Está en la página 1de 6

th

Published in : Proceedings of the 6 Japan-France Materials Science Seminar, 1999, Poitiers, France.
J. Phys. IV France, 10, PR6 151-156 (2000)

Strain hardening in relation to microstructure in precipitation hardening materials

A. Deschamps1, S. Esmaeili2, W.J. Poole2, M. Militzer2

1
LTPCM / ENSEEG, Domaine Universitaire, BP75, 38402 St Martin d'Hères Cedex, France
2
Dept of Metals and materials, University of British Columbia, Vancouver, BC V6T 1Z4, Canada

Abstract : The influence of microstructure on strain hardening is studied through Kocks-Mecking plots in a number
of systems showing precipitation hardening : Al-Zn-Mg, Al-Mg-Si-Cu, and Fe-Cu. The presence of a supersaturated
solid solution is shown to result in an extremely high work hardening rate, due to dynamic precipitation during the
straining. When precipitation occurs, a drastic change in the work hardening capability is observed, which can be
related to the type of precipitate-dislocations interactions and to the residual solute content. Shearable precipitates do
not seem to influence greatly the work hardening behavior, which is then mostly controlled by the solute content.
Non-shearable precipitates induce a high initial hardening rate. However this high initial value cannot be sustained
to high strains due to extensive dynamic recovery in the solute-depleted matrix. From the analysis of the work
hardening rate, it seems that precipitates remain shearable up to very large sizes and to very overaged states in the
Al-Mg-Si-Cu and the Fe-Cu alloys, which has important consequences on the modeling of the hardening curve of
these alloys.

1. INTRODUCTION

The understanding of the evolution of strain hardening with microstructure is relevant to many practical
problems related to the processing and use of materials. For instance, work hardening rate partially
controls formability, ductility and toughness.
In pure metals, a reasonable understanding has been reached for the description of the influence of
temperature and strain rate on the work hardening rate (WHR), both in the low-temperature range and in
the creep range [1]. However, the influence of solid solution and precipitates is not clear, and the existing
theories [1-3] do not account correctly for a lot of experimental data.
The purpose of this paper is to present some experimental data on two aspects :
-i- work hardening of a supersaturated solid solution, and its relation to dynamic precipitation ;
-ii- influence of precipitation on work hardening, and deducing from it the interactions between
dislocations and precipitates (and especially shearing / by-passing).
The first point was investigated in an Al-Zn-Mg alloy (7xxx series), and the second point in three
different alloy systems, namely Al-Zn-Mg, Al-Mg-Si-Cu, and Fe-Cu.
The Al-Zn-Mg system shows a sequence of precipitation through two metastable phases :
α Õ α + GP zones Õ α + η' Õ α + η (=MgZn2)
The Al-Mg-Si-Cu system also shows a complex precipitation sequence including precursors of the
Mg2Si and Q phases.
Finally, precipitates in the Fe-Cu system are pure copper, with their structure changing during aging :
α Õ α + BCC copper Õ α + transition phases Õ α + FCC copper.

2. MATERIALS AND METHODS

The composition of the three alloys is given in the Table 1 below. The processing of the alloys was as
follows :
- Al-Zn-Mg : solution treatment 1h at 475ºC, water quench. Resulting grain structure was 50%
recrystallized, and the average grain size approximately 50 µm. This will be called the as-quenched state
(AQ). 3 days of natural aging (NA state). 30ºC/h heating ramp up to 160ºC, then various times at 160ºC.

1
- Al-Mg-Si-Cu (AA 6111) : solution treatment 10 min. at 560ºC, water quench. Resulting grain structure
was recrystallized with an average grain size of approximately 30 µm. 6 months of natural aging,
followed by various times at 180ºC.
- Fe-Cu : solution treatment 1h at 850ºC, water quench, resulting grain size approximately 50 µm,
followed by various times at 500ºC.

Table 1 : Composition (in wt%) of the three alloys


Al Zn Mg Si Cu Fe C
Al-Zn-Mg bal. 6.1 2.3 <0.1 <0.1 <0.1
Al-Mg-Si-Cu bal. 0.79 0.6 0.7 0.25
Fe-Cu 0.8 bal. <20 ppm

The alloys were strained in tension at a strain rate of 2.10-3 s-1 for Al-Mg-Si and Fe-Cu and 10-4 s-1
for Al-Zn-Mg. The strain hardening θ=dσ/dε was calculated by fitting the true stress - true strain curves
with a polynomial function, and differentiating this function with respect to ε. The evolution of strain
hardening will be plotted in θ vs. (σ-σy) plots (or Kocks-Mecking plots), where σy is the yield stress.
When the influence of temperature is considered, the Kocks-Mecking plots will be represented in a
normalized manner, i.e. as θ/µ vs. (σ-σy)/µ plots, where µ is the temperature-dependent shear modulus.

3. STRAIN HARDENING OF A SUPERSATURATED SOLID SOLUTION

Before considering strain hardening of a solid solution, it is useful to study the strain hardening rate of
pure aluminum. Aluminum of 99.999% purity (grain size 100 µm) was strained at temperatures between
4.2K and 300K, at a strain rate of 10-4 s-1. The resulting stress-strain curves and related strain hardening
rate plots are shown in figure 1. It is found that the yield stress is very low and almost temperature-
independent, whereas the work hardening rate increases substantially when the temperature is lowered,
increasing the ultimate tensile stress from 50 MPa at room temperature up to 350 MPa at 4.2K. The
evolution of WHR with stress follows exactly the framework built by Kocks, Mecking and Estrin [1,4,5] :
initial WHR of about µ/20, linear decrease of θ with σ−σy, the slope of this straight line increasing with
temperature.

600 0.1
a) b)
500 Al-Zn-Mg
0.08
True stress (MPa)

Pure Al, 4.2K


400 Pure Al, 77K Al-Zn-Mg
Pure Al, 300K 0.06
θII=µ/20
θ/µ

300 Al-Zn-Mg, 4.2K


Al-Zn-Mg, 77K 0.04
200 Pure Al Al-Zn-Mg, 300K
Pure Al
100 0.02

0 0
0 5 10 15 20 25 30 35 0 0.005 0.01
True strain (%) (σ-σy)/µ
Figure 1 : a) stress-strain curves and b) normalized work hardening rate plots for 99.999 pure
aluminum and the as-quenched Al-Zn-Mg alloy strained at 4.2K, 77K and 300K.

The as-quenched Al-Zn-Mg alloy shows a quite different behavior, as shown in Figure 1 : the yield
stress is more temperature- sensitive (120 MPa at room temperature, 180 MPa at 4.2K), but the work
hardening rate is completely independent of temperature, over this temperature range. Moreover, the
initial WHR is very high, of the order of µ/10. Besides, one can notice strong thermomechanical
instabilities at the temperature of 4.2K, which can be attributed to adiabatic shear banding.

2
4. STRAIN HARDENING IN THE PRESENCE OF PRECIPITATES

When precipitation occurs, the work hardening behavior can evolve due to the change of a number of
microstructural features :
- The depletion of the solid solution tends to decrease the ability of the material to store high densities of
dislocations, and thus to decrease the work hardening rate. We have seen in section 3 that in
supersaturated solid solutions the WHR can be vastly different from that of a pure metal.
- Precipitates can be either sheared (usually in underaged conditions), or by-passed by the Orowan
mechanism (usually in the overaged conditions). In the first case, shearing of precipitates does not allow
additional dislocation storage, and the alloy should have a low WHR, with a strong tendency for strain
localization. In the second case, by-passing of precipitates leaves dislocation loops, which induce a high
initial WHR related to the storage and relaxation of these geometrically necessary dislocations. Thus, the
detailed study of the initial stages of WHR should enable the distinction between precipitates shearing
and precipitates by-passing.

4.1. The Al-Zn-Mg system


600 3000 0,12
a) NA state
Work hardening rate (MPa)

500 2500 NA + heating to 160°C


0,1
2h at 160°C
True stress (MPa)

400 2000 50h at 160°C


0,08
700h at 160°C
Al-Zn-Mg

θ/µ
300 Al-Zn-Mg 1500 0,06

NA state θ =µ/20
200 1000 ΙΙ 0,04
NA + heating to 160ºC
2h at 160ºC
100 500 0,02
50h at 160ºC
700h at 160ºC b)
0 0 0
0 2 4 6 8 10 12 14 16 0 50 100 150 200 250 300
True strain (%) Stress-yield stress (MPa)

Figure 2 : a) stress-strain curves and b) normalized work hardening rate plots for the Al-Zn-
Mg alloy in different precipitation conditions, tested at room temperature.

We can see in Figure 2 the evolution of the stress-strain curves along the aging treatment. Peak aging in
terms of yield stress is obtained after 2 hours at 160ºC. Several stages can be observed in the evolution of
the strain hardening curves for these conditions :
- the naturally aged material is characterized by a high initial value and a slow decrease of the WHR,
similarly to the as-quenched material ;
- as the solid solution is depleted by the formation of small (presumably shearable) precipitates, the initial
value of the WHR decreases down to the "normal" value of µ/20, for the peak-aged condition. The rate of
decrease of the WHR is not much affected by this microstructural evolution.
- After peak aging, two combined effects can be noticed : the initial WHR increases significantly, and
continuously with aging time (and thus with precipitate size), however a very fast decrease of this WHR
can be observed.

4.2. The Al-Mg-Si-Cu system

The situation in the 6111 alloy is quite different. From figure 3 we can describe several stages in the
evolution of the WHR behavior as a function of aging time, with respect to the peak aging which occurs
after 6 hours at 180ºC :

3
3000 0.12
400
a) as-quenched b)

Work hardening rate (MPa)


350 2500 6 months NA (T4) 0.1
T4 + 1h at 180°C
300
T4 + 6h at 180°C
True stress (MPa)

2000 0.08
250 T4 + 6 months at 180°C

θ/µ
200 as-quenched 1500 Al-Mg-Si-Cu 0.06
6 months NA (T4)
150 T4 + 1h at 180ºC 1000 0.04
T4 + 6h at 180ºC
100 T4 + 6 months at 180ºC
500 0.02
50 Al-Mg-Si-Cu
0 0 0
0 5 10 15 20 0 50 100 150 200
True strain (%) Stress - yield stress (MPa)
Figure 3 : a) stress-strain curves and b) normalized work hardening rate plots for the Al-
Mg-Si-Cu alloy in different precipitation conditions, strained at room temperature.

- From the as-quenched stated to the naturally aged state (6 months at room temperature, T4 material), not
only is the yield strength increased from 50 MPa to 160 MPa, but the initial WHR is also increased to
very high values of more than 2.5 GPa (=µ/10). The slope of decrease of this WHR is identical between
these two states.
- From the T4 state up to the peak-aging condition, a similar behavior as in the Al-Zn-Mg material can be
observed : decrease of the initial WHR down to the normal value of µ/20, and identical rate of decrease of
the WHR with stress.
- Finally, from the peak-aged condition to a very heavily overaged condition (6 months at 180ºC), the
behavior is very different from the Al-Zn-Mg system : the rate of decrease of the WHR increases, but the
initial work hardening rate is constant, thus giving a very low ductility and ultimate tensile strength.

4.3. The Fe-Cu system

500 2500
a) 0.03
As-quenched
2h at 500°C
Work hardening rate (MPa)

400 2000 0.025


7h at 500°C
True stress (MPa)

40h at 500°C 0.02


300 1500 1000h at 500°C
Fe-Cu
θ/µ

0.015
Fe-Cu
200 1000
As-quenched 0.01
2h at 500ºC
100 7h at 500ºC 500
40h at 500ºC 0.005
1000h at 500ºC
0 0 0
0 5 10 15 20 25 0 50 100 150 200 250
True strain (MPa) Stress-yield stress (MPa)
Figure 4 : a) stress-strain curves and b) normalized work hardening rate plots for the Fe-Cu
alloy in different precipitation conditions, strained at room temperature.

The analysis of the Fe-Cu case is slightly more difficult, due to the existence of the Lüders plateau. The
WHR was calculated after the end of this plateau, and the yield stress considered for the work hardening
plots was the lower yield stress (plateau value). The evolution of the work hardening rate can be
discussed from Figure 4, and in some cases substantial differences can be found with the other systems :
- The as-quenched material has the highest hardening capability (note that in this alloy the initial
hardening rate is only of the order of µ/30,indicating no "abnormal" behavior). When the material is in
underaged condition, the first tendency is towards a strong decrease of the initial WHR, without any

4
change in the rate of decrease of the WHR, consistently with the other systems. However, before the peak
hardening is reached (at 40h at 500ºC), the initial WHR increases again, whereas the slope of the WHR
plot increases. Thus, the peak hardening condition shows a higher hardening capability than the
underaged condition.
- When the peak hardening is passed, the behavior changes again : the slope is identical to the peak-aged
condition, but the initial WHR decreases dramatically, leading as in the Al-Mg-Si-Cu system to the
lowest hardening capability of the whole aging curve.

5. DISCUSSION AND CONCLUSIONS

We have seen through these experimental data that the effect of solid solution and of precipitation can be
very diverse from one system to another. It is thus useful to recall what the existing theories predict.
The effect of solid solution has been extensively modeled [1,2,6]. In the absence of special
dislocation storage sites (such as non-shearable precipitates or very small grains or sub-grains), it is
usually considered that the maximum storage rate for dislocations should result in a work hardening rate
of dτ/dγ=µ/200 on a given slip plane [7], which corresponds to approximately µ/20 for macroscopic
values. The effect of solid solution is then supposed to act by reducing the rate of dynamic recovery.
Secondly the effect of shearable precipitates has received very little attention from a theoretical point
of view [1], probably due to the lack of experimental data separating clearly the effect of the presence of
shearable precipitates and the depletion of solid solution. It is considered that the shearing of precipitates
leads to catastrophic shear banding and very low WHR. These states corresponding generally to very high
yield stresses, the resulting ductility provided by Considiere's criterion is very low (since Considiere's
criterion for necking in tension states that the instability occurs when the work hardening rate equals the
flow stress).
Finally, in overaged states it is often presumed that precipitates are by-passed. The work hardening
behavior is treated either in the framework of Ashby [8] by the storage of geometrically necessary
dislocations, or in the framework of Brown and Stobbs [3] by the build-up of internal stresses
proportional to the strain. Both theories predict very high initial WHR, but do not tell much about how
this WHR decreases as strain progresses.

From the experimental data collected in this study, we can both try to show where these theories
succeed or fail, and propose some conclusions on the precipitates-dislocations interactions in the three
systems we have investigated.

5.1. Influence of solute

As it has been observed before [9], the influence of solute content on work hardening behavior seems to
be mostly a change in the initial WHR, and not so much in the rate of decrease of this WHR, which seems
to be more or less constant with solute content. In a previous paper [10], we have shown that the
anomalously high work hardening rates met in supersaturated Al-Zn-Mg solid solutions, as well as the
temperature-independence of these hardening rates, was related to dynamic precipitation during the
tensile test (at least at 300K and 77K). A similar behavior has been found here in the 6111 alloy,
suggesting that this alloy may undergo dynamic precipitation as well. In particular, dynamic precipitation
seems to be the only possible explanation for the higher work hardening rate met in the T4 condition as
compared to the as-quenched material in this alloy. In order to understand fully the effect of solid
solutions on WHR, it would certainly be useful to study a large range of solid solutions, from dilute
undersaturated to strongly supersaturated ones.

5.2. Influence of shearable precipitates

The present study confirms that the presence of shearable precipitates results in a low WHR. However, it
seems that this low work hardening rate does not particularly result from a strong strain localization, but
5
more from the combination of a low solute content (because of the precipitation process) and no ability
for extra dislocation storage provided by non-shearable precipitates :
- When the supersaturation is still high, the presence of shearable precipitates does not result in a low
WHR. In some cases, the presence of GP zones can even be associated with the highest hardening
capability of the material (as for the 6111 alloy).
- When the solute content decreases, the overall work hardening rate decreases, but only through its initial
value : the rate of decrease is not affected, which seems to indicate that the solute depletion is the
governing factor in this decrease.

5.3. Influence of non-shearable precipitates

The influence of non-shearable precipitates is indeed to increase the initial work hardening rate, as can be
seen in the Al-Zn-Mg system. However, the initially high work hardening rate decreases abruptly, due to
the lack of solute : the fast increase in the dislocation density saturates quickly. Thus the resulting
ductility is surprisingly low. When we consider that in overaged states the yield stress has decreased
significantly, the ductility should increase significantly according to Considiere's criterion. This is not the
case : the ductility stays basically constant from peak aging to overaging.

Finally, important consequences on the dislocation-precipitates interactions can be drawn from the
experimental data collected in this study. Notably, it seems that in both the 6111 alloy and the Fe-Cu
alloy, precipitates stay shearable even in the overaged conditions, as shown by the extremely low WHR
met. This means that standard models which associate peak hardening to the shearing / by-passing
transition should not be applied in this case.
Another feature concerns specifically the Fe-Cu alloy. This alloy experiences first a lowering of the
WHR, followed by an increase around the peak hardening condition and a second decrease in the
overaged condition. This behavior can be related to the precipitation sequence : first, coherent, shearable
BCC copper precipitates, then non-shearable complicated transition phases, and finally partially coherent,
shearable FCC copper precipitates obeying the Kurdjumov-Sachs orientation relationship.

ACKNOWLEDGEMENTS

The authors would like to thank Pechiney, Alcan and Dofasco, Inc. for providing the materials for this
study and for financial support. Pr. Bréchet is also warmly thanked for fruitful discussions.

REFERENCES

[1] Y. Estrin, in Unified constitutive laws for plastic deformation, Academic Press, London (1996).
[2] H. Gleiter, Materials Forum 11, 140 (1988).
[3] L.M. Brown, W.M. Stobbs, Phil. Mag. 23, 1185 (1971).
[4] U.F. Kocks, J. Eng. Mat. Techn. 98, 76 (1976).
[5] H. Mecking, B. Nicklas, N. Zarubova, U.F. Kocks, Acta Metall. 29, 1865 (1986).
[6] D.J. Lloyd , D. Kenny, Metall. Trans. A13, 1445 (1982).
[7] Z.S. Basinski, Phil. Mag. 4, 393 (1959).
[8] M.F. Ashby, Phil. Mag. 21, 399 (1970).
[9] A. Deschamps, Y. Bréchet, C.J. Necker, S. Saimoto, J.D. Embury, Mat. Sci. Eng. A207, 143 (1996).
[10] A. Deschamps, M. Niewczas, F. Bley, Y. Brechet, J.D. Embury, L. Le Sinq, F. Livet, J.P. Simon, Phil.
Mag. A 79 (10), 2485 (1999).

También podría gustarte