Plutonium(IV) oxide

Last updated
Plutonium(IV) oxide
Pudioxide.png
Plutonium oxide.png
Names
IUPAC name
Plutonium(IV) oxide
Systematic IUPAC name
Plutonium(4+) oxide
Other names
Plutonium dioxide
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.031.840 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 235-037-3
PubChem CID
  • InChI=1S/2O.Pu/q2*-2;+4 Yes check.svgY
    Key: FLDALJIYKQCYHH-UHFFFAOYSA-N Yes check.svgY
  • [O-2].[O-2].[Pu+4]
Properties
O2Pu
Molar mass 276 g·mol−1
AppearanceDark yellow crystals
Density 11.5 g cm−3
Melting point 2,744 °C (4,971 °F; 3,017 K)
Boiling point 2,800 °C (5,070 °F; 3,070 K)
Structure
Fluorite (cubic), cF12
Fm3m, No. 225
a = 539.5 pm [1]
Tetrahedral (O2−); cubic (PuIV)
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
Radioactive
NFPA 704 (fire diamond)
NFPA 704.svgHealth 4: Very short exposure could cause death or major residual injury. E.g. VX gasFlammability 0: Will not burn. E.g. waterInstability 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g. liquid nitrogenSpecial hazard RA: Radioactive. E.g. plutonium
4
0
0
Radiation warning symbol 3.svg
Flash point non-flammable
Related compounds
Other cations
Uranium(IV) oxide
Neptunium(IV) oxide
Americium(IV) oxide
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

Plutonium(IV) oxide, or plutonia, is a chemical compound with the formula Pu O 2. This high melting-point solid is a principal compound of plutonium. It can vary in color from yellow to olive green, depending on the particle size, temperature and method of production. [2]

Contents

Structure

PuO2 crystallizes in the fluorite motif, with the Pu4+ centers organized in a face-centered cubic array and oxide ions occupying tetrahedral holes. [3] PuO2 owes its utility as a nuclear fuel to the fact that vacancies in the octahedral holes allows room for fission products. In nuclear fission, one atom of plutonium splits into two. The vacancy of the octahedral holes provides room for the new product and allows the PuO2 monolith to retain its structural integrity.[ citation needed ]

Properties

Plutonium dioxide is a stable ceramic material with an extremely low solubility in water and with a high melting point (2,744 °C). The melting point was revised upwards in 2011 by several hundred degrees, based on evidence from rapid laser melting studies which avoid contamination by any container material. [4]

Due to the radioactive alpha decay of plutonium, PuO2 is warm to the touch.[ citation needed ] As with all plutonium compounds, it is subject to control under the Nuclear Non-Proliferation Treaty.

Synthesis

Plutonium spontaneously oxidizes to PuO2 in an atmosphere of oxygen. Plutonium dioxide is mainly produced by calcination of plutonium(IV) oxalate, Pu(C2O4)2·6H2O, at 300 °C. Plutonium oxalate is obtained during the reprocessing of nuclear fuel as plutonium is dissolved in a solution of nitric and hydrofluoric acid. [5] Plutonium dioxide can also be recovered from molten-salt breeder reactors by adding sodium carbonate to the fuel salt after any remaining uranium is removed from the salt as its hexafluoride.

Applications

A pellet of dioxide of plutonium-238 displays incandescence after prolonged time of thermal isolation under asbestos. Plutonium pellet.jpg
A pellet of dioxide of plutonium-238 displays incandescence after prolonged time of thermal isolation under asbestos.

PuO2, along with UO2, is used in MOX fuels for nuclear reactors. Plutonium-238 dioxide is used as fuel for several deep-space spacecraft such as the Cassini, Voyager, Galileo and New Horizons probes as well as in the Curiosity and Perseverance rovers on Mars. The isotope decays by emitting α-particles, which then generate heat (see radioisotope thermoelectric generator). There have been concerns that an accidental re-entry into Earth's atmosphere from orbit might lead to the break-up and/or burn-up of a spacecraft, resulting in the dispersal of the plutonium, either over a large tract of the planetary surface or within the upper atmosphere. However, although at least two spacecraft carrying PuO2 RTGs have reentered the Earth's atmosphere and burned up (Nimbus B-1 in May 1968 and the Apollo 13 Lunar Module in April 1970), [6] [7] the RTGs from both spacecraft survived reentry and impact intact, and no environmental contamination was noted in either instance; in fact, the Nimbus RTG was recovered intact from the Pacific Ocean seafloor and launched aboard Nimbus 3 one year later. In any case, RTGs since the mid-1960s have been designed to remain intact in the event of reentry and impact, following the 1964 launch failure of Transit 5-BN-3 (the early-generation plutonium RTG on board disintegrated upon reentry and dispersed radioactive material into the atmosphere north of Madagascar, prompting a redesign of all U.S. RTGs then in use or under development). [8]

Physicist Peter Zimmerman, following up a suggestion by Ted Taylor, demonstrated that a low-yield (1-kiloton) nuclear weapon could be made relatively easily from plutonium dioxide. [9] Such bomb would require a considerably larger critical mass than one made from elemental plutonium (almost three times larger, even with the dioxide at maximum crystal density; if the dioxide were in powder form, as is often encountered, the critical mass would be much higher still), due both to the lower density of plutonium in dioxide compared with elemental plutonium and to the added inert mass of the air contained. [10]

Toxicology

The behavior of plutonium dioxide in the body varies with the way in which it is taken. When ingested, most of it will be eliminated from the body quite rapidly in body wastes, [11] but a small part will dissolve into ions in acidic gastric juice and cross the blood barrier, depositing itself in other chemical forms in other organs such as in phagocytic cells of lung, bone marrow and liver. [12]

In particulate form, plutonium dioxide at a particle size less than 10 µm [13] is radiotoxic if inhaled due to its strong alpha-emission. [14]

See also

Related Research Articles

<span class="mw-page-title-main">Americium</span> Chemical element, symbol Am and atomic number 95

Americium is a synthetic chemical element; it has symbol Am and atomic number 95. It is radioactive and a transuranic member of the actinide series in the periodic table, located under the lanthanide element europium and was thus named after the Americas by analogy.

The actinide or actinoid series encompasses at least the 14 metallic chemical elements in the 5f series, with atomic numbers from 89 to 102, actinium through nobelium. The actinide series derives its name from the first element in the series, actinium. The informal chemical symbol An is used in general discussions of actinide chemistry to refer to any actinide.

<span class="mw-page-title-main">Curium</span> Chemical element, symbol Cm and atomic number 96

Curium is a synthetic chemical element; it has symbol Cm and atomic number 96. This transuranic actinide element was named after eminent scientists Marie and Pierre Curie, both known for their research on radioactivity. Curium was first intentionally made by the team of Glenn T. Seaborg, Ralph A. James, and Albert Ghiorso in 1944, using the cyclotron at Berkeley. They bombarded the newly discovered element plutonium with alpha particles. This was then sent to the Metallurgical Laboratory at University of Chicago where a tiny sample of curium was eventually separated and identified. The discovery was kept secret until after the end of World War II. The news was released to the public in November 1947. Most curium is produced by bombarding uranium or plutonium with neutrons in nuclear reactors – one tonne of spent nuclear fuel contains ~20 grams of curium.

<span class="mw-page-title-main">Nuclear chain reaction</span> When one nuclear reaction causes more

In nuclear physics, a nuclear chain reaction occurs when one single nuclear reaction causes an average of one or more subsequent nuclear reactions, thus leading to the possibility of a self-propagating series or "positive feedback loop" of these reactions. The specific nuclear reaction may be the fission of heavy isotopes. A nuclear chain reaction releases several million times more energy per reaction than any chemical reaction.

<span class="mw-page-title-main">Nuclear fuel cycle</span> Process of manufacturing and consuming nuclear fuel

The nuclear fuel cycle, also called nuclear fuel chain, is the progression of nuclear fuel through a series of differing stages. It consists of steps in the front end, which are the preparation of the fuel, steps in the service period in which the fuel is used during reactor operation, and steps in the back end, which are necessary to safely manage, contain, and either reprocess or dispose of spent nuclear fuel. If spent fuel is not reprocessed, the fuel cycle is referred to as an open fuel cycle ; if the spent fuel is reprocessed, it is referred to as a closed fuel cycle.

<span class="mw-page-title-main">Nuclear reprocessing</span> Chemical operations that separate fissile material from spent fuel to be recycled as new fuel

Nuclear reprocessing is the chemical separation of fission products and actinides from spent nuclear fuel. Originally, reprocessing was used solely to extract plutonium for producing nuclear weapons. With commercialization of nuclear power, the reprocessed plutonium was recycled back into MOX nuclear fuel for thermal reactors. The reprocessed uranium, also known as the spent fuel material, can in principle also be re-used as fuel, but that is only economical when uranium supply is low and prices are high. Nuclear reprocessing may extend beyond fuel and include the reprocessing of other nuclear reactor material, such as Zircaloy cladding.

Mixed oxide fuel, commonly referred to as MOX fuel, is nuclear fuel that contains more than one oxide of fissile material, usually consisting of plutonium blended with natural uranium, reprocessed uranium, or depleted uranium. MOX fuel is an alternative to the low-enriched uranium fuel used in the light-water reactors that predominate nuclear power generation.

<span class="mw-page-title-main">Radioisotope thermoelectric generator</span> Electrical generator that uses heat from radioactive decay

A radioisotope thermoelectric generator, sometimes referred to as a radioisotope power system (RPS), is a type of nuclear battery that uses an array of thermocouples to convert the heat released by the decay of a suitable radioactive material into electricity by the Seebeck effect. This type of generator has no moving parts and is ideal for deployment in remote and harsh environments for extended periods with no risk of parts wearing out or malfunctioning.

<span class="mw-page-title-main">Fast-neutron reactor</span> Nuclear reactor where fast neutrons maintain a fission chain reaction

A fast-neutron reactor (FNR) or fast-spectrum reactor or simply a fast reactor is a category of nuclear reactor in which the fission chain reaction is sustained by fast neutrons, as opposed to slow thermal neutrons used in thermal-neutron reactors. Such a fast reactor needs no neutron moderator, but requires fuel that is relatively rich in fissile material when compared to that required for a thermal-neutron reactor. Around 20 land based fast reactors have been built, accumulating over 400 reactor years of operation globally. The largest of this was the Superphénix Sodium cooled fast reactor in France that was designed to deliver 1,242 MWe. Fast reactors have been intensely studied since the 1950s, as they provide certain advantages over the existing fleet of water cooled and water moderated reactors. These are:

<span class="mw-page-title-main">Nuclear fuel</span> Material fuelling nuclear reactors

Nuclear fuel is material used in nuclear power stations to produce heat to power turbines. Heat is created when nuclear fuel undergoes nuclear fission.

<span class="mw-page-title-main">Plutonium-239</span> Isotope of plutonium

Plutonium-239 is an isotope of plutonium. Plutonium-239 is the primary fissile isotope used for the production of nuclear weapons, although uranium-235 is also used for that purpose. Plutonium-239 is also one of the three main isotopes demonstrated usable as fuel in thermal spectrum nuclear reactors, along with uranium-235 and uranium-233. Plutonium-239 has a half-life of 24,110 years.

<span class="mw-page-title-main">Uranium dioxide</span> Chemical compound

Uranium dioxide or uranium(IV) oxide , also known as urania or uranous oxide, is an oxide of uranium, and is a black, radioactive, crystalline powder that naturally occurs in the mineral uraninite. It is used in nuclear fuel rods in nuclear reactors. A mixture of uranium and plutonium dioxides is used as MOX fuel. Prior to 1960, it was used as yellow and black color in ceramic glazes and glass.

<span class="mw-page-title-main">Plutonium-238</span> Isotope of plutonium used in radioisotope thermoelectric generators

Plutonium-238 is a radioactive isotope of plutonium that has a half-life of 87.7 years.

<span class="mw-page-title-main">Spent nuclear fuel</span> Nuclear fuel thats been irradiated in a nuclear reactor

Spent nuclear fuel, occasionally called used nuclear fuel, is nuclear fuel that has been irradiated in a nuclear reactor. It is no longer useful in sustaining a nuclear reaction in an ordinary thermal reactor and, depending on its point along the nuclear fuel cycle, it will have different isotopic constituents than when it started.

The Systems Nuclear Auxiliary POWER (SNAP) program was a program of experimental radioisotope thermoelectric generators (RTGs) and space nuclear reactors flown during the 1960s by NASA.

<span class="mw-page-title-main">Plutonium in the environment</span> Plutonium present within the environment

Since the mid-20th century, plutonium in the environment has been primarily produced by human activity. The first plants to produce plutonium for use in cold war atomic bombs were at the Hanford nuclear site, in Washington, and Mayak nuclear plant, in Chelyabinsk Oblast, Russia. Over a period of four decades, "both released more than 200 million curies of radioactive isotopes into the surrounding environment – twice the amount expelled in the Chernobyl disaster in each instance".

<span class="mw-page-title-main">Plutonium</span> Chemical element, symbol Pu and atomic number 94

Plutonium is a chemical element; it has symbol Pu and atomic number 94. It is an actinide metal of silvery-gray appearance that tarnishes when exposed to air, and forms a dull coating when oxidized. The element normally exhibits six allotropes and four oxidation states. It reacts with carbon, halogens, nitrogen, silicon, and hydrogen. When exposed to moist air, it forms oxides and hydrides that can expand the sample up to 70% in volume, which in turn flake off as a powder that is pyrophoric. It is radioactive and can accumulate in bones, which makes the handling of plutonium dangerous.

<span class="mw-page-title-main">Neptunium(IV) oxide</span> Chemical compound

Neptunium(IV) oxide, or neptunium dioxide, is a radioactive, olive green cubic crystalline solid with the formula NpO2. It emits both α- and γ-particles.

<span class="mw-page-title-main">Corium (nuclear reactor)</span> Material in core during nuclear meltdown

Corium, also called fuel-containing material (FCM) or lava-like fuel-containing material (LFCM), is a material that is created in a nuclear reactor core during a nuclear meltdown accident. Resembling lava in consistency, it consists of a mixture of nuclear fuel, fission products, control rods, structural materials from the affected parts of the reactor, products of their chemical reaction with air, water, steam, and in the event that the reactor vessel is breached, molten concrete from the floor of the reactor room.

<span class="mw-page-title-main">Nuclear transmutation</span> Conversion of an atom from one element to another

Nuclear transmutation is the conversion of one chemical element or an isotope into another chemical element. Nuclear transmutation occurs in any process where the number of protons or neutrons in the nucleus of an atom is changed.

References

  1. Christine Guéneau; Alain Chartier; Paul Fossati; Laurent Van Brutzel; Philippe Martin (2020). "Thermodynamic and Thermophysical Properties of the Actinide Oxides". Comprehensive Nuclear Materials 2nd Ed. 7: 111–154. doi:10.1016/B978-0-12-803581-8.11786-2. ISBN   9780081028667.
  2. "Nitric acid processing". Los Alamos Laboratory.
  3. Greenwood, Norman N.; Earnshaw, Alan (1984). Chemistry of the Elements. Oxford: Pergamon Press. p. 1471. ISBN   978-0-08-022057-4.
  4. De Bruycker, F.; Boboridis, K.; Pöml, P.; Eloirdi, R.; Konings, R. J. M.; Manara, D. (2011). "The melting behaviour of plutonium dioxide: A laser-heating study". Journal of Nuclear Materials. 416 (1–2): 166–172. Bibcode:2011JNuM..416..166D. doi:10.1016/j.jnucmat.2010.11.030.
  5. Jeffrey A. Katalenich Michael R. Hartman Robert C. O’Brien Steven D. Howe (Feb 2013). "Fabrication of Cerium Oxide and Uranium Oxide Microspheres for Space Nuclear Power Applications" (PDF). Proceedings of Nuclear and Emerging Technologies for Space 2013: 2. Archived from the original (PDF) on 2016-10-07. Retrieved 2016-07-27.
  6. A. Angelo Jr. and D. Buden (1985). Space Nuclear Power. Krieger Publishing Company. ISBN   0-89464-000-3.
  7. "General Safety Considerations" (PDF). Fusion Technology Institute, University of Wisconsin–Madison. Spring 2000. Archived from the original (PDF lecture notes) on 2018-09-15. Retrieved 2017-10-20.
  8. "Transit". Encyclopedia Astronautica. Archived from the original on June 24, 2002. Retrieved 2013-05-07.
  9. Michael Singer; David Weir & Barbara Newman Canfield (Nov 26, 1979). "Nuclear Nightmare: America's Worst Fear Come True". New York Magazine.
  10. Sublette, Carey. "4.1 Elements of Fission Weapon Design". The Nuclear Weapon Archive. 4.1.7.1.2.1 Plutonium Oxide. Retrieved 20 October 2017. The critical mass of reactor grade plutonium is about 13.9 kg (unreflected), or 6.1 kg (10 cm nat. U) at a density of 19.4. A powder compact with a density of 8 would thus have a critical mass that is (19.4/8)^2 time higher: 82 kg (unreflected) and 36 kg (reflected), not counting the weight of the oxygen (which adds another 14%). If compressed to crystal density these values drop to 40 kg and 17.5 kg.
  11. United States Nuclear Regulatory Commission, Fact sheet on plutonium (accessed Nov 29 2013)
  12. Gwaltney-Brant, Sharon M. (2013-01-01), Haschek, Wanda M.; Rousseaux, Colin G.; Wallig, Matthew A. (eds.), "Chapter 41 - Heavy Metals", Haschek and Rousseaux's Handbook of Toxicologic Pathology (Third Edition), Boston: Academic Press, pp. 1315–1347, ISBN   978-0-12-415759-0 , retrieved 2022-04-10
  13. World Nuclear Society, Plutonium Archived 2015-08-18 at the Wayback Machine (accessed Nov 29 2013)
  14. "Toxicological Profile For Plutonium" (PDF). U.S. Department of Health and Human Services. 2007-09-27. Retrieved 2009-04-23.