Как выбрать гостиницу для кошек
14 декабря, 2021
The thermal vibration of atoms next to free surfaces, to grain boundaries, to the cores of dislocations, etc., make it possible for vacancies to be created and then diffuse into the crystal interior and establish an equilibrium thermal vacancy concentration of 4 — tsV
kBT.
given in atomic fractions. Here, 4 is the vacancy formation enthalpy, and SV is the vacancy formation entropy. The thermal vacancy concentration can be measured by several techniques as discussed in Damask and Dienes,4 Seeger and Mehrer,5 and Siegel,6 and values for 4 have been reviewed and tabulated by Ehrhart and Schultz;7 they are listed in Table 2. When these values for the metallic elements are plotted versus the melt temperature in Figure 4, an approximate correlation is obtained, namely
4 ~ Tm/1067 [3]
4000 3500
2- 3000
о о
Ь 2500
0 >
2000
га
1 1500 “ 1000
500 0
0 500 1000 1500 2000 2500 3000 3500 4000
Melt temperature (K)
Figure 3 Leibfried’s empirical rule between melting temperature and the product of bulk modulus and atomic volume.
Using the Leibfried rule, a new approximate correlation emerges for the vacancy formation enthalpy that has become known as the cBO model8; the constant c is assumed to be independent of temperature and pressure. As seen from Figure 5, however, the experimental values for 4 correlate no better with BO than with the melting temperature.
It is tempting to assume that a vacancy is just a void and its energy is simply equal to the surface area 4nR2 times the specific surface energy g0. Taking the atomic volume as the vacancy volume, that is, O = 4pR 3/3, we show in Figure 6 the measured vacancy formation enthalpies as a function of 4nR2y0, using for g0 the values9 at half the melting temperatures. It is seen that 4 is significantly less, by about a factor of two, compared to the surface energy of the vacancy void so obtained. Evidently, this simple approach does not take into account the fact that the atoms surrounding the vacancy void relax into new positions so as to reduce the vacancy volume 4 to something less than O. The difference
VVel = VV — O [4]
is referred to as the vacancy relaxation volume. The experimental value7 for the vacancy relaxation of Cu is —0.25O, which reduces the surface area of the vacancy void by a factor of only 0.825, but not by a factor of two.
The difference between the observed vacancy formation enthalpy and the value from the simplistic surface model has recently been resolved. It will be shown in Section 1.01.7 that the specific surface energy is in fact a function of the elastic strain tangential to the surface, and when this surface strain relaxes, the surface energy is thereby reduced. At the same time, however, the surface relaxation creates a stress field in the surrounding crystal, and hence a strain energy. As a result, the energy of a void after relaxation is given by
FC [e(R), £*] = 4-pR2g[e(R), £*] + 8nR3me2(R) [5]
The first term is the surface free energy of a void with radius R, and it depends now on a specific surface energy that itself is a function of the surface strain e(R) and the intrinsic residual surface strain e* for a surface that is not relaxed. The second term is the strain energy of the surrounding crystal that depends on its shear modulus m. The strain dependence of the specific surface energy is given by
g[e, e*]=g0 + 2(mS + 1s)(2£* + e)e [6]
Here, g0 is the specific surface energy on a surface with no strains in the underlying bulk material.
However, such a surface possesses an intrinsic, residual surface strain e*, because the interatomic bonding between surface atoms differs from that in the bulk, and for metals, the surface bond length would be shorter if the underlying bulk material would allow the surface to relax. Partial relaxation is possible for small voids as well as for nanosized objects. In addition to the different bond length at the surface, the elastic constants, m and Is, are also different from the corresponding bulk elastic constants. However, they can be related by a surface layer thickness, h, to bulk elastic constants such that
+ Is = (m т i)h = mh/ (1 — 2v) [7]
where l is the Lame’s constant and v is Poisson’s ratio for the bulk solid. Computer simulations on
freestanding thin films have shown10 that the surface layer is effectively a monolayer, and h can be approximated by the Burgers vector b. For planar crystal surfaces, the residual surface strain parameter e* is found to be between 3 and 5%, depending on the surface orientation relative to the crystal lattice. On surfaces with high curvature, however, e* is expected to be larger.
The relaxation of the void surface can now be obtained as follows. We seek the minimum of the void energy as defined by eqn [5] by solving dFc/де = 0. The result is
(ms + ^s)e* h e*
m’R + (ms + Is) (1 — 2v)R + h and this relaxation strain changes the initially unrelaxed void volume
0
0 500 1000 1500 2000 2500 3000 3500 4000
Melting temperature (K)
Figure 4 Vacancy formation energies as a function of melting temperature.
0
0 1 2 3 4 5
Surface energy of a vacancy (eV)
Figure 6 Correlation between the surface energy of a vacancy void and the vacancy formation energy.
3.5 3
2.5 2
1.5 1
0.5 0
0 5 10 15 20 25 30 35 40
Bulk modulus * atomic volume (eV)
Figure 5 Vacancy formation energy versus the product of bulk modulus and atomic volume.
|
|
|
|
|
|
|
|
|
|
|
|
Few experimentally determined values are available for the vacancy relaxation volume, and their accuracy is often in doubt. In contrast, vacancy formation energies are better known. Therefore, we use eqn [12] to determine the vacancy relaxation volumes from experimentally determined vacancy formation energies. The values so obtained are listed in Table 3, and for the few cases7 where this is possible, they are compared with the values reported from experiments. Computed values for the vacancy relaxation volumes are between —0.2O and —0.3O for both fcc and bcc metals. The low experimental values for Al, Fe, and Mo then appear suspect.
The surface energy model employed here to derive eqn [12] is based on several approximations: isotropic, linear elasticity, a surface energy parameter, g0, that represents an average over different crystal orientations, and extrapolation of the energy of large voids to the energy of a vacancy.
Nevertheless, this approximate model provides satisfactory results and captures an important connection between the vacancy relaxation volume and the vacancy formation energy that has also been noted in atomistic calculations.
Finally, a few remarks about the vacancy formation entropy, SV, are in order. It originates from the change in the vibrational frequencies of atoms surrounding the vacancy. Theoretical estimates based on empirical potentials provide values that range from 0.4k to about 3.0k, where k is the Boltzmann constant. As a result, the effect of the vacancy formation entropy on the magnitude of the thermal equilibrium vacancy concentration, CVq, is of the same magnitude as the statistical uncertainty in the vacancy formation enthalpy.
1.02.2.1 Point Defects Compared to Defects of Greater Spatial Extent
In crystallography, we learn that the atoms and ions of inorganic materials are, with the exception of glasses, arranged in well-defined planes and rows.3 This is, however, an idealized representation. In reality, crystals incorporate many types of imperfections or defects. These can be categorized into three types depending on their dimensional extent in the crystal:
1. Point defects, which include missing atoms (i. e., vacancies), incorrectly positioned atoms (e. g., interstitials), and chemically inappropriate atoms (dopants). Point defects may exist as single species or as small clusters consisting of a number of species.
2. Line defects or dislocations, which extend through the crystal in a line or chain. The dislocation line has a central core of atoms, which are located well away from the usual crystallographic sites (in ceramics, this extends, in cylindrical terms, to a nanometer or so). Most dislocations are of edge, screw, or mixed type.4
3. Planar defects, which extend in two dimensions and are atomic in only one direction. Many different types exist, the most common of which is the grain boundary. Other common types include stacking faults, inversion domains, and twins.1,2
The defect types described above are the chemical or simple structural models for the extent of defects. It is critical to bear in mind that all defect types, in all materials, may exert an influence via an elastic strain field that extends well beyond the chemical extent of the defect (i. e., beyond the atoms replaced or removed). This is because the lattice atoms surrounding the defect have had their bonds disrupted. Consequently, these atoms will accommodate the existence of the defect by moving slightly from their perfect lattice positions. These movements in the positions of the neighboring atoms are referred to as lattice relaxation.
As a result of the elastic strain and electrostatic potential (if the defect is not charge-neutral), defects can affect the mechanical properties of the lattice. In addition, defects have a chemical effect, changing the
oxidation/reduction properties. Defects also provide mechanisms that support or impede the movement of ions through the lattice. Finally, defects alter the way in which electrons interact with the lattice, as they can alter the potential energy profile ofthe lattice (whether or not the defect is charged). For example, this may lead to the trapping of electrons. Also, because dopant ions will have a different electronic configuration from that ofthe host atom, defects may donate an electron to a conduction band, resulting in n-type conduction, or a defect may introduce a hole into the electronic structure, resulting in p-type conductivity.
At intermediate temperatures where SIAs and vacancies are mobile, significant solute segregation to point defect sinks can occur. This can lead to precipitation of new phases due to the local enrichment or depletion of solute. These radiation-induced or — enhanced precipitation reactions typically become predominant
phenomena in irradiated ferritic and austenitic steels at elevated temperatures for doses above about 10 dpa, and in irradiated reactor
pressure vessel steels at low dose rates for damage levels above 0.001-0.01 dpa.247,248 Some general aspects of radiation-induced and — enhanced solute segregation and precipitation were described previously in Section 1.03.3.9. The solute segregation and precipitation associated with irradiation can lead to several deleterious effects including property degradation due to grain boundary or matrix embrit-
tlement224,247,249-252 and enhanced susceptibility for localized corrosion or stress corrosion cracking.253-256 Solute segregation and precipitation can lead to either enhanced or suppressed void swelling
behavior.149,257,258 For austenitic stainless steel,
undesirable precipitate phases that generally are associated with high void swelling include the radiation-induced phases M6Ni16Si7 (G), Ni3Si (g0), MP, M2P, and M3P, and the radiation-modified phases M6C, Laves, and M2P200 The undesirable radiation-induced and — modified phases generally are associated with undersized misfits with the lattice, which tends to preferentially attract SIAs and thereby enhance the interstitial bias effect. Figure 26 shows an example of enlarged cavity formation in association with G phase precipitates in neutron- irradiated austenitic stainless steel.106 Potentially desirable radiation-enhanced and — modified phases (when present in the form of finely dispersed precipitates) include M6C, Laves, M23C6, MC, s, and w200
Figure 26 Enlarged cavity formation in association with G phase (Mn6Ni16Si7) precipitates in Ti-modified ‘prime candidate alloy’ austenitic stainless steel following mixed-spectrum fission reactor irradiation at 500°C to 11 dpa that generated 200 appm He. Reproduced from Maziasz, P. J. J. Nucl. Mater. 1989, 169, 95-115. |
To unfault a 1/4 [110] (110) dislocation loop in spinel, we must propagate a 1 /4[112] partial shear dislocation across the loop plane.12 This is described by the following dislocation reaction:
4[110] + i[H2] ! 1[101] [11]
faulted loop partial shear unfaulted loop
This reaction is shown graphically in Figure 5. When we pass a 1 /4[112] shear through a 1/4 [110] (110) dislocation loop, the atomic planes beneath the loop assume new registries, such that in eqn [6], ajpj and a2p2 commute as follows: a1p1 ! a2p2 ! a1p1. The anion layers beneath the loop are left unchanged (B! B, C! C). Also the Al p’ layers are left unchanged (p’ ! p’). Taking the faulted (110) stacking sequence in eqn [6] and assuming that the planes to the right are above the ones on the left, we perform the 1/4[112] partial shear operation as follows:
Figure 5 Spinel unit cell showing the Burgers vectors involved in the partial shear unfaulting reaction for interstitial dislocation loops in spinel. The blue circles represent Mg atoms (Al and O are not shown here). |
(P’S) (a, P,C) (P’S) KP2C) (P’S (a, b,C) |(P’S)| (a, P,C) (P’S) ^C) (faulted)
P’S a2P2C P’S a, P,C
(P’S) (a, P,q (P’S) (a2P2C) (P’S (a, P,C) (P’S) (a2P2C) (P’S) (a, P,C) (unfaulted)
[12]
After propagating the partial shear through the loop, we are left with an unfaulted layer stacking sequence. The Burgers vector of the resultant dislocation loop, 1/2[101], is a perfect lattice vector; therefore, the newly formed dislocation is a perfect dislocation (equal to the Mg-Mg first nearest — neighbor spacing). The resultant 1/2[101](110) dislocation is a mixed dislocation, in the sense that the Burgers vector is canted (not normal) relative to the plane of the loop.
Complex structural materials, such as FMS and NFA, are composed of a wide range of types, morphologies, size scales, and associations of various microstuctural features. A general master model approach to treating He transport, fate, and consequences in such alloys is shown in Figure 3. He transport and clustering reactions within the matrix and in various microstructural regions are treated. In this framework, RT models, or KMC simulations, are used to transport and partition He to the various microstructural regions, such as dislocations, grain and subgrain boundaries, and heterophase interfaces. He is further transported within a region to and from internal subregions. For example, dislocations are regions, and precipitates and jogs on dislocations constitute subregions. He can recycle within or between regions. The accumulation of He in the various sites results in the formation of He bubbles (and in some cases voids).
Preliminary development of a master model code to implement this framework is underway.178 In its current RT formulation, He is generated by transmutations as both Hei and Hes at relative fractions that can be based on a physically motivated parameter that needs to be established. These interstitial and substitutional forms can switch by interaction with vacancies and SIA, respectively. The Hei rapidly diffuses to the various regions where it is trapped or captured by a matrix vacancy to form Hes. Hes diffuses more slowly to the various regions or is displaced by a SIA to form Hei. The differences in the diffusion rate are reflected in the steady-state matrix concentrations of Hei (very low) and Hes (higher).
A CD model (see Section 1.06.3) is used to track the fate of He and bubble evolution by reactions He + Hem! Hem +1 in the various regions and subregions. Although the framework of the CD models allows the disassociation ofclusters by He emission, this is not implemented in the results described below. That is, a He2 cluster is taken as a stable nucleus for the formation and evolution of larger bubbles. A further assumption is that the HemVn clusters grow with He addition as equilibrium bubbles (m/n < 1), along the lowest free-energy path, with a real gas equation of state p = 2g/rb, as described in Section 1.06.3. This approximation is valid at low damage rates and vacancy-rich environments associated with neutron irradiations. The current implementation focuses on bubble evolution, and since the dislocation bias is set to B = 0, the model does not directly treat void formation and swelling. However, this capability will be implemented in future versions of the code.
Helium atoms trapped at precipitates, dislocations, and GBs may either be emitted back to the matrix, at a rate determined by their binding energies, or diffuse to be captured by deeper subregion traps. The He + He! He2 reactions form a bubble nucleus in all regions and subregions, either heterogeneously with other trapped He or homogeneously with reactions between two freely diffusing He atoms. Nucleation of bubbles on dislocations is a very important process. Dislocations are modeled as a size distribution of segments bounded by deeper traps than the dislocation itself, such as junctions, jogs, and attached precipitates. The initial distribution of dislocation segments is resegmented (split) as bubbles homogeneously nucleate on them.
The master model contains many parameters. Where possible, microstructural observations were used to provide microstructural parameters for grain sizes, dislocation densities, and precipitates, as summarized in Table 6. In general, the binding and activation energies were obtained from the models described in Section 1.06.5. Details are presented elsewhere.178
Figure 41 shows an example of the master model predictions of bubble radii (a, c) and number densities (b, d) compared with the ISHI data described previously in this section for 40 appm He/dpa at 500 °C up to 10 dpa for the FMS F82H (a, b) and NFA MA957 (c, d) microstructure variables shown in Table 6. The data are shown for F82H in both as tempered (AT) and 20% CW conditions. The overall agreement is quite good. The model predicts that almost all of the bubbles form on dislocations in F82H and on dislocations and dislocation associated NF in MA957, broadly consistent with observations. The model predicts a smaller number of larger
Table 6 Typical microstructural parameters for FMS and NFA models
|
|
|||
|
|
Figure 41 Model predictions corresponding in situ He-implanter data for the (a) average bubble radius; (b) number density in F82H; (c) average bubble radius; and (d) number density in MA957also show the observations in experiments.
bubbles than observed in the F82H in the AT condition. The agreement is better for MA957, and the model predictions are consistent with the observation that a higher number density of smaller bubbles form in this case. The MA957 model predicts that there is a lower number of smaller bubbles in the matrix and especially on GBs. Note that the models do not yet contain lath boundaries that are observed to contain a high concentration of bubbles in F82H. The predicted size distribution of bubbles is shown in Figure 42. The agreement with the experimental results is again quite good and reflects the significant differences that are observed in the two alloys.
During the migration of a point defect through the crystal lattice, it traverses an energy landscape that is schematically shown in Figure 18. The energy minima are the stable configurations where the defect energy is equal to Ef(r), the formation energy, but modified by the interactions with internal and external strain fields, which in general vary with the defect location r. In order to move to the adjacent energy minimum, the defect has to be thermally activated over the saddle point that has an energy
ES(r)=Ef (r)+Em(r) [69]
where Em(r) is the migration energy. As the properties of the point defect, such as its dipole tensor and its diaelastic polarizability, are not necessarily the same in the saddle point configurations as in the stable configuration, the interactions with the strain fields are different, and the envelope of the saddle point energies follows a different curve than the envelope of the stable configuration energies, as indicated in Figure 18. For a self-interstitial, we
Figure 18 Schematic of the potential energy profile for a migrating defect. |
must also consider the different orientations that it may have in its stable configuration. Accordingly, let Cm(r, t) be the concentration of point defects at the location r and at time t with an orientation m. For instance, the point defect could be the selfinterstitial in an fcc crystal, in which case, there are three possible orientations for the dumbbell axis and m may assume the three values 1, 2, or 3 if the axis is aligned in the xb x2, or x3 direction, respectively. The elementary process of diffusion consists now of a single jump to one adjacent site at r + R, where R is one of the possible jump vectors.
The rate of change with time of the concentration Cm (r, t) is now given by
—C
= Cn(r — R, t)Lnm(r — R | R)
r, v
‘У ‘ Cm(r, t)Lmn(r 1 R) [70]
r, v
Here, the first term sums up all jumps from neighboring sites to site r thereby leading to an increase of Cm (r, t), while the second term adds up all the jumps (really the probabilities of jumps) out of the site r. The frequency (or better the probability) of a particular jump from r to r + R while changing the orientation from m to V is denoted by Lmn(r | R). The eqn [70] applies now to each of the possible orientations, and it appears that this leads to as many diffusion equations as there are possible orientations, and these equations may be coupled if the defect can change its orientation between jumps.
To circumvent this complication, one considers an ensemble of identical systems, all having identical microstructures, and identical internal and external stress fields. The ensemble average of the defect concentration at each site, denoted simply as C(r, t) without a subscript, is now assumed to be the thermodynamic average such that exp(-bEm)
£exp(-bEvf)
= C(r, 0exp(-b4)/N(r)
where the normalization factor N only depends on the location r as do the energies for the stable defect configurations.
Substituting this into eqn [70] on both sides constitutes another assumption. To see this, suppose that the defect concentrations CV(r — R, t) on all neighbor sites happen, at the particular instance t, to be aligned in one direction. Since their new
|
|
|||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
|