Как выбрать гостиницу для кошек
14 декабря, 2021
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
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.
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
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) |
|
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.
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 |
|
|
||
|
|||
|
|||
|
|
||||||||||||||||
|
|||||||||||||||||
|
|
||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
|
|||||||||||||||||
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
|
Uniaxial stress (GPa)
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.
S. J. Zinkle
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.