Polonium-210

Last updated
Polonium-210, 210Po
General
Symbol 210Po
Names polonium-210, 210Po, Po-210,
radium F
Protons (Z)84
Neutrons (N)126
Nuclide data
Natural abundance Trace
Half-life (t1/2)138.376±0.002 d [1]
Isotope mass 209.9828736 [2] Da
Spin 0
Parent isotopes 210Bi  (β)
Decay products 206Pb
Decay modes
Decay mode Decay energy (MeV)
Alpha decay 5.40753 [2]
Isotopes of polonium
Complete table of nuclides

Polonium-210 (210Po, Po-210, historically radium F) is an isotope of polonium. It undergoes alpha decay to stable 206Pb with a half-life of 138.376 days (about 4+12 months), the longest half-life of all naturally occurring polonium isotopes (210–218Po). [1] First identified in 1898, and also marking the discovery of the element polonium, 210Po is generated in the decay chain of uranium-238 and radium-226. 210Po is a prominent contaminant in the environment, mostly affecting seafood and tobacco. Its extreme toxicity is attributed to intense radioactivity, mostly due to alpha particles, which easily cause radiation damage, including cancer in surrounding tissue. The specific activity of 210
Po
is 166 TBq/g, i.e., 1.66 × 1014 Bq/g. At the same time, 210Po is not readily detected by common radiation detectors, because its gamma rays have a very low energy. Therefore, 210
Po
can be considered as a quasi-pure alpha emitter.

Contents

History

The decay chain of uranium-238, known as the uranium series or radium series, of which polonium-210 is a member Decay chain(4n+2, Uranium series).svg
The decay chain of uranium-238, known as the uranium series or radium series, of which polonium-210 is a member
Schematic of the final steps of the s-process. The red path represents the sequence of neutron captures; blue and cyan arrows represent beta decay, and the green arrow represents the alpha decay of Po. It is the short half-lives of Bi and Po that prevent the formation of heavier elements, instead resulting in a cycle of four neutron captures, two beta decays, and an alpha decay. S-R-processes-atomic-mass-201-to-210.svg
Schematic of the final steps of the s-process. The red path represents the sequence of neutron captures; blue and cyan arrows represent beta decay, and the green arrow represents the alpha decay of Po. It is the short half-lives of Bi and Po that prevent the formation of heavier elements, instead resulting in a cycle of four neutron captures, two beta decays, and an alpha decay.

In 1898, Marie and Pierre Curie discovered a strongly radioactive substance in pitchblende and determined that it was a new element; it was one of the first radioactive elements discovered. Having identified it as such, they named the element polonium after Marie's home country, Poland. Willy Marckwald discovered a similar radioactive activity in 1902 and named it radio-tellurium, and at roughly the same time, Ernest Rutherford identified the same activity in his analysis of the uranium decay chain and named it radium F (originally radium E). By 1905, Rutherford concluded that all these observations were due to the same substance, 210Po. Further discoveries and the concept of isotopes, first proposed in 1913 by Frederick Soddy, firmly placed 210Po as the penultimate step in the uranium series. [3]

In 1943, 210Po was studied as a possible neutron initiator in nuclear weapons, as part of the Dayton Project. In subsequent decades, concerns for the safety of workers handling 210Po led to extensive studies on its health effects. [4]

In the 1950s, scientists of the United States Atomic Energy Commission at Mound Laboratories, Ohio explored the possibility of using 210Po in radioisotope thermoelectric generators (RTGs) as a heat source to power satellites. A 2.5-watt atomic battery using 210Po was developed by 1958. However, the isotope plutonium-238 was chosen instead, as it has a longer half-life of 87.7 years. [5]

Polonium-210 was used to kill Russian dissident and ex-FSB officer Alexander V. Litvinenko in 2006, [6] [7] and was suspected as a possible cause of Yasser Arafat's death, following exhumation and analysis of his corpse in 2012–2013. [8] The radioisotope may also have been used to kill Yuri Shchekochikhin, Lecha Islamov and Roman Tsepov. [9]

Decay properties

210Po is an alpha emitter that has a half-life of 138.376 days; [1] it decays directly to stable 206Pb. The majority of the time, 210Po decays by emission of an alpha particle only, not by emission of an alpha particle and a gamma ray; about one in 100,000 decays results in the emission of a gamma ray. [10]

This low gamma ray production rate makes it more difficult to find and identify this isotope. Rather than gamma ray spectroscopy, alpha spectroscopy is the best method of measuring this isotope.

Owing to its much shorter half-life, a milligram of 210Po emits as many alpha particles per second as 5 grams of 226Ra. [11] A few curies of 210Po emit a blue glow caused by excitation of surrounding air.

210Po occurs in minute amounts in nature, where it is the penultimate isotope in the uranium series decay chain. It is generated via beta decay from 210 Pb and 210 Bi.

The astrophysical s-process is terminated by the decay of 210Po, as the neutron flux is insufficient to lead to further neutron captures in the short lifetime of 210Po. Instead, 210Po alpha decays to 206Pb, which then captures more neutrons to become 210Po and repeats the cycle, thus consuming the remaining neutrons. This results in a buildup of lead and bismuth, and ensures that heavier elements such as thorium and uranium are only produced in the much faster r-process. [12]

Production

Deliberate

Although 210Po occurs in trace amounts in nature, it is not abundant enough (0.1 ppb) for extraction from uranium ore to be feasible. Instead, most 210Po is produced synthetically, through neutron bombardment of 209Bi in a nuclear reactor. This process converts 209Bi to 210Bi, which beta decays to 210Po with a five-day half-life. Through this method, approximately 8 grams (0.28 oz) of 210Po are produced in Russia and shipped to the United States every month for commercial applications. [4] By irradiating certain bismuth salts containing light element nuclei such as beryllium, a cascading (α,n) reaction can also be induced to produce 210Po in large quantities. [13]

Byproduct

The production of polonium-210 is a downside to reactors cooled with lead-bismuth eutectic rather than pure lead. However, given the eutectic properties of this alloy, some proposed Generation IV reactor designs still rely on lead-bismuth.

Applications

A single gram of 210Po generates 140 watts of power. [14] Because it emits many alpha particles, which are stopped within a very short distance in dense media and release their energy, 210Po has been used as a lightweight heat source to power thermoelectric cells in artificial satellites; for instance, a 210Po heat source was also in each of the Lunokhod rovers deployed on the surface of the Moon, to keep their internal components warm during the lunar nights. [15] Some anti-static brushes, used for neutralizing static electricity on materials like photographic film, contain a few microcuries of 210Po as a source of charged particles. [16] 210Po was also used in initiators for atomic bombs through the (α,n) reaction with beryllium. [17] Small neutron sources reliant on the (α,n) reaction also usually use polonium as a convenient source of alpha particles due to its comparatively low gamma emissions (allowing easy shielding) and high specific activity.

Hazards

210Po is extremely toxic; it and other polonium isotopes are some of the most radiotoxic substances to humans. [6] [18] With one microgram of 210Po being more than enough to kill the average adult, it is 250,000 times more toxic than hydrogen cyanide by weight. [19] One gram of 210Po would hypothetically be enough to kill 50 million people and sicken another 50 million. [6] This is a consequence of its ionizing alpha radiation, as alpha particles are especially damaging to organic tissues inside the body. However, 210Po does not pose a radiation hazard when contained outside the body. [20] The alpha particles it produces cannot penetrate the outer layer of dead skin cells. [21]

The toxicity of 210Po stems entirely from its radioactivity. It is not chemically toxic in itself, but its solubility in aqueous solution as well as that of its salts poses a hazard because its spread throughout the body is facilitated in solution. [6] Intake of 210Po occurs primarily through contaminated air, food, or water, as well as through open wounds. Once inside the body, 210Po concentrates in soft tissues (especially in the reticuloendothelial system) and the bloodstream. Its biological half-life is approximately 50 days. [22]

In the environment, 210Po can accumulate in seafood. [23] It has been detected in various organisms in the Baltic Sea, where it can propagate in, and thus contaminate, the food chain. [18] 210Po is also known to contaminate vegetation, primarily originating from the decay of atmospheric radon-222 and absorption from soil. [24]

In particular, 210Po attaches to, and concentrates in, tobacco leaves. [4] [22] Elevated concentrations of 210Po in tobacco were documented as early as 1964, and cigarette smokers were thus found to be exposed to considerably greater doses of radiation from 210Po and its parent 210Pb. [24] Heavy smokers may be exposed to the same amount of radiation (estimates vary from 100  µSv [18] to 160 mSv [25] per year) as individuals in Poland were from Chernobyl fallout traveling from Ukraine. [18] As a result, 210Po is most dangerous when inhaled from cigarette smoke. [26]

Polonium-210 has been used in silent killings. Significant amounts were found in the body of Alexander Litvinenko [27] at the time of his murder in London in 2006.

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">Astatine</span> Chemical element, symbol At and atomic number 85

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.

<span class="mw-page-title-main">Alpha decay</span> Type of radioactive decay

Alpha decay or α-decay is a type of radioactive decay in which an atomic nucleus emits an alpha particle and thereby transforms or 'decays' into a different atomic nucleus, with a mass number that is reduced by four and an atomic number that is reduced by two. An alpha particle is identical to the nucleus of a helium-4 atom, which consists of two protons and two neutrons. It has a charge of +2 e and a mass of 4 Da. For example, uranium-238 decays to form thorium-234.

<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.

<span class="mw-page-title-main">Radium</span> Chemical element, symbol Ra and atomic number 88

Radium is a chemical element; it has symbol Ra and atomic number 88. It is the sixth element in group 2 of the periodic table, also known as the alkaline earth metals. Pure radium is silvery-white, but it readily reacts with nitrogen (rather than oxygen) upon exposure to air, forming a black surface layer of radium nitride (Ra3N2). All isotopes of radium are radioactive, the most stable isotope being radium-226 with a half-life of 1,600 years. When radium decays, it emits ionizing radiation as a by-product, which can excite fluorescent chemicals and cause radioluminescence.

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.

<span class="mw-page-title-main">Beta particle</span> Ionizing radiation

A beta particle, also called beta ray or beta radiation, is a high-energy, high-speed electron or positron emitted by the radioactive decay of an atomic nucleus during the process of beta decay. There are two forms of beta decay, β decay and β+ decay, which produce electrons and positrons respectively.

<span class="mw-page-title-main">Radioactive decay</span> Emissions from unstable atomic nuclei

Radioactive decay is the process by which an unstable atomic nucleus loses energy by radiation. A material containing unstable nuclei is considered radioactive. Three of the most common types of decay are alpha, beta, and gamma decay. The weak force is the mechanism that is responsible for beta decay, while the other two are governed by the electromagnetism and nuclear force.

<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.

<span class="mw-page-title-main">Uranium-238</span> Isotope of uranium

Uranium-238 is the most common isotope of uranium found in nature, with a relative abundance of 99%. Unlike uranium-235, it is non-fissile, which means it cannot sustain a chain reaction in a thermal-neutron reactor. However, it is fissionable by fast neutrons, and is fertile, meaning it can be transmuted to fissile plutonium-239. 238U cannot support a chain reaction because inelastic scattering reduces neutron energy below the range where fast fission of one or more next-generation nuclei is probable. Doppler broadening of 238U's neutron absorption resonances, increasing absorption as fuel temperature increases, is also an essential negative feedback mechanism for reactor control.

<span class="mw-page-title-main">Radiochemistry</span> Chemistry of radioactive materials

Radiochemistry is the chemistry of radioactive materials, where radioactive isotopes of elements are used to study the properties and chemical reactions of non-radioactive isotopes. Much of radiochemistry deals with the use of radioactivity to study ordinary chemical reactions. This is very different from radiation chemistry where the radiation levels are kept too low to influence the chemistry.

Uranium (92U) is a naturally occurring radioactive element that has no stable isotope. It has two primordial isotopes, uranium-238 and uranium-235, that have long half-lives and are found in appreciable quantity in the Earth's crust. The decay product uranium-234 is also found. Other isotopes such as uranium-233 have been produced in breeder reactors. In addition to isotopes found in nature or nuclear reactors, many isotopes with far shorter half-lives have been produced, ranging from 214U to 242U. The standard atomic weight of natural uranium is 238.02891(3).

Lead (82Pb) has four observationally stable isotopes: 204Pb, 206Pb, 207Pb, 208Pb. Lead-204 is entirely a primordial nuclide and is not a radiogenic nuclide. The three isotopes lead-206, lead-207, and lead-208 represent the ends of three decay chains: the uranium series, the actinium series, and the thorium series, respectively; a fourth decay chain, the neptunium series, terminates with the thallium isotope 205Tl. The three series terminating in lead represent the decay chain products of long-lived primordial 238U, 235U, and 232Th. Each isotope also occurs, to some extent, as primordial isotopes that were made in supernovae, rather than radiogenically as daughter products. The fixed ratio of lead-204 to the primordial amounts of the other lead isotopes may be used as the baseline to estimate the extra amounts of radiogenic lead present in rocks as a result of decay from uranium and thorium.

Americium (95Am) is an artificial element, and thus a standard atomic weight cannot be given. Like all artificial elements, it has no known stable isotopes. The first isotope to be synthesized was 241Am in 1944. The artificial element decays by ejecting alpha particles. Americium has an atomic number of 95. Despite 243
Am
being an order of magnitude longer lived than 241
Am
, the former is harder to obtain than the latter as more of it is present in spent nuclear fuel.

Induced radioactivity, also called artificial radioactivity or man-made radioactivity, is the process of using radiation to make a previously stable material radioactive. The husband-and-wife team of Irène Joliot-Curie and Frédéric Joliot-Curie discovered induced radioactivity in 1934, and they shared the 1935 Nobel Prize in Chemistry for this discovery.

The decay scheme of a radioactive substance is a graphical presentation of all the transitions occurring in a decay, and of their relationships. Examples are shown below.

<span class="mw-page-title-main">Valley of stability</span> Characterization of nuclide stability

In nuclear physics, the valley of stability is a characterization of the stability of nuclides to radioactivity based on their binding energy. Nuclides are composed of protons and neutrons. The shape of the valley refers to the profile of binding energy as a function of the numbers of neutrons and protons, with the lowest part of the valley corresponding to the region of most stable nuclei. The line of stable nuclides down the center of the valley of stability is known as the line of beta stability. The sides of the valley correspond to increasing instability to beta decay. The decay of a nuclide becomes more energetically favorable the further it is from the line of beta stability. The boundaries of the valley correspond to the nuclear drip lines, where nuclides become so unstable they emit single protons or single neutrons. Regions of instability within the valley at high atomic number also include radioactive decay by alpha radiation or spontaneous fission. The shape of the valley is roughly an elongated paraboloid corresponding to the nuclide binding energies as a function of neutron and atomic numbers.

Naturally occurring radioactive materials (NORM) and technologically enhanced naturally occurring radioactive materials (TENORM) consist of materials, usually industrial wastes or by-products enriched with radioactive elements found in the environment, such as uranium, thorium and potassium and any of their decay products, such as radium and radon. Produced water discharges and spills are a good example of entering NORMs into the surrounding environment.

<span class="mw-page-title-main">Alpha particle</span> Ionizing radiation particle of two protons and two neutrons

Alpha particles, also called alpha rays or alpha radiation, consist of two protons and two neutrons bound together into a particle identical to a helium-4 nucleus. They are generally produced in the process of alpha decay but may also be produced in other ways. Alpha particles are named after the first letter in the Greek alphabet, α. The symbol for the alpha particle is α or α2+. Because they are identical to helium nuclei, they are also sometimes written as He2+
or 4
2
He2+
indicating a helium ion with a +2 charge. Once the ion gains electrons from its environment, the alpha particle becomes a normal helium atom 4
2
He
.

Radium-223 is an isotope of radium with an 11.4-day half-life. It was discovered in 1905 by T. Godlewski, a Polish chemist from Kraków, and was historically known as actinium X (AcX). Radium-223 dichloride is an alpha particle-emitting radiotherapy drug that mimics calcium and forms complexes with hydroxyapatite at areas of increased bone turnover. The principal use of radium-223, as a radiopharmaceutical to treat metastatic cancers in bone, takes advantage of its chemical similarity to calcium, and the short range of the alpha radiation it emits.

References

  1. 1 2 3 Nuclear Data Center at KAERI; Table of Nuclides http://atom.kaeri.re.kr/nuchart/?zlv=1
  2. 1 2 Wang, M.; Audi, G.; Kondev, F. G.; Huang, W. J.; Naimi, S.; Xu, X. (2017). "The AME2016 atomic mass evaluation (II). Tables, graphs, and references" (PDF). Chinese Physics C. 41 (3): 030003-1–030003-442. doi:10.1088/1674-1137/41/3/030003.
  3. Thoennessen, M. (2016). The Discovery of Isotopes: A Complete Compilation. Springer. pp. 6–8. doi:10.1007/978-3-319-31763-2. ISBN   978-3-319-31761-8. LCCN   2016935977.
  4. 1 2 3 Roessler, G. (2007). "Why 210Po?" (PDF). Health Physics News. Vol. 35, no. 2. Health Physics Society. Archived (PDF) from the original on 2014-04-03. Retrieved 2019-06-20.
  5. Idaho National Laboratory (2015). "The Early Years: Space Nuclear Power Systems Take Flight" (PDF). Atomic power in space II: a history of space nuclear power and propulsion in the United States. pp. 2–5. OCLC   931595589.
  6. 1 2 3 4 McFee, R. B.; Leikin, J. B. (2009). "Death by polonium-210: lessons learned from the murder of former Soviet spy Alexander Litvinenko". Seminars in Diagnostic Pathology. 26 (1): 61–67. doi:10.1053/j.semdp.2008.12.003. PMID   19292030.
  7. Cowell, A. (November 24, 2006). "Radiation Poisoning Killed Ex-Russian Spy". The New York Times . Archived from the original on June 19, 2019. Retrieved June 19, 2019.
  8. "Arafat's death: what is Polonium-210?". Al Jazeera . July 10, 2012. Archived from the original on June 19, 2019. Retrieved June 19, 2019.
  9. Sweeney, J. (2022). Killer in the Kremlin. Penguin. p. 120. ISBN   9781804991206.
  10. "210Po A Decay". Korea Atomic Energy Research Institute . Archived from the original on February 24, 2015.
  11. C. R. Hammond. "The Elements" (PDF). Fermi National Accelerator Laboratory. pp. 4–22. Archived (PDF) from the original on 2008-06-26. Retrieved 2019-06-19.
  12. Burbidge, E. M.; Burbidge, G. R.; Fowler, W. A.; Hoyle, F. (1957). "Synthesis of the Elements in Stars". Reviews of Modern Physics . 29 (4): 547–650. Bibcode:1957RvMP...29..547B. doi: 10.1103/RevModPhys.29.547 .
  13. Lim, Solomon (2023). "Neutronic Chain Reactions for Polonium-210 Production". SSRN. doi:10.2139/ssrn.4469519.
  14. "Polonium" (PDF). Argonne National Laboratory. Archived from the original (PDF) on 2012-03-10.
  15. A. Wilson, Solar System Log, (London: Jane's Publishing Company Ltd, 1987), p. 64.[ ISBN missing ]
  16. "Staticmaster Alpha Ionizing Brush". Company 7. Archived from the original on 2018-09-27. Retrieved 2019-06-19.
  17. Hoddeson, L.; Henriksen, P. W.; Meade, R. A. (2004). Critical Assembly: A Technical History of Los Alamos During the Oppenheimer Years, 1943–1945. Cambridge University Press. ISBN   978-0-521-54117-6.
  18. 1 2 3 4 Skwarzec, B.; Strumińska, D. I.; Boryło, A. (2006). "Radionuclides of iron (55Fe), nickel (63Ni), polonium (210Po), uranium (234U, 235U, 238U), and plutonium (238Pu, 239+240Pu, 241Pu) in Poland and Baltic Sea environment" (PDF). Nukleonika. 51: S45–S51. Archived (PDF) from the original on 2019-06-19. Retrieved 2019-06-19.
  19. Ahmed, M. F.; Alam, L.; Mohamed, C. A. R.; Mokhtar, M. B.; Ta, G. C. (2018). "Health risk of polonium-210 ingestion via drinking water: An experience of Malaysia". International Journal of Environmental Research and Public Health. 15 (10): 2056–1–2056–19. doi: 10.3390/ijerph15102056 . PMC   6210456 . PMID   30241360.
  20. "Radiation Studies: CDC - Radiation: Polonium-210 | CDC RSB". www.cdc.gov. 2019-02-11. Retrieved 2022-11-14.
  21. "Penetration Abilities of Different Types of Radiation". www.cdc.gov. Retrieved 2022-11-14.
  22. 1 2 Frequently asked questions about polonium-210 (PDF) (Report). Centers for Disease Control and Prevention. Archived (PDF) from the original on 7 June 2017. Retrieved 19 June 2019.
  23. Richter, F.; Wagmann, M.; Zehringer, M. (2012). "Polonium – on the Trace of a Powerful Alpha Nuclide in the Environment". Chimia International Journal for Chemistry. 66 (3): 131. doi: 10.2533/chimia.2012.131 . Archived from the original on 2019-02-17. Retrieved 2019-06-19.
  24. 1 2 Persson, B. R. R.; Holm, E. (2009). Polonium-210 and Lead-210 in the Terrestrial environment: A historical review. International Topical Conference on Po and Radioactive Pb Isotopes. Seville, Spain.
  25. "F. Typical Sources of Radiation Exposure". National Institute of Health. Archived from the original on 2013-06-13. Retrieved 2019-06-20.
  26. Radford, E. P.; Hunt, V. R. (1964). "Polonium-210: A Volatile Radioelement in Cigarettes". Science. 143 (3603): 247–249. Bibcode:1964Sci...143..247R. doi:10.1126/science.143.3603.247. JSTOR   1712451. PMID   14078362. S2CID   23455633.
  27. "Poisoning of Alexander Litvinenko", Wikipedia, 2023-11-04, retrieved 2023-11-09