Fact-checked by Grok 2 weeks ago

Plutonium

Plutonium is a radioactive chemical element with atomic number 94 and chemical symbol Pu. Classified as an actinide in the periodic table, it is a transuranic element occurring naturally only in trace quantities but primarily produced artificially through neutron irradiation of uranium-238 in nuclear reactors. First synthesized in December 1940 at the University of California, Berkeley, by Glenn T. Seaborg and his team via deuteron bombardment of uranium, plutonium was isolated in February 1941. The element exhibits complex allotropic behavior with six phases at ambient pressure, a silvery-white metallic appearance that rapidly tarnishes in air due to oxidation, and high reactivity including pyrophoricity in finely divided forms. Plutonium-239, the fissile isotope bred in reactors, powers implosion-type nuclear weapons and serves as fuel in mixed-oxide reactor fuel, while plutonium-238 generates heat for space missions via radioisotope decay. Its extreme toxicity, alpha-emitting radioactivity, and proliferation risks underscore stringent handling protocols and international safeguards on production.

Properties

Physical properties


Plutonium is a silvery-white metal that tarnishes rapidly upon exposure to air, forming a dull gray, yellow, or olive-green oxide layer. In its pure form, it exhibits a bright metallic luster initially.
The element exists as a solid at standard temperature and pressure, with a density of 19.816 g/cm³ in its room-temperature alpha phase, though this varies across its six allotropic forms due to differences in crystal packing. Its melting point is 640 °C and boiling point is 3230 °C, reflecting relatively low melting but high boiling characteristics for a heavy actinide. Plutonium has an empirical atomic radius of approximately 151 pm and a van der Waals radius of 200 pm. Unlike typical metals, plutonium displays poor thermal and electrical conductivity, with room-temperature electrical resistivity around 1.5 × 10⁻⁶ Ω·m—high for a metal—and thermal conductivity that remains low even at elevated temperatures up to 822 K. These properties arise from its complex electronic structure and phase instability, contributing to anisotropic expansion and contraction during thermal cycling.
PropertyValue
(α-phase, 25°C)19.816 /cm³
640 °C
3230 °C
Electrical resistivity (RT)1.5 × 10⁻⁶ Ω·m

The - exhibits changes at points, as plutonium undergoes transformations between its allotropes, with the being the densest. This necessitates careful handling to avoid stresses from varying by , such as a linear of about 55 × 10⁻⁶ K⁻¹ in certain modifications.

Allotropes

Plutonium exhibits six distinct allotropic phases under ambient pressure, each characterized by unique crystal structures, densities, and temperature stability ranges, resulting from the complex bonding involving its 5f electrons. These phases lead to substantial volume contractions and expansions during transitions, with density variations spanning approximately 15.92 to 19.86 g/cm³, which poses significant challenges for machining and structural integrity in applications. The α phase, stable below approximately 122 °C, adopts a monoclinic crystal structure with a high density of 19.86 g/cm³, rendering it hard and brittle, akin to ceramic materials. Upon heating, it transforms to the β phase around 125–211 °C, which has a body-centered monoclinic structure and lower density of 17.70 g/cm³. The γ phase follows between roughly 211–310 °C, featuring a face-centered orthorhombic structure at 17.14 g/cm³. Higher-temperature phases include the δ phase, stable from about 310–452 °C, with a face-centered cubic structure and the lowest density of 15.92 g/cm³; this phase is notably ductile when stabilized at room temperature via alloying with elements like gallium (typically 0.5–1 atomic percent), enabling easier fabrication for nuclear components. The δ' phase, body-centered tetragonal with 16.00 g/cm³ density, exists narrowly from 452–475 °C, transitioning to the ε phase (body-centered cubic, 16.51 g/cm³) up to the melting point at 640 °C.
PhaseCrystal StructureDensity (g/cm³)Approximate Stability Range (°C)
αMonoclinic19.86< 122
βBody-centered monoclinic17.70125–211
γFace-centered orthorhombic17.14211–310
δFace-centered cubic15.92310–452
δ'Body-centered tetragonal16.00452–475
εBody-centered cubic16.51475–640
These transformations are first-order, involving latent heats and hysteresis, with volume changes up to 25% that can induce internal stresses and cracking if not managed through alloying or controlled thermal cycling. The δ phase's stabilization is critical for weapons-grade plutonium, as its low symmetry and density contrast with α's high symmetry but brittleness, influencing pit fabrication processes at facilities like Los Alamos.

Chemical properties

Plutonium is a chemically reactive metal that exhibits oxidation states from +3 to +6 under typical conditions, with +4 being the most stable and common in its compounds. Oxidation states +2 and +7 have been achieved through specialized synthetic methods, such as organometallic complexes for +2 and strong oxidants for +7. In aqueous solutions, these states display distinct colors: Pu(III) ranges from green to violet, Pu(IV) from pink to brown, Pu(V) light purple, Pu(VI) light brown, and Pu(VII) dark green. The metal tarnishes rapidly in air, forming a passive oxide layer of PuO₂ that protects the bulk from further oxidation under dry conditions, though finely divided plutonium is pyrophoric and ignites spontaneously. Exposure to moist air or water vapor accelerates oxidation, with water reacting directly to form plutonium hydroxide and hydrogen gas; the reaction rate for massive metal is slow compared to alkali metals but increases with surface area. Plutonium does not react with alkalies but dissolves readily in concentrated hydrochloric, hydroiodic, and perchloric acids; it resists dilute nitric acid due to the stable oxide coating. Plutonium forms a variety of compounds, including oxides like PuO₂ (used in nuclear fuel), hydrides such as PuH₂ (which expand the metal volume), and halides like PuF₄ and PuCl₃. These compounds reflect its actinide chemistry, where f-orbital involvement leads to variable coordination geometries and complex redox behavior, enabling disproportionation reactions in solution, such as 3Pu(IV) → Pu(III) + 2Pu(V).
Oxidation StateTypical IonColor in Aqueous Solution (approx.)Notes
+3Pu³⁺Green-violetReducing conditions
+4Pu⁴⁺Pink-brownMost stable, hydrolyzes easily
+5PuO₂⁺Light purpleUnstable, disproportionates
+6PuO₂²⁺Yellow-brownStable in acidic media

Electronic structure and compounds

Plutonium, atomic number 94, has a ground-state electron configuration of [Rn] 5f⁶ 7s², with the 5f orbitals partially filled and contributing to its chemical behavior as a member of the actinide series. The 5f electrons in plutonium exhibit intermediate localization, distinguishing it from lighter actinides where 5f orbitals are more itinerant and heavier ones where they are more localized, leading to unique bonding properties that blend d- and f-electron characteristics. In compounds, plutonium displays a wide range of oxidation states from +2 to +7, with +3, +4, +5, and +6 being the most prevalent and stable under typical conditions; the +2 state was first synthesized in a cyclopentadienyl complex in 2017, while +7 occurs transiently in strong oxidants. These states arise from the accessibility of 5f and 6d electrons, enabling facile redox changes, and plutonium ions in solution—such as Pu³⁺ (green to violet), Pu⁴⁺ (brown), PuO₂⁺ (+5, pink to blue), and PuO₂²⁺ (+6, yellow)—often coexist, complicating speciation due to hydrolysis and complexation. Key binary compounds include plutonium(IV) oxide (PuO₂), a refractory fluorite-structured solid stable to high temperatures and widely used in nuclear fuels for its low solubility and resistance to radiation damage. Plutonium tetrafluoride (PuF₄) forms green crystals and serves as an intermediate in purification, while plutonium trihalides like PuCl₃ and PuBr₃ exhibit Pu(III) states with layered structures. Plutonium reacts with hydrogen to form hydrides such as PuH₂ and PuH₃, which are pyrophoric and decompose above 400°C, and with nitrogen to yield plutonium nitride (PuN), a potential fast-reactor fuel with high thermal conductivity. Organoplutonium compounds, though less common due to reactivity, include alkyl derivatives like (C₅H₅)₃Pu demonstrating covalent f-orbital involvement.

Nuclear Characteristics

Isotopes and abundance

Plutonium has no stable isotopes, with 15 known radioactive isotopes ranging from mass number 228 to 247. The primary decay modes are alpha decay for most isotopes, supplemented by spontaneous fission in even-mass heavier isotopes and beta minus decay for odd-neutron variants like plutonium-241. Plutonium-244 is the longest-lived isotope, with a half-life of approximately 82 million years, decaying via alpha emission to uranium-240. The table below lists selected plutonium isotopes, their half-lives, and primary decay modes:
IsotopeHalf-lifePrimary decay mode(s)
^{238}Pu87.7 yearsα
^{239}Pu24,000 yearsα
^{240}Pu6,560 yearsα, spontaneous fission
^{241}Pu14 yearsβ⁻
^{242}Pu375,000 yearsα
^{244}Pu82 million yearsα
In nature, plutonium isotopes occur only in trace quantities, primarily from cosmogenic neutron capture on uranium or rare primordial remnants, with concentrations typically below femtograms per kilogram in uranium ores or soils. Plutonium-244 represents the most abundant natural isotope, though its detection requires sensitive mass spectrometry due to low levels. Artificially produced plutonium, generated via neutron capture on uranium-238 in nuclear reactors, exhibits variable isotopic abundance depending on irradiation duration and flux spectrum. Weapons-grade plutonium contains over 93% ^{239}Pu and less than 7% ^{240}Pu to minimize spontaneous fission neutrons, whereas reactor-grade material exceeds 19% ^{240}Pu plus higher even isotopes. In spent fuel from pressurized water reactors, ^{239}Pu comprises roughly 58% of total plutonium, with the remainder distributed among ^{240}Pu, ^{241}Pu, and ^{242}Pu.

Nucleosynthesis and natural occurrence

Plutonium isotopes are primarily synthesized via the rapid neutron-capture process (r-process) in astrophysical events such as core-collapse supernovae and neutron star mergers, where extreme neutron fluxes enable the formation of heavy, neutron-rich nuclei beyond the iron peak. This process involves successive neutron captures on seed nuclei followed by beta decays, producing transuranic elements like plutonium, which cannot form through slower stellar nucleosynthesis pathways such as the s-process due to insufficient neutron densities. The r-process yields favor even-even isotopes, with plutonium-244 (half-life 81.3 million years) being among the more stable, though all plutonium isotopes decay relatively rapidly on geological timescales. On Earth, plutonium does not occur in primordial quantities, as any synthesized during solar system formation would have decayed given half-lives ranging from 14.35 years (Pu-238) to 80.8 million years (Pu-244). Instead, trace natural occurrence arises from secondary processes: neutron capture on uranium-238 in ore deposits, where U-238 absorbs neutrons (from spontaneous fission of U-238 or cosmic rays) to form U-239, which beta-decays to neptunium-239 and then plutonium-239. Concentrations remain exceedingly low, typically on the order of parts per trillion in uranium-rich minerals, far below levels from anthropogenic sources. Additional traces stem from ancient natural nuclear reactors, such as those at Oklo, Gabon, operating about 1.9 billion years ago, where fission chains produced minor plutonium alongside other actinides under self-sustaining conditions. More recently, deep-sea sediment cores reveal influxes of extraterrestrial plutonium-244 from nearby supernovae within the last 10 million years, with detections of approximately 181 atoms per sample indicating r-process ejecta deposition. These cosmic contributions, while confirming ongoing stellar nucleosynthesis, contribute negligibly to Earth's inventory compared to reactor-produced plutonium.

Fission properties

Plutonium-239 is fissile, undergoing induced fission upon capture of low-energy thermal neutrons to release approximately 2.89 neutrons on average per fission event, enabling a self-sustaining chain reaction in suitable configurations. Its thermal neutron fission cross-section measures 747 barns, substantially higher than the 582 barns for uranium-235, contributing to superior neutron economy in reactors. The isotope's prompt neutron multiplicity supports effective multiplication factors (k) exceeding 1 in critical assemblies, with k_infinity values around 2.9 in infinite media under thermal conditions due to low parasitic capture. The bare-sphere critical mass for weapons-grade plutonium-239 (delta phase, density 19.84 g/cm³) is approximately 10 kilograms, corresponding to a radius of about 6.4 cm, far lower than the 52 kilograms for uranium-235 owing to plutonium's higher density and cross-sections. This compactness facilitated implosion-type designs in nuclear weapons, as gun-type assemblies prove impractical for plutonium due to predetonation risks from spontaneous fission. Reflectors and tampers reduce the required mass to 4-6 kg in practical devices by minimizing neutron leakage. Spontaneous fission rates distinguish plutonium isotopes: Pu-239 exhibits negligible rates (half-life ~10^16 years), but Pu-240's rate yields ~415,000 neutrons per second per kilogram, necessitating high-purity plutonium (<7% Pu-240) for weapons to avoid fizzle yields from premature criticality. Pu-241, also fissile with a thermal cross-section of 1010 barns, contributes to reactivity but decays rapidly (half-life 14.3 years), while even isotopes like Pu-242 have low fissionability with thermal neutrons. In fast-spectrum environments, plutonium isotopes show enhanced fission probabilities, with Pu-239's fast fission cross-section enabling breeding ratios >1 in breeder reactors.

Decay heat and radiological behavior

Plutonium isotopes predominantly undergo alpha decay, emitting alpha particles with energies typically exceeding 5 MeV, accompanied by low-energy gamma rays and x-rays below 20 keV, which convert to heat through interaction with surrounding material. This decay mode dominates for isotopes like plutonium-239 (half-life 24,110 years) and plutonium-238 (half-life 87.7 years), with the alpha particles depositing nearly all their kinetic energy locally as thermal energy, typically yielding about 5-6 MeV per decay after accounting for gamma contributions. Plutonium-241, however, decays primarily by beta emission to americium-241 (half-life 14.35 years for the beta process), introducing additional radiological complexity through subsequent alpha decay of the daughter product. Decay heat generation varies significantly by isotope due to differences in half-life and decay energy. Plutonium-238 exhibits the highest specific power, approximately 0.57 watts per gram, arising from its relatively short half-life and efficient alpha decay, making it suitable for radioisotope thermoelectric generators where steady thermal output powers spacecraft. In contrast, plutonium-239 produces far lower heat, on the order of milliwatts per gram, reflecting its longer half-life, while admixtures like plutonium-240 (half-life 6,561 years) contribute modestly to overall heat in reactor-grade or weapons-grade material through both alpha decay and spontaneous fission. This intrinsic heat necessitates cooling in stored plutonium stockpiles to prevent thermal runaway or structural degradation, with total decay heat in mixed-isotope forms scaling roughly with isotopic composition— for instance, weapon-grade plutonium (typically <7% Pu-240) generates about 1-2 watts per kilogram initially. Radiologically, plutonium poses minimal external hazard due to alpha particles' short range (stopped by skin or paper), but its behavior shifts dramatically upon internalization via inhalation or ingestion, where alpha emissions cause dense ionization tracks leading to cellular damage, fibrosis, and elevated cancer risk in organs like lungs or bones. Even-mass isotopes such as plutonium-238, -240, and -242 undergo spontaneous fission at low rates (e.g., 415 neutrons per kilogram per second for pure Pu-240), emitting prompt neutrons that enable criticality assessments but complicate handling by inducing neutron activation in nearby materials. Plutonium-239, while primarily alpha-decaying, contributes fewer neutrons (~2,600 fissions per kilogram per second in pure form), yet its fissionability amplifies radiological output under neutron irradiation. Gamma emissions remain low across isotopes, often <1% of total energy, but daughter products like americium-241 from Pu-241 decay introduce penetrating 60 keV gammas over time. In environmental or dispersal scenarios, plutonium's radiological behavior is governed by its particulate form and solubility; insoluble oxides dominate releases (e.g., from fuel fabrication), lodging in respiratory tissues with prolonged retention (biological half-life >100 days in lungs), whereas soluble forms like Pu(IV) distribute systemically via blood to liver and skeleton. Spontaneous neutron emission from isotopes like Pu-240 limits safe storage densities, as accumulated neutrons can drive subcritical multiplication or material activation, requiring neutron-absorbing diluents in packaging. Overall, these properties—high internal alpha dose potential coupled with manageable external fields—dictate stringent handling protocols, emphasizing containment over shielding for routine operations.

Production

Laboratory synthesis

Plutonium was first synthesized on , 1940, at the of the University of California, Berkeley, through deuteron of using the 60-inch . A team led by Glenn T. Seaborg, including Edwin M. McMillan, Joseph W. Kennedy, and Arthur C. Wahl, induced the nuclear reaction ^{238}U + ^2H → ^{239}U + ^1H, where the resulting uranium-239 underwent successive beta decays: ^{239}U (half-life 23.5 minutes) to ^{239}Np (half-life 2.3 days), and then to ^{239}Pu (half-life 24,110 years). The initial identification relied on tracer-level quantities, with chemical separation of the new element from uranium and fission products via precipitation and solvent extraction techniques, confirming its distinct properties akin to other actinides rather than rare earths. On February 23–24, 1941, the team performed the first chemical identification, verifying plutonium as element 94 through its characteristic oxidation states and precipitation behaviors. This cyclotron method produced microgram quantities insufficient for macroscopic study, limiting yields due to the low cross-section of the (d,p) reaction and beam intensity constraints of the era's accelerator technology. To obtain weighable amounts, Seaborg's group at the Met Lab in Chicago developed carrier-assisted isolation in 1942, using lanthanum fluoride to coprecipitate plutonium from neutron-irradiated uranium samples, yielding the first visible sample on August 20, 1942, and the first weighing (2.77 micrograms of plutonium oxide) the following month. These laboratory techniques, reliant on radiochemical purification and ion-exchange separation, enabled initial characterization of plutonium's metallic properties and reactivity, paving the way for scale-up in production reactors. Laboratory synthesis remained confined to research settings post-discovery, supplanted by reactor-based neutron capture on uranium-238 for bulk production due to vastly higher efficiency.

Industrial-scale production

Industrial-scale production of plutonium relies on the neutron capture by uranium-238 in nuclear reactors to form plutonium-239, followed by chemical separation from irradiated fuel. This process was first achieved at the Hanford Site in Washington state, where the B Reactor began operations on September 26, 1944, producing the first industrial quantities of plutonium for the Manhattan Project. Hanford's nine production reactors and four reprocessing facilities generated nearly two-thirds of all plutonium used by the United States for national security purposes, totaling approximately 67 metric tons over four decades. The core method involves loading aluminum-clad uranium metal "slugs" into reactor channels, irradiating them for periods ranging from days to months to optimize Pu-239 yield while minimizing unwanted isotopes like Pu-240, then discharging and dissolving the fuel in nitric acid. Plutonium is then separated via the PUREX (Plutonium Uranium Redox Extraction) process, a solvent extraction technique using tributyl phosphate in kerosene to isolate plutonium and uranium from fission products and other actinides in aqueous nitric acid solutions. At Hanford, initial separations used bismuth phosphate coprecipitation before transitioning to PUREX in the 1950s for higher efficiency and purity, enabling production rates that peaked at over 10 metric tons per year across multiple sites by the late 1950s. The Savannah River Site in South Carolina complemented Hanford with five heavy-water moderated reactors (R, P, L, K, and C) operational from 1953 onward, producing about 30 metric tons of plutonium, primarily weapons-grade, alongside tritium. Combined U.S. production from 1944 to 2009 reached 111.7 metric tons, with the majority dedicated to nuclear weapons pits—each containing 3 to 6 kilograms of plutonium—fabricated at facilities like Rocky Flats, which processed material at rates exceeding 1,000 pits annually during peak Cold War demand. These operations emphasized weapons-grade plutonium (less than 7% Pu-240) by limiting irradiation times to reduce neutron captures on Pu-239. Post-Cold War, industrial-scale production ceased in the U.S., with no dedicated reactors operating after 1988 at Hanford's N Reactor and 1988 at Savannah River's K Reactor, shifting focus to stockpile stewardship and limited Pu-238 production for space applications at sites like Oak Ridge. Reprocessing of civilian spent fuel via PUREX variants occurs abroad, such as in France's La Hague facility, yielding reactor-grade plutonium (19-23% Pu-240) for mixed-oxide fuel, but U.S. policy prohibits commercial reprocessing to avoid proliferation risks.

Current global production and stockpiles

As of early 2024, the global stockpile of separated plutonium totals approximately 565 metric tons, with about 140 metric tons classified as weapons-usable and the remainder primarily civilian or military-obligated material unsuitable for direct weapons use due to isotopic composition or international safeguards. This inventory reflects cumulative production from nuclear reactors and reprocessing facilities since the mid-20th century, with limited transparency in military programs complicating precise accounting. Major holders include Russia (193 tons total, including 88 tons weapons-usable), the United States (87.6 tons total, 38.4 tons weapons-usable), the United Kingdom (120 tons total, 3.2 tons weapons-usable), and France (102 tons total, 6 tons weapons-usable). China holds about 3 tons, nearly all weapons-usable, while Japan possesses around 45 tons of civilian plutonium stored domestically and abroad.
CountryTotal Stockpile (metric tons)Weapons-Usable (metric tons)
Russia19388
United States87.638.4
United Kingdom1203.2
France1026
China32.9
Japan~45 (civilian)0
Data as of early 2024; military estimates involve uncertainty due to non-disclosure. Current production remains modest compared to Cold War peaks, focused on civilian reprocessing for fuel cycles and limited military restarts. Civilian plutonium separation occurs primarily through reprocessing of spent nuclear fuel at facilities like France's La Hague (capacity ~1,700 tons heavy metal/year, yielding ~10-12 tons Pu annually, much of which is recycled into MOX fuel), Russia's RT-1 plant (400-500 tons/year capacity), and Japan's Rokkasho plant (800 tons/year capacity, operational from 2024). Global civilian inventories grew by about 4 tons net in 2023, reflecting production exceeding MOX fuel fabrication and other uses. Weapons-grade production is constrained; the United States plans 10 plutonium pits (each ~5 kg Pu-239) in 2024 and 20 in 2025 at Los Alamos and Savannah River for warhead modernization, totaling under 0.1 tons annually. Ongoing but undeclared military production continues in India, Pakistan, Israel, and North Korea via dedicated reactors. Efforts to reduce excess stockpiles have stalled. The U.S. holds ~19.7 tons of surplus Cold War-era weapons-grade plutonium designated for downblending or reactor use, with recipients to be announced by December 2025 for advanced reactor testing. Russia withdrew from the 2000 U.S.-Russia Plutonium Management and Disposition Agreement in October 2025, halting joint disposal of 34 tons each of excess weapons plutonium amid geopolitical tensions. These developments underscore persistent proliferation risks from existing inventories, as no comprehensive fissile material cutoff treaty exists.

History

Discovery and pre-war research

Plutonium, element 94, was first synthesized on December 14, 1940, at the University of California, Berkeley, through the bombardment of uranium-238 with deuterons in the 60-inch cyclotron. The reaction produced neptunium-238 as an intermediate, which decayed via beta emission to yield plutonium-238, the first isotope identified. The team, led by Glenn T. Seaborg and including Edwin M. McMillan, Joseph W. Kennedy, and Arthur C. Wahl, isolated trace amounts of the new element chemically using lanthanum fluoride as a carrier and confirmed its presence through radiochemical analysis. Chemical identification as a distinct element occurred on the night of February 23-24, 1941, when Seaborg's group demonstrated that the substance exhibited properties akin to uranium but behaved as a new transuranic element, separable from known actinides via precipitation and solvent extraction techniques. Early experiments revealed plutonium-238's alpha decay with a half-life of approximately 88 years, measured through scintillation counting and ionization chamber data. These findings built on McMillan's prior discovery of neptunium in June 1940, extending the actinide hypothesis Seaborg later formalized, positing that elements beyond actinium formed a series analogous to the lanthanides. The element was named plutonium on March 21, 1942, following the planetary nomenclature with neptunium (after Neptune) and uranium, though the decision stemmed from 1941 deliberations amid growing wartime secrecy. Pre-war research remained limited to microgram-scale samples due to cyclotron constraints and the shift toward classified efforts; initial nuclear cross-section measurements indicated potential neutron capture pathways in uranium leading to plutonium-239, but fission cross-sections for this isotope were not fully explored until 1942. By mid-1941, U.S. authorities classified the work, curtailing open publication and redirecting focus to chain reaction feasibility under military oversight.

Manhattan Project and wartime developments

The Manhattan Project's plutonium production program accelerated in 1942 after metallurgical laboratory tests confirmed plutonium-239's fissile properties exceeded expectations for bomb-grade material, prompting a shift from sole reliance on uranium-235 enrichment. General Leslie Groves selected Hanford, Washington, as the production site due to its isolation, abundant Columbia River water for cooling, and proximity to hydroelectric power. DuPont accepted a fixed-fee contract in October 1942 to design and construct graphite-moderated, water-cooled reactors based on the Chicago Pile-1 prototype, with construction beginning in mid-1943. The B Reactor, Hanford's first production unit, achieved criticality on September 26, 1944, marking the initiation of large-scale plutonium-239 generation via neutron irradiation of uranium-238 fuel slugs. Chemical reprocessing in purpose-built canyon facilities separated plutonium from fission products and unused uranium, yielding the first weapons-grade metal shipment to Los Alamos on February 2, 1945. By mid-1945, Hanford supplied sufficient plutonium for testing and deployment, despite initial challenges like reactor poisoning by xenon-135, resolved through empirical adjustments increasing power output. Plutonium's viability for gun-type assembly, akin to the uranium Little Boy design, proved infeasible due to spontaneous fission from plutonium-240 impurities—inevitable in reactor-bred material—risking predetonation. Los Alamos physicists, led by J. Robert Oppenheimer, pivoted to implosion: symmetrically compressing a subcritical plutonium pit with precisely shaped explosive lenses to achieve supercritical density. This required extensive hydrodynamic simulations, explosive testing at remote sites, and innovations in casting plutonium-gallium alloys for the delta-phase structure, overcoming asymmetries that could fizzle the reaction. The Trinity test on July 16, 1945, at the Alamogordo Bombing Range, detonated a plutonium implosion device ("Gadget") containing approximately 6 kilograms of plutonium-239, yielding 21 kilotons and validating the method despite yield uncertainties from imperfect symmetry. Hanford-derived plutonium powered the Fat Man bomb, an implosion-type weapon airburst over Nagasaki on August 9, 1945, at 1,650 feet altitude, demonstrating plutonium's battlefield efficacy. Wartime developments thus established plutonium as the primary fissile core for subsequent U.S. arsenal designs, prioritizing implosion for its compression advantages over uranium's simpler assembly.

Cold War expansion and applications

Following the success of the Manhattan Project, the United States significantly expanded plutonium production to meet escalating demands for nuclear deterrence amid rising tensions with the Soviet Union. Hanford Site operations grew from the initial B Reactor to include nine production reactors by the early 1960s, with facilities such as D, F, DR, H, C, KE, and KW reactors coming online between 1949 and 1962, alongside the dual-purpose N Reactor activated in 1963 for both plutonium generation and electrical power. These expansions, coupled with four reprocessing plants, yielded nearly two-thirds of all U.S. government plutonium, totaling over 67 metric tons by the end of production. Concurrently, the Savannah River Site in South Carolina initiated operations with its R Reactor in 1953, followed by P and L reactors, contributing additional plutonium until cessation in 1988. The Rocky Flats Plant near Denver, Colorado, began fabricating plutonium pits—the fissile cores for implosion-type nuclear weapons—in 1952, achieving an annual output of 1,000 to 2,000 pits during peak Cold War years to support arsenal growth. This infrastructure enabled the U.S. nuclear stockpile to peak at 31,255 warheads in the late 1960s, with plutonium-239 comprising the primary fissile material in the majority of these devices due to its superior neutron economy compared to uranium-235 for implosion designs. The Soviet Union mirrored this expansion after its first plutonium-based nuclear test, RDS-1 (Joe-1 in the West), on August 29, 1949, at Semipalatinsk, which utilized designs informed by espionage-derived implosion technology. Plutonium production commenced at the Mayak Production Association, with additional reactors built at closed cities like Tomsk-7 (Seversk) and Krasnoyarsk-26 (Zheleznogorsk), sustaining output for an estimated peak stockpile exceeding 40,000 warheads by the 1980s. Soviet facilities emphasized rapid scaling, producing sufficient plutonium for thousands of fission and thermonuclear weapons, though exact yields remain classified; by the Cold War's end, cumulative production approached 100-160 metric tons. Applications during this era centered overwhelmingly on nuclear weapons, where weapons-grade plutonium-239 (containing less than 7% Pu-240) enabled compact, high-yield implosion assemblies essential for deliverable warheads on missiles, bombers, and submarines. Post-1952, plutonium served as the sparkplug in thermonuclear designs, igniting fusion stages in devices like the U.S. Ivy Mike test (yield 10.4 megatons) and subsequent deployable hydrogen bombs, vastly amplifying destructive potential over pure fission weapons. Limited non-weapons uses emerged, including experimental breeder reactor fuel cycles and radioisotope thermoelectric generators (RTGs) powered by plutonium-238 for satellites and remote beacons, with U.S. production of the latter isotope ramping up from 1959 via Savannah River irradiations of neptunium-237. However, military imperatives dominated, with plutonium's role in deterrence credited by strategic analysts for preventing direct superpower conflict through mutually assured destruction. Civilian reactor applications remained negligible until post-Cold War demilitarization efforts, as proliferation risks and high costs deterred widespread adoption.

Post-Cold War management and recent advancements

Following the dissolution of the Soviet Union in 1991, the United States and Russia shifted priorities from expanding plutonium production for nuclear arsenals to managing and disposing of excess weapons-grade material, driven by arms reduction treaties and non-proliferation goals. The U.S. declared approximately 52.7 metric tons of plutonium as surplus to military needs by the mid-1990s, committing to its irreversible disposition to prevent reuse in weapons. This included initial plans for conversion into mixed-oxide (MOX) fuel for commercial reactors or immobilization in glass logs for geological repository storage, though MOX fabrication faced delays due to high costs exceeding $20 billion for the planned facility at Savannah River Site. In September 2000, the U.S. and Russia signed the Plutonium Management and Disposition Agreement (PMDA), obligating each to dispose of at least 34 metric tons of weapons-grade plutonium through irradiation in reactors or equivalent means to achieve irreversibility. Russia suspended implementation in 2016, citing U.S. non-compliance with commitments like removing Syrian chemical weapons and providing Ukraine aid, and formally withdrew from the accord in October 2025 via State Duma approval, reclaiming its 34 tons without mutual disposition progress. The U.S. responded by pivoting to a "dilute and dispose" method for much of its surplus, blending plutonium oxide with inert materials like borosilicate glass to render it unsuitable for weapons, with processing ongoing at sites including Savannah River, where about 7.1 metric tons had been treated by 2018 before partial halts. Parallel to disposition, the U.S. established the Stockpile Stewardship Program (SSP) in 1995 following the 1992 nuclear testing moratorium, relying on advanced simulations, subcritical experiments, and facilities like the National Ignition Facility to certify plutonium pit integrity without full-yield tests. The SSP has enabled refurbishment of existing warheads, such as the W87-1 for the Sentinel ICBM, while addressing plutonium aging effects like phase transitions and helium accumulation from alpha decay. To sustain deterrence, the National Nuclear Security Administration (NNSA) resumed plutonium pit production in 2019 at Los Alamos and plans to reach 80 pits per year by the late 2020s at a new facility in South Carolina, reversing post-Cold War dismantlement that reduced U.S. capacity to near zero. Recent advancements include U.S. efforts to repurpose surplus plutonium for civilian energy, with the Department of Energy set to allocate 19.7 metric tons to reactor fuel fabricators by December 2025, potentially as MOX or advanced fuels to support nuclear expansion amid data center demands. Globally, separated plutonium stocks stood at about 565 metric tons as of early 2024, with civilian holdings (e.g., from reprocessing) surpassing weapons material, prompting multi-recycling strategies in reactors like France's PWRs using high-burnup MOX assemblies. In Russia, testing of MOX fuel for VVER reactors expanded in 2025 to enable closed fuel cycles, reducing waste and leveraging existing stockpiles. These developments reflect a tension between disposal imperatives and renewed interest in plutonium as a fissionable resource, constrained by proliferation risks and infrastructure costs.

Applications

Nuclear weapons and deterrence

Plutonium-239, with a half-life of 24,110 years, is fissile and capable of sustaining an explosive nuclear chain reaction when a neutron induces fission, releasing approximately 2.9 neutrons per fission event on average for thermal neutrons. This property enables its use as the core fissile material in most nuclear weapons, where a subcritical mass is rapidly compressed to supercriticality using symmetric implosion of high explosives. The implosion design became necessary for plutonium because its production in reactors inevitably includes about 1% Pu-240, which undergoes spontaneous fission at a rate 100,000 times higher than Pu-239, causing predetonation in simpler gun-type assemblies and yielding only a low-efficiency fizzle explosion. The plutonium implosion weapon was developed during the Manhattan Project, with the first test, Trinity, on July 16, 1945, at Alamogordo, New Mexico, using 6.2 kilograms of Pu-239 and producing a yield of 21 kilotons of TNT equivalent. This design was deployed operationally in the "Fat Man" bomb, dropped on Nagasaki, Japan, on August 9, 1945, by the B-29 Bockscar, detonating at 1,650 feet altitude and yielding 21 kilotons, which contributed to Japan's surrender in World War II. Postwar, plutonium cores, or "pits," enabled more compact and efficient fission primaries in thermonuclear weapons, boosting yields into the megaton range while fitting into deliverable missiles and bombers. In nuclear deterrence, plutonium pits form the essential fissile component of warheads in the U.S. stockpile, which numbered about 3,708 active and inactive weapons as of 2023, underpinning a strategy of assured retaliation to prevent aggression by demonstrating the credible threat of unacceptable damage. The U.S. maintains plutonium production capabilities, aiming for 80 pits annually by the late 2020s at facilities like Los Alamos and Savannah River, to replace aging pits affected by radiation-induced changes in microstructure and replace those retired under arms control agreements, ensuring long-term stockpile reliability without nuclear testing. This sustainment supports extended deterrence for allies, as plutonium-based weapons provide the high reliability and safety margins needed for global deployment amid peer competitors' nuclear advancements. Critics argue proliferation risks from plutonium handling, but proponents emphasize its irreplaceable role in verifiable, high-confidence deterrence absent alternatives like pure-fusion weapons.

Fissile material in reactors and fuel cycles

Plutonium-239, the primary fissile isotope of plutonium, undergoes fission when absorbing thermal neutrons, releasing approximately 200 MeV of energy per fission event and sustaining chain reactions in nuclear reactors similar to uranium-235. In typical light-water reactors fueled with enriched uranium, Pu-239 is generated in situ through neutron capture by uranium-238 in the fuel rods, with subsequent beta decays: U-238 (n,γ) → U-239 → Np-239 → Pu-239. This process yields about 250-300 kg of plutonium annually in a 1000 MWe reactor operating continuously, contributing roughly one-third of the total fission energy output as the initial U-235 depletes. In nuclear fuel cycles, plutonium enables both once-through and closed-loop strategies to extend uranium resources. Spent fuel from light-water reactors contains approximately 1% plutonium by mass, including fissile Pu-239 (about 50-60% of total Pu) alongside Pu-240, Pu-241, and Pu-242, which can be reprocessed via aqueous methods like PUREX to separate plutonium for reuse. Mixed oxide (MOX) fuel, comprising 3-10% plutonium oxide blended with depleted or reprocessed uranium oxide, allows reactors to burn recovered plutonium, reducing high-level waste volume and utilizing existing infrastructure without breeding net new fissile material in thermal spectra. Over 40 reactors worldwide, primarily in Europe and Japan, have used MOX assemblies since the 1970s, with France recycling about 10% of its annual plutonium inventory into MOX at facilities like La Hague. Fast breeder reactors leverage plutonium's favorable fast-neutron fission cross-section to achieve breeding ratios exceeding 1.0, producing more Pu-239 than consumed by surrounding uranium-238 blankets with fast neutrons that minimize parasitic absorption. In such designs, a plutonium-uranium core initiates fission, with excess neutrons converting U-238 to Pu-239 in the blanket, potentially multiplying fissile resources by a factor of 60-100 over natural uranium's once-through efficiency, as demonstrated in prototypes like Russia's BN-350 (operational 1973-1999) and BN-600 (since 1980). Closed plutonium fuel cycles in breeders support long-term sustainability but require reprocessing to recycle bred plutonium, contrasting with open cycles that discard it as waste. Empirical data from operational breeders confirm net fissile gain, though challenges include higher capital costs and material corrosion from liquid metal coolants like sodium. Plutonium-241, another fissile isotope produced via Pu-240 neutron capture, contributes marginally to reactivity but decays to americium-241, complicating long-term cycle management. Overall, plutonium's integration in fuel cycles hinges on neutron economy: thermal reactors rely on it for burnup extension, while fast systems exploit it for multiplication, grounded in measured cross-sections and isotopic yields from reactor irradiations.

Heat and power sources in remote environments

Plutonium-238 serves as the primary radioisotope in radioisotope thermoelectric generators (RTGs) for providing heat and electricity in remote environments, particularly deep space missions where solar power is insufficient. Its alpha decay generates approximately 0.56 watts of thermal power per gram, enabling continuous operation without mechanical parts or reliance on sunlight. RTGs convert this decay heat into electricity via the Seebeck effect in thermocouples, typically achieving efficiencies of 5-10%. The Multi-Mission RTG (MMRTG), designed for NASA missions, incorporates about 4.8 kg of plutonium-238 dioxide, producing roughly 2 kW thermal and 110 watts electrical at mission start, with output declining due to the isotope's 87.7-year half-life. This design powered the Curiosity rover, launched November 26, 2011, which has operated on Mars since August 6, 2012, and the Perseverance rover, landed February 18, 2021, enabling extended surface exploration in varying temperatures from -140°C to 20°C. Earlier general-purpose heat source RTGs fueled Voyager 1 and 2, launched September 5, 1977, which continue transmitting data from interstellar space as of 2025. Cassini, launched October 15, 1997, used three RTGs for its Saturn orbit until its 2017 end-of-mission dive. Production of plutonium-238 involves irradiating neptunium-237 in reactors, processed at facilities like Oak Ridge National Laboratory and Los Alamos National Laboratory under Department of Energy oversight, with resumed large-scale output since 2015 to meet NASA demands after a post-Cold War hiatus. By 2023, shipments included half-kilogram lots of oxide pellets to support upcoming missions like Dragonfly to Titan, planned for 2028 launch. These systems prioritize safety through iridium-clad fuel pellets encased in multi-layered graphite impact shells, minimizing release risks during launch failures, as demonstrated in historical tests with no significant radionuclide dispersal.

Emerging uses in advanced energy and research

Generation IV nuclear reactors, designed for enhanced sustainability through closed fuel cycles, incorporate plutonium recycling to improve uranium resource utilization and minimize waste. Four of the six Generation IV designs—such as sodium-cooled fast reactors and gas-cooled fast reactors—rely on fast neutron spectra to breed fissile plutonium-239 from uranium-238 while consuming existing plutonium stockpiles, enabling breeding ratios exceeding 1.0 in optimized configurations. These systems support multi-recycling of plutonium in mixed oxide (MOX) or nitride fuels, reducing the need for fresh uranium mining by up to 30 times compared to once-through cycles in light-water reactors. In October 2025, the U.S. Department of Energy announced access to up to 19 metric tons of surplus weapons-grade plutonium from retired warheads for conversion into fuel for advanced reactors, targeting developers of small modular and fast reactors to meet rising energy demands, including data centers. Recipient companies are expected to be selected by year-end 2025, with fuel fabrication requiring several years due to isotopic adjustments for reactor compatibility. This initiative repurposes approximately 10% of the U.S. surplus plutonium inventory, previously designated for disposal, into civilian energy applications while maintaining non-proliferation safeguards. Ongoing research focuses on advanced plutonium fuel fabrication, including models for MOX multi-recycling (MOX-MR) in fast reactors, which simulate isotopic evolution to sustain core performance over multiple cycles without significant degradation. In Russia, trials of REMIX uranium-plutonium fuel in pressurized water reactors advanced to the third cycle at Balakovo Nuclear Power Plant in 2024, demonstrating compatibility with existing infrastructure for partial closed cycles that recycle 95% of actinides. These efforts address proliferation-resistant transmutation of minor actinides alongside plutonium, potentially reducing long-term radiotoxicity of spent fuel by factors of 100 or more through repeated fast-spectrum irradiation.

Risks and Safety Measures

Chemical and radiological toxicity

Plutonium's toxicity arises from both its chemical properties as an actinide heavy metal and its radiological emissions, primarily alpha particles from isotopes such as ^{239}Pu (half-life 24,110 years) and ^{238}Pu (half-life 87.7 years). While external exposure poses negligible risk due to alpha particles' limited penetration (stopped by skin or paper), internal deposition via inhalation, ingestion, or wounds delivers high localized doses, causing cellular damage through ionization and free radical formation. Radiological effects predominate over chemical ones, as alpha radiation's high linear energy transfer (LET) induces genotoxicity, chromosomal aberrations, and carcinogenesis in retained tissues, amplified by plutonium's long biological half-lives (e.g., >50 years in bone). Chemical toxicity, akin to that of lead or mercury, involves ionic disruption and protein binding but is secondary and confounded by radiation in empirical studies. Inhalation represents the primary exposure route, with insoluble forms like PuO_2 retaining in lungs (half-time 14–80 years in humans), leading to radiation pneumonitis, fibrosis, and lung cancer at burdens ≥0.19–0.63 kBq/kg in canine models (equivalent to LOAEL 0.28 kBq/kg for pneumonitis). Human data from Mayak workers exposed to aerosols show excess relative risks (ERR) of 3.9–19 per Gy for lung cancer, with fibrosis at >10 Sv. Soluble compounds (e.g., Pu nitrate, citrate) enhance translocation to blood, concentrating in liver (~45%), skeleton (~45%), and kidneys, where Pu(IV) forms bind transferrin, exacerbating alpha-induced tumors (liver ERR 2.6–29 per Gy; bone ERR 0.76–3.4 per Gy). Chemical effects manifest as hepatic enzyme elevation and nephrotoxicity in animals at similar doses (e.g., LOAEL 0.19 kBq/kg for liver lesions in dogs), though distinguishing from radiation requires non-radioactive analogs, which indicate heavy metal-like renal tubular damage without quantifiable human LD_{50}. Ingestion yields low systemic uptake (<0.1% in adults, up to 10-fold higher in neonates or iron-deficient states), with most excreted fecally, minimizing risks beyond transient gastrointestinal irritation. In neonatal rats, high oral doses (e.g., 5,300 kBq/kg) cause crypt hypertrophy and epithelial damage, blending chemical corrosion and irradiation. Wound contamination delivers soluble plutonium directly, prompting rapid chelation (e.g., DTPA) to reduce retention, as seen in Manhattan Project cases with depositions of 98–3,300 Bq. Overall, occupational limits reflect radiological dominance: chronic inhalation minimal risk level (MRL) at 0.00003 mg/m³, annual limit on intake (ALI) 0.8 µCi for ^{239}Pu. Latency for cancers exceeds 20 years, with Mayak cohorts exhibiting relative risks up to 17 for liver and 7.9 for bone at >7.4 kBq burdens.

Criticality hazards and prevention

Plutonium-239 and plutonium-241, due to their high fission cross-sections, pose criticality hazards when sufficient quantities accumulate in configurations allowing neutron multiplication to exceed losses, potentially leading to an uncontrolled chain reaction and burst of radiation. The bare-sphere critical mass for weapons-grade plutonium-239 metal is approximately 10 kg, though water reflection or moderation can reduce this to 4.9-5.6 kg. Plutonium solutions present elevated risks, with minimum critical masses as low as 0.5 kg in water-moderated geometries at concentrations near 7 g/L, as hydrogen enhances neutron thermalization and slows leakage. Historical accidents underscore these hazards. At Los Alamos in August 1945, a tungsten carbide brick dropped onto a 6.2 kg plutonium core during manual assembly, prompting a supercritical excursion with 10¹⁶ fissions and one fatality from 510 rem exposure. A similar error in May 1946 with a beryllium reflector on the same core yielded 3×10¹⁵ fissions, another fatality from 2,100 rem, and exposures up to 360 rem for seven others. In December 1958, stirring a 3.1 kg plutonium-rich organic layer at Los Alamos caused prompt criticality, delivering ~12,000 rem to operator Cecil Kelley, who died 35 hours later. Solution-handling errors at Hanford (April 1962) and Mayak (March 1953, September 1962, December 1968) produced multiple excursions, injuries from hundreds of rem, and one additional fatality. Prevention employs the double contingency principle, ensuring criticality requires two independent, unlikely failures such as equipment malfunction and procedural violation. Engineering controls prioritize geometry: slab thicknesses limited to 0.65 cm for metal or 5.8 cm for solutions, cylinder diameters under 4.4 cm for metal or 15.7 cm for unreflected solutions, and annular storage to maximize surface-to-volume ratios for neutron escape. Neutron absorbers, including soluble gadolinium nitrate or borosilicate Raschig rings, are integrated to maintain subcriticality (k_eff < 0.95) under credible accidents. Administrative safeguards include fissile mass limits below critical values, solution concentrations capped at <7 g Pu/L without poisons, minimum separations between units, and rigorous training with stop-work authority for operators suspecting unsafe conditions. Criticality alarm systems, featuring neutron-sensitive detectors, trigger evacuations upon detecting fission spikes, limiting potential doses to <2.5 rem for personnel. For metal, dry inert atmospheres prevent hydride formation that could introduce moderation, while avoiding spherical shapes mitigates reflection risks. These measures, validated through calculational benchmarks and historical reviews, have prevented U.S. production-scale plutonium criticality accidents since 1962.

Flammability and handling protocols

Plutonium metal exhibits significant pyrophoricity, particularly in finely divided forms such as powders or turnings, where it can ignite spontaneously upon exposure to air without external ignition sources. Pyrophoric plutonium is defined as material that self-ignites in air at temperatures of 150°C or below in the absence of heat, shock, or friction. Finely divided particles, including 140-mesh powder, ignite at approximately 135°C, while lathe turnings require around 265°C. Massive samples generally necessitate temperatures exceeding 475°C for ignition, often involving oxide reduction to Pu₂O₃ followed by exothermic reoxidation that sustains combustion. In moist air, plutonium reacts to form expanding oxides and hydrides, which flake into pyrophoric powders capable of forming flammable or explosive dust-air mixtures. Handling protocols emphasize confinement and inert environments to mitigate flammability risks. Operations occur within sealed gloveboxes maintained under inert atmospheres like argon or nitrogen to prevent air exposure and spontaneous oxidation. Fine particulates are minimized during machining or processing, with collection systems designed to capture and store residues in non-reactive containers. Personal protective equipment, including respirators, gloves, and clothing, provides barriers against alpha particle emission and potential contamination, though it offers limited shielding for beta particles depending on material thickness. Ignition sources are strictly excluded from plutonium areas, and air-handling systems operate within designated safety envelopes to avoid dust dispersion. For storage and transportation, plutonium is encased in robust, sealed containers under or to inhibit reactivity, with protocols prohibiting exposure to or oxidants that accelerate and formation. Emergency procedures include non--based suppression agents, as water can exacerbate reactions with plutonium or promote criticality in solutions. Regular monitoring for surface oxidation and adherence to operational limits ensure that handling remains within parameters that preclude or fire initiation.

Transportation and storage safeguards

Transportation of plutonium adheres to stringent international and national regulations to mitigate risks of radiation exposure, criticality, and diversion. In the United States, the Nuclear Regulatory Commission (NRC) governs packaging under 10 CFR Part 71, requiring certified containers that withstand hypothetical accident conditions including 30-foot drops, 30-minute fires at 800°C, and water immersion. The Department of Energy's Office of Secure Transportation (OST) manages shipments of special nuclear materials like plutonium using armored tractor-trailers escorted by armed federal agents, with no fatal accidents recorded since operations began in 1975. Internationally, the International Atomic Energy Agency (IAEA) sets standards in SSR-6, emphasizing Type B(U) or Type C packages for plutonium oxide or metal forms, incorporating double containment, shielding, and neutron absorbers to prevent criticality during air, sea, or road transit. Criticality prevention in transport relies on geometric controls, such as limiting plutonium mass per package (often under 5 kg for air shipments) and using boron-lined casks to absorb neutrons, ensuring subcritical configurations even under impact or flooding. Air transport of plutonium faces additional restrictions under 10 CFR 71.88 and 10 CFR Part 871, prohibiting non-exempt shipments unless in specially certified packages or with national security exemptions. Post-9/11 enhancements by the NRC mandate enhanced physical security for transport, including real-time tracking and armed escorts to counter theft risks. Storage safeguards prioritize long-term stability, security, and non-proliferation. Plutonium is typically stabilized per DOE-STD-3013 standards, involving thermal treatment at 950°C in an oxygen-rich atmosphere to convert oxides to a durable form resistant to hydrogen generation or corrosion, packaged in sealed, welded stainless-steel containers under inert argon atmospheres to avert pyrophoricity and radiolysis. Facilities like those at Pantex or Savannah River employ vault storage with continuous monitoring for radiation, temperature, and pressure, alongside material control and accountability (MC&A) programs to track inventories and detect anomalies. IAEA safeguards verify stored plutonium through inspections, seals, and surveillance, ensuring no diversion for weapons use, as applied to civilian stocks exceeding 140 tons globally under agency oversight. Criticality safety in storage mandates spacing, reflectors avoidance, and double contingency principles, where two unlikely failures are required for a critical event.

Controversies and Policy Debates

Proliferation risks versus strategic necessities

Plutonium-239 possesses a bare-sphere critical mass of about 10 kilograms, far lower than the 52 kilograms for uranium-235, enabling more compact and efficient fission weapon designs via implosion compression rather than simpler gun-type assembly suitable only for uranium. This technical superiority allows plutonium-based weapons to achieve higher yields with less material, making it indispensable for advanced thermonuclear devices that integrate plutonium pits as primary stages to trigger fusion secondaries. The proliferation risks stem primarily from plutonium's production in civilian reactors and its separation via reprocessing, which generates stockpiles vulnerable to diversion by states or non-state actors for improvised nuclear devices. Global separated civilian plutonium exceeds 250 metric tons as of recent estimates, with reprocessing programs in nations like Japan and France expanding these quantities despite International Atomic Energy Agency (IAEA) safeguards aimed at material accountancy and containment surveillance. Such safeguards have detected anomalies, as in Iran's undeclared activities, but face challenges in bulk-handling facilities where measurement uncertainties allow potential hidden diversions of kilograms-scale quantities sufficient for weapons. Strategic necessities counterbalance these risks through nuclear deterrence, where plutonium-enabled arsenals underpin mutually assured destruction (MAD) doctrines that empirical records attribute to averting great-power conflicts since 1945, including during the Cold War's multiple crises. Nine states possess nuclear weapons, predominantly plutonium-based, yet none have initiated direct nuclear use against peers, correlating with extended periods of relative peace absent pre-nuclear era world wars. For major powers like the United States, maintaining plutonium production and pit fabrication—restarted at Los Alamos in 2024 after decades—ensures credible second-strike capabilities against expanding threats from Russia and China, whose own plutonium-dependent forces numbered over 2,500 and 500 warheads respectively in 2023 assessments. Policy debates intensify over civilian reprocessing, with proponents of closed fuel cycles arguing it recycles uranium resources while opponents, citing heightened theft risks from separated plutonium, advocate once-through cycles or international fuel banks to centralize handling under stricter controls. Non-proliferation Treaty (NPT) adherents, including the U.S., permit limited military plutonium retention for deterrence but restrict transfers, though enforcement gaps—evident in North Korea's plutonium pathway to weapons—underscore that technological barriers alone fail without robust intelligence and sanctions. Ultimately, while safeguards mitigate but do not eliminate diversion probabilities, the causal link between deployable plutonium arsenals and strategic stability justifies their retention for responsible nuclear states, prioritizing empirical deterrence outcomes over absolute risk aversion.

Environmental impact claims versus empirical data

Environmental advocacy groups and certain media outlets frequently assert that plutonium contamination renders affected areas uninhabitable for millennia, citing its alpha radiation, long half-life of 24,110 years for Pu-239, and purported high bioavailability leading to widespread bioaccumulation in food chains and irreversible ecosystem damage. These claims often extrapolate from worst-case laboratory solubilities or isolated hotspots, overlooking plutonium's geochemical properties that limit actual environmental mobility. Empirically, plutonium predominantly exists as insoluble Pu(IV) oxides like PuO2, which exhibit extremely low solubility in neutral soils and waters (solubility product log Ksp ≈ -57 for Pu(OH)4), strong adsorption to clay minerals and iron oxides, and minimal vertical migration beyond the topsoil layer. Soil-to-plant transfer factors for plutonium typically range from 10^{-5} to 10^{-4} m²/kg, orders of magnitude lower than for more mobile radionuclides like cesium-137, resulting in negligible uptake by vegetation and limited trophic transfer. Global atmospheric nuclear testing from 1945 to 1980 released approximately 5-6 kg of Pu-239 equivalent into the environment, dispersing it as refractory particles that settled primarily in the northern hemisphere, yielding average soil concentrations of 0.1-10 mBq/g—far below levels causing observable ecological disruption. Long-term monitoring by agencies like the United Nations Scientific Committee on the Effects of Atomic Radiation (UNSCEAR) attributes no specific population-level health effects, such as increased cancers, solely to plutonium fallout; overall testing-related risks stem more from volatile fission products like iodine-131 and strontium-90, with plutonium's alpha emissions confined to short-range deposition and low inhalation bioavailability outside hotspots. At sites like the Chernobyl exclusion zone, where ~70 kg of plutonium isotopes were released in 1986 primarily as fuel particles, plutonium hotspots remain localized with dissolution rates under 0.1% per year in soils, contributing minimally to ongoing contamination compared to cesium; biodiversity surveys indicate thriving populations of large mammals and birds, contradicting claims of sterility, as radiation doses to wildlife (often <1 mGy/day) fall below thresholds for population-level effects. U.S. production sites provide further empirical contrast: At Hanford, where ~50 metric tons of plutonium were processed from 1944 to 1987 with documented tank leaks totaling ~1.5 million curies of radioactivity, plutonium has exhibited retarded groundwater migration, with vadose zone retention exceeding 99% due to sorption coefficients (Kd) of 10^3-10^5 mL/g; 2023 monitoring detected plutonium-239/240 in groundwater at <1 pCi/L, well below the 0.15 pCi/L drinking water standard, and no attributable off-site ecological harm, as evidenced by stable fish populations in the Columbia River. Similarly, at Rocky Flats, Colorado, post-1969 fire releases of ~1 kg plutonium led to surface soil hotspots, but cleanup to <50 pCi/g Pu by 2005, combined with epidemiological studies from 1990-1999, found no excess cancers among downwind residents attributable to plutonium exposure, with modeled public doses <1 mrem/year versus natural background of ~300 mrem/year. These data underscore that while initial claims amplified perceived risks—often from sources with institutional incentives to highlight hazards—controlled releases and plutonium's inherent immobility have confined impacts, enabling site repurposing as wildlife refuges without measurable long-term biodiversity loss.

Waste management and recycling efficacy

Plutonium-containing high-level radioactive waste, arising from spent nuclear fuel reprocessing and legacy sites like Hanford and Savannah River, is primarily managed through vitrification, a process that mixes the waste with molten glass to form stable borosilicate logs for immobilization and eventual deep geological disposal. This method enhances chemical durability and reduces leach rates, with vitrification facilities at Hanford beginning operations for low-activity waste on October 15, 2025, processing tank sludges into glass forms suitable for near-surface engineered disposal, while high-level fractions require repository emplacement. Vitrification achieves waste loadings of 20-30% for plutonium-bearing streams, minimizing container volumes, though legacy wastes at Hanford—totaling over 56 million gallons in 177 underground tanks—present ongoing challenges due to corrosion and radionuclide migration risks prior to treatment. Recycling plutonium via aqueous reprocessing, predominantly the PUREX method, separates plutonium and uranium from fission products in spent fuel, enabling reuse in mixed oxide (MOX) fuel assemblies and thereby partitioning waste streams to reduce high-level waste volumes. PUREX recovers over 99% of plutonium, with decontamination factors exceeding 10^4 for key fission products, allowing the extracted plutonium—typically 0.9-1% of spent fuel mass—to be fabricated into MOX pellets for thermal reactors, where it fissions to generate energy and transmute to shorter-lived isotopes. Commercial facilities, such as France's La Hague plant operated by Orano, demonstrate recovery of up to 96% reusable uranium and plutonium per tonne of spent fuel, converting the remaining 4% fission products into a more concentrated high-level waste form. The efficacy of plutonium recycling in waste management is evidenced by volume reductions: reprocessing followed by MOX recycle decreases high-level waste requiring geological disposal by a factor of 5-10 relative to direct spent fuel disposal, as usable actinides are removed and fissioned, lowering long-term heat loads and radiotoxicity by orders of magnitude over 10,000 years. In closed fuel cycles with fast reactors, multiple recycling passes can further diminish waste volumes by up to 90%, effectively burning plutonium and minor actinides while concentrating fission products into manageable forms. However, efficacy is tempered by secondary waste streams from reprocessing—such as nitric acid effluents and hulls—classified as intermediate-level and requiring separate conditioning, though their total volume remains lower than untreated spent fuel assemblies. Empirical data from operational cycles confirm that recycling extends fuel resource utilization without proportionally increasing disposal burdens, provided proliferation safeguards are maintained.

Public health experiments and ethical reviews

During the Manhattan Project in 1945, U.S. researchers conducted non-therapeutic experiments injecting plutonium isotopes into 18 hospital patients to study the element's absorption, distribution, retention in tissues (particularly bone), and urinary and fecal excretion rates, data intended to establish safe exposure limits for atomic workers handling plutonium. These trials, coordinated by the project's Medical Division under Stafford Warren, involved doses of plutonium-239 (typically 0.3 to 5 micrograms) or plutonium-238, administered intravenously at facilities including the University of Rochester's Strong Memorial Hospital, the University of Chicago, and Oak Ridge Hospital. Subjects were selected from vulnerable populations, such as terminally ill cancer patients or those with chronic conditions like hypertension and anemia, under the rationale that their short life expectancies minimized long-term risks and facilitated autopsy access for tissue analysis. The inaugural injection occurred on April 10, 1945, when 57-year-old construction foreman Ebb Cade (code-named HP-1), recovering from severe injuries in a truck accident at Oak Ridge, received 4.7 micrograms of plutonium-239 without his knowledge or consent; subsequent procedures extracted his teeth and surgically removed bone fragments containing plutonium deposits for analysis, though he survived until October 1947, dying from heart disease unrelated to the injection per autopsy findings. Of the 18 subjects, four received plutonium-238 (a higher alpha-emitter for tracer studies), and post-mortem examinations confirmed plutonium's preferential accumulation in bone marrow and liver, with excretion primarily via urine in the initial weeks but long-term skeletal retention exceeding 0.01% of the dose persisting indefinitely. Only one subject, a University of Rochester patient, signed a consent form, which vaguely described the procedure as involving "tracer substances" without disclosing plutonium's identity, radioactivity, or potential carcinogenic risks, reflecting the era's prioritization of national security over individual autonomy. These experiments drew no contemporaneous public scrutiny due to wartime secrecy but were later deemed ethically deficient by the 1994-1995 Advisory Committee on Human Radiation Experiments (ACHRE), established by President Clinton to probe Cold War-era radiation studies; the committee's final report highlighted the absence of voluntary informed consent, deception in subject selection (e.g., portraying injections as routine diagnostics), and failure to weigh harms against non-therapeutic benefits, violating even the nascent Nuremberg Code principles emerging post-WWII. ACHRE noted that while plutonium doses were below acute lethality thresholds (e.g., far under the 5-microgram "maximum permissible body burden" derived from radium analogies), the experiments exploited power imbalances with ill patients, some as young as children in related trials, and prioritized military data over ethical safeguards, contributing to broader distrust in government science. In response, the Department of Energy issued formal apologies to survivors and families in 1995, facilitated limited compensation via a compassionate payment program, and declassified records, underscoring systemic ethical lapses in early nuclear research despite claims of scientific necessity. Subsequent reviews, including congressional hearings, affirmed that modern standards—requiring institutional review boards, full disclosure, and risk minimization—would preclude such unconsented exposures, though defenders cited the experiments' role in averting industrial accidents by informing plutonium's 45-year skeletal half-life.

References

  1. [1]
    Plutonium | Pu (Element) - PubChem - NIH
    Plutonium is a chemical element with symbol Pu and atomic number 94. Classified as an actinide, Plutonium is a solid at 25°C (room temperature).
  2. [2]
    Backgrounder on Plutonium - Nuclear Regulatory Commission
    Plutonium is a radioactive metallic element with the atomic number 94. It was discovered in 1940 by scientists studying how to split atoms to make atomic bombs.
  3. [3]
    [PDF] Some Properties of the Important Plutonium Isotopes ... - OSTI.GOV
    Plutonium is by far the most important of the transuranic elements (elements heavier than uranium). Its use in atomic (fission) weapons has made it a ...<|control11|><|separator|>
  4. [4]
    A History of Plutonium | Los Alamos National Laboratory
    Sep 21, 2022 · Discover plutonium's fascinating history and unique chemical properties, from atomic bombs to nuclear power, space exploration, ...
  5. [5]
    Glenn Seaborg's Greatest Hits
    ... plutonium (atomic number 94, symbol Pu) at U.C. Berkeley on the night of February 23-24, 1941. We found this plutonium in the form of isotope 238 Pu. This ...
  6. [6]
    The Search for "Heavy" Elements
    Aug 9, 2000 · Plutonium (Pu, Z = 94) was discovered in 1941 by bombarding a uranium target with deuterons (a hydrogen nucleus with one proton and one neutron) ...<|separator|>
  7. [7]
    [PDF] Plutonium-An Element at Odds with Itself
    Plutonium is sensitive not only to temperature and pressure but also to chemi- cal additions. In fact, much of its density and structural instability can be ...
  8. [8]
    Atomic number 94 | Los Alamos National Laboratory
    Dec 13, 2021 · The element plutonium was discovered only 81 years ago, but its impact on the world has been monumental.
  9. [9]
    [PDF] Plutonium (Pu) - Washington State Department of Health
    Plutonium-239 will undergo a fission chain reaction if enough of it is concentrated in one place, so it is used at the heart of modern day nuclear weapons and ...
  10. [10]
    Plutonium Element Facts - Chemicool
    Classification: Plutonium is an actinide metal ; Color: silvery-white ; Atomic weight: (244), no stable isotopes ; State: solid ; Melting point: 639.4 oC, 912.5 K.
  11. [11]
    Plutonium | AMERICAN ELEMENTS®
    In its elemental form, CAS 7440-07-5, plutonium has a silvery white appearance. The plutonium atom has a radius of 151.3.pm and its Van der Waals radius is 200.
  12. [12]
    Technical data for the element Plutonium in the Periodic Table
    Overview. Name, Plutonium. Symbol, Pu. Atomic Number, 94. Atomic Weight, 244. Density, 19.816 g/cm3. Melting Point, 640 °C. Boiling Point, 3230 °C.
  13. [13]
    Physical, Nuclear, and Chemical Properties of Plutonium
    Melting point: 641 deg. C. Boiling point: 3232 deg. C. Density: 16 to 20 grams/cubic centimeter. Nuclear Properties of Plutonium. Plutonium belongs to the class ...Missing: radius | Show results with:radius
  14. [14]
    Plutonium - Element information, properties and uses | Periodic Table
    Melting point, 640°C, 1184°F, 913 K. Period, 7, Boiling point, 3228°C, 5842°F, 3501 ... Atomic radius, non-bonded (Å), 2.43, Covalent radius (Å), 1.80. Electron ...
  15. [15]
    Thermal conductivity of plutonium above room temperature
    Mar 18, 2008 · The thermal diffusivity of plutonium metal has been measured from room temperature to 822 K by the laser-flash method, and the thermal ...
  16. [16]
    Plutonium - Periodic Table of Elements - EnvironmentalChemistry.com
    Physical Properties of Plutonium · Atomic Mass Average: 244.0642 · Boiling Point: 3503K 3230°C 5846°F · Coefficient of lineal thermal expansion/K-1: 55E ...
  17. [17]
    [PDF] Plutonium and Its Alloys-From Atoms to Microstructure
    Plutonium is of practical interest principally because the 239 isotope has attractive nuclear properties for energy production and nuclear explosives.
  18. [18]
    Plutonium | Pu | CID 23940 - PubChem - NIH
    Plutonium is a silvery white metal that exists as a solid under normal conditions. It is produced when uranium absorbs an atomic particle.
  19. [19]
    Plutonium - Los Alamos National Laboratory
    Plutonium metal normally has six allotropes or crystal structures; alpha (α), beta (β), gamma (γ), delta (δ), delta prime (δ') and epsilon (ε). It forms a ...
  20. [20]
    In Their Own Words: How Plutonium Alloys Age | LANL
    Dec 1, 2016 · (Delta-phase plutonium, naturally occurring only well above room temperature, is useful because it is ductile, rather than brittle like the room ...
  21. [21]
    [PDF] The Chemistry of Plutonium
    Jan 5, 1983 · Plutonium ions in solution commonly exist in the (III), (IV), (V), and (VI) oxidation states, as Pu³+, Pu4+, PuO2, and PuO2+, respectively.
  22. [22]
    Identification of the Formal +2 Oxidation State of Plutonium
    Feb 24, 2017 · Six Pu oxidation states have been unambiguously confirmed (0 and +3 to +7), and four different oxidation states can exist simultaneously in ...Supporting Information · Author Information · Acknowledgment · References
  23. [23]
    WebElements Periodic Table » Plutonium » properties of free atoms
    The ground state electron configuration of ground state gaseous neutral plutonium is [Rn].5f6.7s2 and the term symbol is 7F0. Kossel shell structure ...
  24. [24]
    Nature of the 5 ⁢ f electronic structure of plutonium
    Mar 24, 2020 · Under ambient temperature and pressure, the α phase is favorable. It crystallizes in a monoclinic structure with 16 Pu atoms within the unit ...<|separator|>
  25. [25]
    Plutonium gets another oxidation state added to its arsenal | Research
    Mar 17, 2017 · A team of US researchers has coaxed a +2 oxidation state out of the radioactive element for the first time.
  26. [26]
    Plutonium Oxidation States in the Waste Isolation Pilot Plant ... - NIH
    Plutonium can exist in solution in the +III, +IV, +V and +VI oxidation states, and the plutonium oxidation state significantly affects its aqueous solubility ...
  27. [27]
    WebElements Periodic Table » Plutonium » compounds information
    Plutonium forms binary compounds with halogens, oxygen, hydrogen, and other elements. The most common oxidation number is 4. Examples include PuH2, PuO2, and ...
  28. [28]
    Plutonium | Radiation Emergencies - CDC
    Apr 17, 2024 · Plutonium is a silvery-gray metal that becomes yellowish when exposed to air. Most plutonium in the environment is in the form of microscopic ...
  29. [29]
    Plutonium - World Nuclear Association
    Aug 16, 2023 · It melts at 640°C into a very corrosive liquid. The alpha phase is hard and brittle, like cast iron, and if finely divided it spontaneously ...
  30. [30]
    Plutonium 239 - an overview | ScienceDirect Topics
    Although plutonium 241 has a half-life of only 14 years, it is important because it decays to americium 241, an alpha emitter with a half-life of 430 years.
  31. [31]
    Plutonium Isotopes - Radioactivity.eu.com
    In the case of conventional pressurized water reactor burning uranium fuel, plutonium 239 corresponds to 58% of plutonium found in the spent fuel.<|separator|>
  32. [32]
    60Fe and 244Pu deposited on Earth constrain the r-process yields ...
    May 14, 2021 · We detect two distinct influxes of 60 Fe to Earth in the last 10 million years and accompanying lower quantities of 244 Pu.
  33. [33]
    Alien radioactive element find prompts creation rethink
    May 13, 2021 · Iron-60 is well-known to form in supernovae, but plutonium-244 formation is not so well understood. If it is formed in supernovae, then one ...
  34. [34]
    Trace seabed plutonium points to stellar forges of heavy elements
    May 13, 2021 · The scientists also detected 181 atoms of plutonium-244 scattered across the 10-million-year span of the sample. They calculated the plutonium- ...<|control11|><|separator|>
  35. [35]
    r-Process Radioisotopes from Near-Earth Supernovae and Kilonovae
    May 11, 2021 · Plutonium-244 is a heavy and long-lived radioactive actinide nucleus, which is only made in a very neutron rich environment via robust r-process ...
  36. [36]
    Do transuranic elements such as plutonium ever occur naturally?
    Mar 23, 1998 · The answer to the question is yes, plutonium does occur naturally. Plutonium appears at very low concentrations in nature.
  37. [37]
    Radionuclide Basics: Plutonium | US EPA
    Feb 6, 2025 · Plutonium is considered a man-made element, although scientists have found trace amounts of naturally occurring plutonium produced under highly ...
  38. [38]
    Facts About Plutonium | Live Science
    Dec 7, 2016 · Plutonium generally isn't found in nature. Trace elements of plutonium are found in naturally occurring uranium ores. Here, it is formed in ...
  39. [39]
    [PDF] Plutonium Fact Sheet - HPS Chapters
    (Extremely small quantities of plutonium were created naturally in sustained underground nuclear reactions estimated to have occurred about 1.9 billion years ...
  40. [40]
    Shrapnel From Relatively Recent Supernovae Found in the Earth's ...
    May 21, 2021 · The specific isotope Dr. Wallner and his team found is actually the most stable of the plutonium isotopes - P-244. However, it doesn't exist ...
  41. [41]
  42. [42]
    A5 Thermal neutron cross sections - IAEA Nuclear Data Services
    A-5. Actinide data: Thermal neutron cross sections, resonance integrals, and Westcott factors. Nuclide, Type, Cross Section σx [barn], Westcott Factor
  43. [43]
    [PDF] 1.3-Fast Reactor Physics - 1 Spectrum and Cross Sections
    Mar 26, 2019 · Spectral Variation of Neutron Cross Sections: Pu-239. ▫ Fission and capture cross section >100X higher in thermal range. ▫ Sharp decrease in ...
  44. [44]
    HEALTH EFFECTS - Toxicological Profile for Plutonium - NCBI - NIH
    Most plutonium isotopes emit a high energy (generally >5 MeV) alpha particle and low energy (<20 keV) gamma and x-rays as they transform into uranium. The ...
  45. [45]
    Properties of Plutonium Isotopes
    Plutonium isotopes are fissionable, with Pu-239 and Pu-241 being fissile. They have varying half-lives, decay modes, and radiological hazards. Pu-241 decays ...
  46. [46]
    [PDF] Chapter 5
    All plutonium produces a significant amount of heat due to its radio- active decay. The plutonium that was used in the 1945 Trinity test consisted mostly of Pu ...
  47. [47]
    Plutonium - Nuclear Museum - Atomic Heritage Foundation
    Jun 5, 2014 · Plutonium was first produced and isolated on December 14, 1940 at the University of California, Berkeley.
  48. [48]
    75th Anniversary: Discovery of Plutonium - ChemistryViews
    Dec 14, 2015 · The team of Glenn T. Seaborg (pictured) used the lab's cyclotron to bombard uranium with deuterons, which generated neptunium, element 93.
  49. [49]
    Manhattan Project: Piles and Plutonium, 1939-1942 - OSTI
    Experiments conducted in early 1941 at the Radiation Laboratory at the University of California, Berkeley, finally completed the link between pile research and ...
  50. [50]
    Seaborg and Plutonium Chemistry, Met Lab, 1942-1944 - OSTI.gov
    Using lanthanum fluoride as a carrier, Seaborg isolated a weighable sample of plutonium in August 1942. At the same time, Isadore Perlman and William J. Knox ...
  51. [51]
    Plutonium Production - Manhattan Project - OSTI.gov
    The production of plutonium bomb cores depended on the mining, milling, and refining of uranium ore as much as the uranium bomb designs did.
  52. [52]
    [PDF] The Plutonium Production Story at the Hanford Site
    Jun 20, 1996 · This document covers the history of plutonium production at Hanford, including fuel fabrication, irradiation, spent fuel handling, plutonium ...
  53. [53]
    [PDF] A Short History of Hanford Waste Generation, Storage, and Release
    Nine nuclear reactors and four reprocessing plants at Hanford produced nearly two-thirds of the plutonium used in the United States for government purposes2.
  54. [54]
    [PDF] The Purex Process, DPSPU 77-11-1.
    The Purex process involves dissolving fuel, then using solvent extraction cycles to separate uranium and plutonium, and then processing uranium to uranium ...
  55. [55]
    [PDF] Savannah River Site www.srs.gov
    The facility historically performed plutonium and neptunium oxide production, including Pu-238 for NASA. The facility is in lay-up state and also awaits further ...
  56. [56]
    [PDF] The United States Plutonium Balance, 1944 - 2009
    Plutonium Mass Balance​​ From 1944 to 2009, DOE and its predecessor organizations produced and acquired a total of 111.7 MT of plutonium (Table 2). During the ...
  57. [57]
    Plutonium Pit Production | Department of Energy
    During the Cold War, the United States could produce more than 1,000 pits per year at the Rocky Flats Plant in Colorado.
  58. [58]
    Plutonium: The First 50 Years - FAS Project on Government Secrecy
    The US plutonium inventory is composed of 85.0 MT of weapon grade, 13.2 MT of fuel grade, and 1.3 MT of reactor grade.
  59. [59]
    Plutonium-238 Production for Space Exploration - National Historic ...
    The US met this challenge by developing nuclear batteries known as radioisotope thermoelectric generators (RTGs) and producing plutonium-238 (Pu-238) as their ...
  60. [60]
    [PDF] Development of Advanced Reprocessing Technologies
    Key steps in the PUREX process. In a PUREX reprocessing facility the spent fuel is thus separated into its four components: uranium, plutonium, high level waste ...
  61. [61]
    Plutonium - International Panel on Fissile Materials
    Apr 28, 2025 · The map shows national stocks of separated plutonium as of the beginning of 2024. The data shows ownership, rather than physical location of the material.
  62. [62]
    International Panel on Fissile Materials
    The global stockpile of separated plutonium was about 565 tons. Of this material, 425 metric tons were produced outside of weapon programs, covered by ...About IPFM · IPFM Blog · Plutonium · Publications
  63. [63]
    [PDF] Global Inventory of Separated Plutonium (Detailed Descriptions By ...
    China is currently constructing two reprocessing plants each with annual capacities of up to 200 tons per year. UK. 3.2. 116.4. The production bases of military ...
  64. [64]
  65. [65]
    Civilian plutonium declarations for 2023 - IPFM Blog
    Mar 7, 2025 · The total amount of plutonium declared as civilian was about 374 tonnes, an increase of about 4 tonnes since the end of 2022.Missing: global stockpiles
  66. [66]
    [PDF] Plutonium Pit Production - Union of Concerned Scientists
    May 29, 2025 · In 2014, Congress mandated that the NNSA produce 10 plutonium pits in 2024, 20 in 2025, and ... from developing new nuclear warheads, to the 2024 ...
  67. [67]
  68. [68]
    Russia moves to withdraw from plutonium agreement with ... - Reuters
    Oct 8, 2025 · After dismantling thousands of warheads after the Cold War, both Moscow and Washington were left with huge stockpiles of weapons-grade plutonium ...
  69. [69]
    Glenn Theodore Seaborg | Science History Institute
    With his colleagues at the University of California at Berkeley, Glenn Seaborg discovered the element plutonium in late 1940. He went on to identify several ...
  70. [70]
    Man's First Glimpse of Plutonium - The New York Times
    Jul 16, 1985 · My colleagues and I synthesized, identified and determined the key nuclear properties of plutonium in late 1940 and early 1941 in Berkeley ...<|control11|><|separator|>
  71. [71]
    Manhattan Project: The Plutonium Path to the Bomb, 1942-1944 - OSTI
    Time Periods · Production Reactor (Pile) Design, 1942 · DuPont and Hanford, 1942 · CP-1 Goes Critical, December 2, 1942 · Seaborg and Plutonium Chemistry, 1942-1944 ...
  72. [72]
    plutonium - Manhattan Project - National Park Service
    Plutonium, discovered in 1940, became essential for atomic weapons. It was produced by transmuting uranium in Hanford, and used in the Fat ...
  73. [73]
    Understand the PAST - Hanford Site
    Jun 17, 2025 · Workers completed design and construction of the world's first plutonium producing reactor (B Reactor) in 11 months. The B Reactor produced ...
  74. [74]
    Hanford Nuclear Reservation
    The first plutonium was delivered to the Army on February 2, 1945. The first test blast was on July 16, in New Mexico, and nuclear bombs were dropped on ...Missing: timeline | Show results with:timeline
  75. [75]
    Manhattan Project: Hanford Becomes Operational, 1943-1944
    Excitement mounted at Hanford as the date for the first operation of a plutonium production reactor approached. On September 13, 1944, Enrico Fermi placed the ...
  76. [76]
    A Tale of Two Bomb Designs | Los Alamos National Laboratory
    Oct 10, 2023 · The insurmountable challenges associated with a plutonium gun-type design prompted the shift to the implosion-style weapon. Creating atomic ...
  77. [77]
    Manhattan Project: The Trinity Test, July 16, 1945 - OSTI.gov
    Little Boy, the uranium bomb, was dropped first at Hiroshima on August 6, while the plutonium weapon, Fat Man, followed three days later at Nagasaki on August 9 ...
  78. [78]
    'Destroyer of Worlds': The Making of an Atomic Bomb | New Orleans
    Jul 10, 2025 · THE TRINITY TEST​​ The test bomb, nicknamed Gadget, contained 13 pounds of plutonium, as well as the implosion-method of detonation. Using a ...
  79. [79]
    Accounting for Unfissioned Plutonium from the Trinity Atomic Bomb ...
    The Trinity test device contained about 6 kg of plutonium as its fission source, resulting in a fission yield of 21 kT. However, only about 15% of the 239 Pu ...
  80. [80]
    Fact Sheet: U.S. Plutonium Pit Production
    Aug 16, 2021 · The United States produced between 1,000 and 2,000 pits per year during the Cold War at the Rocky Flats Plant near Golden, Colorado. The ...
  81. [81]
    The U.S. Nuclear Weapons Stockpile - Department of Energy
    As of 2023, the stockpile stood at 3,748 warheads—a roughly 88 percent reduction in size since its peak of 31,255 warheads in the late 1960s.
  82. [82]
    Nuclear weapon - Soviet Union, Cold War, Arms Race - Britannica
    Sep 23, 2025 · The first Soviet test occurred on August 29, 1949, using a plutonium device (known in the West as Joe-1) with a yield of approximately 20 ...<|separator|>
  83. [83]
    The History of Plutonium Production in Russia
    For almost 50 years, the production and processing of weapon fissile materials was the primary mission of the Soviet nuclear complex. The cloak of secrecy on ...
  84. [84]
    Managing Military Uranium and Plutonium in the United States and ...
    Nov 30, 2016 · Plutonium. 100, 160 ; HEU, 645, 1050 ; Outside of warheads at stockpile of 10,000 ; Plutonium, 65, 125 ; HEU, 420, 825.
  85. [85]
    Fact Sheet: The Civilian and Military Characteristics of Plutonium
    Aug 26, 2021 · Nuclear weapons require fissile materials, that is certain isotopes of uranium and plutonium, to sustain explosive fission chain reactions.
  86. [86]
    U.S. Program for Disposition of Excess Weapons Plutonium
    The United States has declared that 52.7 tons of plutonium, over half of its stockpile, is excess to its military needs, and has decided to pursue a dual-track ...Missing: post- | Show results with:post-
  87. [87]
    What went wrong with US plutonium disposition
    Apr 24, 2018 · Why the US has failed to dispose of weapons-grade nuclear material as planned and promised, and what to do about it.
  88. [88]
    Resources on U.S. Plutonium Disposition
    Oct 6, 2021 · In September 2000, both countries signed the Plutonium Management and Disposition Agreement (PMDA), in which they committed to transform 34 ...
  89. [89]
    Russia withdraws from 25-year-old weapons-grade plutonium ...
    Russia withdraws from 25-year-old weapons-grade plutonium agreement. Russia's lower house of Parliament, the State Duma, approved a measure to withdraw from a ...
  90. [90]
    Plutonium Disposition by “Downblending and Disposal”
    The subject of plutonium disposition has a long history that dates back to the end of the Cold War, combining complex technical, policy, and diplomatic ...
  91. [91]
    Stockpile Stewardship and Management Plan (SSMP)
    Oct 3, 2024 · The report outlines plans to fulfill the requirements to produce a minimum of 80 plutonium pits per year; achieve the First Production Units ...
  92. [92]
    Stockpile Stewardship - Nevada National Security Site
    Stockpile Stewardship ensures the nuclear weapons stockpile is safe, reliable, and secure, using experiments, audits, and simulations, and was established to ...
  93. [93]
    NNSA releases 2025 Stockpile Stewardship and Management Plan
    Oct 3, 2024 · This year's SSMP details NNSA's efforts to establish the capability to produce a minimum of 80 plutonium pits per year; achieve the First ...
  94. [94]
    [PDF] Fiscal Year 2024 - Stockpile Stewardship and Management Plan
    The plan describes how to maintain the nuclear stockpile, produce 80 plutonium pits per year, modernize the stockpile, and recapitalize infrastructure.
  95. [95]
    Expansion of test programme for MOX fuel for VVER reactors
    Jan 23, 2025 · MOX fuel is manufactured from plutonium recovered from used reactor fuel, mixed with depleted uranium. It allows the recycling of used nuclear ...
  96. [96]
    Pitting Nuclear Modernization Against Powering AI: Trump's Plans ...
    Oct 14, 2025 · In 2015, the DOE adopted a new plan for disposing of the old weapons plutonium, namely a dilute and dispose program. The plan assumed that the ...Missing: post- | Show results with:post-
  97. [97]
    Plutonium-239 | Pu | CID 61782 - PubChem - NIH
    It is fissile and can therefore sustain a nuclear chain reaction, leading to applications in nuclear weapons and nuclear reactors. Pu-239 is synthesized by ...
  98. [98]
    Fat Man: Implosion-Type Bomb - Atomic Archive
    The initial design for the plutonium bomb was also based on using a simple gun design (known as the "Thin Man") like the uranium bomb.
  99. [99]
    "Fat Man" Atomic Bomb - Air Force Museum
    A "Fat Man" bomb was dropped over Nagasaki, Japan, on Aug. 9, 1945, near the end of World War II. Released by the B-29 Bockscar, the 10000-pound weapon was ...
  100. [100]
    Atomic Weapons - Manhattan Project National Historical Park (U.S. ...
    The gun-type bomb, named Little Boy, used enriched uranium from Oak Ridge. Fat Man, the implosion-type weapon, used plutonium produced at ...
  101. [101]
    Deterrence defined - Los Alamos National Laboratory
    Apr 2, 2024 · Deterrence simply means preventing something. Most people associate deterrence with preventing war, keeping America safe, and convincing potential adversaries ...<|separator|>
  102. [102]
    [PDF] THE NUCLEAR FUEL CYCLE
    The spent fuel contains uranium (96%), plutonium (1%) and high level waste products (3%). The uranium, with less than 1% fissile 235U and the plutonium can be ...<|separator|>
  103. [103]
    Mixed Oxide (MOX) Fuel - World Nuclear Association
    Oct 10, 2017 · An innovative development in recycling plutonium and uranium as MOX is Russia's REMIX fuel, not yet commercialised. A further alternative is ...MOX use · MOX production · MOX and disposition of... · Dual-component power...
  104. [104]
    Fast Neutron Reactors - World Nuclear Association
    Aug 26, 2021 · Here, fast breeder reactors form stage 2 and use plutonium-based fuel in the core to breed both U-233 from thorium and Pu-239 from U-238 in the ...
  105. [105]
    [PDF] Liquid Metal Fast Breeder Reactors
    In a fast breeder reactor there is fertile material (uranium-238) in the core and in the blanket around the core. The core consists of a mixture of plutonium ...
  106. [106]
    About Plutonium-238 - NASA Science
    Feb 7, 2024 · Plutonium-238 is a special material that emits steady heat due to its natural radioactive decay. Several unique features of plutonium-238 have made it the ...
  107. [107]
    Nuclear Reactors and Radioisotopes for Space
    The high decay heat of Plutonium-238 (0.56 W/g) enables its use as an electricity source in the RTGs of spacecraft, satellites and navigation beacons. Its ...Missing: facts | Show results with:facts
  108. [108]
    Radioisotope Power Systems FAQ - NASA Science
    Jan 5, 2024 · The relatively short half-life of Pu-238 (88 years) means it degrades naturally in a manageable manner over the length of a typical space ...
  109. [109]
    Multi-Mission Radioisotope Thermoelectric Generators, Plutonium ...
    Jul 30, 2020 · The MMRTG will provide heat and power to Perseverance as it spends at least one Mars year (687 Earth days) searching for signs of past microbial life.
  110. [110]
    Radioisotope Power Systems Missions - NASA Science
    The Perseverance rover is powered and warmed by a Multi-Mission Radioisotope Thermoelectric Generator (MMRTG). The rover continues to collect samples on the ...
  111. [111]
    Pu-238 shipment quantity 'opens the tap' for space missions | ORNL
    Sep 15, 2023 · In 2015, NASA contracted with the Department of Energy and ORNL to find a process to again produce Pu-238. The last time large quantities were ...
  112. [112]
    DOE ramps up plutonium oxide production to fuel NASA's deep ...
    Jul 20, 2023 · The Department of Energy recently shipped half a kilogram of plutonium oxide pellets from Oak Ridge National Laboratory to Los Alamos ...<|control11|><|separator|>
  113. [113]
    Generation IV Nuclear Reactors
    Apr 30, 2024 · ... reactors (PHWR) fuelled by natural uranium to generate plutonium. Then fast breeder reactors (FBRs) use this plutonium-based fuel to breed U ...
  114. [114]
    Generation IV Goals, Technologies and GIF R&D Roadmap
    Enhanced sustainability is achieved primarily through the adoption of a closed fuel cycle including the reprocessing and recycling of plutonium, uranium and ...
  115. [115]
  116. [116]
    Pitting Nuclear Modernization Against Powering AI: Trump's Plans ...
    Oct 14, 2025 · The plutonium stockpile is spread across sites around the country, including at Savannah River, the Pantex Plant in Texas, and Los Alamos ...
  117. [117]
    Advanced reactor models for plutonium multirecycling scenario ...
    This paper presents an innovative way to identify the plutonium content in MOX fresh fuel according to its isotopic composition in fuel cycle dynamic ...
  118. [118]
    Russia Advances REMIX Fuel Trials in Boost for Closed Nuclear ...
    Dec 5, 2024 · Russia has launched its third cycle of testing innovative uranium-plutonium REMIX fuel at the Balakovo Nuclear Power Plant in Saratov Oblast.
  119. [119]
    Plutonium Management – PuMMA
    The Generation IV reactors will be able to use and multi-recycle the accumulated Pu but also to overcome the use of fresh uranium coming from ores, thus ...
  120. [120]
    [PDF] ATSDR Plutonium Tox Profile
    As discussed in Section 3.5, Mechanisms of Toxicity, plutonium-induced health effects are considered to be the result of energy deposited by alpha particle ...
  121. [121]
    [PDF] Safe handling and storage of plutonium - The Nuclear Threat Initiative
    The objective of this safety report is to describe in a single publication the best current practices for handling and storing plutonium while providing ...
  122. [122]
    [PDF] A Review of Criticality Accidents - Nuclear Regulatory Commission
    criticality safety: Protection from the consequences of a criticality accident, preferably by prevention of the accident. Encompasses procedures, training ...<|separator|>
  123. [123]
    [PDF] Criticality Safety Basics for INL FMHs and CSOs
    Most, but not all, criticality accidents involved highly enriched uranium or plutonium. But lower enriched systems are not automatically critically safe. The ...
  124. [124]
    Plutonium pyrophoricity (Technical Report) | OSTI.GOV
    Jun 2, 1992 · Plutonium metal that will ignite spontaneously in air at a temperature of 150{degrees}C or below in the absence of external heat, shock, or friction.
  125. [125]
    A mechanism for plutonium pyrophoricity - ScienceDirect.com
    Plutonium must exceed 475°C before self-ignition occurs. External heating of massive samples is necessary to achieve this condition, while finely divided ...
  126. [126]
    [PDF] SAFETY DATA SHEET PLUTONIUM METAL - Department of Energy
    If large amounts become internally deposited, lung cancer, sterility, anemia, leukemia, or bone cancer may occur. SKIN CONTACT: ACUTE EXPOSURE - Alpha radiation ...Missing: emission | Show results with:emission
  127. [127]
    [PDF] Guide of Good Practices for Occupational Radiological Protection in ...
    Plutonium air-handling systems must be operated within the design safety envelope of the system, designated by Operational Safety Requirements (OSRs).
  128. [128]
    [PDF] Regulatory Guide 3.16, General Fire Protection Guide for Plutonium ...
    Special control should be exercised over the handling of flammable, toxic, and explosive gases, chemicals, and materials admitted to the plutonium handling ...
  129. [129]
    [PDF] Radiological Safety Training for Plutonium Facilities
    Because liquid containing plutonium is a special criticality concern, care is required when handling plutonium-bearing liquids. Containers that could hold ...Missing: protocols | Show results with:protocols
  130. [130]
    PART 71—PACKAGING AND TRANSPORTATION OF ...
    Part 71 covers general provisions, package approval standards, and operating controls for the packaging and transportation of radioactive material.
  131. [131]
    Office of Secure Transportation | Department of Energy
    OST is responsible for the safe and secure transport in the contiguous United States of government-owned special nuclear materials.
  132. [132]
    [PDF] Transportation of Safeguards Material - Teppinfo.com
    OST uses secure tractor-trailers, armed federal agents, and specially designed rigs for over-the-road transport, with no fatal accidents since 1975.
  133. [133]
    [PDF] Regulations for the Safe Transport of Radioactive Material
    IAEA safety standards for transport of radioactive material, part of the IAEA Safety Standards Series, are for protection of health and minimization of danger.
  134. [134]
    Transport of Radioactive Material - World Nuclear Association
    Feb 28, 2025 · Plutonium oxide is transported in several different types of sealed package and each can contain several kilograms of material. Criticality is ...
  135. [135]
    [PDF] Plutonium Transportation – Risks and Benefits Larry R. Foulke, PhD ...
    Plutonium transport uses double containment with strict leak limits, and containers are tested for 30-foot drops, fire, and water immersion. Regulations limit ...
  136. [136]
    [PDF] Criticality Safety in the Handling of Fissile Material
    This IAEA publication, SSG-27 (Rev. 1), is a safety guide for criticality safety in handling fissile material, part of the IAEA Safety Standards Series.
  137. [137]
    § 71.88 Air transport of plutonium. | Nuclear Regulatory Commission
    (4) The plutonium is shipped in a package specifically authorized for the shipment of plutonium by air in the Certificate of Compliance for that package issued ...
  138. [138]
    [PDF] U.S. Regulatory Evolution Relative to Secure Transportation
    Since 9/11, the NRC has issued orders to licensees to increase physical security of nuclear and radioactive materials transport. Orders have the force of Law.<|separator|>
  139. [139]
    [PDF] Joint United States/IAEA Proposed Approach for Safeguards During ...
    The stabilization process called for by DOE-STD-3013 involves subjecting the plutonium oxide to a two-hour soak at 950 ЛC in an oxygen rich atmosphere to.
  140. [140]
    [PDF] Nuclear Material Control and Accountability - DOE Directives
    Feb 7, 2023 · This Order establishes requirements for developing, implementing, and maintaining a nuclear material control and accountability (MC&A) program ...<|separator|>
  141. [141]
    2022 civilian plutonium declarations submitted to the IAEA - IPFM Blog
    Mar 25, 2024 · Only about 140 tons of this material are under international (IAEA or Euratom) safeguards.
  142. [142]
    [PDF] Regulatory Guide 3.71, Revision 3, Nuclear Criticality Safety ...
    ANSI/ANS-8.12-1987 (Reaffirmed 2016), “Nuclear Criticality Control and Safety of Plutonium-. Uranium Fuel Mixtures Outside Reactors,” American Nuclear ...
  143. [143]
    Nuclear Weapons Primer
    Both plutonium-239 and uranium-235 have been used as nuclear explosives in fission weapons. Approximately ninety percent of the effort that went into making ...Missing: efficiency | Show results with:efficiency
  144. [144]
    Nuclear Weapons Design Complexity: Beyond 90% Uranium ...
    Jul 6, 2025 · This property distinguishes uranium from plutonium in terms of design options. Spontaneous fission happens when unstable nuclei randomly split ...U-235 Vs U-238: Fission... · Thermal Neutron Absorption... · Subcritical Testing And...
  145. [145]
    Plutonium-239 and Its Relationship with Uranium-235 in ...
    Dec 17, 2024 · To optimize yield and efficiency of uranium-235 and plutonium-239, a plutonium pit is surrounded by a uranium tamper in a thermonuclear weapon.Missing: synthesis deuteron bombardment
  146. [146]
    [PDF] David Wang STS.092 Plutonium vs. Uranium
    The advantage of this. “thermonuclear” device versus the “nuclear” uranium and plutonium bombs is that it produces less radiation, although it is more ...
  147. [147]
    Risks of Civilian Plutonium Programs - The Nuclear Threat Initiative
    Jun 30, 2004 · Growing stockpiles of separated civilian plutonium increase the risk that states or terrorists might acquire this material to build nuclear weapons.Missing: necessities | Show results with:necessities
  148. [148]
    [PDF] Policy Primer on Nuclear Energy: Understanding Proliferation ...
    Sep 6, 2023 · Without a changing security environment, increasing nuclear energy utilization implicitly increases the risk of nuclear weapons proliferation, ...
  149. [149]
    Are IAEA Safeguards on Plutonium Bulk-Handling Facilities Effective?
    The purpose of this paper is to assess the effectiveness of international (IAEA) safeguards at peaceful nuclear fuel cycle facilities which handle plutonium in ...
  150. [150]
    Safeguards to Prevent Nuclear Proliferation
    May 6, 2021 · Safeguards are designed to deter diversion of nuclear material by increasing the risk of early detection. They are complemented by controls on ...Missing: necessities | Show results with:necessities
  151. [151]
    Appendix A: Revising the Practice of Deterrence | Post-Cold War ...
    1 The familiar central doctrine holds that nuclear weapons are maintained to prevent their use and, by extension, any large-scale form of warfare by threatening ...Missing: evidence | Show results with:evidence
  152. [152]
    Public opinion on nuclear energy and nuclear weapons
    The pro-nuclear weapons treatment argued that nuclear deterrence reduces the chances of war between nuclear-armed adversaries, whereas the pro-nuclear power ...
  153. [153]
    Nuclear Fuel Cycle and Proliferation - NMHB 2020 [Revised]
    Plutonium-239 is the most efficient of all the fissile isotopes producible in large quantities and has a relatively long half-life of 24,100 years and a low ...<|separator|>
  154. [154]
    Proliferation is a completely invalid argument against nuclear ...
    Jun 15, 2025 · The proliferation concerns regarding nuclear reprocessing do make sense but they are not a valid argument against reprocessing.Missing: necessities deterrence
  155. [155]
    [PDF] IAEA Safeguards - Serving Nuclear Non-Proliferation
    Radioactive sources that do not contain uranium, plutonium or thorium are not subject to safeguards and need not be reported to the IAEA under a safeguards.
  156. [156]
    Assessing the nuclear weapons proliferation risks in nuclear energy ...
    This research seeks to determine whether future SMR deployment in NEN countries will contribute to nuclear weapons proliferation, and how the risks can be ...Missing: necessities deterrence
  157. [157]
    Influence Without Arms: The New Logic of Nuclear Deterrence
    Dec 5, 2024 · And we also see that latent powers in this circumstance are able to deter their rivals from building nuclear weapons. However, if we see an ...
  158. [158]
    Plutonium | Public Health Statement | ATSDR - CDC
    Plutonium is a silvery-white radioactive metal that exists as a solid under normal conditions. It is produced when uranium absorbs an atomic particle.Missing: radiological | Show results with:radiological
  159. [159]
    [PDF] Toxicological Profile for Plutonium
    Trace amounts of plutonium are found worldwide, mostly due to fallout from atmospheric nuclear testing, which ended in 1980, and released several isotopes ...<|separator|>
  160. [160]
    [PDF] Plutonium and Americium Geochemistry at Hanford: A Site-Wide ...
    This report was produced to provide a systematic review of the state-of-knowledge of plutonium and americium geochemistry at the Hanford Site.
  161. [161]
    Subsurface Transport of Plutonium in Organic and Aqueous Acidic ...
    May 10, 2024 · This study demonstrates that Pu migration can occur both with and without organic solvents at pH < 4, but significantly more Pu can be transported when ...
  162. [162]
    Fallout from Nuclear Weapons Tests and Cancer Risks
    The legacy of open-air nuclear weapons testing includes a small but significant increase in thyroid cancer, leukemia and certain solid tumors.
  163. [163]
    [PDF] Environmental Consequences of the Chernobyl Accident and their ...
    plutonium) were released into the atmosphere primarily in the form of fuel particles. Other radio- nuclides (isotopes of caesium, tellurium, antimony, etc ...
  164. [164]
    Chernobyl: Chapter VI. Agricultural and environmental impacts
    The releases during the Chernobyl accident contaminated about 125 000 km2 of land in Belarus, Ukraine and Russia with radiocaesium levels greater than 37 ...
  165. [165]
    [PDF] Hanford Annual Site Environmental Report for Calendar Year 2023
    Oct 5, 2024 · The Hanford Site was established in 1943 to produce plutonium for atomic weapons during World War II and the Cold War. The site has restricted ...
  166. [166]
    Rocky Flats - historical public exposure studies
    The studies, conducted 1990-1999, estimated cancer risk to residents from 1952-1989, focusing on past off-site releases and how people were exposed.
  167. [167]
    Frequently asked questions about Rocky Flats
    What are the risks to a refuge visitor? Can inhalation of even one particle of plutonium result in cancer in humans? What are the monitoring requirements for ...
  168. [168]
    Assessing The Impacts of Pollution at the Hanford Nuclear Site
    Aug 1, 2019 · The environmental impacts of these hazardous substances are currently being evaluated by NOAA and our partners through a natural resource ...
  169. [169]
    [PDF] 2021 Data Summary Report - Washington State Department of Health
    ... Hanford Site was the production of plutonium for nuclear weapons. Operations resulted in releases of radioactivity to the environment. Today, weapons ...
  170. [170]
    Nuclear Waste Disposal | U.S. GAO - Government Accountability Office
    According to federal law, certain high-level mixed waste must be vitrified—a process in which the waste is immobilized in glass—and disposed of in a deep ...
  171. [171]
    Oct. 15 - Hanford Site waste vitrification begins
    Oct 15, 2025 · Low-activity waste vitrified at the Waste Treatment Plant will be permanently disposed of at the nearby Integrated Disposal Facility.
  172. [172]
    Hanford Vit Plant: Home
    When complete, the Waste Treatment Plant will be used to transform the 56,000,000 gallons of radioactive and chemical wastes being stored in underground tanks ...
  173. [173]
    Processing of Used Nuclear Fuel
    Aug 23, 2024 · Most of the separated plutonium is used almost immediately in mixed oxide (MOX) fuel. World MOX production capacity is currently around 480 ...
  174. [174]
    Recovery and recycling of nuclear fuels - Orano
    Through recycling, up to 96% of the reuseable material in used fuel can be recovered: 1% plutonium, and 95% uranium. The remaining 4% of material are fission ...Missing: via | Show results with:via<|control11|><|separator|>
  175. [175]
    [PDF] Managing Waste With and Without Plutonium Separation - RAND
    Again, since the largest waste volume comes from mill tailings, a reprocessed cycle with less mining and milling would have much less waste volume and thus ...
  176. [176]
    US takes another look at recycling nuclear fuel | Physics Today
    Feb 1, 2024 · Fast reactors and reprocessing could reduce by 90% the volume of nuclear waste that will need to be stored in a geological repository for tens ...
  177. [177]
    Backgrounder on Radioactive Waste
    The uranium and plutonium can be used again as fuel. Most of the high-level waste (other than spent fuel) generated over the last 35 years has come from ...
  178. [178]
    Physics of plutonium recycling - Nuclear Energy Agency (NEA)
    Plutonium recycling is important for economics, resource use, and waste management. It involves thermal and fast reactors, including MOX usage and fast reactor ...
  179. [179]
    Chapter 5: The Manhattan District Experiments
    Cade had been hospitalized since his accident, but the plutonium injection did not take place until April 10. On this date, "HP-12" (the code name HP--"human ...
  180. [180]
    Human Radiation Experiments - Atomic Heritage Foundation
    Jul 11, 2017 · The plutonium tests were intended to determine how excreta could be used to estimate the amount of plutonium remaining in an exposed subject.Missing: LD50 | Show results with:LD50
  181. [181]
    Advisory Committee on Human Radiation Experiments
    Some of the experiments the Committee was asked to investigate, and particu larly a series that included the injection of plutonium into unsus pecting hospital ...
  182. [182]
    Advisory Committee on Human Radiation Experiments
    In chapter 5, we look at the Manhattan Project plutonium-injection experiments and related experimentation. Sick patients were used in sometimes secret ...
  183. [183]
    The World War II plutonium experiments: contested stories and their ...
    During WWII, 18 hospital patients were injected with radioactive plutonium to study its effects on the body.