Category Archives: Comprehensive nuclear materials

Role of Atomic Weight

Materials with low atomic weight, such as aluminum, exhibit more spatially diffuse displacement cascades than high atomic weight materials due to the increase in nuclear and electronic stopping power with in­creasing atomic weight. For example, the calculated average vacancy concentration in Au displacement cascades is about two to three times higher than in Al cascades for a wide range of PKA energies.57 This increased energy density and compactness in the spa­tial extent of displacement cascades can produce enhanced clustering of point defects within the ener­getic displacement cascades of high atomic weight materials. Electrical resistivity isochronal annealing studies of fission neutron-irradiated metals have con­firmed that the amount of defect recovery during Stage I annealing decreases with increasing atomic weight,79 which is an indication of enhanced SIA clustering within the displacement cascades. The importance of atomic weight on defect clustering depends on the material-specific critical energy for subcascade formation compared to the average PKA energy. For example, in the fcc noble metal series Cu, Ag, Au, the subcascade formation energy increases slightly with mass (10, 13, and 14 keV, respectively), and very little qualitative difference exists in the defect cluster accumulation behavior of these three materi — als.13,56 In general, there is not a universal relation between atomic weight and microstructural para­meters such as overall defect production,1 1 defect cluster yield,122,123 or visible defect cluster size.56

Point Defect Evolution and Vacancy Supersaturation

Voids are a consequence of a supersaturation of vacancies in the lattice and the tendency of excess vacancies to condense into higher-order defect com­plexes (either vacancy loops or voids). However, the root cause ofvoid formation is actually not the vacan­cies, but the interstitials. Each atomic displacement event during irradiation produces a pair of defects known as a Frenkel pair. The constituents of a Frenkel pair are an interstitial (i) and a vacancy (v). Interstitials are more mobile than vacancies at most temperatures (at low-to-moderate temperatures, say less than half the melting point (0.5 T^), vacancies are essentially immobile in most materials), such that i-defects freely migrate around the lattice, while v-defects either remain stationary or move much smaller distances than i-defects. Because i-defects are highly mobile, they are able to diffuse to other lattice imperfections, such as dislocations, grain boundaries, and free surfaces, where they often are readily absorbed. This situation leads to a supersatu­ration of vacancies, that is, a condition in which the bulk vacancy concentration exceeds the complemen­tary bulk interstitial concentration. This is a highly undesirable circumstance for a material exposed to displacive radiation damage conditions, because the v-defect concentration will continue to grow (unchecked) at approximately the Frenkel defect production rate, while the i-defect concentration will reach a steady-state concentration, determined by interstitial mobility and by the concentration of extended defects (extended defects presumably serve as sinks for interstitial absorption). The v-defect concentration will inevitably reach a critical stage at which the lattice can no longer support the exces­sive concentration of vacancies, at which point the v-defects will migrate locally and condense to form voids (or vacancy loops or clusters). This entire pro­cess, initiated by the supersaturation of vacancies, causes the material to undergo macroscopic swelling, and the material becomes susceptible to microcrack­ing or failure by other mechanical mechanisms. This, indeed, is the fate suffered by a-Al2O3 when exposed to a neutron (displacive) radiation environment.

It is interesting that a supersaturation of vacancies can even be established in a material devoid of extended defects, such as a high-quality single crystal or a very large-grained polycrystalline material. Single crystal a-Al2O3 (sapphire) is an example of just such a material.6 When freely migrating i-defects are unable to readily ‘find’ lattice imperfections such as grain boundaries and dislocations, they instead ‘find’ one another. Interstitials can bind to form diin­terstitials or higher-order aggregates. Eventually, a new extended defect, produced by the condensation of i-defects, becomes distinguishable as an interstitial dislocation loop (also known as an interstitial Frank loop). Once formed, such a lattice defect acts as a sink for the absorption of additional freely migrating i-defects. With this, the conditions for a supersatura­tion of vacancies and macroscopic swelling are established.

The defect situation just described can be conve­niently summarized using chemical rate equations as described in detail in Chapter 1.13, Radiation Dam­age Theory. In eqn [1], we employ a simplified pair of rate equations to show the time-dependent fate of interstitials and vacancies produced under irradiation for an imaginary single crystal of A atoms:

^ = Pi (Aa! Ai + Va)

—Ri—v (Ai + VA! AA) Jia]

—N (nucleation rate for interstitial loops)

— G(growth rate for interstitial loops)

= pv (aa! Ai + Va) ribl

—Ri—v (Ai + Va! Aa)

where Ci(t) and Cv(t) are the time-dependent concen­trations of interstitials and vacancies, respectively; Pi and Pv are the production rates of interstitials and vacancies, respectively (equal to the Frenkel pair pro­duction rate); Ri-v is the recombination rate of interstitials and vacancies (i. e., the annihilation rate of i and v point defects when they encounter one another in the matrix); N and G are the nucleation and growth rates, respectively, of interstitial loops; Aa is an A atom on an A lattice site; Ai is an interstitial A atom; and Va is a vacant A lattice site (an A vacancy). (This equation for vacancies assumes low or moderate temperatures, such that vacancies are effectively immobile. Under high-temperature irradiation conditions, we would need to add nucleation and growth terms for voids, vacancy loops, or vacancy clusters. Reactions with preexisting defects are also ignored in eqn [1].) Note in eqn [1] that i—v recombination, Ri-v, is a harmless point defect annihilation mechanism (it restores, locally, the perfect crystal lattice). On the contrary, nucleation and growth (N and G) of interstitial loops are harmful point defect annihilation mechanisms, in the sense that these mechanisms leave behind unpaired vacancies in the lattice, thus establishing a supersaturation of vacancies, which is a necessary con­dition for swelling.

It is interesting to compare and contrast the neu­tron radiation damage behavior shown in Figure 1 of alumina (a-Al2O3) and spinel (MgAl2O4) single crys­tals, in terms of the defect evolution described in eqn [1]. Alumina must be described as a highly radiation-susceptible material, due to its tendency to succumb to radiation-induced swelling. Spinel, on the other hand, is to be considered a radiation — tolerant material, in view of its ability to resist radiation-induced swelling. According to eqn [1], we can speculate that mechanistically, nucleation and growth of interstitial dislocation loops are much more pronounced in alumina than in spinel. Also, eqn [1] suggests that harmless i—v recombination must be the most pronounced point defect annihila­tion mechanism in spinel so that a supersaturation of vacancies and concomitant swelling is avoided. Indeed, it turns out that nucleation and growth of dislocation loops are far more pronounced in alu­mina than in spinel, as discussed in detail next. The dislocation loop story described below is rich with the complexities of dislocation crystallography and dynamics. The unraveling of the mysteries of dislo­cation loop evolution in alumina versus spinel should be considered one of the greatest achievements ever in the field of radiation effects in ceramics, even though this was accomplished some 30 years ago! This story also illustrates the tremendous complexity of radiation damage behavior in ceramic materials, wherein point defects are created on both anion and cation sublattices, and where the defects generated often assume significant Coulombic charge states in highly insulating ceramics (alumina and spinel are large band gap insulators).

The earliest stages of the nucleation and growth of interstitial dislocation loops are currently impossible to interrogate experimentally. TEM has been used as a very effective technique for examining the struc­tural evolution of dislocation in irradiated solids but only after the defect clusters have grown to diameters of about 5 nm. Interestingly, important changes in dislocation character probably occur in the early stages of dislocation loop growth, when loop diameters are only between 5 and 50 nm.10 Therefore, we must speculate about the nature ofnascent dislocation loops produced under irradiation damage conditions.

Mechanical Properties of FMS After SPNI

The SPNI have produced a very large database on changes in strength and toughness properties up to «20 dpa, 1800 appm He, and 450 °C. The focus here is on two topics: (1) helium-induced hardening and (2) helium embrittlement.

1.06.4.2.1 Helium effects on tensile properties and He-induced hardening effects

The tensile properties of various FMS and AuSS after SPNI have been extensively investigated in the last decade.17,216,217,220-231 The tests were conducted at either ambient or the irradiation temperature.

The SPNI data are first compared with those for neutron irradiations of FMS, which show significant irradiation-induced hardening at <325 °C, decreas­ing hardening between 325 and 400 °C, and little or no hardening at greater than 400 °C.20 In the neutron case, the hardening measured as changes in the yield stress (Asy), initially increases with the square root of dpa but approaches a saturation level at higher doses. The saturation hardening depends on the irradia­tion temperature, and is «480 MPa at 10dpa and 200-300 °C. The uniform elongation (eu) following neutron irradiation decreases to less than 1% within several dpa. The postnecking strains are less affected, and the corresponding total elongation (et) decreases to between 3 and 10%. Figure 30 compares the SPNI yield strength increase data (Asy) with neutron data trends. Up to «10-13 dpa, the SPNI Asy are gener­ally similar to, or slightly lower than, the neutron data trends. However, the SPNI Asy do not saturate up to the maximum dose of 20 dpa, where the hard­ening reaches remarkable levels in excess of 700 MPa. The higher increment of SPNI hardening is even more pronounced at 350 °C and extends to well above 400 °C.17,231 The additional hardening above 10 dpa is primarily attributed to helium bubbles and, perhaps, with an additional contribution from the

image433

Figure 30 A comparison between the hardening induced by SPN irradiations and neutron irradiations. For neutron irradiations, the trend line ‘200 °C’ is for data irradiated and tested at <200 °C. Model curves are reproduced from Yamamoto, T.; Odette, G. R.; Kishimoto, H.; Rensman, J.-W.; Miao, P. J. Nucl. Mater. 2006, 356, 27. Reproduced from Dai, Y.; etal. J. Nucl. Mater. (2011), doi:10.1016/ j. jnucmat.2011.04.029.

higher loop density. Again note that, in some cases, the refined microstructures may be due to variable temperature history effects noted previously.

Figure 31 shows that the corresponding et of FMS after SPNI (symbols)221,222 is generally similar to, or only slightly less than, for neutron irradiations at 325 °C (solid lines)232 up to about 10-12 dpa for irradiations between 80 and 350 °C and room tem­perature tests (note that the modest differences in total elongation may be at least partly due to differ­ences in the size and geometry of the tensile speci­mens). However, at higher doses between 10 and 18 dpa and «750-1300 appm He, the et of FMS fol­lowing SPNI approaches 0 and, in some cases, the tensile specimens break during elastic loading at frac­ture stresses less than the yield stress, ay The fracture surfaces of the high-dose SPN-irradiated samples show a mixed brittle IG and transgranular cleavage fracture,227 similar to that observed in T91 and EM10 FMS after implantation at 250 °C with 2500 appm He producing 0.4 dpa.28

In contrast to the «20 dpa and 1800 appm He data reported here, high levels of hardening that can be attributed to He have generally not been observed previously, either in high-energy implanta­tion studies, at less than «500 appm He,233,234 or in low-temperature SPNI.225 Indeed, excess hardening was not observed in the LANSCE irradiations at <100 °C at He levels up to 2000 appm. Helium implantation at 200 °C235,236 and 250 °C28 indicated that significant hardening due to He occurred only at high He concentration levels above «5000 appm. All these results suggest that at less than 400 °C and «600 appm He, irradiation hardening is dominated by defect clusters and loops. Coupled with the small size of He-vacancy clusters (< 1 nm), a partial expla­nation may be found in recent molecular dynamics (MD) simulations237,238 showing that He bubbles can cause significant hardening but that their contribu­tions are reduced if they are overpressurized.

The data in Figure 32 more directly show the hardening contributions of He in STIP irradiated alloys that were annealed at 600 ° C for 1 h to remove the defect clusters and loops, while leaving the more stable bubbles unaffected.239 The residual bubble hardening is significant and increases with the square root of dpa and He. Note that the latter pro­vides a crude measure of the volume fraction of bubbles. The data in Figure 32 were combined with TEM measurements of 4> and Nb and were used to evaluate the dislocation obstacle strength (a) of the «1 nm bubbles, based on the relation Acy « 3ADPH

1 | 1 | 1 | ‘

і 1 STIP data

25

— WRT

Ti: 80-350 °C

F82H

9

• Optimax-A

20

F

* F82H EBW

T91, EM10

♦ T91

Ti = 325 °C

* EM10

Eurofer, JLF-1

• Optifer-V

15

9Cr2WVTa

◄ HT9

У = 325 °C

► EP823

© EM10

I

© T91

10

* . ^ ——————————————————-

5

** .V ^

©

©

0

=_____ ,_____ і_____ ,____ I______ . ► ,e©^__________________

_

‘ c_____ і______ I_____ і____ з______ і w____ i_ x.___ і ______ і_____ □

0 5 10 15 20 25

Irradiation dose (dpa)

Figure 31 Dose dependence of total elongation of FMS irradiated in STIP at <350 °C and tested at RT. Trend lines for fission reactor irradiations are shown for comparison. Reproduced from Dai, Y.; et al. J. Nucl. Mater. (2011), 415, 306.

image434

Figure 32 The dose and CHe dependences of the hardening of the F82H and Optimax-A irradiated in STIP-I in the as-irradiated and annealed at 600°C/2h conditions. Reproduced from Peng, L.; Dai, Y. J. Nucl. Mater. (2011), doi:10.1016/ j. jnucmat.2010.12.208.

 

sy « 3ADPH « 3aGb^Nbdb, yielding an estimated a « 0.1. Note that this value of a is much lower than estimates based on MD simulations discussed in Section 1.06.5. This difference may be because strength superposition effects for combining bubble obstacles with preexisting strengthening features in the FMS were not accounted for in this evaluation. Strength superposition effects may also help rationalize the smal­ler hardening from bubbles below about 500appm He. Combining estimates of a > 0.1 with the TEM data
discussed in the previous section suggests that signif­icant hardening by bubbles (and voids) will extend to temperatures up to 500 °C at high He levels.

Vacancy Migration

The atomistic process of vacancy migration consists of one atom next to the vacant site jumping into this site and leaving behind another vacant site. The jump is thermally activated, and transition state theory predicts a diffusion coefficient for vacancy migration in cubic crystals of the form

Dv = Wvd02 exp(sm)exp(—EV/kgT)

= DVexp(—EV1 / kg T) [13]

Here, vLV is an average frequency for lattice vibrations, d0 is the nearest neighbor distance between atoms, S^ is the vacancy migration entropy, and E^ is the energy for vacancy migration. It is in fact the energy of an activation barrier that the jumping atom must over­come, and when it temporarily occupies a position at the height of this barrier, the atomic configuration is referred to as the saddle point of the vacancy. It will be considered in greater detail momentarily.

Values obtained for Щ from experimental mea­surements are shown in Figure 8 as a function of the melting point. While we notice again a trend similar to that for the vacancy formation energy, we find that Em for fcc and bcc metals apparently follow different correlations. However, the correlation for bcc metals is rather poor, and it indicates that E^ may be related to fundamental properties of the metals other than the melting point.

The saddle point configuration of the vacancy involves not just the displacement of the jumping atom but also the coordinated motion of other atoms that are nearest neighbors of the vacancy and of the jumping atom. These nearest neighbor atoms

Table 3

Vacancy relaxation volumes for metals

Metal

go(Jm 2)

m (GPa)

n

HV (eV)

V Vі/V (model)

V Vі1 /V (experiment)

Ag

1.19

33.38

0.354

1.11 ± 0.05

—0.247 ± 0.005

Al

1.1

26.18

0.347

0.67 ± 0.03

—0.311 ± 0.003

—0.05, —0.38

Au

1.45

31.18

0.412

0.93 ± 0.04

—0.262 ± 0.003

—0.15 to —0.5

Cu

1.71

54.7

0.324

1.28 ± 0.05

—0.259 ± 0.005

—0.25

Ni

2.28

94.6

0.276

1.79 ± 0.05

—0.236 ± 0.004

—0.2

Pb

0.57

10.38

0.387

0.58 ± 0.04

—0.282 ± 0.005

Pd

1.91

53.02

0.374

1.7, 1.85

—0.239, —0.225

Pt

2.40

65.1

0.393

1.35 ± 0.05

—0.260 ± 0.003

—0.24, —0.42

Cr

2.23

117.0

0.209

2.0 ± 0.3

—0.218 ± 0.02

a-Fe

2.31

90.4

0.278

1.4, 1.89

—0.278, —0.245

—0.05

Mo

2.77

125.8

0.293

3.2 ± 0.09

—0.191 ± 0.004

—0.1

Nb

2.54

39.6

0.397

2.6, 3.07

—0.284, —0.258

Ta

2.76

89.9

0.324

2.2, 3.1

—0.264, —0.228

V

2.51

47.9

0.361

2.2 ± 0.4

—0.298 ± 0.028

W

3.09

160.2

0.280

3.1, 4.1

—0.201, —0.161

lie at the corners of a rectangular plane as shown in
Figure 9. As the jumping atom crosses this plane,
they are displaced such as to open the channel. This
coordinated motion can be viewed as a particular
strain fluctuation and described in terms of phonon
excitations. In this manner, Flynn11 has derived the
following formula for the energy of vacancy migra-
tion in cubic crystals.

Em = ______ 15CH C44(C11 ~ C12 VX 14i

V 2[C11(C11 — C12) + C44(5C11 — 3 C12)]

Here, a is the lattice parameter, Ch1, C12, and C44 are
elastic moduli, and X is an empirical parameter that
characterizes the shape of the activation barrier and
can be determined by comparing experimental
vacancy migration energies with values predicted by
eqn [14]. Ehrhart eta/.7,12 recommend that X = 0.022
for fcc metals and X = 0.020 for bcc metals.

2.5

 

In the derivation of Flynn,11 only the four nearest neighbor atoms are supposed to move, while all other atoms are assumed to remain in their normal lattice positions. On the other hand, Kornblit eta/.13 treat the expansion of the diffusion channel as a quasistatic elastic deformation of the entire surrounding mate­rial. The extent of the expansion is such that the opened channel is equal to the cross-section of the jumping atom, and a linear anisotropic elasticity cal­culation is carried out by a variational method to determine the energy involved in the channel expan­sion. A vacancy migration energy is obtained for fcc metals of

 

image024

[15]

 

image025
image026
image027

Figure 9 Second nearest neighbor atom (blue) jumping
through the ring of four next-nearest atoms (green) into
adjacent vacancy in a fcc structure.

 

image028image029image030

image031

("£max _ e™1). As a result, Kornblit14 assumes that the vacancy migration energy for bcc metals is given by

 

To determine the preexponential factor for self­diffusion

 

if E^ _ £yin < 2kT if EVlax _ E™11 > 2kT

 

max EV ’ max min 2EV EV ;

 

DSd — VLV*2exp( (4 + sm) /k) [21]

requires the values for the entropy 4 + and for the attempt frequency nLV. Based on theoretical esti­mates, Seeger and Mehrer5 recommend a value of 2.5 k for the former. The atomic vibration of nearest neighbor atoms to the vacancy is treated within a sinusoidal potential energy profile that has a maxi­mum height of Ef. For small-amplitude vibrations, the attempt frequency is then given by

 

m

EV —

 

[18]

 

Using the formulae of Flynn and Kornblit, we compute the vacancy migration energies and compare them with experimental values in Figure 10.

With a few exceptions, both the Flynn and the Kornblit values are in good agreement with the exper­imental results.

The self-diffusion coefficient determines the transport of atoms through the crystal under condi­tions near the thermodynamic equilibrium, and it is defined as

 

image032

Vlv

 

Dsd — D? CVq — VLV«2exp( (sV + Sm)/k) exp(_Qso/kT) — D0Dexp(_QsD/kT) [19]

where the activation energy for self-diffusion is

Qsd — 4 + Em [20]

The most accurate measurements of diffusion coef­ficients are done with a radioactive tracer isotope of the metal under investigation, and in this case one obtains values for the tracer self-diffusion coeffi­cient dTd — fDsD that involves the correlation factor f For pure elemental metals of cubic structure, f is a constant and can be determined exactly by com — putation.15 For fcc crystals, f— 0.78145, and for bcc crystals, f— 0.72149.

 

[22]

 

Vlv

 

crystals where M is the atomic mass.

In contrast, Flynn11 assumes that the atomic vibra­tions can be derived from the Debye model for which the average vibration frequency is

 

image033

2

 

• Evm, Flynn, fcc О Evm, Kornblit, fcc ■ Evm, Flynn, bcc □ Evm, Kornblit, bcc

 

■SI

>,

O)

® 1.5

ф

c

q

оз

D)

1

>s

о

c

я

о

03

>

оз

0.5

ф

 

О

г

 

Q •

 

image034

Intrinsic Disorder Reactions

A number of different point defects can form in all ceramics, but their concentration and distributions are interrelated. In the event of the production of a vacancy by the displacement of a lattice atom, this released atom can be either contained within the crystal lattice as an interstitial species (forming a Frenkel pair), or it can migrate to the surface to form part of a new crystal layer (resulting in a Schottky reaction). Figure 1 represents a Frenkel pair: both cations and anions can undergo this type of disor­der reaction, resulting in cation Frenkel and anion Frenkel pairs, respectively. In ceramic materials, both the vacancy and interstitial defects are usually charged, but the overall reaction is charge-neutral. The energy necessary for this reaction to proceed is the energy to create one vacancy by removing an ion from the crystal to infinity plus the energy to create one interstitial ion by taking an ion from infinity and placing it into the crystal. The implication of remov­ing and taking ions from infinity implies that the two species are infinitely separated in the crystal (unlike the two species shown in Figure 1). As separated species, these are defects at infinite dilution, a

Подпись: Antisite pair well-defined thermodynamic limit. As the two defects are charged, they will interact if not infinitely sepa­rated, a point we will return to later.

As with the Frenkel reaction, the Schottky disorder reaction must be charge-neutral. Here, only vacancies are created, but in a stoichiometric ratio. Thus, for a material of stoichiometry AB, one A vacancy and one B vacancy are created. The displaced ions are removed to create a new piece of lattice. It is impor­tant to realize that we are dealing with an equilibrium process for the whole crystal. Thus, as the tem­perature changes, many thousands of vacancies are created/destroyed and new material containing many thousands of ions is formed. Thus, it is not simply that one new molecule is formed, but there is also an increase in the volume of the crystal, which is why the lattice energy is part of the Schottky reaction.

The energy for the Schottky disorder reaction to proceed in an AB material is the energy to create one A-site vacancy by removing an ion from the crystal to infinity plus the energy to create one B-site vacancy by removing an ion from the crystal to infinity, plus the lattice energy associated with one unit of the AB compound. For example in Al2O3, the energy would be that associated with the sum of two Al vacancies plus three O vacancies plus the lattice energy of one Al2O3 formula unit. Again, the vacancy species are assumed to be effectively infinitely separated.

In a crystalline material with more than one type of atom, each species usually occupies its own sub­lattice. If two different species are swapped, this pro­duces an antisite pair (see Figure 2). For example, in an AB compound, one A atom is swapped with the B atom. While this would be of high energy for an AO compound, where A and O are of opposite charge (e. g., Mg2+ and O2—), in an ABO3 material where A and B may have similar or even identical positive charges, antisite energies can be

chemistry terms, a lattice

Подпись: OAПодпись:image198Подпись:Подпись: @AG'Подпись: T ,PПодпись: NnПодпись:site means a position in the crystal that an ion will usually occupy in that crystal structure), the number of ways, OA, of arranging n A-site vacancies is

N!

n!(N — n)!

As we have n B-site vacancies to distribute over N, B-lattice sites, the total number of configurations is the product of OA and OA:

N! ‘

n!(N — n)!

( N! ‘

n!(N — n)!

where N and n are large, as they are when dealing with crystals, we can invoke Stirling’s approxima­tion, which states that ln(MZ) = M ln(M) — M. Thus,

ASc = 2k[N ln(N) — (N — n)ln(N — n) — nln(n)]

Therefore,

AG=nAgf — 2kT [N ln(N) — (N—n)ln(N—n) — nln(n)]

=nAgf — 2kT

To find the equilibrium number of defects, we need to find the minimum of AG with respect to n (see Figure 3). That is

0=Agf — 2kT ln

Assuming that the number of defects is small in comparison to the number of available lattice sites, then N n N:

image204

Figure 3 Relationship of terms contributing to the defect-free energy.

n Ag Ah As

N =exp — Ш =exp — rn exp 2k

Usually, we assume that the energy associated with the change in vibrational entropy is negligible so that the concentration of defects (n/N) is dominated by the enthalpy of reaction:

M=N=“p(-2T)

However, this is not always a valid assumption and care must be taken. When defect concentrations are measured experimentally, they are presented on an Arrhenius plot of ln(concentration) versus 1/ T which yields straight lines with slopes that are pro­portional to the disorder enthalpy (see Figure 4).

Dimensional Instabilities: Irradiation Growth, Creep, and Swelling

Irradiation growth (due to anisotropic nucleation and growth of dislocation loops on different habit planes) can be of significant practical concern at intermediate temperatures in anisotropic materials such as Zr alloys, Be, BeO, Al2O3, uranium, and graphite.40’126’259-261 Anisotropic growth in individual grains in polycrys­talline materials can produce large grain boundary stresses, leading to loss of strength and grain bound­ary fracture in some materials. Figure 27 shows the large anisotropy in measured lattice parameter change in the basal and prism planes for BeO irradiated near room temperature.262 For neutron fluences above 2 x 1020 cm-2 (~0.2 dpa) with a c-axis expansion >0.5% and an a-axis expansion near 0.1%, a rapid decrease in flexural strength was observed.262,263 In materials with highly textured grains, unacceptable anisotropic growth at the mac­roscopic level can occur. One engineering solution is to use processing techniques to produce randomly aligned, small grain-sized materials.

Irradiation creep occurs in the presence ofapplied stress, due to biased absorption of point defects at cavities and along specific dislocation orientations relative to the applied stress.264 Irradiation creep produces dimensional expansion that acts in addition to normal thermal creep mechanisms and is most

image305

Fast neutron fluence, 1020 ncm-2 (E > 1 MeV)

Figure 27 Effect of fission neutron irradiation near 75 °C on the measured lattice parameter changes for BeO. Adapted from Hickman, B. S. In Studies in Radiation Effects, Series A: Physical and Chemical; Dienes, G. J., Ed. Gordon and Breach: New York, 1966; Vol. 1, pp 72-158.

prominent at temperatures from recovery Stage III up to temperatures where thermal creep deformation becomes rapid (typically above 0.5 TM). The magni­tude ofsteady-state irradiation creep is proportional to the applied stress level and dose, and consists of a creep compliance term and a void swelling term. The magnitude of typical irradiation creep compli­ance coefficients260,265,266 for fcc and bcc metals

is 0.5-1 x 10-12 Pa-1 dpa-1. The irradiation creep compliance for ferritic/martensitic steels appears to be about one-half of that for austenitic steels.109 Accelerated irradiation creep due to differential absorption of point defects at low temperatures (e. g. below recovery Stage V) or at low doses can produce creep deformation rates that are up to 10-100 times larger than the steady-state irradiation

267,268

creep rates.

Volumetric swelling from void formation occurs at temperatures above recovery Stage V in fcc and HCP materials (and above Stage III for bcc materi­als), and typically exhibits a linear increase with dose after an initial transient regime. As summarized in Figure 28 the dose-dependent swelling in fast fission reactor-irradiated austenitic stainless steel progresses

image306

Figure 28 Summary of dose-dependent swelling behavior in 20% cold-worked Type 316 austenitic stainless steel due to fast fission reactor irradiation. Reproduced from Garner, F. A.; Toloczko, M. B.; Sencer, B. H. J. Nucl. Mater. 2000, 276, 123-142.

at a swelling rate of ^1%/dpa without evidence for saturation up to swelling levels approaching 100%.1 Similar high swelling levels without evidence of sat­uration have been observed in pure copper108 and some simple bcc alloys.131 Volumetric swelling levels in structural materials in excess of ^5% are difficult to accommodate by engineering design,269 and addi­tional embrittlement mechanisms may appear in aus­tenitic stainless steel for swelling levels above 10% including void channeling and loss of ductility.270,271 Therefore, there is strong motivation to design struc­tural materials that are resistant to void swelling by introducing a high matrix density of point defect sinks or other techniques. In general, the amount of void swelling is lower in bcc materials compared to

fcc materials.50,92,109 For example, the observed void

swelling in many ferritic/martensitic steels is <2% after fission neutron damage levels of 50 dpa or higher, whereas the void swelling in simple austenitic stainless steels may be 30% or higher.109 The superior swelling resistance in ferritic/martensitic steels is largely due to a higher transient dose before onset of steady-state swelling, along with a lower steady-state swelling rate. For many HCP materials, the amount of void swelling is relatively small compared to fcc materials due to anisotropic point defect migration that tends to pro­mote defect recombination.128 However, the potential for anisotropic swelling associated with cavity formation in HCP materials may induce large stresses and poten­tial cracking at grain boundaries.263,272,273 Figure 29 shows an example of aligned cavity formation and grain boundary separation in Al2O3 following fast fis­sion reactor irradiation.272

Unfaulting of faulted Frank loops V: experimental observations

Having examined the crystallography of unfaulting reactions in alumina and spinel, it is interesting now to compare and contrast what is experimentally observed so far as dislocation loop evolution in irra­diated Al2O3 versus MgAl2O4 is concerned. First, it is observed that the 1/3 [0001] (0001) basal loops and 1/3(1010){1010} prismatic loops readily unfault under irradiation, by the reactions shown in eqns [7,8] and [9,10], respectively.6 These reactions occur when the loops reach ^50 nm diameter,10 and each reaction produces an unfaulted loop with a 1/3(1011) perfect Burgers vector. Once formed, these unfaulted loops grow without bound until they intersect other growing dislocation loops, ultimately forming a dis­location network. Such a dislocation network in neu­tron irradiated Al2O3 is shown in Figure 6.

Once the dislocation network in irradiated alu­mina is formed, it has been demonstrated that the product dislocations within the network are free to climb6 The continuous climb of network disloca­tions in Al2O3 provides unsaturable sinks for Al and O interstitials arriving in stoichiometric proportions. All the conditions for a substantial supersaturation of vacancies are now in place. Al and O interstitials are readily absorbed at network dislocations, leaving behind numerous unpaired Al and O vacancies in the lattice. These unpaired vacancies inevitably con­dense to form voids. Under these conditions, void swelling must be the anticipated radiation response of the material.

Contrast the evolution described above for alumina to the observed microstructural evolution in spinel. The predominant 1/4 (110) {110}

image388

Figure 6 Weak-beam dark-field transmission electron micrograph showing the dislocation network formed in Al2O3 following neutron irradiation at 1015 K to a fluence of 3 x 1025nm~2 (~3dpa). Reproduced from Clinard, F. W., Jr.; Hurley, G. F.; etal. J. Nucl. Mater. 1982, 108/109, 655-670.

dislocations in spinel do not unfault under most experimental conditions tested to date.6 (Kinoshita eta/.51 observed unfaulted 1/2 (110) {110} perfect loops in MgAl2O4 single crystals following neutron irradiations in the JOYO fast breeder test reactor (fast neutron fluences up to 6.5 x 1025nm~2 (equivalent to 6.5 dpa), and temperatures between 673 and 873 K). Kinoshita eta/.51 also proposed a growth pro­cess of loops in spinel as follows: 1/6 [111] (111) 1/4 [110] (111) 1/4 [110] (101) 1/4 [110] (110) 1/2 [110] (110). Notice that this sequence ends in an unfaulted, perfect interstitial loop. This final loop configuration should be a good sink for interstitials, thus promoting a supersaturation of vacancies in the lattice. However, in neutron irradiations of MgAl2O4 single crystals in the fast flux test facility (FFTF), no evidence for 1/2 [110] (110) perfect dislocations was found, for neutron fluences ranging from 2.2 x 1026 to 2.17 x 1027nm~2 (equivalent to 22-217dpa) in the temperature range 658-1023 K.51 Therefore, the proposed progression of spinel interstitial loop characteristics described above has, to date, been confirmed only under the JOYO irradiation condi­tions reported by Kinoshita et a/.51) According to Clinard eta/.:6

Persistence of the 1/4 (110) {110} stacking fault amounts to a failure of a 1 /4(112) partial dislocation to nucleate, sweep across the loop plane, and so remove the fault.

The reason for this failure is paradoxical. Apparently, stacking fault energy cannot be the reason. The stack­ing fault energy estimate for 1/4 (110) {110} stacking faults in spinel, 180 mJ m~2,18 is similar to the energy estimates for 1/3 [0001] (0001) and 1/3(10І0){10І0} stacking faults in alumina (320 and 750 mJ m~2, respectively).10 Therefore, there seems to be a rea­sonable ‘driving force’ available to favor unfaulting of 1/4 (110) {110} stacking faults in spinel. Perhaps the explanation is simply that the magnitude of the par­tial shear vector required to unfault the faulted loops is prohibitively large. In spinel, the magnitude of the unfaulting 1/4(112) vector is ~5A, compared with the 1/3 [0001] (4.32 A) and 1/3(1010) (2.74 A) unfaulting vectors in alumina.

Whatever the reason, spinel 1/4 (110) {110} stacking faults do not unfault, and this leads to void swelling resistance and impressive inherent radiation tolerance in spinel compared alumina. Hobbs and Clinard summarize the situation as follows:

The absence of void swelling ( in spinel ) can be attributed to the failure of the loops to unfault and develop into dislocation networks; they therefore remain less than perfect interstitial sinks since the energy per added interstitial never drops below the fault energy. Vacancy-interstitial recombination thus remains the dominant mode ofdefect accommo­dation, and saturating defect kinetics inevitably ensue.

Therefore, in conclusion, the significant swelling of Al2O3 alumina at high temperatures is attributable to the unfaulting of interstitial dislocation loops and the subsequent formation of dislocation networks, which serve as efficient sinks for the absorption of intersti­tial atoms. This leaves behind a supersaturation of lattice vacancies, that is, an excess of unpaired vacan­cies in the bulk of the Al2O3.

In irradiated MgAl2O4, only high-energy faulted loops are available as sinks for interstitials. Therefore, in this case, interstitial-vacancy (i-v) recombination is the dominant mechanism for defect accommoda­tion, and negligible swelling results.

Summary and Some Outstanding Issues

Developing fusion as a large-scale energy source and high-energy proton accelerator-based technologies are the primary motivations for studying He effects in structural alloys. These environments produce copious quantities of He (and H) by transmutation reactions. High levels of He, coupled with displacement (dpa) radiation damage, lead to a wide variety of property degradation phenomena over a wide range of irradia­tion temperatures, including both severe embrittle­ment and dimensional instabilities of various types.

image462

Indeed, there is growing evidence that He-dpa syner — gisms will severely limit the operating window for leading candidate FMS in fusion first wall structures.

Ultimate resolution of He effect issues in fusion applications will require a dedicated high-energy neutron source to develop an information base to construct and verify rigorous physically based pre­dictive models of the effects of the fusion environ­ment on performance-sustaining properties. These models must account for the interactions of a large number of variables, characterizing both the irra­diation service environment and alloy of interest. However, in the interim, there are a number of irradiation techniques that can be used to simulta­neously introduce high levels of He and dpa into materials. The most notable irradiation techniques include multibeam charged particle irradiations (CPI), ISHI in mixed spectrum fission reactors, and spallation neutron sources. Each of these methods has limitations, but significant progress in understanding He effects has been, and will continue to be, achieved by closely integrating all of these irradiation tools with advanced physical models.

A primary objective of such modeling, and asso­ciated experiments, is to understand and predict the transport, fate, and consequences of He, and its inter­actions with displacement damage. There is a very large historical literature on irradiation effects in AuSS, including the key role played by He. Helium is critical to void swelling and high-temperature embrittlement (HTHE), where the latter is mani­fested as severe reductions in creep rupture times and strains. Standard AuSS provide an excellent basis
for developing an understanding of these phenomena, since they are (a) sensitive to many manifestations of irradiation damage and especially He effects; and (b) generate high levels of He from two-step Ni thermal neutron reactions in mixed spectrum fission reactors (typically of the order 50appm He/dpa) and much lower, but significant, amounts of He (<1 appm He/dpa) in fast reactors. Both void formation and HTHE are characterized by a significant incubation period prior to forming growing cavities followed by rapid swelling or creep rupture.

Helium bubbles are typically the formation sites for both voids and grain boundary creep cavities. RT-based thermodynamic-kinetic models rationalize many important trends in these phenomena. In par­ticular, the critical bubble concept relates the combi­nation irradiation variables, of temperature and dpa rates, along with a number of material-defect vari­ables and parameters, to the critical size and helium content of bubbles that convert to voids, due to dislo­cation bias for SIA or grain boundary creep cavities due to stress. The critical bubble concept rationalizes and provides a basis to quantitatively model the incu­bation periods for both of these phenomena. RT can also be used to model post-incubation swelling rates and creep rupture times and strains. These models predict a number of important observations like bimodal bubble-void cavity size distributions and precipitate-associated bubbles and voids.

Critical bubble-cavity growth models highlight the critical roles played by the overall He bubble microstructures on irradiation effects. Bubbles are necessary for the formation of voids and large
numbers of creep cavities. However, high concentra­tions of bubbles, as generally associated with high He generation rates, can prolong incubation times by increasing the number of He partitioning-trapping sites in small highly subcritical bubbles; and, in the case of swelling and other manifestations of matrix radiation displacement damage. High bubble densi­ties reduce the excess vacancy supersaturations and excess defect fluxes by acting as dominant defect sinks. These insights provide important guidance to developing irradiation-tolerant alloys. Enhanced damage tolerance can be achieved by creating fine — scale and stable microstructural features that can form small, harmless bubbles that sequester high levels of He suppressing swelling and protecting GBs from HTHE and grain boundary decohesion, leading to enormous DBTT shifts. Swelling-resistant advanced stainless steels, with extended incubation times, have used alloy carbide and phosphide phases to manage He in this manner. FMS, with high disloca­tion densities and fine lath structures, are intrinsically more damage resistant than standard austenitic alloys for a variety of reasons. These reasons include low He generation rates in fission reactors. However, the rates of He generation are much higher in D-T fusion spectra, and the irradiation damage tolerance of FMS may be significantly degraded in this case.

SPNI have been the primary source of recent insight into the effects of He in structural alloys, especially mechanical property effects. Perhaps the most significant result of the SPNI studies is that high He-hardening synergisms can lead to enormous shifts in the DBTT shifts and IG fracture in FMS. The mechanisms responsible for such synergistic embrittlement, which can lead to transition tempera­ture elevations of 600 °C or more, are He-induced weakening of GBs as well as enhanced hardening at higher temperatures and higher dpa levels. Thus, there is concern that severe embrittlement and void swelling may close the window for useful application of these alloys in fusion environments.

A parallel set of activities is needed to develop predictive models of He effects. Integrated master models, based on a multiscale-multiphysics paradigm, are being developed to predict He transport, fate, and consequences in realistic alloy microstructures. These master models contain a number ofparameters and must be both mechanistically and microstructu­rally informed. First-principles electronic structure theory and atomistic simulations can provide required model parameters and mechanistic insights. These tools include DFT, embedded atom-based

MD and MS, various Monte Carlo methods, and RT. For example, recent first principles and atomistic research has provided important information on He-vacancy cluster energetics, He interactions with a range of microstructural features, He diffusion mechanisms, and rates in the matrix, along dislocations and in GBs. These models show high He-vacancy binding energies even at the smallest cluster sizes, strong interactions between He and dislocations and dislocation jogs, and both homo — and heterophase interfaces. MD methods have also been used to char­acterize phenomena such as resolutioning of He from bubbles, showing this to be a minor mechanism; and determining the dislocation interaction strength of cavities, ranging from under to overpressurized con­ditions, showing that near-equilibrium bubbles are the strongest obstacles. Recent MD studies have also sug­gested that capillary models may overpredict He gas pressures in equilibrium bubbles.

As noted above, the concepts and models described above can be used to guide the development of irradiation-tolerant alloys. The most promising class of such materials have been dubbed NFA, which contain an ultrahigh density of remarkably stable, nanometer-scale Ti-Y-O NF. The NF are now believed to be primarily a complex oxide Y2Ti2O7 pyrochlore phase. The NF provide remarkable high — temperature creep strength, so that NFA can operate above the displacement damage regime, where recov­ery processes are much faster than defect accumula­tion rates. More significantly the NF trap helium in a very high density of very fine-scale bubbles. In prin­ciple, the bubbles suppress or mitigate essentially all manifestation of radiation damage and property deg­radation. Limited proof in principle validation of the ability of NFA to manage He has been provided by ISHI irradiations that were carried out in the HFIR. For example, an irradiation of NFA MA957 to 9 dpa and 380 appm He produced a very high density of nanometer-scale bubbles on the NFs and dislocations. The same irradiation of the FMS F82H produced roughly an order of magnitude lower density of cav­ities, composed of a bimodal distribution of bubbles and larger voids. Bubbles primarily form on disloca­tions in this case. Further, limited data showed rela­tively bubble-free interfaces in MA957, in contrast to interfaces in F82H that are highly decorated with small bubbles.

The results ofthe ISHI experiments compare very favorably with a He transport and fate master model that is under development. The master model is both microstructurally informed and incorporates parameters derived from the atomistic/electronic models cited above. Preliminary RT models have also been used to extrapolate the ISHI data to predict void swelling at higher dpa and He levels. These models suggest that FMS may experience significant swelling at damage levels greater than 50 dpa, while the NFA will remain void free. Indeed the combination of the experimental and modeling results suggest that He can be transformed from a liability to an asset in NFA.

Although these conclusions represent an optimis­tic view of the status of research on measuring, mod­eling, and managing He-dpa effects in structural alloys, especially for fusion applications, there are a number of outstanding issues that require additional science-based research to resolve. Space does not permit a complete listing, but these issues and research needs can generally be classified into broad categories related to He effects per se and those related to developing, optimizing, and qualifying alloys that can manage He in a way that may provide near immunity to radiation damage. He-related issues and research needs include the following:

• Extension of experimental observations of He transport, fate, and consequences to more alloys for a wide range of microstructures and to higher damage (He and dpa) levels and higher temperatures over a range of He/dpa, using all the irradiation techniques: CPI, ISHI, and SPNI. Characterization of the management of He in NFA at temperatures up to 750 °C, or more, is particularly critical.

• Intercomparisons of accelerated high-rate CPI and lower reactor relevant rate ISHI and SPNI data and development of experimental-modeling approaches that will permit reliable predictions of He-dpa effects at very high damage levels, up to several hundred dpa, that cannot be accessed prac­tically in neutron ISHI and SPNI.

• Very detailed characterization of He and He bub­ble distributions, along with the balance of irra­diation microstructures, with particular emphasis on quantitative evaluations, like the distribution of He and bubbles on interfaces, using a suite of advanced characterization methods tailored to this application.

• Focused mechanism studies that can provide both parameters and insights into key mechanisms that govern the transport, fate, and consequences of He in various complex alloy and model systems.

• Continued development and refinement of multi­scale master models of He transport fate and con­sequences, along with the narrower first principles and atomistic studies (models and experiments) needed to better parameterize and mechanistically inform the master models. One of many examples of modeling needs is the resolution of differ­ences between continuum capillary versus atomis­tic MD-based models of helium-vacancy cluster and bubble properties.

• Developing models and basic experiments that relate the microstructural consequences of the fate of He to the degradation of performance — sustaining mechanical properties. One of many examples is the detailed model of the behavior of He on GBs and the corresponding reduction in the grain boundary fracture strength.

However, it is clear that He must be managed as well as understood. Thus, issues and research needs also include development of practical NFA or alloys with similar attributes. These include the following items:

• Identify alloy composition-synthesis designs and thermal mechanical processing paths that optimize the NF, and the balance of NFA microstructures, so as to provide a balanced suite ofoutstanding and isotropic properties.

• Resolve issues of low fracture toughness and aniso­tropic properties in extruded NFA product forms.

• Demonstrate and understand the thermal and irra­diation stability of far-from-equilibrium NF and NFA microstructures.

• Develop practical fabrication and joining methods, which preserve optimal NFA microstructures and yield defect-free components.

• Reduce costs, improve alloy homogeneity and reproducibility, and establish industrial-scale sup­ply sources.

• Qualify new alloys for nuclear service for extended lifetimes.

We believe that there is a science-based framework to resolve these and other critical issues related to the effects and management of He in practical alloy sys­tems. (see Chapter 1.03, Radiation-Induced Effects on Microstructure and Chapter 1.09, Molecular Dynamics)

Stress-Induced Anisotropic Diffusion in fcc Metals

To evaluate the diffusion tensor and the drift force, we consider here the diffusion of self-interstitials and vacancies in fcc crystals, and begin first with the ideal case of a crystal free of any stresses other than those produced by the migrating defect itself. The jump vectors for both self-interstitials and vacancies coin­cide with the n = 12 nearest neighbor locations of the defect in its stable configuration. Table 12 lists the components Xi of these jump vectors, and they are divided into three groups according to the orienta­tions of the dumbbell axes before and after the jump. These are indicated in the first row of Table 12 by the indices 1, 2, or 3 when the dumbbell axis is aligned with the xb x2, or x3 crystal coordinate direc­tion, respectively.

In the absence of stress, all saddle point energies are equal, say ES, and ES — E = Em is just the defect migration energy. It is then a simple matter to

 

image082

-Kv (r + — R)

 

[741

 

This latter function may be viewed as an analytic function of r, since the saddle point energies vary with strains that are obtained from continuum elas­ticity theory, and they are by definition analytic func­tions of r. Expanding the first term in eqn [73] into a Taylor series up to second order, and then reverting back to the real defect concentration C(r, t), one arrives at the diffusion equation

@C 82 8

87 = @785D(r)C(r — t)-a7iF(r)C(r-t) [75]

with the diffusion tensor defined as

 

Table 12 Components of the jump vectors R in units of

d0/V2.

1 $ 2

2 $ 3

3 $ 1

X!

11 —1 —1

0000

11 —1 —1

X2

1 —11 —1

11 —1 —1

0000

X3

0000

1 —11 —1

1 —11 —1

d0 is the nearest neighbor distance between atoms.

 

image083

V (r)

 

[76]

 

-R

 

Подпись: ESПодпись:Подпись: [81]Подпись: D(111) D33 Подпись: = D0 = D0Подпись:

perform the summations in the definitions of the diffusion tensor and the drift force to show that

Dij = lAod02 exp(-b£m)dy = D0(T)% [79]

and Fi = 0.

Next, let us consider the case of an applied spatially uniform stress field. According to Section 1.01.5, eqn [44], the interaction energy to linear order in the applied stress will change the energy ofthe defect in its stable configurations to

Efm = Ef + Wf = Ef — OeiSj [80]

and in its saddle point configurations to eS + WSn

Here, ef is the transformation strain tensor of the defect in its stable configuration with orientation f. We had specified this tensor for self-interstitials in Section

1.01.4 for an orientation in the [001] direction, that is, for m = 3, although the nonlinear contributions were not included. Below, it will be given with these con­tributions for self-interstitials in Cu and for the same orientation. The corresponding transformation strain tensors for the other two orientations, for m = 1, 2, can be obtained from ej by appropriate coordinate rotations.

Similarly, efn is the transformation strain tensor of the defect under consideration in its saddle point con­figuration while changing its orientation from m to n. Specifying it for one particular jump is sufficient to obtain the transformation strain tensors for all other jumps by appropriate coordinate rotations. The inspec­tion of the jump directions for a particular pair mn reveals that there are equal but opposite jump directions with the same saddle point interaction energy; the only difference for such a pair of jump directions is that the components of R and — R have equal and opposite signs. As a result, the drift force F vanishes again. However, for the diffusion tensor, the two opposite jump directions make positive and equal contribution. Suppose, the applied stress is uniaxial with the only nonvanishing component a033 = a. When the crystal is oriented such that this uniaxial stress is perpendicular to a (001) plane, then Chan et a/.35 have shown that the diffusion within the (001) plane and perpendicular to it are given by

D(001) ^ D(001) ^ D0 3 exp [/i°ef3 a] + exp(eSi + e2S2 )a/2] 11 22 2 2exp[bOef1 a] + exp[bOef3 a]

D(001 ) = D0 3exp[b°(e1S1 + e2S2 )a/2]

33 2exp[bOef1 a] + exp[bOef3 a]

respectively.

When the uniaxial stress is perpendicular to a (111) crystal plane, then the diffusion tensor in the reference frame of the stress tensor has the

components35

D(111)- D(111)

D11 = D22

3exp[bO(2e|2 + eS3)a/3] +exp[bO(2e1S1 + e3S3)a/3]
4exp[bO(e[1 + f2 + ef3)a/3]

exp[b°(2e1S1+4,)а/з] [83]

exp[b°(ef1 + e|2 + e33)a/3]

In order to obtain the transformation strains for the saddle point configurations, Chan, Averback, and Ashkenazy35 carried out molecular dynamics simula­tions of diffusion as a function of the applied stress. Fitting their results to the above equations enabled them to determine the tensors ej and ej. Their princi­pal values are listed in Table 13 for self-interstitials and vacancies in Cu.

The ratio of the diffusion coefficients in the plane perpendicular to the uniaxial stress, namely D11/D0, is shown in Figures 19 and 20 as solid curves, while

Table 13 Principal transformation strain components for self-interstitial atoms and vacancies in Cu

Є11

e22

e33

SIA, stable configuration

0.66

0.66

0.48

SIA, saddle point

0.73

0.32

0.75

Vacancy, stable configuration

-0.08

-0.08

-0.08

Vacancy, saddle point

-0.64

-0.46

1.26

Uniaxial stress (GPa)

Подпись:Подпись: @ Sjj (r) dxk image092Подпись: [84]Подпись: rПодпись:Подпись: [85]Подпись:Подпись:Подпись: e mn P1 Подпись: exp Подпись: [87]diffusion parallel to the stress, D33/D0, is shown as dashed curves. The enhancement of diffusion by tensile (positive) stresses in the (001) planes is much larger for vacancies than for self-interstitials. However, when tensile stress is applied to (111) crystal planes, diffusion in these planes is reduced for self-interstitials but remains almost unaltered for vacancies.

Radiation-Induced Effects on Microstructure*

S. J. Zinkle

1.03.1 Introduction

Irradiation of materials with particles that are suf­ficiently energetic to create atomic displacements can induce significant microstructural alteration, rang­ing from crystalline-to-amorphous phase transitions to the generation of large concentrations of point defect or solute aggregates in crystalline lattices. These microstructural changes typically cause signifi­cant changes in the physical and mechanical properties of the irradiated material. A variety of advanced mi­crostructural characterization tools are available to examine the microstructural changes induced by par­ticle irradiation, including electron microscopy, atom probe field ion microscopy, X-ray scattering and spec­trometry, Rutherford backscattering spectrometry, nuclear reaction analysis, and neutron scattering and spectrometry. ,2 Numerous reviews, which summarize the microstructural changes in materials associated with electron3-6 and heavy ion or neutron4,7-20 irradi­ation, have been published. These reviews have focused on pure metals5-10,12-14,16,19 as well as model alloys,3,9,13,14 steels,11,20 and ceramic3,4,15,17,18 materials.

In this chapter, the commonly observed defect cluster morphologies produced by particle irradia­tion are summarized and an overview is presented on some of the key physical parameters that have a major influence on microstructural evolution of irradiated materials. The relationship between microstructural changes and evolution of physical and mechanical properties is then summarized, with particular em­phasis on eight key radiation-induced property deg­radation phenomena. Typical examples of irradiated microstructures of metals and ceramic materials are presented. Radiation-induced changes in the micro­structure of organic materials such as polymers are not discussed in this overview.