m17h

89
Natural Sciences Tripos Part III MATERIALS SCIENCE M17: Nuclear Materials (Lectures 18) Dr J. H. Gwynne Lent Term 2013-14 III Name............................. College..........................

Upload: miha2filip

Post on 03-Oct-2015

6 views

Category:

Documents


0 download

DESCRIPTION

vc hhjj

TRANSCRIPT

  • Natural Sciences Tripos Part III

    MATERIALS SCIENCE

    M17: Nuclear Materials (Lectures 18)

    Dr J. H. Gwynne

    Lent Term 2013-14

    III

    Name............................. College..........................

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 1 -

    INTRODUCTION

    In existing technologies for nuclear power generation there are many performance, reliability

    and safety issues centred on the materials in use. In particular, materials degradation is the

    principal obstacle to extending the lifetime of an existing plant. In almost every case, materials

    issues pose the greatest challenge in bringing next-generation reactor designs to fruition. Of

    course, many of the challenges in nuclear power generation, however severe, are of a general

    kind; examples are high temperatures and corrosive environments.

    The focus of this part of the course will be the specific effects of radiation on structural

    materials and in the reactor core. The effects of radiation can be dramatic: change in shape,

    swelling by some tens of percent, hardening (more than five-fold), drastic embrittlement and

    reduction in ductility, and accelerated corrosion effects such as environmentally induced

    cracking.

    Useful books

    K.L. Murty & I. Charit, An Introduction to Nuclear Materials Wiley (2013)

    B.R.T. Frost (ed.), Nuclear Materials (Vols 10A & 10B, Materials Science & Technology),

    VCH (1994)

    B.M. Ma, Nuclear Reactor Materials and Applications, Van Nostrand (1983)

    G.S. Was, Fundamentals of Radiation Materials Science, Springer (2007)

    Additional Resources

    DoITPoMS TLP: Materials for nuclear power generation

    http://www.doitpoms.ac.uk/tlplib/nuclear_materials/index.php

    CES Edupack Nuclear Power Edition

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 2 -

    Contents

    1. Introduction 5

    1.1 Fundamentals 5

    1.1.1 Notation 5

    1.1.2 Binding energy 5

    1.1.3 Types of radiation and radioactive decay 6

    1.1.4 Exponential decay and half-life 7

    1.1.5 Neutron classification 8

    1.2 Nuclear Reactions 9

    1.2.1 Interactions of neutrons with matter 10

    1.2.1.1 Elastic and inelastic scattering 10

    1.2.1.2 Neutron capture and activation 10

    1.2.1.3 Fission 10

    1.2.1.4 Neutron cross-section 12

    1.2.2 Fusion 14

    1.3 Nuclear Reactors 15

    1.3.1 Components 15

    1.3.2 General and specific materials considerations 16

    1.3.3 Types of fission reactor 18

    1.3.4 Generation I reactors: Magnox 19

    1.3.5 Generation II reactors 20

    1.3.5.1 Pressurised Water Reactor (PWR) 20

    1.3.5.2 Boiling Water Reactor (BWR) 21

    1.3.5.3 Advanced Gas-cooled Reactor (AGR) 21

    1.3.5.4 Other types of Generation II reactor 22

    1.3.6 Generation III and IV reactors 22

    2. Radiation Damage 23

    2.1 Introduction 23

    2.2 Knock-on atoms and displacement cascades 23

    2.3 Dislocation loops 25

    2.4 Nucleation of cavities and voids 27

    3. Nuclear fuels 28

    3.1 Introduction 28

    3.2 Metallic fuels 30

    3.2.1 Uranium 30

    3.2.1.1 Structure 30

    3.2.1.2 Thermal expansion and thermal cycling growth 31

    3.2.1.3 Thermal conductivity 32

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 3 -

    3.2.1.4 Mechanical properties 33

    3.2.1.5 Corrosion 34

    3.2.1.6 Irradiation growth 34

    3.2.1.7 Irradiation swelling 35

    3.2.1.8 Irradiation creep 37

    3.2.2 Plutonium 38

    3.2.3 Thorium 39

    3.3 Ceramic fuels 40

    3.3.1 UO2 41

    3.3.1.1 Structure 42

    3.3.1.2 Irradiation effects 43

    3.3.2 Carbide and nitride fuels 46

    4. Cladding 47

    4.1 Introduction 47

    4.2 Austenitic stainless steels 47

    4.2.1 Helium production and void swelling 48

    4.2.1.1 Minimising void swelling 52

    4.2.2 Inverse Kirkendall effect 53

    4.2.3 Dislocation densities 54

    4.2.4 Creep 55

    4.2.5 Mechanical properties 56

    4.2.6 Summary 58

    4.3 Ferritic alloys 59

    4.4 Zirconium alloys 60

    4.4.1 Structure 61

    4.4.2 Effects of irradiation 61

    4.4.2.1 Irradiation growth 62

    4.4.2.2 Irradiation creep 63

    4.4.2.3 Mechanical properties 63

    4.4.2.4 Corrosion 64

    5. Moderators 65

    5.1 Introduction 65

    5.2 Graphite 67

    5.2.1 Structure 67

    5.2.2 Effect of irradiation on properties 68

    5.2.3 Wigner energy 69

    5.3 Other solid moderators 70

    5.3.1 ZrH 70

    5.3.2 Beryllium 70

    5.4 Liquid moderators 70

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 4 -

    6. Control rods 71

    6.1 Introduction 71

    6.2 Materials used for control rods 72

    7. Corrosion of structural components 73

    7.1 Introduction 73

    7.2 Stress-corrosion cracking (SCC) 73

    7.3 Irradiation-assisted stress-corrosion cracking (IASCC) 77

    7.3.1 Irradiation effects: radiolysis of water 78

    7.3.2 Irradiation effects: persistent effects 78

    7.3.2.1 Stress 78

    7.3.2.2 Segregation 78

    7.3.2.3 Hardening 80

    8. Summary of radiation damage and effects 81

    Glossary 82

    Abbreviations 84

    Question sheet 1 85

    Question sheet 2 87

    Examples class 88

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 5 -

    1. Introduction

    1.1 Fundamentals

    1.1.1 Notation

    A nuclide (any nucleus, any isotope of any element) can be represented as:

    where: A is the mass number (number of nucleons, i.e. number of protons and neutrons)

    Z is the atomic number (number of protons in the nuclide)

    N is the neutron number

    A = Z + N

    The nuclide can be represented in various ways; for example can also be written as

    , , or U-235.

    1.1.2 Binding energy

    Each nucleus has an associated binding energy. The total binding energy of a nucleus is the

    energy released when a nucleus is assembled from individual nucleons: the greater the energy

    release, the lower the potential energy of the nucleus. It is equivalent to the energy required to

    split a nucleus into its component parts. Therefore the higher the binding energy, the more

    stable the nucleus. If a nucleus is converted to another (or others) of higher binding energy, the

    difference in the total binding energies of the nuclei is released as kinetic energy of the

    resulting particles and gamma rays.

    The graph below shows binding energy per nucleon as a function of atomic mass.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 6 -

    1.1.3 Types of radiation and radioactive decay

    Some large, unstable nuclei can become more stable by spontaneously undergoing radioactive

    decay (others undergo fission, as well see in section 1.2).

    Alpha radiation

    An alpha particle is essentially a helium nucleus: two protons and two neutrons, or .

    The most common source of alpha particles is the alpha decay of heavy atoms: when an atom

    emits an alpha particle, its mass number decreases by four and its atomic number decreases by

    two. This is an example of transmutation: the conversion of one chemical element or isotope

    into another. For example, when undergoes alpha decay, it emits an alpha particle and

    forms :

    Beta radiation

    Beta particles are high-energy, high-speed electrons ( ) or positrons ( ) emitted by certain

    radioactive nuclei. The production of beta particles is known as beta decay, which also

    involves transmutation.

    If an unstable nucleus has an excess of neutrons, it may undergo decay, in which a neutron

    is converted into a proton and an electron (and an antineutrino). An example is the decay of

    into

    :

    Alternatively, if an unstable nucleus has an excess of protons, it may undergo decay, in

    which a proton is converted into a neutron and a positron (and a neutrino). An example is the

    decay of into

    :

    Gamma radiation

    Gamma radiation, , is high-frequency (and therefore high-energy) electromagnetic radiation.

    Gamma rays are produced during gamma decay, which occurs following alpha or beta decay.

    The daughter nucleus is usually left in an excited state, so it can move to a lower energy state

    by emitting a gamma ray.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 7 -

    1.1.4 Exponential decay and half-life

    Radioactive isotopes are subject to exponential decay (so long as the remaining number of

    radioactive atoms is large): in a given sample, the number of radioactive atoms decreases at a

    rate proportional to its current value:

    where N(t) is the quantity of undecayed atoms after time t

    is the decay constant

    The solution to this differential equation is:

    where N0 is the initial number of radioactive atoms

    is the mean lifetime of the radioactive atom

    and are related by

    The half-life, , of a radioactive isotope is the length of time after which there is a 50%

    chance that an atom will have undergone nuclear decay. It can also be thought of as the time

    after which half the radioactive atoms will have decayed. It therefore follows that:

    and can be related to the half-life by:

    Therefore,

    Note that N(t) can also be written as:

    The half-life varies depending on the isotope, and is usually determined experimentally. Half-

    lives can vary from 10-24

    s to 10+30

    s (54 orders of magnitude!). For a list of half-lives, see:

    http://en.wikipedia.org/wiki/List_of_isotopes_by_half-life

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 8 -

    1.1.5 Neutron classification

    So far, we have considered spontaneous decay. However, in a nuclear reactor, decay is

    stimulated by neutron impact so it is useful to consider the way in which neutrons are classified.

    This is usually done based on their kinetic energies. Typical values are given in the table below.

    Category Energy

    Cold neutrons 10 MeV

    Generally, thermal neutrons are associated with a kinetic energy of ~0.025 eV (corresponding

    to a speed of 2200 ms-1

    ).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 9 -

    1.2 Nuclear Reactions

    A nuclear reaction is the process in which two nuclei, or one nucleus and a subatomic particle,

    collide to produce one or more nuclides that are different from the nuclide(s) that began the

    process. A nuclear reaction therefore always involves transmutation. Nuclear reactions may

    involve alpha particles, neutrons, protons, electrons or positrons.

    The first study of a nuclear reaction was carried out by Ernest Rutherford in 1919. He was the

    first person to deliberately transmute one element into another: he used alpha radiation to

    convert nitrogen into oxygen through the following reaction:

    An alpha particle is absorbed and a proton is emitted. This reaction can also be written in the

    compact form:

    Here is regarded as the target nuclide,

    as the product nuclide. This experiment

    showed Rutherford that hydrogen nuclei formed part of nitrogen nuclei (and therefore probably

    other nuclei too) and led him to suggest that a hydrogen nucleus was possibly a fundamental

    building block of all nuclei and perhaps also a new fundamental particle. He named it the

    proton in 1920.

    If a reaction like this is endothermic, then there is a threshold energy for it to be possible.

    There is also the Coulomb barrier, which is the energy needed to overcome the electrostatic

    repulsion of approaching positively charged nuclides.

    Note that nuclear equations describe nuclear reactions without considering charges.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 10 -

    1.2.1 Interactions of neutrons with matter

    Most nuclear reactions that we will consider in this course involve neutrons interacting with

    nuclei, we will now consider different ways in which this can happen.

    1.2.1.1 Elastic and inelastic scattering

    Elastic scattering refers to a collision between a neutron and a nucleus in which kinetic energy

    and momentum are both conserved. In inelastic scattering, momentum is conserved but kinetic

    energy is not: the neutron loses kinetic energy, resulting in the emission of gamma radiation.

    Whether elastic or inelastic scattering occurs depends on factors such as the speed of the

    neutron, and the neutron cross-section of the nucleus (see below).

    Note that this is not a nuclear reaction as defined above, since scattering does not involve

    transmutation.

    1.2.1.2 Neutron capture and activation

    If a nucleus captures a neutron, a heavier nucleus is formed, which may cause it to become

    radioactive. This then makes the material more difficult to handle (repair, replace, recycle,

    dispose of) safely. An example is cobalt, which is a common alloying addition. The usual

    isotope is , but this is activated by neutron irradiation:

    has a half-life of 5.3 years.

    1.2.1.3 Fission

    Nuclear fission was discovered experimentally in December 1938 by Otto Hahn and his

    assistant Fritz Strassmann, and explained theoretically in January 1939 by Lise Meitner and her

    nephew Otto Robert Frisch (Hahn won the 1944 Nobel prize in chemistry for the discovery of

    nuclear fission). During fission, the nucleus of an atom splits into two lighter nuclei.

    Only one naturally occurring nuclide shows spontaneous fission (and not very actively). One

    possible fission reaction is:

    Nuclear power generation relies on fission induced by incident neutrons. If a nucleus can

    undergo fission regardless of the incident neutron energy (even if the probability of this

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 11 -

    occurring is low), the nucleus is referred to as fissile, whereas if there is a threshold energy

    needed for fission to occur, the nucleus is referred to as fissionable. Examples of fissile nuclei

    include , and .

    The most commonly used fissile nuclide in thermal reactors is . When absorbs a

    neutron, it leads to the formation of the unstable radionuclide , which immediately splits

    into two smaller nuclei (typically of unequal mass), known as fission fragments. There are

    many possible fission reactions, for example:

    However, a neutron will not necessarily induce fission if it passes through the nucleus: for

    example, fast neutrons are less likely to induce fission in than thermal neutrons the

    faster a neutron is travelling, the less time it spends inside the nucleus and therefore the less

    opportunity it has to induce fission.

    The fission event must emit more than one neutron if the reaction is to be sustained (and

    therefore to create a chain reaction). Each fission of generates an average of 2.4 neutrons,

    and each fission of gives an average of 2.9 neutrons. Each fission event typically

    releases about 200 MeV of energy, about 35% of which is converted to electrical energy in

    power stations.

    The fission products typically exhibit a decay series. For example , a fission product of the

    above reaction, shows an isobaric (conserving mass number) decay series:

    Neutron economy plays an important role in the design of nuclear reactors, since there are

    several ways in which the released neutrons can be used up:

    Fission of a fissile nucleus

    Non-fission capture by the fuel or other components in the reactor core

    Leakage of neutrons from the core

    There is a certain minimum size of a chain reacting system, called the critical size, for which

    the production of neutrons by fission just balances the loss of neutrons to the other processes,

    and the reaction can be sustained independently. The associated mass is the critical mass. If

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 12 -

    more neutrons are lost than produced, the reactor is said to be subcritical, and if more are

    produced than lost, the reactor is said to be supercritical.

    Reactors can also be used to create (breed) fuel, the most common case being the production of

    from fissionable using fast (E = 1 20 MeV) neutrons:

    is then usable as a fissile fuel in thermal reactors.

    1.2.1.4 Neutron cross-section

    The likelihood of a nuclear reaction taking place is described using the appropriate cross-

    section, . A cross-section is a measure of the degree to which a particular nuclide will interact

    with neutrons of a particular energy and is roughly the effective projected area of the target

    nuclide. In conjunction with the neutron flux (roughly equivalent to the number of neutrons

    travelling through unit area in unit time), it enables calculation of the reaction rate (for example

    to calculate the thermal power of a nuclear power plant). Cross-sections are usually quoted in

    barns (1 barn (b) = 10-28

    m2).

    There are different neutron cross-sections, depending on the process being considered. For

    example, the absorption cross-section, a, describes the likelihood of a neutron being absorbed

    by a nuclide, whereas the scattering cross-section, s, describes the likelihood of a neutron

    being scattered by a nucleus. The total absorption cross-section includes the fission cross-

    section, f, and the capture cross-section (which is approximately equal to the cross-section for

    neutron absorption followed by gamma emission, ). The total cross-section is the sum of the

    individual cross-sections.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 13 -

    The neutron cross-section depends on the target type, the type of nuclear reaction (scattering,

    fission etc) and the kinetic energy of the species involved. To a lesser extent, it also depends on

    the angle between the incident neutron and the target nuclide and the target nuclide temperature.

    Note that represents the microscopic cross-section and is a property of a given nuclide,

    whereas the macroscopic cross-section, , takes into account the number of those nuclides

    present. For mixtures of isotopes and elements, the macroscopic cross-sections add.

    The range of a neutron is the distance it travels before being stopped and is a function of the

    neutron energy as well as the capture cross-section of the material through which the neutron is

    moving.

    For fission to be induced by thermal neutrons, the binding energy of the thermal neutron to the

    fissile nuclide must exceed the energy required for the nuclide to split (the fission barrier).

    Heavy nuclides show very different cross-sections for induced fission, f, and for neutron

    absorption followed by gamma emission, :

    Nuclide f (b) (b) Binding Energy (MeV) Fission Barrier (MeV)

    3 x 10-6 737 4.8 7.5

    530 48 6.8 6.0

    586 99 6.5 5.7

    3 x 10-6 2.7 4.8 5.8

    752 269 6.5 5.0

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 14 -

    1.2.2 Fusion

    Fusion will not be discussed in detail in this course, but it is worth including at this point for

    completeness.

    Energy is given off when a nucleus becomes stable (i.e. when it approaches the maximum on

    the graph in section 1.1.2). Whilst moving from heavier nuclei towards this maximum requires

    the nucleus to split apart (fission), moving from lighter nuclei towards the maximum requires

    two nuclei to combine and form a heavier one (fusion). The energy release per mass of nuclide

    is much higher for fusion than fission.

    The reaction of most interest for power generation, yielding 17.6 MeV per event, and with a

    Coulomb barrier of only 0.68 MeV is the fusion of deuterium and tritium:

    There are, however, many technical challenges and no commercial fusion reactors currently

    exist. It is unlikely that any will be set up for some years, but fusion for power generation is a

    prominent research topic. Experimental reactors are in the process of being built, such as ITER

    (International Thermonuclear Experimental Reactor), which is planned to be completed by

    2018.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 15 -

    1.3 Nuclear Reactors

    1.3.1 Components

    The main components of nuclear reactors will be discussed in more detail in subsequent

    sections, but a summary is given here.

    The fuel can be in metallic, alloy or ceramic form and is generally contained within tubes made

    of a metallic alloy (cladding). The most commonly used fuels are enriched uranium (uranium

    in which the percentage of has been increased), uranium oxides and plutonium oxides.

    This cladding provides mechanical support to the fuel, prevents fission products from leaving

    the fuel element and protects the fuel from corrosion caused by the coolant.

    The fuel elements are typically arranged in a regular pattern (square, hexagonal etc) with the

    moderator.

    The moderator slows down neutrons to sustain the fission reaction with thermal neutrons.

    The fuel-moderator assembly is surrounded by a reflector to direct neutrons towards the core

    and to control neutron leakage (thereby improving neutron economy).

    Outside, the reactor is surrounded by shielding that absorbs neutrons and gamma rays and

    reduces the external radiation intensity to a tolerable level.

    Control rods help to control the chain reaction by absorbing neutrons to maintain a steady

    state of operation. Control rods are made from neutron-absorbing materials such as boron and

    hafnium. There are usually two types of control rod in a nuclear reactor: rods for routine

    control, which can be raised or lowered to increase or decrease the amount of heat being

    generated, and safety rods, which can be lowered to shut the reactor down in an emergency.

    The coolant removes the heat that is continually generated and is used to produce steam to

    drive turbines for electricity generation. The coolant can be gas or liquid: examples include

    light or heavy water, carbon dioxide, liquid metals and molten salts. A careful balance is

    needed between the reduction in neutron economy due to the presence of the coolant and the

    efficiency of heat removal.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 16 -

    1.3.2 General and specific materials considerations

    1.3.2.1 General Considerations

    There are many important general materials considerations when choosing materials for use in

    nuclear reactors.

    Mechanical properties the materials used should be strong enough to bear the normal

    loads that the structure would be subjected to and withstand internal or external stresses,

    and ductile enough to avoid any catastrophic failure.

    Ease of fabrication including forming, welding, machining etc.

    Dimensional stability many components in power reactors are required to work at

    high temperatures for extended times, so should be stable and creep-resistant.

    Corrosion resistance many components are in close contact with reactor fluids (such

    as coolants), so should be corrosion-resistant.

    Heat transfer properties particularly relevant for the fuel and cladding materials. The

    heat generated inside the fuel needs to be able to be conducted away efficiently.

    Availability and cost

    1.3.2.2. Specific Considerations

    Of particular interest in this course is the effect of irradiation on the properties of the materials

    used in nuclear reactors.

    Neutron properties for example, the fuel cladding materials need a low neutron

    absorption cross-section, whereas the control materials need a high neutron absorption

    cross-section.

    Susceptibility to induced radioactivity as described above, absorption of neutrons can

    lead to the formation of different isotopes, which may be radioactive. Considerations

    include the abundance of the isotopes, their half-lives and the type of radiation that they

    produce. For example, isotopes with a short half-life that emit low energy radiation are

    much less of a concern than those with a long half-life that emit high energy radiation.

    Radiation stability as we will see in later sections, radiation damage can lead to a

    variety of effects, including the formation of voids, embrittlement, creep and hardening.

    The requirements of materials used for different components within a nuclear reactor are

    summarised in the table below.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 17 -

    Component Main requirements Possible materials

    Cladding

    material

    Low neutron absorption

    Stability under heat and radiation

    Mechanical strength

    Corrosion resistance

    Good heat transfer properties

    Al, Be, Mg, Zr

    Stainless steels

    Ni-based superalloys

    Refractory metals (Mo, Nb, Ti,

    W etc)

    Moderators and

    reflectors

    Low neutron absoption

    Large energy loss by neutron per

    collision

    High neutron scattering

    Water (light or heavy)

    Beryllium (or BeO)

    Graphite

    Control

    materials

    High neutron absorption

    Adequate strength

    Low mass (for rapid movement)

    Corrosion resistance

    Stability under heat and radiation

    B, Cd, Hf, rare earths (Gd, Eu)

    Coolants Low neutron absorption

    Good heat transfer properties

    Low pumping power (low Tm)

    Stability under heat and radiation

    Low induced radioactivity

    Corrosion resistance

    Gases (air, H2, He, CO2, H2O)

    Liquid water (H2O and D2O)

    Liquid metal (Na, Na-K, Bi)

    Molten salts (-Cl, -OH, -F)

    Organic liquids

    Shielding

    material

    Capacity to slow down neutrons

    Absorption of radiation

    Absorption of neutrons

    Light water

    Concrete

    Most control materials

    Metals (Fe, Pb, Bi, Ta, W, Boral)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 18 -

    1.3.3 Types of Fission Reactor

    There are four generations of fission reactors:

    Generation I these have now mostly been retired

    Generation II these make up the majority of reactors still being used

    Generation III these offer minor improvements to the generation II reactors

    Generation IV these are futuristic designs that are currently being researched,

    although commercial construction is unlikely before 2030.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 19 -

    1.3.4 Generation I Reactors: Magnox

    The worlds first nuclear power station to deliver electricity in commercial quantities was the

    Calder Hall reactor in Sellafield, which had four Magnox reactors (and started operation in

    1956).

    Magnox reactors were used for the production of plutonium (for atomic weapons) as well as

    for electricity generation, but they have all now been decommissioned (or are soon to be

    decommissioned).

    They used a graphite moderator, carbon dioxide as coolant, and natural (i.e. unenriched)

    uranium as fuel. The cladding consisted of thin cylindrical tubes of a non-oxidising magnesium

    alloy (the name Magnox comes from magnesium nonoxidising), typically Mg-0.8Al-

    0.005Be. Mg has a low thermal neutron capture cross-section, and the alloy was creep-resistant,

    resistant to corrosion by CO2 and (unlike pure Al) does not react with uranium. The Al

    provided solid solution strengthening and Be improved oxidation resistance.

    A schematic of a Magnox reactor is shown below. CO2 circulated under pressure through the

    reactor core, heated up and was then sent to the steam generator to produce steam, which in

    turn was used to drive a turbine.

    Schematic of a Magnox reactor (from Wikimedia Commons, attribution: Emoscopes)

    The disadvantages of this type of reactor include limited efficiency and power capacity due to

    relatively low maximum operating temperature (345C) this maximum was imposed due to

    concern about reaction of CO2 with graphite at higher temperatures.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 20 -

    1.3.5 Generation II Reactors: 1960s onwards

    Most commercial reactors operating today are Light Water Reactors (LWRs), of which there

    are two main types: Boiling Water Reactors (BWRs) and Pressurised Water Reactors (PWRs).

    PWRs make up 60% of the total currently operational reactors and BWRs account for another

    21%. As the name suggests, they use water (normal light water as opposed to heavy water)

    as the coolant and moderator.

    1.3.5.1 Pressurised Water Reactor (PWR)

    A PWR consists of two separate light water loops, primary and secondary, and is shown

    schematically below.

    The reactor core is located inside a reactor pressure vessel made of a low-alloy ferritic steel

    (typical dimensions are 5 m diameter, 12 m height and 30 cm wall thickness). The pressure

    vessel is internally lined with a reactor cladding of 308-type stainless steel or Inconel-617 to

    provide corrosion resistance. The primary loop operates at a pressure of 15-16 MPa, so that the

    water doesnt boil, even at temperatures of 320-350C.

    The core contains an array of fuel elements with stacks of slightly enriched UO2 pellets clad in

    Zircaloy-4. These cladding tubes are typically 10 mm in diameter and 0.7 mm in wall thickness.

    About 200 fuel rods are bundled together to form a fuel element, and about 180 elements are

    grouped to form an array creating the reactor core. The control rods are typically Ag-In-Cd or

    B4C.

    The primary loop is transported to a steam generator, where heat is transferred to the secondary

    loop system, creating steam, which is used to drive a turbine. The steam generator is essentially

    a heat exchanger containing thousands of tubes made from a nickel-bearing alloy or nickel-

    based superalloy supported by carbon steel plates.

    Schematic of a PWR (from Wikimedia Commons)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 21 -

    1.3.5.2 Boiling Water Reactor (BWR)

    A schematic of a BWR is shown below. The reactor pressure vessel in a BWR is similar to that

    for a PWR. However, in a BWR, there is only one water loop. The water is at a lower pressure

    than in a PWR (about 7 MPa), so that it boils in the reactor core - the normal steam temperature

    is 290-330C.

    The core consists of a fuel assembly comprising slightly enriched UO2 fuel clad with

    recrystallised Zircaloy-2 cladding tubes (12.5 mm outer diameter). Typically, a BWR fuel

    assembly would contain 62 fuel rods and 2 water rods in an 8 x 8 array.

    The control material is generally B4C dispersed in 304-type stainless steel matrix, or hafnium,

    and takes the form of blades arranged throughout the assembly in a cross-shape. Water passes

    through the reactor core producing steam and is dried at the top of the reactor vessel.

    Schematic of a BWR (from Wikimedia Commons)

    1.3.5.3 Advanced Gas-cooled Reactor (AGR): this type of reactor was developed from the

    Magnox design, but operates at a higher gas temperature to improve thermal efficiency. It was

    designed such that the final steam conditions are identical to those in conventional coal-fired

    power stations, meaning that the same design of turbo-generator plant could be used. It uses

    slightly enriched uranium as the fuel, graphite as the moderator and carbon dioxide as the

    coolant.

    It has a better thermal efficiency than a PWR, but the reactor core is larger for the same power

    output and the fuel is used less efficiently, which counters the thermal efficiency advantage.

    The AGR is designed to be refuelled without being shut down first, but fuel assembly vibration

    problems during refueling at full power meant that refueling is now only performed at lower

    power or when shut down.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 22 -

    1. Charge tubes

    2. Control rods

    3. Graphite moderator

    4. Fuel assemblies

    5. Concrete pressure vessel

    and radiation shielding

    6. Gas circulator

    7. Water

    8. Water circulator

    9. Heat exchanger

    10. Steam

    Schematic of an AGR (from Wikimedia Commons)

    1.3.5.4 Other types of Generation II reactor

    CANDU (Canadian Deuterium Uranium) reactor: this is essentially a pressurised heavy

    water reactor (PHWR). Instead of using enriched uranium as the fuel, it uses natural uranium,

    meaning that there are fewer neutrons available to sustain the reaction. A more efficient

    moderator is therefore needed, so heavy water is used because (deuterium) absorbs neutrons

    less readily than (in light water). This type of reactor also allows refuelling without shutting

    down.

    Fast Breeder Reactor (FBR): production of fissile material actually occurs in the fuel of all

    current commercial nuclear power reactors (towards the end of its life, a PWR fuel element

    produces more power from fission of plutonium than uranium), but a breeder reactor is a

    nuclear reactor in which more new fuel is produced than consumed during its operation. The

    reactor converts fertile material (containing and ) into fissile material ( and

    respectively).

    The most common type of FBR is a Liquid Metal Fast Breeder Reactor (LMFBR), in which

    liquid metal (usually Na) is used to transport the heat generated in the core. LMFBRs have a

    high power density due to lack of a moderator and higher temperatures can be achieved,

    leading to higher efficiency. However, Na must be carefully contained because it reacts readily

    with oxygen and water, and also becomes radioactive as it passes through the reactor core.

    1.3.6 Generation III and IV reactors

    Generation III reactors are mainly advanced LWRs, which are designed to be safer, more

    efficient, longer lasting and have a greater capacity. Generation IV reactors are futuristic

    designs that will offer further improvements.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 23 -

    2. Radiation Damage

    2.1 Introduction

    Radiation damage refers to microscopic defects produced in materials due to irradiation, and

    results in changes to their physical, chemical and mechanical properties. The interaction of

    high energy subatomic particles and radiation with crystal lattices can give rise to a variety of

    defects, including vacancies and self-interstitials. The majority of radiation damage is caused

    by neutrons and fission fragments: other types of radiation generally have insufficient energy

    (or are not produced in large enough quantities) to cause major damage.

    Radiation damage has been widely studied because of its importance in structural components

    in and near the cores of nuclear reactors. Such components are subjected to extreme conditions

    of various kinds, including high temperature, high stress and corrosion, but neutron irradiation

    leads to specific types of damage.

    Note that the term radiation effects is generally used to refer to the effects on the behaviour

    or properties of materials in the aftermath of radiation damage. Radiation effects will be

    covered in more detail in subsequent sections.

    2.2 Knock-on atoms and displacement cascades

    The key point is that the incident radiation is sufficiently energetic to displace atoms from their

    equilibrium sites in the crystal structure, since the binding energy of lattice atoms is typically

    small compared to the energy of the impinging particles.

    This creates vacancies and self-interstitials (one vacancy and one interstitial constitute a

    Frenkel pair). The damage follows the sequence:

    1. An energetic incident particle (usually a fast neutron) strikes an atom in the crystal

    2. The transfer of kinetic energy to the atom in the primary recoil is large enough to

    displace it from its lattice site and it becomes a primary knock-on atom, or PKA,

    leaving behind a vacant site

    3. The PKA moves through the lattice, creating further knock-on atoms in a displacement

    cascade

    4. The PKA and other knock-on atoms eventually come to rest as interstitial atoms

    A displacement cascade (also known as a displacement spike) is illustrated in the image below.

    There is a high density of vacancies in the core, with the surrounding material rich in

    interstitials.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 24 -

    The initial mean free path between collisions is of the order of 1 cm for fast neutrons but

    decreases as energy is dissipated in successive collisions; the result is that displacement

    cascades are concentrated in volumes 1 to 10 nm in diameter. Typical damage rates (measured

    as displacements per atom per second, dpa s-1

    ) range from 10-9

    dpa s-1

    in thermal reactors to

    10-5

    dpa s-1

    in the first wall of proposed fusion reactors.

    The displacement energy is the minimum energy that must be transferred to a lattice atom in

    order for it to be displaced from its lattice site. Generally, an average displacement energy of

    25 eV is assumed, but the actual value depends on a variety of factors including the material

    and its crystallographic structure, the trajectories of the knock-on atoms, and the thermal

    energy of the atoms. Generally, higher melting point metals tend to have higher displacement

    energies, as illustrated below.

    Variation of displacement energy as a function of melting temperature (from Murty & Charit, 2013)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 25 -

    If the energy transferred by a knock-on atom to the struck lattice atom is less than the

    displacement energy, the lattice atom will not be dislodged from its lattice site. It will instead

    vibrate around an equilibrium position and the energy will be dissipated as heat.

    Over the lifetime of a component, each atom could be displaced as many as 100 times. Clearly

    such extreme conditions can have profound effects on the microstructure of the alloys involved.

    These effects include dissolution of precipitates, changes in their morphology, and appearance

    of non-equilibrium phases. In this course, we will focus on dislocations and formation of voids.

    In principle, the vacancies and interstitials generated in displacement cascades might

    recombine to restore the equilibrium structure. However, in practice, several factors can give a

    substantial supersaturation of vacancies:

    Self-interstitial atoms form stable clusters and are ultimately removed by dislocation

    glide to other dislocations and grain boundaries; in effect, radiation damage has a bias

    towards the production of vacancies.

    Typically the vacancies and interstitials are created at a temperature high enough for

    them to be mobile. The interstitials are more mobile, which is an additional factor

    leaving the centres of displacement cascades vacancy-rich.

    Dislocations, through the process known as climb, can act as sinks for vacancies and

    interstitials, but the greater elastic strain around the latter again leads to their

    preferential removal.

    The supersaturation of vacancies leads to the appearance of new microstructural features.

    2.3 Dislocation loops

    Dislocation loops can form at lower temperatures (T < 0.2 Tm, where Tm is the absolute melting

    temperature of the irradiated alloy). The displacement cascade illustrated in section 2.2 can be

    thought of as a core of vacancies surrounded by a shell of interstitials. If the vacancy core or

    the interstitial shell collapse (condense) onto a close-packed plane, dislocation loops can be

    generated: collapse of the vacancy core results in a vacancy loop, whilst collapse of the

    interstitial shell results in an interstitial loop.

    In ccp metals, the stacking sequence of close-packed planes (the {111} planes) can be

    described by ABCABCABC. Both types of dislocation loop disrupt the stacking sequence of

    the planes, resulting in a stacking fault: vacancy condensation produces an intrinsic fault and

    interstitial condensation produces an extrinsic fault.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 26 -

    Removal of a layer of atoms (i.e. following the formation of a vacancy loop) results in an

    intrinsic (or single) fault in the stacking sequence, such as ABCAB/ABCABC (where /

    indicates the missing plane of atoms). Insertion of an extra plane of atoms (following the

    formation of an interstitial loop) produces an extrinsic (double) fault in the stacking sequence,

    such as ABCAB/A/CABC.

    The images below show an end-on view of a stack of close-packed planes in a crystalline metal:

    (a) with a vacancy loop (intrinsic fault); (b) with an interstitial loop (extrinsic fault).

    Dislocation loops (from Was, 2007)

    In these dislocation loops, the Burgers vector is normal to the planes and of magnitude equal to

    the interplanar spacing ( ). The dislocation is sessile (unable to glide).

    The formation of vacancy-type and interstitial-type loops under irradiation is a major

    contribution to the observed increase in dislocation density. The continuing evolution of

    dislocation density under irradiation, however, involves many processes and is not susceptible

    to analysis in terms of loop nucleation.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 27 -

    2.4 Nucleation of cavities and voids

    Irradiation at higher temperatures (T > 0.2 Tm), leads to cavities or voids, the nucleation of

    which is related to supersaturation not only of vacancies but also of dissolved helium atoms.

    This helium arises from irradiation-induced transmutation reactions (of B, Ni and Fe)

    accompanied by the emission of alpha particles: since alpha particles are positively charged,

    they easily pick up electrons from the surrounding lattice and become elemental helium. The

    rate of helium production is in the range of 0.512 atomic parts per million per dpa.

    The precipitation of helium atoms results in the formation of small bubbles, which can

    subsequently act as sinks for vacancies, thereby acting as a nucleation point for voids. The rate

    of swelling is much greater than can be accounted for solely by the helium production rate, and

    is mainly due to the condensation of vacancies.

    Cavity nucleation under irradiation has the feature, not often encountered in metallurgical

    precipitation, that the principal species (vacancies and helium atoms) are under continual

    production; without this, cavities already produced would largely disappear on annealing.

    From the first observations of cavity formation in 1967, it was recognised that this is a

    particularly important form of radiation damage, leading to swelling and distortion of irradiated

    components. At higher temperatures, it can cause hardening of irradiated alloys and associated

    embrittlement and loss of ductility. The development of voids on grain boundaries from initial

    helium bubbles also shortens creep life.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 28 -

    3. Nuclear fuels

    3.1 Introduction

    The basic requirements of a nuclear fuel (apart from being easily fissionable and preferably

    fissile) are:

    good thermal conductivity (to allow heat generated during fission to be removed)

    ideally a high melting temperature (to be able to run as hot as possible without melting)

    mechanical stability

    The design and operation of a reactor depend on the behaviour of the fuel. The relative cost of

    the fuel in nuclear power generation is low, so it is possible to expend resources on optimising

    it.

    There must be enough fuel in a reactor for the chain reaction to be self-sustaining. The total

    mass of fuel would be supercritical (leading to a runaway reaction) if it were in monolithic

    form (i.e. a single block), but it is sub-divided to permit moderation and control, and to allow

    heat extraction from the fuel to be efficient enough to prevent it from melting. The fuel can be

    in the form of plates or rods (mostly rods), and it is isolated from the coolant by cladding (see

    section 4).

    Typically, in a PWR, the fuel rods are 45 m long, and the cladding is 912 mm in diameter

    and 0.60.8 mm thick. Fuel assemblies have 200300 rods and there are 150250 such

    assemblies in the core, giving approximately 80100 tonnes of uranium in the reactor.

    Westinghouse fuel sub-assemblies for a PWR (ca. 1978) (from Frost, 1994; Ma, 1983)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 29 -

    The linear heat generation rate (LHGR) is the heat generation rate (i.e. power) per unit length

    of fuel rod (commonly expressed in kW/m) and is limited by the thermal conductivity of the

    fuel and the need to avoid melting in the centre. Typical values are 1570 kW/m.

    It is desirable to have the fuel rods inserted for longer in the reactor (so that they need to be

    replaced less frequently) and to achieve greater burn-up (to allow more efficient use of the

    fuel). Burn-up is a measure of how much energy is extracted from a primary nuclear fuel

    source and is quoted as GW-days per tonne of U (or equivalent), GWd/t.

    Generation II reactors were designed to achieve ~40 GWd/t. With newer fuel technology, and

    particularly the use of nuclear poisons, these reactors are now capable of ~60 GWd/t.

    Nuclear poisons (or neutron poisons) are neutron absorbers that are inserted into some reactors

    to lower the high reactivity of the fresh fuel (the poisons are said to have negative reactivity): if

    a reactor is designed to operate for a long period of time, more fuel than that needed for exact

    criticality must be used, so neutron absorbers are added to control the reaction. The positive

    reactivity from the excess fuel is balanced by the negative reactivity of the neutron absorber.

    These nuclear poisons are often burnable, meaning that their effect wears off over the

    lifetime of the fuel: ideally, the negative reactivity of the poison should decrease at the same

    rate that the fuels excess positive reactivity is depleted. Examples include B4C-Al2O3,

    borosilicate glass and Gd2O3.

    Some more advanced designs are expected to achieve >90 GWd/t from higher-enriched fuel,

    and eventually >200 GWd/t. Complete fission of all heavy metal in a breeder reactor, not just

    fissile content but also any fissionable or fertile material, would yield ~1,000 GWd/t.

    In a power station, high fuel burn-up is desirable for:

    reducing the downtime needed for refuelling

    reducing the number of fresh nuclear fuel elements required and spent nuclear fuel

    elements generated while producing a given amount of energy

    reducing the potential for diversion of plutonium from spent fuel for use in nuclear

    weapons

    It is also desirable that burn-up should be as uniform as possible both within individual fuel

    elements and from one element to another. Materials issues, particularly in the fuel and in the

    cladding severely restrict the burn-up that can safely be achieved.

    At present 98% of all nuclear-generated electricity comes from oxide fuels, but for historical

    and scientific reasons we first consider metallic fuels.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 30 -

    3.2 Metallic fuels

    3.2.1 Uranium

    Uranium is the basic nuclear fuel and is the basis for breeding new fuel. Pure uranium was used

    in the earliest reactors (such as the Magnox reactors at Calder Hall). Metallic uranium and its

    alloys are still used in teaching and research reactors for low-temperature operation, and are of

    interest for some future designs of reactor.

    In principle, metallic fuels have significant advantages:

    high density of fissile or fissionable nuclides

    good fabrication and machinability

    excellent thermal conductivity

    relative ease of reprocessing (through electrorefining)

    Disadvantages include:

    lower melting points

    various irradiation instabilities

    poor corrosion resistance in reactor fluids

    compatibility issues with the cladding materials

    Irradiation stability and corrosion resistance can, however, be improved through alloying.

    Uranium makes up about 4 ppm of the Earths crust (making it more common than elements

    such as silver and mercury) and the amount of economically recoverable uranium in the world

    has been estimated to be about 5.5 million tonnes.

    Natural uranium has 0.7% and 99.3% (and 0.006% ). It is found in a variety of

    minerals, such as pitchblende (U3O8), and uraninite (UO2). The main producers of uranium are

    Kazakhstan (27%), Canada (20%) and Australia (20%), although a sizeable portion of uranium

    is also produced by reprocessing spent fuel rods.

    3.2.1.1 Structure

    Uranium has three crystalline polymorphs:

    is orthorhombic (at RT: a = 2.852 , b = 5.865 , c = 4.945 )

    is tetragonal (at 720C: a = 10.790 , c = 5.656 )

    is cubic (bcc) (at 850C: a = 3.538 )

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 31 -

    3.2.1.2 Thermal expansion and thermal cycling growth

    In single-crystal form, -U is strongly anisotropic. The temperature dependence of its lattice

    parameters is shown below. Note that the coefficient of thermal expansion (CTE) along the a

    and c axes is positive, whereas that along the b axis is negative (i.e. as temperature increases, it

    expands along the a and c directions but shrinks in the b direction).

    T-dependence of the lattice parameters of: (1) -U; (2) -U-15 at.% Pu (from Frost, 1994)

    In polycrystalline samples, the anisotropic CTE causes problems when the temperature is

    changed, since neighbouring grains (if unconstrained) would change shape in different ways on

    heating or cooling. In a polycrystal, grains therefore exert stresses on each other and local

    plastic flow, dislocation multiplication and hardening (and even failure of the material) can

    occur.

    If a polycrystal has grains in random orientations, no net shape change would be expected on

    thermal cycling. However, rolled -U shows strong crystallographic texture (preferred grain

    orientation), and textured polycrystals show thermal cycling growth. Note that growth in

    this context means shape change at constant volume. The growth arises from relative

    movement between neighbouring grains in different orientations combined with stress

    relaxation in some of the grains by plastic deformation or creep.

    The images below show the effect of thermal cycling growth in highly oriented fine-grained

    -uranium between 50C and 500C for 1300 cycles (top) and 3000 cycles (bottom).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 32 -

    Effect of thermal cycling in uranium (from Murty & Charity, 2013)

    Due to its greater symmetry, the gamma phase of uranium does not exhibit thermal cycling

    growth, so gamma stabilising alloying additions such as Al, Mo and Mg can help avoid this

    effect, as shown below. U-Mo alloys typically contain at least 6 wt% Mo in order to avoid

    thermal cycling growth.

    Thermal-cycling growth in uranium alloys (from Ma, 1983)

    3.2.1.3 Thermal conductivity

    A high thermal conductivity is required to allow heat to be removed from the fuel through the

    cladding, to the coolant. The linear power rating of a fuel element is generally limited by the

    thermal conductivity of the fuel, to avoid it melting. The figure below shows thermal

    conductivity of annealed high-purity polycrystalline uranium as a function of temperature and

    it can be seen that thermal conductivity increases as temperature increases (in practice, this is

    limited by factors such as purity of the material).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 33 -

    Thermal conductivity of annealed uranium (from Murty & Charit, 2013)

    3.2.1.4 Mechanical properties

    Uranium is a relatively ductile metal and is therefore easy to work. A variety of fabrication

    techniques can be used to process uranium, including: rolling, forging, casting, extrusion,

    drawing, machining and powder metallurgy.

    A typical stress-strain curve for uranium is shown below.

    A typical stress-strain curve for uranium (from Ma, 1983)

    It is worth noting that the mechanical properties depend on texture, fabrication history and heat

    treatment. The tensile properties are also affected by impurities such as carbon, fission

    products or alloying elements, and the strength also decreases dramatically with increasing

    temperature.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 34 -

    3.2.1.5 Corrosion

    Uranium reacts rapidly with most environments (air, oxygen, hydrogen, water etc). The UO2

    surface layer that forms is not quite protective: at high temperatures, as the film thickens, it

    cracks and crumbles, exposing fresh uranium underneath. Similarly, in water, the UO2 film

    provides reasonable corrosion resistance at low temperatures (50-70C), but at higher

    temperatures, the oxide becomes porous and the protection is lost.

    Irradiation enhances corrosion, but this will be discussed in more detail in section 7.

    3.2.1.6 Irradiation growth

    Irradiation growth is a form of dimensional instability that occurs under irradiation without the

    need for an applied stress, at relatively low temperatures (~300C). As in thermal cycling

    growth, the volume of the material remains constant during irradiation growth (essentially,

    material is moved from one place to another), so this is different to radiation swelling (see

    section 3.2.1.7).

    Under a neutron flux, even single crystals of -U show irradiation growth (although note that

    -U does not, due to its isotropic nature). The growth is accompanied by hardening and

    embrittlement, and arises from the generation of dislocation loops:

    interstitial loops form on (010)

    vacancy loops form on {110}

    Interstitial and vacancy loops form on different planes because of the thermal spike and the

    anisotropic CTE.

    Consider the effect of heating an individual grain relative to its surroundings: it experiences

    tension parallel to [010] (because the CTE in that direction is negative). Interstitial atoms will

    then preferentially condense on (010), thereby giving an expansion to relieve the tensile stress

    (analagous effects occur for vacancy loops on other planes). There is therefore a net expansion

    parallel to [010].

    If the material is polycrystalline, it will not show irradiation growth if there is no

    crystallographic texture (individual grains would change shape but there would be no net shape

    change), but irradiation growth will occur if it has preferred orientation.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 35 -

    Length changes for a uranium single crystal (from Frost, 1994)

    3.2.1.7 Irradiation swelling

    Irradiation swelling involves an increase in volume of a material as a result of irradiation.

    There are two regimes of irradiation swelling:

    At lower temperatures, swelling occurs because anisotropic irradiation growth causes

    internal stresses, which can lead to cavitation (see section 2.4) and facilitates the

    development of voids and therefore swelling.

    At higher temperatures, fission gas bubbles (primarily and heavy, inert

    gases) form in the phase.

    Swelling rate of uranium phases in the burn-up range 0.20.5 at.% (from Frost (1994))

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 36 -

    As illustrated below, voids in -U have a characteristic faceted shape, whilst those in cubic -U

    are more rounded (due to the more isotropic nature of -U).

    Pores in irradiated -U10Zr (wt%) (left) and in -U10Zr (right) (from Frost (1994))

    As shown in the image below, the volume increases can be extreme!

    Irradiation swelling of uranium and uranium alloys (from Ma (1983))

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 37 -

    3.2.1.8 Irradiation creep

    Creep rates can be accelerated by a factor of 10100 under irradiation, and the problem is

    exacerbated by swelling. There are two slightly different types: radiation-induced creep and

    radiation-enhanced creep, but both arise primarily because of vacancies.

    Radiation-induced creep occurs at lower homologous temperatures than thermal creep. At

    these lower temperatures, the vacancy concentration produced by atomic displacements due to

    irradiation could be large enough to induce creep deformation under an applied stress. The

    creep rate is proportional to the stress and the neutron flux.

    Radiation-enhanced creep occurs at higher temperatures, at which thermal creep can also

    occur. The addition of extra vacancies augments the vacancy concentration and enhances the

    creep rate.

    Note that irradiation affects both primary and secondary creep. In primary creep, the strain rate

    is relatively high, but slows with increasing time - this is due to work hardening. The strain rate

    becomes constant due to the balance between work hardening and annealing. This stage is

    known as secondary or steady-state creep. Stress dependence of this rate depends on the creep

    mechanism.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 38 -

    3.2.2 Plutonium

    Plutonium can also be used as a nuclear fuel in nuclear reactors and in space applications.

    is the major fissile isotope of plutonium and has a high fission cross-section. Plutonium

    is only found naturally in trace quantities and is mainly produced artificially through the

    transmutation of .

    It can be recovered from spent fuel in thermal reactors, and depleted uranium can be kept

    together with plutonium for fuels used in fast breeder reactors. Separated plutonium can also be

    used in plutonium-burning reactors.

    There are 6 phases of plutonium, which are only stable in limited temperature ranges. It is

    monoclinic at room temperature, but also has orthorhombic, fcc, bct and bcc phases, and it also

    readily undergoes martensitic transformations. It has a relatively low melting point of 640C.

    Plutonium has a lower critical mass than uranium, and is also toxic and pyrophoric as well as

    being very sensitive to corrosion. It is therefore difficult to work with, but a variety of

    fabrication techniques can be used. Its mechanical properties depend on its phase (some phases

    are very brittle, but others are relatively ductile), but also on impurity and defect concentrations.

    Its properties therefore do not allow it to be used in pure form, so a variety of alloying

    additions (Al, Ga, Mo, Th, Zr etc) are added. Zr is often added to increase the melting

    temperature and to reduce interdiffusion with the stainless steel cladding. However, the

    temperature gradient in alloy fuel rods leads to composition variation, driven by the Soret

    effect (different atoms exhibit different responses to a temperature gradient).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 39 -

    3.2.3 Thorium

    Thorium is another nuclear fuel that has not yet been used to its full potential. is a fertile

    isotope that could produce fissile upon capturing a neutron, so is therefore an important

    breeder material. is the only naturally-occurring isotope and has a half-life of 14 billion

    years (!).

    Thorium is far more abundant than uranium in nature: most rocks and sands contain minute

    amounts of thorium, and monazite is a rare earth phosphate mineral, containing 67 wt%

    thorium.

    Thorium has two phases, both of which are cubic, and it has a melting point of ~1750C. Its

    mechanical properties are sensitive to impurities, cold work and crystallographic texture, and it

    is sensitive to corrosion (although much less so than plutonium). Its mechanical properties and

    corrosion resistance can be improved by alloying.

    The commercial use of thorium has a number of obstacles to overcome: a large amount of

    R&D and testing are still needed and the costs of fuel fabrication and reprocessing are

    high. also becomes contaminated with , which decays to daughter nuclides that are

    high-energy gamma-emitters (and therefore difficult to handle), such as .

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 40 -

    3.3 Ceramic fuels

    The essential requirements remain that there should be a sufficient number density of fissile

    atoms (and we would like to avoid high levels of U enrichment), and that the other nuclides

    (non-fissile components) should have a low mass number and small neutron absorption

    coefficient.

    Potential advantages of ceramics:

    good irradiation stability (no phase transitions)

    higher fuel and plant operating temperature (higher melting temperature than metals)

    excellent corrosion resistance

    low thermal expansion coefficients

    Disadvantages of ceramics:

    brittle, low fracture strength

    poor thermal conductivity, as illustrated below (especially UO2)

    poor heat transfer to cladding (no metallurgical bond)

    Thermal conductivities of major nuclear fuels (from Frost, 1994)

    The three main ceramic nuclear fuels are UO2, UC and UN, although UO2 is most commonly

    used (largely for historical reasons).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 41 -

    3.3.1 UO2

    UO2 is a precursor to U metal and was first used as a blanket fuel, resistant to high-temperature

    water (a blanket fuel is a layer of material containing fertile isotopes that is placed around the

    reactor core as a reflector or absorber, but which is also used to breed additional fissionable

    material). However, it is now very well studied, and has been in reliable use as a nuclear fuel

    for nearly 50 years.

    In UO2 fuel, either natural or slightly enriched uranium (0.7 4% ) can be used. If enriched

    uranium is used, it is good practice to vary the enrichment across the core, as illustrated below.

    This is because irradiation is usually non-uniform and tends to be higher in the centre of a fuel

    rod, so having a varying enrichment can allow burn-up to occur more evenly.

    Plan of the fuel sub-assembly and reactor core for a PWR (from Frost, 1994)

    Note that Fast Breeder Reactors use a mixed oxide, MOX: (U0.75Pu0.25)O2. The oxides UO2 and

    PuO2 show complete mutual solubility:

    Solidus and liquidus lines in the UO2-PuO2 equilibrium phase diagram (from Frost, 1994)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 42 -

    A high density is desirable to ensure a high density of fissile or fissionable atoms, and good

    thermal conductivity. However, some residual porosity is useful to retain fission product gases

    and to allow densification on heating to offset irradiation swelling.

    UO2 can be processed by conventional ceramic powder sintering into bulk shapes such as

    pellets, rods and tubes with densities typically 9397% of the theoretical density (a lower

    density of 85% is used for Fast Breeder Reactors, to allow greater burn-up). Pellets are usually

    used (see image below), since it is difficult to manufacture a whole fuel rod from a ceramic.

    Sintering must be performed in an inert (or reducing) atmosphere, because sintering in air can

    lead to the formation of other uranium oxides such as U3O8 the different densities of the

    various oxide phases causes problems during sintering.

    3.3.1.1 Structure

    UO2 has the CaF2 structure (fcc, a = 5.47 ): the calcium ions occupy the face-centred lattice

    sites whilst the oxygen anions sit in tetrahedral interstices. The large octahedral interstice in the

    centre of the unit cell is empty and can therefore accommodate fission products. This is

    important for radiation stability (since it prevents the products from diffusing elsewhere and

    forming voids).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 43 -

    UO2 shows no polymorphic phase changes on heating and its melting temperature (2865C) is

    much higher than that of uranium metal. UO2 does restructure by grain growth and void

    migration, especially above 1700C.

    It has a low tensile strength of approximately 35 MPa and a Youngs modulus of

    approximately 170 GPa. It rapidly loses strength and becomes ductile above 11001400C (the

    transition temperature depending on strain rate and grain size), but exhibits cracking on reactor

    start-up and shut-down. This cracking occurs primarily due to thermal stresses rather than due

    to irradiation effects.

    The image below shows crack distribution in a fuel pellet, with a superposed strength vs

    temperature curve for (U, Pu)O2. The pellet is ductile in the centre, where the temperature is

    highest, but becomes brittle closer to the edges, where the temperature is lower.

    3.3.1.2 Irradiation effects

    Upon irradiation, gas release can occur:

    Xe and Kr (insoluble fission products)

    Cs, I, Br, Te (volatile fission products)

    The amount of gas released depends on a variety of factors, including porosity, irradiation time

    and irradiation temperature. The centre temperature of the rod in thermal reactors is typically

    restricted to ~1400C (note that in FBRs the temperature may initially be higher to promote gas

    release.)

    Irradiation swelling of UO2 is low, as illustrated below.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 44 -

    Irradiation swelling: (a) in U metal in a Magnox reactor; (b) in UO2 in a PWR (from Frost, 1994)

    Failure of oxide fuel rods occurs mainly due to pellet-cladding mechanical interaction (PCI).

    This can be alleviated by:

    leaving sufficient space between the fuel and cladding (but note that this can affect the

    efficiency of heat transfer)

    using graphite as a liner to provide lubrication

    using smaller diameter rods at lower LHGR

    On increasing the LHGR, expansion and cracking of UO2 fuel pellets

    can strain the cladding, as illustrated here (from Frost, 1994).

    UO2 can deviate significantly from stoichiometry. This is an important issue affecting

    properties and possible burn-up. As fuel is burned, the oxygen:metal ratio increases (because

    not quite all the oxygen released can combine with fission products). It is good practice to start

    with hypostoichiometric fuel (where oxygen:metal 1.97) as oxygen release is detrimental

    for the cladding.

    Thermal conductivity of perfectly dense (U0.8Pu0.2)O2- (from Frost, 1994)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 45 -

    In the ceramic UO2, transport (i.e. diffusion) rates are low (compared to those in metallic U),

    but the movement of grain boundaries during grain growth sweeps gas from grain interiors to

    grain boundaries, where the bubbles then coarsen by diffusion of gas atoms along the grain

    boundaries, as shown below.

    Gas bubbles on grain boundaries in UO2 tested in an AGR (from Frost, 1994)

    Bubbles start from residual porosity in the sintered pellets and then migrate up the temperature

    gradient in the fuel rod. UO3 is the main volatile species.

    The migration of bubbles up the temperature gradient

    generates a void along the central axis of the fuel pellet.

    This image shows thermal restructuring in a 7.5 mm

    diameter 94% dense fuel pellet of (U, Pu)O2, tested in a

    FBR at a LHGR of 45 kW/m and a burn-up of 40 GWd/t.

    Grain growth, cracking and a central void can be seen (from

    Frost, 1994).

    In MOX fuel rods, similar processes transport U from the centre to the surface, leaving the

    centre Pu-rich.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 46 -

    3.3.2 Carbide and nitride fuels

    There are many carbides and nitrides of U (UC, U2C3, U2C, UN, U2N3, UN3), but those of

    interest for use as fuels are UC and UN, which show complete mutual solubility.

    UC and UN both have the NaCl structure:

    UC: a = 4.961 , Tm = 2780 K

    UN: a = 4.889 , Tm = 3120 K

    They are widely recognised as intrinsically better fuels for the future, in particular as advanced

    fuels for FBRs. They would permit moderate LHGR (~70 kW/m), high burn-up (>150 GWd/t),

    and efficient operation as they have a higher density of metallic atoms than UO2.

    UC has been used in thermal and FBRs, and has been more extensively studied than UN, but

    UN is of particular interest for easier reprocessing (important for the FBR fuel cycle).

    The carbide and nitride fuels show similar irradiation effects to UO2, but they have greater

    stability and better thermal conductivity, thus permitting higher LHGR and greater burn-up.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 47 -

    4. Cladding

    4.1 Introduction

    The cladding must:

    isolate the fuel from the coolant (this requires chemical stability and mechanical

    integrity)

    have a low neutron absorption cross-section (to allow neutrons to pass through the

    cladding without being absorbed)

    be corrosion-resistant, and in general be compatible with the fuel and coolant

    conduct heat well (to allow heat to be transferred from the fuel, through the cladding, to

    the coolant: heat transfer from fuel to cladding should also be considered, as well as

    heat transfer through the cladding)

    have a melting point well above the operating temperature of the reactor (note that new

    designs of reactor may well have higher operating temperatures)

    have adequate strength and ductility, especially to withstand swelling of the fuel

    have high stability under irradiation

    have a low induced radioactivity (to facilitate recycling after use)

    The cladding is one of the most critical components in a reactor, and its performance

    (especially irradiation growth and corrosion resistance) is often the key factor limiting

    attainable burn-up.

    The most commonly used cladding materials are stainless steels (for more extreme conditions

    in FBRs) and zirconium alloys (in thermal reactors).

    4.2 Austenitic stainless steels

    Austenitic (or ) stainless steels have an fcc structure. In low carbon steels, austenite is only

    stable above the eutectoid temperature (730C), but different alloying additions can stabilise

    austenite to lower temperatures.

    Austenitic stainless steels are the cladding materials of choice in FBRs where operating

    conditions are particularly severe:

    irradiation can give as much as 100 dpa in total

    liquid-metal temperatures: 250C to 700C

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 48 -

    The most commonly used austenitic stainless steels are 304 and 316, typical compositions of

    which are tabulated below.

    Fe C Cr Ni Mo Mn Si P S

    304 balance

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 49 -

    Due to their small size, helium atoms can easily diffuse interstitially and form bubbles.

    However, voids that form under irradiation are much larger than can be explained only by the

    production of helium: as described in section 2, helium bubbles act as sinks for vacancies and

    therefore act as nucleation sites for voids.

    Austenitic stainless steels show good dimensional stability under irradiation, until the onset of

    void swelling. The figure below schematically shows void swelling behaviour as a function of

    radiation dose, illustrating the different stages of void swelling. The initial incubation period

    represents the neutron dose required to produce sufficient helium to make void nucleation

    possible: this incubation period depends on a variety of factors, including temperature,

    composition and microstructure.

    Void swelling behaviour (from Murty & Charit, 2013)

    After the incubation and transient period (which is highly variable and depends on many

    factors linked to the distribution of solute elements), the swelling enters steady state. The

    steady state swelling rate in austenitic stainless steels is typically ~1% per dpa (in comparison,

    the rate in ferritic steels is 0.1% per dpa), although the rate is lower at both high and low

    temperatures. This is illustrated in the figure below, which schematically shows the three

    regimes of steady-state swelling rate in 316 stainless steel.

    (from Frost, 1994)

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 50 -

    In region I (at low temperatures) direct recombination of vacancies and interstitials can occur

    (due to lower mobilities at lower temperatures): the swelling rate is therefore low but increases

    with increasing temperature (as mobility increases).

    In region II, the steady state swelling rate is high, and can only be explained by an excess of

    vacancies.

    In region III (at higher temperatures), the swelling rate decreases with increasing temperature,

    since the rate of vacancy emission from voids is greater than the flux of vacancies joining the

    voids.

    Under neutron irradiation the maximum total swelling reported for an austenitic stainless steel

    is 88% (at 510C), although the largest reported swelling is 260% for irradiation by 140 keV

    protons at 625C!

    The images below show void swelling at different temperatures in 316 stainless steel at

    ~1.41027

    neutrons/m2 (En > 0.1 MeV), equivalent to ~70 dpa (from Frost, 1994).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 51 -

    The image below shows swelling (~10% linear, 33% by volume) of a 316 stainless steel

    cladding tube (not containing fuel) irradiated at 1.51027

    neutrons/m2 (En > 0.1 MeV),

    equivalent to ~75 dpa, at 510C. Note that all relative proportions are preserved during

    swelling (from Frost, 1994).

    The swelling of the cladding can be quite variable, as illustrated in the left-hand image (a)

    below. This shows the top of a bundle of fuel rods irradiated to 2.1x1027

    n/m2 (E > 0.1 MeV):

    the rods vary in length as a function of gradients in dose, non-uniformity of composition and

    variations in crystallographic texture (as a result of processing).

    In comparison, the right-hand image (b) shows an undistorted fuel-rod assembly manufactured

    using a non-swelling cladding (see below) that has been irradiated at 1.9x1027

    n/m2 (E > 0.1

    MeV) (from Frost, 1994).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 52 -

    4.2.1.1 Minimising void swelling

    The transient period before the onset of steady state swelling can be prolonged by balanced

    solute additions and appropriate thermomechanical treatments.

    P and Si are particularly useful additions, since these form precipitates, the surfaces of which

    act as vacancy sinks, which reduces the vacancy supersaturation.

    This effect is illustrated below and shows that the degree of cold work and solute additions can

    change the amount of swelling in 316 stainless steel irradiated at 500C (from Frost, 1994).

    Swelling is accelerated by stress, as illustrated

    in this graph, which shows the effect of stress

    on the swelling of two modified 316 stainless

    steels, irradiated as pressurised tubes (from

    Frost, 1994).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 53 -

    4.2.2 Inverse Kirkendall effect

    The Kirkendall effect (discovered in 1947) is the motion of

    the boundary layer between two metals, that occurs as a

    result of the difference in diffusion rates of the metal atoms

    (see IB course A).

    In the Kirkendall experiment two blocks of metal (A and B)

    are joined together with inert markers (fine molybdenum

    wires) at the interface to form a diffusion couple.

    If the diffusivity of A atoms is higher than the diffusivity of

    B atoms, there is a net flux of atoms to the right while the

    inert markers remain stationary.

    The situation can also be viewed in terms of vacancies. The equilibrium number of vacancies

    in A is low and in B it is high. The vacancy flux is therefore higher from B to A than A to B.

    The inverse Kirkendall effect arises when a flux of vacancies drives an opposite flux of more

    mobile atoms. Voids act as sinks for vacancies, so the flux of vacancies towards voids is in fact

    a flux of atoms away from voids. Faster-diffusing species exchange places more often with the

    vacancies than slow-diffusing species. The fast-diffusing solutes are therefore depleted at sinks

    while the concentrations of the slow-diffusing species increase.

    In austenitic stainless steels, different elements have different diffusivities:

    The more mobile Cr and Fe atoms diffuse away from voids to such an extent that the matrix

    away from the voids converts to phase ferrite (bcc) and only the regions close to the voids

    retain the original austenite phase. This occurs because nickel (which is itself fcc) is an

    austenite stabiliser whereas chromium (which is bcc) is a ferrite stabiliser.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 54 -

    4.2.3 Dislocation densities

    Typical dislocation densities in austenitic stainless steels are given below:

    Annealed: 1012 m-2

    Cold-worked: 1015 to 1016 m-2

    Irradiated: (6 3) x 1014 m-2 (at steady state)

    The figure below shows dislocation density as a function of irradiation dose in annealed and

    cold-worked austenitic stainless steels, irradiated at 500C (from Frost, 1994).

    The irradiation drives the dislocation density to the same steady-state value, independent of the

    starting value.

    This steady-state arises due to a balance between irradiation causing an increase in dislocation

    density (as discussed in section 2), whilst also causing an increase in mobility (due to heating

    of the material and therefore annealing), which enables dislocations to be removed and

    therefore decreases the dislocation density: at steady-state, dislocations are generated at the

    same rate as they are removed, meaning that the dislocation density reaches a constant value.

    In an annealed sample, the initial dislocation density is low, so irradiation leads to an overall

    increase in dislocation density. However, in a cold-worked sample, the initial dislocation

    density is already high, so the effect of increased mobility means that dislocations are removed

    faster than they are created (until steady state is reached), so that there is an overall decrease in

    dislocation density.

    The absolute value of the steady-state depends on a variety of factors including temperature

    and neutron energy. The higher the temperature, the more effective the annealing-out of

    damage, so the lower the steady state dislocation density.

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 55 -

    4.2.4 Creep

    Irradiation accelerates creep by orders of

    magnitude, as seen in this graph, which shows the

    influence of neutron irradiation on creep rate of an

    annealed steel (from Frost, 1994). Components can

    be readily designed to avoid normal thermal creep,

    so irradiation creep is the main issue. There is a

    complex interaction of displacement rate and

    temperature, and it is difficult to do in-reactor

    experiments.

    Here, the effect of irradiation on creep can be seen.

    The graph shows results from uniaxial tensile tests

    of 20% cold-worked 316 stainless steel during

    thermal ageing and neutron irradiation (from Frost,

    1994).

    Additionally, creep is accelerated by swelling.

    This graph shows the correlation of irradiation

    creep coefficients with swelling rates for annealed

    304L stainless steel (from Frost, 1994).

  • Part III Materials Science M17: Nuclear Materials Lent 2014

    - 56 -

    4.2.5 Mechanical properties

    Irradiation induces hardening and loss of

    ductility, due to the increased dislocation

    density.

    This graph shows the effects of neutron

    irradiation to 12.5 dpa at 450C on the