Sei sulla pagina 1di 10

Materials Science and Engineering A 460461 (2007) 204213

Thermal stability of ECAP processed pure copper


X. Molodova
a,
, G. Gottstein
a
, M. Winning
b
, R.J. Hellmig
c
a
Institut fuer Metallkunde und Metallphysik, RWTH Aachen, 52056 Aachen, Germany
b
Max-Planck-Institut fuer Eisenforschung GmbH, 40237 Duesseldorf, Germany
c
Institut fuer Werkstoffkunde und Werkstofftechnik, TU Clausthal, 38678 Clausthal, Germany
Received 18 December 2006; received in revised form 9 January 2007; accepted 12 January 2007
Abstract
Microstructure and texture evolution of pure copper (99.95%) after equal-channel angular pressing (ECAP) with route B
c
up to 12 passes
and subsequent heat treatment were investigated. After deformation the samples were annealed at different temperatures. The deformed and
annealed states were characterized by X-ray pole gures, electron back scatter diffraction (EBSD), TEM and microhardness tests. It was shown
that the observed texture and microstructure changes during subsequent annealing have to be associated with discontinuous recrystallization. The
study revealed a very low thermal stability of this ECAP processed pure copper. The observed decrease of the apparent activation energy for
recrystallization is most likely due to improved nucleation conditions of ECAP deformed material.
2007 Elsevier B.V. All rights reserved.
Keywords: ECAP; EBSD; TEM; Texture; Recrystallization
1. Introduction
Equal-channel angular pressing (ECAP) has received consid-
erable attention recently because it provides an easy means to
produce an ultrane grain size in a bulk material for improved
mechanical properties.
Despite muchactivityinthis eld, informationontexture evo-
lution during heat treatment is surprisingly sparse as compared
to other aspects of ECAP [13].
Comparatively little is known about the mechanisms of
microstructural change during annealing. A nonuniform coars-
ening leading to the appearance of a duplex microstructure was
reported by several authors [47]. This is usually attributed to
abnormal grain growth, but equally well it can be due to discon-
tinuous recrystallization.
In the present study, microstructure and texture evolution
of pure copper were investigated with respect to texture and
microstructure after ECAP deformation and subsequent heat
treatment.
The deformed and annealed states were characterized by X-
ray pole gures, EBSD, TEM and microhardness tests. Texture
formation and microstructural changes were recorded and com-

Corresponding author. Tel.: +49 241 8020257; fax: +49 241 8022301.
E-mail address: molodova@imm.rwth-aachen.de (X. Molodova).
pared to the behavior of cold rolled material. The observed
phenomena were analyzed with regard to recrystallization and
grain growth mechanisms.
2. Experimental details
Pure copper with an impurity content of 413 ppmwas used in
the present study. Prior to ECAP deformation, the material was
annealed for 2 h at 450

Cin order to obtain a fully recrystallized


homogeneous microstructure with a grain size of about 20 m.
The corresponding texture comprised mainly the cube orienta-
tion, whichis a typical recrystallizationtexture after deformation
like rolling or extrusion with orthorhombic specimen geome-
try (Fig. 1). Specimens with 10 mm10 mm cross-section and
60 mm length were machined from these annealed bars.
ECAP deformation was carried out at room temperature to
up to 12 passes with a 90

die angle corresponding to a total


equivalent strain of about 13.8 using route B
c
, that includes a
90

rotation of the billet around its longitudinal axis after each


pass with a constant sense of rotation.
After ECAP deformation isochronal annealing was carried
out on select samples at various temperatures ranging from 100
to 300

C in a salt bath, in which the sample was held at each


temperature for 10 min. In addition, samples were subjected to
isothermal annealing at three different temperatures.
0921-5093/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.01.042
X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213 205
Fig. 1. (a) Texture and (b) microstructure of the initial material.
The crystallographic texture and microstructure evolution
during ECAP and subsequent annealing was studied on sam-
ples sliced parallel to the direction of deformation (extrusion
direction, ED) (Fig. 2). All X-ray texture measurements of
ECAP processed samples were performed with Co K radia-
tioninbackreectionmode. Fromthe measured{1 1 1}, {2 0 0},
{2 2 0}and{1 1 3}pole gures, the ODFwas calculatedbyusing
the series expansion method with positivity correction (Dahms
method) [8,9]. The microstructural variation in the samples was
observed by automated indexing of electron back scatter diffrac-
tion (EBSD) patterns in a Leo 1530 FEG scanning electron
microscope. TEManalysis was carried out on a JEM2000FXII.
The microhardness was measured by a Vickers diamond pyra-
midal indenter with a load of 200 g. The scatter of the measured
hardness values was within 6%.
3. Results
3.1. ECAP processed samples
3.1.1. Texture
Fig. 3 shows the calculated ODFs after 1, 2, 4, 8 and 12
ECAP passes. For brevity only the
2
=45

section of the
ODF is shown. This section contains all ideal components of
simple shear textures, which are listed in Table 1 [10]. Bold
points in Fig. 3(a) denote the ideal orientations of simple shear
textures.
As seen in Fig. 3(a), the texture after one pass is very sim-
ilar to that in simple shear, which can be represented by two
bers, referred to as A: {1 1 1} u v w and B: {h k l} 1 1 0
Fig. 2. Denition of reference axes ED, TD, ND and measurement plane.
Fig. 3.
2
=45

section of the ODF after (a) 1, (b) 2, (c) 4, (d) 8 and (e) 12
ECAP passes.
Table 1
Ideal components of simple shear textures [10]
Component Miller indices
A (1 1

1)[1

1 0]

A (

1

1 1)[

1 1 0]
B (1 1

2)[1

1 0]

B (

1

1 2)[

1 1 0]
C (1 0 0)[0

1 1]
A

1
(1 1

1)[2

1 1]
A

2
(1

1 1)[

2

1 1]
{1 1 1} bre {1 1 1}u v w
1 1 0 bre {h k l}1 1 0
206 X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213
Fig. 4. Microhardness evolution as a function of pass number.
(Table 1). It is apparent that during pass one all ideal com-
ponents of simple shear textures had already developed. A
more detailed analysis of the ODFs (Fig. 3), however, shows
slight deviations from the exact positions of the components.
Some ECAP texture components (A

1
) were rotated by about
5

away from the ideal components of simple shear texture. In


the following, the ECAP texture components will be termed
like the simple shear components but using the index E. The
strongest component after the rst pass was the C
E
compo-
nent {1 0 0} 0 1 1 (rotated cube), which showed a maximum
intensity of about 18. After two ECAP passes a signicant
variation of the texture could be observed (Fig. 3(b)). The
major orientations changed within the bres along . The max-
imum ODF intensity increased to about 30 and was located
at the shifted

B
E
orientation (
1
, ,
2
) =(225

, 45

, 45

).
With growing number of passes the texture type remained
the same. After four passes a distinct weakening of all tex-
ture components was evident. During subsequent deformation
the texture intensity increased again. The strongest compo-
nent after 12 ECAP passes with a maximum intensity of
about 20 became an orientation between the C
E
and B
E
components.
3.1.2. Microhardness
Fig. 4 shows the hardness evolution with growing number of
ECAP passes. The hardness increased during the rst pass, but
Fig. 6. Evolution of grain boundary character during ECAP deformation.
saturated after four passes. After 12 passes the microhardness
was about 125 HV.
3.1.3. Microstructure
The microstructure of the material after 1 and 12 passes is
presented in Fig. 5. Boundaries characterized by misorientations
less than 15

are classied as low angle boundaries (LABs).


Boundaries with greater misorientations are termed high angle
boundaries (HABs). Elongated grains developed after the rst
ECAP pass. The fraction of high angle boundaries was about
15%. After 12 passes an almost equiaxed microstructure with
a mean grain size of about 440 nm was established. However,
with about 50% the fraction of LABs remained relatively high
(Fig. 6).
In order to obtain a more detailed information on the dis-
location structure of the deformed state TEM investigations
were conducted on selected specimens. After one pass a lay-
ered structure of almost parallel and long subgrains appeared
similar as revealed by EBSD (Fig. 7). A very high density of
dislocations was evident. After 12 passes the subgrain (cell)
structure became more equiaxed, but some elongated features
were still visible (Fig. 8). The subgrain interior contained fewer
dislocations, and the boundaries became sharper. Neverthe-
less, many dislocations could be observed at the boundaries as
well as inside the grains. In general, the structure appeared
to be very inhomogeneous; the dislocation density remained
Fig. 5. Microstructure after 1 (left) and 12 (right) ECAP passes.
X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213 207
Fig. 7. (a and b) TEM micrographs after one ECAP pass at different magnications.
very high even after 12 passes. Furthermore, some mechan-
ical twinning was observed at a higher number of passes
(Fig. 8(d)).
3.2. Microstructure and texture evolution during annealing
after ECAP
3.2.1. Texture
Fig. 9(a) shows representative textures of select specimens
after one ECAP pass and isothermal annealing at 200

C.
Annealing for 3 min caused no signicant change of the tex-
ture compared to the deformed state, just a slight weakening of
the texture could be observed. During annealing for 30 min the
texture continued to randomize and after 5 h the components
of the deformation texture had almost completely disappeared.
After two, four and eight passes the specimens showed a similar
annealing behavior with respect to the texture evolution, but this
is not presented here.
For comparison the texture evolution during isothermal
annealing at 200

Cfor 12 pass samples is presented in Fig. 9(b).


Also in this case texture randomization took place similar to
one ECAP pass samples. By contrast, some new components
appeared at longer annealing times ((
1
, ,
2
) =(0

, 15

, 45

)).
Furthermore, the rate of texture randomization (degradation of
typical texture components) increased with progressing defor-
mation.
3.2.2. Microhardness
The effect of isochronal annealing on microhardness can be
seen in Fig. 10(a). As the number of passes increased the tem-
perature for the onset of softening dropped from approximately
170 to 100

C. The samples after 8 and 12 passes underwent vir-


tually the same development. A comparison of ECAP deformed
material with cold rolled samples that had undergone a thick-
ness reduction of 86.9%, resulting in an equivalent strain of
about 2 (corresponding approximately to two ECAP passes),
revealed that the softening of the cold rolled material began at
slightly lower annealing temperatures than for the one ECAP
pass sample. Furthermore, the hardness plateau attained after
softening increased with increasing number of passes. At tem-
peratures where this hardness plateau was reached the grain size
of the annealed one pass samples grew from 5.5 to 6.4 m dur-
ing annealing at 230 and 300

C, respectively; by contrast, for


12 pass samples the mean grain size changed from1.2 to 1.5 m
when annealed at 150 or 280

C.
During isothermal annealing at 150, 200 and 280

C similar
trends were observed (Fig. 10(b)). With increasing temperature
the softening of the material was accelerated.
Fig. 8. (ad) TEM micrographs after 12 ECAP passes at different magnications.
208 X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213
Fig. 9.
2
=45

sections of the ODF after (a) 1 and (b) 12 passes and isothermal annealing at 200

C.
Fig. 10. Microhardness evolution during (a) isochronal annealing for 10 min and (b) isothermal annealing at 150

C.
3.2.3. Microstructure
Irrespective of the number of passes a similar microstructure
evolution could be observed during annealing characterized by
the appearance of new larger grains in the deformed structure
(Fig. 11). In the course of annealing the fraction of these larger
grains grew at the expense of the deformed microstructure, and
this proceeded faster with increasing number of ECAP passes.
The corresponding TEM micrographs (Fig. 12) revealed large
defect-free grains in the deformed matrix, and nally the ini-
tial (deformed) microstructure became totally replaced by these
large grains. The annealed grain size decreased with growing
number of passes (Fig. 13).
4. Discussion
4.1. Development of the deformation texture
The study demonstrates that the texture development for
ECAP route B
c
after the rst pass is very similar to a simple
shear texture. However, there are also slight but distinct devi-
ations from the ideal positions of the shear components. This
shift of the texture components has been reported by several
authors [1012] and seems to result from the geometry of the
deformation process, strain hardening and friction conditions.
The texture intensity initially grewduring the rst two passes
(Fig. 14). In the second pass, new major orientations were cre-
ated, which has to be attributed to a change of the strain path
owing to the 90

axial rotation of the billet. With this rota-


tion shearing takes place in two planes. Each individual shear
is reversed after shearing the billet in the alternate plane, i.e.
the third extrusion reverses the rst shear, the fourth reverses
the second, etc. Therefore, the strain becomes reversed every
four extrusion passes. This has, of course, a signicant impact
on texture evolution, involving other texture components every
pass. Furthermore, this is apparently the reason for the observed
weak texture after four ECAP passes because of the virtually
total reversal of the shear each 360

rotation. If the shear would


be perfectly reversible, the initial texture would be restored after
every four passes. This is obviously not the case as evident from
the observed texture evolution and apparently due to the fact
X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213 209
Fig. 11. (ad) Microstructure evolution during annealing.
Fig. 12. (ac) TEM micrographs after 12 ECAP passes and annealing at 120

C for 10 min.
Fig. 13. Microstructure evolution during annealing at 280

C for 10 min after (a) 1 pass and (b) 12 passes (note different magnications).
210 X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213
Fig. 14. Texture intensity evolution during ECAP.
that the reverse shear does not fully compensate the forward
shear.
The current results show both similarities and signicant dif-
ferences to other ECAP studies on Cu with route B
c
. Li et al.
[13] observed a similar evolution of the texture intensity but
their texture characteristics were signicantly different. In our
study, the major texture component at higher number of passes
is located between B
E
and C
E
orientations, whereas they found
a texture maximum close to the A

1E
orientation. Gazder et al.
[3] investigated the texture evolution of pure Cu (99.98 wt.%
purity) up to four ECAP passes using route B
c
. They found a
general weakening of the maximum texture intensity after four
passes similar to our observations.
4.2. Deformed and annealed microstructure
A very ne, relatively equiaxed grain structure (grain size
0.44 m) was obtained after 12 passes (Fig. 5). However, the
fraction of LABs remained high and amounted to about 50%
(Fig. 6). TEMinvestigations revealed that even at higher number
of passes the microstructure remained inhomogeneous with a
high dislocation density (Fig. 8).
A discontinuous change in texture, microstructure and hard-
ness for Al and Cu based materials was observed during heat
treatment also by other authors [1,47,1216], and usually asso-
ciated with discontinuous grain growth. The microstructural
changes in pure Cu, i.e. the replacement of deformed microstruc-
ture by coarser defect-free grains, can occur during aging even
at room temperature as recently reported by Estrin et al. [17].
Similar to the results of the current study, Estrin et al. also
observed an accelerated removal of the deformed microstructure
with increasing number of ECAP passes.
The interpretation of the observed microstructural changes
during annealing in terms of grain growth is based on the tacit
assumption that the deformed structure is a perfect polycrystal
with submicron grain size. In such case grain growth is liable
to occur and this grain growth whether normal or abnormal
is driven by boundary curvature. However, the ECAP deformed
structure is far from a perfect polycrystal. As a matter of fact, it
contains a high dislocation density and comprises a high fraction
of low angle grain boundaries as any typical deformation struc-
ture does. Such a structure is known to undergo recrystallization
which is distinctly different from (abnormal) grain growth since
it is driven by a volume energy differential caused by the defect
structure (dislocation structure) and not by boundary curvature.
With a dislocation density

=10
16
m
2
as typical for a heav-
ily deformed material the driving force is p
dis

=12 MPa. For a


grain size of d =0.5 m and an average grain boundary energy
of 0.5 J/m
2
the driving force is p
GB

=3 MPa, i.e. substantially


smaller but of comparable magnitude to the dislocation energy.
p
total
= p
dis
+p
GB
=
1
2
Gb
2
+
3
d
12 +3 MPa
= 15 MPa (G
Cu
40 GPa, b
Cu
= 2.5 10
10
m)
While the driving force contribution from the (sub)grain bound-
aries is comparable to the driving force due to dislocations, the
physical difference between (discontinuous) recrystallization
and discontinuous grain growth manifests itself in the fact that
recrystallization is not curvature-driven, in fact during recrys-
tallization a boundary typically moves against its curvature.
Therefore, according to the essential features of the deformed
microstructure andthe microstructural changes duringannealing
the ECAP structure undergoes primary static recrystallization
during annealing and not discontinuous grain growth.
4.3. Recrystallization kinetics
Since Cu shows only negligible tendency to static recovery,
the hardness change during annealing can be entirely asso-
ciated with softening by recrystallization. The corresponding
change of the recrystallized volume fraction X with anneal-
ing time t (Fig. 15(a)) can be evaluated according to the
JohsonMehlAvramiKolmogorov (JMAK) theory [18,19]
X = 1 exp

t
t
R

(1)
where t
R
is the characteristic time for recrystallization
X(t
R
)

=63%, and q is the Avrami exponent. Since recrystalliza-


tion comprises thermally activated phenomena, t
R
will depend
on temperature T via a Boltzmann-term (Fig. 15(b))
t
R
exp

Q
R
kT

(2)
where Q
R
is the apparent activation energy for recrystallization.
As can be seen from Fig. 15(c), Q
R
decreases dramatically
with increasing number of ECAP passes. Similar results were
also obtained for rolled Cu by Mengelberg et al. [20], however,
much less pronounced.
To understand this decrease of activation energy one has to
realize that the apparent activation energy is actually a weighted
average of the activation energy for nucleation and grain bound-
ary motion. In a time increment dt

, the not yet recrystallized


volume decreases by the growth of all previously nucleated
grains with the nucleation rate

N
dX
1 X
= dt

N()v
3
(t

)
2
d 4 (3)
X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213 211
Fig. 15. (a and b) Recrystallization kinetics and (c) apparent activation energy.
After integration
X = 1 exp

4v
3

t
0

N()(t

)
2
d dt

(4)
For site saturation, i.e.

N(t) = N (t)

t
0

0
N ()(t

)
2
d dt

= N
t
3
3
(5)
or
X = 1 exp

4
3
Nv
3
t
3

(6)
By comparison with Eq. (1) we nd
t
R
=
1
3

4
3
Nv
3
(7)
and with the growth rate
v = v
0
exp

Q
v
kT

(8)
follows
t
R
exp

3Q
v
3kT

(9)
and thus Q
R
=Q
v
. For any other dependency of the nucleation
rate on annealing time that can be expressed by a power law

N = At
n
(10)
and therefore
X = 1 exp(4v
3

t
0

0
A
n
(t

)
2
d dt

)
= 1exp

4v
3

t
0
At

n+3

1
n +1

2
n +2
+
1
n +3

dt

(11a)
or
X = 1 exp

4v
3
A
n +4
t
n+4

1
n +1

2
n +2
+
1
n +3

(11b)
For n =0, i.e. constant

N and with
A =

N
0
exp

N
kT

(12)
t
R
exp

N
+3Q
v
4kT

(13a)
or
Q
R
=
Q

N
+3Q
v
4
(13b)
Typically, Q
v
Q

N
1 eV for copper, and in fact, Q
R
is
usually found in the order of Q
R
1 eV [20]. In a severely
deformed material with a large fraction of high angle boundaries,
there is virtually no need for thermal activation of the nucle-
ation process, since the boundaries are mobile and can be easily
set off. In the limit Q

N
0 we would obtain Q
R
0.75 eV,
i.e. still higher than the experimental result Q
R
=0.68 eV after
12 passes (Fig. 15(c)). The only way to decrease the activation
energy in this approximation is an increase of n, i.e. an accel-
eration of the nucleation rate during recrystallization. In such
case, however, also the Avrami exponent q should be higher. In
fact, when comparing the recrystallization kinetics for differ-
ent numbers of ECAP passes, the Avrami exponent increases
with increasing number of passes (Fig. 16) when the devel-
opment of the recrystallized volume fraction is considered for
Fig. 16. Evolution of the Avrami exponent q with increasing number of ECAP
passes.
212 X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213
t <t
R
, i.e. where the recrystallized grains can essentially nucle-
ate and grow without constraints of impingement, whereas for
t t
R
, the Avrami exponent is q 1 virtually independent of
the number of passes. However, the time t <t
R
is more rele-
vant for the characteristic processes since the activation energy
is measured for t =t
R
and thus, dominated by the development
during t <t
R
. In the limit Q

N
0 and n =1, Q
R
will even tend
to Q
R
(3/5)Q
v
=0.6 eV. Therefore, in principle, the lowappar-
ent activation energy for recrystallization can be attributed to an
increasing ease of nucleation, probably owing to the high density
of high angle boundaries.
Of course, this is only a hypothesis supported by select exper-
imental results and needs experimental conrmation. Adecrease
of the apparent activation energy for recrystallization might also
be due to a decrease of the activation energy for grain boundary
migration. Even a slight decrease of Q
v
would markedly lower
Q
R
because of its large weight in Eqs. (13a) and (13b). Since
the mechanisms of grain boundary migration are only poorly
understood it is an open question what might govern the rele-
vant thermally activated process that would be affected by large
strain deformation. The increased boundary free volume during
dislocation absorption or an increase of the vacancy concentra-
tion may be instrumental in facilitating boundary migration, but
this is only a supposition without experimental or conceptional
support so far, and specic experiments will be necessary to shed
light on this issue.
Compared to rolled samples ECAP deformed copper stands
out by a surprisingly lowthermal stability. In viewof the analysis
given above this may be due to the improved nucleation con-
ditions. However, in other materials ECAP processed samples
showa thermal stabilitysuperior torolledspecimens [21], hence,
this cannot be the only criterion that controls the propensity for
recrystallization.
4.4. Strength of the recrystallized material
The decrease of recrystallized grain size with growing num-
ber of passes (Fig. 13) has to be attributed to a higher number
of nucleation sites. The ner recrystallized grain size causes
an increase of the hardness plateau value after recrystallization
Fig. 17. Microhardness as a function of annealed grain size at 280

C.
softening with growing number of passes (Fig. 10) and follows
the well-known HallPetch relationship,
H
v
= H
0
+Kd
1/2
(14)
where H
v
is the Vickers hardness, d the grain size and K and H
0
are material constants (Fig. 17).
5. Summary
The texture and microstructure evolution of pure copper
(99.95%) samples deformed by ECAP up to 12 passes using
route B
c
was analyzed with progressing deformation and after
subsequent annealing.
1. The observed deformation textures were very similar to sim-
ple shear textures. The strongest component at higher number
of passes was located between the B
E
and C
E
orientations.
2. A very ne equiaxed grain structure (with a grain size of
about 0.44 m) was obtained by severe deformation of cop-
per. The deformedmaterial compriseda fractionof highangle
boundaries of about 50%, which renders also a high fraction
of low angle boundaries. Even at higher number of passes
the dislocation density remained very high.
3. The effect of heat treatment on texture and microstructure
evolution especially with respect to the thermal stability of
the ECAP deformed samples was analyzed and compared
to the annealing behavior of cold rolled samples. It was
found that the observed softening and degradation of the
deformation texture components can be associated with the
appearance of larger grains, which are shown to be due to
discontinuous recrystallization.
4. This ECAP processed material showed a lower thermal sta-
bility than rolled material. The apparent activation energy for
discontinuous recrystallization of ECAP deformed samples
was found to substantially decrease with increasing number
of passes. This is most likely due to improved nucleation
conditions with increasing number of ECAP passes owing to
the growing fraction of high angle grain boundaries.
Acknowledgements
The authors would like to thank the Deutsche Forschungs-
gemeinschaft for nancial support through the research unit
Mechanical properties and interfaces in ultra-ne grained mate-
rials (TP 4-Go 335/31-1(2)).
References
[1] S. Ferrasse, V.M. Segal, F. Alford, Mater. Sci. Eng. A372 (2004) 235.
[2] M. Haouaoui, K.T. Hartwig, E.A. Payzant, Acta Mater. 53 (2005) 801.
[3] A.A. Gazder, F. Dalla Torre, C.F. Gu, C.H.J. Davies, E.V. Pereloma, Mater.
Sci. Eng. A415 (2006) 126.
[4] W.Q. Cao, A. Godfrey, W. Liu, Q. Liu, Mater. Sci. Eng. A360 (2003) 420.
[5] C.Y. Yu, P.L. Sun, P.W. Kao, C.P. Chang, Mater. Sci. Eng. A366 (2004)
310.
[6] D.G. Morris, M.A. Mu noz-Morris, Acta Mater. 50 (2002) 4047.
[7] H. Mughrabi, H.W. H oppel, M. Kautz, R.Z. Valiev, Z. Metallkd. 94 (2003)
1079.
X. Molodova et al. / Materials Science and Engineering A 460461 (2007) 204213 213
[8] M. Dahms, H.J. Bunge, J. Appl. Crystallogr. 22 (1989) 439.
[9] M. Dahms, H.J. Bunge, Textures Microstruct. 10 (1988) 21.
[10] S. T oth, R.A. Massion, L. Germain, S.C. Baik, S. Suwas, Acta Mater. 52
(2004) 1885.
[11] S. Suwas, S. T oth, J. Fundenberger, A. Eberhardt, W. Skrotzki, Scripta
Mater. 49 (2003) 1203.
[12] S. Li, I.J. Beyerlein, C.T. Necker, D.J. Alexander, M. Bourke, Acta Mater.
52 (2004) 4859.
[13] S. Li, I.J. Beyerlein, D.J. Alexander, S.C. Vogel, Acta Mater. 53 (2005)
2111.
[14] R.K. Islamgaliev, F. Chmelik, R. Kuzel, Mater. Sci. Eng. A237 (1997) 43.
[15] S. Poortmans, B. Verlinden, Mater. Sci. Forum 467470 (2004) 1319.
[16] G. Wang, S.D. Wu, L. Zuo, C. Esling, Z.G. Wang, G.Y. Li, Mater. Sci. Eng.
A346 (2003) 83.
[17] Y. Estrin, N.V. Isaev, S.V. Lubenets, S.V. Malykhin, A.T. Pugachov, V.V.
Pustovalov, E.N. Reshetnyak, V.S. Fomenko, L.S. Fomenko, S.E. Shumilin,
M. Janecek, R.J. Hellmig, Acta Mater. 54 (2006) 5581.
[18] M. Avrami, J. Chem. Phys. 8 (1940) 212.
[19] W.A. Johnson, R.F. Mehl, Trans. AIME 135 (1939) 416.
[20] H.D. Mengelberg, M. Meixner, K. Luecke, Acta Metall. 13 (1965) 835.
[21] X. Molodova, R. Berghammer, G. Gottstein, R.J. Hellmig, Int. J. Mat. Res.
98 (2007) in press.

Potrebbero piacerti anche