You are on page 1of 19

140 Journal of Nuclear Materials 150 (19X7) 140~15X

North-Holland. Amsterdam

MICROSTRUCTURE AND PROPERTIES OF COPPER ALLOYS FOLLOWING 14-Mel’


NEUTRON IRRADIATION *

S.J. ZINKLE
Metals and Ceramics Division, Oak Ridge National Lahoratoty, P.O. Box X, Oak Ridge, TN 37831, USA

Received 10 April 1987; accepted 26 May 1987

The microstructure of copper and three single-phase copper alloys (Cu-5% Al, G-5% Mn, Cu-5% Ni) has been examined
following a room-temperature 14-MeV neutron irradiation up to a maximum fluence of 2X 102t n/m* ( -10-s dpa). The
irradiation produced small defect clusters, many of which were resolvable as stacking fault tetrahedra (SFT) or triangle loops.
Alloying did not affect the overall cluster density or size distribution, but it did cause a significant reduction in the density of
SFT and triangle loops. The overall defect cluster size was independent of fluence with a mean and most probable diameter of
- 2.6 and - 1.7 nm, respectively. There was no evidence of direct cascade overlap in this study. Cascade overlap effects are
estimated to become significant for fluences - 1O22 n/m2. The overall defect cluster density was found to vary as N - J&r
with a short transition period for all four metals. Re-examination of published data indicates that this relationship may be
broadly applicable over the irradiation temperature range of 80 to 400 K. Logarithmic plots of the cluster density versus
fluence showed anomalous curvature due to the nonlinear defect survival rate. Logarithmic plots should not be used
exclusively to determine the fluence dependence of radiation damage processes.

1. Introduction followup investigation utilizing electrical resistivity, mi-


crohardness, and electron microscopy was initiated. The
An extensive radiation effects data base has been resistivity and Vickers microhardness results from this
established based on fission reactor irradiation of later study are described elsewhere [3,4]. The present
structural materials. However, successful prediction of report concentrates on the microstructural changes ob-
radiation damage effects in D-T fusion reactors re- served in copper and the copper alloys as a result of
quires that the differences and similarities between fis- irradiation. Radiation hardening results from the litera-
sion and fusion neutron damage microstructures be ture are also reexamined to determine their fluence
fully understood. The present study is a continuation of dependence. A detailed analysis of the resistivity and
a fission-fusion correlation program utilizing the Rotat- TEM measurements made on pure copper is given in a
ing Target Neutron Source (RTNS-II) 14-MeV neutron companion paper [5].
facility at Lawrence Livermore National Laboratory Irradiation produces point defects in isolated or clus-
that was initiated in 1979 [l]. Copper was included in tered forms in a manner that is directly proportional to
this program as a representative fee metal with a rela- fluence. However, numerous studies have found that the
tively low stacking fault energy. Three single-phase surviving defect concentration dependence is less than
copper allows were also selected in order to study solute linear with fluence during irradiation at temperatures
effects. where at least one of the defect types is mobile (see
Initial results from the copper irradiation program section 4). This behavior has often been proposed to be
were reported by Brager et al. [2]. Several questions evidence of cascade overlap effects. An alternative ex-
remain unanswered following this initial study, so a planation found in the electrical resistivity literature
suggests that this effect is due to uncorrelated recombi-
* Research sponsored by the Office of Fusion Energy, US nation of freely migrating defects. In this paper we will
Department of Energy under contract DE-AC05-840R21400 attempt to determine which of these two explanations is
with Martin Marietta Energy Systems, Inc. most plausible.

0022-3115/87/$03.50 0 Elsevier Science Publishers B.V.


(North-Holland Physics Publishing Division)
S.J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys 141

2. Experimental procedure mA current (6-V dc potential) in a Tenupol twin-jet


polishing unit using a solution of 33% HN0,/67%
Foils of pure copper (> 99.99%) and copper alloyed CH,OH cooled to - 25 o C. The foils were examined in
with 5 at% of either aluminum, nickel, or manganese a JEOL 2000FX electron microscope. Foil thickness
were used in this irradiation study. As shown in table 1, were determined using either thickness fringes or con-
these single-phase alloys contain either undersized (Ni) vergent beam techniques. Characte~zation of the micro-
or oversized (Al, Mn) solutes [6,7]. The addition of 5 structure was performed on foils with thicknesses of 50
at% Al reduces the stacking fault energy of copper by a to 120 nm using bright-field and weak beam dark-field
factor of 2 (refs. [8,9]). These three alloys remain single imaging. The defect cluster size and density measure-
phase at room temperature for solute concentrations up ments were made using a Zeiss particle analyzer on
to at least 15%, although there are some indications of prints with a total magnification of about 1 x 106. A
short-range ordering at lower solute levels. total of 1000 to 4000 defect clusters from at least two
Foils of 250 pm thickness that were used in a different foil regions were counted for each irradiation
previous study [2] were obtained and subsequently condition.
cold-rolled to a thickness of 25 pm. Transmission elec- The visibility of small defect clusters is influenced to
tron microscopy (TEM) disks were punched from the a certain degree by the resolution of the electron micro-
rolled foils. The disks were then recrystallized by an- scope and also by the imaging conditions (foil thickness,
nealing in high-purity argon and allowed to air cool. value of the deviation parameter, order of the weak
The pure copper specimens were annealed at 400 ’ C for beam reflection, etc.). Several investigations have ex-
15 min and the copper alloys were annealed at 750 o C amined the effect of these parameters on the visibility of
for 30 min (41. The four alloys were irradiated at room small dislocation loops and SFT [12-161. We have
temperature (25OC) with 14-MeV neutrons using the generally used the recommended conditions of a large
RTNS-II facility. The TEM disks were positioned away deviation parameter (ss 2 0.2/nm) and a (002) reflec-
from the neutron source along four different isoflux tion to ensure maximum visibility of the defect clusters.
contours. This produced four fluence levels (determined Stathopoulos [16] has reported that 26% of the clusters
from niobium disometry foils) of about 1 x 102’, 4 x in the ion-irradiated copper are invisible for a g = 002
102’, 1 X 1021, and 2 X 1021 n/m2 at the end of the reflection. We have observed some cases of loops and
irradiation. Earlier studies had determined that varia- SFT that go out of contrast at certain beam orientation
tion in the 14-MeV neutron dose rate over three orders even for g. b + 0, but the relative fraction does not
of magnitude did not affect the details of cluster forma- appear to be as high as 26%. The densities reported in
tion in copper irradiation at room temperature [2,10]. this paper refer to the measured visible density, and do
Electrical resistivity foils of the four materials were not contain any correction for invisible defect clusters.
also irradiated at the same time to a maximum fluence
of 3 X 1021 n/m2. Four resistivity measurements were
made by interrupting the irradiation and immersing the 3. Results
sample assembly in liquid helium. The pure copper
resistivity results required a correction for size effects The neutron irradiation produced a high density of
due to surface scattering that was on the order of 10%. small defect clusters in copper and the copper alloys.
Microhardness measurements were made on the periph- All of the visible defect clusters were less than 10 nm in
ery of the TEM disks following the irradiation. Further diameter. There was no evidence of void formation in
details of the specimen preparation and irradiation are any of the irradiated foils. Fig. 1 shows the general
described elsewhere [4,5,11]. microstructure of copper following irradiation to a
The TEM specimens were electropolished with 100 fluence of 2 X 1021 n/m2. This corresponds to a damage

Table 1
Effect of solute additions on the properties of copper

Property Copper Cu-5 at% Al Cu-5 at% Mn Cu-5 at% Ni Ref.


Lattice parameter (nm) 0.36150 0.36290 0.3639 0.36096 [61
(Volume size factor of solute) (+ 19.99%) (+ 34.19%) (- 8.45%) (71
Stacking fault energy (mJ/m*) 50 25 50 52 B91
142

Fig. 1. Weak beam (8,3g) microstmcture of copper irradiated to a fluence of 2~ 102’ n/m’. The foil normai is near [llO]; g = 002
and s = O.Z/nm.

level of - 7 x 10e4 dpa using the calculated displace- Each subcascade region may evolve into a defect clus-
ment cross section of 3690 barns for 14-MeV neutrons ters which is visible by TEM. Fig. 3 shows an example
on copper [17]. The irradiated microstructure of the of these closely spaced subcascade clusters in a
three copper alloys was very similar to that of pure low-fiuence specimen. The number of visible defect
copper. Fig. 2 shows typical damage ~crostructures for clusters associated with each cascade region ranges from
each of the four metals. 1 to 7. The average spacing between clusters in a given
The mean free path between scattering events for a cascade is about 5 nm. The relation between the num-
14-MeV neutron in copper is 4 cm. Therefore, at low ber of visible clusters and the number of subcascade
fluences, defect clusters formed in one displacement lobes per cascade event is rather uncertain [20]. Vacancy
cascade are well separated from other cascades. The clusters can form directly from the collapse of the
14-MeV neutron is sufficiently energetic to allow the vacancy-rich core of the subcascade, whereas interstitial
cascade to develop into distinct subcascade lobes [l&19]. clusters are presumably formed by classical nucleation
S. J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys 143

Fig. 2. Weak beam (g, 3g), g = 002, microstructures of copper and copper alloys irradiated to a fhence of 2 x IO*’ n/m*. The foil
normal is near [I101 for the top two pictures, and B = [OOl]in the bottom two pictures.

and growth and may involve cooperative effects from regions are still well isolated from each other at this
adjacent subcascade lobes. fluence level. (The actual intercascade spacings are even
At sufficiently high fluences (high defect cluster den- larger than is apparent in fig. 4 as the foil thickness is
sity), the newly created cascades begin to physically much greater than the intercascade spacing.) This con-
overlap the existing defect clusters and the cluster den- elusion is important for determining the fluence depen-
sity approaches a saturation value. Experimentally, dence of the defect cluster density. Brager et al. [2] have
cascade overlap becomes evident when the volume frac- similarly concluded that there are no cascade overlap
tion occupied by the defect clusters becomes large (say, effects in these alloys for 14-MeV neutron fluences
1 10%). There is no evidence of physical cascade over- approaching 10 ‘* n/m2.
lap at the maximum fluence studied in this investiga- Fig. 6 shows a high magnification example of the
tion, 2 X 1021 n/m2. Figs. 4 and 5 show that the cascade bright-field microstructure of copper following irradia-
144 S.J. Zmkle / Microstructure andpropertres ojneu~ron rrradrcltrd Cu allo~.s

Fig 3. Isolated groups of defect clusters observed in Cu-5’% Al after irradiation to 4~ 10’” n/n?.

tion to 2 X 10” n/m2. Triangle-shaped defect clusters, proposed by Matthai and Bacon [22]. SFT with edge
which are either stacking fault tetrahedra (SFT) or lengths as large as 8 nm have been observed in high-
triangular loops, are clearly visible in this orientation fluence specimens. The observations made in this study
(B = [110]). The use of other orientations such as B = do not allow a determination to be made of whether
[OOl] has shown that SFT and triangle loops occur in SFT formation occurs directly (Matthai and Bacon) as
roughly equal numbers independent of fluence for the opposed to dissociation of a Frank vacancy loop (Silcox
experimental conditions of this study. According to the and Hirsch).
Silcox and Hirsch model [21], the SFT is formed from The effect of fluence on the damage microstructure
the dissociation of a triangular vacancy loop. of Cu-5% Al is shown in fig. 8. The average defect
Fig. 7 shows one of the largest SFT that was observed cluster size was found to be constant over the fluence
in this study. The defect cluster, which has an edge range investigated for all four alloys. The only micro-
length of 6.5 nm, appears as a square for the beam structural effect associated with increasing fluence was
orientation that was used (B = [OOl]). This SFT was an increase in the defect cluster density.
present in a foil that was irradiated to the lowest fluence The size distribution of defect clusters in the four
studied in this investigation, 1 x 102” n/m2. It is possi- metals following irradiation to 2 x 102’ n/m* is given
ble that the SFT may have formed directly from the in fig. 9. The “total” size distribution includes all visible
vacancy-rich core of the displacement cascade region, as clusters. The “PDL + SFT” distribution includes only
S.J. Zinkle / Microstructure andproperties of neutron irradiated Cu alloys 145

Fig. 4. Groups of defect clusters observed in 0-58 Mn after irradiation to 2 X 10L’ n/m’. The foil thickness is about 110 nm.

resolvable partially dissociated loops (PDL) and trian- The total cluster density is plotted as a function of
gle loops and SFT. The total size distributions for the fluence in figs. 10 and 11. The linear plot of density
four materials are very similar, with a fhrence-averaged versus fluence (fig. 10) exhibits some curvature, indicat-
most probable cluster size of d,, = 1.7 nm and a mean
size of d= 2.6 nm. The size distributions are closely Table 2
approximated by log-normal curves [3]. As shown in Relative fractions of SFT and partially dissociated loops (PDL)
table 2, about half of the visible triangle loops were in irradiated alloys
fully dissociated as SFT. Pure copper contains a larger
Copper cu-5 cu-5 h-5
fraction of defect clusters that are resolvable as triangle
at% Al at% Mn at% Ni
loops or SFT compared to the copper alloys (20 versus
12%). The peak in the size distribution of triangle loops NSFT
p(%) 50 _ 30 50
and tetrahedra is about 2.0 nm for copper, Cu-5% Mn, N sFr+ PDL
and Cu-5% Ni. The corresponding peak in the SFT size
$-=(%)
N 20 14 10 9
distribution for the low stacking fault energy alloy
TOT
Cu-5% Al is d,, - 2.5 nm.
146 S.J. Zinkfe / Microstructure and properties ofneutronirradiated Cu alio,vs

Fig. 5. Isolated groups of defect clusters in copper after irradiation to 2 X 102’ n/m2. The foil thickness is about 30 nm

Fig. 6. Bright-field (g = 004) microstructure of copper irradiated to a fluence of 2 X 1021 n/m’.


S.J. Zinkle / Microsiructure and properties of neutron irradiated Cu alloys 147

Fig. 7. Large stacking fault tetrahedron observed in copper following irradiation to a fluence of 1 x lo*’ n/m*.

ing that the fluence dependence is less than linear. A ted foils is plotted as a function of fluence in fig. 12.
better fit is obtained when the data are plotted versus Uncertainties associated with resolving distinct defect
the square root of fluence (fig. 11). Following a transi- cluster shapes at sixes on the order of 2 nm do not allow
tion fluence of - 1 X lo*’ n/m*, the data for all four the fluence dependence of the cluster density (+t versus
materials is nearly linear with the square root of neu- @) to be accurately determined, although it appears
tron fluence. Higher fluence data are needed to confirm to be slightly sublinear. Unlike the total cluster density,
this square root dependence. There is not a large dif- the density of SFT and triangle loops is significantly
ference in cluster density between the four metals over different for the alloys compared to pure copper. The
the fluence range studied in this investigation. This density of these types of clusters in the alloys is about
implies that solute additions to copper do not have a half of the copper value at a fluence of 2 x lo*’ n/m2
significant effect on the total number of point defects (table 2). The SFT (and triangle loops) are expected to
surviving a room-temperature 14-MeV neutron irradia- be vacancy-type defect clusters for the irradiation con-
tion. ditions that were employed. If this assumption is cor-
The density of SFT and triangle loops in the irradia- rect, then it is clear that substitutional solute additions
(both oversized and undersized) reduce the fraction of pared to copper (fig. 12 and table 2), since the short-
vacancy clusters that survive the overall displacement range ordering process will most likely reduce the free
cascade event. This may be due to preferential trapping vacancy concentration in the cascade due to vacancy.-
of vacancies at solute atoms. solute interactions.
Electrical resistivity foils were irradiated along with Careful examination of the electron diffraction pat-
the TEM disks in this study. The radiation-induced terns of the irradiated alloys did not reveal any indica-
resistivity changes are shown in fig. 13. It was found [3] tion of extra spots associated with the proposed short-
that the resistivity change of pure copper was propor- range ordering (SRO). The only fine structure that was
tional to the square root of neutron fluence with a visible in the diffraction patterns obtained at exact zone
transition fluence of about 1 x 10’” n/m’. in agreement axes was faint streaks through the diffraction spots and.
with the TEM measurements of fig. 11. The resistivity in some instances, spot splitting, This behavior was also
changes of the three copper alloys showed a fiuence- observed in pure copper diffraction patterns and is
dependent behavior that was attributed to short-range probably due to the presence of stacking fault tetra-
ordering effects 13). The electrical resistivity can in- hedra and small faulted loops in the foil [27.28]. The
crease or decrease as a result of ordering and clustering absence of additional diffraction spots in the irradiated
processes, depending on the alloy system and the solute alloys indicates that appreciable clustering ha& not oc-
concentration [23-261. For example, the large resistivity curred. The observed results might be explained by
increase in Cu-5% Ni compared to copper as 3 X IO” assuming that dispersed SRO ]29,30], which does not
n/m2 (fig. 13) can be attributed to the formation of require homogeneous ordering, has occurred. or that the
soiute clusters f24}. On the other band. short-range intensity from the SRO is very weak.
ordering is known to produce the observed resistivity The visibility of very small loops (d < 2 nmf was
decrease in Cu-AI and Cu-Mn alloys (X]. The reason found to depend on the foil thickness, as shown in fig.
for the large resistivity increase in C&-S% RI at high 14. Under similar imaging conditions the visible cluster
fluence is uncertain, The measured resistivity changes density in very thin foils (f -C45 nm) was about 60%
are consistent with the observed low density of higher than that found in relatively thick foils (t - 100
vacancy-type defect clusters (SFT) in the alloys com- nm). Comparison of the two size distributions revealed
S.J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys 149

a greatly enhanced proportion of small clusters in the


thin foils. When the two size distributions are normal-
ized to the total cluster density observed in the thin
foils, a similar size distribution is obtained for cluster
sizes larger than - 2 nm. It is apparent that the ob-
served defect density and the mean and most probable
defect size are interrelated and depend on the foil
thickness. For the sake of consistency, we have made
size and density measurements only on micrographs
that are of similar foil thickness (50 to 110 nm). There-
fore, the relative size and density of defects reported in
this study are though to be fairly accurate. The actual
defect number density and mean size may be somewhat
I I I r higher and lower, respectively, than reported. Similar
Cu-5%Mn Cu-5%Ni foil thickness effects have been reported by Riihle and
Crump [31].

n
TOTAL TOTAL

ii

&
SFT+PDL
Fig. 9. Size distribution of defect clusters in copper and copper
alloys following irradiation to 2 X 102’ n/m2. The curve marked
“Total” includes all visible clusters. The curve marked “PDL
0 2 4 6 0 0 2 4 6 0 +SFT” includes only triangle loops, SFT, and loops that are
IMAGE WIDTH (nm) partially dissociated towards SFT formation.

I
_
25°C IRRADIATION
. COPPER
0 CU-5%AI
_
A Cu-5%Mn
A Cu-5%Ni

/
1:x IdO 4xroao 1x102’ 2 x102’

FLUENCE (n/m’1

Fig. 10. Visible defect cluster density of copper and copper alloys as a function of fluence.
150 S.J. Zinkle / Microstructure and properties ofneutronirradiated Cu alloys

I I

25’c IRRADIATION
l COPPER
0 Cu-5%AI
A Cu-5XMn
A Cu-5%Ni

0 lX!(f” 4xiom 1 XIO“ 2 XIO”

fi h/m*)

Fig. 11. Visible defect cluster density of copper and copper alloys plotted versus the square root of fluence.

1 I I -I
i2
259: IRRADIATION
1 ’
I
0 COPPER
10 0 Cu-5%AI
A Cu -5%Mn
A Cu-S%Ni

I I
0.5 1.0 1.5 2.0 2.5

+t (i02’n/m2)

Fig. 12. Density of SIT and triangle loops versus fluence.


S.J. Zinkle / Microstructure and properties ofneutronirradiated Cu alloys 151

4. Discussion
25 “C IRRADIATION
0.3 4 K MEASUREMENT
4.1. Defect microstructure
Pd: 0.2 . Cu-5% Mn
Al
There have been several recent high resolution TEM
c . Cu-5% NI
Y 0.1 studies of copper and copper alloys following neutron
f irradiation near room temperature [2,32-341. Muncie et
5 0
al. [32] found that about 8% of the defect clusters were
c
in the form of SFT following a fission reactor irradia-
z -0.4
tion to 1.3 X 102’ n/m2 (E > 1 MeV). On the other
g -0.2 hand, Yoshida et al. [33] reported that more than one
half of the visible defect clusters were SFT following
-0.3
14-MeV neutron irradiation to fluences between 8 X lOi
I I I I and 8.5 X 102’ n/m2. We have found in this study that
-0.4 L
(X1020 3X(020 (X(02 3x402’ about 10% of the defect clusters are resolvable as SFT
q h/J) for 14-MeV fluences between 1 X 102’ and 2 x 102’
Fig. 13. Electrical resistivity change versus the square root of n/m’, in agreement with Muncie et al. [32]. An ad-
fluence. ditional 10% of the defect clusters are visible as triangle
loops with their edges along (110) directions.
Yoshida et al. [33] has reported that the average size
of defect clusters in copper decreased slightly with
increasing fluence, whereas Shimomura et al. [34] re-
ported that the defect size increased with fluence. As
discussed earlier, the mean visible cluster size is some-
what dependent on the imaging conditions and foil
1 thickness. Brager et al. [2] stated that the cluster size in
COPPER copper and three copper alloys was independent of
i x io2’ n/m fluence up to +t = 7 X 102t n/m2. This last observation
is in agreement with the present results, where the mean
defect cluster size for each of the four materials was
determined to be constant over the fluence range of
1 X 10” to 2 X 102’ n/m2. A theoretical model devel-
oped by Ghoniem et al. [36] predicts that the mean size
of clusters in copper should be essentially constant for
fluences between 1019 and 1O22 n/m2. It is interesting
to note that the cluster size may be independent of
fluence up to very high damage levels. A recent high-en-
ergy ion irradiation study of copper [36] at 10 dpa
reported SFT and total cluster size distributions that
t= lo-440nm
were essentially identical to those given in fig. 9, which
is for copper irradiated to < 10m3 dpa. Of course, this
direct comparison is not strictly valid since damage rate
and primary revoil spectrum effects may be playing a
role in the development of the microstructure. Similar
size distributions have also been reported by English
[37] for copper irradiated with fission neutrons and 30
keV ions. In all cases [2,16,33,34,36-381, the cluster size
0 1 2 3 4 5 6 7 a 9
distribution for copper irradiated near room tempera-
IMAGE WIDTH (nm)
ture appears to be well described by a log-normal [3]
Fig. 14. Comparison of the visible cluster size distribution in distribution function which peaks at a diameter I 2
thin and thick foils. About 1500 defect clusters were counted nm.
for each of the two size distributions. Larson and coworkers [39,40] have used X-ray dif-
152 S.J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys

fuse scattering to determine the size distributions of 4.2. Cascade overlap and defect retention
interstitial and vacancy clusters in irradiated copper.
Their results are in good agreement with figs. 9 and 14, There is interest in determining at what fluence
with a reported most probable and mean diameter of cascade overlap effects become significant for both
- 2 and 2.6 nm, respectively. Their results also indicate fundamental and practical reasons, particularly in rela-
that there is not a significant number of surviving tion to damage-rate experiments [41-441. Ghoniem et
clusters with a diameter less than 1 nm. This offers the al. [35] have predicted that significant cascade overlap
possibility that TEM is capable of resolving all of the should occur in copper during room-temperature irradi-
defect clusters remaining in copper after a room- ation for 14-MeV fluences greater than - lo*” n/m2.
temperature irradiation if very thin foils are used (e.g., We have not observed any significant amount of physi-
fig. 14). cal cascade overlapping for fluences up to 2 x 10 z’
It appears that stacking fault energy by itself does n/m2 (i.e., the volume that contained defect clusters
not have a strong effect on the probability of 14-MeV was a small fraction of the total volume). Brager et al.
neutron displacement cascades producing visible defect [2] reported that no cascade overlap effects were
clusters in copper alloys. This conclusion is based on observed in copper alloys for I4-MeV fluences ap-
the observation (figs. 9 and 10) that the overall density proaching 10 ** n/m2. A simple estimate of the maxi-
and size distribution of defect clusters in Cu-5% Al mum possible cluster density may be made by assuming
(y = 25 mJ/m*) was the same as for Cu-5% Mn and that the cluster spacing at saturation is comparable to
Cu-5% Ni (y = 50 mJ/m2). The Cu-5% Al alloy did the spacing between visible clusters associated with a
contain a slightly larger density of SFT and triangle single cascade. The mean projection of this spacing has
loops compared to the other two alloys (table 2 and fig. been measured in low-fluence specimens (e.g., fig. 3) to
12), indicating that there may be an effect of stacking be I - 5 nm. This corresponds to saturation density of
fault energy on the formation of these types of clusters. clusters N - le3 = 8 x 1024/m’. The scattering cross
Low-energy (30-keV) ion irradiations of copper alloys section for a 14-MeV neutron in copper is u = 3 barns.
have found that SFT formation is strongly influenced The fluence required to produce 8 x 1O24 cascades/m?
by the stacking fault energy [38]. is therefore (+t),,i = 3 X 1O23 n/m2 (0.1 dpa). Direct
The data from the present study and other re- cascade overlap effects might be expected to become
searchers on the influence of stacking fault energy [38]. evident when the cascade density is 2 lo%, of the
irradiation temperature [16,32,37,38], and knock-on en- maximum value, which would correspond to a fluence
ergy [16,32,33,36,37] on SFT formation in copper may of 3 X 10” n/m* (0.01 dpa). This fluence is an order of
be summarized as follows: low-energy primary knock-on magnitude larger than the highest fluence investigated
atoms (PKAs) do not produce SFT in copper at room in the present study.
temperature. Under these conditions, SFT formation There are conflicting results in the literature regard-
occurs only in low-stacking-fault energy alloys (y - 10 ing the fluence dependence of defect cluster formation
mJ/m*). The probability of SFT formation in copper in fee metals such as copper. Low temperature ( < 20 K)
increases with increasing irradiation temperature up to irradiation studies [41-43,451 have shown that the pro-
330°C. The formation of SFT at room temperature is duction rate of point defects and defect clusters is
enhanced when the mean knock-on energy is increased. directly proportional to the fluence for doses up to 10 ~’
At comparable doses (I 10-s dpa), the fraction of dpa (- 10” n/m2). Saturation effects due to cascade
defect clusters resolvable as SFT varies with knock-on overlap become evident at higher fluences. Electrical
energy in the following way: 0% (30-keV Cu ion), 8% resistivity [46-491 and TEM [49] investigations have
(fission neutron), 10% (14-MeV neutron). There may be conclusively demonstrated that electron irradiation of
a complex interdependence between knock-on energy copper at temperatures where the interstitial is mobile
and stacking fault energy (alloying) effects on SFT (Z 50 K) results in a square root dependence between
formation. For example, Stathopoulos et al. [16,38] the surviving defect density and fluence, N - fi. The
found that copper did not form SFT while 26% of the nonlinear behavior is due to uncorrelated recombina-
defect clusters in Cu-10 at% Al were SFT following tion of Frenkel defects at distant sinks, including defect
30-keV Cu ion irradiation. On the other hand, we have clusters. The results from the present study, using both
found significantly (factor of 2) more SFT in copper TEM (fig. 11) and electrical resistivity [3] methods, and
compared to Cu-5 at% Al following 14-MeV neutron those of Ipohorski and Brown [50] strongly indicate that
irradiation. The mean PKA energies for these two the surviving point defect density continues to be pro-
irradiation conditions are - 1 and - 170 keV, respec- portional to ,/@ for irradiation temperatures as high as
tively.
S.J. Zinkte / Microstructure and properties of neutron irradiated Cu alloys 153

40 o C and doses up to 10” n/m2 (E > 1 MeV). Most apparent change of slope from n = 1 to n = : for fluence
TEM studies of copper irradiated with neutrons near levels around 10” to 102i n/m*. This nonlinear behav-
room temperature have reported a “nearly linear” de- ior cannot easily be attributed to direct cascade overlap
pendence between the visible cluster concentration and since these effects are not expected to be large for
the fluence [2,33,X]. Konstantinov et al. [.52] described fluences less than 102’ n/m2 (see earlier discussion).
their defect cluster data using a fluence exponent of On the other hand, it can be easily shown that
n = 0.7 for a 200 * C proton irradiation. Deviation from fluence-dependent recombination effects can produce a
linear behavior was often observed at moderate fluences; nonlinear curve in the plot of log N versus log (@t). As
this was attributed to cascade overlap effects [33,53]. an example, consider the following expression:
Reexamination of the published data (e.g., ref. [51])
N=A[dm-l]. (1)
often indicates that it would better be described by a
square root dependence. Chaplin and Coltman [54] have This equation gives a good description of the observed
recently reported high-dose damage resistivity data for fluence dependence of the visible cluster density (fig.
copper irradiated with fast neutrons at 60’ C. If their 11) and resistivity [3] results (i.e., a transition period
data are replotted versus the square root of fluence, a following by a square root dependece). For small
linear relationship exists up to +t - 5 X 1O22 n/m2 (E fluences (B+t e t), the logarithm of this expression
> 1 MeV). Concave curvature in the ,/‘&? plot becomes reduces to
evident at higher fluences, presumably because of log N= c-t” log(+t), (2)
cascade overlap effects.
Fig. 15 shows the combined data on defect cluster where C is a constant. At large fluences (B+t x=- l), the
density from Yoshida et al. [33] and the present study as corresponding equation is
a function of fluence. Yoshida has reported a similar log IV=I)+ f log((Pt), (3)
size distribution as that used in this paper, so the
relative densities measured in the two studies should be where D is another constant. The limiting slopes given
compatible. When presented as a double logarithmic by eqs. (2) and (3) are in good agreement with the
plot, the data are not linear but instead exhibit an behavior observed in fig. 15. Curvature in a logarithmic
plot can also be obtained by assuming a bilinear of
more complex mathematical function for defect survival.
Eq. (I) is based on the unsaturable trap model
4023 [46,48], which has been successfully used to describe
defect concentrations in copper for electron irradiation
in the recovery Stage II temperature range (60 to 220
IS). This model is a poor description of the actual
physical processes that occur during fast neutron
(cascade-producing) irradiation at and above Stage III
temperatures ( - 300 K), where vacancies are mobile
along with interstitials. Recombination, nucleation, and
spontaneous in-cascade clustering effects occur simulta-
neously under these conditions, and the overall process
might not be described by such a simple analytical
expression (e.g., it may be bilinear or a nonanalytic
l et 01. (49851
YOSHIOA expression). Nevertheless, the fluence dependence of the
A PRESENT STUDY cluster data in all cases investigated [3,33,49-52,541 is
functionally well described by an expression similar in
form to eq. (1) - namely, a short transition period
4020
I I I I Illll I I I I11111 I I I i illi followed by a regime that is nearly proportional to the
4049 4020 402' 4022 square root of fluence. Although this agreement is prob-
+t (“h2) ably coincidental, it allows a simple empirical descrip-
Fig. 15. Total visible cluster density of pure copper as a tion of the cluster density to be obtained as a function
function of fluence, including data from Yoshida et al. [33]. of fluence. Cascade overlap effects do not appear to
Note the anomalous change in slope that occurs around 5 x 10” become evident until the dose exceeds 1O22n/m* (E > 1
n/m*. MeV).
154 S.J. Zinkle / Microstmcture and properties of neutron irradiated Cu alloys

A logarithmic presentation of the cluster density a transition fluence followed by a one-fourth root de-
data (fig. 15) introduces confusion due to the effect of pendence on fluence. (The data in the original papers
fluence-dependent sink terms which produce a nonlin- had generally been fit to n = + or n = { fluence expo-
ear behavior. In particular, a naive interpretation of fig. nents with varying success.) Deviation from linearity in
15 might suggest that the cluster density is directly the n = $ curve occurs at relatively high doses ( > 10”
proportional to the fluence and that cascade overlap n/m’) and is presumably due to cascade overlap ef-
effects become important for fluences of - 5 x lo*’ fects. An example of radiation hardening data for pure
n/m2 (2 X 10e4 dpa). This interpretation is not in copper plotted versus (+t) iI4 is shown in fig. 16. Fol-
agreement with the results presented in this paper. In lowing a transition fluence of - 1 X lo*’ n/m*, the
summary, logarithmic plots should be used with caution data are linear up to a fluence of 0.5 to 1 x lo** n/m*.
since they may give misleading results with regard to In addition to ambient temperature (20 to 80” C)
the functional dependence of measured quantities. neutron irradiation data on copper, the analyzed
The low-fluence transition regime that occurs in hardening data included results from electron [65] and
plots of N versus fi is determined by the relative low-temperature (77 K) irradiations [67], and fee metals
probability of uncorrelated recombination of point de- other than copper [68,70,71]. Heinisch et al. [60] have
fects compared to other sink and clustering terms. It is recently found that yield strength changes in copper
not a true incubation fluence, as can be seen by ex- and stainless steel following neutron irradiation near
amining eq. (1) or a linear plot of density versus fluence. room temperature (25 and 90°C) were best fit by a
curve proportional to the fourth root of fluence. It
4.3. Radiation hardening therefore appears that the relationships N - fi and
Au - ( +r)“4 may be generally applicable to many fee
A considerable amount of uncertainty exists in the metals over the irradiation temperature range of 100 to
literature regarding the fluence dependence of radiation 400 K.
hardening in fee metals for irradiation temperatures of Fig. 17 shows some recent radiation hardening re-
80 to 400 K. The fluence exponent for hardening in the sults for stainless steel irradiated with 14-MeV neutrons
equation Au - (+t)” was originally measured as n = : [71]. The fluence dependence of hardening as de-
by Blewitt and coworkers [55], and later reported to be termined from a double logarithmic plot of the data
n = 4 by Diehl and coworkers [56]. Subsequent studies (which is common practice in the literature) is given in
have generally used either n = f or n = f to describe the insert of fig. 17(a) to be n = :. However, when the
the fluence dependence of the measured hardening (see data are actually plotted versus &, it can be seen that
refs. [57-591 for reviews). It has also been recently a linear relation does not exist. As shown in fig. 17(b),
proposed that n = a is the appropriate fluence exponent the hardness increase due to irradiation is directly pro-
for radiation hardening [3,4,60]. portional to ($t) ‘I4 following a transition fluence of
Several experimental [52,61] and theoretical [35,62] - 1 X 10” n/m*.
studies have established that the increase in strength of
copper due to neutron irradiation near room tempera- 4.4. Defect production and surviving defect fraction
ture is given by Au - m, where N is the cluster
density and d is the mean defect size. The assumption The fraction of defects which survive recombination
N - +t has commonly been invoked to explain that the events during irradiation depends on several factors. It
yield strength increase should be proportional to the is now well established from irradiations performed at
square root of fluence with no incubation period liquid helium temperatures that athermal incascade re-
[58-59,62,63]. However, plots of do versus fl begin combination can reduce the point defect concentrations
to show curvature at doses well below that where cascade to as low as 30% of the calculated Kin&in-Pease value
overlap should be important. Since this study and others [72-741. The displacement efficiency (5) in copper de-
show that the cluster size is essentially independent of creases from 1.0 to 0.3 as the mean recoil energy in-
fluence, this implies that the fluence exponent for creases, due to enhanced close-pair recombination in
hardening is less than 0.5. The previous discussion on the high energy cascades. The value of .$ is relatively
the fluence dependence of the visible cluster density constant for cascade-producing irradiations such as fast
[N - ( +t)‘/2] suggests that n = a will give the correct and 14-MeV neutrons or energetic heavy ions [18,72-741.
dependence. A reanalysis by this author of the harden- This is due to the formation of subcascades at high
ing data given in refs. [2,56,57,62-711 has shown that primary-recoil energies [l&19,72]. The reduction of point
the change in strength can be consistently described by defects below the calculated Kinchin-Pease number in
S. J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys 155

150 I I
I
I
MITCHELL et al. (4978)

125 259: IRRADIATION /’


0 RTNS
0 E%e(D,n)
(00

2
5
75
2
s

50

25 i
0
0
I - _

I
Jf
0

x 402”
__
f.6 WI” 8.t X 10” I.5r
2

1## win*)

Fig. 16. Change in the yield strength of pure copper plotted versus the one-fourth root of neutron fluence. Data are from Mitchell et
al. [66].

high-energy cascades is thought to occur during the fraction of defects at any temperature.
thermal spike phase (- HI-” s) [75]. For a given irradi- The fraction of surviving point defects from a given
ation spectrum, 6 represents the maximum survival displacement is further reduced by nearly a factor of 2

I I I I

JPCA 240
200 - KINOSHITA et al. 0986)

200
t60 -

160
2
3 120
s
80

40

2
0 i XiP f.6Xi02’ 8.1 X(02’ 2.tW0=

c+tP (n/m*)

Fig. 17. Change in the yield strength of an austenitic stainless steel following 14-MeV neutron irradiation plotted as a function of the
square root (left) and fourth root (right) of fluence. The inset figure on the left shows the original data of Kinoshita et al. [71]
presented as a double logarithmic plot.
156 S.J. Zinkle / Microstructure umipropertwsof neumm irrudruted Cu uIIoI.~

at temperatures where the interstitial is mobile ( r 60 K) Simons [74], the freely migrating defects in copper ;IIC
due to correlated recombination of interstitials and predominantly (4: 1) interstitial for neutron and other
vacancies within a single cascade region [76]. Two re- cascade-producing irradiations. Long-range migration
cent neutron irradiation studies of copper have esti- of interstitials to existing distant cascades can therefore
mated the overall fraction of defects surviving corre- cause shrinkage and eventual annihilation of vacancy
lated recombination (including defects in clusters) to be clusters in these regions. The probability of an intersti-
t 2 0.15 at temperatures of 80 to 300 K [3,76]. Kiritani tial encountering a vacancy cluster (as opposed to
et al. [77] have recently reported a correlated defect another interstitial) increases as the vacancy cluster
survival fraction of e = 0.15 for gold irradiated at room density increases.
temperature with 14-MeV neutrons. The parameter e is Kiritani et al. [77] have found strong evidence of
independent of fluence and represents the maximal long-range interstitial interaction with cascades in a
survival fraction of point defects for temperatures where systematic study that involved irradiations of pre-
at least one of the defect types is mobile. This parame- thinned and bulk specimens. The bulk specimens con-
ter is greater than the commonly measured [76,78,79] tained a lower density of clusters, and the vacancy-type
fraction of freely migrating defects (f), which is also clusters were of smaller size compared to the prethinned
independent of fluence but does not include point de- foils. The density difference became more pronounced
fects that cluster during the cascade quench. with increasing fluence, which indicates that the dose
Long-range uncorrelated recombination of point de- dependence of surviving clusters in a bulk foil irradia-
fects at sinks (interaction with distant cascades) will tion is sublinear. Yoshida et al. [33] have found that
reduce the fraction of surviving defects to a value less small levels of gaseous impurities can cause a significant
than 15%. This recombination effect should not be enhancement of the visible cluster density. This might
termed “cascade overlap” since it occurs at very low be caused by impurity trapping of the freely migrating
fluences and is simply dependent on the relative prob- interstitials, which would reduce the annihilation rate of
ability of an interstitial encountering a vacancy-type vacancy clusters. On the other hand, the addition of
versus interstitial-type point defect or cluster. The substitutional impurities such as Al, Mn, and Ni does
amount of uncorrelated recombination that occurs de- not have any large effect on the surviving total defect
pends on the number of point defects and defect clus- cluster density (figs. 10, 11).
ters that are already existing, and hence is fluence In summary, it appears that the observed nonlinear
dependent. We have empirically found that the number behavior of the defect cluster density versus dose at
of surviving defects in copper is approximately propor- damage levels < 102’ n/m* is not due to physical
tional to the square root of fluence for a room-tempera- overlap of cascades. but is instead attributable to long-
ture irradiation. Therefore the surviving defect fraction range interaction of freely migrating interstitials with
(7) should decrease with increasing fluence. Table 3 existing defect clusters. The empirical relation N - @
gives the surviving defect fraction as determined from is not based on any theoretical models. However, it is a
electrical resistivity measurements made on a copper more realistic physical model than assuming a simple
foil irradiated in this study [5]. Similar values for 17were linear dependence.
obtained from an analysis of the TEM data. Other
studies of neutron-irradiated copper have reported val-
ues for TJof 4 to lo%, but they generally do not address 5. Conclusions
the issue of fluence dependence [2,77]. Kiritani et al.
[77] have found that n = 4 to 5% for gold irradiated to Room-temperature irradiation of copper and the
4 X 10” n/m2. single-phase alloys Cu-5% Al, Cu-5% Mn, Cu-5% Ni
The fluence dependence of the surviving defect frac- with 14-MeV neutrons produces defect clusters that are
tion in pure metals is due to uncorrelated recombina- less than 10 nm in size. There is not an appreciable
tion of point defects at distant cascades. According to effect of alloying on the overall cluster density or size
distribution. However, the alloys contain a significantly
Table 3 lower density of stacking fault tetrahedra and triangle
Fluence-dependent fraction of point defects in copper that dislocation loops compared to pure copper. The lower
survive recombination at room temperature (from ref. [5]) density of these two types of defect clusters in the alloys
may be due to preferential binding of vacancies to
Fluence (n/m2) 1x1020 3x1020 3x102’ 3x102’
solute atoms. There was no TEM evidence of short-range
Surviving fraction, 9 (W) 11.1 10.4 6.6 4.7
ordering in the irradiated alloys, in contrast to previ-
S.J. Zinkle / Microstructure and properties of neutron irradiated Cu alloys 151

ously reported electrical resistivity and microhardness References


results.
The overall defect cluster size was found to be con- (11 R.W. Powell, N.F. Panayotou, G.R. Odette and G.E.
stant over the fluence range of this investigation, 1x Lucas, in: DAFS Quart. Prog. Rep., Jan.-Mar. 1979;
10” to 2 X 102’ n/m2. The total size distribution for all DOE/ET-0065/5, p. 105.
[2] H.R. Brager, F.A. Gamer and N.F. Panayotou, J. Nucl.
four metals was similar to a log-normal distribution,
Mater. 103 & 104 (1981) 995.
with a peak at a diameter of - 1.7 nm. Comparison
[3] S.J. Zinkle and G.L. Kulcinski, J. Nucl. Mater. 122 & 123
with highdose neutron and ion-irradiation size distri- (1984) 449.
butions indicates that the mean cluster size may be [4] S.J. Zinkle and G.L. Kulcinski, in: Proc. Conf. on The
essentially constant up to very high damage levels for a Use of Small-Scale Specimens for Testing Irradiated
room-temperature irradiation. Most of the clusters Material, Eds., W.R. Corwin and G.E. Lucas, Albuquer-
smaller than 2 nm diameter were not visible in foils of que, NM, Sept. 1983, ASTM-STP 888 (1986) p. 141.
thickness greater than 100 nm. Cluster densities should [5] S.J. Zinkle, Electrical resistivitv of small dislocation looos
be compared only under comparable visibility condi- in irradiated copper, submitted to J. Phys. F.
tions. Details of the visible size distribution should be [61 W.B. Pearson, Handbook of Lattice Spacings and Struc-
tures of Metals, Vol. I (1958) and Vol. II (1967) (Per-
included whenever cluster densities are reported.
gamon Press, London, New York).
There was no visible evidence of physical cascade
171 H.W. King, J. Mater Sci. 1 (1966) 79.
overlap occurring for 14-MeV fluences up to 2 x lo*’
PI P.C.J. Gallagher, Met. Trans. 1 (1970) 2429.
n/m2. A simple calculation predicts that cascade over-
191 C.B. Carter and I.L.F. Ray, Philos. Mag. 35 (1977) 189.
lap effects should start to become significant for damage [lOI R.H. Howell, Phys. Rev. B22 (1980) 1722.
levels on the order of 10” n/m2. Ull S.J. Zinkle and G.L. Kulcinski, in: DAFS Quart. Prog.
The visible defect cluster density of all four metals Rep., Oct.-Dec. 1981, DOE/ER-0046/g, p. 48.
was found to be proportional to the square root of (121 M.M. Wilson and P.B. Hirsch, Philos. Mag. 25 (1972) 983.
fluence following a transition period - 1 x 102’ n/m2. 1131 F. Hausermann, K.H. Katerbau, M. Riihle, and M.
Examination of published TEM and yield strength data Wilkins, J. Microscopy 98 (1973) 135.

for copper irradiated between 80 and 400 K indicates 1141 D.J.H. Cockayne, J. Microscopy 98 (1973) 116.
[151 L.J. Chen, K. Seshan, and G. Thomas, Phys. Status Solidi
that this square root dependence between cluster den-
A28 (1975) 309.
sity and fluence is valid over a wide range of experimen-
1161 A.Y. Stathopoulos, Philos. Mag. A44 (1981) 285.
tal conditions. Logarithmic plots should not be used
I171 D.G. Doran and N.J. Graves, Displacement cross-sections
exclusively to determine the fluence dependence of and PKA spectra: Tables and applications, HEDL-TME
processes such as radiation hardening and the number 76-70 (December 1976).
of surviving defect clusters. The actual fluence depen- [I81 T. Muroga, K. Kitajima and S. Ishino, J. Nucl. Mater. 133
dence should be checked by plotting the data versus & 134 (1985) 378.
(@f)“. P91 H.L. Heinisch, Computer simulation of displacement
cascades in copper, HEDL-TME 83-17 (June 1983).
WI H.L. Heinisch, J. Nucl. Mater. 117 (1983) 46.
Acknowledgements 1211 J. Silcox and P.B. Hirsch, Philos. Mag. 4 (1959) 72.
PI C.C. Mathai and D.J. Bacon J. Nucl. Mater. 135 (1985)
173.
The author would like to thank M.W. Guinan for his
v31 A.C. Damask, J. Phys. Chem. Solids 1 (1956) 23.
assistance with designing and performing the irradia- ~241 P.L. Rossiter, J. Phys. Fll (1981) 2105.
tions, R.E. Stoller for stimulating discussions, E.A. WI K. Ziesemer and W. Schiile, Acta Met. 33 (1985) 587.
Kenik for manuscript review, and F. Scarboro for WI W. Pfeiler and R. Reihsner, Phys. Status Solidi (A97
manuscript preparation. The helpful assistance of F.A. (1986) 377.
Garner, C. Rowe, B. Borman, C. Logan, and the RTNS- ~271 M.J. Whelan and P.B. Hirsch, Philos. Mag. 2 (1958) 1303.
II staff is also acknowledged. The foils for this study WI G. Thomas, W.L. Bell and H.M. Otte, Phys. Status Solidi
12 (1965) 353.
were provided by H.R. Brager of the Westinghouse-
Hanford Laboratory. The irradiations were performed ~291 L. Trieb and G. Veith, Acta Met. 26 (1978) 185.
[301 A. Varschavsky and E. Donoso, Met. Trans. Al5 (1984)
while the author was at the University of Wisconsin
1999.
under appointment to the Magnetic Fusion Energy
1311 M. Rihle and J.C. Crump, III, Phys. Status Solidi A2
Technology Fellowship Program. (1970) 257.
1321 J.W. Muncie, B.L. Eyre and CA. English, Philos. Mag.
A52 (1985) 309.
158 S.J. Zinkie / Microstructure andproperties of neutron irradiated Cu al1o.w

[33] N. Yoshida, Y. Akashi, K. Kitajima and M. Kiritani, J. [56] J. Diebl and W. Schilling, in: Proc. 3rd lnt. Conf. on
Nucl. Mater. 133 & 134 (1985) 405. Peaceful Uses of Atomic Energy, Vol. 9 Reactor Materials
[34] Y. Shimomura, H. Yosbida, M. Kiritani, K. Kitagawa and (United Nations, 1965) p. 72.
K. Yamakawa, J. Nucl. Mater. 133 & 134 (1985) 385. 1571 M.J. Makin, in: Radiation Effects, Vol. 37, TMS-AIME
[35] N.M. Ghoniem, J. Alhajji, and F.A. Garner, in: Proc. 11th Conf. Asheville, NC, 1965, Ed. W.F. Sheely (Gordon and
Conf. on Effects of Radiation on Materials, Fds. H.R. Breach, New York, 1967) p. 627.
Brager and J.S. P&n, ASTM STP 782 (1982) 1054. [SS] T.J. Koppenaal and R.J. Arsenault, Met. Rev. (J. Inst.
[36] S.J. Zinkle, G.L. Kulcinski and R.W. Knoll, J. Nucl. Metals) 16 (1971) 175.
Mater. 138 (1986) 46. [59] B.L. Eyre, Int. Metall. Rev. 19 (1974) 240.
[37] CA. English, J. Nucl. Mater. 108 & 109 (1982) 104. [60] H.L. Heinisch, SD. Atkin and C. Martinez, J. Nucl.
[38] A.Y. Stathopoulos, C.A. English, B.L. Eyre and P.B. Mater. 141-143 (1986) 807.
Hirsch, Philos. Mag. A44 (1981) 309. [61] M.J. Makin, F.J. Minter and S.A. Manthorpe, Philos.
1391 B.C. Larson and F.W. Young, Jr., in: Point Defects and Mag. 13 (1966) 729.
Defect Interactions in Metals, Eds. J. Takamura et al. [62] J. Diehl and G.P. Seidel, in: Radiation Damage in Reac-
(University of Tokyo Press, 1982) p. 679. tor Materials, Vol. I (IAEA, Vienna, 1969) p. 187.
[40] B.C. Larson, T.S. Noggle, and J.F. Barhost, in: Advanced [63] H.G. Mohamed, A.M. Hammad and F.H. Hammad,
Photon and Particle Techniques for the Characterization Trans. Indian Inst. Metals 35 (3) (1982) 258.
of Defects in Solid, MRS Symp. Proc. Eds. J.B. Roberto [64] LA. El-Shanshoury, J. Nucl. Mater. 45 (1972/73) 245.
et al., Vol. 41 (MRS, 1985), p. 25. [65] I. Yoshizawa, N. Naramoto and K. Kamada, J. Phys. Sot.
[41] J.A. Horak and T.H. Blewitt, Phys. Status Solidi A9 Japan 36 (1974) 202; K. Kamada et al., Phys. Status Solidi
(1972) 721. Al8 (1973) 377.
(421 R.S. Averback, K.L. Merkle and L.J. Thompson, Radiat. 1661 J.B. Mitchell, R.A. Van Konynenburg, C.J. Ether and
Eff. 51 (1980) 91. D.M. Parkin, Proc. Conf. on Radiation Effects and Tri-
[43] M. Nakagawa et al., Phys. Rev. 816 (1977) 5285. tium Technology for Fusion Reactors, Vol. II, Eds. J.S.
[44] C. Abromeit and S.R. MacEwen, Nucl. Instr. and Meth. Watson and F.W. Wiffen, Gatlinburg, TN (1975) p. 172;
B15 (1986) 770. J.B. Mitchell, Lawrence Livermore Laboratory report
[45] J.B. Roberto et al., Appl. Phys. Lett. 30 (1977) 509. UCRL-52388 (January 1978).
[46] R.M. Walker, in: Radiation Damage in Solids, Proc. Int. [67] H.C. Gonzalez and E.A. Bisogni, Phys. Status Solidi 62
School of Physics (Enrico Fermi) course XVIII, Ed. D.S. (1980) 351.
Billington (Academic Press, London, New York, 1962) p. [68] E.R. Bradley and R.H. Jones, in: The Use of Small-Scale
594. Specimens for Testing Irradiated Material, ASTM STP
[47] H.J. Wollenberg~r, in: Vacancies and Interstitials in 888, Eds., W.R. Corwin and G.E. Lucas (ASTM. Phila-
Metals, Eds. A. Seeger, D. Schumacher, W. Schilling and delphia, 1986) p. 186.
J. Diehl (North-Holland, Amsterdam, 1970) pp. 215--253; 1691 K. Shinohara and S. Kitajima, J. Nucl. Mater. 133 & 134
W. Schilling, ibid., p. 255. (1985) 690.
[48] L. Thompson, G. Youngblood and A. Sosin, Radiat. Eff. [70] A. Okada, T. Yoshiie, S. Kojima and M. Kiritani, J. Nucl.
20 (1973) 111. Mater. 141-143 (1986) 907.
[49] G.P. Scheidler and G. Roth, in: Proc. Int. Conf. on [71] T. Kinoshita, Y. Aono, E. Kuramoto and K. Abe, J. Nucl.
Vacancies and Interstitials in Metals, Jill-Conf-2, Vol. 1 Mater. 141-143 (1986) 893.
(KfA, Jiilich, 1968) p. 391. 1721 R.S. Averback et al., Phys. Rev. B18 (1978) 4156.
[SO] M. Ipohorski and L.M. Brown, Philos. Mag. 22 (1970) 1731 J.H. Kinney, M.W. Guinan and Z.A. Munir, J. Nucl.
931. Mater. 122 & 123 (1984) 1028.
[51] M.J. Makin, A.D. Whapham and F.J. Minter, Philos. [74] R.L. Simons. J. Nucl. Mater. 141-143 (1986) 665.
Mag. 7 (1962) 285. [75] R.S. Averback and M.A. Kirk, in: Surface Alloying by
1.521 1.0. Konstantinov, VS. Kharbarov and V.I. Shcherbak, Ion, Electron and Laser Beams, Eds. L.E. Rehn et al.
Fiz. Khim Obrab Mater. (Phys. and Chem. of Materials (ASM, 1987) p. 91.
Processing) No. 3 (1982) 135-138. [76] U. Theis and H. Wollenberger, J. Nucl. Mater. 88 (1980)
1531 T. Muroga, M. Eguchi, N. Yoshida, N. Tsukuda and K. 121.
Kitajima, J. Nucl. Mater. 141-143 (1986) 865. (771 M. Kiritani, T. Yoshiie and S. Kojima, J. Nucl. Mater.
1541 R.L. Chaplin and R.R. Coltman, Jr., J. Nucl. Mater. 108 141-143.
& 109 (1982) 175. [78] J.A. Goldstone, D.M. Parkin and H.M. Simpson, J. Appl.
[SS] T.H. Blewitt, in: Radiation Damage in Solids, Proc. Int. Phys. 51 (1980) 3690.
School of Physics (Enrico Fermi) Course XVIII, Ed. D.S. [79] L.E. Rehn, P.R. Okamoto and R.S. Averback, Phys. Rev.
Billington (Academic Press, London, New York, 1962) p. B30 (1984) 3073.
630.

You might also like