Fact-checked by Grok 2 weeks ago

High-temperature gas-cooled reactor

A high-temperature is a design that employs gas as a and as a moderator, utilizing TRISO-coated particles to achieve core outlet temperatures of 700–950°C while ensuring through passive heat removal mechanisms. HTGRs feature modular construction in either prismatic block or pebble-bed configurations, with power outputs ranging from small-scale reactors (e.g., 30 MWth) to larger demonstration units (up to 600 MWth per module), enabling high exceeding 45% for and of process heat. The TRISO fuel, consisting of or carbide kernels encased in multiple layers, confines products even at temperatures up to 1600–1800°C, preventing release during normal operation or severe accidents. This design's low and graphite's high allow for natural convection cooling, eliminating the risk of meltdown and providing extended grace periods—often days—without active intervention. Development of HTGRs dates back to the , with early experimental plants like the U.S.'s Peach Bottom (115 MWth, 40 MWe, operational 1967–1974) and Fort St. Vrain (330 MWe, 1976–1989) demonstrating the technology's feasibility, followed by international efforts in , , and . As a Generation IV system, HTGRs support decarbonization by providing high-temperature steam or gas for industrial applications, including via thermochemical processes, , and manufacturing, with economic advantages such as lower power generation costs compared to light-water reactors. As of November 2025, operational HTGRs include Japan's High-Temperature Test Reactor (HTTR, 30 MWth, critical since 1998 and restarted in 2021 for ongoing research) and China's demonstration plant (210 MWe, which entered commercial operation in December 2023), marking the first grid-connected pebble-bed HTGR. Ongoing projects, such as the U.S.-based Xe-100 (targeting deployment by 2027) and international collaborations under the Generation IV International Forum, focus on licensing, fuel supply chains for high-assay low-enriched , and integration with to enhance global .

Fundamentals

Core Principles

A high-temperature gas-cooled reactor (HTGR) is a type of nuclear reactor that employs graphite as a moderator and helium as a coolant to achieve core outlet temperatures typically ranging from 750°C to 950°C, facilitating high-efficiency electricity generation and the provision of process heat for industrial applications. This design leverages the inherent properties of its materials to operate at elevated temperatures without the need for active cooling systems under normal conditions, distinguishing it from light-water reactors. Within international frameworks, HTGRs are classified as a variant of the Very High Temperature Reactor (VHTR) under the Generation IV International Forum (GIF), an initiative coordinated by multiple nations to advance sustainable nuclear technologies, and are recognized by the International Atomic Energy Agency (IAEA) as an advanced gas-cooled reactor system emphasizing fuel efficiency and high-temperature performance. HTGRs maintain a predominantly thermal neutron spectrum, where fast neutrons from fission are slowed down primarily through elastic scattering interactions with carbon atoms in the graphite moderator, enabling efficient fission in low-enriched uranium fuel. This spectrum includes contributions from epithermal neutrons, particularly in the resonance energy range, which influence self-shielding effects in fuel particles and are accounted for in core calculations using multi-group neutron transport methods. The core configuration can be either pebble-bed, featuring spherical fuel elements continuously recirculated for online refueling, or prismatic, using stacked hexagonal graphite blocks with embedded fuel compacts for batch refueling, both of which support the thermal spectrum while optimizing neutron leakage and power distribution. The economy in HTGRs benefits from 's low absorption cross-section, approximately 100 times lower than that of , which minimizes parasitic losses and allows for higher fuel utilization compared to water-moderated reactors. By using to moderate without introducing , HTGRs eliminate risks associated with aqueous coolants and structural materials, enhancing long-term core integrity and operational reliability. This solid moderator also provides structural support and heat conduction paths, contributing to a favorable balance of reactivity control and potential exceeding 100 GWd/t in thermal spectra. The thermal power generated in an HTGR core, denoted as P, arises from the energy released by nuclear fissions and can be derived from fundamental reactor physics principles tailored to graphite-moderated systems. The rate of fissions per unit volume is given by the product of the average neutron flux \phi (in neutrons/cm²·s) and the macroscopic fission cross-section \Sigma_f (in cm⁻¹), which for thermal neutrons in HTGRs is dominated by the ^{235}U fission cross-section of approximately 580 barns at 0.025 eV, adjusted for the graphite's moderating ratio of about 200 (mean logarithmic energy loss per collision). Integrating over the core volume V yields the total fission rate N_f = \phi \Sigma_f V fissions per second. Each fission releases an average recoverable energy E_f of about 200 MeV (or $3.2 \times 10^{-11} J), primarily as kinetic energy of fission products and prompt neutrons, with approximately 7% additional contribution from delayed beta and gamma decays of fission products. Thus, the total power is P = N_f E_f = \phi \Sigma_f E_f V. In HTGR-specific analyses, \phi is typically on the order of $10^{13} to $10^{14} n/cm²·s in the core center due to the low power density (around 5-10 MW/m³) enabled by graphite's thermal properties, ensuring the equation aligns with measured outputs in prototypes like the HTR-10 (10 MWth). P = \phi \Sigma_f E_f V

Thermodynamic Operation

In high-temperature gas-cooled reactors (HTGRs), heat generated by in the TRISO-coated fuel particles is transferred to the primarily through conduction within the matrix of the fuel elements and via the forced flow of through the core channels or pebble bed. Fission energy deposits directly in the fuel kernels, conducting outward through the pyrocarbon and layers into the surrounding , which acts as both moderator and structural material, before being convected to the stream. During normal operation, enters the core at 250–350°C and exits at 750–900°C, with forced dominating the , enhanced by 's high thermal conductivity of approximately 0.3 W/m·K at operating temperatures. Radiation contributes in pebble bed configurations, particularly across voids and gaps, but remains secondary to under steady-state conditions. HTGRs employ a closed for efficient , where the hot coolant from the core outlet directly or indirectly drives a , bypassing traditional steam cycles to achieve compact power conversion. The cycle typically includes a , recuperator, precooler, , and multi-stage , with recirculating at pressures of 6–7 MPa. Turbine inlet temperatures reach up to 850°C in designs like the GT-MHR and PBMR, limited by metallic component materials, enabling direct cycle operation that simplifies the plant layout and reduces capital costs compared to indirect cycles. This integration leverages 's chemical inertness and low neutron absorption, allowing high-temperature operation without issues common in water-cooled systems. The thermodynamic efficiency of the Brayton cycle in HTGRs benefits from elevated temperatures and helium's favorable properties, outperforming light-water reactors (LWRs). The ideal Carnot efficiency provides a theoretical bound, given by \eta = 1 - \frac{T_\text{cold}}{T_\text{hot}} where temperatures are in Kelvin; for typical HTGR conditions with T_\text{hot} \approx 1123 K (850°C) and T_\text{cold} \approx 400 K (compressor inlet after intercooling), \eta \approx 64\%. Actual Brayton efficiencies account for irreversibilities and reach 45–48% in HTGRs, compared to 33% in LWRs, due to helium's low molecular weight (4 g/mol), which minimizes compressor work, and its high thermal conductivity, which improves recuperator effectiveness up to 95%. This efficiency advantage enhances overall plant economics and fuel utilization. In pebble bed HTGRs, the across the core, critical for determining requirements, is governed by the , derived for flow through porous packed beds and adapted for multiphase helium-graphite interactions. The equation combines viscous (laminar) and inertial (turbulent) contributions: \frac{\Delta P}{L} = 150 \frac{(1-\epsilon)^2}{\epsilon^3} \frac{\mu v}{d_p^2} + 1.75 \frac{(1-\epsilon)}{\epsilon^3} \frac{\rho v^2}{d_p} The viscous term originates from the Blake-Kozeny model, treating the bed as a of capillaries where loss scales with \mu, superficial velocity v, bed height L, \epsilon (typically 0.39–0.42 in HTGR pebbles), and particle diameter d_p (60 mm for standard fuel pebbles); it dominates at low Reynolds numbers (Re < 1000) common in HTGR flows. The inertial term, from the Burke-Plummer extension, captures form drag in turbulent regimes (Re > 1000), proportional to \rho and v^2. For HTGR pebble flow, with v \approx 5–20 m/min and \mu \approx 4 \times 10^{-5} Pa·s, total \Delta P is 0.2–0.5 across a 10 m core, ensuring efficient circulation; the KTA correlation refines this for anisotropic beds by incorporating experimental pebble bed data. Core outlet temperature is regulated in HTGRs by modulating helium mass flow via variable-speed circulators and adjusting reactor power through control rods, maintaining exit temperatures at 700–900°C to optimize cycle performance while avoiding fuel limits. In the HTR-10 test reactor, for instance, flow rates of 4.3 kg/s achieve 900°C outlet during high-temperature phases, with bypass valves fine-tuning distribution to prevent hot spots. Helium purity is strictly controlled at 99.99% (total impurities <100 ppm, including H_2O <0.1 ppm and CO/CO_2 <10 ppm) using purification loops with getters and catalysts to remove oxidizing species, thereby preventing chronic graphite corrosion that could compromise core integrity over decades of operation.

Design Components

Moderator and Reflector

High-purity graphite serves as the primary moderator in high-temperature gas-cooled reactors (HTGRs) due to its low neutron absorption cross-section, typically \Sigma_a < 0.0035 \, \text{cm}^{-1}, which minimizes parasitic neutron capture while efficiently slowing fast neutrons to thermal energies through elastic scattering. This material also exhibits exceptional thermal stability, maintaining structural integrity up to 1600°C, well beyond typical HTGR core outlet temperatures of 750–950°C, owing to its high sublimation point and resistance to oxidation in inert helium environments. The reflector in HTGR designs consists of radial and axial layers, often constructed from graphite or beryllium, surrounding the core to minimize neutron leakage and enhance criticality by reflecting escaping neutrons back into the fissile region. Beryllium offers superior reflection efficiency due to its low atomic mass and high scattering cross-section, while graphite provides cost-effective, compatible structural support; these configurations typically achieve an effective multiplication factor k_\text{eff} > 1.05, ensuring sufficient excess reactivity for operational margins. Graphite's performance under operational conditions is influenced by and irradiation-induced dimensional changes, which affect core geometry and neutronics over the reactor's lifetime. The coefficient of thermal expansion (CTE) for nuclear-grade is approximately $4 \times 10^{-6} \, \text{K}^{-1} between 20–120°C, leading to manageable expansion at elevated temperatures, but fast neutron irradiation initially causes densification with a change \Delta \rho / \rho = -0.1\% per $10^{21} \, \text{n/cm}^2 (E > 0.1 MeV), followed by swelling at higher doses due to crystal lattice disruption and void formation. These effects are mitigated through allowances for anisotropic and periodic to prevent excessive or cracking in moderator blocks. Manufacturing standards for HTGR graphite emphasize isostatic pressing to achieve isotropic properties and purity levels below 5 for impurities like to preserve low absorption. A representative grade, IG-110, developed for HTGR applications, features a of 1.78 g/cm³ and in-plane thermal conductivity of 116 W/m·K at (decreasing to approximately 70 W/m·K at 600°C), enabling efficient heat dissipation from the core while supporting structural loads under . In prismatic HTGR cores, the moderator consists of dense hexagonal graphite blocks with integrated fuel compacts and coolant channels, achieving near-solid packing with low void fractions for optimal neutron economy. In contrast, pebble-bed designs employ graphite-moderated fuel spheres in a loosely , resulting in moderator packing densities with 60–70% void fraction to facilitate helium flow and online refueling, though this increases neutron leakage compared to prismatic arrangements.

Fuel and Cladding

The fuel in high-temperature gas-cooled reactors (HTGRs) primarily consists of tri-structural isotropic (TRISO) coated particles, which serve as micro-encapsulated elements designed for exceptional durability under high temperatures and irradiation. Each TRISO particle features a central kernel of (UO₂) or uranium oxycarbide (UCO), typically 350–600 μm in , coated with multiple layers: a porous buffer layer of (PyC) approximately 95 μm thick to accommodate gas swelling and ; an inner dense isotropic PyC (IPyC) layer about 40 μm thick for structural support; a (SiC) layer around 35 μm thick acting as the primary product barrier; and an outer dense isotropic PyC (OPyC) layer of similar thickness to protect the SiC and facilitate particle handling. The SiC layer, with its high and , withstands temperatures up to 1800°C while maintaining integrity, enabling the fuel to operate in HTGR environments exceeding 1000°C without significant degradation. Enrichment levels for HTGR TRISO fuel are tailored to achieve efficient economy in graphite-moderated cores, typically ranging from 8–20% ²³⁵U for low-enriched uranium (LEU) kernels to support standard operations, though mixed oxide (MOX) variants incorporate up to 19.75% ²³⁹Pu blended with for advanced cycles like plutonium disposition or utilization. These enrichment choices balance criticality, potential, and resistance, with LEU dominating modern designs to align with non-proliferation goals. The protective multilayer cladding of TRISO particles eliminates the need for traditional metallic sheaths used in light-water reactors, as the coatings provide inherent containment against corrosion and mechanical stress in the helium environment. HTGR fuels achieve high burnup rates, up to 120 GWd/t, owing to their deep-burn capability, which maximizes fuel utilization through extended without compromising particle integrity. (BU) is the total thermal energy extracted per metric ton of initial heavy metal, quantified in GWd/t. This performance stems from the TRISO design's resistance to kernel migration and coating failure, allowing sustained operation at high fluxes. Deep-burn variants can exceed 150 GWd/t in optimized prismatic or pebble-bed configurations. HTGRs employ two principal fuel assembly types: pebble-bed and prismatic. In pebble-bed designs, fuel is embedded in graphite spheres (pebbles) of 60 mm diameter, each containing approximately 15,000 TRISO particles distributed within a 50 mm fueled zone, enabling online refueling and continuous core circulation for uniform burnup. Prismatic assemblies, by contrast, use hexagonal graphite blocks housing cylindrical compacts with around 10,000 TRISO particles per compact, offering higher power density but requiring batch refueling. The pebble-bed approach enhances thermal-hydraulic stability, while prismatic designs suit modular reactors with fixed geometries. Fission product retention in TRISO fuel relies on diffusion-limited transport models, such as Fick's laws and the Booth model, which predict negligible release through intact coatings at operational temperatures. Below 1600°C, these models indicate a release fraction less than 10⁻⁶ for key products like cesium (Cs) and strontium (Sr), with noble gases like krypton (Kr) showing fractions around 10⁻⁹ due to the low diffusivity in PyC and SiC layers (e.g., Cs diffusion coefficient in UO₂: D = 0.90 \times 10^{-18} \exp(-209 \, \text{kJ/mol}/RT) \, \text{m}^2/\text{s}). This containment ensures radiological safety even during transients, as metallic fission products remain trapped unless coatings fail, which is rare below design limits.

Coolant and Circulation

Helium serves as the primary coolant in high-temperature gas-cooled reactors (HTGRs) due to its chemical inertness, which prevents reactions with core materials such as and fuel cladding, ensuring long-term structural integrity. Its low thermal neutron absorption cross-section of approximately 0.0005 minimizes parasitic losses, supporting efficient chain reactions without significant interference. Additionally, exhibits a high of 5.19 J/g·K at 900°C, facilitating effective removal from the core while maintaining single-phase across operational temperatures up to 950°C. The primary coolant loop in HTGRs employs recirculating blowers to maintain circulation, typically operating at pressures of 5–7 to balance efficiency and system compactness. For a representative 350 MWth , mass flow rates range from 100–300 kg/s, directed through core channels to achieve temperatures around 300–400°C and outlet temperatures of 700–900°C, depending on design. This closed-loop configuration isolates the high-temperature primary from secondary systems, with blowers providing the necessary head to overcome pressure drops in the , , and . Heat transfer from the primary occurs via intermediate heat exchangers (IHXs), which employ an indirect loop to isolate the reactor from power conversion or process fluids, enhancing by preventing cross-contamination. In some designs, such as those coupled to advanced processes, the intermediate loop uses sodium as a secondary for its high thermal conductivity, while others incorporate to avoid reactivity issues in high-temperature applications. These IHXs, often helical-coil or plate-fin types, operate at helium-side pressures matching the primary loop, transferring heat at efficiencies exceeding 90% while maintaining differential pressures to ensure . Impurity management is critical in HTGRs to prevent corrosion of graphite components, with hydrogen and methane levels controlled using getters such as titanium sponges or palladium-based absorbers integrated into the purification system. These impurities, arising from minor leaks or , can otherwise promote graphite oxidation; however, with effective control maintaining concentrations below 1–10 , corrosion rates are limited to less than 0.1 mm/year at 900°C, preserving core integrity over decades of operation. The helium purification system operates on a continuous bypass cycle, diverting 1–15% of the primary flow through a series of components to remove impurities and restore coolant purity. Incoming helium passes first through cartridge filters to capture particulates, followed by copper oxide beds at 250–400°C to oxidize hydrogen to water and carbon monoxide to dioxide. Subsequent molecular sieve traps adsorb water vapor and carbon dioxide, while low-temperature charcoal beds (around -196°C using liquid nitrogen) capture nitrogen, methane, and residual gases. The purified helium is then recombined with the main flow, with system sizing ensuring a purification constant of at least 2.9 × 10⁻⁵ s⁻¹ to achieve full coolant cleanup within 24 hours post-shutdown.

Control Mechanisms

Control mechanisms in high-temperature gas-cooled reactors (HTGRs) ensure stable reactivity during operation, load following, and shutdown. Primary reactivity control is achieved through control rods, which consist of neutron absorbers such as (B₄C) or encased in sleeves to compatibly interface with the -moderated . These rods are positioned in channels within the side reflector or core periphery, allowing precise adjustment of absorption to maintain criticality. For emergency shutdown (), the rods are gravity-driven or electromagnetically released, achieving full insertion in less than 2 seconds, typically around 0.6–0.7 seconds to 80% of effective length, ensuring rapid negative reactivity insertion of several percent Δk/k. Burnable poisons are integrated into the fuel elements to manage initial excess reactivity and prevent excessive peaking early in the core life. Common materials include or compounds, which have high absorption cross-sections that diminish over time as the isotopes burn up, providing gradual reactivity hold-down without compromising long-term fuel utilization. These poisons are dispersed within TRISO fuel particles or fuel compacts, typically at concentrations optimized for equilibrium , such as in prismatic or pebble-bed designs. Reactivity feedback coefficients contribute to inherent stability in HTGRs. The Doppler coefficient, arising primarily from fuel temperature broadening of neutron resonances (mainly in ²³⁸U), is negative at approximately α_D = -0.5 pcm/K, enhancing self-regulation during power transients. The , which could be positive due to reduced coolant density increasing leakage in gas-cooled systems, is mitigated by design features like dense moderation and core geometry, ensuring the overall remains negative (typically -1 to -7 pcm/K total). Reactivity balance is quantified by ρ = (k_eff - 1)/k_eff, where k_eff is the effective multiplication factor; in HTGRs, operational designs maintain |ρ| < 1% (1000 pcm) during load following by balancing positions against fuel depletion, buildup, and temperature feedbacks, enabling flexible power adjustment without instability. Reserve shutdown systems provide diversity beyond primary s, often employing soluble injection as a mechanism to flood channels or reflectors with neutron-absorbing , achieving shutdown margins exceeding 1% Δk/k even if rods fail. This is supplemented in some designs by absorber ball systems using pellets for gravity insertion.

Historical Development

Origins and Early Experiments

The conceptual origins of high-temperature gas-cooled reactors (HTGRs) emerged in the United States during the 1940s amid efforts to develop nuclear propulsion for military aircraft, driven by the need for compact, high-temperature heat sources capable of powering air-breathing engines without conventional fuel. The Aircraft Nuclear Propulsion (ANP) program, launched in 1946 under the joint oversight of the U.S. Air Force and the Atomic Energy Commission, explored gas-cooled reactor designs to achieve outlet temperatures exceeding 800°C for efficient jet propulsion. A key milestone was the Gas Cooled Reactor Experiment (GCRE), conducted from 1957 to 1959 at the National Reactor Testing Station in Idaho, which tested a helium- and nitrogen-cooled, graphite-moderated core at up to 32 MWth and 871°C, validating the thermal and neutronic performance of high-temperature gas cooling systems. These military-focused experiments provided foundational data on materials and heat transfer that influenced subsequent civilian HTGR development, shifting emphasis from propulsion to stationary power generation. The first dedicated HTGR prototype was the Dragon reactor in the , an international collaboration under the () and , constructed at and achieving criticality in 1963 with operations continuing until 1976. Rated at 20 MWth, Dragon featured a helium-cooled, -moderated core using prismatic fuel elements, primarily serving as a test platform for coated-particle fuels, high-temperature components, and helium circulation systems under pressures up to 2 . Over its lifespan, it irradiated more than 250 fuel elements, demonstrating core stability at outlet temperatures around 750°C and contributing critical insights into fission product retention and thermal hydraulics that informed global HTGR designs. The project's success highlighted the viability of helium as a for achieving higher efficiencies than earlier carbon dioxide-cooled reactors. In the United States, early HTGR testing advanced to grid-connected operation with the Peach Bottom Unit 1 reactor in , which became critical in and supplied electricity from 1967 until its shutdown in 1974. This 40 MWe (115 MWth) facility, developed by Philadelphia Electric Company under the Atomic Energy Commission's initiative, was the world's first HTGR to deliver commercial power, using a helium-cooled core with hexagonal fuel blocks at outlet temperatures up to 760°C. It accumulated over 1,349 equivalent full-power days, testing fuel performance and steam cycle integration while confirming the technology's potential for baseload electricity with minimal operational incidents. Peach Bottom's data on core physics and component reliability directly supported larger-scale HTGR concepts. Germany pioneered the pebble-bed variant of HTGR with the AVR reactor at Jülich, operational from 1967 to 1988 as a 15 MWe (46 MWth) experimental unit that demonstrated continuous fuel recirculation. Cooled by helium at an average outlet temperature of 950°C—elevated from an initial 850°C in 1974—this design used spherical fuel elements containing thousands of coated uranium carbide particles, achieving high burnups and inherent safety through negative temperature coefficients. The AVR's 21-year operation provided essential validation of pebble-bed flow dynamics, dust management, and high-temperature materials, influencing subsequent modular HTGR architectures despite challenges like metallic impurity contamination. In the during the 1960s, HTGR research diverged from the dominant design—which employed but for dual civilian and plutonium roles—toward helium-cooled concepts at facilities like the of . Early efforts focused on theoretical studies and small-scale loop tests for high-temperature gas cooling, aiming to leverage graphite's neutron economy for efficient power and process heat, though full-scale prototypes remained conceptual until the 1970s VGR-50 project. This parallel path underscored the Soviet emphasis on versatile graphite-based systems while addressing distinct challenges in helium purity and fuel fabrication.

Major Milestones and International Projects

The Fort St. Vrain (FSV) nuclear generating station in the United States represented a significant in scaling up prismatic-block high-temperature gas-cooled reactor (HTGR) to commercial power generation levels. Constructed by and owned by Company of Colorado, the 330 (842 MWth) plant achieved initial criticality in 1974 and began commercial operation in 1976, operating until its permanent shutdown in 1989. Despite demonstrating the feasibility of -cooled prismatic fuel assemblies with thorium-uranium cycles, FSV encountered operational challenges, particularly repeated steam ingress events from the intermediate heat exchangers into the primary circuit, which led to oxidation and required extensive maintenance. These incidents provided critical lessons on material compatibility and system integrity under high-temperature conditions, influencing subsequent HTGR designs to prioritize advanced steam generators and ingress mitigation strategies. In , the marked the first full-scale deployment of pebble-bed HTGR technology for electricity production, advancing the modular concept toward commercialization. The 300 MWe (750 MWth) prototype, developed by Hochtemperatur-Kernkraftwerk GmbH (HKG) and commissioned in 1983 near Hamm, utilized thorium-highly fuel pebbles and achieved over 16,000 hours of operation before its shutdown on September 1, 1989. Key operational hurdles included difficulties with the continuous pebble refueling system, which experienced blockages and handling incidents that increased and maintenance costs. Although technically successful in validating pebble-bed core physics and safety margins, the plant's closure was precipitated by economic pressures and public opposition following these fuel handling issues, underscoring the challenges of integrating complex online refueling in commercial settings. Japan's High-Temperature Test Reactor (HTTR), a 30 MWth prismatic HTGR built by the Japan Atomic Energy Agency (JAEA) at the Oarai Research and Development Center, achieved criticality in 1998 and has since served as a cornerstone for high-temperature applications research. In June 2004, the HTTR first reached its design outlet temperature of 950°C, but a landmark milestone came in March 2010 with the completion of a 50-day continuous operation at this temperature, demonstrating long-term stability for process heat utilization such as and industrial . This achievement validated the reactor's features and fuel performance under prolonged high-temperature exposure, providing data that informed international HTGR safety standards and advanced the Generation IV very high-temperature reactor (VHTR) framework. The Pebble Bed Modular Reactor (PBMR) project in aimed to commercialize modular pebble-bed HTGRs for flexible, factory-assembled deployment, drawing on AVR and heritage. Initiated in 1993 by and the PBMR Pty Ltd consortium, the design targeted 165 MWe per module with helium outlet temperatures up to 900°C, emphasizing economic viability through and high fuel burnup. However, escalating development costs, coupled with the global and inability to secure customers or firm orders, led to progressive funding cuts by the South African government—first in March 2010 and fully in September 2010—resulting in the project's cancellation. In November 2025, the South African government announced plans to revive the project, lifting it from care and maintenance. Despite the termination, the PBMR effort yielded valuable insights into modular , fuel qualification, and economic modeling, which have influenced subsequent pebble-bed initiatives elsewhere. China's , a 10 MWth pebble-bed test reactor at the Institute of Nuclear and New Energy Technology (INET) in , , became operational in 2000 and has played a pivotal role in reviving and advancing HTGR technology on an industrial scale. Achieving full-power operation in 2003, the HTR-10 demonstrated safe shutdown without active systems during loss-of-coolant tests and provided essential data on pebble-bed neutronics, thermohydraulics, and fuel integrity at 750°C outlet temperatures. This experimental success directly paved the way for the demonstration project, a 210 MWe twin-module plant at Shidao Bay that entered commercial operation in December 2023, by validating the scalability of modular pebble-bed designs for commercial power and .

Safety Characteristics

Inherent Safety Features

High-temperature gas-cooled reactors (HTGRs) incorporate several features that rely on physical properties and design principles to prevent damage and product release without the need for active or operator action. These features stem from the reactor's materials, form, and characteristics, ensuring self-regulation and passive management during normal operation and postulated accidents such as loss of or loss of power. A key inherent safety mechanism in HTGRs is the strong negative temperature coefficient of reactivity, which provides automatic self-regulation of the reactor power. As the core temperature rises, the reactivity decreases due to the of neutron absorption resonances in the and thermal expansion effects in the graphite moderator, leading to a reduction in power without control rod insertion. This coefficient remains negative across the full range of operating and accident temperatures, ensuring the reactor shuts down passively even in scenarios with limited flow. The TRISO (tristructural-isotropic) coated particle fuel represents another fundamental element, designed to maintain integrity and retain products under extreme conditions. Each particle consists of a kernel surrounded by multiple layers of and , which act as a robust and diffusion barrier. This fuel form retains products with very low release fractions—typically less than 0.1% for key isotopes—up to temperatures of 1600°C, well beyond the of conventional fuels, preventing significant radioactive release even if overheats. HTGRs operate at a low core power density, typically 3–5 kW/L, which is approximately 20–30 times lower than the 100 kW/L in light-water reactors (LWRs). This low density, combined with the large thermal mass of the graphite moderator, allows for extended passive removal of decay heat following shutdown, as the heat generation rate per unit volume remains manageable without forced cooling. The design ensures that fuel temperatures stay below critical limits for prolonged periods, facilitating natural dissipation to the environment. The core structure further enhances through its excellent thermal conductivity and high , enabling passive heat dissipation via conduction and after a (LOCA). In the absence of active cooling, conducts through the graphite matrix to the and is radiated to surrounding structures or convected by residual gas, maintaining peak temperatures below 1600°C indefinitely. This passive mechanism eliminates the risk of core meltdown, as demonstrated in safety tests on the , where the core remained intact during simulated complete blackout and LOCA scenarios with no active systems.

Accident Mitigation Systems

High-temperature gas-cooled reactors (HTGRs) incorporate engineered accident mitigation systems to address beyond-design-basis events, ensuring and minimizing radiological releases through redundant, passive, and active features. These systems complement the inherent safety characteristics of HTGRs, such as high thermal margins and negative reactivity feedback, by providing additional barriers and cooling pathways during severe accidents like loss-of-coolant or external hazards. Containment structures in modular prismatic HTGR designs primarily rely on steel pressure vessels (RPVs) to enclose the reactor core, primary circuit, and associated components, maintaining integrity under accident conditions. The RPV serves as the final confinement barrier, designed to withstand internal pressures from primary coolant leaks or depressurization events. For instance, in designs like the General Atomics modular HTGR, the RPV holds helium pressures up to 6 MPa while accommodating thermal expansions and seismic loads, preventing uncontrolled release of fission products. This structure also integrates liners and insulation to limit heat transfer and corrosion, ensuring long-term stability during extended accident scenarios. Emergency cooling systems focus on passive decay heat removal to prevent core damage without relying on active power sources. Natural circulation loops utilize the helium coolant's buoyancy to transfer residual heat from the core to external heat exchangers during loss-of-forced-cooling events, maintaining fuel temperatures below 1600°C for extended periods. Complementing this, the reactor vessel auxiliary cooling system (RVACS) employs air-cooled heat exchangers surrounding the reactor vessel, dissipating up to 1-2% of full power as decay heat through natural convection and radiation, as demonstrated in analyses for modular HTGRs where RVACS alone suffices for indefinite cooling post-shutdown. These systems are integral to prismatic and pebble-bed configurations, providing multi-layered redundancy. Hydrogen management addresses potential generation from graphite moderator oxidation in steam-ingress accidents, using passive autocatalytic recombiners (PARs) to mitigate combustion risks. These devices, strategically placed within the , catalytically recombine and oxygen into without external power, maintaining concentrations below 4% to prevent . In advanced HTGR designs, PARs are integrated into the RPV or , drawing on natural for effective distribution, as validated in severe simulations where they reduce hydrogen buildup by over 90% within hours. Seismic and flood protections are designed to withstand extreme external events, with HTGRs specifying a design basis typically of 0.2–0.3g (), depending on site-specific hazards, with margins to higher loads such as 0.5g to ensure structural integrity of the RPV and supports. Seismic and damping features, such as base mats and flexible piping, limit accelerations to below equipment qualification thresholds, as seen in the prototype where analyses confirmed no disruption under 0.5g horizontal loads. For flooding, elevated siting raises critical components above probable maximum flood levels—typically 5-10 meters above in coastal designs like —incorporating watertight barriers and drainage to prevent inundation of safety systems. Following the 2011 Fukushima-Daiichi accident, modern HTGRs like China's have implemented post-Fukushima upgrades, including enhanced venting systems with filtered paths to manage buildup and recombiners for severe . These incorporate advanced , such as real-time sensors and seismic detectors integrated into the and , ensuring early detection and automated response to multi-unit or external events. Such enhancements, verified through re-evaluations and commissioning tests in 2021–2023, confirm the HTR-PM's ability to maintain core integrity without off-site power for over 72 hours.

Advantages and Applications

Performance and Economic Benefits

High-temperature gas-cooled reactors (HTGRs) achieve thermal efficiencies of up to 50%, significantly higher than the approximately 33% typical of light-water reactors (LWRs), due to their high core outlet temperatures of 700–950°C that enable advanced Brayton or combined cycles. This elevated efficiency reduces fuel consumption per unit of electricity generated, with burnups reaching 80,000–150,000 MWd/tU, compared to 40,000–50,000 MWd/tU in LWRs, thereby minimizing the volume of spent fuel and by factors of up to four relative to LWRs through more complete of . Additionally, HTGRs support extended fuel cycles of 3–6 years between refuelings, facilitated by robust TRISO-coated particle fuel that withstands high temperatures without failure, allowing for operational flexibility in both prismatic and pebble-bed configurations. Economically, modular HTGR designs benefit from factory fabrication of standardized components, such as reactor vessels, heat exchangers, and fuel elements, which reduces on-site construction time to approximately 3 years—half that of traditional large-scale reactors—while improving and lowering labor costs through serial production. (LCOE) estimates for nth-of-a-kind modular HTGRs range from $60–80/MWh in 2019 USD (escalated to similar ranges for 2025 projections), competitive with or lower than LWRs at $70–90/MWh, owing to high capacity factors exceeding 90% and reduced outage durations of 30–60 days per cycle. These factors contribute to overall cost advantages, with potentially 20–30% lower than non-modular designs due to in . From an environmental perspective, HTGRs exhibit lifecycle of approximately 10 g CO₂ eq./kWh, among the lowest for technologies, primarily from and processing rather than operations. In deep-burn modes, where transuranic elements from LWR spent fuel are incorporated, HTGRs transmute long-lived actinides into short-lived fission products, eliminating most long-lived components and further reducing radiotoxicity and disposal burdens compared to conventional cycles.

Industrial and Hydrogen Uses

High-temperature gas-cooled reactors (HTGRs) are particularly suited for delivering process heat at temperatures up to 950°C, enabling decarbonization in energy-intensive sectors such as , , and chemical processing. These reactors can supply high-quality heat through intermediate heat exchangers, reducing reliance on fossil fuels for processes like reduction in (requiring 800–900°C) and in production (around 900°C). In chemical industries, HTGR heat supports endothermic reactions, such as synthesis or production, with modular designs like the Xe-100 providing 200 MWth of thermal output tailored for such applications. For example, in March 2025, Dow and submitted a construction permit application to the for a proposed Xe-100 HTGR project at Dow's site, aimed at providing power and steam for , with NRC review ongoing as of November 2025. A key application is via thermochemical , where HTGRs integrate with cycles like the sulfur-iodine (S-I) process to achieve efficiencies of 40–50% on a higher heating value basis. The S-I cycle involves high-temperature decomposition of , represented by the reaction: \mathrm{H_2SO_4 \rightarrow SO_2 + H_2O + \frac{1}{2}O_2} at approximately 800°C, followed by lower-temperature steps for iodine recycling and generation, all powered by HTGR heat without direct electrical input. For instance, a 200 MWth HTGR module can support S-I cycle operations to produce significant volumes, with studies showing net thermal efficiencies up to 47.6% when coupled to reactors outlet temperatures of 950°C. This approach yields zero-carbon suitable for fuels, chemicals, and . HTGRs also enable for and seawater , leveraging their high outlet temperatures for efficient multi-purpose operation. The High-Temperature Test Reactor (HTTR) in demonstrated this capability by achieving 950°C outlet temperatures and testing steam generation for , producing up to 10 tons of fresh water per day in coupled systems. Such configurations allow simultaneous electricity, , and water production, with via multi-stage flash processes benefiting from HTGR's stable thermal supply at 80–150°C after heat cascading. For production (CO + H₂), HTGRs facilitate high-temperature steam methane reforming at 800–1000°C, enhancing over conventional methods by providing clean to endothermic reactions. This process supports downstream applications like Fischer-Tropsch synthesis for synthetic fuels, with nuclear-assisted reforming reducing CO₂ emissions by up to 35% compared to natural gas-fired plants. Market projections indicate that sources, including HTGRs, could supply up to 10% of heat demand by 2030, driven by decarbonization goals and the technology's ability to meet high-temperature needs in hard-to-abate sectors. By , initial deployments in pilots are expected to demonstrate scalability, with the market growing to support this transition.

Deployment and Status

Operational and Decommissioned Reactors

The High-Temperature Gas-cooled Reactor Pebble-bed Module (HTR-PM) in , located at Shidao Bay in Province, represents the world's first commercial-scale modular HTGR deployment. This plant features two 250 MWth reactor modules coupled to a single , delivering a net electrical output of 210 MWe with a coolant outlet temperature of 750°C. It achieved initial grid connection on December 20, 2021, and entered full commercial operation in December 2023, demonstrating stable performance in and planned co-generation applications. The High-Temperature Test Reactor (HTTR) in , operated by the Japan Atomic Energy Agency (JAEA) at the Oarai Research Establishment, is a prismatic-block HTGR dedicated to . With a thermal capacity of 30 MWth and no net electrical generation, it has been operational since achieving initial criticality in November 1998, reaching full-power tests at 950°C outlet temperature by 2004. As of 2025, the HTTR continues intermittent operations for validation, irradiation, and demonstrations, with restarts following regulatory approvals in 2021 and ongoing plans for facility expansions. Among decommissioned HTGRs, the Fort St. Vrain (FSV) plant in the United States, a prismatic-block design with a reactor vessel, operated from 1976 to 1989 as the only commercial-scale HTGR in . Rated at 842 MWth and 330 MWe, it faced challenges including helium circulator failures and leaks, achieving a lifetime below 20% before permanent shutdown in August 1989 due to economic factors. Decommissioning commenced in 1990, with fuel removal completed by 1992 and full site restoration achieved by 2011 using dry storage for spent fuel. The High-Temperature Reactor () in , a pebble-bed prototype, provided key operational data from its grid connection in 1985 until shutdown on September 1, 1989. With a thermal capacity of approximately 750 MWth and 300 MWe output, it utilized thorium-uranium fuel elements and operated for over 16,000 hours, generating about 1.67 billion kWh before closure due to financial and political pressures rather than technical failures. Decommissioning involved core defueling by 1995, transitioning to safe enclosure () status, with partial dismantling ongoing as of 2025 under German nuclear phase-out policies. Russia's VGR-50 prototype, a loop-type HTGR at the site, served as an early experimental platform integrating elements during its operational phases from 1972 to 1987. Rated at around 50 MWth, it focused on high-temperature testing and services without commercial power production, contributing to subsequent designs like the VGM modular concept before decommissioning in the late . Globally, HTGRs have accumulated limited operational experience as of , totaling several GW-years across historical and current facilities in , process heat testing, and demonstrations, though constrained by historical scale and regional programs.
ReactorCountryTypeThermal Power (MWth)Electrical Output (MWe)Operational PeriodKey Notes
Pebble-bed500 (twin modules)2102021–presentCommercial modular deployment; co-generation capable.
HTTRPrismatic300 (test)1998–presentR&D focus on high-temperature applications.
Fort St. VrainPrismatic8423301976–1989Decommissioned; economic challenges.
THTR-300Pebble-bed~7503001985–1989Decommissioned; fuel testing.
VGR-50Loop-type~500 ()1972–1987Decommissioned; integration tests.

Proposed and Under-Development Designs

China's HTR-PM600 project represents a significant expansion of the high-temperature gas-cooled reactor (HTGR) technology demonstrated by the initial units, featuring six 250 MWth pebble-bed modules connected to a single for a net capacity of 600 MWe. This design maintains the inherent safety features of the original while scaling up for commercial power generation and potential applications. The preliminary safety analysis report for the HTR-PM600 was submitted to regulators in , with ongoing reviews supporting commercialization efforts expected in the post-2030 timeframe. In the United States and , X-energy's Xe-100 is a pebble-bed HTGR (SMR) designed for 80 MWe per unit, with a standard four-unit configuration delivering 320 MWe for flexible deployment at industrial sites or power grids. The uses TRISO for high-temperature operation up to 750°C, enabling electricity production alongside process heat for applications like . A completed in September 2025 confirmed the Xe-100's suitability for repurposing coal sites in , with first-of-a-kind deployment targeted for the early 2030s through partnerships like the one with . In November 2025, X-energy announced a supply agreement for initial deployment and began vertical on its TRISO-X fabrication facility to support Xe-100 fueling. The Micro Modular Reactor (MMR), originally developed by Ultra Safe Nuclear Corporation, is a compact HTGR variant producing up to 15 or 35 MWth, emphasizing process heat for remote or industrial uses such as and , with a focus on TRISO fuel encapsulation for enhanced safety. Following USNC's in 2024, the MMR technology was acquired by NANO Nuclear Energy in December 2024 and rebranded as the MMR, with plans to advance demonstrations including a prototype at the and a U.S. Air Force contract. Although earlier proposals for an MMR deployment at Idaho National Laboratory were not pursued, the acquisition supports continued development toward operational prototypes in the late 2020s. As of 2025, the (IAEA) projects continued growth in advanced nuclear technologies, including HTGRs, with global nuclear capacity under construction reaching approximately 64 GW across all types and further expansions anticipated by 2030 to meet decarbonization goals; specific HTGR contributions are expected from ongoing projects in and , potentially adding several gigawatts in planning or early development phases.

Research and Innovations

Current R&D Efforts

In , the Institute of Nuclear and New Energy Technology (INET) at is conducting ongoing fuel qualification efforts for the reactor, including tests to validate performance at high levels exceeding 100 GWd/t, with recent analyses in 2024–2025 focusing on calibration and history dependencies to support extended fuel cycles. These tests build on prior qualifications and aim to confirm TRISO particle integrity under operational conditions up to 950°C, contributing to the reactor's commercial scalability. The Japan Atomic Energy Agency (JAEA) is advancing hydrogen production demonstrations using the High-Temperature Test Reactor (HTTR), with 2025 activities including preparations for integrating high-temperature helium gas at 950°C with steam methane reforming processes to produce clean hydrogen. This initiative, in collaboration with Mitsubishi Heavy Industries, targets operational tests by 2028 but involves 2025 regulatory applications and mock-up validations to demonstrate efficient heat transfer for decarbonized hydrogen generation. While primary focus is on steam reforming, exploratory synergies with high-temperature electrolysis are under consideration to enhance efficiency. Under the Euratom-funded SafeG project (2020–2024), research on the gas-cooled fast reactor demonstrator incorporates synergies with high-temperature gas-cooled reactor (HTGR) technologies, particularly in for high-temperature gas cooling environments. Key outcomes, reported in publications dated 2025, include evaluations of cladding and structural materials like composites to withstand coolant at temperatures above °C, drawing on HTGR expertise to improve resistance and thermal performance in gas-cooled systems. These efforts strengthen features such as removal, with project reports emphasizing cross-technology knowledge transfer for modular designs. The U.S. Department of Energy's Advanced Reactor Demonstration Program (ARDP) requested approximately $55 million in the FY 2026 budget (as of August 2025) for Advanced Reactor Technologies, including research and development in very high-temperature reactors (VHTRs)/HTGRs such as graphite and fuel qualification, with $10 million specifically for the Xe-100 HTGR demonstration and $58 million under Next Generation Fuels for TRISO fuel; this extends from 2025 planning and supports modeling via $28.6 million in Advanced Modeling and Simulation. FY2026 appropriations enacted in September 2025 provided $1.9 billion overall for the Office of , exceeding the request, though specific VHTR allocations remain consistent with the proposal as of November 2025. This funding supports TRISO fuel testing at high and multiphysics simulations for VHTRs, aiming to reduce deployment risks through virtual prototyping integrated with experimental data from facilities like the National Reactor Innovation Center. As of November 2025, reported continued advancements in the Xe-100 HTGR, focusing on fuel qualification and regulatory progress for 2027 deployment. The (IAEA) is leading a Coordinated Research Project (CRP) on modular high-temperature gas-cooled reactor safety design, active from 2024 through 2026, with a focus on multiphysics modeling to assess features under accident scenarios. Participants, including JAEA and INET, are developing integrated models for neutronics, thermal-hydraulics, and fuel behavior to propose updated safety criteria, with interim updates presented in early 2025 emphasizing in gas-cooled systems. This CRP facilitates international collaboration on validation benchmarks, ensuring robust computational tools for future HTGR deployments. In November 2025, Dutch engineering firm announced progress on adapting high-temperature gas-cooled reactor (HTGR) technology for small modular reactors (SMRs) in applications, targeting integration with industries to reduce global shipping emissions by up to 5% through high-temperature process heat for decarbonization.

Future Technological Advances

The Very High-Temperature Reactor (VHTR), a key Generation IV design, aims to achieve core outlet temperatures exceeding 950°C to enable high-efficiency and advanced process heat applications, with potential extensions beyond 1000°C in future iterations. This evolution supports closed fuel cycles incorporating thorium-uranium mixtures, enhancing resource utilization and waste reduction compared to traditional once-through approaches. Hybrid configurations integrating VHTR thermal spectra with fast-spectrum elements are also under conceptual exploration to optimize and fuel breeding. Advancements in structural materials, particularly silicon carbide fiber-reinforced silicon carbide (SiC/SiC) composites, are poised to enable VHTR operations at temperatures up to 1200°C by providing superior oxidation resistance and thermal stability in environments. These composites mitigate material degradation at elevated temperatures, potentially allowing reduced coolant pressures or purity requirements while maintaining core integrity. Integration of and for represents a transformative operational enhancement for HTGRs, with implementations anticipated between 2025 and 2035 to monitor component health in real-time and preempt failures. Such systems leverage sensor data to optimize maintenance schedules, drawing on ongoing demonstrations in advanced reactor prototypes. Proliferation-resistant features in future HTGR designs emphasize once-through fuel cycles using low-enriched , where buildup remains below 5% of heavy metal content due to high rates exceeding 100 GWd/t, rendering extracted material less suitable for weapons. The TRISO-coated particle architecture further complicates diversion by encasing in durable matrices. Market projections indicate substantial growth for HTGRs within the sector, with global investments forecasted to reach several billion dollars by the early 2030s, driven by demand for high-temperature heat in decarbonized industries. By 2040, HTGRs are expected to capture a notable share of SMR deployments, supported by international R&D collaborations targeting commercial viability.

References

  1. [1]
    Very High Temperature Reactor (VHTR) | GIF Portal
    The High-Temperature Gas-Cooled Reactor (HTGR) has a history spanning over 50 years, marked by successes and challenges. The initial experimental plants ...
  2. [2]
    [PDF] Gas-cooled Reactors and Industrial Heat Applications
    Jun 16, 2022 · The high-temperature gas-cooled reactor (HTGR) is a helium-cooled graphite-moderated nuclear fission reactor technology that uses fully ceramic ...<|control11|><|separator|>
  3. [3]
    What is HTGR ?
    Outline of High Temperature Engineering Test Reactor · Establishment of technical bases of high temperature gas-cooled reactor with the HTTR · Various Hydrogen ...
  4. [4]
    [PDF] Chapter 4 THE HIGH TEMPERATURE GAS COOLED REACTOR ...
    A reactor core design which ensures that the maximum fuel element temperature limit cannot be exceeded in any accident. The reactor and the steam generator are ...
  5. [5]
    [PDF] High Temperature Gas Cooled Reactor Fuels and Materials
    This document broadly covers several aspects of coated particle fuel technology, namely: manufacture of coated particles, compacts and elements; design-basis; ...
  6. [6]
    [PDF] High Temperature Gas- cooled Reactors: Core Design
    Combined correlations have been developed to capture heat transfer in pebble bed core via conduction, convection and radiation. ... Investigation of local Heat ...
  7. [7]
    [PDF] Heat Transport and Afterheat Removal for Gas Cooled Reactors ...
    The scope includes experimental and analytical investigations of heat transport by natural convection conduction and thermal radiation within the core and ...
  8. [8]
    [PDF] Parametric Investigation of Brayton Cycle for High Temperature Gas ...
    This paper includes preliminary calculations of the steady state overall Brayton cycle efficiency based on the pebble bed reactor reference design (helium used ...
  9. [9]
    [PDF] Current status and future development of modular high temperature ...
    During the 1980s, the modular high temperature gas cooled reactor (HTGR) concept was developed, primarily in Germany and the United States of America.
  10. [10]
    [PDF] HTGR Technology Family Assessment for a Range of Fuel Cycle ...
    Aug 23, 2010 · Third, the solid moderator in an HTGR (graphite) represents more of a waste management issue than the liquid moderator in an LWR (water).
  11. [11]
    [PDF] SAM Code Validation on Frictional Pressure Drop through Pebble ...
    predict frictional pressure drop through porous pebble beds: (1) the classical Ergun correlation;. (2) the KTA correlation, which is widely used in pebble ...
  12. [12]
    [PDF] Role of Nuclear Grade Graphite in Controlling Oxidation in Modular ...
    The reflector dimensions in modular HTGR cores are sized to obtain optimal neutron economy. (i.e., to maximize neutron thermalization and backscatter into ...
  13. [13]
    [PDF] High Temperature Gas Cooled Reactor HTGR Training Materials
    Jul 17, 2019 · Forced and mixed convection heat transfer at high pressure and high temperature in a graphite flow channel. J. Heat Transfer, 140, pp.
  14. [14]
    Small modular high temperature reactor optimisation – Part 1: A ...
    This article aims to provide a comparison between Beryllium oxide and nuclear graphite as a neutron reflector for high temperature reactor.Missing: HTGR | Show results with:HTGR
  15. [15]
    [PDF] Appendices to the Assessment of Graphite Properties and ...
    “Dimensional Change, Irradiation Creep and Thermal/Mechanical Property Changes in Nuclear. Graphite,” International Materials Reviews, 61(3):155–182, 2016 ...
  16. [16]
    [PDF] NUCLEAR GRAPHITE FOR HIGH TEMPERATURE REACTORS B.J. ...
    Abstract: The cores and reflectors in modern High Temperature Gas Cooled Reactors (HTRs) are constructed from graphite components. There are two main designs; ...
  17. [17]
    [PDF] Review of Graphite Properties Relevant to Micro-Reactor Design
    Nov 1, 2021 · 2.2.3 Thermal Conductivity. Reference 8 provides a correlation of burn-off and thermal conductivity for IG-110 stratified by sample orientation.
  18. [18]
    None
    Below is a merged summary of HTGR fuel information from IAEA-TECDOC-1645, consolidating all details from the provided segments into a comprehensive response. To maximize density and clarity, key information is presented in tables where appropriate, followed by narrative summaries for aspects that are better suited to prose. All unique details from each segment are retained, with references to page numbers or sections where available.
  19. [19]
    [PDF] triso-coated fuel particle performance
    The fuel in an HTGR core is contained in billions of coated particles, each of which acts as its own containment. The small kernels of fuel are each coated with ...
  20. [20]
    [PDF] Coated Particle Fuels for High Temperature Gas Cooled Small ...
    The main advantage of TRISO fuel is its capability to withstand extreme conditions like high temperature and prolonged irradiation. Research programmes have ...
  21. [21]
    [PDF] Fuel-Cycle and Nuclear Material Disposition Issues Associated with ...
    6) MEU cycle: U-235 enriched to 20% (max) with thorium. Highly proliferating cycle, with incentive to recycle U-233. Some of the issues associated with HTGR ...
  22. [22]
    Nuclear Fuel Cycle Overview
    Sep 23, 2025 · Burn-up in GWd/t is the conventional measure for oxide fuels, and 60 GWd/t U is equivalent to about 6.5 atomic percent burn-up (i.e. about 6.5% ...Missing: HTGR equation
  23. [23]
    [PDF] Benchmark Specification for HTGR Fuel Element Depletion.
    HTGR designs utilize graphite-moderated fuel forms and helium gas as a coolant. There are two main forms of HTGR fuels: pebbles are used in the Pebble Bed.Missing: void | Show results with:void
  24. [24]
    [PDF] TRISO-Coated Particle Fuel Phenomenon Identification and ...
    May 3, 2015 · The Booth diffusion model has been used to estimate the release of fission gases via these mechanisms and has been used to describe fission.
  25. [25]
    None
    Below is a merged summary of the HTGR (High Temperature Gas-cooled Reactor) core principles, neutronics, graphite moderation, and equations, consolidating all information from the provided segments into a single, comprehensive response. To maximize detail and clarity, I’ve organized the content into sections with tables where appropriate (e.g., for equations and key data). The response retains all mentioned information while avoiding redundancy and ensuring a dense, structured format.
  26. [26]
    helium - the NIST WebBook
    Gas Phase Heat Capacity (Shomate Equation) ; 800. 20.79, 146.7 ; 900. 20.79, 149.1 ; 1000. 20.79, 151.3 ; 1100. 20.79, 153.3 ...
  27. [27]
    [PDF] High-Temperature Gas- Cooled Test Reactor Point Design
    The lower outlet temperature was selected to ensure sufficient thermal margin under normal operating conditions to prevent melting of ...
  28. [28]
    [PDF] Next Generation Nuclear Plant Intermediate Heat Exchanger ...
    The HTGR concept is considered to be the nearest-term reactor design that has the capability to efficiently produce hydrogen. The plant size, reactor thermal ...Missing: sodium | Show results with:sodium
  29. [29]
    [PDF] Guidance for Developing Principal Design Criteria for Advanced ...
    Two IHXs are located in the reactor vessel. These are connected to one intermediate heat transport loop which contains sodium as the heat transport fluid. The ...
  30. [30]
    [PDF] NGNP Reactor Coolant Chemistry Control Study
    Nov 1, 2010 · The corrosion rate in the HTGR is heavily dependent on operating temperatures, the impurity levels of the helium coolant, and the material ...
  31. [31]
    [PDF] DETERMINATION OF THE SAFETY ROD DROP TIME IN THE ...
    The time required to insert a safety rod after a manual scram ranged from 588 to 658 msec; the predicted time was 720 msec.
  32. [32]
    [PDF] seismic scrammanuty of httr control rods - INIS-IAEA
    Especially, accurate scram time estimation is very important for assuring reactor safety. The High Temperature Engineering Test Reactor (HTTR), a. HTGR of Japan ...
  33. [33]
    Feasibility study on burnable poison credit concept to HTGR fuel ...
    As a BP, boron, gadolinium, erbium, and hafnium are investigated. As a result, it is found that boron and gadolinium suit this concept, and the 14 wt% fuel can ...
  34. [34]
    The use of burnable absorbers integrated into TRISO/QUADRISO ...
    Dec 1, 2021 · Regarding fuel type, the HTGR employs two main types: prismatic block and pebble-bed types. Both types are moderated by graphite and fuelled by ...
  35. [35]
    [PDF] HTGR reactor physics and fuel cycle studies
    Nevertheless, the temperature reactivity coefficients of the reactor (both Doppler and moderator coefficient) were found to be sufficiently negative over the ...
  36. [36]
    Full article: Small, long-life high temperature gas-cooled reactor free ...
    On the other hand, if the reactor operates with high positive reactivity, equal to or greater than 1$ (ρ ≥ 1$), it will be uncontrollable as there will be a ...
  37. [37]
    [PDF] General Atomics, Transmittal of Gas Turbine-Modelar Helium ...
    The reserve shutdown control material consists of 40 % wt natural boron in B4C granules dispersed in graphite matrix and formed into pellets. The reserve ...
  38. [38]
    [PDF] IAEA Coordinated Research Project on HTGR Physics, Thermal ...
    Should the control rods become inoperable, a backup reserve shutdown control (RSC) is provided in the form of borated pellets that may be released into channels ...Missing: soluble | Show results with:soluble
  39. [39]
    [PDF] Chapter 13.qxd - Idaho National Laboratory
    Aircraft Nuclear Propulsion crew poses in front of Heat Transfer Reactor Experiment No. 1. INEEL 89-79. Page 3. southeast Idaho in July 1952 that a nuclear ...
  40. [40]
    [PDF] When Nuclear Power Was Flying High - LeeHite.org
    Between 1946 and 1961, the Air Force and the Atomic Energy Commission spent a billion dollars in an effort to develop nuclear-powered airplanes. Although.
  41. [41]
    Dragon Project - Nuclear Energy Agency (NEA)
    The Dragon was the first experimental high temperature gas-cooled reactor (HTGR) built in the 1960s. Dragon used helium gas coolant and graphite as the neutron ...Missing: 1957-1976 | Show results with:1957-1976
  42. [42]
    peach bottom-1 - PRIS - Reactor Details
    Reference Unit Power (Net Capacity), Design Net Capacity, Gross Capacity, Thermal Capacity. 40 MWe. 40 MWe. 42 MWe. 115 MWt. Construction Start Date, First ...Missing: output MWth<|separator|>
  43. [43]
    [PDF] Peach Bottom Atomic Power Station, Unit 1
    May 13, 2024 · PBAPS-1 was the first prototype installation of a High-Temperature Gas-Cooled Reactor. (HTGR) in the United States under the Atomic Energy ...
  44. [44]
    AVR Juelich - World Nuclear Association
    AVR Juelich was a High Temperature Gas-Cooled Reactor (HTGR) pebble bed prototype, with a net capacity of 13 MWe, and permanently shut down in 1988.
  45. [45]
    [PDF] The Importance of the AVR Pebble-Bed Reactor For the Future ...
    The A VR pebble-bed high temperature gas-cooled reactor (HTGR ) at Juelich,. Germany operated from 1967 to 1988 and was certainly the most important HTGR.
  46. [46]
    [PDF] GCRA 89-001, Summary of Gas-Cooled Reactor Programs.
    Jan 26, 1989 · The USSR has had an ongoing HTGR research and development program for many years. Most of the work is performed at the. Kurchatov Institute in ...
  47. [47]
    [PDF] The status of high-temperature gas-cooled reactor development and ...
    Most of the early development centered on low-temperature systems using graphite moderator, metal-clad fuel and carbon-dioxide coolant. Commercial deployment of ...Missing: Soviet Union
  48. [48]
    [PDF] Fort Saint Vrain Gas Cooled Reactor Operational Experience
    This study provides an analysis of the operational experience history of FSV as depicted in the monthly FSV operating experience reports produced by ORNL. It ...
  49. [49]
    [PDF] gas cooled reactor decommissioning, fuel storage and waste disposal
    The pebble bed high temperature reactor THTR 300 was shutdown on. 01.09.89 after more than 16,000 h in operation. The THTR 300 is a prototype reactor project ...
  50. [50]
    [PDF] THTR 300 MWe Prototype Reactor - Safety Assessment 1. Main ...
    Sep 1, 1989 · Main incidents and problems. The operation of the THTR 300 showed some incidents and problems: * Difficulties with the refuelling system ...
  51. [51]
    THTR-300 - World Nuclear Association
    THTR-300 was a High Temperature Gas-Cooled Reactor (HTGR) pebble bed reactor, with a net capacity of 296 MWe, and permanently shut down on 29 September 1988.Missing: incidents | Show results with:incidents
  52. [52]
    [PDF] JAEA-Technology-2009-063.pdf
    The HTTR achieved the first criticality in 1998, the reactor outlet coolant temperature of 950°C in 2004, and 30 days continuous operation in 2007. Since 2002, ...
  53. [53]
    [PDF] High Temperature (Gas Cooled) Reactors IAEA Activities
    Sep 5, 2016 · Japan. • HTTR completed of continuous 50-day high temperature. (950oC) operation in March 2010. • Several safety demonstrations done and ...
  54. [54]
    Small Nuclear Power Reactors - World Nuclear Association
    Financial constraints led to delays and in September 2010 the South African government confirmed it would stop funding the project and closed it down. The ...
  55. [55]
    Nuclear Power in South Africa
    The key project is construction of new nuclear power plants with the ... In March 2010, the government drastically cut funding for the PBMR, then in ...
  56. [56]
    South Africa 2022
    In 2010 the South African government announced the closure of the PBMR company largely due to financial constraints and inability to secure customers. The ...
  57. [57]
    Nuclear Power in China
    The first of the twin reactors of the demonstration HTR-PM achieved first criticality in September and the second in November 2021. The HTR-PM 600 reactor units ...
  58. [58]
    Coated particle fuel: Historical perspectives and current progress ...
    ... fission product retention under extremely severe conditions, including temperatures of 1600 °C for hundreds of hours. The fuel constitutes one of the key ...
  59. [59]
    [PDF] Safety related design and economic aspects of HTGRs
    This advanced nuclear power reactor has the capability to provide high temperature energy for the generation of electricity and for industrial process heat ...<|separator|>
  60. [60]
    [PDF] OOUF- 700117- -3 - OSTI.gov
    The AVR LOCA test was performed to demonstrate that inherent characteristics of small HTGRs enable them to withstand highly unlikely LOCA conditions in which ...
  61. [61]
    Use of DRACS to Enhance HTGRs Passive Safety and Economy
    all the HTGR designs use Reactor Vessel Auxiliary Cooling System (RVACS) for passive decay heat removal. (1) The decay heat first is transferred to core ...
  62. [62]
    [PDF] Advanced nuclear plant design options to cope with external events
    ▫ Emergency heat sinks outside the pre-stressed concrete vessel (PCV), i.e. emergency tanks & heat exchangers. 4.4. Inherent safety features: ▫ Self ...Missing: prestressed | Show results with:prestressed
  63. [63]
    [PDF] IAEA Nuclear Energy Series Design Features to Achieve Defence in ...
    • Soluble boron free core design eliminates boron dilution accidents. Level ... • Passive shutdown through injection of poison to the moderator, using ...
  64. [64]
    [PDF] the seventh national report under the convention on nuclear safety
    distributed pattern can avoid hydrogen gathering and explosion in the containment. After the Fukushima nuclear accident, safety improvements were made as.
  65. [65]
    [PDF] IAEA TECDOC SERIES
    Section 3 discusses various designs of the engineered safety features of advanced reactors and SMRs which includes diverse trip system, residual heat removal ...
  66. [66]
    [PDF] Literature Review of Advanced Reactor Cost Estimates
    The review collected information relating to estimated overnight costs, Operation & Maintenance (O&M or OPEX), and levelized costs of electricity (LCOE).
  67. [67]
    [PDF] Advanced Nuclear Process Heat for Industrial Decarbonization
    For instance, the temperature out- put of a High-Temperature Gas-Cooled Reactor (HTGR) can reach 950 degrees C (1,742 degrees C).28 This means that this ...
  68. [68]
    Powering Heavy Industry & Manufacturing with Advanced Nuclear ...
    Jul 18, 2025 · X-energy's Xe-100 High-Temperature Gas-cooled Reactor (HTGR) is designed to deliver 565°C steam for industrial applications, making it an ideal ...
  69. [69]
    [PDF] Hybrid sulfur cycle flowsheets for hydrogen production using high ...
    One flowsheet can be used with a 950°C reactor to make hydrogen at 44.0% to 47.6% net thermal efficiency, HHV basis. • The other flowsheet can be used with a ...
  70. [70]
    Progress of nuclear hydrogen production through the iodine–sulfur ...
    This paper presents the progress of nuclear hydrogen production through the iodine–sulfur (IS) process over the past 10 years.Missing: MWth | Show results with:MWth
  71. [71]
    Economic Feasibility of Hydrogen Generation Using HTR-PM ... - MDPI
    This study examines two advanced hydrogen production methods: high-temperature steam electrolysis (HTSE) and the sulfur–iodine (SI) thermochemical cycle. Both ...<|separator|>
  72. [72]
    Nuclear Process Heat for Industry
    In 2019 there were 79 nuclear reactors used for desalination, district heating, or process heat, with 750 reactor-years of experience in these, mostly in Russia ...
  73. [73]
    [PDF] Guidance on Nuclear Energy Cogeneration
    Low temperature heat applications such as district heating and seawater desalination require nuclear heat in the range of 230–250°C. Currently operating water- ...
  74. [74]
    [PDF] Nuclear non-electric applications are a
    Nuclear hydrogen production by steam reforming (saving 35% natural gas/CO2) can be coupled to HTTR for test today ! 15. Steam reforming H2 production plan on ...
  75. [75]
    [PDF] Gas-cooled Reactors and Industrial Heat Applications - OECD
    Some HTGR designs are also scalable in multi-unit installations (NEA, 2021). Figure 2.4: Life cycle CO2 emissions in g/kWh of different electricity sources.
  76. [76]
    Gas Cooled Reactor Market Report | Global Forecast From 2025 To ...
    The global gas-cooled reactor market size was estimated at approximately $5.6 billion in 2023 and is projected to reach nearly $9.8 billion by 2032, ...
  77. [77]
    [PDF] GOV/INF/2025/8-GC(69)/INF/4 - International Atomic Energy Agency
    Aug 25, 2025 · China brought a modular high temperature gas cooled reactor (HTR-PM) into commercial operation in December 2023, generating 200 MW of ...
  78. [78]
    Testing the feasibility of multi-modular design in an HTR-PM nuclear ...
    Mar 21, 2025 · The HTR-PM power plant is situated in the Shidao Bay area of China Shandong province, has been integrated to the grid since December 20, 2021, ...
  79. [79]
    Application for changes to reactor installation of the High ...
    Mar 27, 2025 · JAEA plans to establish a hydrogen production facility connected to the HTTR and to confirm the hydrogen production technology by directly using ...
  80. [80]
    Japan 2018
    The initial criticality of the HTTR was achieved in November 1998. The HTTR attained full power operation of 30 MWth and a gas temperature of 950°C at the ...
  81. [81]
    [PDF] Licensing Experiences of HTTR - Nuclear Regulatory Commission
    ▫ HTTR restarted its operation on July 30, 2021. E. Ishitsuka, “Experience of HTTR Licensing for Japan's New Nuclear Regulation,” GIF Webinar Series 52 ...
  82. [82]
    [PDF] "Preliminary Decommissioning Plan for Fort St Vrain Nuclear ...
    that in light of its operating history, Fort St. Vrain was not "used and useful" in rendering a utility service. The OCC sought to remove. Fort St. Vrain ...
  83. [83]
    [PDF] Decommissioning of the Thorium High Temperature Reactor (THTR ...
    I. Introduction. The pebble bed high temperature reactor THTR 300 was shutdown on. 01.09.89 after more than 16,000 h in operation. The THTR 300 is a ...
  84. [84]
    Decommissioning of the thorium high temperature reactor (THTR 300)
    Jan 15, 2025 · The prototype Thorium-High-Temperature-Reactor (THTR 300) was decommissioned using the option of safe enclosure. Decision was made in 1989.
  85. [85]
    [PDF] Update of HTR R&D Progress in China
    Mar 21, 2023 · HTR-PM600. ◇Six reactor modules connect to one steam turbine, comparing with HTR-. PM demonstration plant: ◇ the same safety features ...Missing: expansion | Show results with:expansion<|separator|>
  86. [86]
    IAEA Convenes Global Experts to Assesses High Temperature Gas ...
    Dec 22, 2021 · After the HTR-PM demonstration operation, China expects to move on to commercialization sometime after 2030 of the HTR-PM600, which will consist ...Missing: timeline | Show results with:timeline
  87. [87]
    Xe-100: High-Temperature Gas-Cooled Nuclear Reactors (HTGR)
    Xe-100 is a 80 MWe reactor that can be scaled into a 'four-pack' 320 MWe power plant—with our modular design, the scale can grow even larger as needed. Our base ...Missing: 2028-2030 | Show results with:2028-2030
  88. [88]
    X-energy Confirms Feasibility of Xe-100 Advanced Small Modular ...
    Sep 25, 2025 · X-energy's Xe-100 SMR is an 80-megawatt (“MW”) high-temperature gas-cooled reactor uniquely suited to Alberta's energy needs. In addition to ...Missing: MWe four- pack 320 2028-2030
  89. [89]
    Study confirms feasibility of Xe-100 SMR for Alberta
    Sep 26, 2025 · The study by X-energy Canada has confirmed the feasibility and benefits of repurposing an existing thermal generation site in Alberta with ...Missing: 80 MWe four- pack 320 2028-2030
  90. [90]
    NANO Nuclear acquires Micro Modular Reactor (MMR®) and Pylon ...
    Dec 24, 2024 · The system, which is under development, could be used to provide carbon-free, high-quality process heat for co-located industrial applications, ...Missing: MWe | Show results with:MWe
  91. [91]
    NANO announces MMR rebrand on completion of USNC technology ...
    Jan 14, 2025 · NANO Nuclear Inc's newly completed USD85 million acquisition of Ultra Safe Nuclear Corporation's nuclear technology will now be known as the KRONOS MMR.
  92. [92]
    Idaho National Laboratory Demonstration nuclear power plant ...
    Status, Nameplate capacity, Reactor type, Model, Owner. Cancelled (inferred June 2025), 15 MW, Gas-cooled reactor, MMR, Ultra Safe Nuclear Corp. Location.
  93. [93]
    [PDF] Hermes CP Approved Safety Evaluation
    The test reactor will support development of Kairos's fluoride salt-cooled, high-temperature reactor technology. The Hermes reactor will be a 35-megawatt ...
  94. [94]
    Technology - Kairos Power
    The intrinsic low pressure in Kairos Power reactors enhances safety and eliminates the need for bulky and expensive high-pressure containment structures. Kairos ...Missing: HTGR influences
  95. [95]
    IAEA increases nuclear growth projections - World Nuclear News
    Sep 15, 2025 · In addition to the high case scenario, the IAEA has also increased its low case projection with a 50% capacity increase to 561 GW by 2050. One ...Missing: HTGR 5-10
  96. [96]
    [PDF] Nuclear Technology Review 2025
    At the end of 2024, the 70 years of worldwide cumulative operating experience amounted to over 20166 reactor-years, provided by 653 reactors with a total ...
  97. [97]
    Burnup calibration and its dependence on irradiation history in high ...
    Aug 9, 2025 · HTR-PM fuel pebble irradiation qualification in the high flux reactor in Petten ... fuel qualification will be performed, namely heating tests ...
  98. [98]
    HTR-PM fuel pebble irradiation qualification in the high flux reactor ...
    Apr 1, 2018 · For qualification, the pebbles were required to reach central temperatures of 1050 ± 50 °C, with a pebble surface temperature as homogeneous as ...
  99. [99]
    JAEA, MHI team up for HTTR hydrogen project - World Nuclear News
    Apr 25, 2022 · The Japan Atomic Energy Agency and Mitsubishi Heavy Industries will establish a demonstration hydrogen production project at the High-Temperature Test Reactor ...Missing: electrolysis | Show results with:electrolysis<|separator|>
  100. [100]
    [PDF] Development of High Temperature Gas-cooled Reactor for ...
    For this reason, JAEA is promoting an HTTR heat application test in connection with the HTTR, which has achieved a reactor coolant outlet temperature of 950°C, ...
  101. [101]
    Safety of GFR through innovative materials, technologies ... - CORDIS
    Sep 22, 2025 · The EU-funded SafeG project aims to connect developers of the ALLEGRO reactor with experts in GFRs and high-temperature reactors.Missing: HTGR synergies
  102. [102]
    Outcomes of the Euratom SafeG Project on ALLEGRO Research ...
    The paper describes R&D activities suported by Euratom funded projects (SafeG, TREASURE) focused on advancing the development of the Gas-cooled Fast Reactor ( ...Missing: HTGR | Show results with:HTGR
  103. [103]
    SafeG Project
    The main objectives of the SafeG project are: To strengthen safety of the GFR demonstrator ALLEGRO; To review the GFR reference options in materials and ...Missing: HTGR synergies 2025
  104. [104]
    [PDF] FY 2026 Congressional Justification - Department of Energy
    The FY 2026 Budget Request provides $50 million to fund research, development, and ... FY 2026 funding will also support development of digital twin models ...Missing: VHTR | Show results with:VHTR
  105. [105]
    Advanced Reactor Demonstration Program | Department of Energy
    ARDP will provide $160 million in initial funding, available through the ARD Funding Opportunity Announcement . Applicants can receive support through three ...Advanced Reactor... · DOE Awards $20 million for... · Energy Department’s...Missing: VHTR | Show results with:VHTR
  106. [106]
    Modular High Temperature Gas Cooled Reactor Safety Design | IAEA
    ... reactors known for their inherent and passive safety features, such as the absence of core melt scenarios and minimal reliance on active safety systems ...
  107. [107]
    Update on the status of the CRP on Modular HTGR Safety Design
    Jan 17, 2025 · The specific outcomes as defined: 1) Definition of Modular HTGRs. 2) Discussion of the inherent or intrinsic characteristics of the fuel and ...Missing: 2024-2026 multiphysics
  108. [108]
    The IAEA coordinated research project on modular HTGR safety ...
    Nov 15, 2016 · Title: The IAEA coordinated research project on modular HTGR safety design ... physics models and robust, efficient, and accurate codes. One way ...Missing: 2024-2026 | Show results with:2024-2026
  109. [109]
    [PDF] Thorium fuel cycle — Potential benefits and challenges
    In Germany, two Pebble Bed HTGRs, namely AVR 15 MW(e) and THTR 300. MW(e), successfully operated till the late 1980s after which they were terminated. In Pebble.<|separator|>
  110. [110]
    Design strategies for LWR and VHTR test environments in a fast ...
    A fast spectrum is especially advantageous because proper use of moderators and absorbers can tailor the flux to any spectrum softer than the original fast ...
  111. [111]
    SiC/SiC Composites for 1200 C and Above
    Nov 1, 2004 · This chapter presents information concerning processes and properties for five silicon carbide (SiC) fiber-reinforced SiC matrix composite systems recently ...
  112. [112]
    The critical issues of SiC materials for future nuclear systems
    A more promising choice is SiC/SiC composites, which permit an outlet temperature maximum of ~ 1000 °C. SiC/SiC composites also appear to be very promising ...
  113. [113]
    AI-Enabled Predictive Maintenance Digital Twins for Advanced ...
    Project Description. Advanced reactors must be designed to be financially competitive with fossil fuel power plants to gain a foothold in future energy markets.Missing: machine 2025-2035 implementations
  114. [114]
  115. [115]
    [PDF] Very High-Temperature Reactor (VHTR) Proliferation Resistance ...
    The Russians are simultaneously pursuing the GT-MHR as a “deep-burn” option for weapon- grade plutonium disposition [9]. The use of high enriched uranium (HEU) ...
  116. [116]
    High Temperature Gas Cooled Reactor Analysis 2025-2033 ...
    The High Temperature Gas-Cooled Reactor (HTGR) market is poised for significant growth, driven by increasing demand for advanced nuclear energy solutions ...
  117. [117]
    High-Temperature Gas-Cooled Reactor Market Research Report 2033
    The market is expected to expand at a CAGR of 8.1% from 2025 to 2033, with the forecasted market size projected to reach USD 2.87 billion by 2033. This growth ...Reactor Type Analysis · Coolant Type Analysis · Regional Outlook