You are on page 1of 7

Journal of Materials Processing Technology 100 (2000) 116122

The strain rate and temperature dependence of the dynamic


impact properties of 7075 aluminum alloy
Woei-Shyan Lee*, Wu-Chung Sue, Chi-Feng Lin, Chin-Jyi Wu
Department of Mechanical Engineering, National Cheng Kung University, Tainan 70101, Taiwan, ROC
Received 24 August 1998

Abstract
The dynamic impact properties of 7075 aluminum alloy are studied experimentally using a split Hopkinson bar. Cylindrical specimens of
10 mm height and 10 mm diameter are compressed dynamically at temperatures ranging from 25 to 3008C and at constant strain rates of
from 103 to 5103 s1. The inuence of strain rate and temperature on the mircrostructural evolution, the fracture mechanisms and the
occurrence of shear localization is investigated. It is found that the compressive stressstrain response depends sensitively on the applied
strain rate and test temperature. Considering the effects of strain rate, temperature, strain hardening, rate sensitivity and thermal softening of
the material, a constitutive equation is used successfully to describe the dynamic impact deformation behavior of 7075 Al alloy.
Microstructural observations reveal that the size of the initial coarse equi-axial grains is reduced as the strain rate and temperature increase
due to dynamic recrystallization. In contrast, the second phase increases in size in response to increasing strain rate and temperature. SEM
observation of the fracture surfaces makes evident an adiabatic shearing mechanism along the fracture planes accompanying crack
formation. # 2000 Elsevier Science S.A. All rights reserved.
Keywords: 7075 Al alloy; Split Hopkinson bar; Strain rate and temperature effects; Adiabatic shearing mechanism; Grain renement

1. Introduction
7075 aluminum alloy is one of the most important engineering alloys and has been utilized extensively in aircraft
structures because of its high strength-to-density ratio. A
considerable amount of work has been carried out on the
plastic ow of this material under low strain rates and
various temperatures [1,2]. However, up to now, there has
been little work concerning the systematic effects of strain
rate and temperature on the plastic ow response, as well as
the evolution of the microstructure, during dynamic impact
deformation. From the deformability viewpoint and for
structural design purposes, it is necessary to characterize
the mechanical properties of 7075 Al alloy over a wide range
of temperatures and strain rates up to impact loading.
The compression split Hopkinson bar (SHPB) is used
widely to determine the mechanical properties of structural
materials under high loading rates [3]. Using this technique,

Corresponding author. Tel.: 886-6-2757575 (ext. 62174); fax: 8866-2352973.


E-mail address: wslee@mail.ncku.edu.tw (W.-S. Lee).

the impact response of metals, alloys and, more recently,


many non-metallic and composite materials, has been studied, and several reviews of this eld have been published
[47]. It is clear that, to a greater or lesser extent, most
materials show a signicant change in mechanical response
under increased strain rates or temperatures. Several
mechanisms such as dislocation damping have been proposed to understand high velocity deformation. It is generally recognized that thermally-activated strain-rate
analysis can provide fundamental insight into temperature
and strain rate effects on stress [8,9].
It is also essential to clarify the structural evolution during
deformation over a wide range of temperatures and strain
rates. Several authors have found that the signicant strainrate and temperature dependence of the ow stress of
materials such as copper and titanium alloy in the high
strain-rate ranges can be understood in connection with the
evolution of the structure during deformation [10,11]. Actually, the mechanical behavior of a material depends not only
on the strain rate and temperature but also on its current
microstructure, and changes in microstructure result in
changes of the plastic ow behavior. Hence, the establishment of more physically-based constitutive models to

0924-0136/00/$ see front matter # 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 9 2 4 - 0 1 3 6 ( 9 9 ) 0 0 4 6 5 - 3

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

describe the complex loading processes of 7075 alloy


requires a knowledge of the coincident inuences of temperature, strain rate and microstructure on the high-strainrate mechanical responses of the alloy.
The objective of this study is to characterize the behavior
of 7075 aluminum alloy during dynamic compression using
a split Hopkinson bar over a temperature range of from 25 to
3008C, at constant strain rates of from 103 to 5103 s1. The
stressstrain relation, the evolution of the microstructure and
the fracture characteristics are discussed in terms of the test
conditions. A constitutive equation expressing the plastic
ow behavior is developed by analysis and regression of the
test results.
2. Experimental procedure
The material tested was a commercially produced 7075
aluminum alloy with the following chemical composition:
Zn 5.65, Mg 2.7, Cu 1.58, Cr 0.2, Fe 0.07, Ti 0.02, Mn 0.05,
balance Al (all in mass pct). This material was supplied as
hot-rolled plates of 12.7 mm thickness, which were formed
in the annealed (O temper) condition and subsequently heat
treated to the T6 temper (1208C for 24 h). Cylindrical
compression specimens of height 10 mm and diameter
10 mm were prepared from the plates with the compression
axis parallel to the rolling direction. All of the specimens
presented a maximum atness defect of 0.01 mm and a
maximum parallelism defect of 0.05 mm between their
faces.
Dynamic compression tests were conducted in the strainrate range of from 103 to 5103 s1 and at temperatures of
25 to 3008C utilizing a split Hopkinson method [3]. Fig. 1
illustrates schematically the apparatus and the measuring
circuits. The striker, the input and the output bars are

117

Ti6A14V rods of 12.7 mm diameter. A good straightness


was achieved and tests have shown the ability of such
materials to support high impact stresses without plastic
deformation or rupture. A short rise time was used for the
incident wave in order to secure a rapid material response.
The stress waves in the input and output bars were sensed by
strain gages cemented at the positions indicated in Fig. 1 and
were recorded in a Nicolet Pro 60 oscilloscope. The stress,
strain, and strain-rate were analyzed in digital form using
one-dimensional stress-wave theory. A full description of
the utilization of this analytic technique for obtaining
dynamic properties data is included in [12].
Temperature variations of between 25 and 3008C on the
split Hopkinson bar were achieved utilizing a clam shell
radiant-heating furnace with an internal diameter of 25 mm
and a heating element of 300 mm length. The sample
temperature was monitored using a ChromelAlumel thermocouple attached directly to the sample. The microstructure was examined using optical and scanning electron
microscopy (SEM). Specimens for optical metallography
were sectioned perpendicular to the loading axis from the
deformed samples, polished conventionally and etched
using an acidic solution (10 ml H2O, 10 ml HNO3, 10 ml
HCl and 5 ml HF). An MeF3 optical microscope was used
for microstructural observations. Fracture specimens were
examined using a Jeol JXA-840 scanning electron microscope operating at an accelerating potential of 20 kV.
3. Results and discussion
3.1. Stressstrain behavior
The true compressive stressstrain curves of 7075 Al
alloy deformed at four temperatures under strain rates of

Fig. 1. Schematic presentation of the compression split Hopkinson bar apparatus.

118

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

Fig. 2. True stressstrain curves of 7075 Al alloy deformed at 1.3103 s1


for all of the temperature tested.

Fig. 4. As for Fig. 2, but for a strain rate of 3.1103 s1.

1300, 2400 and 3100 s1 are shown in Figs. 24, respectively. The ow stress as well as the shape of the ow curves
are sensitively dependent on temperature and strain rate. For
all of the specimens, after initial yielding, the ow stress
increases monotonically with different strain-hardening
rates. Comparing these curves with one another, it is found
that, for a specic strain rate, the ow stress decreases
markedly with temperature. Further, changes in temperature
have a signicant effect on the work-hardening rate: the
degree of work hardening is considerably smaller during
deformation at the higher temperature of 3008C, at this
temperature a nearly horizontal line being obtained, suggesting that the rate of work-hardening is being balanced by
the rate of thermal softening. In contrast, for a xed temperature, the ow stress generally increases as the strain rate
increases due to an increase of dislocation density and the
dislocation multiplication rate. When the ow stress relative
to the temperature is compared to the ow stress relative to
the strain rate, there is no doubt that the effect of temperature
on the ow stress is more pronounced than that of the strain

rate on ow stress. Flow softening appears more clearly as


the temperature increases, even when the specimen is loaded
at a high strain rate.

As can be seen from the stressstrain curves shown in


Figs. 24, the strain rate has obvious effects on the ow
stress for all temperatures. Fig. 5 shows a plot of ow stress
versus log strain-rate at various temperatures for a xed true
strain of 0.1. It is clear that the relationship between the true
stress and the strain rate at any temperature is represented by
a single straight line, the slope of which gives information
about the rate sensitivity of the ow stress. This type of
stressstrain rate relationship has been reported for many
metals [1318]. It is shown that, over a wide range of strain
rates, two regions of strain-rate sensitivity exist. When the
strain-rate is in the range e_ < 102 103 s1 , the strain rate
has only a slight inuence on the ow stress. In this range,
the stress is proportional to the logarithm of the strain rate.

Fig. 3. As for Fig. 2, but for a strain rate of 2.4103 sl.

Fig. 5. Inuence of the strain rate on the ow stress at a constant true


strain of 0.1 as a function of temperature.

3.2. Strain-rate effect

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

119

This is generally acknowledged to be a consequence of the


role of thermal activation in the control of deformation
mechanics. Above 103 s1, the ow stress increases more
rapidly with strain rate, with an approximately linear relationship. This behavior can be interpreted to be an indication
that velocity-dependent ``drag'' on dislocation motion is
causing an additional limitation to deformation. On the basis
of the data in Fig. 5, it is clear that, under the present testing
conditions, the deformation behavior of 7075 Al alloy is
dominated by a dislocation viscous drag mechanism. In
addition, at 3008C, the ow stress is found to be relatively
less sensitive to the strain rate, which suggests that thermal
softening plays a major role in the strain-rate effect for
higher temperature regimes.
3.3. Temperature effect
A substantial decrease in the ow stress with increasing
temperature for 7075 Al alloy is observed in Figs. 24.
Indeed, the manner in which the ow stress decreases with
increasing temperature, at a given strain level, can be seen
more clearly by plotting the ow stress directly as a function
of the test temperature. Figs. 6 and 7 illustrate such behavior
for the two strain rates of 1.3103 and 3.1103 s1. As can
be seen, for both of the strain rates, the ow stress decreases
moderately up to about 2008C and then decreases fairly
dramatically. This difference in slopes is sufciently large as
to imply that different temperature sensitivity and microstructural evolution appear in the material at theses two
temperature ranges. With regard to the effect of strain rate on
the temperature sensitivity, it is found that the ow stress
variations with temperature for both of the strain rates are
very similar. This observation demonstrates clearly that the
temperature sensitivity of the tested material is independent
of strain rate.
Actually, on the micro-scale, the slope of the ow stress
versus the temperature indicates a variation of microstruc-

Fig. 7. Variation of the ow stress as a function of the temperature for a


strain rate of 3.1103 s1 for different true strains.

ture and an augmentation of the dislocation annihilation rate


with increasing temperature. Since 7075 Al alloy is strengthened by precipitates that form during the decomposition of a
supersaturated solid solution, the strength of this material
results from lattice strain interactions between precipitates
and dislocations, and is dependent highly on the type,
distribution and size of the precipitates. During high-temperature loading, dissolution of the precipitates is involved
in the thermally-activated process. Increasing the temperature accelerates precipitate dissolution, the rate of dislocation annihilation and, subsequently, the rate of softening.
These physical and metallurgical reactions result in a reduction of ow stress and, correspondingly, the material
becomes relatively softer and weaker. Although high rate
deformation leads to an enhancement of strength due to the
increase of work hardening and dislocation density, it is
obvious visually that the deformation resistance of 7075 Al
alloy at the high temperature range is dominated by the
temperature effect.
3.4. Deformation constitutive equation

Fig. 6. Variation of the ow stress as a function of the temperature for a


strain rate of 1.3103 s1 for different true strains.

Computer simulations have been used extensively to


predict the responses of complex structures at high strain
rates. In order to obtain realistic data from these simulations,
it is necessary to formulate an accurate constitutive relationship that describes the material behavior over a range of
strains, strain rates, and temperatures. Many constitutive
equations have been developed to describe the behavior of
materials under dynamic loading. Some of these models
describe only the behavior of the yield stress with strain-rate
changes, whilst others describe strain and strain-rate hardening effects without softening effects. A collection of these
equations is discussed by Meyer [19]. Kobayashi and Dodd
[20] proposed the following equation with a term for temperature softening:
s s0 en e_ m 1 bDT

(1)

120

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

Fig. 8. Comparison between the predicted and measured stressstrain


curves for a strain rate of 3.1103 s1.

where s is the ow stress, e the strain, n the work-hardening


coefcient, e_ the strain rate, m the strain-rate sensitivity
index, T the temperature and s0 and b are constants.
This equation is used to predict the high rate and high
temperature deformation behavior of 7075 Al alloy. The data
in the plots of Figs. 24 are used to determine the coefcients and exponents of the constitutive equations by a
regression analysis technique that determines the best-t
parameters for all of the data for a given alloy. To t the
experimental data to Eq. (1), the values of four specic
material parameters for 7075 Al alloy are determined as
s0510.58 MPa, n6.8102, m0.144 and b1.35
103. By putting these four obtained material parameters
into the proposed constitutive equation, the high rate and
high temperature deformation behavior of 7075 Al alloy
under mathematical conditions that are equivalent to the
present actual experimental test conditions is predicted. Fig.
8 shows an example for specimens tested at a strain rate of
3.1103 sl, for which a good agreement with the experimental data using Eq. (1) is noted.
3.5. Evolution of the microstructure during impact
deformation
The specimens after impact loading are examined by
optical microscopy in order to determine the effects of strain
rate and temperature on the evolution of the microstructure.
Fig. 9(a) shows the microstructure of the specimens
deformed at room temperature under a strain rate of
1.3103 s1. Under these conditions, the microstructure
consists of coarse stable equi-axed grains of Al matrix with
an average diameter of 40 mm. Each grain is composed of
many ne precipitates of the major alloying elements (Zn,
Mg, Al). A nearly continuous network of second phase
surrounding the equi-axed grains can also be seen. Along
the grain boundaries, many lath-like precipitates are found,
and a few precipitate particles larger than about 15 mm

Fig. 9. Optical micrographs taken from specimens deformed at: (a) 258C
with a strain rate of 1.3103 s1; and (b) 3008C with a strain rate of
3.1103 s1.

appear within grains and grain boundaries. These lath and


particular precipitates of MgZn2 are insoluble at the solutetreated temperature.
With deformation under higher strain rate and temperature conditions, a distinct difference in the microstructural
morphology is observed, as shown in Fig. 9(b), which
corresponds to a specimen deformed at a strain rate of
3.1103 sl at 3008C. Clearly, high-rate and high-temperature loading leads to a growth of second phase in order to
reduce the total boundary energy. The large difference in the
amount of second phase may be attributed to the augmentation of the test temperature and impact heating of the
specimen during deformation. If the grain morphology is
examined, it is found that the Al matrix grains remain
roughly equal-axed, but the initial coarse grains have been
uniformly rened. The average grain size decreases dramatically from 110 mm at a strain rate of 1.3103 s1 and a
temperature of 258C, to 40 mm at a strain rate of 3.1103 s1
and a temperature of 3008C. This result suggests the possibility of dynamic recrystallization taking place during
deformation. Fig. 10 shows the changes of average grain
size in terms of strain rate and temperature by using a linear

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

121

Fig. 12. Optical micrograph of the 7075 Al alloy deformed at 1008C and
3.1103 s1, showing arc-type adiabatic shear bands.
Fig. 10. Changes in grain size with strain rate at a true strain of 0.15 for
different temperature conditions.

line intercept method against the obtained optical micrographs.


3.6. Fracture feature observations
Scanning electron micrographs show that fractures
always occurs at an angle of 458 with respect to the
compression axis, and that they are accompanied by adiabatic shear bands. An example of the fracture appearance is
shown in Fig. 11, which corresponds to a specimen
deformed at 2008C and 3.1103 s1. In this case, the
adiabatic shear bands observed in the specimens are narrow
and consist of sharp cracks that lead to the separation of the
specimen. Fig. 12 shows an optical micrograph of the
sheared area for a specimen deformed at 1008C and
3.1103 s1. Examination of the cross-sections of the tested
specimens indicates that adiabatic shear bands form near one
periphery and run diagonally through the depth of the
specimen to the opposite periphery. The adiabatic shear
bands as observed on any cross-section appear as arcs with
their radii towards the specimen center. Hence, adiabatic
shear bands after initiation propagate in two directions, the

Fig. 11. Fracture appearance of the 7075 Al alloy deformed at 2008C with
a strain rate of 3.1103 s1.

faster one along the center of inversion into the depth of the
specimen to the opposite side, the other as a slow, concurrent
lateral growth. The faster propagates at an angle of approximately 458 to the perpendicular plane.
Under dynamic impact loading, an adiabatic shear band
can be dened as a stress localization phenomenon generally
caused by a plastic instability, in turn caused by thermal
softening during adiabatic or quasi-adiabatic deformation.
At strain rates above 103 sl, plastic deformation in the
impacted material does not have enough time to occur
thoroughly. Therefore, the impact energy has to be dissipated rapidly by an elevation of temperature within the
material. This adiabatic temperature rise in the material
can be very high and, furthermore, because of the short
duration of the event, is concentrated within narrow bands at
locations of maximum stress. As a consequence of this
temperature rise, the material within the shear band can
possibly melt or, more usually, its microstructure will be
changed dramatically with the formation of the transformed
white phase. Dynamic impact material failure can thus differ
from the other dynamic fracture modes of void formation
and coalescence. Fig. 13 shows fracture surface features

Fig. 13. Fracture features of specimen tested at 3008C and 3.1103 s1 as


observed by SEM.

122

W.-S. Lee et al. / Journal of Materials Processing Technology 100 (2000) 116122

produced by separation along the adiabatic shear bands. It is


clear that the fracture surface is relatively smooth, with ne
striations running parallel to the shear direction. Also, cracks
and secondary micro-cracks are seen in the fracture surface,
indicating the brittleness of the material.
4. Conclusions
Dynamic impact experiments on 7075 Al alloy have been
conducted to investigate the inuence of the loading rate and
the temperature on the mechanical properties and microstructure variations. The compressive stressstrain response
of this material is found to depend on both the strain rate and
the temperature. By means of the experimentally-determined material parameters, a proposed deformation constitutive equation is used successfully to describe the behavior
of the material under dynamic loading. Fracture analysis
show that adiabatic shear bands play a key role in the
dynamic failure process of 7075 Al alloy. The fractures
exhibit relatively smooth surfaces, which contains many
cracks that are characteristic of brittle fracture. Microstructural observations reveal that grain renement and secondphase growth are induced during high-rate and high-temperature deformation.
Acknowledgements
The authors would like to acknowledge both their department and the National Science Council of the Republic of

China for their nancial support. The grant from the NSC is
numbered 86-2212-E006-015.
References
[1] A.K. Ghosh, G. Gandhi, in: H.J. McQueen et al. (Eds.), Strength of
Metals and Alloys (ICSMA7), Pergamon Press, Oxford, 1985, p.
2065.
[2] M.A. Zaidi, J.A. Wert, in: A.K. Vasudevan, R.D. Doherty (Eds.),
Aluminum Alloys Contemporary Research and Applications,
Academic Press, New York, 1989, p. 137.
[3] U.S. Lindholm, J. Mech. Phys. Solids 12 (1964) 317.
[4] P.S. Follansbee, G.T. Gray, Metall. Trans. 20A (1989) 863.
[5] M. Meyers, Y.J. Chen, F.D.S. Marquis, D.S. Kim, Metall. Mater.
Trans. 26A (1995) 2493.
[6] A.M. Bragov, A.K. Lomunov, Int. J. Impact Eng. 16(2) (1995) 321.
[7] W.S. Lee, S.T. Chiou, Composites: part B 27B (1996) 193.
[8] K. Tanaka, K. Ogawa, T. Nojima, in: J. Shioiri, K. Kawata (Eds.),
High Velocity Deformation of Solids, Springer, Berlin, 1979, p. 98.
[9] K. Ogawa, T. Nojima, J. Mater. Sci. Jpn. 37 (1988) 41.
[10] W.S. Lee, C.F. Lin, J. Mater. Process. Technol. 75 (1998) 127.
[11] C.Y. Chiem, J. Duffy, Mater. Sci. Eng. A 57 (1983) 233.
[12] W.S. Lee, C.F. Lin, Mater. Sci. Eng. A 241 (1998) 48.
[13] R.C. Dorward, K.R. Hause, J. Mater. Eng. Perform. 4(2) (1995) 216.
[14] E. Cerri, E. Evangelista, A. Forellese, H.J. McQueen, Mater. Sci.
Eng. A 197 (1995) 181.
[15] D.E. Albert, G.T. Gray III, Acta Mater. 45(1) (1997) 343.
[16] G.R. Johnson, J.M. Hoegfeldt, U.S. Lindholm, A. Nagy, Trans.
ASME, J. Eng. Mater. Technol. 105 (1983) 48.
[17] A.J. Holzer, Trans. ASME, J. Eng. Mater. Technol. 101 (1979) 231.
[18] B.C. Muddle, Metall. Trans. A 15A (1984) 1089.
[19] L.W. Meyer, in: M. Meyer et al. (Eds.), Shock Wave and High Strain
Rate Phenomena in Materials, Marcel Dekker, New York, 1992.
[20] H. Kobayashi, B. Dodd, Int. J. Impact Eng. 8 (1989) 1.

You might also like