You are on page 1of 42

Unit I - NUCLEAR PHYSICS The nuclear model of the atom describes how the three basic sub atomic

particles, the proton, the neutron and the electron are arranged. The nucleus is the centre of the atom and is positive in charge. It is made up of protons and neutrons. Negative electrons orbit the atom. The atom is made up mostly of empty space. The nuclear model of the atom consists of a nucleus (meaning: 'nut' or 'kernel') which is surrounded by orbiting electrons. The atom is made up mostly of empty space. The nucleus is made up of protons and neutrons. Protons are positive, neutrons are neutral and electrons are negative. In a neutral atom the number of protons (positive charge) = the number of electrons (negative charge) Protons determine the identity of an element. The number of protons is called the Atomic Number. Each element has a unique Atomic Number. eg. All atoms of Carbon have an Atomic Number of 6. ie. they all contain 6 protons. All atoms of oxygen contain 8 protons. ie. They have an Atomic Number of 8. The Atomic Number for each element can be found in the Periodic Table. Neutrons help stabilise atoms. If there are too many or too few neutrons the atom becomes unstable. Atoms of the same element that contain a different number of neutrons are called isotopes. Electrons are involved in chemical reactions. During a reaction electrons are either transferred or shared between chemical species. The noble gases are very unreactive because they have a complete number of electrons in their outer shell.

Equivalent weight (also known as gram equivalent) is a term which has been used in several contexts in chemistry. In its most general usage, it is the mass of one equivalent, that is the mass of a given substance which will: supply or react with one mole of hydrogen cations H+ in an acidbase reaction; or supply or react with one mole of electrons e in a redox reaction.[1] Equivalent weight has the dimensions and units of mass, unlike atomic weight, which is dimensionless. Equivalent weights were originally determined by experiment, but (insofar as they are still used) are now derived from molar masses. Additionally, the equivalent weight of a compound can be calculated by dividing the molecular weight by the number of positive or negative electrical charges that result from the dissolution of the compound. Nuclei are made up of protons and neutron, but the mass of a nucleus is always less than the sum of the individual masses of the protons and neutrons which constitute it. The difference is a measure of the nuclear binding energy which holds the nucleus together. This binding energy can be calculated from the Einstein relationship: Nuclear binding energy = mc2

For the alpha particle m= 0.0304 u which gives a binding energy of 28.3 MeV.

The enormity of the nuclear binding energy can perhaps be better appreciated by comparing it to the binding energy of an electron in an atom. The comparison of the alpha particle binding energy with the binding energy of the electron in a hydrogen atom is shown below. The nuclear binding energies are on the order of a million times greater than the electron binding energies of atoms.

Nuclear Binding Energy Curve The binding energy curve is obtained by dividing the total nuclear binding energyby the number of nucleons. The fact that there is a peak in the binding energy curve in the region of stability near iron means that either the breakup of heavier nuclei (fission) or the combining of lighter nuclei (fusion) will yield nuclei which are more tightly bound (less mass per nucleon). The binding energies of nucleons are in the range of millions of electron voltscompared to tens of eV for atomic electrons. Whereas an atomic transition might emit a photon in the range of a few electron volts, perhaps in the visible light region, nuclear transitions can emit gamma-rays with quantum energies in the MeV range.

Radioactivity Radioactivity refers to the particles which are emitted from nuclei as a result of nuclear instability. Because the nucleus experiences the intense conflictbetween the two strongest forces in nature, it should not be surprising that there are many nuclearisotopes which are unstable and emit some kind of radiation. The most common types of radiation are called alpha, beta, and gamma radiation, but there are several other varieties of radioactive decay. Radioactive decay rates are normally stated in terms of their half-lives, and the half-life of a given nuclear species is related to its radiation risk. The different types of radioactivity lead to different decay paths which transmute the nuclei into other chemical elements. Examining the amounts of the decay products makes possibleradioactive dating. Radiation from nuclear sources is distributed equally in all directions, obeying theinverse square law. Alpha Radioactivity

Composed of two protons and two neutrons, the alpha particle is a nucleus of the element helium. Because of its very large mass (more than 7000 times the mass of the beta particle) and its charge, it has a very short range. It is not suitable for radiation therapy since its range is less than a tenth of a millimeter inside the body. Its main radiation hazard comes when it is ingested into the body; it has great destructive power within its short range. In contact with fast-growing membranes and living cells, it is positioned for maximum damage.

Alpha particle emission is modeled as a barrier penetration process. The alpha particle is the nucleus of the helium atom and is the nucleus of highest stability. Alpha Barrier Penetration The energy of emitted alpha particles was a mystery to early investigators because it was evident that they did not have enough energy, according to classical physics, to escape the nucleus. Once an approximate size of the nucleus was obtained by Rutherford scattering, one could calculate the height of the Coulomb barrier at the radius of the nucleus. It was evident that this energy was several times higher than the observed alpha particle energies. There was also an incredible range of half livesfor the alpha particle which could not be explained by anything in classical physics. The resolution of this dilemma came with the realization that there was a finite probability that the alpha particle could penetrate the wall by quantum mechanical tunneling. Using tunneling, Gamow was able to calculate a dependence for the half-life as a function of alpha particle energy which was in agreement with experimental observations. Alpha Binding Energy The nuclear binding energy of the alpha particle is extremely high, 28.3 MeV. It is an exceptionally stable collection of nucleons, and those heavier nuclei which can be viewed as collections of alpha particles (carbon-12, oxygen-16, etc.) are also exceptionally stable. This contrasts with a binding energy of only 8 MeV for helium-3, which forms an intermediate step in the proton-proton fusion cycle.

Alpha, Beta, and Gamma Historically, the products of radioactivity were called alpha, beta, and gamma when it was found that they could be analyzed into three distinct species by either a magnetic field or an electric field.

Penetration of Matter Though the most massive and most energetic of radioactive emissions, the alphaparticle is the shortest in range because of its strong interaction with matter. The electromagnetic gamma ray is extremely penetrating, even penetrating considerable thicknesses of concrete. The electron of beta radioactivity strongly interacts with matter and has a short range.

Half-life, abbreviated t, is the period of time it takes for the amount of a substance undergoing decay to decrease by half. The name was originally used to describe a characteristic of unstable atoms (radioactive decay), but it may apply to any quantity which follows a set-rate decay. The original term, dating to 1907, was "half-life period", which was later shortened to "half-life" in the early 1950s.[1] Half-lives are used to describe quantities undergoing exponential decayfor example, radioactive decay where the half-life is constant over the whole life of the decay, and is a characteristic unit (a natural unit of scale) for the exponential decay equation. However, a half-life can also be defined for non-exponential decay processes, although in these cases the half-life varies throughout the decay process. For a general introduction and description of exponential decay, see the article exponential decay. For a general introduction and description of non-exponential decay, see the article rate law. Corresponding to sediments in environmental processes, if the half-life is greater than the residence time, then the radioactive nuclide will have enough time to significantly alter the concentration. The converse of half-life is doubling time. An exponential decay process can be described by any of the following three equivalent formulas:

where is the initial quantity of the substance that will decay (this quantity may be measured in grams, moles, number of atoms, etc.), is the quantity that still remains and has not yet decayed after a time t, is the half-life of the decaying quantity, is a positive number called the mean lifetime of the decaying quantity, is a positive number called the decay constant of the decaying quantity. The three parameters , , and are all directly related in the following way:

where ln(2) is the natural logarithm of 2 (approximately 0.693). By plugging in and manipulating these relationships, we get all of the following equivalent descriptions of exponential decay, in terms of the half-life:

Regardless of how it's written, we can plug into the formula to get as expected (this is the definition of "initial quantity") as expected (this is the definition of half-life)

, i.e. amount approaches zero as t approaches infinity as expected (the longer we wait, the less remains). Decay by two or more processes Some quantities decay by two exponential-decay processes simultaneously. In this case, the actual half-life T1/2 can be related to the half-lives t1 and t2 that the quantity would have if each of the decay processes acted in isolation:

For three or more processes, the analogous formula is:

In nuclear fission reactors, neutrons cause the fission. When a neutron hits the fuel nucleus, say, U-235, the neutron is absorbed and an isotope, U-236 is formed. The U-236 is invariably in its excited state and has to de-excite. As said earlier, one possibility of de-excitation is fission, and the entire reactor programme depends on this possibility. Let us see the types of interactions that are possible, and how these are quantified. A nuclear reaction is generally symbolised as Target (projectile, ejectile) Residue. Many types of reactions are possible when a neutron hits a nucleus. The probability of each type of reaction happening depends on the nucleus and is very sensitive to the energy of the neutron. In this regard, two neighbouring nuclides (say, of masses A and A+1) may differ drastically. Some of the prominent reactions are explained below: Elastic scattering: The neutron and the nuclide collide and share a part of their kinetic energies. They rebound with speeds different from the original speeds, such that the total kinetic energy before and after the collision remains the same. If the nucleus is stationary before collision, it will gain energy from the neutron and start moving, and the neutron gets slowed down due to loss of kinetic energy. However, the residual nucleus is not excited but is in its ground state.

This is the type of reaction that mostly helps fast neutrons to be slowed down to low energies in a reactor. Can the neutron gain energy? Yes. However, this is obviously possible when the nucleus has higher kinetic energy than the neutron. Inelastic scattering: The neutron and the nuclide collide and rebound with speeds different from the original speeds, but the rebounding nuclide is left in an excited energy state. Hence the total kinetic energy after the collision is less than that before the collision, and this difference accounts for the energy of excitation. If the nucleus is stationary before collision, the neutron must have kinetic energy exceeding the excitation energy, so that such a reaction is possible. Hence inelastic scattering is said to be a threshold reaction, the threshold being the minimum kinetic energy of the neutron required for the reaction to be possible. The excited nucleus subsequently de-excites by emitting radiation. Heavy nuclides have lower thresholds than light nuclides. Though the probability of inelastic scattering is generally lower than elastic, the energy loss to the neutron is higher in an inelastic collision. Inelastic scattering in heavy nuclides degrades fission neutron energies heavily.

Capture: The neutron is absorbed by the target nucleus to form the next higher isotope (of mass A+1), in an excited state of energy. The new isotope de-excites by emitting rays. The neutron is thus lost in this reaction. This is often known as radiative capture.

(n,x) reaction: In this reaction, n represents neutron, and x represents any particle like neutron, proton, deutron, particle, etc. or a combination of such particles. It means that a neutron interaction with a nuclide results in emission of the particle(s) represented by x. (e.g.) if the emitted particle is 2 neutrons are emitted, it is then (n,2n) reaction. Such reactions are generally threshold reactions. Fission: This is the most important reaction upon which the present day nuclear energy programme depends. Nuclear fission is a phenomenon in which a heavy nucleus, splits into two smaller nuclei, called the fission fragments, mostly of unequal masses, one often with nearly half the mass as the other, and rarely of equal masses. This reaction gives off a large amount of energy and emits two or more neutrons, and gamma rays. When a neutron hits a heavy nuclide like U-235, the neutron gets absorbed in the heavy nuclide that gets energetically agitated (or excited). If the new energy state of the heavy nuclide is sufficient for it to split, then it can split to cause fission. The neutrons produced in fission are fast, with an average energy of 2 MeV.

It must be noted that the fission fragments themselves are in excited state, and they de-excite generally by and neutron emissions. The neutron emitted during fission are called prompt neutrons, and those emitted by the fragments after a delay are called delayed neutrons. Similary, prompt and delayed gammas are also emitted. About 80 % of the energy released in fission is carried away by the fission products (and the rest by the other particles), which in turn transfer the energy to the surroundings, making the energy recoverable. Some energy is carried away by particles known as neutrinos, which are chargeless and light, do not interact with any material, and hence their energy is not recoverable. The approximate distribution of fission energy through various emissions is given below: As a fission caused by a neutron involves production of further neutrons, a fission chain reaction becomes possible, and such a chain is ensured in the design of a reactor. The nuclear energy released in fission is

about a million times the chemical energy released in burning a block of coal of equal mass. The following figures show the different stages in fission.

Probabilities of various reactions: A neutron-nuclear collision may result in a variety of reactions. As explained earlier, a neutron-nuclear collision could give rise to elastic or inelastic scattering, fission, or capture. A quantity called the microscopic cross-section is defined to represent the measure of the probabilities of each such reaction. The term cross-section generally refers to a plane surface or area of a cut-out section. The basic meaning of cross-section remains the same in the nuclear jargons as well, though it is used to represent the probability of an interaction. Imagine a projectile approaching a target as in the figure below. Obviously the chance (probability) of a collision depends on the surface area projected by the target. i.e. The probability of collision is larger if the area is larger. If there are a large number of projectiles, the number of projectiles that could hit the surface is larger if the surface area is larger. Arguing in this manner, it is easy to visualize that the probability of collision (i.e. interaction) between the target and the projectile could be expressed by the effective surface area available for the collision.

The idea is extended to represent probabilities of nuclear interactions, but a precise definition consistent with physics is employed. As the boundaries of sub-atomic systems are not sharp, the surface area supposed to be presented by a target can be larger than its geometric surface area. Thus it is now clear, in the nuclear jargons, that the term cross-section represents a measure of the probability of interaction between a projectile and a target. This quantity is usually called microscopic cross-section, represented by the symbol , and is expressed in area units. The practical unit of microscopic cross-section is a barn, where 1 barn = 10-24cm2 (i.e 10-28 m2). The neutron-nuclear microscopic cross-sections vary significantly from nuclide to nuclide and drastically with respect to neutron energy. Knowledge of cross-sections is essential to understand the physics behind the design of a nuclear reactor and also to select the composition for reactor materials. Properly averaged cross-sections are usually used. The table below gives representative (average) values of microscopic cross-sections for various materials of importance to thermal and fast reactors. Cross-sections of some important reactor materials Cross-sections in barns Material Thermal neutrons (E = 0.0253 eV) Fission 92-U-233 Fissile 92-U-235 94-Pu-239 Fertile Clad 92-U-238 90-Th-232 40-Zr Steel Light Water Coolant Heavy Water 11-Na-23 Control Rod Fission Products 5-B-10 48-Cd 54-Xe-135 36-Kr-83 62-Sm-149 528.45 585.086 747.401 0.00001177 0.0 Capture 45.76 98.6864 270.329 2.71692 7.4 0.185396 3.08668 0.664 0.0013 0.528 3840.0 2524.15 2636300.0 207.667 40144.3 Fast neutrons (E > 100 eV) Fission 2.7323 1.9041 1.7973 0.042758 0.010193 Capture 0.27176 0.55549 0.49614 0.33188 0.38157 0.0265766 0.0170228 0.00051554 0.00011459 0.0027511 2.73462 0.266766 0.0059985 0.235944 1.91883

Fission chain reaction: As indicated earlier, the availability of neutrons from fission helps continue the chain of fission reactions. Since more than one neutron is generated in fission, it is possible to achieve the fission chain reaction, even after accounting for other reactions of neutron, like absorption. Thus one neutron must be effectively available, after every fission for continuing the chain. The rate of production of fission neutrons depend on the cross-sections, and the composition of the fissioning material. It can be seen from the cross-section table that the cross-sections are higher when the neutron is slow than when they are fast. We must appropriately choose the fuel composition and neutron energy region to work with, to achieve the chain reaction A cross section is the effective area which governs the probability of some scattering or absorption event. Together with particle density and path length, it can be used to predict the total scattering probability via the Beer-Lambert law. In nuclear and particle physics, the concept of a cross section is used to express the likelihood of interaction between particles. When particles in a beam are thrown against a foil made of a certain substance, the cross section is a hypothetical area measure around the target particles of the substance (usually its atoms) that represents a surface. If a particle of the beam crosses this surface, there will be some kind of interaction. The term is derived from the purely classical picture of (a large number of) point-like projectiles directed to an area that includes a solid target. Assuming that an interaction will occur (with 100% probability) if the projectile hits the solid, and not at all (0% probability) if it misses, the total interaction probability for the single projectile will be the ratio of the area of the section of the solid (thecross section, represented by ) to the total targeted area. This basic concept is then extended to the cases where the interaction probability in the targeted area assumes intermediate values - because the target itself is not homogeneous, or because the interaction is mediated by a non-uniform field.

Unit II - NUCLEAR REACTIONS AND REACTION MATERIALS The Fission Process When an incident subatomic projectile is absorbed by a heavy nucleus, the resulting compound nucleus is produced in a very excited state. The excitation energy E* of the compound nucleus equals the binding energy of the incident particle to it plus the kinetic energy of the incident projectile, less a negligible recoil energy of the compound nucleus. The excited compound nucleus is very unstable, with a lifetime of about 10~14 s. The "nuclear fluid" of such an excited nucleus undergoes large oscillations and deformations in shape. If the compound nucleus is sufficiently excited, it may, during one of its oscillations in shape, deform into an elongated or dumbbell configuration in which the two ends Coulombically repel each other and the nuclear forces, being very short ranged, are no longer able to hold the two ends together. The two ends then separate (scission) within ~ 10~20 s into two nuclear pieces, repelling each other with such tremendous Coulombic force that many of the orbital electrons are left behind. Two highly charged fission fragments are thus created.6 For example, in neutron-induced fission of 235U, the reaction sequence, immediately following scission of the compound nucleus, is n + 2gU > 2^U* > Y+n + Y+m + (n + m)e~ , (6.32) where the Y# and YL indicate, respectively, the heavy and light primary fission products, and the ionic charge n and m on the fission fragments is about 20. The primary fission fragments are produced in such highly excited nuclear states that neutrons can "boil" off them. Anywhere from 0 to about 8 neutrons, denoted by i^p, evaporate from the primary fission fragments within about 10~17 s of the scission or splitting apart of the compound nucleus. These neutrons are called prompt fission neutrons, hence the subscript p in z/p, in order to distinguish them from delayed neutrons emitted at later times during radioactive decay of the fission products. Following prompt neutron emission, the fission fragments are still in excited states, but with excitation energies insufficient to cause particle emission. They promptly decay to lower energy levels only by the emission of prompt gamma rays 7p. This emission is usually completed within about 2 x 10~14 s after the emission of the prompt neutrons. The highly charged fission fragments pass through the surrounding medium causing millions of Coulombic ionization and excitation interacts with the electrons of the medium. As the fission fragments slow, they gradually acquire electrons reducing their ionization charge, until, by the time they are stopped, they have become electrically neutral atoms. This transfer of the fission fragments' kinetic energy to the ambient medium takes about 10~12 s. Thus, after about 10~12 s following scission, the fission reaction may be written 2jgU > 255U* > YH + YL + i/p(Jn) + 7p- (6-33) In the fission process, the number of neutrons and protons must be conserved. For neutron-induced fission of 2g|U this requires AL + AH + VP = 236 NL + NH + vp = 144 (6.34) ZL + ZH = 92 After prompt neutron and gamma ray emission, the fission fragments are termed fission products. The fission products are still radioactively unstable since they still have too many neutrons compared to protons to form stable atoms. They thus form the start of decay chains whose members radioactively decay, usually by isobaric /3~ decay, (constant A) until a stable end-nuclide is reached. The half-lives of the fission product daughters range from fractions of a second to many thousands of years. The proper management of these long-lived fission products is a major challenge to proponents of nuclear fission power. Fusion Reactions An alternative to the fission of heavy nuclides for extracting nuclear energy is the fusion of light nuclides. Some possible fusion reactions involving the lightest nuclides are listed below. ?D + ?D>?T + }H, Q = 4.03 MeV ?D + 2D^He + Jn, Q = 3.27 MeV ?D + ?T > ^He + Jn, Q = 17.59 MeV ?D + 3He^He+}H, Q = 18.35 MeV 3T + sT>4He + 2in) g = 11.33 MeV }H + |Li > *He + f>He, Q = 4.02 MeV 5B > 3(|He), Q = 8.08 MeV Although these fusion reactions are exoergic, they have threshold energies because of the repulsive Coulomb forces between the two reactants. Consequently, for these reactions to occur, it is necessary that the reactants have sufficient incident kinetic energy to overcome the Coulomb barrier and allow the two reacting nuclides to reach each other, thereby enabling the nuclear forces in each particle to interact and precipitate the fusion reaction. The kinetic energy needed is typically a few keV to several hundred keV. The needed incident kinetic energy could be provided by a particle accelerator to give one of the reactants sufficient kinetic energy and to direct it towards a target composed of the second reactant. Indeed, this is done in many laboratories in which fusion reactions are studied. However, the energy needed to operate the accelerator far exceeds the energy released by the subsequent fusion reactions, and this approach is not a practical source of fusion energy.

Thermonuclear Fusion The reactants can also be given sufficient kinetic energy to interact if they are raised to high temperature T so that their thermal motion provides the necessary kinetic energy (average kinetic energy Eav kT), where k is Boltzmann's constant). Fusion reactions induced by the thermal motion of the reactants are often called thermonuclear reactions. However, the temperatures needed are very high, typically 10-300 xlO6 K. At these temperatures, the reactants exist as a plasma, a state of matter in which the electrons are stripped from the nuclei so that the hot plasma consists of electrons and positively charged nuclei. The challenge to produce fusion energy in useful quantities is to confine a high temperature plasma of light elements for sufficiently long times and at sufficiently high nuclide densities to allow the fusion reactions to produce more energy than is required to produce the plasma. Three approaches are available. Gravitational Confinement One pilasma confinement approach, the basis of life itself, uses gravity to confine a plasma of light nuclei. Such is the method used by stars where, deep in their interiors, extreme temperatures and pressures are produced. For example, at our sun's core, the temperature is about 15 MK and the pressure is 4 x 1016 MPa (~ 400 billion earth atmospheres). Obviously, this is not a very useful approach to use with earthbound energy-generating systems. Magnetic Confinement The second confinement approach, is to use magnetic fields to confine the plasma. Since a plasma is composed of fast moving charged particles, their motion can be affected by external magnetic fields. Currently, several multinational efforts are underway to design fusion plasma devices with clever arrangements of electromagnetic fields to heat and confine a plasma. The hot plasma fluid in these devices is generally very tenuous (~ 1015 particles per cm3) and is notoriously unstable, exhibiting many modes of breaking up and allowing the plasma to come into contact with the walls of the reaction chamber where it is immediately chilled to normal temperatures and any possibility of thermonuclear reactions ceases. These instabilities increase dramatically as the density and temperature of the plasma increase. Moreover, the magnetic pressure needed to confine a plasma is enormous. For example, to confine a plasma with an average thermal energy of 10 keV per particle and a particle density equal to atmospheric density (~ 1019 cm""3) requires pressures exceeding 10 atm [Mayo 1998]. The external magnetic field coils and their support structure must be designed to withstand such pressures. Finally, the reaction chamber must be surrounded with some system to cool the chamber walls, which are heated by the charged fusion products, and to stop and absorb the more penetrating fusion neutrons. The heat generated by the fusion products must then be converted into electrical energy before these magnetic confinement devices can be considered a practical source of nuclear energy. All current experimental research into fusion power is based on the D-T fusion reaction, which requires the lowest plasma temperature to overcome the Coulomb barrier problem. This reaction is ID + ?T > ^He (3.54 MeV) + Jn (14.05 MeV). (6.46) The radioactive tritium (7\/2 = 12.3 y) needed for this reaction can be generated by having the fusion neutron produced in this reaction interact with lithium through the following reactions: Jn + aLi > 4 2Re + JT Q = 4.78 MeV, in + l L i ^ H e , 3 T , i Q = 4.78 MeV. Although great advances have been made in magnetic confinement, no system using this approach has yet to produce more energy than that used to operate the system. Yet the enormous amount of energy that would be made available if practical fusion devices could be fabricated motivates the current research efforts. Radioactivity Radioactive nuclei and their radiations have properties that are the basis of many of the ideas and techniques of atomic and nuclear physics. We have seen that the emission of alpha and beta particles has led to the concept that atoms are composed of smaller fundamental units. The scattering of alpha particles led to the idea of the nucleus, which is fundamental to the models used in atomic physics. The discovery of isotopes resulted from the analysis of the chemical relationships between the various radioactive elements. The bombardment of the nucleus by alpha particles caused the disintegration of nuclei and led to the discovery of the neutron and the present model for the composition of the nucleus. The discovery of artificial, or induced, radioactivity started a new line of nuclear research and hundreds of artificial nuclei have been produced by many different nuclear reactions. The investigation of the emitted radiations from radionuclides has shown the existence of nuclear energy levels similar to the electronic energy levels. The identification and the classification of these levels are important sources of information about the structure of the nucleus. A number of radioactive nuclides occur naturally on the earth. One of the most important is fgK, which has an isotopic abundance of 0.0118% and a half-life of 1.28 x 109 y. Potassium is an essential element needed by plants and animals, and is an important source of human internal and external radiation exposure. Other naturally occurring radionuclides are of cosmogenic origin. Tritium (fH) and l$C are produced by cosmic ray interactions in the upper atmosphere, and also can cause measurable human exposures. *|C (half life 5730 y), which is the result of a neutron reaction with ^N in

the atmosphere, is incorporated into plants by photosynthesis. By measuring the decay of 14C in ancient plant material, the age of the material can be determined. Other sources of terrestrial radiation are uranium, thorium, and their radioactive progeny. All elements with Z > 83 are radioactive. Uranium and thorium decay into daughter radionuclides, forming a series (or chain) of radionuclides that ends with a stable isotope of lead or bismuth.

In all nuclear interactions, including radioactive decay, there are several quantities that are always conserved or unchanged by the nuclear transmutation. The most important of these conservation laws include: Conservation of charge, i.e., the number of elementary positive and negative charges in the reactants must be the same as in the products. Conservation of the number of nucleons, i.e., A is always constant. With the exception of EC and (3^ radioactive decay, in which a neutron (proton) transmutes into a proton (neutron), the number of protons and neutrons is also generally conserved. Conservation of mass/energy (total energy). Although, neither rest mass nor kinetic energy is generally conserved, the total (rest-mass energy equivalent plus kinetic energy) is conserved. Conservation of linear momentum. This quantity must be conserved in all inertial frames of reference. Conservation of angular momentum. The total angular momentum (or the spin) of the reacting particles must always be conserved. Nuclear Chain Reactions A chain reaction refers to a process in which neutrons released in fission produce an additional fission in at least one further nucleus. This nucleus in turn produces neutrons, and the process repeats. The process may be controlled (nuclear power) or uncontrolled (nuclear weapons).

U235 + n fission + 2 or 3 n + 200 MeV If each neutron releases two more neutrons, then the number of fissions doubles each generation. In that case, in 10 generations there are 1,024 fissions and in 80 generations about 6 x 10 23 (a mole) fissions. NUCLEAR CYCLES The procurement, preparation, utilization, and ultimate disposition of the fuel for a reactor is called the reactor fuel cycle. Those parts of a fuel cycle that precede the use of the fuel in a reactor are collectively called the front end of the fuel cycle. That portion related to the fuel after it has been withdrawn from a reactor is called the back end of the fuel cycle. The mass flows indicated in the figure are in kilograms for a nominal 1 ,000 MWe ( l GWe) plant operated for 1 year at a capacity factor of O.75-that is, the masses correspond to 0.75 GWe-year. The fuel cycle for a BWR plant is entirely similar except that the mass flows are somewhat different. The cycle begins with the mining of uranium ore. This ore contains uranium in the form of a number of complex oxides and is reduced to the oxide U3 0s , which is then converted to uranium hexafluoride, UF6 , the form in which uranium is accepted at current isotope enrichment plants. Approximately 0.5% of the uranium is lost in the conversion to UF6. Following enrichment to about 3 W /0 in 23S U, uranium can be used in the blanket. This uranium, together with the uranium and plutonium from the reprocessing plant, is fabricated into core and blanket assemblies and introduced into the reactor. The spent fuel, after a cooling period in the spent fuel pool, is sent to be reprocessed. Since an LMFBR produces more fissile plutonium than it needs for its own operation, the excess plutonium is sold for use in other LMFBRs. The fuel cycles for breeders operating on thorium and 233 U are essentially the same as that in Fig. 4.40, except that the primary fuel is now thorium and the reprocessed product is Th02 and 233 U02 . In this cycle, there is no enrichment necessary since the reactor is fueled with natural uranium. The Canadians also do not reprocess their spent fuel in view of the abundant uranium resources in Canada. Should enriched uranium ever be used in CANDU-like reactors, UF6 conversion and enrichment steps would have to be introduced into the cycle.

Uranium Production The discovery of fission led to two potential routes to the production of fissile material for the first nuclear weapons by the United States in the 1940s. The first involved separating uranium-235 from uranium-238 isotopes in natural uranium by gaseous diffusion. The second path produced plutonium-239 by bombarding fertile uranium-238 in a nuclear reactor. But both approaches began with mining of uranium ore. Today, the production of fissile fuel for nuclear power reactors uses many methods originally developed for producing nuclear weapons. This unit addresses the metallurgy of uranium, its conversion into gaseous uranium hexafluoride required for enrichment processes, and the fabrication of fuel rods from the enriched uranium hexafluoride. The enrichment processes are covered in a separate unit. Uranium, the heaviest naturally occurring element, is about 500 times more prevalent than gold and about as abundant as tin. However, it is usually found in trace concentrations. The most common mineral containing uranium is pitchblend which is composed of UO2 in the presence of smaller amounts of UO3. If the concentration of pitchblend is great enough for it to be extracted economically, the material is known as an ore. Deposits containing more than 0.1% pitchblend are economically viable. Deposits containing more than 20% pitchblend are rare. In 2007, Canada, Australia, and Kazakhstan accounted for over half of the worlds uranium production. The cost of uranium is determined by the concentration of uranium in the ore: the higher the concentration, the lower the cost. Mining and Preparation of Yellowcake The objective of uranium extraction chemistry is the preparation of U3O8, called yellowcake (Figure 1). Extraction of uranium is often difficult, and the metallurgical procedures vary with the geological environment of the ore. Traditional methods of open pit or underground mining are used to extract uranium ore. More recently, in situ leaching has also been used to extract and concentrate the ore. This technique circulates oxygenated groundwater through a porous ore body to dissolve the uraniumcontaining compounds and bring them to the surface.

Drums of Yellowcake

The ore is first crushed and ground to liberate mineral particles (Figure 2). An amphoteric oxide is then leached with sulfuric acid. UO3(s) + 2H+(aq) UO22+(aq) + H2O

UO22+(aq) + 3SO42-(aq) UO2(SO4)34-(aq) A basic oxide is converted by a similar process to the water-soluble UO2(CO3)34-(aq) ion. Preparation of Yellowcake

Preparation of yellow cake, purified U3O8(s) Two methods are used to concentrate and purify the uranium: ion exchange and solvent extraction. Solvent extraction, the more common method, uses tertiary amines in an organic kerosene solvent in a continuous process. First the amines, R3N, react with sulfuric acid: 2 R3N(org) + H2SO4(aq) (R3NH)2SO4(org) Then the amine sulfate extracts the uranyl ions into the organic phase while the impurities remain in the aqueous phase. In the case of the uranyl sulfate ion, the following reaction occurs: (R3NH)2SO4(org) + UO2(SO4)34-(aq) (R3NH)4UO2(SO4)3(org) + 2SO42-(aq)

The solvents are removed by evaporation in a vacuum, and ammonium diuranate, (NH 4)2U2O7, is precipitated by adding ammonia to neutralize the solution. The diuranate is then heated to yield solid U3O8.

Refining and converting U3O8 to UF 6 At the refinery, the yellowcake is dissolved in nitric acid. The resulting solution of uranium nitrate, UO2(NO3)2 6H2O, is fed into a continuous solvent extraction process. The uranium is extracted into an organic phase (kerosene) with tributyl phosphate, and the impurities remain again in the aqueous phase. After this purification, the uranium is washed out of the kerosene with dilute nitric acid and concentrated by evaporation to pure UO2(NO3)26H2O. Heating yields pure UO3. The initial separation and refining processes generate large volumes of acid and organic waste. It is necessary to enrich the U-235 isotope concentration from its natural composition of 0.7% for use as reactor fuel or weapons components. Reactor grade uranium contains from 0.8 to 8.0% U-235, while weapons grade uranium contains more than 90% of the lighter U-235 isotope. Because the uranium isotopes have identical chemical properties, the processes employed for enrichment must use physical techniques which take advantage of the slight differences in their masses. The two enrichment methods used today, centrifugation and diffusion, require that the uranium be in a gaseous form, uranium hexafluoride, UF6(g). Although enrichment involves physical processes, chemistry plays an important role in synthesizing UF6 gas and returning the UF6 enriched in U-235 to a solid, UO2. Conversion to the hexafluoride involves The UO3 is reduced with hydrogen in a kiln: UO3(s) + H2(g) ) UO2(s) + H2O(g) the following sequence of reactions.

The uranium dioxide is then reacted with hydrogen fluoride to form uranium tetrafluoride: UO2(s) + 4HF(g) ) UF4(s) + 4H2O(g)

The tetrafluoride is then fed into a fluidized bed reactor and reacted with gaseous fluorine to obtain the hexafluoride: UF4(s) + F2(g) ) UF6(g) Uranium hexafluoride is now suitable feedstock for the gaseous diffusion or centrifugationenrichment processes. Production of uranium metal Production of solid fuel rods from uranium hexafluoride gas enriched in U-235 requires another series of chemical and metallurgical processes (Figure 3). Production of Fuel Rods from UF6

The uranium hexafluoride is first reduced to uranium tetrafluoride with hydrogen. UF6(g) + H2(g) UF4(s) + 2HF(g)

Uranium metal is then produced by reducing the uranium tetrafluoride with either calcium or magnesium, both active group IIA metals that are excellent reducing agents. UF4(s) + 2Ca(s) U(s) + 2CaF2(s)

Production of uranium dioxide, often used as a reactor fuel, from uranium hexafluoride can be accomplished by the following reaction. UF6(g) + 2H2O(g) + H2(g) UO2(s) + 6HF(g) Reactor fuel consists of ceramic pellets formed from pressed uranium oxide, which is sintered (baked) at a high temperature (over 1400C). The pellets are then placed in metal tubes made of a zirconium alloy or stainless steel and sealed in an atmosphere of helium to form fuel rods. The fuel rods are then grouped in clusters to form the fuel assemblies, which are placed into the reactor core (Figure 4). The individual rods for a pressurized water reactor (PWR) are about 1 inch in diameter and 4 meters in length. Fuel

assemblies for PWRs contain from 179 to 264 rods, and a fully fueled PRW will contain from 121 to 193 assemblies. A PWR must be shut down for refueling. This occurs at intervals of 1 to 2 years, when about a third of the fuel assemblies are replaced. The spent fuel assemblies are removed to cooling pools at the reactor site. Uranium Purification Natural uranium contains 0.7205% U-235, the fissile isotope of uranium. The remaining mass includes 99.274% U-238 and a small amount of U-234 (0.0055%). Uranium-238 does not contribute to slow neutron fission; however, it can react with neutrons to form a fissile isotope of plutonium, Pu-239. Thus U-238 is known as a fertile material, i.e., one that can produce fissile materials. Although U-235 and U238 are chemically identical, they differ slightly in their physical properties, most importantly mass. This small mass difference allows the isotopes to be separated and makes it possible to increase (enrich) the percentage of U-235 in uranium. Most civilian power reactors use enriched uranium fuel containing 0.8 to 8.0% U-235, known as low enriched uranium (LEU). Weapons grade uranium must contain highly enriched uranium (HEU) with an isotopic concentration greater than 90% U-235. In producing U-235 for the first atomic bomb, Manhattan Project scientists considered four physical processes for uranium enrichment: gaseous diffusion (effusion), electromagnetic separation, liquid thermal diffusion, and centrifugation. During the project the first three were employed at Oak Ridge to produce enriched uranium for the bomb used at Hiroshima. Centrifugation was abandoned because the technology and materials required to spin corrosive uranium hexafluoride with a rotator at high speeds were not practical for industrial, large-scale separations. However, advances in technology and materials make centrifugation the preferred method of enrichment today. Zirconium - Nuclear applications Consuming about 1% of the Zr supply, zirconium is used for cladding nuclear reactor fuels. [16] For this purpose, it is mainly used in the form of zircaloys. The benefits of Zr alloys is their low neutroncapture cross-section and good resistance to corrosion under normal service conditions.[5][6] The development of efficient methods for the separation of zirconium from hafnium was required for this application. One disadvantage of zirconium alloys is their reactivity toward water at high temperatures leading to the formation of hydrogen gas and to the accelerated degradation of the fuel rod cladding: Zr + 2 H2O ZrO2 + 2 H2 This exothermic reaction is very slow below 100 C, but at temperature above 900 C the reaction is rapid. Most metals undergo similar reactions. The redox reaction is relevant to the instability offuel assemblies at high temperatures,This reaction was responsible for a small hydrogen explosion first observed inside the reactor building of Three Mile Island accidented nuclear power plant in 1979, but then, the containment building was not damaged. The same reaction occurred in the reactors 1, 2 and 3 of the Fukushima I Nuclear Power Plant (Japan) after the reactors cooling was interrupted by the earthquake and tsunami disaster of March 11, 2011 leading to the Fukushima I nuclear accidents. After venting of hydrogen in the maintenance hall of these three reactors, the explosive mixture of hydrogen with air oxygen detonated, severely damaging the installations and at least one of the containment buildings. To avoid explosion, the direct venting of hydrogen to the open atmosphere would have been a preferred design option. Now, to prevent the risk of explosion in many pressurized water reactor (PWR) containment buildings, a catalyst-based recombinator is installed to rapidly convert hydrogen and oxygen into water at room temperature before explosivity limit is reached. Thorium as a nuclear fuel Benefits and challenges The naturally occurring isotope thorium-232 is a fertile material, and with a suitable neutron source can be used as nuclear fuel in nuclear reactors, including breeder reactors. In 1997, the U.S. Energy Department underwrote research into thorium fuel, and research was also begun in 1996 by the International Atomic Energy Agency (IAEA), to study the use of thorium reactors. Nuclear scientist Alvin Radkowsky of Tel Aviv University in Israel founded a consortium to develop thorium reactors, which included other companies: Raytheon Nuclear Inc., Brookhaven National Laboratory, and the Kurchatov Institute in Moscow. Radkowsky was chief scientist in the U.S. nuclear submarine program directed by Admiral Hyman Rickover and later headed the design team which built the USA's first civilian nuclear power plant at Shippingport, Pennsylvania, which was a scaled-up version of the first naval reactor. The third Shippingport core, initiated in 1977, bred Thorium. [21] Even earlier examples of reactors using fuel with thorium exist, including the first core at the Indian Point Energy Center in 1962. Some countries, including India, are now investing in research to build thorium-based nuclear reactors. A 2005 report by the International Atomic Energy Agency discusses potential benefits along with the challenges of thorium reactors.[23] India has also made thorium-based nuclear reactors a priority with its focus on developing fast breeder technology. Some benefits of thorium fuel when compared with uranium were summarized as follows Weapons-grade fissionable material (233U) is harder to retrieve safely and clandestinely from a thorium reactor;

Thorium produces 10 to 10,000 times less long-lived radioactive waste; The fissionable thorium cycle uses 100% of the isotope as coming out of the ground, which does not require enrichment, whereas the fissile uranium cycle depends on only the 0.7% fissile U-235 of the natural uranium. The same cycle could also use the fissionable U-238 component of the natural uranium, and also contained in the depleted reactor fuel; Thorium cannot sustain a nuclear chain reaction without priming so fission stops by default. However, when used in a breeder-like reactor, unlike uranium-based breeder reactors, thorium requires irradiation and reprocessing before the above-noted advantages of thorium-232 can be realized, which makes thorium fuels initially more expensive than uranium fuels.But experts note that "the second thorium reactor may activate a third thorium reactor. This could continue in a chain of reactors for a millennium if we so choose." They add that because of thorium's abundance, it will not be exhausted in 1,000 years. The Thorium Energy Alliance (TEA), an educational advocacy organization, emphasizes that "there is enough thorium in the United States alone to power the country at its current energy level for over 1,000 years." Thorium energy fuel cycle Like 238U, 232Th is not fissile itself, but it is fertile: it will absorb slow neutrons to produce, after two beta decays, 233U, which is fissile.[17] Also, preparation of thorium fuel does not requireisotopic separation. The thorium fuel cycle creates 233U, which, if separated from the reactor's fuel, can be used for making nuclear weapons. This is why a liquid-fuel cycle (e.g., Molten Salt Reactor or MSR) is preferred only a limited amount of 233U ever exists in the reactor and its heat-transfer systems, preventing any access to weapons material; however the neutrons produced by the reactor can be absorbed by a thorium or uranium blanket and fissile 233U or 239Pu produced. Also, the 233U could be continuously extracted from the molten fuel as the reactor is running. Neutrons from the decay of uranium-233 can be fed back into the fuel cycle to start the cycle again. The neutron flux from spontaneous fission of 233U is negligible. 233U can thus be used easily in a simple gun-type nuclear bomb design.In 1977; a light-water reactor at the Shippingport Atomic Power Station was used to establish a 232Th-233U fuel cycle. The reactor worked until its decommissioning in 1982.[31][32][33] Thorium can be and has been used to power nuclear energy plants using both the modified traditional Generation III reactor design and prototype Generation IV reactor designs. The use of thorium as an alternative fuel is one innovation being explored by the International Project on Innovative Nuclear Reactors and Fuel Cycles (INPRO),conducted by the International Atomic Energy Agency (IAEA). Unlike its use in MSRs, when using solid thorium in modified light water reactor (LWR) problems include: the undeveloped technology for fuel fabrication; in traditional, once-through LWR designs potential problems in recycling thorium due to highly radioactive 228Th; some weapons proliferation risk due to production of 233U; and the technical problems (not yet satisfactorily solved) in reprocessing. Much development work is still required before the thorium fuel cycle can be commercialized for use in LWR. The effort required has not seemed worth it while abundant uranium is available, but geopolitical forces (e.g. India looking for indigenous fuel) as well as uranium production issues, proliferation concerns, and concerns about the disposal/storage of radioactive waste are starting to work in its favor. Beryllium - Nuclear applications Thin plates or foils of beryllium are sometimes used in nuclear weapon designs as the very outer layer of the plutonium pits in the primary stages of thermonuclear bombs, placed to surround thefissile material. These layers of beryllium are good "pushers" for the implosion of the plutonium-239, and they are also good neutron reflectors, just as they are in beryllium-moderated nuclear reactors.[62] Beryllium is also commonly used as a neutron source in laboratory experiments in which relatively few neutrons are needed (rather than having to use a nuclear reactor). For this purpose, a target of beryllium9 is bombarded with energetic alpha particles from a radio-isotope such as polonium-210, radium226, plutonium-239, or americium-241. In the nuclear reaction that occurs, a beryllium nucleus is transmuted into carbon-12, and one free neutron is emitted, traveling in about the same direction as the alpha particle was heading. Such neutron sources, named "urchin" neutron initiators, were used some in early atomic bombs.[62] Beryllium is also used at the Joint European Torus nuclear-fusion research laboratory, and it will be used in the more advanced ITER to condition the components which face the plasma. [63]Beryllium has also been proposed as a cladding material for nuclear fuel rods, owing to its good combination of mechanical, chemical and nuclear properties.[5] Beryllium fluoride is one of the constituent salts of the eutectic salt mixture FLiBe, which is used as a solvent, moderator and coolant in many hypothetical molten salt reactor designs.[64]

Unit III REPROCESSING Spent Fuel By far the most problematic of all radioactive wastes is that of spent fuel. During the three to four years a uranium fuel rod spends in a power reactor, much of the 235U and a small amount of the 238U are converted into fission products and transuranic isotopes. Of the hundreds of different radionuclides produced as fission products, only seven have half-lives greater than 25 years: 90Sr (29.1 y), 137Cs (30.2 y), "Te (0.21 My), 79Se (1.1 My), 93Zr (1.5 My), 135Cs (2.3 My), and 129I (16 My). The latter five, with such long half-lives, are effectively stable, and thus the long-term activity of fission-product waste is determined solely by 137Cs and 90Sr (in secular equilibrium with its 54-hour half-life daughter 90Y). After 1000 years, the fission product activity will have decreased by a factor of exp[(1000 y In2)/30 y] ~ 10~10, an activity less than the ore from which the uranium was extracted However, some transuranic isotopes in the spent fuel have much longer half-lives than 90Sr and 137Cs, notably 239Pu with a 24,000 y half-life. It is these transuranic actinides that pose the greatest challenge for permanent disposal of spent fuel, requiring isolation of this HWL from the biosphere for several hundred thousand years. Spent fuel reprocessing, as currently practiced by some countries such as France and England, allows fissile isotopes to be recovered and used in new fuel. In addition, this reprocessing of spent fuel allows the fission products to be separated from the transuranic radionuclides. As discussed above, storage of the fission products requires isolation for only about a thousand years; the transuranic nuclides can be recycled into new reactor fuel and transmuted or fissioned into radionuclides with much shorter half-lives. Although spent-fuel reprocessing can greatly reduce the length of time the radioactive waste must be safely stored, it poses nuclear weapon proliferation problems since one of the products of reprocessing is plutonium from which nuclear weapons can be fabricated. By keeping the plutonium in the highly radioactive spent fuel, it is less likely that it will be diverted and used for weapons. For this reason the U.S. currently elects not to reprocess its spent fuel. In addition to fission products and transuranic nuclides. spent fuel accumulates radioactive daughters of the almost stable 238U and 235U. which were also present in the original uranium ore.

Solvent Extraction Liquidliquid solvent extraction technology is used in numerous industrial processes such as petroleum and petrochemical processing, organic chemical and pharmaceutical production, food and fuel industries, effluent treatment, and hydrometallurgy.Although solvent extraction technology had its beginning more than 100 years ago, it owes a significant amount of its present prominence as a separation technique to its successful applications in the field of nuclear technology, when demands for simple separation processes for many elements, previously considered only laboratory curiosities, arose. The use of thorium, uranium, plutonium and other actinides in the nuclear energy programs and interest in metals such as hafniumfree zirconium as a reactor construction material and bismuth, gallium and indium as heat transfer media have led to the rapid development of many solvent extraction processes that display selectivity, simplicity and speed. This paper briefly reviews the various liquidliquid solvent extraction reagents used in several nuclear technological areas, such as purification of reactor materials, radionuclide production, and the recovery and purification of thorium, uranium, and plutonium. New reagents, especially bifunctional organophosphorus compounds for actinide removal from acidic nuclear waste solutions will be discussed with a call for better, more selective, and less expensive reagents for use in the field. Reactor materials Materials employed in reactor technology along with some of the successful solvent extraction reagents used for their separation are summarized in Table 1. Of the processes connected with the elements listed none have more challenged the nuclear chemist and engineer than those of zirconium and hafnium separation and production. Each element is always associated with the other and their chemistry is very similar. Among the extractants used for the successful separation of the two elements are tributyl phosphate (ref. 10,11), thenoyltrifluoroacetone and trinoctylamine ,while the addition of thiocyanate to the aqueous phase extends the range of applicable solvents to methylisobutyl ketone ,diethyl ether, nbutanol, cyclohexanone, nbutyl acetate, pentyl acetate, methyl propyl ketone, and ethylisopropyl

ketone. The solvents commonly used in industrial processes involve methylisobutyl ketone and tributyl phosphate. They have the advantage of high separation Tactors. Uranium recovery from ores and other sources The major stages in the conventional processes for the milling of uranium ores include leaching of uranium by acid or alkaline solutions, concentration and purification by solvent extraction or ion exchange, and precipitation. Solvent extraction has displaced many of the initial ion exchange processes for the concentration and purification of uranium from acid leaching processes. While it is primarily used with clarified acid solutions, considerable development work has been done on alkaline solutions and on solventinpulp applications. The two types of solvent employed are tertiary amines (Alamine 336) and alkylphosphoric acids (di(2ethylhexyl) phosphoric acid). Generally, amine solvents have superseded the alkylphosphoric acids because of their selectivity, although the latter are preferred when vanadium is present in the feed stream. The fertilizer industry produces large quantities of wet process phosphoric acid solution containing 0.10.2 g/l of uranium, which represent a large potential source of uranium. A twocycle process was developed for the recovery of uranium from these phosphoric acid solutions by extraction with di(2ethylhexyl)phosphoric acid plus trioctylphosphine oxide in an aliphatic diluent .Later studies simplified the solution chemistry of the process by replacing the first extraction cycle with the commercial solvent mixture, mono and dioctylphenylphosphoric acid (octylphenylacidphosphate) Solvent extraction technology plays a vital role in actinide recovery and purification, radionuclide production, the preparation of reactor naterials, and other aspects of the nuclear field. There are challenging research areas for the separations chenist and synthetic organic chemist to wotk together to improve solvent extraction processes and develop energy saving process technology. One major area alluded to in this paper is through the preparation and testing of new reagents which have a far greater selectivity for the solute species, better overall recovery, improved physical properties, less toxicity, and lower costs. Solvent extraction equipment Liquid-liquid extraction (also called solvent extraction) was initially utilized in the petroleum industry beginning in the 1930s. It has since been utilized in numerous applications including petroleum, hydrometallurgical, pharmaceutical, and nuclear industries. Liquid-liquid extraction describes a method for separating components of a solution by utilizing an unequal distribution of the components between two immiscible liquid phases. In most cases, this process is carried out by intimately mixing the two immiscible phases, allowing for the selective transfer of solute(s) from one phase to the other, then allowing the two phases to separate. Typically, one phase will be an aqueous solution, usually containing the components to be separated, and the other phase will be an organic solvent, which has a high affinity for some specific components of the solution. The process is reversible by contacting the solvent loaded with solute(s) with another immiscible phase that has a higher affinity for the solute than the organic phase. The transfer of solute from one phase into the solvent phase is referred to as extraction and the transfer of the solute from the solvent back to the second (aqueous) phase is referred to as backextraction or stripping. The two immiscible fluids must be capable of rapidly separating after being mixed together, and this is primarily a function of the difference in densities between the two phases. While limited mass transfer can be completed in a single, batch equilibrium contact of the two phases, one of the primary advantages of liquid-liquid extraction processes is the ability to operate in a continuous, multistage countercurrent mode. This allows for very high separation factors while operating at high processing rates. Countercurrent operation is achieved by repeating single-stage contacts, with the aqueous and organic streams moving in opposite directions as shown in Figure

Mixer-Settlers This device consists of a small mixing chamber followed by a larger gravity settling chamber as shown in Figure.

Each mixer-settler unit provides a single stage of extraction. The two phases enter the mixing section where they are mixed using an impeller. The two-phase solution flows into the settling section where they are allowed to separate by gravity due to their density differences. Typical mixer settlers have mixing times on the order of a few minutes and settling times of several minutes. The separate phases exit the settling section by flowing over a weir (organic solution) or through an underflow then over a weir (aqueous phase). The separation interface is controlled by the height of the weirs on the outlets of the settler section. Only minimal instrumentation is required and mechanical maintenance is limited to occasional mixing motor replacement. In a countercurrent process, multiple mixer settlers are installed with mixing and settling chambers located at alternating ends for each stage (since the outlet of the settling sections feed the inlets of the adjacent stages mixing sections). Mixer-settlers are used when a process requires longer residence times and when the solutions are easily separated by gravity. They require a large facility footprint, but do not require much headspace, and need limited remote maintenance capability for occasional replacement of mixing motors.

Unit IV NUCLEAR REACTOR Nuclear reactors produce energy through a controlled fission chain reaction (see The First Chain Reaction). While most reactors generate electric power, some can also produce plutonium for weapons and reactor fuel. Power reactors use the heat from fission to produce steam, which turns turbines to generate electricity. In this respect they are similar to plants fueled by coal and natural gas. The components common to all nuclear reactors include a fuel assembly, control rods, a coolant, a pressure vessel, a containment structure, and an external cooling facility. The speed of the neutrons in the chain reaction determines the reactor type (Figure 1). Thermal reactors use slow neutrons to maintain the reaction. These reactors require a moderator to reduce the speed of neutrons produced by fission. Fast neutron reactors, also known as fast breeder reactors (FBR), use high speed, unmoderated neutrons to sustain the chain reaction. Types of Nuclear Reactors

Thermal reactors operate on the principle that uranium-235 undergoes fission more readily with slow neutrons than with fast ones. Light water (H2O), heavy water (D2O), and carbon in the form of graphite are the most common moderators. Since slow neutron reactors are highly efficient in producing fission in uranium-235, they use fuel assemblies containing either natural uranium (0.7% U-235) or slightly enriched uranium (0.9 to 2.0% U-235) fuel. Rods composed of neutron-absorbing material such as cadmium or boron are inserted into the fuel assembly. The position of these control rods in the reactor core determines the rate of the fission chain reaction. The coolant is a liquid or gas that removes the heat from the core and produces steam to drive the turbines. In reactors using either light water or heavy water, the coolant also serves as themoderator. Reactors employing gaseous coolants (CO2 or He) use graphite as the moderator. The pressure vessel, made of heavy-duty steel, holds the reactor core containing the fuel assembly, control rods, moderator, and coolant. The containment structure, composed of thick concrete and steel, inhibits the release of radiation in case of an accident and also secures components of the reactor from potential intruders. Finally, the most obvious components of many nuclear power plants are the cooling towers, the external components, which provide cool water for condensing the steam to water for recycling into the containment structure. Cooling towers are also employed with coal and natural gas plants. Reactor Fundamentals It is important to realize that while the U-235 in the fuel assembly of a thermal reactor is undergoing fission, some of the fertile U-238 present in the assembly is also absorbing neutrons to produce fissile Pu-239. Approximately one third of the energy produced by a thermal power reactor comes from fission of this plutonium. Power reactors and those used to produce plutonium for weapons operate in different ways to achieve their goals. Production reactors produce less energy and thus consume less fuel than power reactors. The removal of fuel assemblies from a production reactor is timed to maximize the amount of plutonium in the spent fuel (Figure 2). Fuel rods are removed from production reactors after only several months in order to recover the maximum amount of plutonium-239. The fuel assemblies remain in the core of a power reactors for up to three years to maximize the energy produced. However it is possible to recover some plutonium from the spent fuel assemblies of a power reactor. Thermal Reactors Currently the majority of nuclear power plants in the world are water-moderated, thermal reactors. They are categorized as either light water or heavy water reactors. Light water reactors use purified natural water (H2O) as the coolant/moderator, while heavy water reactors employ heavy water, deuterium oxide (D2O). In light water reactors, the water is either pressured to keep it in superheated form (in a

pressurized water reactor, PWR) or allowed to vaporize, forming a mixture of water and steam (in a boiling water reactor, BWR). In a PWR (Figure 3), superheated water flowing through tubes in the reactor core transfers the heat generated by fission to a heat exchanger, which produces steam in a secondary loop to generate electricity. None of the water flowing through the reactor core leaves the containment structure. In a BWR (Figure 4), the water flowing through the core is converted to directly to steam and leaves the containment structure to drive the turbines. Light water reactors use low enriched uranium as fuel. Enriched fuel is required because natural water absorbs some of the neutrons, reducing the number of nuclear fissions. All of the 103 nuclear power plants in the United States are light water reactors; 69 are PWRs and 34 are BWRs. Pressurized Water Reactor

Boiling Water Reactor

(1) the core inside the reactor vessel creates heat, (2) a steam-water mixture is produced when very pure water (reactor coolant) moves upward through the core, absorbing heat, (3) the steam-water mixture leaves the top of the core and enters the two stages of moisture separation where water droplets are removed before the steam is allowed to enter the steam line, and (4) the steam line directs the steam to the main turbine, causing it to turn the turbine generator, which produces electricity.

Heavy water reactors use D2O as the coolant/moderator, allowing natural, unenriched uranium to be used as the fuel. This is possible because D2O absorbs fewer neutrons than H2O. The heat transfer system is similar to that of the PWR, with the steam generator located within the containment structure. The heavy water reactor shown in Figure 5, known as CANDU, was developed in Canada and sold globally. The cost advantage in fuel is offset by the expense of producing D2O by a chemical exchange process or electrolysis. The individual fuel assemblies of a heavy water reactor can be replaced without shutting down the reactor, thus eliminating the down time involved with refueling a light water reactor. However, spent fuel produced by a heavy water reactor contains more plutonium and tritium than that from light water reactors. This, coupled with the difficulty in monitoring a continuously fueled reactor, causes concerns about the proliferation of nuclear weapons. Thus, heavy water is classified as a "sensitive material" because a nation possessing it can produce plutonium directly from natural uranium, eliminating the need for uranium enrichment. Heavy Water Reactor

Another type of thermal reactor is graphite moderated and gas cooled. Twenty-six Magnox reactors (Figure 6), employing pressurized carbon dioxide as the coolant, were built in the United Kingdom but are currently being phased out of service. These gas cooled reactors have the same advantages as the heavy water reactors in that they can use natural uranium fuel and be fueled continuously. Gas Cooled Reactor

The pebble bed represents a new design for a gas cooled reactor (Figure 7). It uses helium as the coolant and a fuel consisting of low enriched uranium dioxide coated with silicon carbide and pyrolitic carbon encased in small graphite spheres. The graphite serves as the moderator. Fresh fuel is added at the top of the reactor and the spent fuel removed from the bottom, allowing continuous operation. The helium

coolant is maintained at high temperatures and pressures, which increases the efficiency of heat transfer and power production. South Africa is leading the effort to develop this reactor technology. Pebble Bed Reactor

Fast Neutron Reactors In contrast to thermal reactors, the neutrons in a fast neutron reactor (or fast breeder reactor, FBR) are not slowed by the presence of a moderator (Figure 8). The coolant, usually a liquid sodium or lead, is a substance that does not slow or absorb neutrons. It also has excellent heat transfer properties, which allow the reactor to be operated at lower pressures and higher temperatures than thermal reactors. Fast Neutron Breeder Reactor

An FBR is configured and operated to produce more fuel than it consumes. Fast neutrons are readily absorbed by fertile uranium-238, which then can undergo successive beta emissions to become fissile Pu239. Thorium-232 is another fertile isotope that can absorb neutrons and produce fissile uranium-233 by beta emissions. These fissile isotopes can be reprocessed for nuclear reactor fuel or weapons. Because

fast neutrons are not as efficient in producing fission as slow ones, FBRs require uranium oxide containing 20% U-235, plutonium oxide, or a mixture of these oxides, known as MOX, as fuel. Originally FBRs were thought to be a means of extending global uranium resources by producing fissile Pu-239 or U-233 as reactor fuel. However, problems with reactor operations and material components combined with the discovery of new uranium deposits mean that FRBs are not economically competitive with existing thermal reactors. FBR research has produced technical advances but the limiting factor continues to be the price of FBR-produced reactor fuel versus the cost of uranium fuel. FBRs are more complex than other types of reactors and also raise concerns about the proliferation of plutonium for use in nuclear weapons. Summary of Reactor Types Table 1 summarizes the characteristics of the reactors discussed above. Table 1 Reactor Characteristics Reactor Type Function Coolant Moderator Chemical Form Fuel of Fuel Enrichment Level* Thermal Boiling Water electricity light water light water uranium dioxide low enriched uranium Pressurized Water electricity, light water light water uranium dioxide low enriched nautical power uranium Heavy Water electricity, heavy water heavy water uranium dioxide or natural, unenriched plutonium uranium metal uranium production Gas Cooled electricity, carbon graphite uranium dicarbide slightly enriched or Graphite plutonium dioxide or or uranium metal natural uranium Moderated production helium Water Cooled electricity, light water graphite uranium dicarbide slightly enriched Graphite plutonium or uranium metal uranium Moderated production Pebble Bed Gas electricity pressurized graphite and uranium dioxide or low enriched Cooled Graphite helium silicon carbide thorium dioxide uranium Moderated** Fast Neutron Fast Neutron electricity, molten none required various mixtures various mixtures of Breeder plutonium sodium or of plutonium plutonium dioxide production lead dioxide and and uranium dioxide uranium dioxide *Percentage of U-235 isotope in the fuel compared to U-238 isotope. Natural uranium contains 0.7% U235, slightly enriched uranium from 0.8 to 3.0% U-235, and low enriched uranium from 3.0 to 5.0% U235. HEAT GENERATION IN REACTORS The starting point in the design of a reactor cooling system is the determination of the spatial distribution of the heat produced within the reactor. the energy released in fission appears in several fOnTIs-as kinetic energy of the fission neutrons, as prompt fission y-rays, as y-rays and f3-rays from the decay of fission products, and in the emission of neutrinos. With the exception of the neutrinos, virtually all of this energy is ultimately absorbed somewhere in the reactor. However, because these various radiations are attenuated in different ways by matter, their energy tends to be deposited in different locations. In the following discussion of the deposition of energy, it is assumed that the reactor fuel is in the form of individual fuel rods. This is the case for virtually all power reactor concepts except the molten salt breeder reactor Heat Production in Fuel Elements (Fuel Rods) the fission fragments have a kinetic energy of about 1 68 Me V per fission. These highly charged particles have an extremely short range ,and therefore their energy is deposited near the site of the fission within the fuel. Similarly, most of the 8 MeV of the fission product fJ-rays is also deposited in the fuel. However, many of the y-rays from the decaying fission products and those emitted directly in fission pass out of the fuel since they are less strongly attenuated than charged particles. Some of these y-rays are absorbed in the surrounding coolant and/or moderator, in the thermal shield, or in the radiation shielding that surrounds the reactor. However, because of the proximity of fuel rods in most reactors, many of the y rays are intercepted by and absorbed in neighboring rods. The prompt neutrons are emitted with a total kinetic energy of about 5 Me V per fission. In a thermal reactor, the bulk of this energy is deposited in the moderator as the neutrons slow down. The capture yrays emitted following the absorption of these neutrons in nonfission reactions are therefore produced and absorbed throughout the reactor. In a fast reactor, the fission neutrons slow down very little before they are absorbed, and their kinetic energy appears as an addition to the energy of the capture y-rays. The

delayed neutrons contribute negligibly to the energy of a reactor. As already noted, none of the energy of the neutrinos is retained within a reactor. It should be clear from the foregoing remarks that the spatial deposition of fission energy depends on the details of the reactor's structure. Nevertheless, for preliminary calculations, it may be assumed that approximately one-third of the total y -ray energy-about 5 MeV-is absorbed in the fuel. This, together with the 1 68 MeV from the fission fragments and 8 MeV from the fJ-rays, gives 1 8 1 MeV per fission (about 90% of the recoverable fission energy), which is deposited in the fuel, most in the immediate vicinity of the fission site. The remainder (about 20 Me V) of the recoverable energy is deposited in the coolant and/or moderator, in various structural materials, and in the blanket, reflector, and shield. Radiation Heating As noted already, roughly 1 0% of the recoverable energy of fission is absorbed outside the fuel. In thermal reactors, the kinetic energy of the fission neutrons is deposited in the surrounding moderator and coolant in more or less the same spatial distribution as the fissions from which these neutrons originate. However, only 2% to 3 % of the fission energy appears in this form, and it is often assumed that this kinetic energy is deposited uniformly throughout the core. The calculation of the energy deposition from the longer Fission Product Decay Heating After a few days of reactor operation, the f3- and y -radiation emitted from decaying fission products amounts to about 7% of the total thermal power output of the reactor. When the reactor is shut down, the accumulated fission products continue to decay and release energy within the reactor. This fission product decay energy can be quite sizable in absolute terms, and a means for cooling the reactor core after shutdown must be provided in all reactors except those operating at very low power levels. If this is not done, the temperature of the fuel may rise to a point where the integrity of the fuel is compromised and fission products are released. Consider a reactor that has been operating at a constant thermal power Po long enough for the concentrations of the radioactive fission products to come to equilibrium. Since the rate of production of the fission products is proportional to the reactor power, it follows that the activity of the fission products at any time after the reactor has been shut down is also proportional to Po. The ratio P (ts) / Po, where P (ts ) is the power (rate of energy release) emanating from the fission products at the time ts after shutdown, is therefore independent of Po. HEAT FLOW BY CONDUCTION Energy is removed from a reactor by two fundamentally different heat transfer processes-conduction and convection. In conduction, heat is transmitted from one location in a body to another as a result of a temperature difference existing in the body-there is no macroscopic movement of any portion of the body. It is by this mechanism, as shown in this section, that heat produced in a fuel rod is transferred to the surface of the rod. Heat convection involves the transfer of heat to a moving liquid or gas, again as the result of a temperature difference and the later rejection of this heat at another location. Thus, the heat conducted to the surface of a fuel rod is carried into the coolant and out of the system by convection. BOILING HEAT TRANSFER Up to this point, it has been assumed that a liquid coolant does not undergo a change in phase as it moves along a coolant channel absorbing heat from the fuel. It was assumed that the coolant does not boil. However, there are some distinct advantages in permitting a reactor coolant to boil. For one thing, the coolant pressure is much lower when the coolant is allowed to boil than when boiling must be prevented. In addition, for a given flow rate and heat flux, lower cladding and fuel temperature are required for a boiling than for a nonboiling coolant. For these reasons, boiling of a restricted nature is now pennitted in pressurized water reactors, although steam is not produced directly in these reactors. With boiling water reactors not only is advantage taken of higher heat transfer rates, by producing steam within these reactors, the entire secondary coolant loop of the PWR can be eliminated. Boiling coolants other than ordinary water have been considered in a number of reactor concepts, but none of these has reached a practical stage of development. Therefore, the following discussion pertains largely to watercooled reactors. However, the principles to be considered also apply to other types of liquid-cooled reactors. Nuclear Shelding Since a nuclear reactor is "highly radioactive" especially when operating at high power levels, producing a lot of penetrating gamma rays (like high energy x-rays) and neutron radiation, people could not work in the area unless there was some way of mitigating/reducing the radiation levels. Thus, the designers put in some "shielding" materials of some kind, depending on the design of the reactor and the power levels. Most commercial nuclear plants can rely on using concrete and distance (the farther away you are, the less damaging the radiation) to reduce the radiation levels where people work. Another good shielding material (at least for gamma radiation) is lead. It is a bit more expensive, so it tends to be relied on more for places where there is less space, such as on ships that have nuclear power plants. Water is also a good shield for neutrons since the neutrons bounce off the hydrogen and give up a lot of their energy with each collision.

A couple of feet of water or concrete are considered enough to reduce the neutron radiation by a factor of ten. Two inches of lead are viewed as being able to reduce gamma radiation by a factor of ten. The (cylindrical) shield/shielding is placed around the outside of the reactor pressure vessel, as close as possible to minimize the amount of material needed to get the radiation levels down. Just for information, many nuclear reactors also have what is called a "thermal shield" to protect the reactor pressure vessel from temperature profiles and excessive embrittlement induced by the radiation streaming out of the reactor core. That is part of the reactor system design, shielding the pressure vessel itself from the intensive radiation levels. Fusion Reactors Reactors for nuclear fusion are of two main varieties, magnetic confinementreactors and inertial confinement reactors. The strategies for creating fusion reactors are largely dictated by the fact that the temperatures involved in nuclear fusion are far too high to be contained in any material container. The strategy of the magnetic confinement reactor is to confine the hot plasma by means of magnetic fields which keep it perpetually in looping paths which do not touch the wall of the container. This is typified by the tokamak design, the most famous example of which is the TFTR at Princeton. The strategy of the inertial confinement reactor is to put such high energy density into a small pellet of deuterium-tritium that it fuses in such a short time that it can't move appreciably. The most advanced test reactors involve laser fusion, particularly in the Shiva and Nova reactors at Lawrence Livermore Laboratories Fusion Reactors: Magnetic Confinement There are two ways to achieve the temperatures and pressures necessary for hydrogen fusion to take place: Magnetic confinement uses magnetic and electric fields to heat and squeeze the hydrogen plasma. The ITER project in France is using this method. Inertial confinement uses laser beams or ion beams to squeeze and heat the hydrogen plasma. Scientists are studying this experimental approach at theNational Ignition Facility of Lawrence Livermore Laboratory in the United States. Let's look at magnetic confinement first. Here's how it would work: Microwaves, electricity and neutral particle beams from accelerators heat a stream of hydrogen gas. This heating turns the gas into plasma. This plasma gets squeezed by super-conducting magnets, thereby allowing fusion to occur. The most efficient shape for the magnetically confined plasma is a donut shape (toroid). A reactor of this shape is called a tokamak. The ITER tokamak will be a self-contained reactor whose parts are in various cassettes. These cassettes can be easily inserted and removed without having to tear down the entire reactor for maintenance. The tokamak will have a plasma toroid with a 2-meter inner radius and a 6.2-meter outer radius.

Unit V SAFETY AND DISPOSAL Nuclear safety covers the actions taken to prevent nuclear and radiation accidents or to limit their consequences. This covers nuclear power plants as well as all other nuclear facilities, the transportation of nuclear materials, and the use and storage of nuclear materials for medical, power, industry, and military uses. The nuclear power industry has improved the safety and performance of reactors, and has proposed new safer (but generally untested) reactor designs but there is no guarantee that the reactors will be designed, built and operated correctly.Mistakes do occur and the designers of reactors at Fukushima in Japan did not anticipate that a tsunami generated by an earthquake would disable the backup systems that were supposed to stabilize the reactor after the earthquake.According to UBS AG, the Fukushima I nuclear accidents have cast doubt on whether even an advanced economy like Japan can master nuclear safety.Catastrophic scenarios involving terrorist attacks are also conceivable. An interdisciplinary team from MIT have estimated that given the expected growth of nuclear power from 2005 2055, at least four serious nuclear accidents would be expected in that period.To date, there have been five serious accidents (core damage) in the world since 1970 (one at Three Mile Island in 1979; one at Chernobyl in 1986; and three at Fukushima-Daiichi in 2011), corresponding to the beginning of the operation of generation II reactors. This leads to on average one serious accident happening every eight years worldwide Reactor protection system (RPS) A reactor protection system is composed of systems that are designed to immediately terminate the nuclear reaction. While the reactor is operating, the nuclear reaction continues to produce heat and radiation. By breaking the chain reaction, the source of heat can be eliminated, and other systems can then be used to continue to remove decay heat from the core. All plants have some form of the following reactor protection systems: Control rods Control rods are a series of metal rods that can be quickly inserted into the core to absorb neutrons and rapidly terminate the nuclear reaction. See control rods for more information. Safety injection / standby liquid control A nuclear reaction can also be stopped by injecting a liquid that absorbs neutrons directly into the core. In boiling water reactors this usually consists of a solution containing boron (such as boric acid), which can be injected to displace the water in the core. A signature of pressurized water reactors is that they use a boron solution in addition to control rods to control the reaction, and so the concentration is simply increased to slow or stop the reaction.

Cooling tower at the Philippsburg Nuclear Power Plant, Germany The essential service water system (ESWS) circulates the water that cools the plants heat exchangers and other components before dissipating the heat into the environment. Because this includes cooling the systems that remove decay heat from both the primary system and the spent fuel rodcooling ponds, the ESWS is a safety-critical system.[2] Since the water is frequently drawn from an adjacent river, the sea, or other large body of water, the system can be endangered by large volumes of seaweed, marine organisms, oil pollution, ice and debris.[2][3] In locations without a large body of water in which to dissipate the heat, water is recirculated via a cooling tower. The failure of half of the ESWS pumps was one of the factors that endangered safety in the 1999 Blayais Nuclear Power Plant flood,[4][5] while a total loss occurred during the Fukushima I and Fukushima II nuclear accidents in 2011.[6][5]

HPCI and LPCI as a part of active ECCS

An emergency core cooling system comprises a series of systems that are designed to safely shut down a nuclear reactor during accident conditions. Under normal conditions, heat is removed from a nuclear reactor by condensing steam after it passes through the turbine. In a boiling water reactor, condensed steam (water) is fed back into the reactor. In a pressurized water reactor, it is fed back through the heat exchanger. In both cases, this keeps the reactor core at a constant temperature. During an accident, the condenser is not used, so alternate methods of cooling are required to prevent damage to the nuclear fuel. These systems allow the plant to respond to a variety of accident conditions, and additionally introduce redundancy so that the plant can be shut down even with one or more subsystem failures. In most plants, ECCS is composed of the following systems: High pressure coolant injection system (HPCI) This system consists of a pump or pumps that have sufficient pressure to inject coolant into the reactor vessel while it is pressurized. It is designed to monitor the level of coolant in the reactor vessel and automatically inject coolant when the level drops below certain setpoints. This system is normally the first line of defense for a reactor since it can be used while the reactor vessel is still highly pressurized. Depressurization system (ADS)

Passive ECCS This system consists of a series of valves which open to vent steam several feet under the surface of a large pool of liquid water (known as the wetwell or torus) in pressure suppression type containments, or directly into the primary containment structure, in other types of containments, such as large-dry, icecondenser, and sub-atmospheric containments. The actuation of these valves depressurizes the reactor vessel and allows lower pressure coolant injection systems to function, which have very large capacities in comparison to high pressure systems. Some depressurization systems are automatic in function but can be inhibited, some are manual and operators may activate if necessary. Low pressure coolant injection system (LPCI) This system consists of a pump or pumps which inject additional coolant into the reactor vessel once it has been depressurized. In some nuclear power plants, LPCI is a mode of operation of a residual heat removal system (RHR or RHS). LPCI is generally not a stand-alone system. Corespray system This system uses spargers (special spray nozzles) within the reactor pressure vessel to spray water directly onto the fuel rods, suppressing the generation of steam. Reactor designs can include corespray in high-pressure and low-pressure modes. Containment spray system This system consists of a series of pumps and spargers which spray coolant into the primary containment structure. It is designed to condense the steam into liquid water within the primary containment structure to prevent overpressure, which could lead to involuntary depressurization. Isolation cooling system This system is often driven by a steam turbine, and is used to provide enough water to safely cool the reactor if the reactor building is isolated from the control and turbine buildings. Steam turbine driven cooling pumps with pneumatic controls can run at mechanical governor controlled adjustable speeds, without battery power, emergency generator, or off-site electrical power. The Isolation cooling system is a defensive system against a condition known as station blackout. Emergency electrical systems Under normal conditions, nuclear power plants receive power from off-site. However, during an accident a plant may lose access to this power supply and thus may be required to generate its own power to supply its emergency systems. These electrical systems usually consist of diesel generators and batteries. Diesel generators

Diesel generators are employed to power the site during emergency situations. They are usually sized such that a single one can provide all the required power for a facility to shut down during an emergency. Facilities have multiple generators for redundancy. Additionally, systems which are not required to shut down the reactor have separate electrical sources (often separate generators) so that they do not affect shutdown capability. Motor generator flywheels Loss of electrical power can occur suddenly, and it can damage or undermine equipment. To prevent damage, motor-generators can be tied to flywheels which can provide uninterrupted electrical power to equipment for a brief period of time. Often they are used to provide electrical power until the plant electrical supply can be switched to the batteries and/or diesel generators. Batteries Batteries often form the final redundant backup electrical system and are also capable of providing sufficient electrical power to shut down a plant. The DC power generated by batteries can be converted to AC power to run AC devices such as motors using an electrical inverter. Containment systems Containment systems are designed to prevent the release of radioactive material into the environment. Fuel cladding The fuel cladding is the first layer of protection around the nuclear fuel and is designed to protect the fuel from corrosion that would spread fuel material throughout the reactor coolant circuit. In most reactors it takes the form of a sealed metallic or ceramic layer. It also serves to trap fission products, especially ones that are gaseous at the temperatures reached within the reactor, such as krypton, xenon and iodine. Cladding does not constitute shielding, and must be developed such that it absorbs as little radiation as possible. For this reason, materials such as magnesium and zirconium are used for their low neutron capture cross sections. Reactor vessel The reactor vessel is the first layer of shielding around the nuclear fuel and usually is designed to trap most of the radiation released during a nuclear reaction. The reactor vessel is also designed to withstand high pressures. Primary containment The primary containment system usually consists of a large metal and concrete structure (often cylindrical or bulb shaped) which contains the reactor vessel. In most reactors it also contains all of the radioactive contaminated systems. The primary containment system is designed to withstand strong internal pressures resulting from a leak or intentional depressurization of the reactor vessel. Secondary containment Some plants have a secondary containment system which encompasses the primary system. This is very common in BWRs because most of the steam systems, including the turbine, contain radioactive materials. Core catching In case of a full melt-down, the fuel would most likely end up on the concrete floor of the primary containment building. Concrete can withstand very much heat, so the thick flat concrete floor in the primary containment will often be sufficient protection against the so-called China Syndrome. The Chernobyl plant didn't have a containment building, but the core was eventually stopped by the concrete foundation. Due to concerns that the core would melt its way through the concrete, a "core catching device" was invented, and a mine was quickly dug under the plant with the intention to install such a device. The device contains a quantity of metal which would melt, diluting the corium and increasing its heat conductivity; the diluted metallic mass could then be cooled by water circulating in the floor. Today, all new Russian-designed reactors are equipped with core-catchers in the bottom of the containment building.[7] Non-containable events Nuclear events outside of the primary containment building will not be contained. Any accident involving the spent fuel pool, which is outside of the primary containment, will not be contained. Standby gas treatment A standby gas treatment (SBGT) system is part of the secondary containment system of a nuclear power plant. When called upon to operate, the SBGT system filters and pumps air from secondary containment to the environment and maintains a negative pressure within the secondary containment in order to limit the release of radioactive material. Each SBGT train generally consists of a mist eliminator/roughing filter; an electric heater; a prefilter; two absolute (HEPA) filters; an activated charcoal filter; an exhaust fan; and associated valves, ductwork, dampers, instrumentation, and controls. The signals that trip the SBGT system are plant-specific; however, automatic trips are generally associated with the electric heaters and a high temperature condition in the charcoal filters. Ventilation and radiation protection In case of a radioactive release, most plants have a system designed to remove radiation from the air to reduce the effects of the radiation release on the employees and public. This system usually consists of the following:

Containment ventilation This system is designed to remove radiation and steam from primary containment in the event that the depressurization system was used to vent steam into primary containment. Control room ventilation This system is designed to ensure that the operators who are required to operate the plant are protected in the event of a radioactive release. This system often consists of activated charcoalfilters which remove radioactive isotopes from the air. Changes and consequences of accident Accident types For a list of many of the most important accidents see the International Atomic Energy Agency site. Loss of coolant accident A loss-of-coolant accident (LOCA) is a mode of failure for a nuclear reactor; if not managed effectively, the results of a LOCA could result in reactor core damage. Each nuclear plant'semergency core cooling system (ECCS) exists specifically to deal with a LOCA. Nuclear reactors generate heat internally; to remove this heat and convert it into useful electrical power, a coolant system is used. If this coolant flow is reduced, or lost altogether, the nuclear reactor's emergency shutdown system is designed to stop the fission chain reaction. However, due to radioactive decay the nuclear fuel will continue to generate a significant amount of heat. Thedecay heat produced by a reactor shutdown from full power is initially equivalent to about 5 to 6% of the thermal rating of the reactor.[1] If all of the independent cooling trains of the ECCS fail to operate as designed, this heat can increase the fuel temperature to the point of damaging the reactor. If water is present, it may boil, bursting out of its pipes. (For this reason, nuclear power plants are equipped with pressure-operated relief valves and backup supplies of cooling water.) If graphite and air are present, the graphite may catch fire, spreading radioactive contamination. This situation exists only in AGRs, RBMKs, Magnox and weapons-production reactors, which use graphite as a neutron moderator. (see Chernobyl disaster.) The fuel and reactor internals may melt; if the melted configuration remains critical, the molten mass will continue to generate heat, possibly melting its way down through the bottom of the reactor. Such an event is called a nuclear meltdown and can have severe consequences. The so-called "China syndrome" would be this process taken to an extreme: the molten mass working its way down through the soil to the water table (and below) - however, current understanding and experience of nuclear fission reactions suggests that the molten mass would become too disrupted to carry on heat generation before descending very far; for example, in the Chernobyl disaster the reactor core melted and core material was found in the basement, too widely dispersed to carry on a chain reaction (but still dangerously radioactive). Some reactor designs have passive safety features that prevent meltdowns from occurring in these extreme circumstances. The Pebble Bed Reactor, for instance, can withstand extreme temperature transients in its fuel. Another example is the CANDU reactor, which has two large masses of relatively cool, low-pressure water (first is the heavy-water moderator; second is the light-water-filled shield tank) that act as heat sinks. Another example is the Hydrogen Moderated Self-regulating Nuclear Power Module, in which the chemical decomposition of the uranium hydride fuel halts the fission reaction by removing the hydrogen moderator.[2] The same principle is used in TRIGA research reactors. Under operating conditions, a reactor may passively (that is, in the absence of any control systems) increase or decrease its power output in the event of a LOCA or of voids appearing in its coolant system (by water boiling, for example). This is measured by the coolant void coefficient. Most modern nuclear power plants have a negative void coefficient, indicating that as water turns to steam, power instantly decreases. Two exceptions are the Russian RBMK and the Canadian CANDU. Boiling water reactors, on the other hand, are designed to have steam voids inside the reactor vessel. Criticality accidents A criticality accident (also sometimes referred to as an "excursion" or "power excursion") occurs when a nuclear chain reaction is accidentally allowed to occur in fissile material, such as enriched uranium or plutonium. The Chernobyl accident is an example of a criticality accident. This accident destroyed a reactor at the plant and left a large geographic area uninhabitable. In a smaller scale accident at Sarov a technician working with highly enriched uranium was irradiated while preparing an experiment involving a sphere of fissile material. The Sarov accident is interesting because the system remained critical for many days before it could be stopped, though safely located in a shielded experimental hall. This is an example of a limited scope accident where only a few people can be harmed, while no release of radioactivity into the environment occurred. A criticality accident with limited off site release of both radiation (gamma and neutron) and a very small release of radioactivity occurred at Tokaimura in 1999 during the production of enriched uranium fuel. Two workers died, a third was permanently injured, and 350 citizens were exposed to radiation. Decay heat Decay heat accidents are where the heat generated by the radioactive decay causes harm. In a large nuclear reactor, a loss of coolant accident can damage the core: for example, at Three Mile Island a

recently shutdown (SCRAMed) PWR reactor was left for a length of time without cooling water. As a result the nuclear fuel was damaged, and the core partially melted. The removal of the decay heat is a significant reactor safety concern, especially shortly after shutdown. Failure to remove decay heat may cause the reactor core temperature to rise to dangerous levels and has caused nuclear accidents. The heat removal is usually achieved through several redundant and diverse systems, and the heat is often dissipated to an 'ultimate heat sink' which has a large capacity and requires no active power, though this method is typically used after decay heat has reduced to a very small value. However, the main cause of release of radioactivity in the Three Mile Island accident was a pilot-operated relief valve on the primary loop which stuck in the open position. This caused the overflow tank into which it drained to rupture and release large amounts of radioactive cooling water into the containment building. In 2011, an earthquake and tsunami caused a loss of power to two plants in Fukushima, Japan, crippling the reactor as decay heat caused 90% of the fuel rods in the core of the Daiichi Unit 3 reactor to become uncovered. As of May 30, 2011, the removal of decay heat is still a cause for concern. Transport Transport accidents can cause a release of radioactivity resulting in contamination or shielding to be damaged resulting in direct irradiation. In Cochabamba a defective gamma radiography set was transported in a passenger bus as cargo. The gamma source was outside the shielding, and it irradiated some bus passengers. In the United Kingdom, it was revealed in a court case that in March 2002 a radiotherapy source was transported from Leeds to Sellafield with defective shielding. The shielding had a gap on the underside. It is thought that no human has been seriously harmed by the escaping radiation. Equipment failure Equipment failure is one possible type of accident, recently at Biaystok in Poland the electronics associated with a particle accelerator used for the treatment of cancer suffered a malfunction.[43]This then led to the overexposure of at least one patient. While the initial failure was the simple failure of a semiconductor diode, it set in motion a series of events which led to a radiation injury. A related cause of accidents is failure of control software, as in the cases involving the Therac-25 medical radiotherapy equipment: the elimination of a hardware safety interlock in a new design model exposed a previously undetected bug in the control software, which could lead to patients receiving massive overdoses under a specific set of conditions. Human error

A sketch used by doctors to determine the amount of radiation to which each person had been exposed during the Slotin excursion Many of the major nuclear accidents have been directly attributable to operator or human error. This was obviously the case in the analysis of both the Chernobyl and TMI-2 accidents. At Chernobyl, a test procedure was being conducted prior to the accident. The leaders of the test permitted operators to disable and ignore key protection circuits and warnings that would have normally shut the reactor down. At TMI-2, operators permitted thousands of gallons of water to escape from the reactor plant before observing that the coolant pumps were behaving abnormally. The coolant pumps were thus turned off to protect the pumps, which in turn led to the destruction of the reactor itself as cooling was completely lost within the core. A detailed investigation into SL-1 determined that one operator (perhaps inadvertently) manually pulled the 84-pound (38 kg) central control rod out about 26 inches rather than the maintenance procedure's intention of about 4 inches. An assessment conducted by the Commissariat lnergie Atomique (CEA) in France concluded that no amount of technical innovation can eliminate the risk of human-induced errors associated with the operation of nuclear power plants. Two types of mistakes were deemed most serious: errors committed during field operations, such as maintenance and testing, that can cause an accident; and human errors made during small accidents that cascade to complete failure. [5] In 1946 Canadian Manhattan Project physicist Louis Slotin performed a risky experiment known as "tickling the dragon's tail"which involved two hemispheres of neutron-reflective beryllium being brought together around a plutonium core to bring it to criticality. Against operating procedures, the hemispheres were separated only by a screwdriver. The screwdriver slipped and set off a chain reaction criticality accident filling the room with harmful radiation and a flash of blue light (caused by

excited, ionized air particles returning to their unexcited states). Slotin reflexively separated the hemispheres in reaction to the heat flash and blue light, preventing further irradiation of several coworkers present in the room. However Slotin absorbed a lethal dose of the radiation and died nine days afterwards. The infamous plutonium mass used in the experiment was referred to as the demon core. Lost source Lost source accidents,[46][47] also referred to as an orphan source are incidents in which a radioactive source is lost, stolen or abandoned. The source then might cause harm to humans. For example, in 1996 sources were left behind by the Soviet army in Lilo, Georgia.Another case occurred at Yanango where a radiography source was lost, also at Samut Prakarn a phosphorusteletherapy source was lost[ and at Gilan in Iran a radiography source harmed a welder. The best known example of this type of event is the Goinia accident which occurred in Brazil. The International Atomic Energy Agency has provided guides for scrap metal collectors on what a sealed source might look like. The scrap metal industry is the one where lost sources are most likely to be found. Trafficking in radioactive and nuclear materials Information reported to the International Atomic Energy Agency (IAEA) shows "a persistent problem with the illicit trafficking in nuclear and other radioactive materials, thefts, losses and other unauthorized activities".From 1993 to 2006, the IAEA confirmed 1080 illicit trafficking incidents reported by participating countries. Of the 1080 confirmed incidents, 275 incidents involved unauthorized possession and related criminal activity, 332 incidents involved theft or loss of nuclear or other radioactive materials, 398 incidents involved other unauthorized activities, and in 75 incidents the reported information was not sufficient to determine the category of incident. Several hundred additional incidents have been reported in various open sources, but are not yet confirmed Criteria for safety. One relatively prevalent notion in discussions of nuclear safety is that of safety culture. The International Nuclear Safety Advisory Group, defines the term as the personal dedication and accountability of all individuals engaged in any activity which has a bearing on the safety of nuclear power plants. [61] The goal is to design systems that use human capabilities in appropriate ways, that protect systems from human frailties, and that protect humans from hazards associated with the system. [61] At the same time, there is some evidence that operational practices are not easy to change. Operators almost never follow instructions and written procedures exactly, and the violation of rules appears to be quite rational, given the actual workload and timing constraints under which the operators must do their job. Many attempts to improve nuclear safety culture were compensated by people adapting to the change in an unpredicted way.[61] According to Areva's Southeast Asia and Oceania director, Selena Ng, Japan's Fukushima nuclear disaster is "a huge wake-up call for a nuclear industry that hasn't always been sufficiently transparent about safety issues". She said "There was a sort of complacency before Fukushima and I don't think we can afford to have that complacency now".[62] An assessment conducted by the Commissariat lnergie Atomique (CEA) in France concluded that no amount of technical innovation can eliminate the risk of human-induced errors associated with the operation of nuclear power plants. Two types of mistakes were deemed most serious: errors committed during field operations, such as maintenance and testing, that can cause an accident; and human errors made during small accidents that cascade to complete failure. [60] According to Mycle Schneider, reactor safety depends above all on a 'culture of security', including the quality of maintenance and training, the competence of the operator and the workforce, and the rigour of regulatory oversight. So a better-designed, newer reactor is not always a safer one, and older reactors are not necessarily more dangerous than newer ones. The 1978 Three Mile Island accident in the United States occurred in a reactor that had started operation only three months earlier, and the Chernobyl disaster occurred after only two years of operation. A serious loss of coolant occurred at the French Civaux-1 reactor in 1998, less than five months after start-up.[63] However safe a plant is designed to be, it is operated by humans who are prone to errors. Laurent Stricker, a nuclear engineer and chairman of the World Association of Nuclear Operators says that operators must guard against complacency and avoid overconfidence. Experts say that the "largest single internal factor determining the safety of a plant is the culture of security among regulators, operators and the workforce and creating such a culture is not easy".[63] Nuclear waste-types of waste and its disposal Like other industrial processes, generating electricity from nuclear power or making nuclear weapons creates waste. These radioactive and chemically toxic wastes result from the mining and processing of uranium as well as from storing or reprocessing spent reactor fuel. Waste from Uranium Production The tailings or waste produced by the extraction or concentration of uranium from its ore contain radioactive isotopes of uranium, thorium, and radium as well as significant concentrations of heavy metal including chromium, lead, molybdenum, and vanadium. More than 200 pounds of tailings are produced

for each pound of uranium. This sandy waste material must be contained in carefully monitored sites known as tailings piles Containment Site on the Colorado River near Moab, Utah

Waste from Conversion, Enrichment, and Fuel Fabrication Processes Uranium production processes do not affect the level of radioactivity and do not produce significant chemical waste. An enrichment process for one ton of uranium hexafluoride produces 130 kg of UF6 (3.5% U-235) and 870 kg of depleted UF6 containing U-238. Depleted uranium has few applications. However, its high density 18.7 g/cm3 makes it useful in armor plating and radiation shielding. It is also a potential energy source for fast breeder reactors. Waste from Reactors Spent fuel in the open and closed fuel cycles generates radioactive waste. The components of spent reactor fuel can either be treated as waste (in the open fuel cycle) or reprocessed (in the closed fuel cycle). In either case, spent fuel from the reactor is initially stored in cooling ponds under water at the reactor site to allow a decrease in radioactivity and a corresponding decrease in temperature. The amount of time the spent fuel remains in a pool is determined by whether it is to be kept as waste or reprocessed for either fuel or nuclear weapons. If it is to be treated as waste, it can remain in the pool indefinitely. If it is to be reprocessed to recover plutonium for weapons, it is removed after several months. Waste from Nuclear Power Generation The Open Fuel Cycle In the open fuel cycle, the spent fuel rods remain in the pools under at least 20 feet of water (Figure 2), which protects the surroundings from radiation. Storage of Spent Fuel at the Reactor

After a minimum of one year, the rods may be removed from the pool and placed in a cylinder in a chemically inert atmosphere of helium gas (Figure 3). The cylinder is then sealed and encased in steel and concrete to contain the radiation and enhance security for storage or transportation to a permanent repository. In 2008 there were 160,000 assemblies containing 45,000 tons of spent fuel from nuclear power reactors in the United States. The majority of these are stored at the reactor sites in reactor pools with only about 5 percent in dry casks. Each year about 7,800 additional used fuel assemblies are placed

into storage. If all of the current assemblies were collected in a single location they would cover a football field to about a height of five and half yards. Dry Cast Storage

Waste from Reprocessing Spent Fuel The Closed Fuel Cycle Spent fuel to be reprocessed for mixed oxide (MOX) fuel remains in the pool for several years before removal. The PUREX (link to unit on plutonium production)LINK process, used both for extracting plutonium and uranium in the closed fuel cycle and plutonium for weapons, generates large volumes of chemical and radioactive waste. In addition, the small amount of highly radioactive material remaining in spent reactor fuel after extraction of the uranium and plutonium poses a significant waste management problem. Lastly, today the mixed uranium and plutonium oxides (MOX) from reprocessing are used only once in thermal reactors due to the buildup of neutron absorbing Pu-240. Thus, this spent MOX fuel becomes waste to be managed. Waste from Nuclear Weapons Production The highly radioactive liquid waste from reactors used to produce plutonium for the nuclear weapons of the United States is stored in tanks (Figure 4) at the Hanford, Washington, and Savannah River, South Carolina. The Hanford site manages the largest volume of high-level waste, but the Savannah River site contains more total radioactivity. At Hanford, high-level waste alkaline liquid, salt cake, and sludge are stored in 149 single-shell and 28 double-shell underground tanks, while Savannah River has 51 tanks. These tanks contain approximately 88 million gallons of liquid, which is not only radioactive but also chemically toxic. The composition of the liquid varies from tank to tank. These facilities produced a combined total of 120 tons of plutonium for 20,000 nuclear warheads. Waste Storage Tank

The U.S. Department of Energy has begun a process of mixing this waste with sand at high temperatures to form a liquid glass mixture, which is poured into stainless steel canisters where it solidifies and is sealed for permanent storage. This method of stabilization, known as vitrification, has also been used to process waste from power reactors. Types of Nuclear Waste According to the U.S. Department of Energy (DOE), the four major elements of the environmental legacy of nuclear weapons production are waste, contaminated environmental media, surplus facilities, and materials in inventory. We will focus on the first two components. As we have seen in previous modules, nuclear weapons production in the United States was a complex series of manufacturing operations that generated large quantities of nuclear and chemical wastes. The term waste is defined as solids or liquids that are radioactive, chemically hazardous, or both. This waste consists of materials that have been disposed of previously, await disposal, or have been retrieved in site cleanups and are currently in storage. Waste is measured in terms of its volume (cubic meters) and its radioactivity (curies). Waste from nuclear weapons production managed by DOE includes 24 million cubic meters containing 900 million curies. The major categories of waste are high-level waste, transuranic waste, low-level waste, mixed low-level waste, 11e(2) byproduct material, hazardous waste, and ther waste. High-level waste is the highly radioactive waste resulting from spent nuclear fuel from production or power reactors, as well as from the chemical processing of spent nuclear fuel and irradiated target assemblies. The radioactivity comes from fission fragments and their daughter products resulting from the fission of U235 in production reactors. Although radiation from short-lived fission products (fragments and their daughters) will decrease dramatically in the next hundred years, radiation risks associated with the long-lived products will remain high for thousands of years. In the initial decay period, most of the radioactivity is due to Cs-137, Sr-90, and their short-lived daughter products. Plutonium, americium, uranium, and their daughter products are the major contributors to long-term radioactivity (Figure 5).

Radioactive Decay of High-level Waste from Reprocessing One Tonne of Spent Reactor Fuel

Transuranic (TRU) waste contains alpha-emitting transuranic elements or actinides with half-lives of greater than 20 years and a combined activity of greater than 100 nanocuries per gram of waste. Because of the long half-lives of many TRU isotopes, TRU waste can remain radioactive for hundreds of thousands of years. Some common isotopes found in TRU are Pu-238, Pu-239, Pu-240, Pu-241, Pu-242, Am-241, and Cu-244. TRU waste results from the fabrication of plutonium components, recycling of plutonium from scrap, retired weapons, and chemical separation of plutonium. Unlike high-level waste, which results from a few specific processes with a narrow range of physical matrices and chemical characteristics, TRU waste exists in many forms with a spectrum of chemical properties. A small percentage of TRU waste exhibits high direct exposure hazards and is referred to as "remotehandled" TRU waste. The majority of TRU waste emits low levels of direct radiation and is called "contact-handled" waste. The chief hazard of "contact-handled" waste is due to alpha radiation. Alpha particles cannot penetrate the skin but cause serious localized tissue damage when inhaled or ingested. When inhaled, TRU elements tend to accumulate in the lungs; soluble TRU compounds migrate through the body, accumulating in the bone marrow and liver. Mixed low-level waste contains both chemically hazardous waste subject to the Resource Conservation and Recovery Act (RCRA) and radioactive materials. The radioactive component of mixed low-level waste is similar to low-level waste and thus less radioactive than high-level or TRU waste. Hazardous chemical components present in mixed waste include toxic heavy metals, explosives, halogenated organic compounds, and acids. By-product materials include waste from uranium production described above. The other category is defined by government regulations. A variety of materials not covered previously fall into these categories. These materials include polychlorinated biphenyls, asbestos, and byproduct materials that have been mixed with chemically hazardous substances. Disposal More and more nations are now taking to nuclear energy as it is a cleaner and more efficient source of energy. But like any other method, there are drawbacks associated with it. One issue is the problem of nuclear waste disposal. Why is nuclear waste an issue? Nuclear waste is a cause for concern because it is not bio-degradable, meaning it does not decompose naturally under the affect of the atmosphere. Secondly, it causes a number of health hazards for anyone who comes into contact with the radiation from this waste. Harmful radioactive emissions can cause skin cancer and genetically alter the DNA of people coming into contact with them, the effects of which will be passed on to the descendants of these victims for many generations to come.

Sources of nuclear waste Nuclear waste is mainly created by the following sources: 1. Spent fuel from nuclear reactors; 2. Waste left after reprocessing of spent fuel; 3. Waste obtained from dismantling of nuclear weapons; and 4. Waste from industrial, medical and other sectors. Methods used for disposal of nuclear waste 1. Deep ocean disposal: In this method, containers made of borosilicate glass are filled up with nuclear waste. This glass has the capacity to prevent any nuclear radiation from leaking out. The container is enclosed in yet another water-tight metal container and dumped into the ocean. Though these containers are said to be leak-proof, there is a speculation that a minor quantity of radiation does manage to escape from these containers. 2. Deep geological burial: The containers used in this method are similar to the ones used in the deep ocean disposal system. But in this case, the containers are buried deep underground, in less-populated areas. This method depends on the natural decaying ability of the radioactive material for its success. The materials are buried deep under the earth for thousands of years and allowed to settle into a safe level of radioactivity. 3. Nuclear waste recycling is a new waste disposal method being looked into, in which the uranium, plutonium and other fission products are separated into different streams using chemical processes. The advantage of doing this is that these products can be re-used or disposed of more easily. However, the recycling process is not feasible at present because many countries do not have proper facilities to implement this method, and it is also a costly process. Radiation Hazards and prevention The radiation produced by nuclear reactions interacts with living tissue in many ways depending on the type of radiation. This radiation includes high-energy, charged particles (alpha and beta), neutrons of various energies, and photons (gamma and x-rays). In addition to this primary radiation, fission also produces radioactive isotopes of many elements, which in turn can emit particles and photons, known as secondary radiation. Many of these isotopes, such as strontium and iodine, can enter the body, where they replace nonradioactive elements and remain there emitting ionizing radiation. In many ways, the presence of these radioactive isotopes is more insidious than direct radiation from external sources that is more easily detected and reduced by proper shielding. Living tissue contains large amounts of water and the light elements, hydrogen, carbon, nitrogen, and oxygen, with lesser quantities of phosphorus and sulfur. The abundance of these light nuclei affects the interaction of nuclear radiation with living tissue. Alpha and beta particles, protons, and fission products of sufficient energy, as well as photons (gamma and x-rays), can remove the valence electrons of the elements in living materials. This process, known as ionization, changes the chemical reactivity of the affected atoms. Molecules containing ionized atoms may react to form substances that are detrimental to life. The ionizing particles and fission products may undergo thousands of collisions with the atoms before stopping thus they may cause extensive ionization in the target material.

The interaction of a high-energy photon with matter is very different than that experienced by a charged particle. Instead of losing its energy through large numbers of collisions in a relatively short distance, a photon loses its energy in a single interaction with an atom of the target material. The outcome of this interaction depends on the energy of the photon; it can produce secondary electrons and photons that can continue the transfer of energy to nearby atoms. Because gamma and x-ray photons, as well as neutrons, transfer energy to neighboring atoms by means of secondary electrons and photons, the radiation they produce is known as indirect ionizing radiation. Conversely, the radiation produced by alpha particles and protons is known as direct ionization because its energy is transferred directly to the target material.

Because neutrons have short half-lives (approximately 640 seconds) they do not exist in the absence of fission. In collisions with the light nuclei present in living material, fast neutrons produce indirect ionizing radiation. When slow neutrons are captured by target nuclei, most of their energy is dissipated as heat and thus little ionization is produced. Thus, the net effect is that neutrons damage tissue by causing both ionization and heating. All the processes described so far are physical and occur within approximately 10 -16 of a second of interacting with living tissue. After this, some energized atoms and molecules can dissipate their energy by becoming free radicals and other excited molecules. Free radicals are atoms or molecules that are electrically neutral and have an unpaired valence (outer) electron; they are extremely reactive chemical species. Because living tissue contains a large amount of water, the primary free radicals are H and OH; the former is a strong reducing agent and the latter, which predominates, is a strong oxidizingagent. These free radicals are known to cause abnormal chemical reactions in living substances. There are at least three ways to measure the effect of radiation. Each method has specific units of measurement associated with it: Curies(Ci) and becquerels(Bq) measure the activity as number of disintegrations per second. Activity is a property of the source of radiation; it indicates nothing about the effect of the radiation on the target material. Radsand grays (Gy) (Gy) are used in measuring the energy of radiation absorbed by the target material in joules per kilogram. Sieverts (Sv)or rems measure dose equivalents, a quantity used in radiological protection. The units of sieverts are also joules per kilogram. The sievert and gray both measure the effect of radiation on the target, but the sievert takes into account the effects of different types of radiation on human tissue. Equal exposure to different types of radiation does not necessarily produce equal biological effects. For example, one gray of alpha radiation will have a greater effect than one gray of beta radiation. One sievert of radiation produces a constant biological effect regardless of the type of radiation. The gray results from direct measurements made on tissue or cells, whereas the sievert is a radiological protection unit based on the best available evidence that includes quality and weighting factors.

The ability to penetrate matter differs greatly among the various types of nuclear radiation. A sheet of paper, a layer of clothing, or an inch of air can stop relatively slow moving, heavy alpha particles. Thus, it is easy to shield against alpha radiation, unless the alpha-emitting substance enters the body. Beta particles are lighter and travel faster than alpha particles. They can penetrate a fraction of an inch in solids and liquids and several feet in air. Gamma rays and neutrons are electrically neutral and thus not slowed by collisions with the target materials. They do not interact strongly as the charged alpha and beta particles do and are therefore highly penetrating. Their ability to penetrate the target material depends upon their energy. High-energy gamma rays may require several feet of material for adequate shielding. When radiation strikes complex biological molecules, such as proteins or nucleic acids, it may fracture the molecules and prevent their proper functioning. This can result in loss of cell vitality, decreased enzyme activity, initiation of cancer, and genetic mutations. The immediate effects of acute exposure to radiation are caused by free radicals rupturing the cell membranes. This rupturing causes the cells to lose their contents and die. If enough cells are killed, functions associated with the cells cease. Death occurs because of the direct loss of vital organs or because of secondary infections resulting from the breakdown of the immune system. The effect depends on the dose of radiation received. High doses of more than 100 gray affect the central nervous system, resulting in loss of coordination (including breathing problems), with death occurring within 1 or 2 days. Doses from 9 to 100 grays damage the gastrointestinal tract, causing nausea, vomiting, and diarrhea. Progressive dehydration can result in death within several weeks. Lower doses (from 3 to 9 grays) damage the bone marrow and other haematopoietic tissues. This can lead to loss of appetite and hair, hemorrhaging, inflammation, and secondary infections such as pneumonia. These effects are also found in patients undergoing radiation therapy. Doses of less than 3 grays are rarely lethal, but cause symptoms that include loss of appetite and hair, hemorrhaging, and diarrhea. Long-term risks of radiation exposure center on the incidence of cancer and genetic mutations. Both effects have proven difficult to determine since cigarette smoking, diet, and sunlight exposure are also known to cause cancer and possibly to promote genetic mutations. Analysis of approximately 100,000 survivors of Hiroshima and Nagasaki shows a slight increase in genetic mutations over what would be expected for a normal population, but this increase is so small that it may not be statistically significant.

Hiroshima survivors who received more than 2 gray showed a slight increase in the instances of cancer over a normal population. There is a good deal of controversy over the contribution of natural background radiation and man-made radiation to the incidence of cancer in humans. The dosage threshold below which radiation has no effect is also controversial. Background radiation comes from three sources: cosmic rays, naturally occurring radioactive elements such as radon-222, and solar radiation. The amount of exposure to this natural radiation depends on a number of factors, such as geographic location, house construction materials, medical treatments, and occupation. The average exposure for a U.S. resident is 0.36 rem per year. The rapidly dividing cancer-producing cells are most susceptible to radiation damage. Thus radiation is used to treat many types of cancer. It has, however, proven most effective against individual tumors. Focused radiation beams can shrink or eliminate cancerous tumors while doing less damage to surrounding tissue. In most cases, the benefits of radiation therapy outweigh the risks to the patients health. Radiation is also used in medical diagnosis. X-ray imaging and radioactive tracer studies provide information useful in treating patients.

Radiation therapy Radiation, like other aspects of nuclear science, can be both destructive and beneficial. The intelligent use of radiation for the treatment of cancer, medical diagnosis, food preservation, and other useful applications requires an informed public. Likewise, the solutions to the storage of nuclear waste also necessitate public understanding of the effects of nuclear radiation. Nuclear Proliferation Nuclear weapons have spread from nation to nation since the development of the first atomic bombs by the United States during World War II (Figure 1). Eight countries currently possess these weapons along with Israel, which has never admitted possession but is thought to have a small arsenal (Table 1). The primary motivations for a nation to obtain these weapons are fear of rival nations, international prestige, and national pride.

Nations Possessing Nuclear Weapons (2009) Country Program Initiated

United States 1942 Soviet Union/Russia 1945 United Kingdom 1942 France 1949 China 1955 India 1965 Pakistan 1972 North Korea 1971(?) Israel 1952 *Operational and stockpiled All numbers are estimates and are further described in the Nuclear Notebook in theBulletin of the Atomic Scientists, and the nuclear appendix in the SIPRI Yearbook. Additional reports are published on the FAS Strategic Security Blog. Unlike those publications, this table is updated continuously as new information becomes available. Current update: April 2, 2009. Since its inception in 1975, the Nuclear Non-proliferation Treaty (NPT) has helped to limit the number of nuclear armed nations to the five that possessed these weapons at the time it came into force (the United States, Russia, the United Kingdom, France, and China) along with four other nuclear states (India, Israel, Pakistan, and North Korea) that have refused to join the treaty. The NPTs security framework formed the basis for diplomatic efforts that have led several states to abandon their nuclear weapons programs (Argentina, Brazil, Sweden, and Libya). They have also resulted in four nations giving up their nuclear weapons (South Africa, Belarus, Kazakhstan, and Ukraine). However, today uncertainty concerning the purpose of the nuclear program of Iran (a signatory of the NPT), the nuclear weapons program of North Korea (a country who withdrew from the NPT), and the recent nuclear agreement between the United States and India are challenging the treaty. Among the responsibilities of the International Atomic Energy Agency (IAEA), created in 1957 as an agency of the United Nations, is that of assuring that the nuclear materials and activities of NPT signatory nations are not intended for military purposes. To this end, IAEA inspects and verifies nuclear facilities and activities in these countries. The agency has no enforcement powers but rather reports to the UN Security Council, which can then take action against violators. Twenty-one countries with civilian nuclear facilities have signed the NPT as non-nuclear-weapons states and have accepted IAEA inspections. Requirements for Producing Nuclear Weapons A nation seeking to produce nuclear weapons must complete the following basic steps: Develop a weapon design or obtain one from an external source. Produce highly enriched uranium or plutonium for the core of the device or obtain this material from an external source. Fabricate this material into the fissile component of the weapon. Fabricate or obtain from external sources the non-nuclear components of the weapon. These include high explosives and a triggering mechanism to detonate the nuclear core. Verify the reliability of all components individually and as a system. Assemble the components into a deliverable weapon (Figure 2). The elements of design and fabrication are complex but publicly available. The major barrier to producing a nuclear device is obtaining the fissile material, either highly enriched uranium (HEU) or plutonium. The IAEA states that 25 kg of HEU (> 90% U-235) or 8 kg of plutonium are the minimum amounts required for a 20-kiloton explosion (equivalent to the Nagasaki bomb). However, a nation with more sophisticated technology could build the same weapon with as little as 5 kg of HEU or 3 kg of plutonium. The production of HEU and plutonium are presented in units 8 and 10 (links to these units)

First Weapon Tested Current Number of Weapons (2009)* 1945 ~9,400 1949 ~13,000 1952 185 1960 300 1964 240 1974 60 1998 60 2006 <10 1981(?) 80

Figure 2 Components of a Nuclear Weapon

You might also like