Как выбрать гостиницу для кошек
14 декабря, 2021
whereas in future fusion reactors employing deuterium (D) and tritium (T) as fuel, the neutrons are the result of DT fusion. Spallation neutron sources, which are used for a variety of material research purposes, generate neutrons as a result of spallation reactions between a high-energy proton beam and a heavy metal target. Neutron exposure can lead to substantial changes in the microstructure of the materials, which are ultimately manifested as observable changes in component dimensions and changes in the material’s physical and mechanical properties as well. For example, radiation-induced void swelling can lead to density changes greater than 50% in some grades of austenitic stainless steels1 and changes in the ductile-to-brittle transition temperature greater than 200 °C have been observed in the low-alloy steels used in the fabrication of reactor pressure vessels.2,3 These phenomena, along with irradiation creep and radiation-induced solute segregation are discussed extensively in the literature4 and in more detail elsewhere in this comprehensive volume (e. g., see Chapter 1.03, Radiation-Induced Effects on Microstructure; Chapter 1.04, Effect of Radiation
on Strength and Ductility of Metals and Alloys; and Chapter 1.05, Radiation-Induced Effects on Material Properties of Ceramics (Mechanical and Dimensional)). The objective of this chapter is to describe the process of primary damage production that gives rise to macroscopic changes. This primary radiation damage event, which is referred to as an atomic displacement cascade, was first proposed by Brinkman in 1954.5,6 Many aspects of the cascade damage production discussed below were anticipated in Brinkman’s conceptual description.
In contrast to the time scale required for radiation — induced mechanical property changes, which is in the range of hours to years, the primary damage event that initiates these changes lasts only about 10~ns. Similarly, the size scale of displacement cascades, each one being on the order of a few cubic nanometers, is many orders of magnitude smaller than the large structural components that they affect. Although interest in displacement cascades was initially limited to the nuclear industry, cascade damage production has become important in the solid state processing practices of the electronics industry also.7 The cascades of interest to the electronics industry arise from the use of ion beams to fabricate, modify, or analyze materials for electronic devices. Another related application is the modification of surface layers by ion beam implantation to improve wear or corrosion resistance of materials.8 The energy and mass of the particle that initiates the cascade provide the principal differences between the nuclear and ion beam applications. Neutrons from nuclear fission and DT fusion have energies up to about 20 MeV and 14.1 MeV, respectively, while the peak neutron energy in spallation neutron sources reaches as high as the energy of the incident proton beam, ~1 GeV in modern sources.9 The neutron mass of one atomic mass unit (1 a. m.u.^1.66 X 10-27 kg) is much less than that of the mid-atomic weight metals that comprise most structural alloys. In contrast, many ion beam applications involve relatively low-energy ions, a few tens of kiloelectronvolts, and the mass of both the incident particle and the target is typically a few tens of atomic mass unit. The use of somewhat higher energy ion beams as a tool for investigating neutron irradiation effects is discussed in Chapter 1.07, Radiation Damage Using Ion Beams.
This chapter will focus on the cascade energies of relevance to nuclear energy systems and on iron, which is the primary component in most of the alloys employed in these systems. However, the description of the basic physical mechanisms of displacement cascade formation and evolution given below is generally valid for any crystalline metal and for all of the applications mentioned above. Although additional physical processes may come into play to alter the final defect state in ionic or covalent materials due to atomic charge states,10 the ballistic processes observed in metals due to displacement cascades are quite similar in these materials. This has been demonstrated in molecular dynamics (MD) simulations in a range of ceramic materials.11-15 Finally, synergistic effects due to nuclear transmutation reactions will not be addressed; the most notable of these, helium production by (n, a) reactions, is the topic of Chapter
1.6, The Effects of Helium in Irradiated Structural Alloys.
The RDT is frequently but inappropriately called ‘the rate theory.’ This is due to the misunderstanding of the role of the transition state theory (TST) or (chemical reaction) RT (see Laidler and King38 and Hanggi et at39 for reviews) in the RDT. The TST is a seminal scientific contribution of the twentieth century. It provides recipes for calculating reaction rates between individual species of the types which are ubiquitous in chemistry and physics. It made major contributions to the fields of chemical kinetics, diffusion in solids, homogeneous nucleation, and electrical transport, to name a few. TST provides a simple way of formulating reaction rates and gives a unique insight into how processes occur. It has survived considerable criticisms and after almost 75 years has not been replaced by any general treatment comparable in simplicity and accuracy. The RDT uses TST as a tool for describing reactions involving radiation — produced defects, but cannot be reduced to it. This is true for both the mean-field models discussed here, and the kinetic Monte Carlo (kMC) models that are also used to simulate radiation effects (see Chapter 1.14, Kinetic Monte Carlo Simulations of Irradiation Effects).
The use of the name RT also created an incorrect identification of the RDT with the models that emerged in the very beginning, which assumed the production of only FPs and 3D migrating PDs to be the only mobile species, that is, FP3DM. It failed to appreciate the importance of numerous contradicting experimental data and, hence, to produce significant contribution to the understanding of neutron irradiation phenomena (see Barashev and Golubov35 and Section 1.13.6). A common perception that the RDT in general is identical to the FP3DM has developed over the years. So, the powerful method was rejected because of the name of the futile model. This caused serious damage to the development of RDT during the last 15 years or so. Many research proposals that included it as an essential part, were rejected, while simulations, for example, by the kMC etc. were aimed at substituting the RDT. The simulations can, of course, be useful in obtaining information on processes on relatively small time and length scales but cannot replace the RDT in the large — scale predictions. The RDT and any of its future developments will necessarily use TST.
An important approximation used in the theory is the MFA. The idea is to replace all interactions in a
Interaction of energetic particles with a solid target is a complex process. A detailed description is beyond the scope of the present paper (Robinson41). However, the primary damage produced in collision events is the main input to the RDT and is briefly introduced here. Energetic particles create primary knock-on (or recoil) atoms (PKAs) by scattering either incident radiation (electrons, neutrons, protons) or accelerated ions. Part of the kinetic energy, EPKa, transmitted to the PKA is lost to the electron excitation. The remaining energy, called the damage energy, Td, is dissipated in elastic collisions between atoms. If the Td exceeds a threshold displacement energy, Ed, for the target material, vacancy-interstitial (or Frenkel) pairs are produced. The total number of displaced atoms is proportional to the damage energy in a model proposed by Norgett et at42 and known as the NRT standard
Epka(£)
2 Ed
The Monte Carlo method was originally developed by von Neumann, Ulam, and Metropolis to study the diffusion of neutrons in fissionable material on the Manhattan Project9,10 and was first applied to simulate radiation damage of metals more than 40 years ago by Besco,11 Doran,12 and later Heinisch and coworkers.13,14
Monte Carlo utilizes random numbers to select from probability distributions and generate atomic configurations in a stochastic process,15 rather than the deterministic manner of MD simulations. While different Monte Carlo applications are used in computational materials science, we shall focus our attention on KMC simulation as applied to the study of radiation damage.
The KMC methods used in radiation damage studies represent a subset of Monte Carlo (MC) methods that can be classified as rejection-free, in contrast with the more classical MC methods based on the Metropolis algorithm.9,10 They provide a solution to the Master Equation which describes a physical system whose evolution is governed by a known set of transition rates between possible states.16 The solution proceeds by choosing randomly among the various possible transitions and accepting them on the basis of probabilities determined from the corresponding transition rates. These probabilities are calculated for physical transition mechanisms as Boltzmann factor frequencies, and the events take place according to their probabilities leading to an evolution of the microstructure. The main ingredients of such models are thus a set of objects (which can resolve to the atomic scale as atoms or point defects) and a set of reactions or (rules) that describe the manner in which these objects undergo diffusion, emission, and reaction, and their rates of occurrence.
Many of the KMC techniques are based on the residence time algorithm (RTA) derived 50 years ago by Young and Elcock17 to model vacancy diffusion in ordered alloys. Its basic recipe involves the following: for a system in a given state, instead of making a number of unsuccessful attempts to perform a transition to reach another state, as in the case of the Metropolis algorithm,9, 0 the average time during which the system remains in its state is calculated. A transition to a different state is then performed on the basis ofthe relative weights determined among all possible transitions, which also determine the time increment associated with the selected transition. According to standard transition state theory (see for instance Eyring1 ) the frequency Гx of a thermally activated event x, such as a vacancy jump in an alloy or the jump of a void can be expressed as:
Гх =exp(-йг) [11
where nx is the attempt frequency, kB is Boltzmann’s constant, Г is the absolute temperature, and Ea is the activation energy of the jump.
During the course of a KMC simulation, the probabilities of all possible transitions are calculated and one event is chosen at each time-step by extracting a random number and comparing it to the relative probability. The associated time-step length dt and average time-step length At are given by:
= and At = 1 [2]
Гх Г x И
n n
where r is a random number between 0 and 1. The RTA is also known as the BKL (Bortz, Kalos, Lebowitz) algorithm.19 Other techniques are possible, as described by Chatterjee and Vlachos.20 The basic steps in a KMC simulation can be summarized thus:
1. Calculate the probability (rate) for a given event to occur.
2. Sum the probabilities of all events to obtain a cumulative distribution function.
3. Generate a random number to select an event from all possible events.
4. Increase the simulation time on the basis of the inverse sum of the rates of all possible events
( |
At — , where — is a random deviate that
N, R,
assures a Poisson distribution in time-steps and N and R are the number and rate of each event i.
5. Perform the selected event and all spontaneous events as a result of the event performed.
6. Repeat Steps 1-4 until the desired simulation condition is reached.
A common formalism for excess energy expressions is the Redlich-Kister-Muggianu relation, which for a binary system can be written as
Gex = X, XjX Lk, ij(X, — Xj)k [11]
k
where L is the coefficient of the expansion in k, which can also have a temperature dependence typically of the form a + bT. Thus, a regular solution is defined as k equals zero leaving a single energetic term. This approach is related to the Bragg-Williams description, with random mixing of constituents yet with enthalpic energetic terms such that
Gex = XaXbEbw [12]
Here, XAXB represents a random mixture of A and B components and is thus the probability that A-B is a nearest-neighbor pair, and EBW is the Bragg — Williams model energetic parameter.
In a relevant example, Kaye eta/.16 have generated a solution model for the five-metal white phase noted above and more extensively discussed in Chapter 2.20, Fission Product Chemistry in Oxide Fuels. A binary constituent of the model is the fcc-structure Pd-Rh system, which at elevated temperatures forms a single solid solution across the entire compositional range. The phase diagram of Figure 2 also shows a low — temperature miscibility gap, that is, two coexisting identically structured phases rich in either end member. The excess Gibbs free energy expression for
the fcc phase was determined from an optimization using tabulated thermochemical information together with the phase equilibria and yielded
Gex = XPdXRh [21247 + 2199XRh
-(2.74 — 0.56XRh)T] [13]
The knowledge of the phenomenological coefficients L, including their dependence on the chemical composition, allows the prediction of RIS phenomena. Unfortunately, in practice, it is very difficult to get such information from experimental measurements, especially for concentrated and multicomponent alloys, and for the diffusion by interstitials. As we have seen, it is also quite difficult to establish the exact relationship between the phenomenological coefficients and the atomic jump frequencies because of the complicated way in which they depend on the local atomic configurations and because correlation effects are very difficult to be fully taken into account in diffusion theories. An alternative approach to analytical diffusion equations, then, is to integrate point defect jump mechanisms, with a realistic description ofthe frequencies in the complex energetic landscape of the alloy, in atomistic-scale simulations such as mean-field equations, or Monte Carlo simulations (molecular dynamics methods are much too slow — by several orders of magnitude — for microstructure evolution governed by thermally activated migration of point defects).
Atomic-scale methods are appropriate techniques to simulate nanoscale phenomena like RIS. They are all based on an atomic jump frequency model. From this point of view, the difficulties are the same as for the modeling of other diffusive phase transformations (such as precipitation or ordering during thermal aging), complicated by the point defect formation and annihilation mechanisms and by the self-interstitial jump mechanisms, which are usually more complex than the vacancy ones.76
The parameter commonly used to correlate the damage produced by different irradiation environments is the total number of displacements per atom (dpa). Kinchin and Pease7 were the first to attempt to determine the number of displacements occurring during irradiation and a modified version of their model known as the Norgett-Robinson-Torrens (NRT) model8 is generally accepted as the international standard for quantifying the number of atomic displacements in irradiated materials.9 According to the NRT model, the number of Frenkel pairs (FPs), nNRT(T), generated by a primary knock-on atom (PKA) of energy T is given by
kEd(T )
2 Ed
where ED(T) is the damage energy (energy of the PKA less the energy lost to electron excitation), Ed is the displacement energy, that is, the energy needed to displace the struck atom from its lattice position, and к is a factor less than 1 (usually taken as 0.8). Integration of the NRT damage function over recoil spectrum and time gives the atom concentration of displacements known as the NRT displacements per atom (dpa): f(E)vNRT(T)a(E, T)dTdE where f(E) is the neutron flux and s(E, T) is the probability that a particle of energy E will impart a recoil energy Tto a struck atom. The displacement damage is accepted as a measure of the amount of change to the solid due to irradiation and is a much better measure of an irradiation effect than is the particle fluence. As shown in Figure 1, seemingly different effects of
irradiation on low temperature yield strength for the same fluence level (Figure 1 (a)) and disappear when dpa is used as the measure of damage (Figure 1(b)).
A fundamental difference between ion and neutron irradiation effects is the particle energy spectrum that arises because of the difference in the way the particles are produced. Ions are produced in accelerators and emerge in monoenergetic beams with very narrow energy widths. However, the neutron energy spectrum in a reactor extends over several orders of magnitude in energy, thus presenting a much more complicated source term for radiation damage. Figure 2 shows the considerable difference in neutron and ion energy spectra and also between neutron spectra in different reactors and at different locations within the reactor vessel.
Distance into solid (m)
Another major difference in the characteristics of ions and neutrons is their depth of penetration. As shown in Figure 3, ions lose energy quickly because of high electronic energy loss, giving rise to a spatially nonuniform energy deposition profile caused
|
|
||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|
||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||
|
|||||