You are on page 1of 9

Acta mater.

48 (2000) 2665±2673
www.elsevier.com/locate/actamat

DEFORMATION TEXTURE TRANSITION IN BRASS:


CRITICAL ROLE OF MICRO-SCALE SHEAR BANDS
E. EL-DANAF 1{, S. R. KALIDINDI 1{, R. D. DOHERTY 1 and C. NECKER 2
1
Department of Materials Engineering, Drexel University, Philadelphia, PA 19103-2875, USA and 2Los
Alamos National Laboratory, Los Alamos, NM 87545, USA

(Received 28 December 1999; accepted 29 January 2000)

AbstractÐThe transition in deformation textures between low stacking fault energy f.c.c. metals (e.g. brass
textures) and the medium to high stacking fault energy f.c.c. metals (e.g. copper textures) is addressed. A
detailed microscopy investigation was conducted in parallel with texture measurements on deformed
samples of copper and 70/30 brass to di€erent strain levels in three di€erent deformation paths, namely,
plane strain compression, simple compression, and simple shear. The objective of the study was to identify
the speci®c trends in the transition between the brass textures and the copper textures that correlated with
the onset of deformation twinning and those that correlated with the onset of micro-scale shear banding. It
was found that several important transitions in the evolution of the deformation textures, especially in the
rolled samples, correlated not with the onset of deformation twinning but with the onset of micro-scale
shear banding. These results strongly suggest that the critical feature in texture transition is not twinning
directly, but the shear banding promoted by the high strain hardening rates of low stacking fault energy
f.c.c. metal. 7 2000 Acta Metallurgica Inc. Published by Elsevier Science Ltd. All rights reserved.

Keywords: Shear bands; Texture; Deformation twinning; Crystal structure; Copper alloys

1. INTRODUCTION that this transition shifted systematically to lower


Stacking fault energy (SFE) plays an important role strains with increasing the solute level from 2.5 to
in f.c.c. metals and alloys in determining the pre- 30% zinc in pure copper. These observations were
vailing mechanisms of plastic deformation. Low attributed to the faster formation of mechanical
SFE metals exhibit a tendency for mechanical twin- twins in the higher solute Cu±Zn alloys.
ning besides slip during large plastic deformations}. Le€ers and Kayworth [6] conducted a trans-
Based largely on this correlation, Wassermann [3] mission electron microscopy investigation on
hypothesized that deformation twinning was re- deformed samples of brass and argued that the
sponsible for the transition between the so-called volume fraction of the twinned regions was so small
copper-type and the brass-type textures. A rigorous that the development of the brass-type texture can-
analysis of the evolution of rolling textures in cop- not be a volume e€ect of deformation twins as
per and copper±zinc (brass) alloys, the latter having suggested by Wassermann [3] and Hirsch et al. [4,
a rapidly falling SFE as the zinc content increases, 5]. In 40% rolled 85/15 brass, Le€ers and
was conducted by Hirsch et al. [4, 5]. These authors Kayworth [6] and Le€ers and Sorenson [7] found
demonstrated that the copper±brass transition of that the fraction of the twinned material was only
the rolling texture can be characterized by an about 2%. Duggan et al. [8] studied 76% rolled 70/
abrupt decrease of the copper component …f111g  30 brass and also came to the conclusion that the
h112i† accompanied by an increase in the orien- volume fraction of the twinned material was insu-
tation density of the copper-twin orientation TC cient to have a decisive bulk e€ect on the texture,
…f255gh511i† (as predicted by Wassermann [3]), and at least directly. Le€ers and Jensen [9] have asserted
that neither the extent of twinning nor the degree of
orientation selectivity of twinning is suciently
great to account for the di€erences between copper
{ Present address: Department of Mechanical Design
and Production, Cairo University, Cairo, Egypt.
texture and the brass texture.
{ To whom all correspondence should be addressed. Le€ers [10] proposed an alternative model for
} Recent research by our group has re-examined the con- evolution of brass rolling texture. According to this
nection between SFE and twinning in f.c.c. alloys, and model, in early stages of deformation, 50% of the
clari®ed the physics behind this correlation [1, 2]. grains would have twin lamellae and that would

1359-6454/00/$20.00 7 2000 Acta Metallurgica Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S 1 3 5 9 - 6 4 5 4 ( 0 0 ) 0 0 0 5 0 - 1
2666 EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION

result in mainly planar slip (slip parallel to twin al. [4] found that the part of the b-®ber (covering
lamellae which in f.c.c. lie on {111} matrix planes) orientations from the ``copper'' component to the
in the twinned grains. In the other 50% of the ``brass'' component through the ``S'' orientation in
grains, that are free of twins, slip takes place on Euler angle space), found near the ``copper'' orien-
multiple {111} slip planes in order to maintain tation …f111gh112i† should twin most easily. Thus,
strain continuity. At later stages of deformation, the copper orientation might be expected to be
grains with {111} planes approximately parallel to transformed to the TC (twin copper) orientation
the rolling plane have a dense population of defor- …f255gh511i). According to Hirsch et al. [4], sub-
mation twins, which makes homogeneous slip on sequent deformation (rolling) causes TC to rotate
these planes extremely dicult, and leads to in- to an intermediate orientation Y …f111gh112i† which
homogeneous deformation (shear bands) that pro- then rotates to the Goss orientation G …f110gh001i†
duces a texture di€erent from that produced by a by the process of shear bands. These authors hy-
more homogeneous deformation. This process pothesize that the Goss orientation is only a meta-
involves a non-Taylor (non-homogeneous) defor- stable orientation and will then rotate to the brass
mation ®eld at the micro-scale that, according to orientation. This model, for the development of
Le€ers [10], leads to a texture di€erent from the brass orientation in brass-type textures, relies on a
copper-type texture. However, these authors have volume e€ect of deformation twinning and is there-
not stated the strain levels at which the two distinct fore very questionable in light of the ®ndings from
stages of texture evolution outlined above operate. Le€ers and co-workers described above. On the
Furthermore, Le€ers remarked [10] that a single other hand, Le€ers and Sorenson [7] argue that the
model (a modi®ed Sachs model assuming slip pre- development of the brass component in brass-type
dominantly on a single {111} plane) provides good rolling texture is a natural consequence of the non-
predictions of deformed textures in brass during homogeneous deformation ®eld promoted by the
both stages of texture evolution, although the production of deformation twins, and that this fea-
micro-scale deformation processes are entirely ture is captured well by their modi®ed Sachs model.
di€erent in the two stages. This raises serious ques- However, the main problem with this hypothesis, in
tions on whether there are indeed two separate and the opinion of the present authors, is that this
distinct stages of texture evolution, as proposed by model should result in a grossly heterogeneous
these authors. strain ®eld at the grain-scale (due to the di€erences
There is considerable disagreement and confusion in orientations of twins in the di€erent crystal
in literature on when the brass rolling texture starts orientations) and give rise to large di€erences in
to deviate from copper rolling texture. Wassermann grain morphologies in the deformed microstructure
[3], Kallend and Davis [11], Hirsch and Lucke [4], for di€erently oriented grains. There is no evidence
and Singh et al. [12], reported that up to moderate for such features in the deformed microstructures in
rolling deformations …050% rolling or 00:7 true the studies reported to date.
strain in plane strain compression) the texture in Sevillano et al. [15] argued that shear bands are
brass develops by normal slip in the same manner present in rolled f.c.c. metals of low SFE from the
as copper, to produce a spread of orientations from point where the brass texture starts to form, and
f112gh111i (copper orientation) to f110gh112i (brass investigated if the shear bands in¯uenced the tex-
orientation). However, Le€ers and Kayworth [6] ture evolution signi®cantly in both low and high
disagreed with the interpretation of the experimen- SFE f.c.c. metals. Their numerical study of texture
tal observations since the studies reported typically evolution in these metals led them to conclude that
refer to the {111} pole ®gures that show the di€er- shear bands did not play a signi®cant role in the
ence less clearly than {200} pole ®gures. Le€ers [13] texture evolution. On the other hand, Duggan et al.
argued that there are distinct di€erences in textures [8] conducted a thorough experimental investigation
reported by Kallend and Davis [11] for copper and of microstructure evolution in brass and concluded
for 90/10 brass even after a reduction of only 20%. that shear bands did play a major role in the defor-
On the basis of ODFs, Tobisch and Mucklich [14] mation process and the texture evolution.
showed that the texture of 73/27 brass deviates sig- The above discussion provides a brief summary
ni®cantly from that of copper at rolling reduction of the contradicting views held in current literature
of 30%. It might be noted, however, that Kallend on the origin of the transition from the copper-type
and Davis [11], and Hirsch et al. [4] have also used textures to brass-type textures in f.c.c. polycrystals.
ODF techniques to study these deformation tex- The main handicap in resolving some of the issues
tures. mentioned is a lack of quantitative information on
Another unresolved issue in the evolution of the when (i.e. at what strain levels) deformation twin-
brass-type rolling texture is an understanding of the ning and micro-scale shear banding initiate in the
mechanism responsible for the higher brass com- deformed microstructure in the di€erent metals dis-
ponent …f110gh112i† in these textures compared with cussed. If the information on the initiation and
the copper-type texture. By studying orientation evolution of these microstructural features is gath-
Schmid factors for twinning and for slip, Hirsch et ered in conjunction with texture evolution, then it
EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION 2667

should be possible to clarify some of the issues tilt in 58 sample rotation and tilt increments with a
raised above and resolve them in an unambiguous 1 s counting time.
manner. In recent studies, our group has demon- The data were analyzed using the popLA soft-
strated there are clear correlations between in- ware [17], to produce corrected and recalculated
itiation of deformation twinning and initiation of pole ®gures and orientation distributions. The raw
micro-scale shear banding in the deformed micro- data were corrected for background and defocusing
structure and certain characteristic features in the using a correction ®le generated from a copper
strain hardening plots [1, 2, 16]. The purpose of this sample with a random texture. The data were then
study is to extend these correlations to changes in run through a harmonics algorithm to extrapolate
texture evolution, and thereby explore if any of the the outer fringes of the pole ®gure data and then
transitions in the texture evolution can be speci®- the data were re-normalized. These data were run
cally correlated with either the initiation of defor- through the WIMV algorithm to produce recalcu-
mation twinning or the initiation of micro-scale lated pole ®gures and orientation distributions.
shear banding in the deformed microstructures.
3. RESULTS
2. EXPERIMENTAL DETAILS
3.1. Stress±strain responses
2.1. Mechanical testing
Figure 1(a) shows the equivalent true stress±
The materials used in this study are mainly cop- equivalent true strain responses in simple com-
per and 70/30 brass, which are both f.c.c. metals pression, plane strain compression and simple shear
and are representative of low SFE and medium for 70/30 brass for a grain size of about 30 mm, as
SFE metals, respectively. The as-received materials reported in our previous work [16]. In this plot, the
were heat treated to provide equiaxed grain struc- equivalent stress and its work±conjugate equivalent
tures with an average grain size of 30±40 mm (the strain were de®ned using the von Mises de®nitions.
heat treatment details to obtain di€erent grain sizes Figure 1(b) shows the normalized strain hardening
in these metals are provided in Ref. [2]). The initial rate plotted against normalized stress parameter
textures in the annealed samples were nominally …s ÿ s0 †=G (s0 being the initial yield strength of the
random. Simple compression, plane strain com- alloy, and G the shear modulus). We have pre-
pression, and simple shear tests were conducted on viously demonstrated that the initiation of the ®rst
the annealed samples. The details of these exper- plateau in the strain hardening curve (labeled as
iments and the measured stress±strain responses, Stage B in our previous work [1]) correlated with
have already been discussed by the authors in a the onset of deformation twinning in the micro-
recent publication [16]. In this study, these exper- structure [16]. Furthermore, we have also veri®ed
iments were repeated to produce samples subjected that the cross-over point between the simple com-
to di€erent strain levels in all three modes of defor- pression and plane strain compression stress±strain
mation, and the evolution of underlying texture for curves in Fig. 1(a) (corresponding to an equivalent
each deformation path was documented (details of strain of about 0.5 and an equivalent stress of
texture measurements are presented in the next sec- about 450 MPa) correlated with initiation of exten-
tion). Since the strain levels corresponding to in- sive micro-scale shear banding in the deformed
itiation of deformation twinning and micro-scale microstructures [16]. Note that this cross-over point
shear banding in these materials (with the chosen also corresponds to the point where there is a
grain sizes) have already been established in our noticeable drop in the strain hardening rate plot for
previous study [1, 16], we selected deformation plane strain compression compared with simple
levels such that we can speci®cally examine the compression [Fig. 1(b)]. We have also reported pre-
changes in texture evolution brought upon by onset viously that in simple compression deformed
of deformation twinning and by onset of micro- samples of 70/30 brass micro-scale shear bands in-
scale shear banding in the low SFE f.c.c. metals. itiated at about a true strain of ÿ0:6, and that the
numbers of these shear bands in simple compression
deformed samples were signi®cantly much lower
2.2. Texture measurements
than those in plane strain compression deformed
Crystallographic textures were measured by X- samples [1, 16].
ray re¯ection technique on a Scintag X1 ®ve-axis
pole ®gure goniometer using Cu-Ka radiation. The 3.2. Texture evolution
beam is a point source, circularly collimated to
0.8 mm. The detector is a Peltier cooled solid state The evolution of texture in plane strain com-
detector with a slit of 2 mm. The (111), (200) and pression of copper and 70/30 brass was documented
(220) pole ®gures were measured. The samples were by measuring texture at various strain levels in the
oscillated 4 mm to collect data from as many grains range of ÿ0:25 to ÿ1:50: The (111), (100) and (110)
as possible. Data were collected out to 808 sample equal-area projection pole ®gures for both copper
2668 EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION

Fig. 1. (a) Equivalent true stress±equivalent true strain responses (von Mises) in simple compression,
plane strain compression and simple shear for 70/30 brass for a grain size of about 30 mm. (b)
Normalized strain hardening rate plotted against normalized stress parameter.

and 70/30 brass corresponding to strains of ÿ0:25 akin to a brass-type texture. At the low strain
and ÿ1:50 are shown in Figs 2 and 3, respectively. levels, however, there is only a slight di€erence, in
As expected, at high strain levels the deformation agreement with the observations made previously
texture in copper resembles a copper-type texture by Le€ers and Jensen [9]. In order to compare
while the corresponding texture in 70/30 brass is quantitatively the texture evolution between these

Fig. 2. (111), (100) and (110) equal-area projection pole ®gures for both copper and 70/30 brass at a
strain of ÿ0:25 in plane strain compression.
EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION 2669

Fig. 3. (111), (100) and (110) equal-area projection pole ®gures for both copper and 70/30 brass at a
strain of ÿ1:50 in plane strain compression.

low and medium SFE metals, the volume fraction ponent in 70/30 brass increases with strain up to
of grains oriented along the most important orien- about a strain of 0.5, and then there is a distinct
tations in the f.c.c. rolling textures, namely the cop- change in that it decreases with strain beyond this
per orientation, the brass orientation, and the S strain level. Figure 5 shows the evolution of the
orientation, were calculated from the experimental volume fraction variation of the brass orientation
data. The volume fractions were calculated using …011†‰112Š in both 70/30 brass and copper with the
the following relation [18]: imposed strain. It is seen that the general trends are
the same for both materials except that the rate of
DV X increase of the brass component is somewhat higher
ˆ f … g†Dg …1†
V in 70/30 brass compared with copper (especially at
where g represents an orientation in the Euler angle low strains up to a strain of about 0.5). Figure 6
space, f( g ) denotes the normalized orientation den- shows the evolution of the volume fraction of the S
sity output from the computer code popLA (pre-
ferred orientation package from Los Alamos [17]),
and the summation is performed on all orientations
within a 15 deg Rodriguez angle of the selected
ideal orientation. Since the output from popLA
code for f( g ) is provided in a discretized orien-
tation space (5 deg bins in all three Euler angles), a
weighting function is used on orientations that are
between 15 and 16 deg Rodriguez angle from the
ideal orientation such that the orientations at
15 deg are weighted with a factor of 1.0 and the
orientations at 16 deg are weighted with a factor of
0.0, and the weighting function is varied linearly in
between these two limits.
Figure 4 shows the evolution of the volume frac-
Fig. 4. Evolution of the volume fraction variation of the
tion of the copper orientation …112†‰111Š in both 70/ copper orientation …112†‰111Š in both 70/30 brass and cop-
30 brass and copper with the imposed strain. It can per with the imposed strain for a bin size of 15 deg of
be seen that the volume fraction of the copper com- Rodriguez angle in plane strain compression.
2670 EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION

ponent is quite small in both the metals, it declined


much more quickly in 70/30 brass than in copper.
The textures from the simple shear deformed
samples were also measured. However, at the strain
levels we investigated (up to a shear strain of ÿ1:6),
there were no signi®cant di€erences in the deformed
textures measured in 70/30 brass and copper
samples. In the interest of space, these have not
been presented here, but interested readers can refer
to Ref. [19].

4. DISCUSSION
Fig. 5. Evolution of the volume fraction variation of the
brass orientation …011†‰112Š in both 70/30 brass and cop- 4.1. Plane strain compression/rolling
per with the imposed strain for a bin size of 15 deg of
Rodriguez angle in plane strain compression. A comparison of the texture evolution in plane
strain compression deformed samples of 70/30 brass
orientation …123†‰634Š with the imposed strain for and copper (of roughly the same initial average
both copper and 70/30 brass. It is seen that the grain size) reveals the following points (Figs 4±6):
evolution of the S orientation in 70/30 brass shows 1. There is a drastic change in the evolution of the
some of the trends seen in the evolution of the cop- volume fraction of the copper component at
per orientation (Fig. 4). Speci®cally, the volume about a strain of ÿ0:5: A similar change also
fraction of the S orientation in 70/30 brass starts to occurs for the volume fraction of the S orien-
decline around a strain of 0.5, although the e€ect tation, although the e€ect is less severe. These
here is not as severe as was seen earlier for the cop- changes correlate very well with the onset of
per orientation. micro-scale shear banding in the deformed
The texture evolution in simple compression microstructures [16]. Note that deformation twin-
deformed samples of both 70/30 brass and copper ning initiates at about a strain of ÿ0:08 in 70/30
was also studied systematically. Figure 7 shows the brass. Since the change in the evolution of the
(111), (100) and (110) equal-area projection pole volume fractions of the copper and S orien-
®gures for both 70/30 brass and copper at a true tations occur well beyond the initiation of defor-
strain of ÿ1:0: In order to make quantitative com- mation twinning and coincide with the onset of
parisons, we computed the volume fractions of shear banding, it is reasonable to attribute these
grains in the (110), (111) and (100) ®ber texture changes as consequences of micro-scale shear
components in both 70/30 brass and copper as a banding in the material.
function of the imposed strain. The results of these 2. There is a noticeable di€erence in the rate of
calculations are shown in Figs 8±10. As expected, increase of the brass component between the two
the results indicate a strong buildup of the (110) metals, especially at low strains. The brass com-
®ber components. However, an interesting feature ponent builds up faster in 70/30 brass compared
observed was that there was a slow but distinct with copper, and this is attributable as a conse-
increase in the (111) ®ber component in 70/30 brass quence of deformation twinning. The nature of
at high strain levels beyond a strain of about ÿ0:55: this in¯uence, we believe, is unlikely to be a bulk
It was also noted that although the (100) ®ber com- volume fraction e€ect of twinning, but it could

Fig. 6. Evolution of the volume fraction variation of the S orientation …123†‰634Š in both 70/30 brass
and copper with the imposed strain for a bin size of 15 deg of Rodriguez angle in plane strain com-
pression.
EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION 2671

Fig. 7. (111), (200) and (220) equal-area projection pole ®gures for both 70/30 brass and copper at a
true strain of ÿ1:0 in simple compression.

be an indirect e€ect of twinning on the slip pro- this change in texture evolution also correlates to
cess, as argued by Le€ers [10]. a slower buildup of the (110) ®ber component in
70/30 brass compared with copper at the same
strain levels. This change in evolution of the
4.2. Simple compression (111) ®ber component once again correlates very
well with the onset of micro-scale shear banding
A comparison of the texture evolution in simple in the deformed microstructures [1]. Note that
compression deformed samples of 70/30 brass and deformation twinning initiates in 70/30 brass in
copper (of roughly the same initial average grain simple compression deformed samples at about a
size) reveals the following points (Figs 8±10): strain of ÿ0:07: Since the change in the evolution
1. There is a slight increase in the volume fraction of the (111) ®ber component occurs well beyond
of (111) ®ber component at large strains in 70/30 the initiation of deformation twinning and co-
brass, beyond a strain of about ÿ0:55: Note that incides with the onset of shear banding, it is
reasonable to attribute this change as a conse-

Fig. 8. Volume fractions of grains in the (110) ®ber texture Fig. 9. Volume fractions of grains in the (111) ®ber texture
components in both 70/30 brass and copper as a function components in both 70/30 brass and copper as a function
of the imposed strain for a bin size of 15 deg of Rodriguez of the imposed strain for a bin size of 15 deg of Rodriguez
angle in simple compression. angle in simple compression.
2672 EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION

quence of micro-scale shear banding in the ma- paper that some of the major changes in texture
terial. Note that the results presented here are in evolution described above between 70/30 brass and
agreement with the reported data on the evol- copper are clearly correlated with onset of extensive
ution of the ratios of intensities of the di€erent micro-scale shear banding and not with the onset of
peaks presented previously by Asgari et al. [1]. twinning. That is not to say that twinning and low
However, since the previous data were reported stacking fault energy play no indirect role in texture
as ratios of intensities, there were no quantitative di€erences. Low stacking fault energy promotes the
data on the evolution of the individual ®ber com- condition for twinning as discussed for example by
ponents, such as presented in this study. El-Danaf et al. [2], and twinning in turn promotes
2. There is a noticeable di€erence in the rate of increased strain hardening, see for example Asgari
decrease of the (100) ®ber component between et al. [1]. Numerous studies, for example by
the two metals at low strains. The (100) ®ber Duggan et al. [8] and Korbel et al. [22], have shown
component drops much faster in 70/30 brass that alloys that twin show much enhanced shear
compared with copper, and this is attributable as banding. Therefore, there is clearly an indirect link
a consequence of deformation twinning. This between twinning and texture changes but the key
result has been noted in previous studies [20], structural feature that correlates with the texture
and has been attributed to the fact that the change is the onset of strain localization, shear
grains in the (100) ®ber texture are ideally banding. Other studies, for example by Baumann
oriented for deformation twinning (with the lar- and Doherty [21], have shown that solute enhanced
gest resolved shear stress on the twinning sys- strain hardening in aluminum promotes shear band-
tems). ing. Therefore, an important test for the ideas pre-
sented here will be to see if heavy cold rolling of
high solute aluminum alloys also shows a change
4.3. Simple shear from a copper to a brass type rolling texture.
Future work also appears necessary to try to
It was observed that the shear textures in 70/30 understand why the shear banding observed in
brass and copper were quite similar at least to the brass causes the texture changes seen here for plane
strain level of ÿ1:6 as investigated in this study. strain compression (or rolling) and for simple com-
Optical microscopy investigation of the simple pression deformation.
shear deformed samples in 70/30 brass revealed an
absence of any micro-scale shear bands. However,
our previous studies [16] demonstrated that there 5. CONCLUSIONS
was a substantial amount of deformation twinning This study demonstrates that several of the im-
in the same samples as were investigated for texture portant di€erences occurring in the evolution of de-
here. Furthermore, the amount of deformation formation textures between low SFE f.c.c. metals
twinning in the shear samples was found to be com- such as brass and medium SFE f.c.c. metals such as
parable with that in simple compression samples copper correlate with onset of micro-scale defor-
[16]. So, in spite of the high frequency of defor- mation banding and not with the onset of defor-
mation twins in simple shear deformation there mation twinning in the deformed microstructures.
appears to be no major in¯uence of deformation Speci®cally, these include:
twinning on texture evolution. This observation
provides support for the arguments made in this 1. The decrease of the copper component in 70/30
brass at high strain levels (beyond a strain level
of ÿ0:5 for a 30 mm grain-sized material) in
plane strain compression.
2. The decrease of the S component in 70/30 brass
at high strain levels (beyond a strain level of
ÿ0:5 for a 30 mm grain-sized material) in plane
strain compression.
3. The increase of the (111) ®ber component in 70/
30 brass at high strain levels (beyond a strain
level of ÿ0:55 for a 30 mm grain-sized material)
in simple compression.
It was also found that in simple shear where there
was twinning but no shear banding, the two metals,
copper and brass, each with very di€erent SFE
showed similar textures to a strain of 1.6.
Fig. 10. Volume fractions of grains in the (100) ®ber tex-
ture components in both 70/30 brass and copper as a func-
tion of the imposed strain for a bin size of 15 deg of AcknowledgementsÐThis work was supported ®nancially
Rodriguez angle in simple compression. by NSF (CMS 9503943). Cold-rolled plates of 70/30 brass
EL-DANAF et al.: DEFORMATION TEXTURE TRANSITION 2673

were kindly provided for this work by Dr James Breedis 12. Singh, C. D., Ramaswamy, V. and Suryanarayana,
of Olin Corporation, CT. C., Textures Microstruct., 1992, 19, 101.
13. Le€ers, T., Textures Microstruct., ICOTOM 3, 1993,
22, 53.
REFERENCES 14. Tobisch, J. and Mucklich, A., Texture, 1974, 1, 211.
1. Asgari, S., El-Danaf, E., Kalidindi, S. R. and 15. Sevillano, J. G., Houtte, P. V. and Aernoudt, E.,
Doherty, R. D., Metall. Trans., 1997, 28A, 1781. Scripta metall., 1977, 11, 581.
2. El-Danaf, E., Kalidindi, S. R. and Doherty, R. D., 16. El-Danaf, E., Kalidindi, S. R. and Doherty, R. D.,
Metall. Trans., 1999, 30A, 1223. Int. J. Plasticity, in press.
3. Wassermann, G., Z. Metallk., 1963, 54, 61. 17. Kallend, J. S., Kocks, U. F., Rollett, A. D. and
4. Hirsch, J., Lucke, K. and Hatterly, M., Acta metall., Wenk, H.-R., Mater. Sci. Engng, 1991, A132, 1.
1988, 36, 2905. 18. Bunge, H. J., in Quantitative Texture Analysis, ed. H.
5. Hirsch, J. and Lucke, K., Acta metall., 1988, 36, 2863. J. Bunge and C. Esling. Duetsche Gesellschaft fuÈr
6. Le€ers, T. and Kayworth, P., Soc. Francaise Metall., metallkunde±Societe Franc° aise de MeÂtallurgie,
ICOTOM 3, 1973, 149. Oberursel, 1982.
7. Le€ers, T. and Sorenson, J. B., Acta metall., 1990, 38, 19. El-Danaf, E., Ph.D. thesis, Drexel University, 1998.
1917. 20. Chin, G. Y., in Textures in Research and Practice, ed.
8. Duggan, B. J., Hatherly, M., Hutchinson, W. B. and J. Greuen and J. Wassermann. Springer-Verlag,
Wake®eld, P. T., Metals Sci., 1978, 12, 343. Berlin, 1969, p. 51.
9. Le€ers, T. and Jensen, G., Trans. metall. Soc. 21. Baumann, S. F. and Doherty, R. D., in Aluminum
A.I.M.E., 1968, 242, 314. Alloys for Packaging, ed. J. G. Morris. TMS,
10. Le€ers, T., ICOTOM 11, 1996, 1, 300. Warrendale, PA, 1993, p. 369.
11. Kallend, J. S. and Davis, G. J., Phil. Mag., 1972, 25, 22. Korbel, A., Embury, J. D., Hatherly, M., Martin, P.
471. L. and Erbsloh, H. W., Acta metall., 1986, 34, 1999.

You might also like