Sunteți pe pagina 1din 10

Acta mater. Vol. 46, No. 5, pp.

1827±1836, 1998
Acta Metallurgica Inc.
Published by Elsevier Science Ltd
Printed in Great Britain
PII: S1359-6454(97)00365-0 1359-6454/98 $19.00 + 0.00

MICROSTRUCTURE AND FLOW STRESS OF


POLYCRYSTALS AND SINGLE CRYSTALS
N. HANSEN and X. HUANG
Materials Research Department, Risù National Laboratory, DK-4000, Roskilde, Denmark

(Received 12 May 1997)

AbstractÐThe relation between polycrystal deformation and single crystal deformation has been studied in
this work. A pure aluminium polycrystal having an average grain size of 300 mm has been strained in ten-
sion at room temperature. The ¯ow stress has been determined at four di€erent strains (0.05, 0.14, 0.22
and 0.34) and deformation microstructures have been characterized qualitatively and quantitatively by
transmission electron microscopy. Improved experimental techniques have allowed large foil areas to be
characterized and in total 89 grains have been examined. At the four strains examined a classi®cation of
the deformation microstructures into three di€erent types has shown a correlation between the grain orien-
tation and the type of deformation microstructure which develops during straining. The dislocation density
has been calculated at each strain and by assuming that the shear stress is proportional to the square root
of the dislocation density the shear stress±strain relationship has been derived for each of the three groups
of grains showing di€erent deformation microstructures. The stress±strain curves show a strain hardening
behaviour which depends on the orientation of the grain. The behaviour of the grains embedded in the
polycrystal is compared with the behaviour of single crystals and the stress±strain curve of the polycrystal
is estimated with good accuracy from single crystal data, which are weighted based on a quantitative tex-
ture analysis of the polycrystal. # 1998 Acta Metallurgica Inc.

1. INTRODUCTION embedded in a polycrystal information can be


Based on classical papers, e.g. by Sachs [1], obtained about the slip pattern of individual grains.
Taylor [2] and Kocks [3, 4] the relationship between It has also been possible to estimate the dislocation
polycrystal deformation and single crystal defor- density of the grains allowing a calculation of the
mation has been studied extensively over many shear stress taken to be proportional to the square
root of the dislocation density. This experimental
years [5±9]. The general problem when comparing
approach can therefore give information both about
polycrystals and single crystals is that the slip pat-
the slip pattern and the ¯ow stress of the grains
tern of grains embedded in a polycrystal is a€ected
embedded in a polycrystal allowing a comparison
by grain interaction which is absent when single
with single crystal behaviour. In applying this ex-
crystals are tested. This problem has been addressed
perimental approach it cannot be excluded that
by Kocks who suggested basing comparisons on
local grain interaction e€ects in the polycrystal may
single crystals deforming in multislip [4]. Thereby
reduce the e€ect of grain orientation e€ects making
the slip pattern in the single crystal might simulate
such a comparison dicult. To minimize such an
the slip pattern of the polycrystal where strain ac-
e€ect and at the same time ensure that polycrystal
commodation between grains requires several slip
behaviour is studied, aluminium having a fairly
systems to operate, e.g. at least ®ve as suggested in
large grain size has been chosen. Aluminium being
the Taylor model [2, 5]. In tension such multislip
a f.c.c. metal with a high stacking fault energy is
may be activated in free single crystals having their
considered suitable as the many slip systems active,
tensile axes in the h100i and h111i directions, [3].
reduce the e€ect of grain size on the slip pattern
The shear stress±strain relationship for these two
and on the ¯ow stress [10].
crystal orientations di€ers signi®cantly and it was
evident that the best simulation of polycrystal beha-
viour was obtained by choosing the crystal of [111]
2. EXPERIMENTAL
orientation. A larger strain hardening rate and a
higher level of shear stress compared to the beha- Pure aluminium (99.996%) specimens having a
viour of a real polycrystal showed, however, that a grain size of approximately 300 mm have been pro-
de®nitive answer to the problem of comparing poly- duced by cold rolling and recrystallization.
crystal deformation and single crystal deformation Specimens of square cross section 48 mm2 and
was not reached [4, 6]. with gauge length 20 mm have been strained in ten-
A di€erent approach has been tried in this work. sion in an Instron testing machine to true strains of
By careful, microstructural analysis of the grains 0.05, 0.14, 0.22 and 0.34 (engineering strains of 5,
1827
1828 HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS

15, 25 and 40%). The strain rate used was where Sv is the dislocation boundary area per unit
8.3  10ÿ4 sÿ1. To ensure that the stress±strain beha- volume and rA is the dislocation density per unit
viour is representative for true polycrystalline beha- area of boundary [14]. Sv does not depend on the
viour the stress±strain curve has been compared spatial arrangement of the boundaries and equals
with previous data obtained from specimens having 2/Dr [13]. rA equals 1/h for a simple tilt wall where
a larger ratio between the specimen cross section h is the spacing between the dislocations and for a
and the grain size [10]. Very good agreement has twist wall rA=2/h [15]. For small angles of misor-
been found both in shape and level of the stress±
ientation h 2 b/ym and by averaging over tilt and
strain curves indicating that an e€ect of surface
twist boundaries rA is approximated by 1.5 ym/b.
grains is negligible in the present specimens. For
transmission electron microscopy (TEM) obser- The dislocation density in the boundaries is then:
vations, thin foils have been cut from the two side rb ˆ 3ym =Dr  b …2†
planes of the tensile specimens and subsequently
thinned and electropolished using standard Besides the dislocations present in dislocation
methods. In this preparation, care has been exerted boundaries, dislocations are also found in the
in obtaining foils from the interior of the specimens volumes between the boundaries. The density of
to reduce possible surface e€ects. On each foil the these dislocations (ro) has been determined from
tensile axis has been marked allowing the orien- TEM micrographs by counting the number of inter-
tation of crystallites to be determined with respect sections between dislocations and a set of random
to this axis. In order to obtain accurate structural test lines (PL). The dislocation density is then
parameters large areas have been examined of the
order of 2000±6000 mm2 of each grain. In total 89 ro ˆ 2PL =t …3†
grains have been examined at the four strains. where t is the foil thickness [13]. This parameter has
Local orientation measurements of crystallites
been determined from intensity oscillations in con-
within grains have been made using a semiauto-
vergent beam di€raction patterns. In general, the
matic Kikuchi line technique in the areas where the
determination of ro is not very accurate but as ro is
montages of the microstructure are taken [11].
Linear and area scanning have been used when only a rather small fraction of the total dislocation
measuring the orientation of the crystallites allow- density, the accuracy of measuring the dislocation
ing the orientation measurement to be related to density is predominantly determined by the pre-
the microstructural observations. One such relation- cision in measuring ym and Dr.
ship concerns the crystallographic characteristic of The texture of the tensile specimens in the unde-
the boundaries in which the dislocations have accu- formed state and after straining has been deter-
mulated, i.e. a dislocation boundary should be mined by neutron di€raction [16]. For each sample
classi®ed as either a crystallographic or a non-crys- three complete pole ®gures have been measured and
tallographic boundary [12]. A boundary in the ®rst the three-dimensional crystallite orientation distri-
category is taken to be parallel to within 58 with a bution function (ODF) has been calculated using
slip plane, whereas boundaries in the second cat- the series expansion method [17]. Table 1 shows the
egory make a larger angle with a slip plane. A pre-
volume fractions for the [100], [111] and [110] orien-
liminary classi®cation is based on measurements of
tations which have been determined by integrating
the angle between the dislocation boundary and the
the density over an angle of 158 around the ideal
trace of the slip plane in the ®eld of observations.
The analysis was then re®ned by tilting the foil to position of the component in the stereographic tri-
achieve the minimum image width for the dislo- angle. Table 1 shows that the texture changes are
cation boundaries. The inclination of this boundary not very pronounced over the strain range investi-
plane to the nearest slip plane was then determined gated. There is a tendency that the [111] component
and if the inclination was determined to be less strengthens and that the [110] component weakens.
than 58 than the boundary was con®rmed as of This corresponds well to predictions by the Taylor
crystallographic type. model [2, 5] that unstable crystal orientations rotate
The orientation measurements are used to calcu- towards either [100] or [111] directions for straining
late the mean misorientation angle (ym) across the in tension.
dislocation boundaries. Also the spacing between
the boundaries (Dr) is determined by counting the Table 1. Volume fractions of texture components
number of intersections between the boundaries and
h100i h111i h110i
a random set of test lines [13]. From the two Strain 0±158 0±158 0±158
parameters ym and Dr the volume density of dislo-
0 0.280 0.132 0.100
cations in boundaries (rb) can be obtained from the 0.05 0.240 0.109 0.106
equation: 0.14 0.282 0.236 0.048
0.22 0.256 0.284 0.023
rb ˆ Sv rA …1† 0.34 0.277 0.226 0.033
HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS 1829

Fig. 2. Type 2 microstructure. TEM micrograph at a


strain of 0.14. The structure is typically subdivided by
ordinary cell boundaries having no crystallographic or
macroscopic orientation.

Next it has been found [12] that the microstructure


of a grain relates to the orientation of the tensile
axis of that grain. This is illustrated in Fig. 4,
Fig. 1. Type 1 microstructure. TEM micrograph at a
where the average orientation of the tensile axis of
strain of 0.14. The microstructure is subdivided by the 89 grains is plotted in an inverse pole ®gure. In
extended straight dense dislocation walls and microbands this ®gure the microstructural di€erences are intro-
into cell blocks which contain ordinary dislocation cells. duced by using di€erent symbols for the three types
The extended boundaries have been termed crystallo- of microstructures. Figure 4 shows that Type 1 and
graphic boundaries (see text). The tensile axis is marked
TA in this side plane view. Type 2 structures are only observed within rela-
tively small regions of the standard triangle, in an
area near the boundary line between [100] and [111]
3. RESULTS and near [100], respectively. By contrast, Type 3
structures have been observed for many di€erent
3.1. Microstructure
The microstructural observations show that the
deformation microstructure of the individual grains
is fairly homogeneous, but that signi®cant grain-to-
grain variations exist. It has been shown [12] that
these variations are not random allowing the
observed microstructures to be grouped into three
types as shown in Figs 1±3. The characteristics of
these structures are:
. Type 1: The microstructure is subdivided by
extended, crystallographic dislocation boundaries
de®ning cell blocks which contain ordinary dislo-
cation cells.
. Type 2{: The microstructure is subdivided by
ordinary dislocation boundaries de®ning a three-
dimensional cell structure.
. Type 3{: The microstructure is subdivided as for
Type 1, but with extended non-crystallographic
boundaries.
Fig. 3. Type 3 microstructure. TEM micrograph at a
{Note that Type 2 and Type 3 in [12] were called Type strain of 0.14. The microstructure is as for Type 1 subdi-
3 and Type 2, respectively. vided into cell blocks by dense dislocation walls and
{Note that Type 2 and Type 3 in [12] were called Type 3 microbands. The extended boundaries have been termed
and Type 2, respectively. non-crystallographic boundaries. (See text.)
1830 HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS

3.2. Volume concentration and grain orientation


The di€erent microstructures observed for the
three groups of grains make it relevant to estimate
the volume concentration of these groups. From
this it will be possible to estimate the individual
contributions of the groups to the behaviour of the
bulk polycrystal. The determination of these volume
concentrations has been based on a discretization of
the ODFs (see Section 2) into about 6000 orien-
tations which are introduced into an inverse pole
®gure. The next step of the calculation has been
based on the overlap observed (see Fig. 5) between
the groups of grains showing di€erent microstruc-
tures and the regions marked by dashed boundaries,
which subdivide the standard triangle. These
regions originate from theoretical calcula-
tions [2, 5, 18] based on the Taylor model and the
Fig. 4. Inverse pole ®gure showing the tensile axis orien- ®ve regions in Fig. 5 are each characterized by a
tation of 89 grains embedded in polycrystalline specimens
strained in the range 0.05±0.34. R Grains with crystallo-
speci®c set of possible slip systems. From Fig. 5 the
graphic dislocation boundaries (Type 1 microstructure). q following qualitative correspondence between
Grains containing equiaxed cells (Type 2 microstructure). groups of grains and regions is obtained:
q Grains with non-crystallographic dislocation boundaries
(Type 3 microstructure). From Ref. [12]. (i) Grains showing Type 1 microstructure have
orientations concentrated in the lower half of
region 2 with M-factors less than 3.0.
(ii) Grains showing Type 2 microstructure are con-
grain orientations. It has also been observed at all
centrated in Region 1 and in the left part of
four strains, that the orientations of grains are not Region 3 with M-factors less than 2.5.
randomly distributed across the triangle. The con- (iii) Grains showing Type 3 microstructure are dis-
centration of orientations is largest near the bound- tributed over the remaining area of the tri-
ary line between [100] and [111]. This concentration angle.
re¯ects the starting texture as well as the textural Based on this observation the volume concentration
evolution taking place during tensile straining. and the average M-factors for the three groups of

Fig. 5. Inverse pole ®gure as in Fig. 4, but subdivided into ®ve regions (see text).
HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS 1831

Table 2. Volume fractions ( f ) and average M-factors for three groups of grains showing di€erent types of deformation microstructure
Type 1 Type 2 Type 3
Strain f1 M1 f2 M2 f3 M3

0 0.19 2.75 0.28 2.40 0.53 3.23


0.05 0.17 2.75 0.24 2.40 0.59 3.21
0.14 0.18 2.76 0.27 2.40 0.55 3.30
0.22 0.19 2.75 0.24 2.39 0.57 3.30
0.34 0.22 2.76 0.27 2.39 0.51 3.28
With reference to Fig. 5: Type 1. Region 2 for M < 3.0; Type 2. Region 1 + Region 3 M < 2.5; Type 3. Region 4 + Region 5 + Region
2 (M e 3.0) + Region 3 (M e 2.5).

grains have been calculated, see Table 2. In this cal- above 0.2 in the dislocation density of grains with
culation it has been assumed that the texture of the Type 2 structures, i.e. grains with orientations near
specimens shows rotational symmetry. This intro- the h100i corner position in Fig. 4. From the dislo-
duces a discrepancy when comparing with calcu- cation density and the volume fraction (Table 2) of
lations where the full ODF information is used. A the three groups of grains, the total dislocation den-
comparison shows that this discrepancy is small sity (rt) of the specimen can be calculated. rt is
about 3%, and it has therefore been neglected. given in Table 3 together with the measured tensile
Table 2 shows that neither the volume fraction nor ¯ow stress.
the M-values are much a€ected by the tensile strain. The ¯ow stress in terms of the resolved shear
However, this does not mean that grain rotations stress (t) has been estimated for the three types of
have not taken place during straining. This is illus- structures based on the empirical relation by taking
trated in Fig. 6 showing the volume concentrations the friction stress to, equal to zero and applying
of grains in the ®ve regions in Fig. 5. Figure 6 p
clearly shows that the concentration of grains near t ˆ abG ro ‡ rb …4†
the boundary line between [100] and [111] increases
where b is Burgers vector (b = 0.286  10ÿ9 m), G is
with strain in accordance with model predictions
the shear modulus (G = 26,000 MPa), and a is a
(see Section 2). Also to be noted in Fig. 6 is that
the largest changes appear to be rotations from number of the order of 0.2±0.35 [19, 20]. The val-
region 5 to region 4, i.e. the grains rotate through idity of equation (4) has been shown in many stu-
orientations, all representing the group of grains dies and apart from a variation in a, equation (4) is
developing the Type 3 microstructure. quite insensitive to changes in materials and process
parameters. In order to derive an estimate of a in
the present experiment, equation (4) has been used
3.3. Dislocation density and ¯ow stress to calculate the ¯ow stress (s) of the polycrystalline
For each of the three groups of grains the dislo- specimen taking s = Mt, thus
cation density (rb+ro) has been determined as a p
s ˆ MabG rt …5†
function of strain, see Fig. 7. It is estimated that
the densities given are determined with an accuracy where M is the average Taylor factor which is cal-
of about 10% and it should be noted that rb only culated to 2.94 from M-factors for three groups of
accounts for the geometrically necessary dislo- grains and for four strains (Table 2). The data in
cations. Figure 7 shows that the dislocation density Table 3 allows a calculation of a, which also is
of grains with Type 3 structures is signi®cantly lar-
ger than that of grains with Type 1 and Type 2
structures. Also note the small increase for strains

Fig. 7. The calculated dislocation densities for groups of


Fig. 6. The volume fraction of grains having orientations grains having the three types of microstructure shown in
within the ®ve regions in Fig. 5. (See text.) Figs 1±3.
1832 HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS

Table 3. Flow stress and dislocation density nominal shear stress±strain behaviour for the poly-
Dislocation crystal, using an M factor of 2.94 when transform-
Tensile Tensile ¯ow density ing from tensile stress/strain to shear stress/strain.
strain stress (MPa) 1013(mÿ2) a
Here, the strain hardening behaviour of the poly-
0.05 34.9 3.2 0.28 crystal di€ers signi®cantly from the behaviour of
0.14 48.5 9.8 0.22 the three groups of grains which together comprises
0.22 55.2 15.1 0.21
0.34 61.5 19.0 0.20 the structure of the polycrystalline specimen.
G = 26,000 MPa; b = 0.286  10ÿ9 m; M = 2.94.
4. DISCUSSION
given in Table 3. a is observed to decrease with
strain, but a reasonable approximation of the exper- This experiment shows for grains embedded in a
imental stress±strain curve can be obtained for the polycrystal a clear correlation between the crystallo-
average value of a equal to 0.24, which will be used graphic orientation, the deformation microstructure
and the estimated shear stress±strain behaviour.
in the following. The good agreement between this
This has also been observed for single crystals
®gure and the a-values reported in the literature
deformed in tension where the microstructure [21±
(e.g. Refs [18] and [19]) gives good con®dence in the
27] and the shear stress±strain behaviour [3, 4, 28±
applied method of determining the dislocation den-
31] has been examined for di€erent crystal orien-
sity. That the experimental a value of 0.24 is rela-
tations. A comparison between polycrystalline beha-
tively low compared to published values indicates viour and single crystal behaviour will therefore be
that the contribution to the total dislocation density carried out in the following based on (i) microstruc-
from geometrically necessary dislocations appar- tural observations and (ii) estimates and measure-
ently outweighs the contribution from redundant ments of the stress±strain behaviour. This
dislocations in the boundaries. comparison is followed by a discussion of polycrys-
The shear stress for the grains having the three tal deformation, especially the stress±strain curve of
di€erent types of microstructure is plotted in Fig. 8 the polycrystal.
vs the shear strain. This strain has been estimated
by multiplying the tensile strain with an average 4.1. Microstructure and slip pattern
M factor for each of the three groups of grains Deformation microstructures have not been stu-
averaged over four strain values. These died to a great extent in aluminium single crystals
factors are M1(av) = 2.76, M2(av) = 2.40 and strained in tension. Copper crystals are, however,
M3(av) = 3.26 (Table 2). Note in Fig. 8 that three expected to show a comparable behaviour and such
groups of grains having the three di€erent types of crystals have been studied in more detail [21±26]. A
microstructure show signi®cantly di€erent shear general ®nding has been [21±24] that the defor-
stress±strain curves. Most marked is the typical mation structure of the [100] crystal is a character-
Stage ``III'' behaviour, i.e. parabolic hardening of istic three-dimensional, uniform cell structure,
group 3 grains compared to the behaviour of group whereas a banded structure of dislocation bound-
2 grains where a rapid strain hardening at low aries is generally observed for other
strain develops into a low hardening region at med- orientations [24±26]. For such banded structures it
ium to high strain. In Fig. 8 is also plotted the has been found that the deviations of the plane of
the dislocation boundary from the primary slip
plane increases with increasing multiplicity of slip.
An example of this e€ect is a deviation of 298
measured for the dislocation boundaries in a [111]
crystal, which has six slip systems [25]. These obser-
vations have been further examined by a Schmid
factor analysis [27], which shows that if two systems
on one or two planes have high Schmid factors the
crystals tend to form crystallographic dislocation
boundaries. However, if more than two slip planes
have high Schmid factors, the crystals tend to form
non-crystallographic boundaries [27].
For the present polycrystal, a Taylor analysis in
general agrees with a Schmid factor analysis for
single crystals as the group of grains developing
Type 1 structures predominantly slip on two planes,
whereas the group of grains developing Type 3
Fig. 8. The shear stress±strain curves for groups of grains
structures slip on three or more planes. For the
having di€erent types of microstructure (see text) and for group of grains developing the Type 2 structure the
the polycrystalline specimen. number of slip planes is four as in a [100] single
HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS 1833

crystal strained in tension. Finally, note the large


spread of orientations which characterizes the
groups of grains, which develop the Type 3 struc-
ture. For many of these orientations, single glide
will dominate in a single crystal with the formation
of crystallographic dislocation boundaries as a
result [19, 21]. In contrast, grains embedded in a
polycrystal develop the Type 3 structure, i.e. non-
crystallographic dislocation boundaries are formed.
This di€erence shows that glide on a limited num-
ber of systems in a single crystal is replaced by mul-
tislip when a crystal of the same orientation is
embedded in a polycrystal.
In correlating the microstructure with the slip
pattern the coincidence should be further explored
between grain orientations grouped according to
microstructure and grain orientations grouped Fig. 9. The e€ect of the grain orientation on the nominal
according to speci®c slip system combinations (see shear stress±strain curves of single crystals of pure alu-
minium. The orientations of the crystals are shown on the
Fig. 5). The speci®c slip system combinations in insert of the standard stereographic triangle. The data are
the ®ve regions of this ®gure are calculated on the from Ref. [29].
basis of the Taylor model [2, 5, 18], but other
models may divide the triangle in a di€erent way. strongly orientation dependent and that the shear
One goal for future research will therefore be to stress±strain behaviour can be divided into three
establish correlations between the microstructural characteristic types. (i) Type 1 shows hardening in
evolution in tension and the active slip systems as three stages and a decreasing length of the easy
has recently been done for f.c.c. metals deformed glide stage with a decreasing angle between the in-
by rolling [32]. itial orientation of the crystal and the boundary
To conclude this section it must be pointed out between [100] and [111]. (ii) Type 2 shows a rapid
that grain rotations during deformation create a initial work hardening rate, which decreases to a
general problem when microstructures of polycrys- very low value at a shear strain of about 0.1. This
tals have to be correlated with orientations of the type is characteristic of crystals having orientations
individual grains. The reason is that an actual in the [100] corner of the triangle. (iii) Type 3
microstructure may re¯ect the total deformation shows parabolic stress±strain curves with a high
history and that previous orientation changes are work hardening rate and a high ¯ow stress. This
unknown. For more detailed studies of correlations type is characteristic for crystals having orientations
between microstructure and crystallographic orien- in the [111] corner of the triangle.
tations, studies of single crystals should therefore be The shear stress±strain curves in Fig. 9 show a
encouraged. clear resemblance to the shear±stress curves in Fig. 8
for the three groups of grains comprising the poly-
4.2. Stress±strain curves of single crystals and poly- crystal. This resemblance of the behaviour of single
crystals crystals and grains embedded in a polycrystal will
In contrast to the limited information on defor- be discussed in the following text based on shear
mation microstructures of aluminium single crys- stress±strain curves of single crystals represented in
tals, the stress±strain behaviour has been studied the literature [28±31]. Here relatively good agree-
for many di€erent orientations [28±31]. Generally, ment is found between di€erent investigations for
it has been found that crystals of the three corner example by LuÈcke and Lange [28, 31],
orientations of the stereographic triangle, have dis- Staubwasser [30] and Fortunier et al. [31]. In this
tinctly di€erent shear stress±strain behaviour. For last paper the stress±strain behaviour has been fol-
positions along the boundaries between the corners lowed to relatively large strain values and thus this
a continued change in the stress±strain behaviour is work has been chosen as the basis for most of the
observed from one corner to the next and for crys- comparisons here polycrystal.
tals having orientations within the triangle, an easy In Figs 10±12, the polycrystal shear stress±strain
glide stage is a characteristic feature. Figure 9 curves are compared with single crystal curves of
shows a characteristic shear stress±strain curve for speci®c orientations. (i) The group of grains show-
two corner positions [111] and [100] of the stereo- ing Type 1 microstructures and having orientations
graphic triangle, for one position on the boundary at or near the boundary between [100] and [111]
between [100] and [111] and for two positions (see Fig. 4) is compared in Fig. 10 with the beha-
within the triangle [29]. The selection of these orien- viour of two crystals numbered 14 and 15 [30] at
tations by Lange and LuÈcke [29] shows that the the boundary between [100] and [111]. These crys-
shear stress±strain behaviour of a single crystal is tals show Type 1 behaviour according to [29],
1834 HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS

Fig. 10. The nominal shear stress±strain curve for grains Fig. 12. The nominal shear stress±strain curve for grains
having Type 1 microstructure compared with the stress± having Type 3 microstructure compared with the stress±
strain curve for two crystals numbers 14 and 15 in Ref. strain curves for a [111] crystals from Ref. [31].
[30] having an orientation on the boundary between [100]
and [111] in the standard triangle [30].

4.3. Polycrystal deformation and polycrystal stress±


which also characterizes the stress±strain behaviour
strain curve
of the grains embedded in the polycrystal. (ii) The
group of grains showing Type 2 microstructures The analysis of the orientation dependence of the
and having orientations near the [100] corner is in microstructural evolution and the shear stress±
Fig. 11 compared with the behaviour of a [100] strain behaviour shows a clear resemblance between
crystal [31]. This crystal shows Type 2 behaviour the behaviour of grains embedded in a polycrystal
according to [29], which also characterizes the and single crystals which deform by double slip and
stress±strain curves of the grains embedded in the multislip. However, for orientations within the stan-
polycrystal. The higher ¯ow stress and the higher dard triangle the behaviour of polycrystal grains
strain hardening rates of the grains in the poly- and single crystals di€ers. The typical single glide
behaviour is not observed in the polycrystal exper-
crystal may be related to the fact that the stress±
iment showing that an embedded grain has to
strain curve for the grains of the polycrystal aver-
deform on more systems to maintain macroscopic
age the behaviour of grains having a spread of
compatibility with its neighbours. Originally this led
orientations around [100], whereas the single crys-
Kocks to suggest [3, 4] that comparisons between
tal curve represents the behaviour of an ideal [100] polycrystal and single crystal behaviour should be
crystal. (iii) The group of grains showing Type 3 based on single crystals deforming by multislip, i.e.
microstructures and having orientations over a in tension in the [100] and [111] orientations. In
large part of the stereographic triangle, including fact, it was found [4] that the polycrystal curve fell
orientations near the [111] corner are compared in between the [100] and [111] curves for both alu-
Fig. 12 with the behaviour of a [111] crystal [31]. minium and iron.
This crystal shows Type 3 behaviour according The present experiments support the idea that the
to [29] which also characterizes the grain multislip of a single crystal in the [111] orientation
embedded in the polycrystal. is relevant to consider in comparisons of embedded
grains having a large orientation spread in the
stereographic triangle which further show a shear
stress±strain behaviour quite comparable to that of
the [111] crystal. Further the experiment has shown
that other slip patterns can ensure macroscopic
strain accommodations of the grains in the poly-
crystal. Fewer slip systems may suce, as in groups
of grains showing a Type 1 microstructure. This is
in general accord with observations of grain subdi-
vision during deformation which has led to the sug-
gestion that volume elements of a grain can deform
with fewer slip systems than the ®ve required in the
Taylor model to ensure strain compatibility. Each
volume element can therefore not ful®l this cri-
Fig. 11. The nominal shear stress±strain curve for grains terion, but neighbouring elements may ful®l it
having Type 2 microstructure compared with a stress± collectively [33]. Deviations from the Taylor model
strain curve for a [100] crystal from Ref. [31]. may also have their cause in unpredicted slip and in
HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS 1835

Fig. 13. Tensile ¯ow stress±strain curves for single crystals of corner orientations ([100] and [111]) from
Ref. 31 and of an orientation crystals numbered 14 and 15 in Ref. [30] on the boundary between [100]
and [111]. (The latter curve has been obtained by transforming the shear stress±strain curve in Fig. 10
by the use of a Taylor-factor equal to 2.8.) The contributions of the three single crystals to the tensile
¯ow stress of a polycrystalline specimen have been weighted and added to give the curve marked
Polycryst. (calc.). On this ®gure is also shown the tensile stress±strain curve for the polycrystal tested in
this work, marked Polycryst. (exp.).

activation of extra slip systems near grain bound- M-factor for an ideal h111i crystal which is 3.67
aries and triple junctions. (18) and the average M-factor for group 3 grains
Based on the observed correlations between poly- which is 3.26. The calculated curve is shown in
crystalline and single crystal behaviour the polycrys- Fig. 13 together with the experimental stress±strain
talline tensile stress±strain curve may be calculated curve for the polycrystalline specimens. Good agree-
based on the stress±strain curves for single crystals, ment is observed both when it comes to the shape
see Fig. 13. This calculation of the stress±strain of the curve and the magnitude of the ¯ow stress,
curve of the polycrystal is based on the assumption although there is a certain overprediction of the
that the contribution for the three groups of grains ¯ow stress at large strain.
of the polycrystal can be added linearly after being
In the calculation of the polycrystalline curve in
weighted based on the volume fractions given in
Fig. 13 a contribution of grain boundary strength-
Table 2. It is further assumed as in the Taylor
ening is not included. This contribution is, however,
model that the tensile strain is the same for the
not large, of the order of 3±4 MPa for the present
embedded grains of the polycrystal as for the bulk
grain size [10].
polycrystalline specimen.
The polycrystalline curve is calculated by using In conclusion, it appears to be a crucial obser-
the crystals numbered 14 and 15 in Ref. [30] and vation that the stress±strain behaviour of grains in
the [100] curve to calculate the contributions of a polycrystal is orientation dependent and shows
grains of group 1 and group 2, respectively. No cor- many similarities with single crystals oriented for
rection is made for the change in the Taylor M-fac- double slip and multislip. To explore the relation-
tor as it is quite similar for the single crystals 2.8 ship between polycrystals and single crystals, the
and 2.45, respectively and for the groups of behaviour of grains embedded in a polycrystal must
embedded grains of the polycrystal 2.76 and 2.40, be studied further and also as a function of the
respectively (from Table 4). For grains of group 3 grain size. Such investigations can be greatly facili-
the h111i curve is used, however, after an M-factor tated, if the grains in the polycrystal can be classi-
correction due to the signi®cant di€erence in the ®ed into groups showing a similar microstructural
and mechanical behaviour. A microstructural analy-
Table 4. M-factors sis can be used in these studies, but a faster tech-
Type of grain M-factor nique is desirable. One such technique is high-
14, 15 crystals* 2.8 (estim.)
energy X-ray di€raction, where the behaviour of in-
[100] crystal 2.45 dividual grains embedded in a polycrystal can be
[111] crystal 3.67 studied both during and after deformation [34]. Of
Group 1 grains 2.76 (aver.)
Group 2 grains 2.40 (aver.) special interest will be the combined observations of
Group 3 grains 3.26 (aver.) microstructural and crystallographic changes during
*See text and Ref. [30]. in situ deformation of embedded grains in poly-
1836 HANSEN and HUANG et al.: MICROSTRUCTURE AND FLOW STRESS

crystals which may then be compared with the 7. Aernoudt, E., Gil, Sevillano J. and Van Houtte, P., in
behaviour of single crystals. Constitutive Relations and Their Physical Basis, ed. S.
I. Andersen, J. B. Bilde-Sùrensen, N. Hansen, T.
Le€ers, H. Lilholt, O. B. Pedersen and B. Ralph. Risù
5. CONCLUSIONS National Laboratory, Roskilde, 1987, p. 1.
8. Dawson, P. R., Beaudoin, A. J. and Mathur, K. K.,
Pure aluminium polycrystals have been strained in Numerical Predictions of Deformation Processes and
in tension at room temperature and based on a the Behaviour of Real Materials, ed. S. I. Andersen, J.
B. Bilde-Sùrensen, T. Lorentzen, O. B. Pedersen and
microstructural analysis the following conclusions N. J. Sùrensen. Risù National Laboratory, Roskilde,
have been reached: 1994, p. 33.
9. Hansen, N., Metall. Trans., 1985, 16A, 2167.
. Three groups of grains have been identi®ed show- 10. Hansen, N., Acta Metall., 1977, 25, 863.
ing di€erent types of deformation microstructures 11. Liu, Q., J. Appl. Cryst., 1994, 27, 755.
and a correlation has been observed between 12. Huang, X. and Hansen, N., Scripta mater, 1997, 1.
microstructure and grain orientation. The dislo- 13. Underwood, E., Quantitative Stereology, Addison
cation density has been estimated for the three Wesley, Reading, MA, 1970, p. 274.
14. Hansen, N., in Numerical Predictions of Deformation
groups of grains and the shear stress±strain re- Processes and the Behaviour of Real Materials, ed. S. I.
lationship has been obtained by taking the shear Andersen, J. B. Bilde-Sùrensen, T. Lorentzen, O. B.
stress proportional to the square root of the dis- Pedersen and N. J. Sùrensen. Risù National
location density. Each of the three groups of Laboratory, Roskilde, 1994, p. 325.
grains shows a characteristic shear stress±strain 15. Read, W. T. and Shockley, W., Phys. Rev., 1950, 78, .
16. Juul, Jensen D. and Randle, V., in Materials
relationship. Architecture, 1989, ed. J. B. Bilde, N. Sùrensen, N.
. The deformation behaviour of single crystals and Hansen, Jensen D. Juul, T. Le€ers, H. Lilholt and O.
grains embedded in a polycrystal has been corre- B. Pedersen. Risù National Laboratory, Roskilde,
lated by taking into account the di€erence in the 1989, p. 103.
slip pattern between a single crystal and an 17. Bunge, H. J. and Esling, C., Quantitative Texture
Analysis, DGM, Oberursel, 1982, p. 551.
embedded grain in a polycrystal. The correlations 18. Chin, G. Y., in Work Hardening in Tensison and
obtained have allowed the tensile stress±strain Fatigue, ed. A. W. Thompson. American Institute of
curve for the polycrystalline specimen to be calcu- Mining, Metallurgical and Petroleum Engineers, New
lated with good accuracy from single crystal data York, 1977, p. 45.
19. Nabarro, F. R. N., Basinski, Z. S. and Holt, D. R.,
which are weighted based on a quantitative tex-
Adv. Phys., 1964, XII, 193.
tural analysis of the polycrystal. 20. Neuhaus, R. and Schwink, Ch, Phil. Mag., 1992, 65,
1463.
21. Steeds, J. W., Proc. Roy. Soc. A, 1966, 292, 343.
22. GoÈttler, E., Phil. Mag., 1973, 28, 1087.
AcknowledgementsÐWe thank Drs B. Bay, J. B. Bilde- 23. Ambrosi, P., GoÈttler, E. and Schwink, Ch, Scripta
Sùrensen, D. A. Hughes, D. Juul Jensen, Q. Liu, Y. L. Metall., 1974, 8, 1093.
Liu, and T. Le€ers for many fruitful discussions. J. 24. Kawasaki, Y., Jap. J. Appl. Phys., 1979, 18, 1429.
Lindbo is gratefully acknowledged for his very skilled 25. Kawasaki, Y. and Takeuchi, T., Scripta Metall., 1980,
preparation of the TEM foils. Finally, we thank T. Skov 14, 183.
for drawing the ®gures and E. Sùrensen for typing the 26. Kawasaki, Y., in Strength of Materials, ed. H.
manuscript. The present work is performed under the Oikawa, K. Maruyama, S. Takeuchi and M.
Danish Materials Technology Programme ®nanced by the Yamaguchi. The Jap. Inst. of Metals, 1994, p. 187.
Danish Agency for Development of Trade and Industry, 27. Liu, Q. and Hansen, N., Phys. Stat. Sol. (b), 1995,
The Danish Natural Science Research Council and the 149, 187.
Danish Technical Research Council. 28. LuÈcke, K. and Lange, H., Z. Metallk., 1952, 43, 55.
29. Lange, H. and LuÈcke, K., Z. Metallk., 1953, 44, 183.
REFERENCES 30. Staubwasser, W., Acta Metall., 1959, 7, 43.
31. Fortunier, R., Orlans-Joliet, B., Montheiller, F. and
1. Sachs, G., Z. Verein Deut. Ing., 1928, 72, 734. Driver, J. H., in Constitutive Relations and Their
2. Taylor, G. I., J. Inst. Met., 1938, 62, 307. Physical Basis, ed. S. I. Andersen, J. B. Bilde-
3. Kocks, U. F., Acta Metall., 1960, 8, 345. Sùrensen, N. Hansen, T. Le€ers, H. Lilholt, O. B.
4. Kocks, U. F., Metall. Trans. AIME, 1970, 1, 1121. Pedersen and B. Ralph. Risù National Laboratory,
5. Bishop, J. F. W. and Hill, R., Phil. Mag., 1951, 42, 1987, p. 317.
414 and 1298. 32. Winter, G., Juul, Jensen D. and Hansen, N., Acta
6. Le€ers, T., in Deformation of Polycrystals: mater., 1997, 45, 5059.
Mechanisms and Microstructure, ed. N. Hansen, A. 33. Bay, B., Hansen, N., Hughes, D. A. and Kuhlmann-
Horsewell, T. Le€ers and Hans Lilholt. Risù National Wilsdorf, D., Acta Metall. Mater., 1992, 40, 205.
Laboratory, Roskilde, 1981, p. 55. 34. Juul, Jensen D., (Private communication).

S-ar putea să vă placă și