Liquid fluoride thorium reactor

Last updated

Liquid FLiBe salt FLiBe.png
Liquid FLiBe salt

The liquid fluoride thorium reactor (LFTR; often pronounced lifter) is a type of molten salt reactor. LFTRs use the thorium fuel cycle with a fluoride-based molten (liquid) salt for fuel. In a typical design, the liquid is pumped between a critical core and an external heat exchanger where the heat is transferred to a nonradioactive secondary salt. The secondary salt then transfers its heat to a steam turbine or closed-cycle gas turbine. [1]

Contents

Molten-salt-fueled reactors (MSRs) supply the nuclear fuel mixed into a molten salt. They should not be confused with designs that use a molten salt for cooling only (fluoride high-temperature reactors, FHRs) and still have a solid fuel. [2] Molten salt reactors, as a class, include both burners and breeders in fast or thermal spectra, using fluoride or chloride salt-based fuels and a range of fissile or fertile consumables. LFTRs are defined by the use of fluoride fuel salts and the breeding of thorium into uranium-233 in the thermal neutron spectrum.

The LFTR concept was first investigated at the Oak Ridge National Laboratory Molten-Salt Reactor Experiment in the 1960s, though the MSRE did not use thorium. The LFTR has recently been the subject of a renewed interest worldwide. [3] Japan, China, the UK and private US, Czech, Canadian [4] and Australian companies have expressed the intent to develop, and commercialize the technology.

LFTRs differ from other power reactors in almost every aspect: they use thorium that is turned into uranium, instead of using uranium directly; they are refueled by pumping without shutdown. [5] Their liquid salt coolant allows higher operating temperature and much lower pressure in the primary cooling loop. These distinctive characteristics give rise to many potential advantages, as well as design challenges.

Background

Thorium is relatively abundant in the Earth's crust. NAMrad Th let.gif
Thorium is relatively abundant in the Earth's crust.
Tiny crystals of thorite, a thorium mineral, under magnification. Amr-thorite.jpg
Tiny crystals of thorite, a thorium mineral, under magnification.
Molten salt reactor at Oak Ridge MSRE Reactor.JPG
Molten salt reactor at Oak Ridge

By 1946, eight years after the discovery of nuclear fission, three fissile isotopes had been publicly identified for use as nuclear fuel: [6] [7]

Th-232, U-235 and U-238 are primordial nuclides, having existed in their current form for over 4.5 billion years, predating the formation of the Earth; they were forged in the cores of dying stars through the r-process and scattered across the galaxy by supernovas. [9] Their radioactive decay produces about half of the Earth's internal heat. [10]

For technical and historical [11] reasons, the three are each associated with different reactor types. U-235 is the world's primary nuclear fuel and is usually used in light water reactors. U-238/Pu-239 has found the most use in liquid sodium fast breeder reactors and CANDU Reactors. Th-232/U-233 is best suited to molten salt reactors (MSR). [12]

Alvin M. Weinberg pioneered the use of the MSR at Oak Ridge National Laboratory. At ORNL, two prototype molten salt reactors were successfully designed, constructed and operated. These were the Aircraft Reactor Experiment in 1954 and Molten-Salt Reactor Experiment from 1965 to 1969. Both test reactors used liquid fluoride fuel salts. The MSRE notably demonstrated fueling with U-233 and U-235 during separate test runs. [13] :ix Weinberg was removed from his post and the MSR program closed down in the early 1970s, [14] after which research stagnated in the United States. [15] [16] Today, the ARE and the MSRE remain the only molten salt reactors ever operated.

Breeding basics

In a nuclear power reactor, there are two types of fuel. The first is fissile material, which splits when hit by neutrons, releasing a large amount of energy and also releasing two or three new neutrons. These can split more fissile material, resulting in a continued chain reaction. Examples of fissile fuels are U-233, U-235 and Pu-239. The second type of fuel is called fertile. Examples of fertile fuel are Th-232 (mined thorium) and U-238 (mined uranium). In order to become fissile these nuclides must first absorb a neutron that's been produced in the process of fission, to become Th-233 and U-239 respectively. After two sequential beta decays, they transmute into fissile isotopes U-233 and Pu-239 respectively. This process is called breeding. [5]

All reactors breed some fuel this way, [17] but today's solid fueled thermal reactors don't breed enough new fuel from the fertile to make up for the amount of fissile they consume. This is because today's reactors use the mined uranium-plutonium cycle in a moderated neutron spectrum. Such a fuel cycle, using slowed down neutrons, gives back less than 2 new neutrons from fissioning the bred plutonium. Since 1 neutron is required to sustain the fission reaction, this leaves a budget of less than 1 neutron per fission to breed new fuel. In addition, the materials in the core such as metals, moderators and fission products absorb some neutrons, leaving too few neutrons to breed enough fuel to continue operating the reactor. As a consequence they must add new fissile fuel periodically and swap out some of the old fuel to make room for the new fuel.

In a reactor that breeds at least as much new fuel as it consumes, it is not necessary to add new fissile fuel. Only new fertile fuel is added, which breeds to fissile inside the reactor. In addition the fission products need to be removed. This type of reactor is called a breeder reactor. If it breeds just as much new fissile from fertile to keep operating indefinitely, it is called a break-even breeder or isobreeder. A LFTR is usually designed as a breeder reactor: thorium goes in, fissile products come out.

Reactors that use the uranium-plutonium fuel cycle require fast reactors to sustain breeding, because only with fast moving neutrons does the fission process provide more than 2 neutrons per fission. With thorium, it is possible to breed using a thermal reactor. This was proven to work in the Shippingport Atomic Power Station, whose final fuel load bred slightly more fissile from thorium than it consumed, despite being a fairly standard light water reactor. Thermal reactors require less of the expensive fissile fuel to start, but are more sensitive to fission products left in the core.

There are two ways to configure a breeder reactor to do the required breeding. One can place the fertile and fissile fuel together, so breeding and splitting occurs in the same place. Alternatively, fissile and fertile can be separated. The latter is known as core-and-blanket, because a fissile core produces the heat and neutrons while a separate blanket does all the breeding.

Reactor primary system design variations

Oak Ridge investigated both ways to make a breeder for their molten salt breeder reactor. Because the fuel is liquid, they are called the "single fluid" and "two fluid" thorium thermal breeder molten salt reactors.

Single fluid reactor

Simplified schematic of a single fluid reactor. MSRCutAway.png
Simplified schematic of a single fluid reactor.

The one-fluid design includes a large reactor vessel filled with fluoride salt containing thorium and uranium. Graphite rods immersed in the salt function as a moderator and to guide the flow of salt. In the ORNL MSBR (molten salt breeder reactor) design [18] a reduced amount of graphite near the edge of the reactor core would make the outer region under-moderated, and increased the capture of neutrons there by the thorium. With this arrangement, most of the neutrons were generated at some distance from the reactor boundary, and reduced the neutron leakage to an acceptable level. [19] Still, a single fluid design needs a considerable size to permit breeding. [20]

In a breeder configuration, extensive fuel processing was specified to remove fission products from the fuel salt. [13] :181 In a converter configuration fuel processing requirement was simplified to reduce plant cost. [19] The trade-off was the requirement of periodic uranium refueling.

The MSRE was a core region only prototype reactor. [21] The MSRE provided valuable long-term operating experience. According to estimates of Japanese scientists, a single fluid LFTR program could be achieved through a relatively modest investment of roughly 300–400 million dollars over 5–10 years to fund research to fill minor technical gaps and build a small reactor prototype comparable to the MSRE. [22]

Two fluid reactor

The two-fluid design is mechanically more complicated than the "single fluid" reactor design. The "two fluid" reactor has a high-neutron-density core that burns uranium-233 from the thorium fuel cycle. A separate blanket of thorium salt absorbs neutrons and slowly converts its thorium to protactinium-233. Protactinium-233 can be left in the blanket region where neutron flux is lower, so that it slowly decays to U-233 fissile fuel, [23] rather than capture neutrons. This bred fissile U-233 can be recovered by injecting additional fluorine to create uranium hexafluoride, a gas which can be captured as it comes out of solution. Once reduced again to uranium tetrafluoride, a solid, it can be mixed into the core salt medium to fission. The core's salt is also purified, first by fluorination to remove uranium, then vacuum distillation to remove and reuse the carrier salts. The still bottoms left after the distillation are the fission products waste of a LFTR.

The advantages of separating the core and blanket fluid include:

  1. Simpler fuel processing. Thorium is chemically similar to several fission products, called lanthanides. With thorium in a separate blanket, thorium is kept isolated from the lanthanides. Without thorium in the core fluid, removal of lanthanide fission products is simplified.
  2. Low fissile inventory. Because the fissile fuel is concentrated in a small core fluid, the actual reactor core is more compact. There is no fissile material in the outer blanket that contains the fertile fuel for breeding, other than that which has been bred there. Because of this, the 1968 ORNL design required just 315 kilograms of fissile materials to start up a 250 MW(e) two fluid MSBR reactor. [24] :35 This reduces the cost of the initial fissile startup charge, and allows more reactors to be started up on any given amount of fissile material.
  3. More efficient breeding. The thorium blanket can effectively capture leaked neutrons from the core region. There is nearly zero fission occurring in the blanket, so the blanket itself does not leak significant numbers of neutrons. This results in a high efficiency of neutron use (neutron economy), and a higher breeding ratio, especially with small reactors.

One weakness of the two-fluid design is the necessity of periodically replacing the core-blanket barrier due to fast neutron damage. [25] :29 ORNL chose graphite for its barrier material because of its low neutron absorption, compatibility with the molten salts, high temperature resistance, and sufficient strength and integrity to separate the fuel and blanket salts. The effect of neutron radiation on graphite is to slowly shrink and then swell it, causing an increase in porosity and a deterioration in physical properties. [24] :13 Graphite pipes would change length, and may crack and leak.

Another weakness of the two-fluid design is its complex plumbing. ORNL thought a complex interleaving of core and blanket tubes was necessary to achieve a high power level with acceptably low power density. [24] :4 ORNL chose not to pursue the two-fluid design, and no examples of the two-fluid reactor were ever constructed.

However, more recent research has questioned the need for ORNL's complex interleaving graphite tubing, suggesting a simple elongated tube-in-shell reactor that would allow high power output without complex tubing, accommodate thermal expansion, and permit tube replacement. [1] :6 Additionally, graphite can be replaced with high molybdenum alloys, which are used in fusion experiments and have greater tolerance to neutron damage. [1] :6

Hybrid "one and a half fluid" reactor

A two fluid reactor that has thorium in the fuel salt is sometimes called a "one and a half fluid" reactor, or 1.5 fluid reactor. [26] This is a hybrid, with some of the advantages and disadvantages of both 1 fluid and 2 fluid reactors. Like the 1 fluid reactor, it has thorium in the fuel salt, which complicates the fuel processing. And yet, like the 2 fluid reactor, it can use a highly effective separate blanket to absorb neutrons that leak from the core. The added disadvantage of keeping the fluids separate using a barrier remains, but with thorium present in the fuel salt there are fewer neutrons that must pass through this barrier into the blanket fluid. This results in less damage to the barrier. Any leak in the barrier would also be of lower consequence, as the processing system must already deal with thorium in the core.

The main design question when deciding between a one and a half or two fluid LFTR is whether a more complicated reprocessing or a more demanding structural barrier will be easier to solve.

Calculated nuclear performance of 1000-MW(e) MSBR design concepts [25] :29
Design conceptBreeding ratioFissile inventory
Single-fluid, 30-year graphite life, fuel processing1.062300 kg
Single-fluid, 4-year graphite life, fuel processing1.061500 kg
1.5 fluid, replaceable core, fuel processing1.07900 kg
Two-fluid, replaceable core, fuel processing1.07700 kg

Power generation

An LFTR with a high operating temperature of 700 degrees Celsius can operate at a thermal efficiency in converting heat to electricity of 45%. [23] This is higher than today's light water reactors (LWRs) that are at 32–36% thermal to electrical efficiency. In addition to electricity generation, concentrated thermal energy from the high-temperature LFTR can be used as high-grade industrial process heat for many uses, such as ammonia production with the Haber process or thermal Hydrogen production by water splitting, eliminating the efficiency loss of first converting to electricity.

Rankine cycle

Rankine steam cycle Rankine cycle layout.png
Rankine steam cycle

The Rankine cycle is the most basic thermodynamic power cycle. The simplest cycle consists of a steam generator, a turbine, a condenser, and a pump. The working fluid is usually water. A Rankine power conversion system coupled to a LFTR could take advantage of increased steam temperature to improve its thermal efficiency. [27] The subcritical Rankine steam cycle is currently used in commercial power plants, with the newest plants utilizing the higher temperature, higher pressure, supercritical Rankine steam cycles. The work of ORNL from the 1960s and 1970s on the MSBR assumed the use of a standard supercritical steam turbine with an efficiency of 44%, [25] :74 and had done considerable design work on developing molten fluoride salt – steam generators. [28]

Brayton cycle

Closed-cycle gas turbine schematic Schem turb gaz3 en-simple.svg
Closed-cycle gas turbine schematic

The Brayton cycle generator has a much smaller footprint than the Rankine cycle, lower cost and higher thermal efficiency, but requires higher operating temperatures. It is therefore particularly suitable for use with a LFTR. The working gas can be helium, nitrogen, or carbon dioxide. The low-pressure warm gas is cooled in an ambient cooler. The low-pressure cold gas is compressed to the high-pressure of the system. The high-pressure working gas is expanded in a turbine to produce power. Often the turbine and the compressor are mechanically connected through a single shaft. [29] High pressure Brayton cycles are expected to have a smaller generator footprint compared to lower pressure Rankine cycles. A Brayton cycle heat engine can operate at lower pressure with wider diameter piping. [29] The world's first commercial Brayton cycle solar power module (100 kW) was built and demonstrated in Israel's Arava Desert in 2009. [30]

Removal of fission products

The LFTR needs a mechanism to remove the fission products from the fuel. Fission products left in the reactor absorb neutrons and thus reduce neutron economy. This is especially important in the thorium fuel cycle with few spare neutrons and a thermal neutron spectrum, where absorption is strong. The minimum requirement is to recover the valuable fissile material from used fuel.

Removal of fission products is similar to reprocessing of solid fuel elements; by chemical or physical means, the valuable fissile fuel is separated from the waste fission products. Ideally the fertile fuel (thorium or U-238) and other fuel components (e.g. carrier salt or fuel cladding in solid fuels) can also be reused for new fuel. However, for economic reasons they may also end up in the waste.

On site processing is planned to work continuously, cleaning a small fraction of the salt every day and sending it back to the reactor. There is no need to make the fuel salt very clean; the purpose is to keep the concentration of fission products and other impurities (e.g. oxygen) low enough. The concentrations of some of the rare earth elements must be especially kept low, as they have a large absorption cross section. Some other elements with a small cross section like Cs or Zr may accumulate over years of operation before they are removed.

As the fuel of a LFTR is a molten salt mixture, it is attractive to use pyroprocessing, high temperature methods working directly with the hot molten salt. Pyroprocessing does not use radiation sensitive solvents and is not easily disturbed by decay heat. It can be used on highly radioactive fuel directly from the reactor. [31] Having the chemical separation on site, close to the reactor avoids transport and keeps the total inventory of the fuel cycle low. Ideally everything except new fuel (thorium) and waste (fission products) stays inside the plant.

One potential advantage of a liquid fuel is that it not only facilitates separating fission-products from the fuel, but also isolating individual fission products from one another, which is lucrative for isotopes that are scarce and in high-demand for various industrial (radiation sources for testing welds via radiography), agricultural (sterilizing produce via irradiation), and medical uses (Molybdenum-99 which decays into Technetium-99m, a valuable radiolabel dye for marking cancerous cells in medical scans).

Details by element group

The more noble metals (Pd, Ru, Ag, Mo, Nb, Sb, Tc) do not form fluorides in the normal salt, but instead fine colloidal metallic particles. They can plate out on metal surfaces like the heat exchanger, or preferably on high surface area filters which are easier to replace. Still, there is some uncertainty where they end up, as the MSRE only provided a relatively short operating experience and independent laboratory experiments are difficult. [32]

Gases like Xe and Kr come out easily with a sparge of helium. In addition, some of the "noble" metals are removed as an aerosol. The quick removal of Xe-135 is particularly important, as it is a very strong neutron poison and makes reactor control more difficult if unremoved; this also improves neutron economy. The gas (mainly He, Xe and Kr) is held for about 2 days until almost all Xe-135 and other short lived isotopes have decayed. Most of the gas can then be recycled. After an additional hold up of several months, radioactivity is low enough to separate the gas at low temperatures into helium (for reuse), xenon (for sale) and krypton, which needs storage (e.g. in compressed form) for an extended time (several decades) to wait for the decay of Kr-85. [18] :274

For cleaning the salt mixture several methods of chemical separation were proposed. [33] Compared to classical PUREX reprocessing, pyroprocessing can be more compact and produce less secondary waste. The pyroprocesses of the LFTR salt already starts with a suitable liquid form, so it may be less expensive than using solid oxide fuels. However, because no complete molten salt reprocessing plant has been built, all testing has been limited to the laboratory, and with only a few elements. There is still more research and development needed to improve separation and make reprocessing more economically viable.

Uranium and some other elements can be removed from the salt by a process called fluorine volatility: A sparge of fluorine removes volatile high-valence fluorides as a gas. This is mainly uranium hexafluoride, containing the uranium-233 fuel, but also neptunium hexafluoride, technetium hexafluoride and selenium hexafluoride, as well as fluorides of some other fission products (e.g. iodine, molybdenum and tellurium). The volatile fluorides can be further separated by adsorption and distillation. Handling uranium hexafluoride is well established in enrichment. The higher valence fluorides are quite corrosive at high temperatures and require more resistant materials than Hastelloy. One suggestion in the MSBR program at ORNL was using solidified salt as a protective layer. At the MSRE reactor fluorine volatility was used to remove uranium from the fuel salt. Also for use with solid fuel elements fluorine volatility is quite well developed and tested. [31]

Another simple method, tested during the MSRE program, is high temperature vacuum distillation. The lower boiling point fluorides like uranium tetrafluoride and the LiF and BeF carrier salt can be removed by distillation. Under vacuum the temperature can be lower than the ambient pressure boiling point. So a temperature of about 1000 °C is sufficient to recover most of the FLiBe carrier salt. [34] However, while possible in principle, separation of thorium fluoride from the even higher boiling point lanthanide fluorides would require very high temperatures and new materials. The chemical separation for the 2-fluid designs, using uranium as a fissile fuel can work with these two relatively simple processes: [35] Uranium from the blanket salt can be removed by fluorine volatility, and transferred to the core salt. To remove the fissile products from the core salt, first the uranium is removed via fluorine volatility. Then the carrier salt can be recovered by high temperature distillation. The fluorides with a high boiling point, including the lanthanides stay behind as waste.

Optional protactinium-233 separations

The early Oak Ridge's chemistry designs were not concerned with proliferation and aimed for fast breeding. They planned to separate and store protactinium-233, so it could decay to uranium-233 without being destroyed by neutron capture in the reactor. With a half-life of 27 days, 2 months of storage would assure that 75% of the 233Pa decays to 233U fuel. The protactinium removal step is not required per se for a LFTR. Alternate solutions are operating at a lower power density and thus a larger fissile inventory (for 1 or 1.5 fluid) or a larger blanket (for 2 fluid). Also a harder neutron spectrum helps to achieve acceptable breeding without protactinium isolation. [1]

If Pa separation is specified, this must be done quite often (for example, every 10 days) to be effective. For a 1 GW, 1-fluid plant this means about 10% of the fuel or about 15 t of fuel salt need to go through reprocessing every day. This is only feasible if the costs are much lower than current costs for reprocessing solid fuel.

Newer designs usually avoid the Pa removal [1] and send less salt to reprocessing, which reduces the required size and costs for the chemical separation. It also avoids proliferation concerns due to high purity U-233 that might be available from the decay of the chemical separated Pa.

Separation is more difficult if the fission products are mixed with thorium, because thorium, plutonium and the lanthanides (rare earth elements) are chemically similar. One process suggested for both separation of protactinium and the removal of the lanthanides is the contact with molten bismuth. In a redox-reaction some metals can be transferred to the bismuth melt in exchange for lithium added to the bismuth melt. At low lithium concentrations U, Pu and Pa move to the bismuth melt. At more reducing conditions (more lithium in the bismuth melt) the lanthanides and thorium transfer to the bismuth melt too. The fission products are then removed from the bismuth alloy in a separate step, e.g. by contact to a LiCl melt. [36] However this method is far less developed. A similar method may also be possible with other liquid metals like aluminum. [37]

Advantages

Thorium-fueled molten salt reactors offer many potential advantages compared to conventional solid uranium fueled light water reactors: [8] [20] [38] [39] [40] [41]

Safety

Economy and efficiency

Comparison of annual fuel requirements and waste products of a 1 GW uranium-fueled LWR and 1 GW thorium-fueled LFTR power plant. Lwrvslftr2.png
Comparison of annual fuel requirements and waste products of a 1 GW uranium-fueled LWR and 1 GW thorium-fueled LFTR power plant.

Disadvantages

LFTRs are quite unlike today's operating commercial power reactors. These differences create design difficulties and trade-offs:

Recent developments

The Fuji MSR

The FUJI MSR was a design for a 100 to 200 MWe molten-salt-fueled thorium fuel cycle thermal breeder reactor, using technology similar to the Oak Ridge National Laboratory Reactor Experiment. It was being developed by a consortium including members from Japan, the United States, and Russia. As a breeder reactor, it converts thorium into nuclear fuels. [98] An industry group presented updated plans about FUJI MSR in July 2010. [99] They projected a cost of 2.85 cents per kilowatt hour. [100]

The IThEMS consortium planned to first build a much smaller MiniFUJI 10 MWe reactor of the same design once it had secured an additional $300 million in funding, but IThEMS closed in 2011 after it was unable to secure adequate funding. A new company, Thorium Tech Solution (TTS), was founded in 2011 by Kazuo Furukawa, the chief scientist from IThEMS, and Masaaki Furukawa. TTS acquired the FUJI design and some related patents.

Chinese thorium MSR project

The People's Republic of China has initiated a research and development project in thorium molten-salt reactor technology. [101] It was formally announced at the Chinese Academy of Sciences (CAS) annual conference in January 2011. Its ultimate target is to investigate and develop a thorium based molten salt nuclear system in about 20 years. [102] [103] An expected intermediate outcome of the TMSR research program is to build a 2 MW pebble bed fluoride salt cooled research reactor in 2015, and a 2 MW molten salt fueled research reactor in 2017. This would be followed by a 10 MW demonstrator reactor and a 100 MW pilot reactors. [104] [105] The project is spearheaded by Jiang Mianheng, with a start-up budget of $350 million, and has already recruited 140 PhD scientists, working full-time on thorium molten salt reactor research at the Shanghai Institute of Applied Physics. An expansion of staffing has increased to 700 as of 2015. [106] As of 2016, their plan is for a 10MW pilot LFTR is expected to be made operational in 2025, with a 100MW version set to follow in 2035. [107]

At the end of August 2021, the Shanghai Institute of Applied Physics (SINAP) completed the construction of a 2MW (thermal) experimental thorium molten salt reactor in Wuwei, Gansu, known as the TMSR-LF1. [108] China plans to follow up the experiment with a 373MW version by 2030. [109]

Flibe Energy

Kirk Sorensen, former NASA scientist and Chief Nuclear Technologist at Teledyne Brown Engineering, has been a long-time promoter of thorium fuel cycle and particularly liquid fluoride thorium reactors. He first researched thorium reactors while working at NASA, while evaluating power plant designs suitable for lunar colonies. Material about this fuel cycle was surprisingly hard to find, so in 2006 Sorensen started "energyfromthorium.com", a document repository, forum, and blog to promote this technology. In 2006, Sorensen coined the liquid fluoride thorium reactor and LFTR nomenclature to describe a subset of molten salt reactor designs based on liquid fluoride-salt fuels with breeding of thorium into uranium-233 in the thermal spectrum. In 2011, Sorensen founded Flibe Energy, a company that initially intends to develop 20–50 MW LFTR small modular reactor designs to power military bases; Sorensen noted that it is easier to promote novel military designs than civilian power station designs in the context of the modern US nuclear regulatory and political environment. [110] [111] An independent technology assessment coordinated with EPRI and Southern Company represents the most detailed information so far publicly available about Flibe Energy's proposed LFTR design. [112]

Thorium Energy Generation Pty. Limited (TEG)

Thorium Energy Generation Pty. Limited (TEG) was an Australian research and development company dedicated to the worldwide commercial development of LFTR reactors, as well as thorium accelerator-driven systems. As of June 2015, TEG had ceased operations.

Alvin Weinberg Foundation

The Alvin Weinberg Foundation was a British charity founded in 2011, dedicated to raising awareness about the potential of thorium energy and LFTR. It was formally launched at the House of Lords on 8 September 2011. [113] [114] [115] It is named after American nuclear physicist Alvin M. Weinberg, who pioneered the thorium molten salt reactor research.

Thorcon

ThorCon nuclear reactor is a proposed floating molten salt reactor, by the US-based Thorcon company. The two-reactor unit is designed to be manufactured on an assembly line in a shipyard, and to be delivered via barge to any ocean or major waterway shoreline. The reactors are to be delivered as a sealed unit and never opened on site. All reactor maintenance and fuel processing is done at an off-site location.

Nuclear Research and Consultancy Group

On 5 September 2017, the Dutch Nuclear Research and Consultancy Group announced that research on the irradiation of molten thorium fluoride salts inside the Petten high-flux reactor was underway. [116]

See also

Related Research Articles

<span class="mw-page-title-main">Nuclear reactor</span> Device used to initiate and control a nuclear chain reaction

A nuclear reactor is a device used to initiate and control a fission nuclear chain reaction or nuclear fusion reactions. Nuclear reactors are used at nuclear power plants for electricity generation and in nuclear marine propulsion. Heat from nuclear fission is passed to a working fluid, which in turn runs through steam turbines. These either drive a ship's propellers or turn electrical generators' shafts. Nuclear generated steam in principle can be used for industrial process heat or for district heating. Some reactors are used to produce isotopes for medical and industrial use, or for production of weapons-grade plutonium. As of 2022, the International Atomic Energy Agency reports there are 422 nuclear power reactors and 223 nuclear research reactors in operation around the world.

<span class="mw-page-title-main">Pressurized water reactor</span> Type of nuclear reactor

A pressurized water reactor (PWR) is a type of light-water nuclear reactor. PWRs constitute the large majority of the world's nuclear power plants. In a PWR, the primary coolant (water) is pumped under high pressure to the reactor core where it is heated by the energy released by the fission of atoms. The heated, high pressure water then flows to a steam generator, where it transfers its thermal energy to lower pressure water of a secondary system where steam is generated. The steam then drives turbines, which spin an electric generator. In contrast to a boiling water reactor (BWR), pressure in the primary coolant loop prevents the water from boiling within the reactor. All light-water reactors use ordinary water as both coolant and neutron moderator. Most use anywhere from two to four vertically mounted steam generators; VVER reactors use horizontal steam generators.

<span class="mw-page-title-main">Pebble-bed reactor</span> Type of very-high-temperature reactor

The pebble-bed reactor (PBR) is a design for a graphite-moderated, gas-cooled nuclear reactor. It is a type of very-high-temperature reactor (VHTR), one of the six classes of nuclear reactors in the Generation IV initiative.

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

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

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

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

<span class="mw-page-title-main">Breeder reactor</span> Nuclear reactor generating more fissile material than it consumes

A breeder reactor is a nuclear reactor that generates more fissile material than it consumes. These reactors can be fueled with more-commonly available isotopes of uranium and thorium, such as uranium-238 and thorium-232, as opposed to the rare uranium-235 which is used in conventional reactors. These materials are called fertile materials since they can be bred into fuel by these breeder reactors.

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

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

Passive nuclear safety is a design approach for safety features, implemented in a nuclear reactor, that does not require any active intervention on the part of the operator or electrical/electronic feedback in order to bring the reactor to a safe shutdown state, in the event of a particular type of emergency. Such design features tend to rely on the engineering of components such that their predicted behaviour would slow down, rather than accelerate the deterioration of the reactor state; they typically take advantage of natural forces or phenomena such as gravity, buoyancy, pressure differences, conduction or natural heat convection to accomplish safety functions without requiring an active power source. Many older common reactor designs use passive safety systems to a limited extent, rather, relying on active safety systems such as diesel-powered motors. Some newer reactor designs feature more passive systems; the motivation being that they are highly reliable and reduce the cost associated with the installation and maintenance of systems that would otherwise require multiple trains of equipment and redundant safety class power supplies in order to achieve the same level of reliability. However, weak driving forces that power many passive safety features can pose significant challenges to effectiveness of a passive system, particularly in the short term following an accident.

<span class="mw-page-title-main">Integral fast reactor</span> Nuclear reactor design

The integral fast reactor is a design for a nuclear reactor using fast neutrons and no neutron moderator. IFR would breed more fuel and is distinguished by a nuclear fuel cycle that uses reprocessing via electrorefining at the reactor site.

<span class="mw-page-title-main">Molten-salt reactor</span> Type of nuclear reactor cooled by molten material

A molten-salt reactor (MSR) is a class of nuclear fission reactor in which the primary nuclear reactor coolant and/or the fuel is a mixture of molten salt with a fissionable material.

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

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

Uranium-233 is a fissile isotope of uranium that is bred from thorium-232 as part of the thorium fuel cycle. Uranium-233 was investigated for use in nuclear weapons and as a reactor fuel. It has been used successfully in experimental nuclear reactors and has been proposed for much wider use as a nuclear fuel. It has a half-life of 160,000 years.

Generation IVreactors are nuclear reactor design technologies that are envisioned as successors of generation III reactors. The Generation IV International Forum (GIF) – an international organization that coordinates the development of generation IV reactors – specifically selected six reactor technologies as candidates for generation IV reactors. The designs target improved safety, sustainability, efficiency, and cost. The World Nuclear Association in 2015 suggested that some might enter commercial operation before 2030.

<span class="mw-page-title-main">Thorium fuel cycle</span> Nuclear fuel cycle

The thorium fuel cycle is a nuclear fuel cycle that uses an isotope of thorium, 232
Th
, as the fertile material. In the reactor, 232
Th
is transmuted into the fissile artificial uranium isotope 233
U
which is the nuclear fuel. Unlike natural uranium, natural thorium contains only trace amounts of fissile material, which are insufficient to initiate a nuclear chain reaction. Additional fissile material or another neutron source is necessary to initiate the fuel cycle. In a thorium-fuelled reactor, 232
Th
absorbs neutrons to produce 233
U
. This parallels the process in uranium breeder reactors whereby fertile 238
U
absorbs neutrons to form fissile 239
Pu
. Depending on the design of the reactor and fuel cycle, the generated 233
U
either fissions in situ or is chemically separated from the used nuclear fuel and formed into new nuclear fuel.

<span class="mw-page-title-main">Molten-Salt Reactor Experiment</span> Nuclear reactor, Oak Ridge 1965–1969

The Molten-Salt Reactor Experiment (MSRE) was an experimental molten salt reactor research reactor at the Oak Ridge National Laboratory (ORNL). This technology was researched through the 1960s, the reactor was constructed by 1964, it went critical in 1965, and was operated until 1969. The costs of a cleanup project were estimated at $130 million.

Hybrid nuclear fusion–fission is a proposed means of generating power by use of a combination of nuclear fusion and fission processes.

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

FLiBe is the name of a molten salt made from a mixture of lithium fluoride (LiF) and beryllium fluoride. It is both a nuclear reactor coolant and solvent for fertile or fissile material. It served both purposes in the Molten-Salt Reactor Experiment (MSRE) at the Oak Ridge National Laboratory.

<span class="mw-page-title-main">Thorium-based nuclear power</span> Nuclear energy extracted from thorium isotopes

Thorium-based nuclear power generation is fueled primarily by the nuclear fission of the isotope uranium-233 produced from the fertile element thorium. A thorium fuel cycle can offer several potential advantages over a uranium fuel cycle—including the much greater abundance of thorium found on Earth, superior physical and nuclear fuel properties, and reduced nuclear waste production. One advantage of thorium fuel is its low weaponization potential. It is difficult to weaponize the uranium-233 that is bred in the reactor. Plutonium-239 is produced at much lower levels and can be consumed in thorium reactors.

<span class="mw-page-title-main">Integral Molten Salt Reactor</span>

The Integral Molten Salt Reactor (IMSR) is a nuclear power plant design targeted at developing a commercial product for the small modular reactor (SMR) market. It employs molten salt reactor technology which is being developed by the Canadian company Terrestrial Energy. It is based closely on the denatured molten salt reactor (DMSR), a reactor design from Oak Ridge National Laboratory. In addition, it incorporates some elements found in the SmAHTR, a later design from the same laboratory. The IMSR belongs to the DMSR class of molten salt reactors (MSR) and hence is a "burner" reactor that employs a liquid fuel rather than a conventional solid fuel. This liquid contains the nuclear fuel as well as serving as the primary coolant.

The Molten-Salt Demonstration Reactor (MSDR) was a semi-commercial-scale experimental molten salt reactor (MSR) design developed at Oak Ridge National Laboratory (ORNL).

References

  1. 1 2 3 4 5 6 7 8 9 10 LeBlanc, David (2010). "Molten salt reactors: A new beginning for an old idea" (PDF). Nuclear Engineering and Design. 240 (6): 1644. doi:10.1016/j.nucengdes.2009.12.033.
  2. Greene, Sherrel (May 2011). Fluoride Salt-cooled High Temperature Reactors – Technology Status and Development Strategy. ICENES-2011. San Francisco, CA.
  3. Stenger, Victor (12 January 2012). "LFTR: A Long-Term Energy Solution?". Huffington Post.
  4. Williams, Stephen (16 January 2015). "Molten Salt Reactors: The Future of Green Energy?". ZME Science. Retrieved 12 August 2015.
  5. 1 2 Warmflash, David (16 January 2015). "Thorium Power Is the Safer Future of Nuclear Energy". Discover Magazine. Archived from the original on 21 January 2015. Retrieved 22 January 2015.
  6. UP (29 September 1946). "Atomic Energy 'Secret' Put into Language That Public Can Understand". Pittsburgh Press . Retrieved 18 October 2011.
  7. UP (21 October 1946). "Third Nuclear Source Bared". The Tuscaloosa News . Retrieved 18 October 2011.
  8. 1 2 3 4 5 6 7 8 9 10 11 12 13 Hargraves, Robert; Moir, Ralph (July 2010). "Liquid fluoride thorium reactors: an old idea in nuclear power gets reexamined" (PDF). American Scientist. 98 (4): 304–313. doi:10.1511/2010.85.304. Archived from the original (PDF) on 8 December 2013.
  9. Synthesis of heavy elements. Gesellschaft für Schwerionenforschung. gsi.de
  10. The KamLAND Collaboration; Gando, Y.; Ichimura, K.; Ikeda, H.; Inoue, K.; Kibe, Y.; Kishimoto, Y.; Koga, M.; Minekawa, Y.; et al. (17 July 2011). "Partial radiogenic heat model for Earth revealed by geoneutrino measurements" (PDF). Nature Geoscience. 4 (9): 647–651. Bibcode:2011NatGe...4..647K. doi:10.1038/ngeo1205.
  11. "Lab's early submarine reactor program paved the way for modern nuclear power plants". Argonne's Nuclear Science and Technology Legacy. Argonne National Laboratory. 1996.
  12. Sorensen, Kirk (2 July 2009). "Lessons for the Liquid-Fluoride Thorium Reactor" (PDF). Mountain View, CA. Archived from the original (PDF) on 12 December 2011.
  13. 1 2 Rosenthal, M.; Briggs, R.; Haubenreich, P. "Molten-Salt Reactor Program: Semiannual Progress Report for Period Ending August 31, 1971" (PDF). ORNL-4728. Oak Ridge National Laboratory.{{cite journal}}: Cite journal requires |journal= (help)
  14. MacPherson, H. G. (1 August 1985). "The Molten Salt Reactor Adventure". Nuclear Science and Engineering. 90 (4): 374–380. Bibcode:1985NSE....90..374M. doi:10.13182/NSE90-374. Archived from the original on 4 June 2011.
  15. Weinberg, Alvin (1997). The First Nuclear Era: The Life and Times of a Technological Fixer. Vol. 48. Springer. pp. 63–64. Bibcode:1995PhT....48j..63W. doi:10.1063/1.2808209. ISBN   978-1-56396-358-2.{{cite book}}: |journal= ignored (help)
  16. "ORNL: The First 50 Years - Chapter 6: Responding to Social Needs". Archived from the original on 16 September 2012. Retrieved 12 November 2011.
  17. "Plutonium". World Nuclear Association. March 2012. Archived from the original on 30 March 2010. Retrieved 28 June 2012. The most common isotope formed in a typical nuclear reactor is the fissile Pu-239 isotope, formed by neutron capture from U-238 (followed by beta decay), and which yields much the same energy as the fission of U-235. Well over half of the plutonium created in the reactor core is consumed in situ and is responsible for about one third of the total heat output of a light water reactor (LWR).(Updated)
  18. 1 2 3 4 Rosenthal; M. W.; et al. (August 1972). "The Development Status of Molten-Salt Breeder Reactors" (PDF). ORNL-4812. Oak Ridge National Laboratory.{{cite journal}}: Cite journal requires |journal= (help)
  19. 1 2 3 Rosenthal, M. W.; Kasten, P. R.; Briggs, R. B. (1970). "Molten Salt Reactors – History, Status, and Potential" (PDF). Nuclear Applications and Technology. 8 (2): 107–117. doi:10.13182/NT70-A28619.
  20. 1 2 Section 5.3, WASH 1097 "The Use of Thorium in Nuclear Power Reactors", available as a PDF from Liquid-Halide Reactor Documents Accessed 11/23/09
  21. Briggs, R. B. (November 1964). "Molten-Salt Reactor Program Semiannual Progress Report For Period Ending July 31, 1964" (PDF). ORNL-3708. Oak Ridge National Laboratory.{{cite journal}}: Cite journal requires |journal= (help)
  22. Furukawa; K. A.; et al. (2008). "A road map for the realization of global-scale Thorium breeding fuel cycle by single molten-fluoride flow". Energy Conversion and Management. 49 (7): 1832. doi:10.1016/j.enconman.2007.09.027.
  23. 1 2 Hargraves, Robert; Moir, Ralph (January 2011). "Liquid Fuel Nuclear Reactors". Forum on Physics & Society. 41 (1): 6–10.
  24. 1 2 3 Robertson, R. C.; Briggs, R. B.; Smith, O. L.; Bettis, E. S. (1970). "Two-Fluid Molten-Salt Breeder Reactor Design Study (Status as of January 1, 1968)". ORNL-4528. Oak Ridge National Laboratory. doi: 10.2172/4093364 .{{cite journal}}: Cite journal requires |journal= (help)
  25. 1 2 3 Robertson, R. C. (June 1971). "Conceptual Design Study of a Single-Fluid Molten-Salt Breeder Reactor" (PDF). ORNL-4541. Oak Ridge National Laboratory.{{cite journal}}: Cite journal requires |journal= (help)
  26. LeBlanc, David (May 2010). "Too Good to Leave on the Shelf". Mechanical Engineering. 132 (5): 29–33. doi: 10.1115/1.2010-May-2 .
  27. Hough, Shane (4 July 2009) Supercritical Rankine Cycle. if.uidaho.edu
  28. "Oak Ridge National Laboratory: A New Approach to the Design of Steam Generators for Molten Salt Reactor Power Plants" (PDF). Moltensalt.org. Retrieved 24 October 2012.
  29. 1 2 Sabharwall, Piyush; Kim, Eung S.; McKellar, Michael; Anderson, Nolan (April 2011). Process Heat Exchanger Options for Fluoride Salt High Temperature Reactor (PDF) (Report). Idaho National Laboratory. Archived from the original (PDF) on 8 August 2014. Retrieved 4 May 2012.
  30. ""Flower power" has been inaugurated in Israel" (News). Enel Green Power. 10 July 2009.{{cite journal}}: Cite journal requires |journal= (help)[ dead link ]
  31. 1 2 "Pyrochemical Separations in Nuclear Applications: A Status Report" (PDF). Retrieved 24 October 2012.
  32. Forsberg, Charles W. (2006). "Molten-Salt-Reactor Technology Gaps" (PDF). Proceedings of the 2006 International Congress on Advances in Nuclear Power Plants (ICAPP '06). Archived from the original (PDF) on 29 October 2013. Retrieved 7 April 2012.
  33. 1 2 3 "LIFE Materials: Molten-Salt Fuels Volume 8" (PDF). E-reports-ext.11nl.gov. Retrieved 24 October 2012.
  34. "Low-Pressure Distillation of Molten Fluoride Mixtures: Nonradioactive Tests for the MSRE Distillation Experiment;1971, ORNL-4434" (PDF). Retrieved 24 October 2012.
  35. "Design Studies of 1000-Mw(e) Molten-Salt Breeder Reactors; 1966, ORNL-3996" (PDF). Retrieved 24 October 2012.
  36. "Engineering Tests of the Metal Transfer Process for Extraction of Rare-Earth Fission Products from a Molten-Salt Breeder Reactor Fuel Salt; 1976, ORNL-5176" (PDF). Retrieved 24 October 2012.
  37. Conocar, Olivier; Douyere, Nicolas; Glatz, Jean-Paul; Lacquement, Jérôme; Malmbeck, Rikard & Serp, Jérôme (2006). "Promising pyrochemical actinide/lanthanide separation processes using aluminium" . Nuclear Science and Engineering. 153 (3): 253–261. Bibcode:2006NSE...153..253C. doi:10.13182/NSE06-A2611. S2CID   91818903.
  38. "Molten Salt Reactors: A New Beginning for an Old Idea" (PDF). Archived from the original (PDF) on 4 October 2013. Retrieved 24 October 2012.
  39. "Potential of Thorium Fueled Molten Salt Reactors" (PDF). Archived from the original (PDF) on 22 January 2012. Retrieved 24 October 2012.
  40. "6th Int'l Summer Student School on Nuclear Physics Methods and Accelerators in Biology and Medicine (July 2011, JINR Dubna, Russia)" (PDF). Uc2.jinr.ru. Archived from the original (PDF) on 15 May 2013. Retrieved 24 October 2012.
  41. Cooper, N.; Minakata, D.; Begovic, M.; Crittenden, J. (2011). "Should We Consider Using Liquid Fluoride Thorium Reactors for Power Generation?". Environmental Science & Technology. 45 (15): 6237–8. Bibcode:2011EnST...45.6237C. doi: 10.1021/es2021318 . PMID   21732635.
  42. 1 2 3 4 5 6 7 8 Mathieu, L.; Heuer, D.; Brissot, R.; Garzenne, C.; Le Brun, C.; Lecarpentier, D.; Liatard, E.; Loiseaux, J.-M.; Méplan, O.; et al. (2006). "The Thorium molten salt reactor: Moving on from the MSBR" (PDF). Progress in Nuclear Energy. 48 (7): 664–679. arXiv: nucl-ex/0506004 . doi:10.1016/j.pnucene.2006.07.005. S2CID   15091933.
  43. 1 2 "Engineering Database of Liquid Salt Thermophysical and Thermochemical Properties" (PDF). Inl.gov. Archived from the original (PDF) on 8 August 2014. Retrieved 24 October 2012.
  44. "Chapter 13: Construction Materials for Molten-Salt Reactors" (PDF). Moltensalt.org. Retrieved 24 October 2012.
  45. "Thermal- and Fast Spectrum Molten Salt Reactors for Actinide Burning and Fuel Production" (PDF). Archived from the original (PDF) on 19 January 2012. Retrieved 24 October 2012.
  46. 1 2 Devanney, Jack. "Simple Molten Salt Reactors: a time for courageous impatience" (PDF). C4tx.org. Archived from the original (PDF) on 23 September 2015. Retrieved 24 October 2012.
  47. Moir, R. W. (2008). "Recommendations for a restart of molten salt reactor development" (PDF). Energy Conversion and Management. 49 (7): 1849–1858. doi:10.1016/j.enconman.2007.07.047.
  48. Leblanc, D. (2010). "Molten salt reactors: A new beginning for an old idea". Nuclear Engineering and Design. 240 (6): 1644. doi:10.1016/j.nucengdes.2009.12.033.
  49. "The Influence of Xenon-135 on Reactor Operation" (PDF). C-n-t-a.com. Retrieved 24 October 2012.
  50. 1 2 3 "Assessment of Candidate Molten Salt Coolants for the Advanced High-Temperature Reactor (AHTR)- ORNL-TM-2006-12" (PDF). Archived from the original (PDF) on 26 September 2012. Retrieved 24 October 2012.
  51. "A Modular Radiant Heat-Initiated Passive Decay-Heat-Removal System for Salt-Cooled Reactors" (PDF). Ornl.gov. Archived from the original (PDF) on 21 October 2008. Retrieved 24 October 2012.
  52. Thorium Fuel Cycle, AEC Symposium Series, 12, USAEC, Feb. 1968
  53. "Using LTFR to Minimize Actinide Wastes" (PDF). Thoriumenergyaslliance.com. Archived from the original (PDF) on 15 May 2013. Retrieved 24 October 2012.
  54. 1 2 Engel, J. R.; Grimes, W. R.; Bauman, H. F.; McCoy, H. E.; Dearing, J. F.; Rhoades, W. A. (1980). Conceptual design characteristics of a denatured molten-salt reactor with once-through fueling (PDF). Oak Ridge National Lab, TN. ORNL/TM-7207. Archived from the original (PDF) on 14 January 2010. Retrieved 22 November 2011.
  55. Hargraves, Robert & Moir, Ralph (27 July 2011). "Liquid Fuel Nuclear Reactors". Aps.org. Retrieved 3 August 2012.
  56. "for nuclear energy looms". Archived from the original on 22 July 2016. Retrieved 26 January 2016.
  57. 1 2 Sylvain, David; et al. (March–April 2007). "Revisiting the Thorium-Uranium nuclear fuel cycle" (PDF). Europhysics News. 38 (2): 24–27. Bibcode:2007ENews..38b..24D. doi: 10.1051/EPN:2007007 .
  58. "Image based on". Thoriumenergyalliance.com. Archived from the original (PDF) on 5 April 2012. Retrieved 24 October 2012.
  59. Evans-Pritchard, Ambrose (29 August 2010) Obama could kill fossil fuels overnight with a nuclear dash for thorium. Telegraph. Retrieved on 24 April 2013.
  60. 1 2 3 "Oak Ridge National Laboratory: Abstract" (PDF). Energyfromthorium. Retrieved 24 October 2012.
  61. "Denatured Molten Salt Reactors" (PDF). Coal2nuclear.com. Retrieved 24 October 2012.
  62. "Estimated Cost of Adding a Third Salt-Circulating System for Controlling Tritium Migration in the 1000-Mw(e) MSBR [Disc 5]" (PDF). Retrieved 24 October 2012.
  63. 1 2 3 4 Bonometti, J. "LFTR Liquid Fluoride Thorium Reactor-What fusion wanted to be!" Presentation available in www.energyfromthorium.com (2011)
  64. "Critical issues of nuclear energy systems employing molten salt fluorides" (PDF). Archived from the original (PDF) on 26 April 2012. Retrieved 24 October 2012.
  65. Peterson, Per F.; Zhao, H. & Fukuda, G. (5 December 2003). "Comparison of Molten Salt and High-Pressure Helium for the NGNP Intermediate Heat Transfer Fluid" (PDF). U.C. Berkeley Report UCBTH-03-004. Archived from the original (PDF) on 11 August 2014.
  66. "Products". Flibe Energy. Archived from the original on 28 June 2013. Retrieved 24 October 2012.
  67. Bush, R. P. (1991). "Recovery of Platinum Group Metals from High Level Radioactive Waste" (PDF). Platinum Metals Review. 35 (4): 202–208. Archived from the original (PDF) on 24 September 2015. Retrieved 9 March 2013.
  68. "Thorium fuel cycle – Potential benefits and challenges" (PDF). International Atomic Energy Agency . Retrieved 27 October 2014.
  69. Chiang, Howard; Jiang, Yihao; Levine, Sam; Pittard, Kris; Qian, Kevin; Yu, Pam (8 December 2014). Liquid Fluoride Thorium Reactors: Traditional Nuclear Plant Comparison Analysis and Feasibility Study (PDF) (Technical report). University of Chicago.
  70. "Thorium". World Nuclear.
  71. Peterson, Per F. & Zhao, Haihua (29 December 2005). "Preliminary Design Description for a First-Generation Liquid-Salt VHTR with Metallic Vessel Internals (AHTR-MI)" (PDF). U.C. Berkeley Report UCBTH-05-005. Archived from the original (PDF) on 1 January 2014.
  72. 1 2 Fei, Ting; et al. (16 May 2008). "A MODULAR PEBBLE-BED ADVANCE D HIGH TEMPERATURE REACTOR" (PDF). U.C. Berkeley Report UCBTH-08-001. Archived from the original (PDF) on 1 January 2014. Retrieved 24 October 2012.
  73. "The Thorium Molten Salt Reactor: Launching The Thorium Cycle While Closing The Current Fuel Cycle" (PDF). Retrieved 24 October 2012.
  74. "The Aircraft Reactor Experiment-Physics" (PDF). Moltensalt.org. Retrieved 24 October 2012.
  75. National Research Council (U.S.). Committee on Remediation of Buried and Tank Wastes. Molten Salt Panel (1997). Evaluation of the U.S. Department of Energy's alternatives for the removal and disposition of molten salt reactor experiment fluoride salts. National Academies Press. p. 15. ISBN   978-0-309-05684-7.
  76. 1 2 "Fluorine Production and Recombination in Frozen MSR Salts after Reactor Operation [Disc 5]" (PDF). Retrieved 24 October 2012.
  77. Forsberg, C.; Beahm, E.; Rudolph, J. (2 December 1996). Direct Conversion of Halogen-Containing Wastes to Borosilicate Glass (PDF). Symposium II Scientific Basis for Nuclear Waste Management XX. Vol. 465. Boston, Massachusetts: Materials Research Society. pp. 131–137.
  78. "Costs of decommissioning nuclear power plants" (PDF). Iaea.org. Archived from the original (PDF) on 6 August 2009. Retrieved 24 October 2012.
  79. "Oak Ridge National Laboratory: Graphite Behaviour and Its Effects on MSBR Performance" (PDF). Moltensalt.org. Retrieved 24 October 2012.
  80. 1 2 "IAEA-TECDOC-1521" (PDF). Retrieved 24 October 2012.
  81. "Semiannual Progress Report for Period Ending February 28, 1970" (PDF). ORNL-4548: Molten-Salt Reactor Program. p. 57. Archived from the original (PDF) on 29 June 2011. Retrieved 6 June 2015.
  82. Rodriguez-Vieitez, E.; Lowenthal, M. D.; Greenspan, E.; Ahn, J. (7 October 2002). Optimization of a Molten-Salt Transmuting Reactor (PDF). PHYSOR 2002. Seoul, Korea.
  83. 1 2 "Nuclear Weapons Archive – Useful Tables" . Retrieved 31 August 2013.
  84. "Thorium Fuel Has Risks" . Retrieved 16 October 2015.
  85. 1 2 "Neptunium 237 and Americium: World Inventories and Proliferation Concerns" (PDF). Isis-online.org. Retrieved 24 October 2012.
  86. 1 2 "Distribution and Behavior of Tritium in the Coolant-Salt Technology Facility [Disc 6]" (PDF). Retrieved 24 October 2012.
  87. Manely; W. D.; et al. (1960). "Metallurgical Problems in Molten Fluoride Systems". Progress in Nuclear Energy. 2: 164–179.
  88. Heung, L.K. (31 August 2012). "Titanium for long-term tritium storage" (PDF). Osti.gov. doi:10.2172/10117162 . Retrieved 24 October 2012.{{cite journal}}: Cite journal requires |journal= (help)
  89. Robertson, R.C. (31 August 2012). "Conceptual Design Study of a Single-Fluid Molten-Salt Breeder Reactor" (PDF). Osti.gov. doi:10.2172/4030941 . Retrieved 24 October 2012.{{cite journal}}: Cite journal requires |journal= (help)
  90. Moir; R. W.; et al. (2002). "Deep-Burn Molten-Salt Reactors" (Application under Solicitation). LAB NE 2002-1. Department of Energy, Nuclear Energy Research Initiative.{{cite journal}}: Cite journal requires |journal= (help)
  91. "Status of materials development for molten salt reactors" (PDF). Retrieved 24 October 2012.
  92. (52 MB) Intergranular Cracking of INOR-8 in the MSRE,
  93. "Potential of Thorium Molten Salt Reactors: Detailed Calculations and Concept Evolutions in View of a Large Nuclear Energy Production" (PDF). Hal.archives-ouvertes.fr. Retrieved 24 October 2012.
  94. Zhao, H. & Peterson, Per F. (25 February 2004). "A Reference 2400 MW(t) Power Conversion System Point Design for Molten-Salt-Cooled Fission and Fusion Energy Systems" (PDF). U.C. Berkeley Report UCBTH-03-002. Archived from the original (PDF) on 1 January 2014.
  95. Hee Cheon No; Ji Hwan Kim; Hyeun Min Kim (2007). "A review of helium gas turbine technology for high-temperature gas-cooled reactors". Nuclear Engineering and Technology. 39 (1): 21–30. doi: 10.5516/net.2007.39.1.021 .
  96. "Conceptual Design study of a Single Fluid Molten Salt Breeder Reactor" (PDF). Energyfromthorium.com. Retrieved 24 October 2012.
  97. "Heat Transfer Salt for High Temperature Steam Generation [Disc 5]" (PDF). Retrieved 24 October 2012.
  98. Fuji MSR pp. 821–856, Jan 2007
  99. "IThEO Presents International Thorium Energy & Molten-Salt Technology Inc". International Thorium Energy Organisation. 20 July 2010. Archived from the original on 27 July 2010.
  100. "Chapter X. MSR-FUJI General Information, Technical Features, and Operating Characteristics" (PDF).
  101. Martin, Richard (1 February 2011). "China Takes Lead in Race for Clean Nuclear Power". Wired Science.
  102. "未来核电站 安全"不挑食"". Whb.news365.com.cn. 26 January 2011. Archived from the original on 17 July 2012. Retrieved 24 October 2012.
  103. Clark, Duncan (16 February 2011). "China enters race to develop nuclear energy from thorium". The Guardian. London.
  104. "Kun Chen from Chinese Academy of Sciences on China Thorium Molten Salt Reactor TMSR Program". YouTube. 10 August 2012. Retrieved 24 October 2012.
  105. Halper, Mark (30 October 2012). "Completion date slips for China.s thorium molten salt reactor". Weinberg Foundation. Archived from the original on 21 April 2017. Retrieved 17 April 2013.
  106. Evans-Pritchard, Ambrose (6 January 2013). "China blazes trail for 'clean' nuclear power from thorium". The Daily Telegraph .
  107. Brian Wang (11 October 2016). "Update on the Liquid Fluoride Thorium Reactor projects in China and the USA". Next Big Future. Retrieved 27 June 2017.
  108. "Chinese molten-salt reactor cleared for start up". 9 August 2022.
  109. Mallapaty, Smriti (9 September 2021). "China prepares to test thorium-fuelled nuclear reactor". Nature. 597 (7876): 311–312. Bibcode:2021Natur.597..311M. doi:10.1038/d41586-021-02459-w. PMID   34504330. S2CID   237471852.
  110. "Flibe Energy". Flibe Energy. Retrieved 24 October 2012.
  111. "New Huntsville company to build Thorium-based nuclear reactors". Huntsvillenewswire.com. 27 September 2011. Archived from the original on 6 April 2012. Retrieved 24 October 2012.
  112. "Program on Technology Innovation: Technology Assessment of a Molten Salt Reactor Design – The Liquid-Fluoride Thorium Reactor (LFTR)". EPRI. 22 October 2015. Archived from the original on 10 March 2016. Retrieved 10 March 2016.
  113. Clark, Duncan (9 September 2011). "Thorium advocates launch pressure group". The Guardian. London.
  114. "The Weinberg Foundation – London: Weinberg Foundation to heat up campaign for safe, green,…". Mynewsdesk. 8 September 2011. Archived from the original on 30 October 2011. Retrieved 24 October 2012.
  115. "New NGO to fuel interest in safe thorium nuclear reactors". BusinessGreen. 8 September 2011. Retrieved 24 October 2012.
  116. "NRG: Detail".

Further reading

Videos