Fact-checked by Grok 2 weeks ago

Nuclear reactor physics

Nuclear reactor physics is the field of physics that examines , , and related phenomena to enable the design, operation, and safety analysis of nuclear fission reactors. It focuses on achieving a controlled, self-sustaining where each event produces sufficient to induce further fissions, typically through the effective multiplication factor k_{\text{eff}}, defined such that k_{\text{eff}} = 1 for criticality, k_{\text{eff}} > 1 for supercriticality, and k_{\text{eff}} < 1 for subcriticality. Fundamental to this are interactions including moderation to thermal energies for efficient in fuels like uranium-235, absorption by control materials, and leakage from the core, balanced to maintain neutron economy. Central concepts include the four-factor formula for neutron multiplication in thermal reactors, k = \eta f p \varepsilon, where \eta is the neutrons produced per absorption in fuel, f the thermal utilization factor, p the resonance escape probability, and \varepsilon the fast fission factor, providing a first-principles basis for core design calculations. Reactor kinetics govern dynamic behavior, with delayed neutrons—originating from fission product decay and comprising about 0.65% of total neutrons for —crucial for controllability, as their ~10-60 second half-lives prevent rapid power excursions and allow insertion of absorbers like boron or cadmium rods to adjust reactivity. Notable achievements encompass enabling high-efficiency energy extraction, with each fission releasing approximately 200 MeV primarily as kinetic energy of fragments convertible to heat, yielding energy densities orders of magnitude beyond chemical fuels. Defining characteristics include inherent safety features from physics, such as negative reactivity coefficients from of resonances and moderator density changes, though historical accidents underscore the need for precise modeling of transients and poison buildup like . This discipline underpins diverse reactor types, from light-water moderated systems to fast breeders, supporting scalable, low-carbon power generation.

Fundamental Concepts

Nuclear Fission and Chain Reactions

Nuclear fission is a nuclear reaction in which the nucleus of a heavy isotope, such as (^235U), absorbs a neutron and subsequently splits into two lighter nuclei known as fission products, releasing binding energy in the form of kinetic energy, gamma rays, and typically 2 to 3 prompt neutrons per event. In reactor applications, this process is induced primarily by low-energy thermal neutrons, which have a high probability of absorption by ^235U, forming an excited ^236U compound nucleus that deforms and divides asymmetrically, often yielding products like barium-141 and krypton-92 or xenon-140 and strontium-94, along with the neutrons. The total recoverable energy released per fission is approximately 200 MeV, with about 168 MeV as kinetic energy of the fission fragments, 7 MeV in prompt gamma rays, and the remainder in neutron kinetic energy and beta decays of products. The prompt neutrons emitted, with an average multiplicity of around 2.4 for thermal-neutron-induced fission of ^235U, possess energies primarily between 0.5 and 2 MeV, enabling them to potentially interact with other fissile nuclei. This sets the stage for a chain reaction, wherein neutrons from one fission event are absorbed by additional ^235U nuclei, inducing further fissions and potentially leading to exponential growth in the neutron population if the average number of neutrons producing subsequent fissions exceeds one. In an uncontrolled scenario, such as in nuclear explosives, rapid supercritical multiplication occurs, but nuclear reactors are designed to sustain a controlled chain reaction at or near criticality, where the fission rate remains steady, by incorporating neutron absorbers (control rods) and moderators to manage neutron flux and spectrum. The feasibility of a sustained chain reaction depends on the neutron economy, influenced by factors like the probability of neutron absorption leading to versus capture without , but fundamentally arises from the excess neutrons beyond those needed to maintain the reaction. also produces delayed neutrons from subsequent decays of certain , which, though fewer (about 0.65% of total for ^235U), provide essential controllability by slowing the reaction kinetics and allowing human intervention to prevent runaway excursions. This combination of prompt and delayed neutrons distinguishes reactor chain reactions from purely explosive ones, enabling safe power generation through heat extraction via coolant systems.

Neutron Cross Sections and Interactions

Neutron cross sections represent the effective probability of interaction between an incident neutron and a target nucleus, quantified as an apparent geometrical area in units of barns, where 1 barn equals 10^{-28} square meters. The microscopic cross section, denoted σ, applies to individual nuclides and varies by reaction type and neutron energy, while the macroscopic cross section, Σ = N σ (with N as atomic number density), governs bulk material attenuation and determines the mean free path λ = 1/Σ for neutrons traversing a medium. Interactions fall into scattering and absorption categories. Elastic scattering conserves kinetic energy between neutron and nucleus via momentum transfer, predominant for fast neutrons (energies above ~1 MeV) and essential for moderation in reactors, where light nuclei like hydrogen in water reduce neutron speeds through repeated collisions. Inelastic scattering excites the nucleus to higher energy states, emitting gamma rays upon de-excitation, becoming significant above ~1 MeV and contributing to heating in structural materials. Absorption encompasses radiative capture (n,γ), forming a compound nucleus that decays by gamma emission without fission, and induced fission (n,f) in fissile isotopes like ^{235}U, where the nucleus splits into fragments, releasing 2-3 neutrons and ~200 MeV energy. Cross sections exhibit strong energy dependence. At thermal energies (~0.025 eV), absorption cross sections for many nuclides follow the 1/v law, inversely proportional to neutron velocity due to the constant s-wave scattering amplitude in quantum mechanics, yielding high values such as ~584 barns for ^{235}U fission. In the epithermal range (1 eV to ~100 keV), resonant absorption peaks occur when incident neutron energy aligns with excitation levels of the compound nucleus, modeled by the Breit-Wigner formula σ(E) ∝ Γ_n Γ_γ / [(E - E_r)^2 + (Γ/2)^2], where E_r is resonance energy and Γ denotes widths for neutron capture or fission channels; these resonances, spaced ~0.1-10 eV apart, cause rapid σ variations and influence reactor fuel utilization. For fast neutrons (>1 MeV), cross sections approach geometric limits (~πR^2, with R the nuclear radius ~1.2 A^{1/3} fm, yielding ~1-10 barns) dominated by classical shadowing, decreasing as E^{-1/2} for per optical model approximations. Evaluated libraries like ENDF/B-VIII.0 provide parameters and cross section data up to 20 MeV for , derived from integral experiments and theoretical models to ensure accuracy in predicting economy. In light water reactors, ^{238}U exhibits negligible thermal (σ_f < 0.001 barns) but strong resonant capture (~2.7 barns average in resolved range), acting as a parasitic absorber that depletes neutrons unless compensated by fast spectrum breeding. These interactions underpin criticality calculations, with total cross sections dictating diffusion and absorption probabilities central to sustaining chain reactions.

Neutron Economy and Criticality

Multiplication Factor and Criticality

The effective multiplication factor, k_{\mathrm{eff}}, represents the ratio of the number of neutrons produced by fission in one neutron generation to the number of neutrons lost (through absorption or leakage) from the preceding generation in a finite reactor core. This factor determines whether a chain reaction can be sustained, with k_{\mathrm{eff}} accounting for geometric leakage effects absent in the infinite multiplication factor k_\infty. In practice, k_{\mathrm{eff}} is computed using transport or diffusion theory codes calibrated against benchmark experiments, as direct measurement relies on observable neutron populations during startup. The value of k_{\mathrm{eff}} is expressed through the six-factor formula k_{\mathrm{eff}} = \eta f p \varepsilon P_{\mathrm{FNL}} P_{\mathrm{TNL}}, where \eta is the reproduction factor (average neutrons emitted per thermal neutron absorbed in fissile material, approximately 2.42 for uranium-235), f is the thermal utilization factor (fraction of thermal neutrons absorbed in fuel versus all materials), p is the resonance escape probability (fraction of fast neutrons slowing to thermal energies without resonance capture, typically 0.95–0.99), \varepsilon is the fast fission factor (additional neutrons from fast fission beyond thermal-induced, around 1.03 in light-water reactors), P_{\mathrm{FNL}} is the fast non-leakage probability, and P_{\mathrm{TNL}} is the thermal non-leakage probability. These factors derive from microscopic cross-sections and macroscopic reactor composition, with k_\infty = \eta \varepsilon f p for an infinite lattice ignoring leakage. Criticality occurs precisely when k_{\mathrm{eff}} = 1, balancing neutron production exactly with losses to maintain a steady-state chain reaction, though delayed neutrons introduce kinetic delays preventing instantaneous equilibrium. A subcritical configuration (k_{\mathrm{eff}} < 1) results in of the neutron population, as losses exceed production, requiring an external source for detectable flux. Supercriticality (k_{\mathrm{eff}} > 1) drives , with the neutron doubling time inversely proportional to \ln k_{\mathrm{eff}} per generation (typically microseconds for prompt neutrons). Reactivity \rho, defined as \rho = \frac{k_{\mathrm{eff}} - 1}{k_{\mathrm{eff}}}, quantifies deviation from criticality in fractional units, often expressed in percent \Delta k/k (1% = 0.01) or pcm (1 pcm = 10^{-5} \Delta k/k). Thus, \rho = 0 at criticality, \rho < 0 for subcritical states, and \rho > 0 for supercritical ones; control systems adjust \rho via or poisons to maintain safe operation near zero.

Subcritical Multiplication


Subcritical multiplication refers to the amplification of neutron flux in a nuclear system where the effective factor k < 1, driven by an external neutron source that initiates limited fission chains. In this regime, neutron production from fission is insufficient to sustain the chain reaction independently, but source neutrons generate additional fission neutrons over successive generations, resulting in a total neutron population exceeding the source level.
The total steady-state neutron population N satisfies the balance equation where losses equal the sum of source neutrons and fission-produced neutrons, yielding N = \frac{S}{1 - k}, with S as the external source strength. This formula derives from the geometric series expansion of neutron generations: source neutrons S produce kS in the first fission generation, k^2 S in the second, and so forth, summing to S \sum_{n=0}^{\infty} k^n = \frac{S}{1 - k} since k < 1. The subcritical multiplication factor M = \frac{1}{1 - k} thus represents the ratio of total neutrons to source neutrons. Reactivity \rho, defined as \rho = \frac{k - 1}{k} , is negative for subcritical conditions (\rho < 0). Solving for k = \frac{1}{1 - \rho} shows M = \frac{1 - \rho}{\rho (1 - \rho)/k}, but precisely M = \frac{1 + |\rho|}{|\rho|} where \rho = -|\rho|, approximating to M \approx -\frac{1}{\rho} for small |\rho| \ll 1. This approximation holds as k approaches 1 from below. During reactor startup from a deeply subcritical state, external sources such as or maintain detectable flux levels for instrumentation. Neutron detector counts, inversely proportional to M, are plotted against control rod positions to extrapolate the critical configuration where k = 1 and M \to \infty. This "1/M method" predicts criticality without exceeding it, ensuring safety; for instance, in one experimental setup, counts versus rod height intersected zero at the critical blade position of approximately 8 inches. Subcritical multiplication also underpins accelerator-driven systems, where high-intensity spallation sources achieve effective power production despite k \approx 0.95 - 0.98, enhancing safety by allowing prompt shutdown upon source cessation.

Neutron Moderation and Spectrum

Moderators and Thermalization

In thermal reactors, moderators slow fast neutrons emitted from fission—averaging approximately 2 MeV in energy—to thermal energies around 0.025 eV at typical operating temperatures, thereby increasing the fission cross-section for isotopes such as uranium-235. This thermalization occurs primarily through elastic scattering collisions with moderator nuclei, where neutrons lose kinetic energy incrementally until reaching thermal equilibrium with the surrounding medium. The process relies on the kinematics of neutron-nucleus collisions, with the fractional energy loss maximized for low-mass nuclei; in a head-on collision, a neutron transfers up to \frac{4A}{(A+1)^2} of its energy to a target nucleus of mass number A, favoring elements like hydrogen (A=1) over heavier ones. Effective moderators must exhibit high macroscopic scattering cross-sections (\Sigma_s) relative to absorption cross-sections (\Sigma_a), minimizing neutron capture while efficiently decelerating particles. The average logarithmic energy decrement per collision, \xi = 1 + \frac{(A-1)^2}{2A} \ln \left( \frac{A-1}{A+1} \right), quantifies slowing-down efficiency, with \xi \approx 1 for hydrogen enabling rapid moderation in roughly 18 collisions from 2 MeV to thermal energies. Deuterium (\xi \approx 0.51) requires about 35-40 collisions, while carbon (\xi \approx 0.158) demands around 114, leading to longer migration distances and potential leakage in reactor designs. The overall merit is captured by the moderating ratio \xi \Sigma_s / \Sigma_a, which accounts for parasitic absorption; low values necessitate enriched fuel or larger moderator volumes to achieve criticality. Common moderators balance these properties with practical considerations like cost, stability, and compatibility with coolants. Light water (H_2O) has a moderating ratio of approximately 62 but suffers from higher absorption by hydrogen-1, requiring enriched uranium fuel in pressurized and boiling water reactors. Heavy water (D_2O), with a ratio exceeding 7,000 due to deuterium's negligible thermal absorption, permits natural uranium fueling in , though its production cost limits widespread use. Graphite, offering a ratio around 200-300 and solid form for structural integrity, serves in advanced gas-cooled and but requires separate cooling to manage low thermal conductivity and potential radiolytic gas formation. Beryllium, with \xi \approx 0.209 and low absorption, provides compact moderation but is rarely used alone due to toxicity, expense, and (n, 2n) reactions increasing neutron economy challenges.
Material\xi (approx.)Moderating Ratio (approx.)Key Advantages/Disadvantages
Light Water0.9562Dual coolant/moderator; high absorption limits fuel choice.
Heavy Water0.51>7,000Low absorption enables ; high cost.
0.158200-300Structural, high temperature tolerance; slow moderation, needs cooling.
0.209High (~100+)Compact, low absorption; expensive, toxic.
Thermalization also involves diffusion of epithermal neutrons, described by the Fermi age \tau, which measures the mean-squared distance traveled during slowing down; low-mass moderators yield smaller \tau, reducing spatial separation between fast and thermal fluxes. In practice, impurities or reactor poisons can degrade performance by elevating \Sigma_a, necessitating purification and monitoring to sustain neutron economy.

Fast vs. Thermal Reactors

Thermal reactors utilize moderators like light water or to thermalize neutrons to energies around 0.025 eV, where the cross-section of peaks at approximately 580 barns, facilitating sustained chain reactions with low-enriched fuel dominated by the fissile isotope. This soft spectrum maximizes probability for but incurs neutron losses during moderation—up to 20-30% via without —and high parasitic capture in fertile resonances (1 eV to 100 eV range), limiting breeding and fuel efficiency to the natural 0.7% fissile content. Fast reactors, by contrast, dispense with moderators to preserve a hard neutron spectrum with average energies above 0.1 MeV (typically peaking near 0.1-1 MeV from the fission distribution), enabling fissions in both fissile materials like plutonium-239 and fertile isotopes such as uranium-238 above a ~1 MeV threshold, while minimizing absorptions in cladding and coolant due to lower fast-neutron cross-sections in those materials. Fission cross-sections drop to 1-2 barns for uranium-235 in this regime, necessitating higher initial fissile loading (>20% plutonium-239 or highly enriched uranium) to achieve criticality, as the eta factor (neutrons emitted per absorption) for plutonium-239 reaches ~2.7, supporting a positive neutron balance despite reduced per-fission probability. The hard spectrum underpins fast reactors' superior neutron economy for resource utilization, with breeding ratios often exceeding 1.0—up to 1.3 in sodium-cooled designs—by converting (99.3% of ) to via (n,γ) capture, yielding over times the extractable from compared to thermal systems' sub-unity ratios (<0.6 typically). Thermal reactors' conversion ratios remain below 1 due to spectrum-induced losses, confining them to once-through cycles without reprocessing. Operationally, fast reactors' physics demand compact cores for spectrum hardening and exhibit altered kinetics: delayed neutron fractions are similar (~0.3-0.4% effective), but prompt neutron lifetimes are shorter (~10^{-7} s vs. ~10^{-3} s thermal), accelerating transients, while feedback relies more on fuel expansion than Doppler broadening, which is weaker in fast fluxes. This enables actinide transmutation but requires precise control to avoid positive reactivity insertions from voids.

Reactor Kinetics and Control

Delayed Neutrons and Controllability

Delayed neutrons arise from the beta decay of certain fission products, known as neutron precursors, which are unstable isotopes produced during nuclear fission; these precursors decay with half-lives ranging from fractions of a second to about a minute, emitting neutrons with a total yield of approximately 0.016 neutrons per fission event in uranium-235. Unlike prompt neutrons, which are emitted directly and instantaneously during the fission process, delayed neutrons introduce a temporal lag in the neutron population growth, typically modeled in six precursor groups with distinct decay constants λ_i and fractional yields β_i, where the total delayed neutron fraction β = ∑ β_i ≈ 0.0065 (or 0.65%) for thermal neutron-induced fission of ^{235}U. This small fraction β is isotope-dependent and slightly varies with the neutron spectrum, being lower for plutonium-239 (β ≈ 0.0021) and higher for certain minor actinides. The prompt neutron lifetime, or mean generation time Λ, in thermal reactors is on the order of 10^{-4} seconds, representing the average time from neutron birth to subsequent ; without delayed neutrons, any supercritical reactivity insertion ρ > 0 would lead to exponential power growth governed solely by prompt neutrons, with a τ_p ≈ (Λ / ρ) ln(2), resulting in uncontrollable excursions for even small ρ (e.g., ρ = 0.001 yields τ_p ≈ 0.07 seconds). Delayed neutrons mitigate this by decoupling a portion β of the neutron production from the immediate chain, effectively requiring ρ > β for ; reactors are designed to operate with ρ << β (sub-prompt critical), where the reactor period τ ≈ β / ρ provides seconds to minutes for control systems—such as neutron-absorbing rods—to respond and stabilize power. This inherent delay enhances controllability, as evidenced by point kinetics equations incorporating precursor concentrations C_i: dn/dt = [(ρ - β)n / Λ] + ∑ λ_i C_i and dC_i/dt = (β_i n / Λ) - λ_i C_i, where the delayed terms dominate the initial response to positive reactivity. In practice, the effective delayed neutron fraction β_eff, which weights β_i by the importance of delayed neutrons (typically lower energy than prompt) in sustaining the chain reaction, is used in heterogeneous reactor designs; for light-water reactors fueled with enriched uranium, β_eff ≈ 0.0064–0.0075, decreasing slightly with fuel burnup due to isotopic shifts toward lower-β plutonium. Without delayed neutrons, reactor control would rely on impossibly fast mechanical interventions, rendering sustained chain reactions infeasible for power generation; their presence allows stable operation across a range of ρ values, with safety margins defined by the prompt critical threshold (k_eff ≈ 1 + β_eff). This physics underpins regulatory limits on maximum reactivity insertions, ensuring human or automatic systems can avert prompt supercriticality.

Reactor Kinetics Equations

The point kinetics equations provide a simplified model for the time-dependent neutron population and power level in a nuclear reactor, derived from the space- and energy-dependent neutron diffusion equations by assuming a separable form for the neutron flux, \phi(\mathbf{r}, E, t) = p(t) \psi(\mathbf{r}, E, t), where p(t) is the time-dependent amplitude and \psi is the slowly varying spatial-energy shape function approximated as the initial steady-state distribution. This point approximation neglects spatial flux distortions and is valid for transients where the core-averaged behavior dominates, such as control rod movements or temperature feedback effects on timescales longer than the prompt neutron lifetime but shorter than significant fuel depletion. The model incorporates six groups of delayed neutron precursors to capture the stabilizing effect of delayed neutrons, which originate from beta decay of fission fragments with half-lives ranging from seconds to minutes. The fundamental neutron balance equation is \frac{dn(t)}{dt} = \frac{\rho(t) - \beta}{\Lambda} n(t) + \sum_{i=1}^{6} \lambda_i C_i(t), where n(t) is the core-average neutron density, \rho(t) is the reactivity (a measure of deviation from criticality), \beta = \sum_{i=1}^{6} \beta_i is the total effective delayed neutron fraction (typically 0.0065 for ^{235}U-dominated fission and 0.0021 for ^{239}Pu, decreasing to 0.0055–0.0070 over a light-water reactor fuel cycle due to isotope evolution), \Lambda is the mean prompt neutron generation time (on the order of $10^{-5} to $10^{-4} s in thermal reactors), \lambda_i are the group decay constants, and C_i(t) are the precursor concentrations. The corresponding precursor equations are \frac{dC_i(t)}{dt} = \frac{\beta_i}{\Lambda} n(t) - \lambda_i C_i(t), \quad i = 1, \dots, 6. These form a system of seven coupled ordinary differential equations, often solved numerically for arbitrary \rho(t), with power proportional to n(t) under adiabatic assumptions. Reactivity enters as \rho(t) = \frac{k_{\text{eff}}(t) - 1}{k_{\text{eff}}(t)}, where k_{\text{eff}} is the effective multiplication factor; positive \rho drives supercritical growth, while the -\beta term reduces the prompt eigenvalue, ensuring controllability—for instance, \rho < \beta (about 0.65% for ^{235}U) yields slower excursions manageable by control systems. In the prompt approximation (neglecting precursors for very fast transients), the neutron equation simplifies to \frac{dn}{dt} = \frac{\rho}{\Lambda} n, yielding exponential behavior n(t) \propto e^{\alpha t} with \alpha = \rho / \Lambda > 0 for supercriticality, but this omits the dominant delayed contribution for \rho \ll \beta. Steady-state solutions for constant \rho satisfy the inhour equation, relating \rho to the inverse period $1/T = \alpha + \sum \frac{\beta_i \lambda_i}{\lambda_i - 1/T}, where T is the reactor period. The model's parameters vary with fuel composition and spectrum; for example, effective \beta decreases in mixed-oxide fueled reactors due to plutonium's lower yield.

Burnup, Poisons, and Depletion

Reactor Poisons: Short- and Long-Lived

Reactor poisons, also known as poisons, are isotopes, primarily products, that possess exceptionally high thermal cross-sections, thereby reducing the economy and reactivity by competing with for s. These poisons arise from the process and subsequent chains, accumulating in the core and necessitating reactivity compensation through control rods, burnable absorbers, or operational adjustments. Among them, short-lived poisons exhibit rapid buildup and decay, influencing transient behavior, while long-lived poisons persist, affecting steady-state and long-term operations. The principal short-lived reactor poison is (Xe-135), which forms via the of iodine-135 (I-135), a direct product with a cumulative yield of approximately 6% in U-235 . I-135 has a of 6.57 hours, decaying to Xe-135, which itself has a of 9.14 hours and a neutron cross-section of about 2.6 × 10^6 barns. During steady-state at high power, Xe-135 reaches an equilibrium concentration where its production rate balances removal via decay and neutron-induced to Xe-136; however, this equilibrium yields a significant reactivity penalty, equivalent to roughly -2,800 pcm (percent milli-k) at full power in typical light-water reactors. Xe-135 poisoning manifests prominently in transients, such as reactor shutdowns or power reductions, where drops abruptly, halting burnup of Xe-135 while I-135 continues, causing a post-shutdown peak in Xe-135 concentration after about 10-12 hours. This "xenon buildup" can depress reactivity sufficiently to prevent restart for up to 24-48 hours or longer, depending on initial power level and , a phenomenon observed and modeled in reactor physics simulations. Power increases can also induce spatial xenon oscillations due to tilting, requiring careful control to avoid instability, as the poison's high cross-section amplifies local reactivity variations. In contrast, long-lived poisons like samarium-149 (Sm-149) accumulate irreversibly due to their stability, originating from the decay of promethium-149 (Pm-149), which has a of 53 hours and arises from direct with a yield of about 1.4%. Sm-149 possesses a neutron absorption cross-section of 42,000 barns and does not decay radioactively, leading to a gradual buildup that reaches equilibrium only after prolonged operation, typically weeks to months. Unlike Xe-135, Sm-149 transients are minimal because Pm-149's longer buffers rapid changes, but its persistence contributes to a cumulative reactivity loss of around -1,000 to -2,000 pcm over a fuel cycle, influencing strategies and requiring initial core loading with excess reactivity or burnable poisons. The differential impacts of these poisons underscore the need for designs to incorporate margins for effects; for instance, Xe-135 dominates short-term challenges, while Sm-149 drives long-term depletion considerations, with both modeled via point kinetics equations extended for poison concentrations to predict k_eff evolution. Other minor poisons exist, but Xe-135 and Sm-149 account for the majority of fission product reactivity holdback in thermal reactors.

Fuel Burnup and Reactivity Evolution

Fuel burnup quantifies the extent of nuclear fuel consumption, expressed as the released per unit mass of initial heavy metal atoms, commonly in gigawatt-days per metric (GWd/t). In pressurized water reactors (PWRs), typical discharge burnups range from 40 to 60 GWd/t, reflecting the progressive of fissile isotopes and subsequent isotopic changes in the fuel. During irradiation, the fissile (U-235) content depletes from an initial enrichment of 3-5 wt% to below 1 wt% at high , directly reducing the probability of events and thus the neutron multiplication factor. (Pu-239) accumulates through on U-238, reaching approximately 0.8-1 wt% of by 50 GWd/t, partially offsetting the U-235 loss due to its higher thermal cross-section (around 750 barns versus 580 barns for U-235). However, net fissile inventory declines, as Pu-240 and higher isotopes build up with parasitic absorption. Fission products accumulate as a , with neutron-absorbing isotopes like samarium-149 (Sm-149, capture cross-section ~40,000 barns) and rare earth elements contributing a reactivity penalty equivalent to several percent Δk over a typical . Short-lived absorbers such as transiently exacerbate this, but long-term effects dominate -driven evolution. The infinite multiplication factor (k∞) may rise modestly in early stages from Pu buildup and spectral hardening but decreases overall, necessitating design provisions for reactivity hold-down. Reactivity evolution manifests as a gradual decrease in effective multiplication factor (k_eff), compensated by initial excess reactivity of several percent Δk/k, controlled via soluble in the and burnable absorbers like gadolinium-155/157 (cross-sections exceeding 60,000 barns). In PWRs, boron concentration typically drops by about 100 per GWd/MTU to maintain criticality, from ~1000-1200 at beginning-of-cycle (BOC) to near-zero at end-of-cycle (EOC). Burnable absorbers deplete through (n,γ) reactions, releasing positive reactivity worth over the cycle, while control rods provide fine adjustments. This dynamic ensures stable power operation but requires precise modeling to avoid positive voids or peaking factor excursions at high .

Fuel Cycle Physics

Uranium Enrichment Physics

Uranium enrichment physically separates the fissile isotope (²³⁵U) from the more abundant (²³⁸U) to increase the ²³⁵U concentration beyond its natural abundance of 0.711% atomic fraction in , enabling sustained chain reactions in most commercial reactors that rely on thermal neutrons. Light water reactors, the dominant type, require low-enriched (LEU) with 3-5% ²³⁵U, as natural 's low fissile content insufficiently sustains criticality without graphite moderation or blankets. The mass difference—²³⁵U at 235.044 u and ²³⁸U at 238.051 u—drives separation, with ²³⁸U comprising 99.274% of and acting primarily as a fertile but non-fissile absorber in thermal spectra. The fundamental physics exploits isotopic mass disparities in (UF₆) gas, where molecular weights differ by ~0.86%: 349 u for ²³⁵UF₆ versus 352 u for ²³⁸UF₆. Separation factors (α), defined as the ratio of enriched-to-depleted ratios post-stage, are small (near 1), necessitating multi-stage cascades to achieve target assays while minimizing tails (, typically 0.2-0.3% ²³⁵U). The separative work unit (SWU) quantifies the thermodynamic effort, derived from the value function V(x) = (2x - 1) ln[x/(1 - x)], where x is the ²³⁵U ; one SWU equals the work to produce unit masses of product, feed, and tails with V(P) + V(W) - 2V(F) = 1 kg-SWU, reflecting entropy-of-mixing costs rather than direct energy. Producing 1 kg of 3.6% LEU from 0.711% feed with 0.25% tails requires ~7.9 SWU. Gas centrifugation, the dominant method since the 1980s, rotates UF₆ at 50,000-90,000 rpm in cylindrical rotors ~1-2 m long, generating radial accelerations of 10⁵-10⁶ g to impose a centrifugal potential μ(r) ≈ (1/2) m ω² r², where m is molecular mass, ω angular velocity, and r radius. Heavier ²³⁸UF₆ migrates outward via this force, while lighter ²³⁵UF₆ concentrates axially or inward; countercurrent axial flow, driven by thermal transpiration or scoops, amplifies separation beyond equilibrium diffusion. Single-stage α ≈ 1.05-1.2, far exceeding gaseous diffusion's ~1.004, due to integrated radial gradients and flow dynamics, though cascades of 10-20 stages per machine and thousands total are needed for LEU. Enrichment efficiency scales with rotor speed squared and mass ratio, but viscosity, Coriolis effects, and supersonic flows limit practical α. Historical , phased out for energy intensity, forced UF₆ through semi-porous barriers under pressure gradients, leveraging Graham's effusion law where lighter molecules traverse pores faster, yielding α = √(M₂₃₈F₆ / M₂₃⁵F₆) ≈ 1.0043 ideally, reduced to ~1.002-1.003 in practice by back-diffusion and non-ideal pores. Stages numbered ~4,000 for LEU, with each consuming ~2,500 kWh/SWU versus centrifuges' ~50 kWh/SWU, highlighting diffusion's inefficiency from repeated pressure drops across membranes. methods, like molecular photoexcitation, selectively ionize or dissociate ²³⁵UF₆ via tuned wavelengths matching vibrational differences, but remain developmental due to scaling challenges. Enrichment physics underscores risks, as identical processes yield highly (>20% ²³⁵U) for weapons with fewer SWU but heightened safeguards needs.

Advanced Fuel Isotopes and Breeding

Advanced fuel isotopes in nuclear reactors extend beyond the primary fissile material to include and , which enable enhanced fuel utilization through processes. , produced from via followed by decays (U-238(n,γ)→U-239→Np-239→Pu-239), exhibits a high fission cross-section for both and fast neutrons, making it suitable for mixed-oxide (MOX) fuels in light-water reactors or as driver fuel in breeders. , derived from through Th-232(n,γ)→Th-233→Pa-233→, offers a favorable neutron economy in spectra due to its low parasitic in , potentially supporting self-sustaining cycles with minimal initial fissile inventory. These isotopes address resource limitations by leveraging abundant fertile materials, but their risks and from precursors like Pa-233 necessitate careful isotopic management. Breeding refers to the transmutation of fertile isotopes into fissile ones at a rate exceeding consumption, quantified by the breeding ratio (BR), defined as the number of fissile atoms produced per fissile atom destroyed. Achieving BR > 1 requires a neutron excess from fission (typically 2.4-2.9 neutrons per fission event) to overcome parasitic captures on structure, coolant, and fission products while sustaining the chain reaction and enabling net production. In thermal reactors, BR typically falls below 1 (conversion ratio ≈0.6 for uranium-plutonium), as soft neutron spectra promote captures over fissions in fertile isotopes; fast spectra in breeder designs harden the flux, reducing (n,γ) losses and boosting fission probabilities for Pu-239 production from U-238. The process hinges on precise neutron economy: for U-Pu breeding, two of the ~2.5 excess neutrons per fission must convert U-238 without undue loss, yielding BR values of 1.1-1.3 in optimized fast oxide or metal-fueled cores. The uranium-plutonium cycle dominates fast breeder physics, where depleted uranium (mostly U-238) blankets surround fissile cores to capture leakage neutrons, producing Pu-239 with yields approaching 1.2 fissile atoms per consumed in equilibrium. Parasitic effects, such as Pu-240 buildup from secondary captures, degrade BR over time by increasing non-fissile content and spontaneous fission sources, necessitating reprocessing to separate Pu isotopes. In contrast, the thorium-uranium cycle leverages Th-232's lower capture-to-fission ratio in produced U-233, enabling thermal breeding in heavy-water or molten-salt configurations, though protactinium-233's 27-day half-life allows neutron capture losses if not chemically separated. Experimental data from Shippingport (1977-1982) demonstrated thermal Th-U breeding with BR ≈1.07 using seed-blanket designs, confirming viability but highlighting Pa-233 removal challenges for higher gains. Fast thorium cycles achieve BR >1 more readily, as elevated energies minimize threshold reactions, but proliferation concerns from pure U-233 (lacking even Pu-240 discriminants) persist. Minor actinides like neptunium-237 and represent advanced fuels for , incinerating long-lived waste via fast-spectrum fissions, though their high absorption cross-sections demand BR-compensating designs to avoid reactivity penalties. Overall, physics underscores causal trade-offs: while enabling indefinite fuel extension from (0.7% U-235), it amplifies radiotoxicity from higher actinides and requires spectrum control for multiplication exceeding unity. Demonstrated in prototypes like Russia's BN-800 (operational since , BR ≈1.1), these cycles pivot on empirical cross-section , with fast fluxes yielding 20-30% higher than thermal due to reduced captures.

Natural and Historical Physics Examples

Oklo Natural Fission Reactor

The natural reactors consist of sixteen zones within the deposits of the Oklo mine in , , where self-sustaining nuclear chain reactions occurred approximately 1.7 billion years ago. These reactions were enabled by geological conditions that concentrated in layers saturated with , which served as a moderator to slow s for in (U-235). At that epoch, contained about 3% U-235, compared to 0.72% today, due to the differential decay rates of U-235 ( 704 million years) and U-238 ( 4.47 billion years), providing sufficient for criticality without enrichment. Each zone operated at low thermal power levels of around 20 kilowatts, with reactions proceeding in pulsed cycles: moderation initiated , generating heat that boiled the water, reducing neutron moderation and halting the chain reaction until the zone cooled and refilled. The reactors were discovered in May 1972 when routine analysis of shipments from the mine revealed anomalously low U-235 concentrations, averaging 0.44% in affected samples—nearly 40% below the natural abundance—prompting investigations that confirmed reactions. This finding validated theoretical predictions by geochemist Paul K. Kuroda, who in 1956 proposed that natural reactors could form under prehistoric conditions with higher U-235 fractions and suitable . Post-discovery studies identified reaction zones as slab-like ore bodies, typically 10-20 meters thick and hundreds of meters long, with enriched in U-238 post- and depleted in lighter isotopes, alongside migration patterns of fission products over geological timescales. Evidence of fission includes the isotopic signatures of like xenon-129, produced in ratios matching thermal fission spectra, and the presence of rare earth elements such as and in anomalous abundances consistent with and chains. products like cesium and isotopes were partially retained in the matrix, demonstrating natural confinement over billions of years, though some migrated via , informing models of long-term behavior. These reactors operated without human intervention, relying on inherent physics: positive reactivity from fissile concentration balanced by from of resonances and loss of moderator density, preventing runaway excursions. In reactor physics terms, exemplifies the prerequisites for criticality—neutron economy governed by the effective multiplication factor k_{eff} > 1—achieved through ore geometry minimizing leakage, moderation by hydrogen in water, and minimal neutron absorption by impurities. The site's total energy output is estimated at around 100 megawatt-years per zone, with levels comparable to modern light-water reactors, underscoring that natural systems can sustain controlled fission via self-regulating thermal-hydraulic mechanisms. No similar reactors have operated since, as contemporary U-235 levels preclude criticality in water-moderated without deliberate enrichment or reflector designs.

Safety and Misconception Physics

Inherent Physical Safety Mechanisms

In nuclear reactors, inherent physical safety mechanisms arise from fundamental neutronics and thermodynamic behaviors that automatically counteract power excursions, reducing reactivity without active intervention or external . These include negative reactivity feedback coefficients tied to and changes, which ensure self-stabilization by decreasing the effective multiplication factor k_{eff} as conditions deviate from . Such mechanisms are intrinsic to the core's material properties and neutron spectrum, providing prompt responses on sub-second to minute timescales. The in fuel contributes a strongly negative fuel temperature coefficient, typically on the order of -1 to -5 pcm/°C in light-water reactor fuels. As fuel temperature increases, thermal motion broadens the neutron absorption resonances of U-238 (and other even-mass isotopes), enhancing parasitic neutron capture at the expense of fission in U-235 or Pu-239; this resonance shift occurs rapidly due to atomic vibrations, inserting negative reactivity proportional to the square root of temperature change. In pressurized water reactors (PWRs), this feedback alone can compensate for reactivity insertions up to several dollars, preventing supercritical transients. Moderator temperature and void effects provide additional in thermal-spectrum reactors. Elevated moderator temperature reduces its density, hardening the neutron spectrum by diminishing thermalization efficiency; this favors over , yielding a of -10 to -30 pcm/°C in boiling water reactors (BWRs). Similarly, the —negative in most light-water designs at -50 to -200 pcm/%void—arises as steam bubbles form, further reducing moderation and increasing leakage; for instance, in PWRs, void formation during a loss-of-coolant event inherently suppresses sustainability. These effects contrast with designs like the , where positive void coefficients (due to moderation and low-enriched fuel) amplified the 1986 Chernobyl excursion, underscoring the causal role of spectrum-dependent feedback in accident physics. Fuel and core expansion offer slower but stabilizing geometric feedbacks. Thermal expansion of fuel pellets decreases fission density, while clad and assembly dilation increases neutron leakage; in sodium-cooled fast reactors, axial fuel expansion can yield negative coefficients exceeding -0.5 pcm/°C, though spectrum hardening may introduce positive components at low power, necessitating design mitigation for overall negativity. Collectively, these mechanisms ensure \rho < 0 for temperature rises, with total core coefficients often designed below -20 pcm/°C to bound excursions within delayed neutron fractions (\beta \approx 0.0065 in thermal fuels), as verified in operational data from reactors like the Shippingport PWR, where feedbacks limited transients to benign levels.

Physics of Reactor Accidents

Nuclear reactor accidents fundamentally involve disruptions in the balance of neutron production and loss, or in heat removal from fission and decay processes, potentially leading to core damage or radionuclide release. In terms of neutronics, accidents can stem from reactivity excursions where the effective multiplication factor k_{eff} > 1, resulting in \rho = \frac{k_{eff} - 1}{k_{eff}} > 0, causing exponential power growth governed by the point kinetics equations: \frac{dP}{dt} = \frac{\rho - \beta}{\Lambda} P + \sum_i \lambda_i C_i, where \beta \approx 0.0065 is the delayed neutron fraction, \Lambda is the prompt neutron lifetime (microseconds in thermal reactors), and C_i are delayed neutron precursors. If \rho > \beta, prompt criticality ensues, yielding power doubling times under a second, overwhelming thermal feedback and control systems. Thermal-hydraulic failures, such as , arise when or diminishes, impairing from clad to the secondary side. Even post-scram (fission chain reaction halted by control rods), from fission product chains—primarily short-lived isotopes like (t_{1/2} = 8 days) and cesium-137 (t_{1/2} = 30 years)—persists at 6-7% of full-power rating immediately after shutdown, decaying to ~1.5% after 1 hour and ~0.4% after 1 day via and gamma emissions. Without cooling, clad temperatures exceed 1200°C, triggering zirconium-water oxidation: \ce{Zr + 2H2O -> ZrO2 + 2H2}, generating explosive hydrogen and potential clad breach, followed by melting at ~2800°C if core geometry permits. In the 1986 RBMK-1000 accident, a positive —arising from and 's role as both coolant and weak absorber—amplified reactivity during a low-power test: voids displaced , reducing parasitic more than loss, yielding \Delta \rho \approx +4.7\beta initially, escalating to supercriticality (\rho > \beta) as power surged from 200 MWt to over 30,000 MWt in seconds, vaporizing fuel and destroying . This design-specific feedback, absent in - reactors with negative , underscores causal role of core physics in excursion severity. Contrastingly, the 1979 Three Mile Island Unit 2 partial meltdown involved no reactivity excursion but a LOCA variant: a stuck-open pressurizer depleted primary , reducing core flow and causing departure from (DNB), with elevating clad temperatures to ~45% of the , though prevented ; operator misdiagnosis delayed high-pressure injection, but negative reactivity feedbacks (, \Delta \rho \approx -2\% per 1000°C) limited power rise. The 2011 Fukushima Daiichi event exemplified prolonged accumulation sans fission restart: post-earthquake , tsunami-induced station blackout halted recirculation pumps, uncovering in Units 1-3 over hours; integrated to ~10-20% of initial energy inventory boiled residual water, yielding oxidation and buildup, with explosions breaching containments but no recriticality due to flooding and disruption. Physics models confirm core melting initiated ~4-6 hours post-loss of , driven solely by conductive/convective limits in degraded .

References

  1. [1]
    Reactor Physics | Definition & Applications | nuclear-power.com
    Nuclear reactor physics is the field of physics that studies and deals with the applied study and engineering applications of neutron diffusion and fission ...Reactor Kinetics · Reactor Control · Reactor Criticality · Nuclear Safety
  2. [2]
    [PDF] DOE Fundamentals Handbook Nuclear Physics and Reactor Theory ...
    Another basic concept of matter that the Greeks debated was whether matter was continuous or discrete. That is, whether matter could be continuously divided and ...
  3. [3]
    Physics of Uranium and Nuclear Energy
    May 16, 2025 · Nuclear reactors work by containing and controlling the physical process of nuclear fission. · Radioactive decay of both fission products and ...
  4. [4]
    Fundamentals of Nuclear Reactor Physics - ScienceDirect.com
    This book covers how fission reactors work, reactor design, safe operation, atomic physics, neutron reactions, kinetics, and spatial distribution.
  5. [5]
    The Fission Process - MIT Nuclear Reactor Laboratory
    Each time a U-235 nucleus splits, it releases two or three neutrons. Hence, the possibility exists for creating a chain reaction. Nuclear fission. The MIT ...
  6. [6]
    Uranium-235 Chain Reaction - HyperPhysics Concepts
    A U-235 chain reaction occurs when a neutron from fission strikes another nucleus, initiating fission. If at least one neutron does, the chain continues. ...
  7. [7]
    Fission and nuclear reactors - Physics
    When a thermal neutron hits a U-235 nucleus and is absorbed, the nucleus generally splits into two big pieces and a few neutrons. Two of the possible reactions ...
  8. [8]
    Nuclear Fission Energy
    Aug 9, 2000 · One fission event results in the release of about 200 MeV of energy, or about 3.2 ´ 10-11 watt-seconds. Thus, 3.1 ´ 1010 fissions per second ...
  9. [9]
    Nuclear Fission Fragments - HyperPhysics
    Fission Fragment Example​​ A common pair of fragments from uranium-235 fission is xenon and strontium: Highly radioactive, the xenon decays with a half-life of ...
  10. [10]
    Fission Chain Reaction - Chemistry LibreTexts
    Jan 29, 2023 · Nuclear chain reactions are a simple, yet powerful method which to produce both constructive and destructive forces.Introduction · Chain Reaction · Understanding Fission · Controlled Chain Reactions
  11. [11]
    Nuclear Chain Reactions - Atomic Archive
    A chain reaction refers to a process in which neutrons released in fission produce an additional fission in at least one further nucleus.
  12. [12]
    NUCLEAR 101: How Does a Nuclear Reactor Work?
    The main job of a reactor is to house and control nuclear fission—a process where atoms split and release energy. Fission vs Fusion with multi-colored ...
  13. [13]
    Control of Fission Chain Reactions - HyperPhysics Concepts
    In the operation of a nuclear reactor, fuel assemblies are put into place and then the control rods are slowly lifted until a chain reaction can just be ...
  14. [14]
  15. [15]
    [PDF] Module 3: Neutron Induced Reactions Dr. John H. Bickel
    Neutron reaction cross sections. • Total microscopic neutron cross section is expressed as: σ = dN/dt / [(I/A) n A ∆x] • Defining neutron flux as: φ = I/A ( ...
  16. [16]
    [PDF] Interaction of Neutrons with Matter.
    Apr 2, 2011 · They slow down due to scattering (referred to as moderation) until they reach thermal energies. Finally, they are absorbed by a target nucleus.
  17. [17]
    Cross-Sections - DoITPoMS
    Inelastic scattering (n,n'): a neutron is briefly absorbed by a nucleus, leaving it in an excited state. The nucleus can later return to its ground state, ...<|separator|>
  18. [18]
    [PDF] ENDF- 103 (BAW - 3 16) May 1967 U-238 NEUTRON CROSS ...
    The data in section 2 are further subdivided into a low-energy range (from 0.00 1 to 5 eV) , a resolved- resonance energy range (from 5 to 3920 eV), and an ...
  19. [19]
    Neutron Cross Section - an overview | ScienceDirect Topics
    Resonances occur when the energy of an incoming neutron matches the energy difference ΔE of two discrete energy states of the nucleus, the neutron is absorbed ...
  20. [20]
    Neutron Cross-section Standards
    Neutron cross-section standards are important for measuring and evaluating other neutron reactions, and most cross sections are measured relative to them.Missing: ORNL | Show results with:ORNL
  21. [21]
    ENDF/B-VIII.0: The 8th Major Release of the Nuclear Reaction Data ...
    ENDF/B-VIII.0 includes new IAEA standards, improved thermal neutron scattering, CIELO project data, and updated data for light nuclei, structural materials, ...
  22. [22]
    ENDF/B-VII.1 Evaluated Nuclear Data Library
    It contains evaluated cross sections for 163 materials (all isotopes) mostly up to 140 MeV. The sublibrary was supplied by Los Alamos National Laboratory (LANL) ...
  23. [23]
  24. [24]
    [PDF] DOE Fundamentals Handbook Nuclear Physics and Reactor Theory ...
    A measure of the increase or decrease in neutron flux in an infinite reactor is the infinite multiplication factor, k∞. The infinite multiplication factor is ...
  25. [25]
    Calculating Criticality | Los Alamos National Laboratory
    Nov 1, 2023 · In nuclear reactor theory, the neutron multiplication factor k is a key ratio that represents the average number of neutrons from one fission ...
  26. [26]
    None
    Summary of each segment:
  27. [27]
    [PDF] Subcritical Multiplication and Reactor Startup - DSpace@MIT
    Nuclear fission reactors operate by maintaining a precise neutron balance. The reactor is in a critical condition if the number of neutrons created by the ...
  28. [28]
    Features of a subcritical nuclear reactor - ScienceDirect.com
    A subcritical nuclear reactor is a device where the nuclear-fission chain reaction is initiated and maintained using an external neutron source.
  29. [29]
    [PDF] Module 5: Neutron Thermalization Dr. John H. Bickel
    • Slowing down fission emitted neutrons to thermal energies increases fission reaction rate. • Principle way to slow down neutrons is via collisions. Page 6. 6.
  30. [30]
    Neutron Thermalization - Stanford University
    Mar 6, 2024 · You have to find some way to slow the neutrons down to increase the chance of fission. This process of slowing down neutrons is known as neutron thermalization.
  31. [31]
    Moderators - DoITPoMS
    A moderator is designed to slow down fast neutrons such that they are more easily absorbed by fissile nuclei. There are two main factors in choosing a moderator ...
  32. [32]
    Neutron Energy - an overview | ScienceDirect Topics
    eV by elastic scattering for hydrogen is 27 number of collisions, whereas carbon requires 119 collisions and Uranium requires about 2175 collisions (Rinard, ...
  33. [33]
    Moderating Ratio - MR | nuclear-power.com
    The moderating ratio or moderator quality is the complete measure of a moderator's effectiveness because it also takes into account the absorption effects.
  34. [34]
    [PDF] A comparative study of neutron moderation characteristics via direct ...
    Commonly used moderators include light water (H₂O), heavy water. (D₂O), and graphite, each offering unique advantages and limitations depending on the reactor.<|control11|><|separator|>
  35. [35]
    Moderator Material - an overview | ScienceDirect Topics
    Next to heavy water, graphite has the highest moderating ratio and is extensively used as moderator and reflector for gas cooled thermal reactors. The world's ...
  36. [36]
    [PDF] Advanced Moderator Material Handbook - Idaho National Laboratory
    Sep 30, 2020 · Moderators are used in nuclear reactors to thermalize, or slow down, neutrons so they may more efficiently participate in fission reactions in ...
  37. [37]
    [PDF] Status of Fast Reactor Research and Technology Development
    1. Liquid metal fast breeder reactors – Safety measures. 2. Nuclear power plants – Design and construction. 3. Nuclear reactors – Decommissioning.
  38. [38]
    [PDF] 1.3-Fast Reactor Physics - 1 Spectrum and Cross Sections
    Mar 26, 2019 · – Impacts of fast energy spectrum. • Fission-to-capture, enrichment, etc. – Comparison to LWR neutron physics. ▫ Nuclear Data and Validation. – ...
  39. [39]
    Slow and fast neutrons - THE PHENOMENON - radioactivity.eu.com
    The use of fast neutrons in so-called 'fast reactors' allows for the production of more fissile nuclei than are destroyed, as the absorption of at least one ...
  40. [40]
    Fast Neutron Reactors - World Nuclear Association
    Aug 26, 2021 · About 20 fast neutron reactors (FNR) have already been operating, some since the 1950s, and some supplying electricity commercially.
  41. [41]
    [PDF] AN EVALUATION OF THE FAST MIXED SPECTRUM REACTOR
    In this hard spectrum the value of eta (n) for Pu-239 is high enough (-2.7) to maintain the chain reaction and to breed. On the other hand, the fast fission.
  42. [42]
    Fast reactors - International Atomic Energy Agency
    The fast neutron spectrum allows fast reactors to largely increase the energy yield from natural uranium as compared to thermal reactors.
  43. [43]
    Shrinking nuclear waste and increasing efficiency for a sustainable ...
    Fast reactors can also produce or 'breed' more fuel than they consume and burn off some of the waste contained in spent fuel, such as minor actinides, which ...
  44. [44]
    [PDF] RADIATION DETECTION FROM FISSION - INFO
    For example, there are ~2.5 prompt neutrons emitted in 235U fission and. 0.016 delayed neutrons. For gamma rays there are ~7 prompt, and ~7 delayed gamma rays.
  45. [45]
    [PDF] DELAYED FISSION NEUTRONS - IAEA-NDS
    Decay curve of non-halogen delayed neutron precursors in fission of U by thermal neutrons. ... In 235U thermal systems this fraction is often near 0. 0065 (the ...
  46. [46]
    [PDF] Week 3: Controlling a Nuclear Reactor - MIT ESP
    Delayed neutrons are a tiny fraction (for U-235 fission, about 0.65%) of neutrons produced, but they are very important as they make controlling a nuclear chain ...<|separator|>
  47. [47]
    [PDF] Principles of Fission Nuclear Power - OSTI.GOV
    Fraction of delayed neutrons varies with isotope and (slightly) with neutron spectrum. ▫. Average delayed neutron energy is significantly less than average.
  48. [48]
    Prompt Neutron Lifetime - PNL | nuclear-power.com
    The typical prompt neutron lifetime in thermal reactors is on the order of 10−4 seconds. Generally, the longer neutron lifetimes occur in systems in which the ...
  49. [49]
    [PDF] -1- TOPIC: 292003 KNOWLEDGE: K1.01 [2.9/3.0] QID: B124 A ...
    The delayed neutron fraction for U-235 will decrease. B. The delayed neutron ... Delayed neutrons make up a larger fraction of fission neutrons than prompt ...
  50. [50]
    [PDF] Reactor Kinetics and Control - MIT OpenCourseWare
    Reactor Kinetics and Control. Page 2. Prof. Andrew C. Kadak, 2008. Page 2 ... Delayed neutrons are produced at about ½ the energy of prompt neutrons.
  51. [51]
    [PDF] Measurement of effective delayed neutron fraction for nasa zero ...
    Actual delayed neutron fraction for uranium 235. 0.0064. integral by the ... thermal power fraction ratio of thermal fission rate to total fission rate.
  52. [52]
    Delayed Neutrons | Definition & Characteristics | nuclear-power.com
    Compared with fission neutrons, which make a self-sustaining chain reaction possible, delayed neutrons make reactor control possible, and photoneutrons are ...
  53. [53]
    [PDF] MEASURED AND CALCULATED EFFECTIVE DELAYED ... - OSTI.gov
    In general, delayed neutrons are more effective to control the reactor than prompt neutrons because they are born at lower energy compared to prompt (fission) ...
  54. [54]
    None
    ### Extracted Point Kinetics Equations
  55. [55]
    [PDF] Modeling of Reactor Kinetics and Dynamics - INL Digital Library
    of prompt and delayed neutrons. Because delayed neutrons are born with less ... Due to the feedback mechanisms and nonlinearities present in the reactor kinetics ...
  56. [56]
    [PDF] ELEMENTARY PHYSICS OF REACTOR CONTROL MODULE D
    • two fission products, Xenon-135 and Samarium-149 have particular importance in terms of reactor operation and control: these two nuclei have very large ...
  57. [57]
    [PDF] Westinghouse Technology 2.1 Reactor Physics Review.
    Effective Multiplication Factor. K effective is a measure of the change in fission neutron population from one neutron generation to the next, or of fission ...
  58. [58]
    Samarium 149 | Production & Equilibrium | nuclear-power.com
    Samarium 149 is an indirect product of U-235 fission and has a very large neutron capture cross-section (about 42,000 barns). It does not decay since it is ...
  59. [59]
    [PDF] 0518 - R304B - GE BWR_4 Technology - 1.7 Reactor Physics.
    The principle of operating a nuclear reactor is based on neutron economy, which is the accounting of the number of neutrons between successive generations.
  60. [60]
    [PDF] The physics of nuclear fission reactors with a focus on the effect of ...
    The two fission products Xenon-135 and Samarium-149 neutron poisons have a substantial effect on the power reactor operation and design. This is because ...Missing: lived
  61. [61]
    [PDF] Very High Burn-ups in Light Water Reactors - Nuclear Energy Agency
    The reasons for this gradual evolution of burn-ups are first that the levelised fuel cycle cost unambiguously decreases with increasing discharge burn-up for ...<|separator|>
  62. [62]
    [PDF] Impact of extended burnup - on the nuclear fuel cycle
    Increase of burnup reflects on the requirements for natural uranium, enrichment and fuel fabrication, reactor core configuration and its control, fuel.
  63. [63]
    [PDF] Tier 2 Chapter 04 - Reactor - Section 4.3 Nuclear Design
    This excess reactivity depletes with fuel burnup, but early in the cycle it is controlled by the addition of soluble boron to the coolant and burnable absorber ...
  64. [64]
    Nuclear Fuel Facts: Uranium | Department of Energy
    Natural Uranium – contains a 238U concentration of 99.27 percent, 235U concentration of 0.711 percent and very little 234U. · Low Enriched Uranium – contains a ...
  65. [65]
    Uranium Enrichment - World Nuclear Association
    Jun 6, 2025 · It involves forcing uranium hexafluoride gas under pressure through a series of porous membranes or diaphragms.
  66. [66]
    [PDF] 0559 - F204S - Uranium Enrichment Processes - Gaseous Diffusion.
    The separation factor represents the enrichment per stage. Figure 2-12. Typical Cascade Schematic. Page 28. Module 2.0: Gaseous Diffusion.
  67. [67]
    [PDF] 0559 - F204S - Uranium Enrichment Processes - 04 - Gas Centrifuge.
    The centrifuge separation process uses the principle of centrifugal force to create a density gradient in gaseous uranium hexafluoride. (UF6) that contains ...Missing: physics | Show results with:physics
  68. [68]
    [PDF] Characteristics of the Gas Centrifuge for Uranium Enrichment and ...
    An excellent derivation of physical principles of the gas centrifuge for uranium enrichment available in the open literature is the one developed by Ernst Rätz.
  69. [69]
    Uranium Enrichment | Nuclear Regulatory Commission
    Laser Separation​​ Uranium can be enriched by separating isotopes of uranium with lasers. Molecules can be excited by laser light; this is called photoexcitation ...
  70. [70]
    [PDF] Uranium Enrichment and Nuclear Weapon Proliferation - SIPRI
    Table 1.2 Uranium enrichment techniques, according to physical principles and mechanisms used. Property of. Physical isotope principle. Physical mechanism.<|separator|>
  71. [71]
    [PDF] Thorium fuel cycle — Potential benefits and challenges
    232Th and 233U are the best 'fertile' and 'fissile' materials respectively for thermal neutron reactors and. 'thermal breeding' has been demonstrated for (Th, U) ...
  72. [72]
    Thorium - World Nuclear Association
    May 2, 2024 · In the closed fuel cycle, the driver fuel required for starting off is progressively replaced with recycled U-233, so that on reaching ...
  73. [73]
    Conversion Factor - Breeding Ratio - Nuclear Power
    The conversion factor is defined as the ratio of fissile material created to fissile material consumed, and it is often referred to as the breeding ratio.
  74. [74]
    [PDF] BREEDING RATIO AND DOUBLING TIME CHARACTERISTICS OF ...
    The Clinch River Breeder Reactor, per the current design, will have a breeding ratio of 1.20 at the start-of-life and 1.23 at the end of the first.
  75. [75]
    [PDF] Advanced fuels with reduced actinide generation
    The scope of the meeting included advanced fuels for all types of nuclear reactors: light water reactors, heavy water reactors, high temperature reactors, fast ...
  76. [76]
    Meet Oklo, the Earth's Two-billion-year-old only Known Natural ...
    the rock was evidence of natural fission that occurred over two billion years ago. ... “Like in a man- ...
  77. [77]
    Nature's Nuclear Reactors: The 2-Billion-Year-Old Natural Fission ...
    Jul 13, 2011 · Unfortunately for science, the sixteen natural nuclear reactors at Oklo have been destroyed, completely mined out for their rich uranium ore.Missing: facts | Show results with:facts
  78. [78]
    The Cosmic Origins of Uranium - World Nuclear Association
    May 16, 2025 · Each operated at about 20 kW thermal. At that time the concentration of U-235 in all natural uranium was 3.7 percent instead of 0.7 percent as ...
  79. [79]
    Meet Oklo, the Earth's 2 billion-year-old natural nuclear reactor
    Scientists have determined that Oklo's uranium deposits consisted of approximately 3% U235 billions of years ago, the same enrichment found in today's ...
  80. [80]
    Evidence of Natural Fission at Oklo - Stanford
    Mar 16, 2024 · The main evidence of natural fission occurring at Oklo is the depletion of U-235, and the anomalous abundance of fission decay products, particularly the rare ...
  81. [81]
    Discovery of fissionogenic Cs and Ba capture five years after Oklo ...
    Aug 13, 2018 · The Oklo natural nuclear reactors provide a wealth of information regarding the migration and retention of fission products in nuclear ...
  82. [82]
    Reactivity Coefficients - Reactivity Feedbacks | nuclear-power.com
    A major factor in the prompt negative temperature coefficient for the TRIGA cores is the core spectrum hardening that occurs as the fuel temperature increases.
  83. [83]
    Safety of Nuclear Power Reactors
    Feb 11, 2025 · The main safety features of most reactors are inherent – negative temperature coefficient and negative void coefficient. The first means ...
  84. [84]
    Fuel Temperature Coefficient - Doppler Coefficient - Nuclear Power
    The Doppler broadening of resonances is a very important phenomenon, which improves reactor stability. Assume, that the Doppler coefficient in a reactor is -3 ...
  85. [85]
    Improving prompt temperature feedback by stimulating Doppler ...
    Doppler feedback is a strong feedback able to counteract reactivity insertions of many dollars as shown later in this document.
  86. [86]
    [PDF] Module 10 - Power Reactor Feedback Effects Rev 01.
    Doppler and moderator density reactivity feedback. Start with simplified ... • Doppler reactivity defect is independent of reactor type. • Recall our ...
  87. [87]
    [PDF] Safety and Security Defense-in- Depth for Nuclear Power Plants
    Inherent safety refers to characteristics of the reactor design that are physically intrinsic to the design. For example, PWR and BWR design have a negative ...
  88. [88]
    Physics of Plutonium Recycling - Volume III - Nuclear Energy Agency
    Nov 10, 2020 · ... void coefficient in mixe ... reactivity feedback associated with steam formation should always remain negative for inherent safety reasons.
  89. [89]
    Multi-physics analysis promotes safe operation of advanced nuclear ...
    Jan 2, 2025 · They demonstrated how inherent safety feedback mechanisms can prompt safe reactor shutdown without any human operation or automatic control rod ...
  90. [90]
    [PDF] Safety margins of operating reactors
    The most important safety margins relate to physical barriers against release of radioactive material, such as fuel matrix and fuel cladding, ...<|separator|>
  91. [91]
    The negative temperature coefficient as a natural safety feature of ...
    Dec 31, 1989 · The negative temperature coefficient as a natural safety feature of SLOWPOKE reactors. Conference · Sun Dec 31 23:00:00 EST 1989 ...
  92. [92]
    Physics of nuclear reactor safety - IOP Science
    Reactivity properties of a reactor core depend strongly on the regions of neutron energy in which the nuclear reactions, relevant for the neutron balance, take ...
  93. [93]
    [PDF] Reactivity Accidents - INIS-IAEA
    (1) Reactivity accidents must be prevented; (2) If they nevertheless occur, systems are provided to protect against them; and (3) If any fission products are ...
  94. [94]
    Decay Heat - Decay Energy - Nuclear Power
    The decay heat produced after a reactor shutdown from full power is initially equivalent to 6 to 7% of the rated thermal power.
  95. [95]
    Loss of Coolant Accidents - an overview | ScienceDirect Topics
    A Loss-of-Coolant Accident (LOCA) is defined as an event resulting from a break in the reactor coolant pressure boundary, leading to a rise in reactor core ...
  96. [96]
    [PDF] Nuclear Fuel Behaviour in Loss-of-coolant Accident (LOCA ...
    The 1986 report was limited to the oxidation, embrittlement and deformation of pressurised water reactor (PWR) fuel in a loss-of-coolant accident (LOCA). Since ...
  97. [97]
    Chernobyl: Chapter I. The site and accident sequence
    The most important characteristic of the RBMK reactor is that it possesses a "positive void coefficient". This means that if the power increases or the flow ...
  98. [98]
    A simplified analysis of the Chernobyl accident | EPJ N
    Jan 20, 2021 · This process will bring about changes in reactor reactivity. We know (see INSAG-7) that there was xenon poisoning in the reactor at the time of ...
  99. [99]
    A reactor physicist explains Chernobyl - American Nuclear Society
    Apr 28, 2022 · “This is not a safe design feature,” he said, adding that reactors in the United States are not allowed to have a positive void coefficient.
  100. [100]
    Three Mile Island Accident - World Nuclear Association
    Oct 11, 2022 · In 1979 at Three Mile Island nuclear power plant in USA a cooling malfunction caused part of the core to melt in the #2 reactor.
  101. [101]
    Backgrounder on the Three Mile Island Accident
    A combination of equipment malfunctions, design-related problems and worker errors led to TMI-2's partial meltdown and very small off site releases of ...
  102. [102]
    Fukushima Daiichi Accident - World Nuclear Association
    All three cores largely melted in the first three days. The accident was rated level 7 on the International Nuclear and Radiological Event Scale, due to high ...Missing: physics | Show results with:physics
  103. [103]
    Full article: Research activities on nuclear reactor physics and ...
    Just after the Fukushima accident, concerned issues related to the reactor physics were possibility of recriticality and quantification of radioactive materials ...
  104. [104]
    Benchmark Study of the Accident at Fukushima (BSAF)
    It is widely believed that all three reactor cores experienced some melting, although the extent is as yet unknown. The consequent release of radioactive ...