F2, Cl2, Br2, I2, and At2". The Journal of Chemical Physics. 104 (22): 9040–9046. Bibcode:1996JChPh.104.9040V. doi: 10.1063/1.471636 .
  36. Glushko, V. P.; Medvedev, V. A.; Bergma, G. A. (1966). Termicheskie Konstanty Veshchestv (in Russian). Vol. 1. Nakua. p. 65.
  37. Kugler & Keller 1985, pp. 116–117.
  38. Smith, A.; Ehret, W. F. (1960). College chemistry. Appleton-Century-Crofts. p. 457.
  39. 1 2 Champion, J.; Seydou, M.; Sabatié-Gogova, A.; Renault, E.; Montavon, G.; Galland, N. (2011). "Assessment of an Effective Quasirelativistic Methodology Designed to Study Astatine Chemistry in Aqueous Solution" . Physical Chemistry Chemical Physics . 13 (33): 14984–14992 (14984). Bibcode:2011PCCP...1314984C. doi:10.1039/C1CP20512A. PMID   21769335.
  40. Milanov, M.; Doberenz, V.; Khalkin, V. A.; Marinov, A. (1984). "Chemical Properties of Positive Singly Charged Astatine Ion in Aqueous Solution". Journal of Radioanalytical and Nuclear Chemistry. 83 (2): 291–299. doi:10.1007/BF02037143. S2CID   97361684.
  41. 1 2 Lavrukhina & Pozdnyakov 1970, p. 234.
  42. Milesz, S.; Jovchev, M.; Schumann, D.; Khalkin, V. A. (1988). "The EDTA Complexes of Astatine". Journal of Radioanalytical and Nuclear Chemistry. 127 (3): 193–198. doi:10.1007/BF02164864. S2CID   93032218.
  43. 1 2 Guérard, F.; Gestin, J.-F.; Brechbiel, M. W. (2013). "Production of [211At]-Astatinated Radiopharmaceuticals and Applications in Targeted α-Particle Therapy". Cancer Biotherapy and Radiopharmaceuticals. 28 (1): 1–20. doi:10.1089/cbr.2012.1292. PMC   3545490 . PMID   23075373.
  44. Champion, J.; Alliot, C.; Renault, E.; Mokili, B. M.; Chérel, M.; Galland, N.; Montavon, G. (2010). "Astatine Standard Redox Potentials and Speciation in Acidic Medium". The Journal of Physical Chemistry A. 114 (1): 576–582 (581). Bibcode:2010JPCA..114..576C. doi:10.1021/jp9077008. PMID   20014840. S2CID   15738065.
  45. Dolg, M.; Kuchle, W.; Stoll, H.; Preuss, H.; Schwerdtfeger, P. (1991). "Ab Initio Pseudopotentials for Hg to Rn: II. Molecular Calculations on the Hydrides of Hg to At and the Fluorides of Rn". Molecular Physics. 74 (6): 1265–1285 (1265, 1270, 1282). Bibcode:1991MolPh..74.1265D. doi:10.1080/00268979100102951.
  46. Saue, T.; Faegri, K.; Gropen, O. (1996). "Relativistic Effects on the Bonding of Heavy and Superheavy Hydrogen Halides". Chemical Physics Letters. 263 (3–4): 360–366 (361–362). Bibcode:1996CPL...263..360S. doi:10.1016/S0009-2614(96)01250-X.
  47. Barysz, M. (2010). Relativistic Methods for Chemists. Springer. p. 79. ISBN   978-1-4020-9974-8.
  48. Thayer, J. S. (2005). "Relativistic Effects and the Chemistry of the Heaviest Main-group elements". Journal of Chemical Education. 82 (11): 1721–1727 (1725). Bibcode:2005JChEd..82.1721T. doi:10.1021/ed082p1721.
  49. Wulfsberg, G. (2000). Inorganic Chemistry. University Science Books. p. 37. ISBN   978-1-891389-01-6.
  50. Smith, D. W. (1990). Inorganic Substances: A Prelude to the Study of Descriptive Inorganic Chemistry . Cambridge University Press. p.  135. ISBN   978-0-521-33738-0.
  51. Leimbach, D.; Sundberg, J.; Yangyang, G.; et al. (February 2020). "The electron affinity of astatine". Nature Communications. 11 (1): 3824. arXiv: 2002.11418 . Bibcode:2020NatCo..11.3824L. doi:10.1038/s41467-020-17599-2. PMC   7393155 . PMID   32733029.
  52. Anders, E. (1959). "Technetium and astatine chemistry". Annual Review of Nuclear Science . 9: 203–220. Bibcode:1959ARNPS...9..203A. doi: 10.1146/annurev.ns.09.120159.001223 .(subscription required)
  53. "Superheavy Element 117 Confirmed – On the Way to the "Island of Stability"". GSI Helmholtz Centre for Heavy Ion Research. Archived from the original on 3 August 2018. Retrieved 26 July 2015.
  54. 1 2 3 4 Nefedov, V. D.; Norseev, Yu. V.; Toropova, M. A.; Khalkin, Vladimir A. (1968). "Astatine". Russian Chemical Reviews. 37 (2): 87–98. Bibcode:1968RuCRv..37...87N. doi:10.1070/RC1968v037n02ABEH001603. S2CID   250775410.(subscription required)
  55. Aten, A. H. W. Jr.; Doorgeest, T.; Hollstein, U.; Moeken, H. P. (1952). "Section 5: Radiochemical Methods. Analytical Chemistry of Astatine". Analyst. 77 (920): 774–777. Bibcode:1952Ana....77..774A. doi:10.1039/AN9527700774.(subscription required)
  56. 1 2 3 4 5 6 Zuckerman & Hagen 1989, p. 31.
  57. 1 2 Zuckerman & Hagen 1989, p. 38.
  58. Chatterjee, Sayandev; Czerwinski, Kenneth R.; Fitzgerald, Hilary A.; Lakes, Andrew L.; Liao, Zuolei; Ludwig, Russell C.; McBride, Katie M.; Vlasenko, Vladislav P. (2020). "Novel Platforms for Drug Delivery Applications". Woodhead Publishing Series in Biomaterials. Woodhead Publishing. Subchapter 16.4.2: Redox behavior. doi:10.1016/B978-0-323-91376-8.00012-4.
  59. Kugler & Keller 1985, pp. 213–214.
  60. Kugler & Keller 1985, pp. 214–218.
  61. Kugler & Keller 1985, p. 211.
  62. Kugler & Keller 1985, pp. 109–110, 129, 213.
  63. Davidson, M. (2000). Contemporary boron chemistry. Royal Society of Chemistry. p. 146. ISBN   978-0-85404-835-9.
  64. 1 2 3 4 Zuckerman & Hagen 1989, p. 276.
  65. Elgqvist, J.; Hultborn, R.; Lindegren, S.; Palm, S. (2011). "Ovarian cancer: background and clinical perspectives". In Speer, S. (ed.). Targeted Radionuclide Therapy. Lippincott Williams & Wilkins. pp. 380–396 (383). ISBN   978-0-7817-9693-4.
  66. 1 2 3 Zuckerman & Hagen 1989, pp. 190–191.
  67. Brookhart, M.; Grant, B.; Volpe, A. F. (1992). "[(3,5-(CF3)2C6H3)4B]-[H(OEt2)2]+: a convenient reagent for generation and stabilization of cationic, highly electrophilic organometallic complexes". Organometallics . 11 (11): 3920–3922. doi:10.1021/om00059a071.
  68. Kugler & Keller 1985, p. 111.
  69. Sergentu, Dumitru-Claudiu; Teze, David; Sabatié-Gogova, Andréa; Alliot, Cyrille; Guo, Ning; Bassel, Fadel; Da Silva, Isidro; Deniaud, David; Maurice, Rémi; Champion, Julie; Galland, Nicolas; Montavon, Gilles (2016). "Advances on the Determination of the Astatine Pourbaix Diagram: Predomination of AtO(OH)2 over At in Basic Conditions". Chem. Eur. J. 22 (9): 2964–71. doi:10.1002/chem.201504403. PMID   26773333.
  70. Kugler & Keller 1985, p. 222.
  71. Lavrukhina & Pozdnyakov 1970, p. 238.
  72. 1 2 Kugler & Keller 1985, pp. 112, 192–193.
  73. Kugler & Keller 1985, p. 219.
  74. Zuckerman & Hagen 1989, pp. 192–193.
  75. Zuckerman & Hagen 1990, p. 212.
  76. Brinkman, G. A.; Aten, H. W. (1963). "Decomposition of Caesium Diiodo Astatate (I), (CsAtI2)". Radiochimica Acta. 2 (1): 48. doi:10.1524/ract.1963.2.1.48. S2CID   99398848.
  77. Zuckerman & Hagen 1990, p. 60.
  78. Zuckerman & Hagen 1989, p. 426.
  79. Appelman, E. H.; Sloth, E. N.; Studier, M. H. (1966). "Observation of Astatine Compounds by Time-of-Flight Mass Spectrometry". Inorganic Chemistry. 5 (5): 766–769. doi:10.1021/ic50039a016.
  80. Pitzer, K. S. (1975). "Fluorides of Radon and Element 118". Journal of the Chemical Society, Chemical Communications. 5 (18): 760b–761. doi:10.1039/C3975000760B.
  81. Bartlett, N.; Sladky, F. O. (1973). "The Chemistry of Krypton, Xenon and Radon". In Bailar, J. C.; Emeléus, H. J.; Nyholm, R.; et al. (eds.). Comprehensive Inorganic Chemistry. Vol. 1. Pergamon. pp. 213–330. ISBN   978-0-08-017275-0.
  82. Ball, P. (2002). The Ingredients: A Guided Tour of the Elements. Oxford University Press. pp. 100–102. ISBN   978-0-19-284100-1.
  83. Lavrukhina & Pozdnyakov 1970, pp. 227–228.
  84. Allison, F.; Murphy, E. J.; Bishop, E. R.; Sommer, A. L. (1931). "Evidence of the Detection of Element 85 in Certain Substances". Physical Review. 37 (9): 1178–1180. Bibcode:1931PhRv...37.1178A. doi:10.1103/PhysRev.37.1178.(subscription required)
  85. "Alabamine & Virginium". Time . 15 February 1932. Archived from the original on 30 September 2007.
  86. Trimble, R. F. (1975). "What Happened to Alabamine, Virginium, and Illinium?". Journal of Chemical Education. 52 (9): 585. Bibcode:1975JChEd..52..585T. doi:10.1021/ed052p585.(subscription required)
  87. MacPherson, H. G. (1934). "An Investigation of the Magneto-optic Method of Chemical Analysis". Physical Review. 47 (4): 310–315. Bibcode:1935PhRv...47..310M. doi:10.1103/PhysRev.47.310.
  88. 1 2 3 Mellor, J. W. (1965). A Comprehensive Treatise on Inorganic and Theoretical Chemistry. Longmans, Green. p.  1066. OCLC   13842122.
  89. 1 2 Burdette, S. C.; Thornton, B. F. (2010). "Finding Eka-Iodine: Discovery Priority in Modern Times" (PDF). Bulletin for the History of Chemistry. 35: 86–96. Archived (PDF) from the original on 9 October 2022.
  90. Scerri, E. (2013). A Tale of 7 Elements (Google Play ed.). Oxford University Press. pp.  188–190, 206. ISBN   978-0-19-539131-2.
  91. Karlik, B.; Bernert, T. (1942). "Über Eine Vermutete β-Strahlung des Radium A und die Natürliche Existenz des Elementes 85" [About a Suspected β-radiation of Radium A, and the Natural Existence of the Element 85]. Naturwissenschaften (in German). 30 (44–45): 685–686. Bibcode:1942NW.....30..685K. doi:10.1007/BF01487965. S2CID   6667655.(subscription required)
  92. Leigh-Smith, A.; Minder, W. (1942). "Experimental Evidence of the Existence of Element 85 in the Thorium Family". Nature. 150 (3817): 767–768. Bibcode:1942Natur.150..767L. doi:10.1038/150767a0. S2CID   4121704.(subscription required)
  93. Corson, MacKenzie & Segrè 1940.
  94. 1 2 Davis, Helen Miles (1959). The Chemical Elements (PDF) (2nd ed.). Science Service, Ballantine Books. p. 29. Archived from the original (PDF) on 23 August 2017. Retrieved 14 August 2016.
  95. Karlik, B.; Bernert, T. (1943). "Eine Neue Natürliche α-Strahlung" [A New Natural α-radiation]. Naturwissenschaften (in German). 31 (25–26): 298–299. Bibcode:1943NW.....31..298K. doi:10.1007/BF01475613. S2CID   38193384.(subscription required)
  96. Karlik, B.; Bernert, T. (1943). "Das Element 85 in den Natürlichen Zerfallsreihen" [The Element 85 in the Natural Decay Chains]. Zeitschrift für Physik (in German). 123 (1–2): 51–72. Bibcode:1944ZPhy..123...51K. doi:10.1007/BF01375144. S2CID   123906708.(subscription required)
  97. Lederer, C. M.; Hollander, J. M.; Perlman, I. (1967). Table of Isotopes (6th ed.). John Wiley & Sons. pp. 1–657.
  98. 1 2 Corson, D. R. (2003). "Astatine". Chemical & Engineering News. 81 (36): 158. doi:10.1021/cen-v081n036.p158.
  99. Corson, MacKenzie & Segrè 1940, pp. 672, 677.
  100. Hamilton, J. G.; Soley, M. H. (1940). "A Comparison of the Metabolism of Iodine and of Element 85 (Eka-Iodine)". Proceedings of the National Academy of Sciences. 26 (8): 483–489. Bibcode:1940PNAS...26..483H. doi: 10.1073/pnas.26.8.483 . PMC   1078214 . PMID   16588388.
  101. Neumann, H. M. (1957). "Solvent Distribution Studies of the Chemistry of Astatine". Journal of Inorganic and Nuclear Chemistry. 4 (5–6): 349–353. doi:10.1016/0022-1902(57)80018-9.
  102. Johnson, G. L.; Leininger, R. F.; Segrè, E. (1949). "Chemical Properties of Astatine. I". Journal of Chemical Physics. 17 (1): 1–10. Bibcode:1949JChPh..17....1J. doi:10.1063/1.1747034. hdl: 2027/mdp.39015086446914 . S2CID   95324453.
  103. Dreyer, I.; Dreyer, R.; Chalkin, V. A. (1979). "Cations of Astatine in Aqueous Solutions; Production and some Characteristics". Radiochemical and Radioanalytical Letters (in German). 36 (6): 389–398.
  104. 1 2 Aten, A. H. W. Jr. (1964). The Chemistry of Astatine. Advances in Inorganic Chemistry and Radiochemistry. Vol. 6. pp. 207–223. doi:10.1016/S0065-2792(08)60227-7. ISBN   978-0-12-023606-0.
  105. Nefedov, V. D.; Norseev, Yu. V.; Toropova, M. A.; Khalkin, V. A. (1968). "Astatine". Russian Chemical Reviews. 37 (2): 87–98. Bibcode:1968RuCRv..37...87N. doi:10.1070/RC1968v037n02ABEH001603. S2CID   250775410.
  106. 1 2 Fry, C.; Thoennessen, M. (2013). "Discovery of the astatine, radon, francium, and radium isotopes". Atomic Data and Nuclear Data Tables . 09 (5): 497–519. arXiv: 1205.5841 . Bibcode:2013ADNDT..99..497F. doi:10.1016/j.adt.2012.05.003. S2CID   12590893.
  107. Kokkonen, Henna. "Decay properties of the new isotopes 188At and 190At" (PDF). University of Jyväskylä. Retrieved 8 June 2023.
  108. 1 2 3 4 Lavrukhina & Pozdnyakov 1970, p. 229.
  109. Rankama, K. (1956). Isotope Geology (2nd ed.). Pergamon Press. p. 403. ISBN   978-0-470-70800-2.
  110. Lide, D. R., ed. (2004). CRC Handbook of Chemistry and Physics (85th ed.). CRC Press. pp.  14–10. ISBN   978-0-8493-0485-9.
  111. 1 2 Stwertka, Albert. A Guide to the Elements, Oxford University Press, 1996, p. 193. ISBN   0-19-508083-1
  112. Lavrukhina & Pozdnyakov 1970, p. 228–229.
  113. Asimov, I. (1957). Only a Trillion. Abelard-Schuman. p. 24.
  114. Kolthoff, I. M.; Elving, P. J., eds. (1964). Treatise on Analytical Chemistry. Part II: Analytical Chemistry of the Elements. Vol. 4. New York: Interscience Encyclopedia. p. 487.
  115. Kugler & Keller 1985, p. 4.
  116. Maiti, M.; Lahiri, S. (2011). "Production cross section of At radionuclides from 7Li+natPb and 9Be+natTl reactions". Physical Review C. 84 (6): 07601–07604 (07601). arXiv: 1109.6413 . Bibcode:2011PhRvC..84f7601M. doi:10.1103/PhysRevC.84.067601. S2CID   115321713.
  117. Greenwood & Earnshaw 2002, p. 796.
  118. Kugler & Keller 1985, p. 5.
  119. Barton, G. W.; Ghiorso, A.; Perlman, I. (1951). "Radioactivity of Astatine Isotopes". Physical Review. 82 (1): 13–19. Bibcode:1951PhRv...82...13B. doi:10.1103/PhysRev.82.13. hdl:2027/mdp.39015086480574.(subscription required)
  120. 1 2 3 4 5 6 Zalutsky, M. R.; Pruszynski, M. (2011). "Astatine-211: Production and Availability". Current Radiopharmaceuticals. 4 (3): 177–185. doi:10.2174/1874471011104030177. PMC   3503149 . PMID   22201707.
  121. 1 2 Larsen, R. H.; Wieland, B. W.; Zalutsky, M. R. J. (1996). "Evaluation of an Internal Cyclotron Target for the Production of 211At via the 209Bi (α,2n)211At reaction". Applied Radiation and Isotopes. 47 (2): 135–143. doi:10.1016/0969-8043(95)00285-5. PMID   8852627.
  122. Nagatsu, K.; Minegishi, K. H.; Fukada, M.; Suzuki, H.; Hasegawa, S.; Zhang, M. (2014). "Production of 211At by a vertical beam irradiation method". Applied Radiation and Isotopes. 94: 363–371. doi:10.1016/j.apradiso.2014.09.012. PMID   25439168.
  123. Barbet, J.; Bourgeois, M.; Chatal, J. (2014). "Cyclotron-Based Radiopharmaceuticals for Nuclear Medicine Therapy". In R. P.; Baum (eds.). Therapeutic Nuclear Medicine. Springer. pp. 95–104 (99). ISBN   978-3-540-36718-5.
  124. 1 2 Wilbur, D. S. (2001). "Overcoming the Obstacles to Clinical Evaluation of 211At-Labeled Radiopharmaceuticals". The Journal of Nuclear Medicine. 42 (10): 1516–1518. PMID   11585866.
  125. 1 2 3 4 Lavrukhina & Pozdnyakov 1970, p. 233.
  126. Gopalan, R. (2009). Inorganic Chemistry for Undergraduates. Universities Press. p. 547. ISBN   978-81-7371-660-7.
  127. Stigbrand, T.; Carlsson, J.; Adams, G. P. (2008). Targeted Radionuclide Tumor Therapy: Biological Aspects. Springer. p. 150. ISBN   978-1-4020-8695-3.
  128. Gyehong, G.; Chun, K.; Park, S. H.; Kim, B. (2014). "Production of α-particle emitting 211At using 45 MeV α-beam". Physics in Medicine and Biology. 59 (11): 2849–2860. Bibcode:2014PMB....59.2849K. doi:10.1088/0031-9155/59/11/2849. PMID   24819557. S2CID   21973246.
  129. Kugler & Keller 1985, pp. 95–106, 133–139.
  130. Lavrukhina & Pozdnyakov 1970, pp. 243–253.
  131. Kugler & Keller 1985, p. 97.
  132. Lindegren, S.; Bäck, T.; Jensen, H. J. (2001). "Dry-distillation of Astatine-211 from Irradiated Bismuth Targets: A Time-saving Procedure with High Recovery Yields". Applied Radiation and Isotopes. 55 (2): 157–160. doi:10.1016/S0969-8043(01)00044-6. PMID   11393754.
  133. Yordanov, A. T.; Pozzi, O.; Carlin, S.; Akabani, G. J.; Wieland, B.; Zalutsky, M. R. (2005). "Wet Harvesting of No-carrier-added 211At from an Irradiated 209Bi Target for Radiopharmaceutical Applications" . Journal of Radioanalytical and Nuclear Chemistry. 262 (3): 593–599. doi:10.1007/s10967-005-0481-7. S2CID   93179195.
  134. 1 2 Balkin, Ethan; Hamlin, Donald; Gagnon, Katherine; Chyan, Ming-Kuan; Pal, Sujit; Watanabe, Shigeki; Wilbur, D. (18 September 2013). "Evaluation of a Wet Chemistry Method for Isolation of Cyclotron Produced [211At]Astatine". Applied Sciences. 3 (3): 636–655. CiteSeerX   10.1.1.383.1903 . doi: 10.3390/app3030636 . ISSN   2076-3417.
  135. 1 2 3 Vértes, Nagy & Klencsár 2003, p. 337.
  136. Zalutsky, Michael; Vaidyanathan, Ganesan (1 September 2000). "Astatine-211-Labeled Radiotherapeutics An Emerging Approach to Targeted Alpha-Particle Radiotherapy". Current Pharmaceutical Design. 6 (14): 1433–1455. doi:10.2174/1381612003399275. PMID   10903402.
  137. Wilbur, D. Scott (20 February 2013). "Enigmatic astatine". Nature Chemistry. 5 (3): 246. Bibcode:2013NatCh...5..246W. doi: 10.1038/nchem.1580 . PMID   23422568.
  138. Vértes, Nagy & Klencsár 2003, p. 338.
  139. 1 2 Fisher, D. (1995). "Oral History of Dr. Patricia Wallace Durbin, PhD". Human Radiation Studies: Remembering the Early Years. United States Department of Energy, Office of Human Radiation Experiments. Retrieved 25 March 2015.
  140. Vaidyanathan, G.; Zalutsky, M. R. (2008). "Astatine Radiopharmaceuticals: Prospects and Problems". Current Radiopharmaceuticals. 1 (3): 177–196. doi:10.2174/1874471010801030177. PMC   2818997 . PMID   20150978.
  141. Lavrukhina & Pozdnyakov 1970, pp. 232–233.
  142. Odell, T. T. Jr.; Upton, A. C. (2013) [Softcover reprint of the hardcover 1st edition 1961]. "Late Effects of Internally Deposited Radioisotopes". In Schwiegk, H.; Turba, F. (eds.). Radioactive Isotopes in Physiology Diagnostics and Therapy[Radioaktive Isotope in Physiologie Diagnostik Und Therapie]. Springer-Verlag. pp. 375–392 (385). ISBN   978-3-642-49477-2.
  143. Keller, Cornelius; Wolf, Walter; Shani, Jashovam. "Radionuclides, 2. Radioactive Elements and Artificial Radionuclides". Ullmann's Encyclopedia of Industrial Chemistry . Weinheim: Wiley-VCH. doi:10.1002/14356007.o22_o15. ISBN   978-3527306732.

Bibliography