Sie sind auf Seite 1von 9

Materials Science and Technology

ISSN: 0267-0836 (Print) 1743-2847 (Online) Journal homepage: http://www.tandfonline.com/loi/ymst20

Effects of strain rate and test temperature on flow


behaviour and microstructural evolution in AA
8090 AlLi alloy
W. Fan, B.P. Kashyap & M.C. Chaturvedi
To cite this article: W. Fan, B.P. Kashyap & M.C. Chaturvedi (2001) Effects of strain rate and test
temperature on flow behaviour and microstructural evolution in AA 8090 AlLi alloy, Materials
Science and Technology, 17:4, 431-438, DOI: 10.1179/026708301101510005
To link to this article: http://dx.doi.org/10.1179/026708301101510005

Published online: 19 Jul 2013.

Submit your article to this journal

Article views: 15

View related articles

Citing articles: 10 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=ymst20
Download by: [IIT Indian Institute of Technology - Mumbai]

Date: 25 August 2016, At: 03:16

Effects of strain rate and test temperature on


flow behaviour and microstructural evolution in
AA 8090 Al Li alloy
W. Fan, B. P. Kashyap, and M. C. Chaturvedi
Tensile specimens of superplastic forming grade AA 8090 Al Li alloy were deformed at constant strain rates in the
range 161025 161022 s21 and at constant temperatures in the range 298 843 K, to investigate their effects on
the nature of stress strain (s e) curves and on the concurrent microstructures, substructures, and microtextures.
The s e curves exhibited flow hardening in the early part of deformation, the rate of which was found to increase
with an increase in strain rate and decrease in temperature. Grain growth, cavitation, dislocation interactions, and
texture weakening were observed to occur during superplastic deformation. The early part of deformation involved a
substantial dislocation slip contribution to flow hardening and to the mechanism for superplastic flow. However, the
microstructural evolution facilitated the conventional mechanism of grain boundary sliding and its accommodation
by the diffusional process, as suggested by the strain rate sensitivity index and activation energy determined.
MST/4804
At the time the work was carried out the authors were in the Department of Mechanical and Industrial Engineering, University of
Manitoba, Winnipeg, Canada, R3T 5V6 (mchat@cc.umanitoba.ca), where Professor Kashyap was on leave from the
Department of Metallurgical Engineering and Materials Science, Indian Institute of Technology, Bombay, Mumbai 400 076,
India. Manuscript received 1 May 2000; accepted 5 July 2000.
# 2001 IoM Communications Ltd.

Introduction
The need for materials of high strength and low density (i.e.
having high specific strength) in aerospace applications has
led to the development of several aluminium alloys. While
alloying, heat treatment, and mechanical working are
known to improve the strength of aluminium alloys, the
procedure is usually accompanied by a significant loss in
ductility. This limits the formability of high strength
aluminium alloys. In fact, most of the commercially
promising aluminium alloys are known to exhibit much
lower ductility than that of aluminium at ambient
temperature. However, after suitable thermomechanical
processing, several commercial aluminium alloys can be
made superplastic, whereby they exhibit tensile ductility
exceeding several hundred per cent. One aluminium alloy to
have attracted a large number of research activities leading
to successful implementation of superplasticity in aerospace
applications in the past decade is AA 8090 Al Li alloy.
This alloy has been extensively studied to characterise
superplastic flow in terms of strain rate sensitivity index and
ductility,1,2 mechanisms for superplastic flow,3,4 superplastic formability,5 and texture evolution.6 8
Lloyd1 and Ridley et al.2 reported maximum ductility to
be associated with maximum values of strain rate sensitivity
.
index m at intermediate strain rates e and elevated
1
temperatures T. While Lloyd found the maximum ductility
.
to be 640% at e~1.761024 s21 and T~753 K, Ridley et al.2
pointed out that concurrent grain growth during superplastic deformation shifted the strain rate for maximum
ductility and m towards a lower value. The adverse effect of
concomitant grain growth on ductility, however, could be
compensated by performing tensile tests at a constant
crosshead speed (constant initial strain rate) instead of at a
constant true strain rate, whereby the strain rate automatically decreases with increasing strain. The constant
crosshead speed type tests at 793 K thus led to a maximum
ductility of more than 850% with an initial strain rate of
y161024 s21. In constant true strain rate tests, such grain
growth is also responsible for an apparent increase in flow
ISSN 0267 0836

stress with increase in strain. Unless some correction based


on grain growth is applied to the experimental values of
flow stress, a wide variation in magnitude of the parameters
of the constitutive relationship for superplastic deformation
can be envisaged. This is because the development of the
constitutive relationship per se and the evaluation4 of its
parameters are based on a condition of steady state. In
addition to grain growth, AA 8090 Al Li alloy is also
known to exhibit cavitation1 and textural evolution6 during
superplastic deformation. While cavitation generally
becomes significant only at large strains, the texture changes
more rapidly in the early part of deformation, as
summarised by Bowen.6 All of these microstructural
changes can influence superplastic behaviour to varying
extents. Studies of the mechanisms for superplastic
deformation of AA 8090 Al Li alloy did not support the
conventional mechanisms unambiguously. The initial
microstucture (elongated grains and low angle grain
boundaries) and the evolution of microstructure and
crystallographic preferred orientation did not correspond
with grain boundary sliding and grain rotation. This led
Blackwell and Bate3 to suggest that intragranular slip
participated in superplastic deformation more than merely
as an accommodating process to grain boundary sliding. Pu
et al.4 reported high temperature and low temperature
superplasticity in AA 8090 Al Li alloy. Using experimental values for the parameters of the constitutive relationship
(such as strain rate sensitivity index and activation energy),
they supported the conventional mechanism for superplastic deformation, with grain boundary sliding acting as a
primary deformation mechanism and dislocation creep as
an accommodating process. However, despite a large
number of publications on different aspects of superplasticity in this material, the variation in microstructural
parameters and their influence on the nature of the stress
strain curves during superplastic deformation does not seem
to have received adequate attention. Therefore, the aim of
the present work was to investigate the flow behaviour of
AA 8090 Al Li alloy and correlate it with the concurrent
microstructural evolution during superplastic deformation
over a range of strain rates and temperatures.
Materials Science and Technology

April 2001 Vol. 17

431

432 Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090

were first mechanically polished, and then electropolished


using 25% nitric acid and 75% methanol at 298 K using an
18 V power supply. Specimens for transmission electron
microscopy (TEM) were prepared by mechanical thinning
followed by electropolishing. The electrolyte used was the
same as for OIM but the electropolishing condition was
different, namely 253 K and 17 V. Examination by TEM
was done using a Jeol 2000FX microscope operating at
200 kV.

Results

1 Three-dimensional composite microstructure to represent initial microstructure obtained on heating to and


soaking at test temperature of 803 K for 20 min

Experimental procedure
The superplastic forming (SPF) grade of Al Li alloy of
composition (wt-%) Al 2.5Li 1.4Cu 1.2Mg 0.11Zr was
obtained from Alcan in the form of a sheet 1.8 mm in
thickness. Tensile specimens of 10 mm gauge length and
5 mm gauge width were machined, with the tensile axis
being kept parallel to the rolling direction. True constant
strain rate tensile tests were carried out with a computerised
Instron universal testing machine. Test temperatures were
controlled to an accuracy of 1 K. Before straining the
specimens, 20 min was allocated for heating to and soaking
at the test temperatures. After deformation, the specimens
were rapidly quenched within 100 ms using a specially
designed in situ quenching facility.9
Metallographic specimens were mechanically polished.
Etching was done with Kellers reagent, comprising 1%HF,
1.5%HCl, and 2.5%HNO3 in water. Grain size was
measured in the longitudinal section by the mean linear
incept method. This was carried out manually during
examination of the microstructure using a Leitz TAS plus
image analyser. Orientation image microscopy (OIM) was
carried out on the longitudinal surface using a TSL
orientation image microscope attached to a Jeol 840
scanning electron microscope. For this, the specimens

The representative microstructure of the tensile specimen


just before the start of deformation at 803 K is shown in
Fig. 1. It indicates fine elongated grains in the midthickness
layer in the longitudinal and transverse sections, sandwiched between two surface layers with a coarse and nearly
equiaxed microstructure. The grain size in the surface layer
was measured to be 5.6 mm. For the elongated grains in the
midthickness layer, measurements taken both along the
direction of grain elongation and perpendicular to it gave a
mean grain size of 5.0 mm. In these two types of layer, the
texture components were also found to be different. The
surface layer was dominated by S {123}n634m type texture
whereas the centre layer was dominated by Bs {110}n112m
type texture. The major texture components present in top
and bottom surface layers shown in Fig. 1 were cube (Cb)
{100}n001mvBs {110}n112mvcopper (Cu) {112}n111m
vS{123}n634m, in increasing order of their volume
fractions. In the centre layer, the texture components
present in increasing order of their volume fractions were
CuvCbvSvBs. Figure 2 shows TEM images illustrating
the variation in substructures in the surface region
(examined in the sheet normal direction, Fig. 2a) and in
the midthickness region (examined on the longitudinal
surface, Fig. 2b). As discussed below in the Microstructural
evolution subsection, the midthickness layer contained
predominantly subgrains whereas the surface layer exhibited well developed grains. As the microstructure, substructure, and texture were different in the surface and
midthickness layers, the full thickness of the sheet was
divided into three layers: two surfaces symmetrical about a
centre layer, constituting a compositelike structure as
shown in Fig. 1. The flow behaviour and microstructural
evolution during superplastic deformation of tensile specimens containing such initial microstuctures are described
below. However, the initial microstructure may vary slightly
depending on the test temperature, although the period of

b
a rolling; b midlongitudinal

2 Substructures seen on given surfaces representing initial condition as given in Fig. 1

Materials Science and Technology

April 2001 Vol. 17

Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090 433

(a)

4 Variation in normalised stress s/E at onset of plastic


strain e~0.0 as function of homologous test tempera.
ture T/Tm at two strain rates e~161023 and
24 21
1610 s

(b)

temperatures, the flow stresses were greater at the higher


strain rate of 161023 s21.
In another series of experiments, the effect of strain rate
on the nature of the stress strain curves was investigated
over the strain rate range 161025 161022 s21 at six test
temperatures in the range 623 803 K. The stress strain
curves at two temperatures are shown in Fig. 5. As
expected, the flow stress increases with increasing strain
rate. Furthermore, within the strain rate range investigated,
flow hardening is noted to increase whereas the elongation
at failure can be seen to decrease with an increase in strain

(a)

.
.
a e~161023 s21; b e~161024 s21

3 True stress true strain curves at various test temperatures ranging from room temperature to 843 K

20 min was kept constant for heating to and soaking at the


test temperatures before actual straining.

FLOW BEHAVIOUR
Individual tensile specimens were deformed to failure in the
temperature range 298 843 K at the constant strain rates
161023 s21 and 161024 s21. The resulting stress strain
curves at these two strain rates are plotted in Fig. 3. As
expected, the flow stress increases with a decrease in test
temperature. Figure 3 also indicates a decrease in elongation to failure and an increase in flow hardening rate with a
decrease in test temperature. At higher test temperatures of
773 K and above, the s e curves suggest that a steady state
of flow stress has been attained, whereby the stress no
longer varies as a function of strain. The transitional flow
behaviour, beyond which the flow stress becomes independent of strain, appears to extend to a larger strain level as
the strain rate of deformation is increased.
The flow stresses at various temperatures were normalised (s/E) by Youngs moduli E at the corresponding
temperatures, according to the work of Pu et al.,4 and were
plotted as a function of homologous temperature T/Tm,
where Tm is the absolute melting point, at the fixed strain
levels. As shown in Fig. 4, at the onset of plastic strain, the
decrease in flow stress with increasing temperature was
found to be less sensitive to temperature towards the lower
and upper ends of the temperature range investigated. At all

(b)

a T~623 K; b T~803 K

5 True stress true strain curves at various strain rates

Materials Science and Technology

April 2001 Vol. 17

434 Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090

Ln (TRUE STRESS) (MPa)

(a)

(b)

8 Plot of strain rate sensitivity index m as function of


test temperature: values of m were determined from
Fig. 7 using regression analysis

Ln (STRAIN RATE) (s_1)


a T~773 K; b T~803 K

6 Plot of ln(stress) ln(strain rate) at various strain levels

Ln (TRUE STRESS) (MPa)

rate. At certain levels of strain, the stress strain curves also


suggest the occurrence of steady state flow behaviour at the
lower strain rates and the higher test temperature.
From the stress strain curves such as those shown in
Fig. 5, stresses at fixed strain levels were read at the various
.
strain rates to plot ln s ln e data. At various strain levels in
the range from the onset of plastic strain e~0.0 to e~1.5,
such plots were examined at strain intervals of 0.1. At
.
strains of less than 0.3, the ln s ln e plot did not follow the
expected linear behaviour because of the large and somewhat unsystematic flow hardening exhibited by the stress
strain curves such as shown in Fig. 5. In fact, the slope

appeared to be negative at intermediate strain rates.


.
However, ln s ln e plots at larger strains were found to
follow the normal trend of superplastic behaviour. Exam.
ples of ln s ln e plots at two test temperatures of 773 and
.
803 K are illustrated in Fig. 6. The slope of the ln s ln e
curve is known as the strain rate sensitivity index m.
According to the nature of the stress strain curves (Fig. 5),
m would be expected to vary with strain, but the variation in
m at the higher test temperatures and the larger strains was
not very significant. For example, in the strain range
0.30 1.50 and at the test temperature 803 K, the values of
.
m, obtained by applying regression analysis to the ln s ln e
data, were found to vary between 0.41 and 0.45. However,
.
ln s ln e plots at different temperatures clearly showed a
variation in slope as a function of test temperature. Such a
.
ln s ln e plot at a selected strain of 0.70 is shown in Fig. 7.
It can be seen that there is not only an increase in flow stress
with a decrease in test temperature but also a change in the
value of m. The variation in m as a function of test
temperature is plotted in Fig. 8, which shows that the
maximum value of my0.45 occurs at the temperature of
773 K.

MICROSTRUCTURAL EVOLUTION
To investigate microstructural evolution, metallographic,
TEM, and OIM specimens were prepared from the gauge
and shoulder sections of the tensile specimens after
deformation to various strain levels. Also included were
those specimens obtained by heating to and soaking at the
test temperatures, before straining. Grain growth, evolution
of the midthickness elongated grains into equiaxed grains,
cavitation, dislocation interactions, texture evolution, and
the transformation of low angle boundaries into high angle
boundaries were observed. While some of these features
were quantified, others were evaluated only qualitatively.
Described below are some of the typical microstructural
observations made in the present study.

Grain growth

Ln (STRAIN RATE) (s_1)


7 Plot of ln(stress) ln(strain rate) at selected strain level
of 0.7 and various temperatures in range 623 843 K

Materials Science and Technology

April 2001 Vol. 17

Grain growth, cavitation, and transformation from elongated grains into equiaxed grains are illustrated in Fig. 9
(cf. Fig. 1). The micrographs of longitudinal sections of the
tensile specimens shown in Fig. 9 represent the microstructures developed upon deformation to varying strain
levels, under the temperature and strain rate condition of
803 K and 161023 s21. After deformation to a large strain,
the microstructure became more or less uniform on all three
mutually perpendicular surfaces (Fig. 10).
Grain growth that occurred at the various test temperatures is plotted in Fig. 11, which shows the grain size results

Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090 435

a e~0.5; b e~0.75; c e~1.0

9 Microstructures of longitudinal sections illustrating grain growth, change from elongated grains to equiaxed grains,
and cavitation with progress in superplastic deformation at temperature of 803 K and strain rate of 161023 s21

of heating to and soaking at the test temperatures dI, grain


size in the shoulder sections dS, and grain size in the gauge
sections dG of the specimens. For the deformed specimens,
the strain rate was 161023 s21 and the strain was 1.0. From
the data of these three basic curves, the static and dynamic
grain growth at various temperatures was evaluated as
follows. The difference between dS and dI yields the
contribution of the time involved in deformation at the
test temperature dSt, which is the same as static annealing.
The difference between dG and dS yields the contribution of
strain alone to grain growth de at elevated temperatures,
which is same as the effect of dynamic annealing. Also
plotted in Fig. 11 are the grain sizes contributed by such
static and dynamic annealing. It can be seen that grain
growth is rapid at temperatures above 803 K. Also,

10 Three-dimensional
microstructure
of
specimen
deformed to e~1.75 at temperature of 803 K and
23 21
strain rate of 1610 s

deformation induced grain growth de contributes more to


the observed grain sizes than grain growth contributed by
static annealing alone dSt.
The grain sizes obtained during superplastic deformation
in the gauge sections are contributed by static and dynamic
annealing (i.e. dG~dStzde) during superplastic deformation. Grain growth behaviour as a function of strain at
various test temperatures is shown in Fig. 12. It can be seen
that grain sizes at all strain levels increase with an increase
in test temperature. The variation in grain size with strain,

11 Plot of grain size as function of test temperature,


showing grain growth during heating to and soaking
at various test temperatures before straining dI, grain
growth in shoulder dS and gauge dG sections upon
tensile testing to a strain of 1.0, and contributions of
static dSt
. and dynamic de annealing for period of
testing: e~161023 s21

Materials Science and Technology

April 2001 Vol. 17

436 Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090

12 Plot of grain size in gauge section


. as function of
strain at various test temperatures: e~161023 s21

based on regression analysis, suggested the following


relationship
dG ~Ken

: :

: (1)

The values of grain growth coefficient K and strain


exponent of grain growth n at the various test temperatures
are listed in Table 1.

Substructural evolution
Transmission electron microscopy was carried out on the
tensile specimens deformed to e~0.5 at the strain rate of
161023 s21 and at constant test temperatures of 623, 723,
803, and 843 K. For this, thin foils of the midthickness layer
were prepared, and they were examined in the normal
direction of the original sheet. Three examples are shown in
Fig. 13 to illustrate the effect of test temperature on the
deformed substructures. At 623 K, T2, b9, silicon rich, and
iron rich phases were found to be prevalent (Fig. 13a).
However, at higher test temperatures of 773 K and above,
T2 was found to have dissolved but b9, silicon rich, and iron
rich phases were still present (Fig. 13b and c). Figure 13
shows evidence of dislocation activities during superplastic
deformation. However, it remains to be examined whether
the appearance of dislocations is enhanced by the presence
of precipitates or some thermal contraction associated with
the fast rate of cooling. The extent of contraction can be
evaluated from the following analysis. The coefficient
of thermal expansion for aluminium is 23.661026 K21
and the cooling rate to freeze the substructure was
y1273 K s21. This can introduce a compressive strain
(despite being physically unloaded upon completion of
tensile tests) at a strain rate of y361022 s21. For the test
temperature of 843 K, for example, the compressive strain
can be substantial (ey20.019) and can influence the prior
dislocation structures.

Texture evolution
Tensile specimens were deformed to selected strain levels at
.
T~803 K and e~161023 s21, to follow the texture
evolution. Some of the results have been presented earlier.10
Table 1 Values of K and
. n of equation (1) at various test
temperatures: e~161023 s21
T, K

K, mm

723
803
823
843

1.52
2.15
2.25
2.52

0.28
0.21
0.44
0.34

Materials Science and Technology

April 2001 Vol. 17

c
a T~623 K; b T~773 K; c T~843 K

13 Precipitate and dislocation structure in specimens


deformed to e~0.5 at given test temperatures:
.
e~161023 s21

However, it may be restated here that substantial texture


weakening occurred during superplastic deformation. A
decrease in the maximum intensity of pole figures derived
from the midthickness layer with superplastic deformation
is shown in Fig. 14. Values of the maximum intensities were
obtained from the combined list of intensity variations
given for (111), (110), and (100) pole figures in the output of
the TSL orientation image microscope. Furthermore, as
pointed out for the initial microstructure (Fig. 1), the
midthickness layer contained fine elongated grains, whose

Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090 437

14 Variation in maximum intensity. of pole figures (PF)


as function of strain: T~ 803 K, e~161023 s21

misorientation angles h suggested them to be predominantly


subgrains. Histograms showing the distribution of h values
at various strain levels10 delineated a change from low angle
(hf15) to high angle (ho15) boundaries. From these
histograms, the proportion of low angle boundaries was
obtained as a function of strain. This is plotted in Fig. 15. It
can clearly be seen that the change from low angle to high
angle boundaries occurs rapidly in the early part of
deformation. Then, at large superplastic strains, the
proportion of low with respect to high angle boundaries
tends to attain a steady state, with only y6% low angle
boundaries.

Discussion
FLOW HARDENING
The stress strain curves shown in Figs. 3 and 5 indicate
flow hardening in the early part of deformation, which is
more pronounced towards the lower temperatures and the
higher strain rates. At low temperatures, hardening is
attributed to conventional dislocation related strain hardening,11 but flow hardening during superplastic deformation is attributed to concurrent grain growth. In the present
study, no noticeable grain growth was observed up to the
test temperature of 623 K, but rapid grain growth occurred
at temperatures of 803 K and above. Therefore, flow
hardening at the test temperatures from ambient to 623 K
is suggested to be the result of dislocation dislocation
and dislocation precipitate interactions (Fig. 13a). At
higher test temperatures, concurrent grain growth contributed to flow hardening. However, the evaluation of
the flow hardening contribution of the instantaneous
grain size during superplastic deformation, according
to a method discussed elsewhere,12 suggested that grain
growth could contribute only partly to the observed
extent of flow hardening. Therefore, some flow hardening
in the present material even towards the upper temperature
limit of testing (843 K) may have come from other sources.
One such source appears to be the conventional strain
hardening mechanisms, as suggested by the presence of
significant dislocations (Fig. 13b and c). Although no such
systematic analysis was done for the variation in flow
hardening with strain rate, the greater flow hardening rate
at higher strain rates (Fig. 5) may also be associated with
conventional strain hardening. This is based on observations that the flow behaviour and nature of microstructural
evolution are similar at higher strain rates and lower
temperatures.13,14

15 Variation in proportion of low angle boundaries


.
(LAB) as function of strain: T~803 K, e~161023 s21

STRAIN INDUCED MICROSTRUCTURAL


EVOLUTION
Grain growth and the weakening of texture are substantially
enhanced during superplastic deformation, compared with
changes occurring during static annealing.15,16 In the present
work, the texture components Bs and S (in both the surface
and centre layers) underwent continuous weakening during
superplastic deformation. If the grains are equiaxed, superplastic deformation is dominated by grain boundary sliding,
which is accommodated by diffusion and dislocation creep.
These accommodating mechanisms are manifested in the
form of grain boundary migration, grain rotation, grain
switching, and grain rearrangements during superplastic
deformation. While the microstructure in the surface layer
(Fig. 1) is conducive to such a mechanism for superplastic
deformation from the start, the centre layer may be
dominated by dislocation mechanisms in the early part of
deformation. Once the entire microstructure becomes uniform, the same mechanism of superplastic deformation may
be operative in the bulk of the tensile specimen. Every step of
grain boundary sliding during superplastic deformation
causes local stress concentrations at triple points or at
irregularities such as ledges in the grain boundaries, which
creates a condition for the grain boundary to move normal
to itself. Thus, the process of grain growth is accelerated.
However, there appeared to be an increased tendency for
grain growth above T~803 K (Fig. 11), although the
tendency for the material to exhibit superplasticity appeared
to decline above this temperature (Fig. 8). Grain growth at
such high temperatures may be associated not only with the
mechanism for superplastic deformation, but also with
decreasing grain boundary precipitate interactions. In the
present material, the fine grain size is maintained by the
pinning of grain boundaries by precipitates. As some of
the precipitates are dissolved at higher temperatures
(Fig. 13), the reduced number of barriers can lead to
higher grain growth. Also, accelerated grain growth at
higher temperatures can arise from the mechanism for grain
growth per se, which follows an exponential type relationship with temperature17 whereas m does not.
The presence of strong textures (S, Cu, and Bs in the
surface layer; Bs and S in the centre layer) at the beginning
of deformation suggests that the dislocation mechanism
might not be acting merely as an accommodating process
for grain boundary sliding. Instead, the dislocation
mechanism itself could be dominant over grain boundary
sliding during superplastic deformation. In fact, the
presence of a large proportion of low angle boundaries
(Fig. 15) is not favourable for grain boundary sliding.18
However, during high temperature deformation, the
Materials Science and Technology

April 2001 Vol. 17

438 Fan et al. Effects of strain rate and temperature on flow and microstructure of AA 8090

elongated grains break into equiaxed grains which, in turn,


can promote grain boundary sliding and grain rotation.
This process can then contribute to a weakening of texture
(Fig. 14) and an increase in misorientation angles of the low
angle boundaries (Fig. 15).

APPARENT PARAMETERS OF CONSTITUTIVE


RELATIONSHIP
In view of the microstructural evolution and strain
sensitivity of the flow stress, the parameters of the
constitutive relationship for superplastic deformation,
developed on the basis of steady state flow, cannot be
unambiguously determined. This problem appears to be
true for a wide range of superplastic materials, in which
cases apparent values of the parameters have been used to
characterise the flow behaviour and understand provisionally the mechanism for superplastic deformation. It is
surprising that the value of m did not vary much with strain
in the present work. There is, however, some effect of
temperature on m (Fig. 8). In the literature on superplasticity,19,20 m is reported to increase with temperature in
some materials whereas in others m is independent of
temperature. Whether m changes with temperature or not,
ductility is invariably reported to increase with temperature.
In the present study, both ductility and m (Figs. 3 and 8)
were found to have maximum values at (intermediate) high
temperatures. The decrease in both of these properties,
towards the upper limit of the temperature range investigated, is attributed to the excessive grain growth (Fig. 11).
Values of activation energy for superplastic deformation
Q were evaluated from the stress values at fixed strains and
strain rates but at varying temperatures, following an
Arrhenius type relationship.19 For example, at the strain
rate of 161023 s21, Q was found to be 88.2 and
78.3 kJ mol21 at the strains of 0.3 and 0.7, respectively.
As the difference between the stresses at any two
temperatures changed with increasing strain (Fig. 3), the
values along the y axis (ln s) of the Arrhenius plot changed
for the same values of x (i.e. 1/T), resulting in a change in
the slope of the ln s versus 1/T plot. Such a variation in
slope yields different values of activation energy.21,22
However, at large strains, the flow stress does not vary
with strain, whereby the activation energy could represent
the flow behaviour of a pseudo-steady state. The values of Q
reported here are comparable to the activation energies for
grain boundary diffusion (84 kJ mol21), as well as dislocation pipe diffusion (82 kJ mol21) in aluminium.23 Thus, the
values of m and Q support the conventional mechanism for
superplastic deformation, which involves grain boundary
sliding and its accommodation by grain boundary diffusion.
This could be possible by the microstructural evolution that
changed the initial layered structure into a uniform
equiaxed microstructure in the early part of deformation.

Conclusions
The study of flow behaviour and microstructural evolution
in AA 8090 Al Li alloy over the strain rate range
161025 161022 s21 and temperature range 298 843 K
led to the following conclusions.
1. Flow hardening occurred in the early part of
deformation, the rate of which increased with an increase
in strain rate and a decrease in temperature. The flow
hardening was suggested to result from the combined effects
of dislocation activity and grain growth.
2. Both strain rate sensitivity index and tensile ductility
were noted first to increase and then to decrease with an
increase in test temperature. However, grain growth
increased continuously with temperature.
Materials Science and Technology

April 2001 Vol. 17

3. During the early part of superplastic deformation,


elongated grains in the midthickness layer changed into
equiaxed grains, with a simultaneous weakening of texture
and increase in the proportion of high angle grain
boundaries.
4. During superplastic deformation, the values of m and
Q were marginally dependent on strain, but with the typical
values of y0.40 and 83.2 kJ mol21, respectively. This
suggests that, subsequent to the initial deformation induced
microstructural homogenisation, the conventional mechanism for superplastic deformation became operative,
as opposed to dominance of the dislocation related
mechanisms.

Acknowledgements
The authors would like to thank the consortium of
Manitoba aerospace industries and the Natural Science
and Engineering Research Council of Canada for financial
support. Technical assistance from D. Mardis and J. Van
Dorp is appreciated.

References
1. d. j. lloyd: Can. Metall. Q., 1986, 25, 257 263.
2. n. ridley, d. w. livesey, and j. pilling: J. Phys. (France),
1987, 48, C-251 256.
3. p. l. blackwell and p. s. bate: Metall. Trans. A, 1993, 24A,
1085 1093.
4. h. p. pu, f. c. liu, and j. c. huang: Metall. Mater. Trans. A,
1995, 26A, 1153 1166.
5. t. r. chen, j. c. huang, y. m. hwang, and j. m. liauo: Scr.
Metall. Mater., 1994, 31, 309 314.
6. a. w. bowen: Mater. Sci. Technol., 1990, 6, 1058 1071.
7. x. h. zeng, m. ahmad, and o. engler: Mater. Sci. Technol.,
1994, 10, 581 591.
8. a. k. singh, g. g. saha, a. a. gokhale, and r. k. ray: Metall.
Mater. Trans. A, 1998, 29A, 665 675.
9. w. fan: PhD thesis, University of Manitoba, Winnipeg,
Canada, 1998.
10. w. fan, m. c. chaturvedi, n. c. goel, and n. l. richards: in
Superplasticity in advanced materials ICSAM-97, (ed. A. H.
tikon-Zurich, Switzerland, Trans.
Chokshi), 563 568; 1997, U
Tech. Publications.
11. m. a. meyers and k. k. chawla: Mechanical metallurgy
principles and applications, 353; 1984, Englewood Cliffs, NJ,
Prentice-Hall.
12. b. p. kashyap and k. k. tangri: Metall. Trans. A, 1987, 18A,
417 424.
13. h. j. mcqueen and j. j. jonas: in Plastic deformation of metals,
(ed. R. J. Arsenault), 393 493; 1975, New York, Academic
Press.
14. w. roberts: in Deformation, processing and structure,
(ed. G. Krauss), 109 184; 1984, Metals Park, OH, American
Society for Metals.
15. b. p. kashyap, a. arieli, and a. k. mukherjee: J. Mater. Sci.,
1985, 20, 2661 2686.
16. b. p. kashyap and m. c. chaturvedi: Mater. Sci. Eng. A, 2000,
A281, 88 95.
17. r. e. reed-hill: Physical metallurgy principles, 202 207;
1964, Toronto, D. Van Nostrand Co.
18. h. gleiter and b. chalmers: Prog. Mater. Sci., 1972, 16, 1
272.
19. k. a. padmanabhan and g. j. davies: in Superplasticity,
71 74; 1980, Berlin, Springer-Verlag.
20. t. g. langdon: Metall. Trans. A, 1982, 13A, 689 701.
21. b. p. kashyap and a. k. mukherjee: Scr. Metall., 1982, 16,
541 546.
22. b. p. kashyap and a. k. mukherjee: Mater. Sci. Technol., 1985,
1, 291 296.
23. h. j. frost and m. f. ashby: Deformation-mechanism maps,
20 29; 1982, Oxford, Pergamon Press.

Das könnte Ihnen auch gefallen