Sei sulla pagina 1di 10

Formation of persistent dislocation loops by ultra-high

strain-rate deformation during cold spraying


Christine Borchers
*
, Frank Gartner, Thorsten Stoltenho, Heinrich Kreye
Department of Mechanical Engineering, Helmut Schmidt University, D-22039 Hamburg, Germany
Received 27 January 2005; received in revised form 14 February 2005; accepted 17 February 2005
Available online 19 April 2005
Abstract
Pure copper was deformed at ultra-high strain-rates by the technique of cold spraying. It was found that after annealing the hard-
ness of the samples does not fall as low as is known for heavily deformed copper at normal strain rates, e.g., cold rolling. The reason
for this is the presence of defects that do not heal out after annealing at 600 C. These defects were identied as dislocation loops of
the extrinsic and intrinsic kind. Possible mechanisms for the formation of these loops are discussed.
2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Cold spraying; High-resolution electron microscopy; High-speed deformation; Point defects; Hardness
1. Introduction
When copper undergoes heavy deformation and sub-
sequent annealing, the temperature dependence of phys-
ical properties like Vickers hardness and resistivity is
well known, if the strain rate is lower than about 100/s
[1]. In that regime, the mechanisms of formation and
migration of dislocations upon deformation are well
understood [2,3]. High strain-rate deformation of cop-
per has also been studied extensively [412]. Most of
these studies focus on dynamic recrystallization, disloca-
tion arrangement, and deformation twinning. As early
as 1962, Hornbogen reported shock-induced dislocation
loops in aFe [13]. Meyers and co-workers [12] observed
a large number of dislocation loops after laser-induced
shock compression. To our knowledge, no annealing
studies of copper after shock loading have been per-
formed until now.
Cold spraying (CS) is a coating technique, where
bonding of particles is due to extensive plastic deforma-
tion and resulting phenomena at the interface [14
19,10,11]. In this process, solid particles smaller than
50 lm are introduced into a stream of inert gas before
being accelerated in a convergingdiverging de Laval
type nozzle. The particles reach velocities ranging typi-
cally from 500 to 1000 m/s and build up dense coatings
only due to their kinetic energy upon impact. As was
shown in extensive modelling of the process [20], in
the rst 200 ns after impact stresses reach up to about
0.5 GPa and strains up to about 10, corresponding to
strain rates of 10
6
/s to 5 10
9
/s. These values are
strongly dependent on the distance from the interface,
and the location on the circumference of the spherical
impinging particle [20]. In Fig. 1, a Cu particle 50 ns
after impinging on a Cu substrate is sketched.
The contours indicate iso-strains. Strain values in the
contact surface vary continuously and reach their max-
imum at the equator (marked A), whereas they are
smallest at the south pole (S). The quality of the strain
exhibits an analoguous behaviour: at the equator it
has a large amount of shear components, at the south
1359-6454/$30.00 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2005.02.048
*
Corresponding author. Present adress: Institute of Materials
Physics, University of Go ttingen, D-37077 Go ttingen, Germany.
Tel.: +49 551 395 584; fax: +49 551 395 012.
E-mail address: chris@ump.gwdg.de (C. Borchers).
Acta Materialia 53 (2005) 29913000
www.actamat-journals.com
pole it is mostly compressive. Near the equator, where
the temperature approaches the melting point, the mod-
elling suggests that the material loses its shear strength
and there is a decrease in stress going along with a
change from plastic to viscous ow [20]. This adiabatic
shear instability is accompanied by the occurrence of a
fast travelling material jet, as is indicated near A in
Fig. 1, and is well known from, e.g., explosive welding
[2123].
The microstructure of cold sprayed copper coatings
has been extensively studied, showing a wide range of
dierent features [11,24,25], ranging from dislocation
walls to nanosized grains formed during dynamic recrys-
tallization to rather large defect-free grains formed by
recrystallization during cooling of the coatings. This
work will focus on the occurrence of vacancy and self-
interstitial type dislocation loops in CS Cu coatings.
Since such loops were in the past mainly studied after
irradiation of the material [2631], the results obtained
in this study and the discussion in the light of the corre-
sponding literature should supply a more general under-
standing of that type of lattice defect.
2. Experimental
Copper coatings were cold sprayed by using nitrogen
as process and carrier gas. To obtain a maximum depo-
sition eciency and dense coatings, preheating of the
process gas to 300 C and a gas inlet pressure of
3.0 MPa were required. As feedstock, gas atomized
99.8% pure Cu-powder with an oxygen content of less
than 0.1 wt.% in the size range between 5 and 25 lm
was used. More general aspects of the coating technique
are reported in [1416,19]. The coatings were annealed
isochronically for 1 h at 100600 C in steps of 100 to
investigate the thermal inuence on defect density and
coating performance. Moreover, a set of coatings was,
after annealing for 1 h at 600 C, cold rolled to 90% size
reduction and then subjected to the same annealing pro-
gramme again to study the stability of intrinsic defects.
Three states were investigated by transmission electron
microscopy (TEM): the as-sprayed, after 1 h annealing
at 600 C, and the cold rolled state after annealing.
The macroscopic eect of annealing treatment on CS
Cu coatings was studied by Vickers hardness measure-
ments and resistivity measurements, both carried out
at room temperature.
Microstructural investigations were performed by
scanning electron microscopy (SEM), TEM, and high-
resolution transmission electron microscopy (HRTEM).
To investigate the sprayed coatings by TEM and
HRTEM, discs of 3 mm diameter were punched from
sheets of the coatings, and then polished, dimpled, and
ion milled. SEM was performed with a Philips XL40,
TEM was performed with a Philips EM420 ST, and
HRTEM was performed with a Philips CM200 FEG.
Hardness measurements on all samples were performed
according to Vickers under a load of 2.943 N (HV 0.3).
Hardness measurements were carried out under the
standard DIN 50133 (now DIN EN ISO 6507), the val-
ues were gained over 10 readings. The load was chosen
to obtain overall information about the samples; inden-
tations had a diameter of about 20 lm or larger, i.e.,
most indentations covered more than one powder parti-
cle. The standard deviation of hardness measurements
was <2%. Resistivity measurements were performed
with a four-point device using a Digistant 6425 precision
current-generator and a Keithley 181 nanovoltmeter
and a measured section of 5 mm.
3. Results
Fig. 2 shows the microstructure of the feedstock pow-
der. The powder particles are spherical with the size
range given above, see the SEM micrograph Fig. 2(a).
The microstructure of a feedstock particle as obtained
by TEM is shown in Fig. 2(b), where a very high-dislo-
cation density and a grain size of several microns are
revealed.
In Fig. 3(a), the Vickers hardness of dierent Cu sam-
ples after isochronic annealing is shown. The straight
line shows the hardness of an as-sprayed coating, the
dashed line represents the hardness of a bulk Cu sheet,
cold rolled to 90%, and the dotted line represents the
hardness of a CS sample after annealing for 1 h at
600 C and subsequent cold rolling to 90%. All three
curves illustrate a decrease in hardness with rising
annealing temperature. The initial hardness of the as-
sprayed CS coating is somewhat higher than that of
Fig. 1. Sketch of a Cu particle 50 ns after impinging on a Cu substrate.
The contours indicate iso-strains. It can be seen that at the south pole,
marked S, the strain is mostly compressive but appreciably smaller
than near the equator, marked A. Here, the strain is highest, moreover
it is shear strain.
2992 C. Borchers et al. / Acta Materialia 53 (2005) 29913000
the cold rolled sheet or the annealed and rolled CS sam-
ple. The as-sprayed as well as the annealed and rolled
CS samples feature a residual hardness after heat treat-
ment at 600 C that is almost twice as high as the corre-
sponding residual hardness of the cold rolled Cu sheets.
In all three cases, a marked decrease in hardness is ob-
served between 100 and 200 C. Fig. 3(b) shows the
resistivity of CS Cu coatings at RT after dierent iso-
chronic annealing treatments. Both the hardness and
the electrical resistivity fall after annealing, but the resid-
ual resistivity after annealing at 600 C is only about
10% higher than the value for bulk Cu, shown as a
straight horizontal line in the diagram.
The TEM-micrographs, Fig. 4, show the microstruc-
ture of an as-sprayed copper coating. Fig. 4(a) is an
overview. A particleparticle boundary is indicated by
black arrows. The wide range of dierent microstruc-
tural features mentioned above is clearly illustrated.
Some grains are nanosized, whereas others are microm-
eter-sized. Some of the latter, indicated by white arrows,
exhibit coee-bean-like contrast. A close-up of such a
grain can be seen in Fig. 4(b). The coee-beans are
aligned on crystallographic planes. The local density of
these defects can be estimated to 10
23
/m
3
, but it was
not possible to establish a reliable size distribution for
these defects in the as-sprayed state. The regions show-
ing such coee-beans have a very low dislocation den-
sity. The volume fraction of areas with loops can be
estimated as roughly 25%, though as always with
TEM the total examined volume is very small.
In the TEM micrographs of Fig. 5 the microstruc-
ture of a cold sprayed copper coating after heat-treat-
ment for 1 h at 600 C is shown. As shown in the
overview Fig. 5(a), there are no more nanosized
grains, but a rather more recrystallized microstructure
with micrometer-sized grains and recrystallization
twins. Some of the grains exhibit a coee-bean-like
contrast, similar to that of the as-sprayed coating, as
marked with a white arrow. Fig. 5(b) shows a close-
up. The local defect density seems to be somewhat
smaller than in the as-sprayed sample, estimated as
3 10
22
/m
3
.
The TEM micrographs in Fig. 6 show the microstruc-
ture of a coating that was cold rolled to 90% after
annealing for 1 h at 600 C. Here, the grain size is on
Fig. 3. Physical properties of Cu samples at RT and after isochronic
annealing for 1 h at 100600 C in steps of 100 and subsequent
cooling to RT: (a) Vickers hardness. The straight line represents an as-
sprayed coating, the dashed line a bulk sheet cold rolled to 90%, and
the dotted line represents a CS sample after annealing for 1 h at 600 C
and subsequent cold rolling to 90%; (b) RT resistivity of CS Cu. The
resistivity of bulk Cu is given by a straight horizontal line.
Fig. 2. Powder microstructures: (a) SEM image of the Cu feedstock
powder; (b) microstructure of a feedstock particle as shown by TEM.
A very high dislocation density and a grain size of several microns are
revealed.
C. Borchers et al. / Acta Materialia 53 (2005) 29913000 2993
the order of some 100 nm, and the dislocation density is
again very high. Some grains, like the one marked with
an arrow, exhibit a coee-bean-like contrast, very much
like in the as-sprayed and the annealed coatings. A
close-up can be seen in Fig. 6(b). Again the coee-beans
correlate with a low dislocation density. The defect den-
sity can be roughly estimated to be 10
22
/m
3
, although it
is dicult to determine exact values within the small ob-
served volumes.
Fig. 7 shows HRTEM images of two of the coee-
bean-like contrasts in the sample shown in Fig. 6. The
lattice spacing is Cu {111}. In both micrographs, dislo-
cation loops can be seen. Fig. 7(a) shows an intrinsic
loop, Fig. 7(b), an extrinsic one. Both loops have an
extension of about 3.5 nm. Because of restricted tilting
in the electron microscope, it is not possible to deter-
mine unambiguously the exact Burgers vectors and ha-
bit planes of the loops.
4. Discussion
The development of the hardness of deformed bulk
copper upon annealing is well understood. In the tem-
perature range up to 100200 C, the initially high
hardness barely changes, while at temperatures between
100 and 400 C, there is an abrupt decrease in hardness
[1]. This decrease in hardness is attributed to primary
recrystallization [1], and the respective recrystallization
temperatures are determined by purity, the amount of
Fig. 4. TEM-micrographs of the microstructure of an as-sprayed Cu coating. (a) Overview. A particleparticle boundary is indicated by black
arrows. Some grains are nanosized, whereas others are micrometer-sized. Some of the latter, indicated by white arrows, exhibit coee-bean-like
contrast. (b) Close-up of a grain with coee-bean contrast.
2994 C. Borchers et al. / Acta Materialia 53 (2005) 29913000
strain, and the initial microstructure of the copper. The
same abrupt decrease in hardness is also found for CS
Cu, see Fig. 3(a). Here too, primary recrystallization is
observed, as can be seen when Fig. 4(a) (as-sprayed) is
compared to Fig. 5(a) (annealed for 1 h at 600 C). It
was somewhat surprising, however, to nd that the
hardness of cold sprayed copper coatings does not fall
to a value comparable to that of cold rolled bulk Cu
after annealing at 600 C. The distinguishing micro-
structural feature of the as-sprayed coatings compared
to cold rolled bulk metal is the existence of numerous
dislocation loops in the sprayed coatings, even after
heat treatment. This can explain high residual hardness
of the coatings. When dislocation loops are generated
in Cu by irradiation with neutrons, protons or ions,
hardening due to these loops was reported long ago,
e.g., by McReynolds et al. in 1955 [32]. They also per-
formed annealing studies of irradiation hardened Cu,
and found a sharp decrease of the critical shear stress
at about 300 C. Unfortunately, no values are given,
so one can not compare the values to those of unirra-
diated Cu.
Fig. 5. TEM micrographs of a CS Cu coating after heat-treatment of 1 h at 600 C. (a) Overview. There are no more nanosized grains, rather a
recrystallized microstructure with micrometer-sized grains and recrystallization twins. Some have a coee-bean-like contrast, very much like in the as-
sprayed coating, one of which is marked with a white arrow. (b) Close-up of a grain with coee-bean contrast.
C. Borchers et al. / Acta Materialia 53 (2005) 29913000 2995
High-resolution studies of cold rolled coatings, see
Fig. 7, revealed extrinsic and intrinsic dislocation loops,
i.e., agglomerations of interstitials and vacancies. The
loops have an estimated average diameter of about
3.5 nm, or about 200 agglomerated point defects per
loop, which corresponds to a local point defect concen-
tration of the order of 10
5
, keeping in mind that these
loops are not found in every part of the examined coat-
ings. These vacancies and interstitials must be formed
during or immediately after the cold spray process.
Two mechanisms of vacancy/interstitial formation are
conceivable:
1. During cold spraying, there is a temperature rise to
near the melting point at heating rates of about
10
9
K/s [20]. Thermal vacancies might emerge
through the formation of Frenkel pairs. Under strain,
the formation energy E
F
of Frenkel pairs in Cu
decreases from 4.8 to about 2.2 eV [33]. So the follow-
ing estimation gives the possible concentration c
F
of
Frenkel pairs in quenched-up strained Cu:
c
F
exp
S
F
k

exp
E
F
kT

10
9
;
where k is the Boltzmann constant, T = 1200 K the
temperature, and S
F
the formation entropy
S
F
= 2.5k of a Frenkel pair [34]. This value alone can-
not account for the high density of dislocation loops
observed in CS Cu.
2. The process of cold spraying is accompanied by
heavy deformation of the powder particles at extre-
mely high strain rates of around 10
9
/s reaching stres-
ses up to 8 10
11
Pa [20]. This deformation is
eectuated mainly by dislocations. These glide, climb,
and intersect. The latter mechanism leads to jogged
dislocations, the glide of which generates point
defects, which dislocation climbing does as well.
Hirth and Lothe [35] estimated a critical stress r
crit
above which the jogged dislocations can move with-
out the aid of thermal uctuations:
r
crit

E
F
2bla
2:5 10
8
Pa;
Fig. 6. TEM micrographs of a coating that was cold rolled to 90% after annealing for 1 h at 600 C. (a) Overview. Here, the grain size is on the order
of some 100 nm, and the dislocation density is again very high. Some grains, like the ones marked with arrows, exhibit a coee-bean-like contrast,
very much like in the as-sprayed and the annealed coatings. (b) Close-up of a grain with coee-bean contrast.
2996 C. Borchers et al. / Acta Materialia 53 (2005) 29913000
where b = 2.5 10
10
m is the Burgers vector,
l = 10
8
m the mean intersection length for the initial
dislocation density of estimated 10
16
/m
2
, and
a = 2.5 10
10
m a unit step width of a moving dislo-
cation. E
F
2.2 eV is taken as above. This value is at
least two orders of magnitude below the peak stresses
reached in the cold spray process, making the gener-
ation of athermal point defects highly probable. Ko-
vacs and Zsoldos [36] gave a possibility of estimating
the concentration c
P
of point defects formed by these
mechanisms:
c
P

1
2l
Z
s dc 3 10
2
;
where l = 42 GPa is the shear modulus of Cu [37],
and
R
s dc 1=2s
max
c
max
with the maximum shear
stress s
max
= 0.5 GPa and the maximum shear strain
c
max
= 10 taken from Assadi et al. [20]. This estima-
tion seems at rst glance to be very high, but on the
one hand the formation of a high density of vacancies
and interstitials after high strain-rate deformation is
reported for copper [38], and on the other hand one
can expect a large number of these point defects to
Fig. 7. HRTEM images of two of the coee-bean-like contrasts in a cold sprayed copper coating which was annealed for 1 h at 600 C, and then cold
rolled to 90%. The lattice spacing is Cu {111}: (a) intrinsic dislocation loop; (b) extrinsic dislocation loop.
C. Borchers et al. / Acta Materialia 53 (2005) 29913000 2997
be annihilated even in the very short time disposable
during or after the cold spray process.
With both of the proposed mechanisms of point de-
fect formation, most of the defects should recombine.
However, if only 1% of these defects do not recombine
but rather diuse and agglomerate, they most probably
form extrinsic or intrinsic dislocation loops, as observed
in the sprayed copper coatings. In CS Cu coatings, the
presence of dislocation loops is always accompanied
by the absence of straight dislocations. Since the feed-
stock powder has a very high-dislocation density, the
formation of these loops has to go back to the annihila-
tion of straight dislocations, a process that seems to be
related to some special combination of strain, type of
strain, strain-rate and temperature rise in the material.
Since these quantities are not known for a given region
in a coating, and they are in particular highly non-
uniform over the whole coating, any guess on this spe-
cial combination would be quite speculative.
Osetzky and co-workers [26] analysed the congura-
tions of interstitials and their clusters by molecular dy-
namic simulations. They report several dierent loop
congurations with {111} and {110} habit planes,
and Burgers vectors b = 1/3111, Frank loops, and
b = 1/2110, the so-called perfect loops. Frank loops
are sessile, while perfect loops are glissile. All are stable
for a wide range of temperatures [26]. Perfect loops can
dissociate on their glide prism to form extended loops,
i.e., three-dimensional defects with sides composed of
stacking faults. Furthermore, they nd that perfect
loops with {111} habit planes rotate towards {110}
during relaxation. According to continuum mechanics
[36], dislocation loops of these types are stable, if their
radius R is bigger than about 4a, where a is the lattice
constant, which corresponds to 1.5 nm. Dislocation
loops are persistent even after heat treatment [36], if
their diameter exceeds a critical value. For copper, this
value seems to be in the vicinity of 3 nm, since the small-
est observed loops after heat treatment retain this size.
Smaller loops tend to emit vacancies and vanish [36],
hence the observed lower loop density in the annealed
coatings. Puigvi and co-workers [29] studied the interac-
tion between point defects and point defect clusters
(loops) inter alia in copper by computer modelling. They
found that vacancyself-interstitial recombination is
inhibited, and hence the loops are stable. In addition,
they found that single point defects can prevent cluster
mobility, which has been conrmed by Hudson et al.
[39]. The above considerations can explain the presence
of dislocation loops in cold sprayed copper coatings
after heat treatment, even though the exact nature of
the loops is not elucidated in this work.
Quite a lot of work has been done to nd out how dis-
location loops behave during heavy deformation. Rod-
ney and Martin [27] studied the interaction between
moving dislocations and small (R<1 nm) interstitial
Frank and perfect loops in Ni. They found that about
one third of the loops are absorbed by dislocations,
and most of the rest ip their Burgers vector to a direc-
tion parallel to the glide plane of the dislocation. Rodney
[31] found in a subsequent molecular dynamic study per-
formed for Ni that Frank loops can be either unfaulted
according to the reaction a=3111 a=6112 !
a=2110, or sheared and subsequently reconstructed.
Yang and co-workers [28] found that Frank loops may
or may not unfault during plastic deformation of Cu,
while Wolfer et al. [30] report gliding and rotating perfect
loops. All of these studies correspond well with the
observation presented in this work that dislocation loops
are persistent in CS Cu after cold rolling.
After irradiation of Cu, periodic walls of defects,
alternating with walls completely free of defects are ob-
served [40]. The walls are aligned along {001}, i.e., elas-
tically soft planes: they seem to minimize the total elastic
energy of the elastically anisotropic system [40]. Robach
and co-workers [41] studied the tensile deformation of
irradiated Cu in situ in a TEM. They found that de-
fect-free channels appeared only after deformation, but
not under irradiation. Moreover, the observed channels
were arranged along {111} planes, conrming observa-
tions made by Diaz de la Rubia et al. [42] before. In the
case of cold sprayed Cu coatings, no such channeling
was observed, neither after spraying, nor after subse-
quent annealing and cold rolling. One can assume that
the high amount of shear stress in cold spraying as well
as in cold rolling inhibits the formation of such defect-
free walls. However, it must be pointed out that the
microstructure of cold sprayed Cu coatings is highly
non-uniform and some microstructural features like de-
fect-free channels might have merely escaped our
attention.
The dierence in residual hardness between CS Cu
coatings and cold rolled bulk Cu after annealing at
600 C is pronounced. The mechanical properties seem
to be much more sensitive to dislocation loops than
the electrical properties. This is easy to explain. The
agglomeration of point defects reduces the electrical
resistivity of metals as the main free path for electrons
rises. In contrast, the yield strength and consequently
the hardness of metals rises when point defects agglom-
erate to larger faults, as is the case in precipitation hard-
ening and dispersion hardening, since larger defects
represent stronger obstacles to the motion of
dislocations.
5. Conclusions
The present work has shown that during cold spray-
ing, a large number of intrinsic and extrinsic dislocation
loops are formed. The generation of Frenkel pairs dur-
2998 C. Borchers et al. / Acta Materialia 53 (2005) 29913000
ing up-quenching of metals to satisfy the thermody-
namic need for vacancies is conceivable; however, a
numerical estimation of the generated number of Fren-
kel pairs yields a density of only 10
9
. The most proba-
ble formation mechanism for these loops is rather the
agglomeration of point defects generated by dislocation
climb during or immediately after deformation, leading
to annihilation of the latter. This assumption is substan-
tiated by the local absence of straight dislocations,
whereas these do exist in great number where loops
are absent. The observed dislocation loops are persistent
after annealing up to 600 C and subsequent cold roll-
ing, leading to a residual hardness almost twice as high
as that of cold rolled and annealed bulk copper. The
residual resistivity after annealing, when bulk copper is
compared to CS copper, is only about 10% higher than
the value for the bulk. To conclude, shock-deformed
copper and, particularly, CS Cu coatings are not apt
to soften upon heat treatment as bulk copper is, while
the resistivity, after annealing, is only 10% above bulk
value, which are facts that might be important for engi-
neering applications.
Acknowledgments
We thank Hamid Assadi for enlightening and fruitful
discussions. This work was supported by the Deutsche
Forschungsgemeinschaft under the Grant No. KR
1109/3-2, which is gratefully acknowledged.
References
[1] Cahn RW. Recovery and recrystallization. In: Cahn RW, Haasen
P, editors. Physical metallurgy. Amsterdam: North-Holland;
1996. p. 2401.
[2] Argon AS. Mechanical properties of single-phase crystalline
media: deformation at low temperatures. In: Cahn RW, Haasen
P, editors. Physical metallurgy. Amsterdam: North-Holland;
1996. p. 1878.
[3] Argon AS. Mechanical properties of single-phase crystalline
media: deformation in the presence of diusion. In: Cahn RW,
Haasen P, editors. Physical metallurgy. Amsterdam: North-Hol-
land; 1996. p. 1958.
[4] Shiori J, Satoh K, Nishimura K. Experimental studies on the
behaviour of dislocations in copper at high rates of strain. In:
Kawata K, Shiori J, editors. High velocity deformation of
solids. Berlin: Springer; 1978. p. 50.
[5] Stelly M, Dormeval R. Some results on the dynamic deformation
of copper. In: Kawata K, Shiori J, editors. High velocity
deformation of solids. Berlin: Springer; 1978. p. 82.
[6] Andrade U, Meyers MA, Vecchio KS, Chokshi AH. Acta Metall
Mater 1994;42:3183.
[7] Meyers MA, LaSalvia JC, Nestorenko VF, Chen YJ, Kad BK.
Dynamic recrystallization in high strain rate deformation. In:
McNelley TR, editor. Proceedings of ReX96: the third interna-
tional conference on recrystallization and related phenom-
ena. Monterey (CA): Monterey Institute of Advanced Studies;
1997. p. 279.
[8] Atroshenko SA. Shock induced recrystallization in metals. In:
Gottstein G, Molodov DA, editors. Recrystallization and grain
growth, vol. 2. Berlin: Springer; 2001. p. 931.
[9] Meyers MA, Benson DJ, Vo hringer O, Kad BK, Xue Q, Fu HH.
Mater Sci Eng A 2002;322:194.
[10] Borchers C, Stoltenho T, Gartner F, Kreye H, Assadi H. Mater
Res Soc Symp Proc 2002;673:7.10.
[11] Borchers C, Stoltenho T, Gartner F, Kreye H, Assadi H. J Appl
Phys 2003;93:10064.
[12] Meyers MA, Gregori F, Kad BK, Schneider MS, Kalantar DH,
Remington BA, et al. Acta Mater 2003;51:1211.
[13] Hornbogen E. Acta Metall 1962;10:978.
[14] Alkhminov AP, Klinkov SV, Kosarev VF, Papyrin AN. J Appl
Mech Technol Phys 1997;38:324.
[15] Alkhminov AP, Kosarev VF, Papyrin AN. J Appl Mech Technol
Phys 1998;39:318.
[16] Papyrin AN, Klinkov SV, Kosarev VF, Alkhminov AP. Exper-
imental study of interaction of supersonic gas jet with a substrate
under cold spray process. In: Bernd CC, Khor KA, Lugscheider
EF, editors. Thermal spray 2001: new surfaces for a new
millennium. Materials Park (OH): ASM International; 2001. p.
423.
[17] Vlcek J, Huber H, Voggenreiter H, Fischer A, Lugscheider E,
Hallen H, et al. Kinetic powder compaction applying the cold
spray process a study on parameters. In: Bernd CC, Khor KA,
Lugscheider EF, editors. Thermal spray 2001: new surfaces for a
new millennium. Materials Park (OH): ASM International;
2001. p. 417.
[18] Stoltenho T, Kreye H, Richter HJ, Assadi H. Optimization of
the cold spray process. In: Bernd CC, Khor KA, Lugscheider
EF, editors. Thermal spray 2001: new surfaces for a new
millennium. Materials Park (OH): ASM International; 2001. p.
409.
[19] Stoltenho T, Richter H, Kreye H. J Thermal Spray Technol
2002;11:542.
[20] Assadi H, Gartner F, Stoltenho T, Kreye H. Acta Mater
2003;51:4379.
[21] Kreye H. Wdg J Res Suppl 1977;56:154.
[22] Kreye H, Wittkamp I, Richter U. Z Metallkd 1976;67:141.
[23] Hammerschmidt M, Kreye H. Microstructure and bonding
mechanism in explosive welding. In: Meyers MA, Murr LE,
editors. Shock waves and high strain-rate phenomena in met-
als. New York: Plenum Press; 1981. p. 961.
[24] McCune RC, Donlon WT, Popoola OO, Cartwright EL. J
Thermal Spray Technol 2000;9:73.
[25] Borchers C, Stoltenho T, Gartner F, Kreye H. J Appl Phys
2004;94:4288.
[26] Osetzky YN, Serra A, Singh BN, Golubov SI. Philos Mag A
2000;80:2131.
[27] Rodney D, Martin G. Phys Rev B 2000;61:8714.
[28] Yang Y, Abe H, Sekimura N. Phys Lett A 2003;315:293.
[29] Puigvi MA, Serra A, de Diego N, Osetsky YN, Bacon DJ. Philos
Mag Lett 2004;84:257.
[30] Wolfer WG, Okita T, Barnett DM. Phys Rev Lett
2004;92:085507.
[31] Rodney D. Acta Mater 2004;52:607.
[32] McReynolds AW, Augustyniak A, McKeown M. Phys Rev
1955;98:418.
[33] Sato K, Yoshiie T, Satoh Y, Xu Q, Kuramoto E, Kiritani M. Rad
E Def Sol 2002;157:171.
[34] Dederichs PH, Zeller R. Dynamical properties of point defects in
metals. In: Ho hler G, editor. Point defects in metals II. Ber-
lin: Springer; 1980. p. 160.
[35] Hirth JP, Lothe J. Theory of dislocations. New York: McGraw-
Hill; 1968. p. 545.
[36] Kovacs I, Zsoldos L. Dislocations and plastic deforma-
tion. Oxford: Pergamon; 1973. p. 249.
C. Borchers et al. / Acta Materialia 53 (2005) 29913000 2999
[37] Frost HJ, Ashby MF. Deformation mechanism
maps. Oxford: Pergamon; 1982. p. 21.
[38] Kiritani M, Satoh Y, Kizuka Y, Arakawa K, Ogasawara Y, Arai
S, et al. Philos Mag Lett 1999;79:797.
[39] Hudson TS, Dudarev SL, Sutton AP. Proc Roy Soc Lond A
2004;460:2457.
[40] Jager W, Ehrhart P, Schilling W, Dworschak F, Gadalla AA,
Tsukuda N. Mater Sci Forum 1987;1518:881.
[41] Robach JS, Robertson IM, Wirth BD, Arsenlis A. Philos Mag A
2003;83:955.
[42] Diaz de la Rubia T, Zbib HM, Khraishi TA, Wirth BD, Victoria
M, Caturia MJ. Nature 2000;406:874.
3000 C. Borchers et al. / Acta Materialia 53 (2005) 29913000

Potrebbero piacerti anche