Iodine pit

Last updated

The iodine pit, also called the iodine hole or xenon pit, is a temporary disabling of a nuclear reactor due to buildup of short-lived nuclear poisons in the reactor core. The main isotope responsible is 135Xe, mainly produced by natural decay of 135I. 135I is a weak neutron absorber, while 135Xe is the strongest known neutron absorber. When 135Xe builds up in the fuel rods of a reactor, it significantly lowers their reactivity, by absorbing a significant amount of the neutrons that provide the nuclear reaction.

Contents

The presence of 135I and 135Xe in the reactor is one of the main reasons for its power fluctuations in reaction to change of control rod positions.

The buildup of short-lived fission products acting as nuclear poisons is called reactor poisoning, or xenon poisoning. Buildup of stable or long-lived neutron poisons is called reactor slagging.

Fission products decay and burnup

One of the common fission products is 135Te, which undergoes beta decay with half-life of 19 seconds to 135I. 135I itself is a weak neutron absorber. It builds up in the reactor in the rate proportional to the rate of fission, which is proportional to the reactor thermal power. 135I undergoes beta decay with half-life of 6.57 hours to 135Xe. The yield of 135Xe for uranium fission is 6.3%; about 95% of 135Xe originates from decay of 135I.

135Xe is the most powerful known neutron absorber, with a cross section for thermal neutrons of 2.6×106  barns, [1] so it acts as a "poison" that can slow or stop the chain reaction after a period of operation. This was discovered in the earliest nuclear reactors built by the Manhattan Project for plutonium production. As a result, the designers made provisions in the design to increase the reactor's reactivity (the number of neutrons per fission that go on to fission other atoms of nuclear fuel). [2] 135Xe reactor poisoning played a major role in the Chernobyl disaster. [3]

By neutron capture, 135Xe is transformed ("burned") to 136Xe, which is effectively [lower-alpha 1] stable and does not significantly absorb neutrons.

The burn rate is proportional to the neutron flux, which is proportional to the reactor power; a reactor running at twice the power will have twice the xenon burn rate. The production rate is also proportional to reactor power, but due to the half-life time of 135I, this rate depends on the average power over the past several hours.

As a result, a reactor operating at constant power has a fixed steady-state equilibrium concentration, but when lowering reactor power, the 135Xe concentration can increase enough to effectively shut down the reactor. Without enough neutrons to offset their absorption by 135Xe, nor to burn the built-up xenon, the reactor has to be kept in shutdown state for 1–2 days until enough of the 135Xe decays.

135Xe beta-decays with half-life of 9.2 hours to 135Cs; a poisoned core will spontaneously recover after several half-lives. After about 3 days of shutdown, the core can be assumed to be free of 135Xe, without it introducing errors into the reactivity calculations. [4]

The inability of the reactor to be restarted in such state is called xenon precluded start up or dropping into an iodine pit; the duration of this situation is known as xenon dead time, poison outage, or iodine pit depth. Due to the risk of such situations, in the early Soviet nuclear industry, many servicing operations were performed on running reactors, as downtimes longer than an hour led to xenon buildup that could keep the reactor offline for significant time, lower the production of 239Pu, required for nuclear weapons, and would lead to investigations and punishment of the reactor operators. [5]

Xenon-135 oscillations

The interdependence of 135Xe buildup and the neutron flux can lead to periodic power fluctuations. In large reactors, with little neutron flux coupling between their regions, flux nonuniformities can lead to formation of xenon oscillations, periodic local variations of reactor power moving through the core with a period of about 15 hours. A local variation of neutron flux causes increased burnup of 135Xe and production of 135I, depletion of 135Xe increases the reactivity in the core region. The local power density can change by a factor of three or more, while the average power of the reactor stays more or less unchanged. Strong negative temperature coefficient of reactivity causes damping of these oscillations, and is a desired reactor design feature. [4]

Iodine pit behavior

Development of (1) concentration of Xe and (2) reactor reactivity after reactor shutdown. (Until shutdown the neutron flux was ph = 10 neutrons/m s.) Reactor shutdown xe chart en.png
Development of (1) concentration of Xe and (2) reactor reactivity after reactor shutdown. (Until shutdown the neutron flux was φ = 10 neutrons/m s.)

The reactivity of the reactor after the shutdown first decreases, then increases again, having a shape of a pit; this gave the "iodine pit" its name. The degree of poisoning, and the depth of the pit and the corresponding duration of the outage, depends on the neutron flux before the shutdown. Iodine pit behavior is not observed in reactors with neutron flux density below 5×1016 neutrons m−2s−1, as the 135Xe is primarily removed by decay instead of neutron capture. As the core reactivity reserve is usually limited to 10% of Dk/k, thermal power reactors tend to use neutron flux at most about 5×1013 neutrons m−2s−1 to avoid restart problems after shutdown. [4]

The concentration changes of 135Xe in the reactor core after its shutdown is determined by the short-term power history of the reactor (which determines the initial concentrations of 135I and 135Xe), and then by the half-life differences of the isotopes governing the rates of its production and removal; if the activity of 135I is higher than activity of 135Xe, the concentration of 135Xe will rise, and vice versa.

During reactor operation at a given power level, a secular equilibrium is established within 40–50 hours, when the production rate of iodine-135, its decay to xenon-135, and its burning to xenon-136 and decay to caesium-135 are keeping the xenon-135 amount in the reactor constant at a given power level.

The equilibrium concentration of 135I is proportional to the neutron flux φ. The equilibrium concentration of 135Xe, however, depends very little on neutron flux for φ > 1017 neutrons m−2s−1.

Increase of the reactor power, and the increase of neutron flux, causes a rise in production of 135I and consumption of 135Xe. At first, the concentration of xenon decreases, then slowly increases again to a new equilibrium level as now excess 135I decays. During typical power increases from 50 to 100%, the 135Xe concentration falls for about 3 hours. [6]

Decrease of the reactor power lowers production of new 135I, but also lowers the burn rate of 135Xe. For a while 135Xe builds up, governed by the amount of available 135I, then its concentration decreases again to an equilibrium for the given reactor power level. The peak concentration of 135Xe occurs after about 11.1 hours after power decrease, and the equilibrium is reached after about 50 hours. A total shutdown of the reactor is an extreme case of power decrease. [7]

Design precautions

If sufficient reactivity control authority is available, the reactor can be restarted, but a xenon burn-out transient must be carefully managed. As the control rods are extracted and criticality is reached, neutron flux increases many orders of magnitude and the 135Xe begins to absorb neutrons and be transmuted to 136Xe. The reactor burns off the nuclear poison. As this happens, the reactivity increases and the control rods must be gradually re-inserted or reactor power will increase. The time constant for this burn-off transient depends on the reactor design, power level history of the reactor for the past several days (therefore the 135Xe and 135I concentrations present), and the new power setting. For a typical step up from 50% power to 100% power, 135Xe concentration falls for about 3 hours. [6]

The first time 135Xe poisoning of a nuclear reactor occurred was on September 28, 1944, in Pile 100-B at the Hanford Site. Reactor B was a plutonium production reactor built by DuPont as part of the Manhattan Project. The reactor was started on September 27, 1944, but the power dropped unexpectedly shortly after, leading to a complete shutdown on the evening of September 28. Next morning the reaction restarted by itself. The physicist John Archibald Wheeler, working for DuPont at the time, together with Enrico Fermi were able to identify that the drop in the neutron flux and the consequent shutdown was caused by the accumulation of 135Xe in the reactor fuel. Fortunately, the reactor was built with spare fuel channels that were then used to increase the normal operating levels of the reactor, thus increasing the burn-up rate of the accumulating 135Xe. [8]

Reactors with large physical dimensions, e.g. the RBMK type, can develop significant nonuniformities of xenon concentration through the core. Control of such non-homogeneously poisoned cores, especially at low power, is a challenging problem. The Chernobyl disaster occurred after recovering Reactor 4 from a nonuniformly poisoned state. Reactor power was significantly reduced in preparation for a test, to be followed by a scheduled shutdown. Just before the test, the power plummeted in part due to the accumulation of 135Xe as a result of the low burn-up rate at low power. Operators withdrew most of the control rods in an attempt to bring the power back up. Unbeknownst to the operators, these and other actions put the reactor in a state where it was exposed to a feedback loop of neutron power and steam production. A flawed shutdown system then caused a power surge that led to the explosion and destruction of reactor 4.

The iodine pit effect has to be taken in account for reactor designs. High values of power density, leading to high production rates of fission products and therefore higher iodine concentrations, require higher amount and enrichment of the nuclear fuel used to compensate. Without this reactivity reserve, a reactor shutdown would preclude its restart for several tens of hours until 135I/135Xe sufficiently decays, especially shortly before replacement of spent fuel (with high burnup and accumulated nuclear poisons) with fresh one.

Fluid fuel reactors cannot develop xenon inhomogeneity because the fuel is free to mix. Also, the Molten Salt Reactor Experiment demonstrated that spraying the liquid fuel as droplets through a gas space during recirculation can allow xenon and krypton to leave the fuel salts. [lower-alpha 2]

Notes

  1. Xenon-136 undergoes double beta decay with an extremely long half-life of 2.165×1021 years.
  2. Removing 135Xe from neutron exposure also means that the reactor will produce more of the long-lived fission product 135Cs. However, thanks to the short 3.8-minute half-life of 137Xe it will mostly decay in the fuel, leaving the 135Cs significantly less contaminated with the far more active 137Cs and so more suitable for treatments like nuclear transmutation.

Related Research Articles

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

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

<span class="mw-page-title-main">Nuclear 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">Scram</span> Emergency shutdown of a nuclear reactor

A scram or SCRAM is an emergency shutdown of a nuclear reactor effected by immediately terminating the fission reaction. It is also the name that is given to the manually operated kill switch that initiates the shutdown. In commercial reactor operations, this type of shutdown is often referred to as a "scram" at boiling water reactors (BWR), a "reactor trip" at pressurized water reactors and EPIS at a CANDU reactor. In many cases, a scram is part of the routine shutdown procedure, which serves to test the emergency shutdown system.

<span class="mw-page-title-main">Natural nuclear fission reactor</span> Naturally occurring uranium self-sustaining nuclear chain reactions

A natural nuclear fission reactor is a uranium deposit where self-sustaining nuclear chain reactions occur. The conditions under which a natural nuclear reactor could exist were predicted in 1956 by Paul Kuroda. The remnants of an extinct or fossil nuclear fission reactor, where self-sustaining nuclear reactions have occurred in the past, are verified by analysis of isotope ratios of uranium and of the fission products. This was first discovered in 1972 in Oklo, Gabon by Francis Perrin under conditions very similar to Kuroda's predictions.

<span class="mw-page-title-main">Control rod</span> Device used to regulate the power of a nuclear reactor

Control rods are used in nuclear reactors to control the rate of fission of the nuclear fuel – uranium or plutonium. Their compositions include chemical elements such as boron, cadmium, silver, hafnium, or indium, that are capable of absorbing many neutrons without themselves decaying. These elements have different neutron capture cross sections for neutrons of various energies. Boiling water reactors (BWR), pressurized water reactors (PWR), and heavy-water reactors (HWR) operate with thermal neutrons, while breeder reactors operate with fast neutrons. Each reactor design can use different control rod materials based on the energy spectrum of its neutrons. Control rods have been used in nuclear aircraft engines like Project Pluto as a method of control.

<span class="mw-page-title-main">Nuclear fission product</span> Atoms or particles produced by nuclear fission

Nuclear fission products are the atomic fragments left after a large atomic nucleus undergoes nuclear fission. Typically, a large nucleus like that of uranium fissions by splitting into two smaller nuclei, along with a few neutrons, the release of heat energy, and gamma rays. The two smaller nuclei are the fission products..

A neutron reflector is any material that reflects neutrons. This refers to elastic scattering rather than to a specular reflection. The material may be graphite, beryllium, steel, tungsten carbide, gold, or other materials. A neutron reflector can make an otherwise subcritical mass of fissile material critical, or increase the amount of nuclear fission that a critical or supercritical mass will undergo. Such an effect was exhibited twice in accidents involving the Demon Core, a subcritical plutonium pit that went critical in two separate fatal incidents when the pit's surface was momentarily surrounded by too much neutron reflective material.

In nuclear engineering, the void coefficient is a number that can be used to estimate how much the reactivity of a nuclear reactor changes as voids form in the reactor moderator or coolant. Net reactivity in a reactor is the sum total of multiple contributions, of which the void coefficient is but one. Reactors in which either the moderator or the coolant is a liquid typically will have a void coefficient value that is either negative or positive. Reactors in which neither the moderator nor the coolant is a liquid will have a void coefficient value equal to zero. It is unclear how the definition of "void" coefficient applies to reactors in which the moderator/coolant is neither liquid nor gas.

<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.

Naturally occurring samarium (62Sm) is composed of five stable isotopes, 144Sm, 149Sm, 150Sm, 152Sm and 154Sm, and two extremely long-lived radioisotopes, 147Sm (half life: 1.06×1011 y) and 148Sm (6.3×1015 y), with 152Sm being the most abundant (26.75% natural abundance). 146Sm is also fairly long-lived, but is not long-lived enough to have survived in significant quantities from the formation of the Solar System on Earth, although it remains useful in radiometric dating in the Solar System as an extinct radionuclide. A 2012 paper revising the estimated half-life of 146Sm from 10.3(5)×107 y to 6.8(7)×107 y was retracted in 2023. It is the longest-lived nuclide that has not yet been confirmed to be primordial.

Naturally occurring xenon (54Xe) consists of seven stable isotopes and two very long-lived isotopes. Double electron capture has been observed in 124Xe and double beta decay in 136Xe, which are among the longest measured half-lives of all nuclides. The isotopes 126Xe and 134Xe are also predicted to undergo double beta decay, but this has never been observed in these isotopes, so they are considered to be stable. Beyond these stable forms, 32 artificial unstable isotopes and various isomers have been studied, the longest-lived of which is 127Xe with a half-life of 36.345 days. All other isotopes have half-lives less than 12 days, most less than 20 hours. The shortest-lived isotope, 108Xe, has a half-life of 58 μs, and is the heaviest known nuclide with equal numbers of protons and neutrons. Of known isomers, the longest-lived is 131mXe with a half-life of 11.934 days. 129Xe is produced by beta decay of 129I ; 131mXe, 133Xe, 133mXe, and 135Xe are some of the fission products of both 235U and 239Pu, so are used as indicators of nuclear explosions.

<span class="mw-page-title-main">Isotopes of iodine</span> Nuclides with atomic number of 53 but with different mass numbers

There are 37 known isotopes of iodine (53I) from 108I to 144I; all undergo radioactive decay except 127I, which is stable. Iodine is thus a monoisotopic element.

Caesium (55Cs) has 41 known isotopes, the atomic masses of these isotopes range from 112 to 152. Only one isotope, 133Cs, is stable. The longest-lived radioisotopes are 135Cs with a half-life of 1.33 million years, 137
Cs
with a half-life of 30.1671 years and 134Cs with a half-life of 2.0652 years. All other isotopes have half-lives less than 2 weeks, most under an hour.

<span class="mw-page-title-main">Nuclear reactor physics</span> Field of physics dealing with nuclear reactors

Nuclear reactor physics is the field of physics that studies and deals with the applied study and engineering applications of chain reaction to induce a controlled rate of fission in a nuclear reactor for the production of energy. Most nuclear reactors use a chain reaction to induce a controlled rate of nuclear fission in fissile material, releasing both energy and free neutrons. A reactor consists of an assembly of nuclear fuel, usually surrounded by a neutron moderator such as regular water, heavy water, graphite, or zirconium hydride, and fitted with mechanisms such as control rods which control the rate of the reaction.

In applications such as nuclear reactors, a neutron poison is a substance with a large neutron absorption cross-section. In such applications, absorbing neutrons is normally an undesirable effect. However, neutron-absorbing materials, also called poisons, are intentionally inserted into some types of reactors in order to lower the high reactivity of their initial fresh fuel load. Some of these poisons deplete as they absorb neutrons during reactor operation, while others remain relatively constant.

<span class="mw-page-title-main">Fission products (by element)</span> Breakdown of nuclear fission results

This page discusses each of the main elements in the mixture of fission products produced by nuclear fission of the common nuclear fuels uranium and plutonium. The isotopes are listed by element, in order by atomic number.

<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 about $130 million.

Xenon-135 (135Xe) is an unstable isotope of xenon with a half-life of about 9.2 hours. 135Xe is a fission product of uranium and it is the most powerful known neutron-absorbing nuclear poison, with a significant effect on nuclear reactor operation. The ultimate yield of xenon-135 from fission is 6.3%, though most of this is from fission-produced tellurium-135 and iodine-135.

Shutdown is the state of a nuclear reactor when the fission reaction is slowed significantly or halted completely. Different nuclear reactor designs have different definitions for what "shutdown" means, but it typically means that the reactor is not producing a measurable amount of electricity or heat, and is in a stable condition with very low reactivity.

<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. It also incorporates 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 and also serves as primary coolant.

References

  1. Stacey, Weston M. (2007). Nuclear Reactor Physics. Wiley-VCH. p. 213. ISBN   978-3-527-40679-1.
  2. Staff. "Hanford Becomes Operational". The Manhattan Project: An Interactive History. U.S. Department of Energy, Office of History and Heritage Resources. Archived from the original on October 14, 2010. Retrieved 2013-03-12.
  3. Pfeffer, Jeremy I.; Nir, Shlomo (2000). Modern Physics: An Introductory Text. Imperial College Press. pp. 421 ff. ISBN   1-86094-250-4.
  4. 1 2 3 "Xenon-135 Oscillations". Nuclear Physics and Reactor Theory (PDF). Vol. 2 of 2. U.S. Department of Energy. January 1993. p. 39. DOE-HDBK-1019/2-93. Retrieved 2014-08-21.
  5. Kruglov, Arkadii (15 August 2002). The History of the Soviet Atomic Industry. CRC Press. pp. 57, 60. ISBN   0-41526-970-9.
  6. 1 2 Xenon decay transient graph
  7. DOE Fundamentals Handbook: Nuclear Physics and Reactor Theory Volume 2 (PDF). U.S. Department of Energy. January 1993. pp. 35–42. Archived from the original (PDF) on 2012-11-09. Retrieved 2013-03-12.
  8. "John Wheeler's Interview (1965)". www.manhattanprojectvoices.org. Retrieved 2019-06-19.