You are on page 1of 7

Scripta Materialia, Vol. 34, No. 1. pp.

89-95, 1996
Elsevier Science Ltd
Pergamnon Copyright 0 1995 Acta Metallurgica Inc.
Printed in the USA. All rights reserved
13594462196 $12.00 + .oo

0956-716X(95)00479-3

STRAIN AGEING IN HEAVILY DRAWN


EUTECTOID STEEL WIRES
P. Watt&, J. Van Humbeeckt, E. Aernoud$ and I. Lefeves
7 Department of Metallurgy and Materials Engineering, KU Leuven
$ N.V. Bekaert S.A., Zwevegem
(Received May 10,1995)

Static strain ageing includes the change in mechanical and physical properties of a deformed solid as a function
of time and temperature. When former property changes are concurrent with plastic deformation the effects
are called dynamic strain ageing. It has been widely accepted that the extent of strain ageing relates to the con-
centration and mobility of interstitial or substitutional elements, to the presence and mobility of lattice defects
and to the microstructure [ I].
The interest in strain ageing of severely drawn eutectoid steel wires has revived in industry because of its
impact on procertig and product quality. Strain ageing turns out to be important at different production lev-
els. During storage, e.g., as-drawn wire might undergo room temperature static strain ageing which appears
to deteriorate the ductility. Besides, as-drawn wire is sometimes submitted to several kinds of ageing heat
treatments in order to meet some product specifications, e.g. to enhance the elastic limit and the yield strength
[2]. An understanding of the kinetics of strain ageing and insight in the influence of the chemical composition
on strain ageing phenomena can facilitate the appropriate choice of former heat treatments or allow to develop
wire with a low strain ageing susceptibility.
In the past, strain ageing in severely drawn wires was investigated by means of various techniques ranging
from tensile testing [ 1,3-41 and electrical measurements [6-81 to ion probe microscopy [4,9] and Mdssbauer
spectroscopy [ 10,111. The investigation in this paper is devoted to a quantitative study of the heat effects
induced by ageing heavily drawn wire. The advent of powerful calorimetric techniques opened some new per-
spectives to study static strain ageing [ 121, which are also applicable to tine wires. We studied the kinetics
of static strain ageing by means of dXerential scanning calorimetry (DSC) and thermoelectrical methods
(TEP). It is shown that DSC can be easily adopted to di&minate between the stages during which strain
ageing proceeds. This approach enables an elegant determination of the activation energies of the stages of
strain ageing. In addition, the thermopower of wire was measured in order to disclose the mechanism of strain
ageing which operates below 250°C.

ExDerimental

The wire under investigation was pearlitic steel wire, manufactured by N.V. Bekaert S.A. The starting ma-
terial was Stelmor cooled wire rod. The wire rod was pickled and drawn to semifinished wire in a dry drawing
process. This xmifinished wire is subsequently patented, which consists of an austinitisation treatment at
950°C followed by a quench to 590°C in a lead bath This wire further drawn to a strain of 3.9 in a wet draw-
ing process at industrial drawing speeds. The average reduction per pass was about 20 per cent and the die

89
90 STRAIN AGEING Vol. 34, No. 1

angles were of the order of 10 o After the last die exit it was quenched in ice and further stored under frozen
conditions until the performance of the DSC-experiments.
The DSC-experiments were perhormed on a commercially available calorimeter of TA-systems type DSC-
10. Composite samples were prepared by filling ahnninium pans with wires that were cut from a batch of aa-
received wire. Heat flow signals were recorded relative to a reference sample. The latter was cut from the
batch of wire under investigation and subjected to a heating cycle from room temperature to 600°C in the
DSC-cell. Such sample can be used as a reference since it appeared to be fully chemically inert. The meas-
urements were performed in a protective atmosphere of Argon gas which was led through the DSC-cell.
Emphasis has been put on the determination of the maximum peak temperatures in the spectra, to incorporate
them in a Kissinger analysis. Besides, the thermopower of samples which were artilicially aged in a silicone
oil bath was measured. The samples were isothermally aged at temperatures below the end temperature of
the first peak in the DSC-spectra. The measurements were performed relative to a calibration sample with
absolute thermopower of 7.3 pV/K. During the measurement a stable temperature gradient was applied. The
sample ends were held at a temperature T,=lS “C and T,=25 “C, respectively.

Results

Figure 1 displays the DSC-spectra of the sample under investigation, Irrespective of the heating rate, all spec-
tra show 4 maxima, which can be brought in connection with the different stages of strain ageing which occur
in severely drawn wires. It can be inferred from these spectra that there is a considerable overlapping of the
stages. At the present time, a fitting procedure is being developed in order to deeonvolute the spectra. As such,
we only report on a Kissinger analysis that was performed. The estimation of the activation energies of the
ageing mechanisms is a tirst step towards a characterisation of the stages of strain ageing. In order to deter-
mine the latter, we adopted the hypothesis that the reaction rate dx/dt associated to the ageing mechanisms can
be expressed by:

0.06 -
40 K/tin

0.05

0.04

B
5 0.03
g
E 0.02
3
s

I I I-l-

400 500 600 700 800


temperature [“C]

Figure 1. DSGspectra of eutectoid steel wire drawn to a total drawing strain of 3.9. The heat flow signal is proportional to the adopted
heating rates of 7,15,20,25,30,35 and 4OK/min.
Vol. 34, No. 1 STRAIN AGEING 91

dx
- = A.(1 - x)‘.exp -
Ea (1)
dt R.T1
in which x is the fraction of the reaction which is completed atter a time t and r is the order of the reaction,
which is of no hather importance in what follows, T is the temperature measured in the DSC-cell and R is the
gas constant. From equation 1 one can derive that the maximum peak temperature T,,,,relates to the activation
energy, E. according to:

d[ln O/T;] 4
(2)
d[l/T,] = - y

in which @=dT/dt is the heating rate which was adopted in the DSC-experiments. We found that only the
maximum peak temperatures of the t%t and the second peak fit well in the framework of the Kissinger theory
[ 131,whereas the correlation is too poor for the third and the fourth peak. Figure 2 displays the Kissinger plots
for the first and the second peak in the DSC-spectrum. Fitting to equation 2 yields an activation energy of
68kJ/mole for the first peak, whereas the activation energy for the second peak is significantly higher, namely
81 k&role. In the following paragraph we comment on the mechanisms which give rise to these peaks. We
principally focus on the first peak since the combination of DSC and TEP measurements enables to disclose
the mechanism which is involved. Several authors [3,7] have attempted to explain the mechanism of strain
ageing in the temperature interval covered by the second peak; however, their interpretations are rather de-
batable. Most ofthem suppose that the second peak should be due to static recovery or precipitation, whereas
Tarui et al. [4] 1eRthe possibility that it is caused by spheroidisation of the cementite. Because of lack of con-
vincing microscopic data in literature we only pay attention to the activation energy associated to the second
peak.
Note that the activation energy owing to the first peak almost matches the value for interstitial carbon dif-
fbsion in the Fe matrix, 78kJ/mole [14]. Nevertheless we believe that within the temperature interval covered
by the first peak some other mechanism besides interstitial solute diffusion is involved. This was contirmed

-12.5

-13

0.0016 0.002 0.0024

l/Tm [K]

Figure 2. Kissinger plot ofthe maximum peak temperatures of the first and the second peak in the DSC-experiments represented
in figure 1.
92 STRAIN AGEING Vol. 34, No. 1

I I I , / , I , I
o T=120°C

0 T=140”C

o T=I60”C,

0
3 4 5 6 7 8

In(t)

Figre 3. Johnson-Avrami-Meld plot of the them~opower measurements on wire that was artikially aged between 120°C and 180°C.

by combining the DSC-experiments with measurements of the thermopower at RT. The rise in thermopower
af?.erart&&l ageing was fitted to a Johnson-Avrami-Mehl equation. The Avrami-plots are pictured in figure
3. According to this theory, the fraction of the solid state reaction, f(t), which evolves during isothermal heat
treatment can be described as:

f(f) = 1 - exp(-(W”> (3)

S(t) - 3, .
WeWfW= s _s mwhich S(t) is the thermopower after an artiticial ageing time t, S, is the

thermopower at sa&%ion an!\ S, is the thermopower of as-received wire. The time exponent n is charac-
teristic for the intrinsic nature of the ageing process, it was estimated by means of linear regression. As can
be seen from figure 4 we find that n is temperature independent and ranges between 0.23 and 0.3.

100 120 140 160 180 200

artificial ageing temperature (“C)

Figure 4. Evolution ofthe time eqonent in the Avarmi equation with ageing temperature.
Vol. 34, No. 1 STRALNAGEING 93

Discussion

In order to interpret our measurements we refer to the stages of strain ageing reported in earlier studies which
shed a light on strain ageing in high carbon steel wires.

Stape I of Strah Apeiag

We believe that the mechanism underlying the first peak is the segregation of C atoms to ferrite dislocations.
Alter patenting, the ferrite phase in eutectoid steel is supersaturated with interstitially dissolved carbon. Be-
cause of thermal activation these carbon atoms will migrate to ferrite dislocations to pin them. In literature
this is called stage I of strain ageing [3,5]. This process is completed when the reservoir of interstitial carbon
atoms in ferrite is exhausted. Because of the relatively small density of interstitially dissolved carbon atoms
in the ferrite this will take place before the dislocationsubstructure in the ftite is fully decorated. The density
of carbon atoms in the ferrite phase of patented and drawn eutectoid wire was measured by Tarui et al. [4]
using field ion microscopy (FIM). Tarui et al. observed a signiticant increase in the carbon content in the fer-
rite atler drawing of patented wire. Some Mossbauer studies [ 10,l l] allude to the existence of a deformation
induced decomposition of the cementite phase. In these studies it appears that the integrated areas under the
cementite sextet in the Mossbauer spectra of cold deformed samples decrease with increasing drawing strain.
This observation agrees with the FIM experiments of Tarui et al. It gives evidence for a loss of carbon atoms
ti-omthe cementite phase and suggests a rise in the carbon concentration in the ferrite after plastic deformation.
This issues from the fact that the equilibrium carbon concentration in the ferrite depends on the dislocation
density [151. During plastic deformation ferrite dislocations might drag carbon atoms from the cementite to
the ftite on ther trajectory on the slip systems and across the cementite [ 161. It is not clear whether the car-
bon enrkhmenll: in the ferrite is accompanied with the formation of Hagg carbides, Fe,&, in the cementite
which are known to form below 230°C [17].
Stage I of strain ageing was attributed by Kemp [3] and Yamada [5] to the migration of these interstitially
dissolved carbon atoms to ferrite dislocations. The carbon concentration in the ferrite is not sufEcient to lock
all ferrite dislocations despite its increase after plastic deformation. This led Kemp and Yamada to conclude
that some other mechanism than interstitial solute diffusion must operate. We share this opinion since inter-
stitial C difhrsionin pearlitic ferrite is unlikely to yield the extremely high heat fluxes we encountered for the
first peak in the DSC-experiments.

Stage II of Strain AQeing

Arguments have been given [3,5] that stage I is followed by a stage which is characterised by a partial de-
composition of the cementite phase, which is temperature induced. The elucidation of some ageing features
reported by Kemp and Yamadarequired the assumptionthat during stage II a transfer of carbon atoms can take
place ti the cementite towards the ftite phase. Those carbon atoms are furnished by the cementite to main-
tainthecarbonsequilibrium concentrationin the ferrite phase. The decomposition of the cementite is accom-
panied with the building up of carbon atoms near dislocations and the depletion of carbon in the ferrite inter-
stitial lattice sites. When the concentration of carbon in the ferrite matrix in contact with cementite is drained
below the equilibrium level further decomposition of the carbon will take place. We believe that in the tem-
perature interval covered by the first peak stage I interferes with stage II. The thermopower of wire artificially
aged at the corresponding temperatures was measured in order to justify this statement. The time exponent
in the Avrami equation was estimated to be of the order of 113. This implies that the dominant mechanism of
strain ageing below 250°C differs from matrix diffusion of C interstitials which is character&d by a P
dependency [18,191. Time exponents of the order of l/2 are expected when carbon interstitials are captured
in cylindricalzones around dislocations,which was used as the basis for the models of Cottrell[20] and Harp-
er [2 11.According to the model of Lenient and Cohen [22] a time exponent of l/3 is in favour of a decom-
94 ST&UN AGEING Vol. 34, No. 1

position mechanism of one phase. In their model to describe the kinetics of such a decomposition, they as-
sumed a planar flow of interstitials from one phase to another. This is applicable to eutectoid pearl&e where
this flow takes place ti-om the cementite perpendicular towards the ferrite-cement&e interface. The assumption
of Lement and Cohen is in agreement with the mechanism of cementite decomposition, proposed herealter.
The mechanisms of strain ageing can be inferred from the energy diagram pictured in figure 5. During
wire drawing plastic deformation induces the massive production of dislocations. The energy diagram illu-
strated in figure 5 is a schematic representation of the energy configuration of carbon atoms in interstitial sites
in the ferrite (left) and in the cementite (right) and in the core of ferrite dislocations after plastic deformation.
The presented con@uration is favourable to the transfer of carbon atoms to the ferrite dislocations. It allows
that during stage I carbon atoms leave their interstitial sites to lock ferrite dislocations. From microstructural
point of view stage II is to some extent akin to stage I. Both result in the formation of Cottrell atmospheres
around ferrite dislocations. In contrast to stage I, during stage II carbon atoms are supplied by the decompos-
ing cementite as illmated in the energy diagram on the right. The atom movement across the ferrite-cemen-
tite interface results in a net depletion of C in the ferrite. As such, there will be a net flow of carbon atoms to
ftite dislocations to occupy sites with a specific energy. At equilibrium, the net flow of carbon atoms ceases
as ill~ated in figure 5. In this state, carbon atoms in the cementite and around the ferrite dislocations occupy
sites with a similar energy.
Using the energy diagram in figure 5 we can give a physical interpretation to the activation energy of the
first peak which we estimated in this study. The energy diagram on the right represents a single pinning pro-
cess which is inherent to the cementite decomposition mechanism. Cementite decomposition, however, is

f.
maintained by multipie dislocation pinning whereby each pinning centre has its own activation energy Some
scatter on former activation energy is suspected, caused by the heterogeneity on the heat of solution of cemen-
titi along arrays of pinning centres and by the interaction between dislocations and interstitial solutions. This
interaction is thought to depend on the presence of alloying elements [ 1,10,23]. Hence, the activation energy
associated to the hrst peak we e&mated from our calorimetric data is an average over all pinning events which

energy

v
>

inter&
site

I, cementite
ferrite
dislocation I
ferritedislocation
J&Gtion position r

Figure 5. Scematic representation of an energy diagram which explains the transfer of carbon atoms 6om left a intemtitial site to a
dislocation during stage I of strain ageinp, right the tram&r from cement&e.to ferrite dislocations during the decomposition of cementite
during stage II of strain ageing. E b isthebindingenergyofacarbonatomtoFeinthecementite, E~isthebiienergyb&wen
a ferrite dislocation and a C atom.
Vol. 34, No. 1 STRAIN AGEING 95

are comprised in the decomposition of the cementite and stage I of strain ageing. Consequently, our approach
offers some interesting perspectives to study the effect of alloying elements on the decomposition of cementite
in steel. Extensive DSC work is in progress in order to trace which alloying elements promote or inhibit ce-
mentite decomposition.

Conclusion

The DSC-technique turns out to be a valuable tool to assess strain ageing in severely cold drawn steel wire.
It offers several advantages with respect to the classical way to assess strain ageing by means of electrical
transport measurements or mechanical tests. From a DSC-spectrum one obtains insight in the stages during
which strain ageing proceeds. Unlike former classical approaches one can get evidence for the overlapping
of the stages. DSC-data of severely drawn wires can be easily interpreted in the hamework of a Kissinger
analysis to yield the activation enw of these stages. This can contribute to a better understanding of the me-
chanisms which are involved. As such, the DSC-technique can provide insight to optimise the chemical com-
position of steel wire rod in order to reduce the susceptibility of wire to strain ageing. By combining DSC-data
and thermopow6r data we disclosed the me&an&m of strain ageing below 250 “C as the migration of C atoms
which are interstitially dissolved in the ferrite to dislocations. When the reservoir of carbon atoms in the ferrite
is exhausted, partial decomposition of the cementite will take place to maintain the locking of ferrite dislo-
cations.

AcknowledPments

The authors would like to thank J. Gil Sevillano at C.E.I.T. San Sebastian for the stimulating discussions about
the subject. J. Van Humbeeck acknowledges the NFWO (National Foundation for Scientific Research).

References

1. J.D. Baird, Metall. Rev., 16,1, (1971)


2. W. Van Raemdonck, I. L.efever,N.V. Bekaert LA, Zwevegem, private communication
3. LP. Kemp, G. Polhud, AN. Bran&y, Mat. SC. & Techo., 6,381, (1990)
4. T. Tarui, T. Takeha& S. Ohashi and R Uemori, Wire Industry, September1994,25
5. Y. Yamada, Trans. ISIJ, 16,417, (1976)
6. AH. CottreIl and AT. Churchman, Jourt~ ofthe Iron and Steel Inst., 271, (1949)
7. H. Abe and T. Suzuki, Tram. ISIJ, 20,690, (1980)
8. J. Campbell rrndH. Coorad, Scripta Met_,31,69, (1994)
9. AR Waugh and S. Paetke and D.V. Edmonds, Metallography, 14,237, (1981)
10. V.N. Gridnev, V.G. Gawilyuk, Phys. Metals, 4,3,531, (1982)
11. C. Dauwe, RU Gent, private communication.
12. M. Van Rooyen and E.J. Mittemijer, Scripta Met.,l6,1255, (1982)
13. H. E. Kissinger, Joum. ofRes.oftheNat. BureauofStaudards,57,4.2712,(1956)
14 R.J. Borg and G.J. Dienes in ‘Introductionto Solid State Diffusion’,Academic Press Inc., (1988)
15 in ‘Steel: A handbook for Materials Research and Engineering Volume 1: Fundamentals, ed. by Die Verein Deutscher
Eisenhtlttenleute,Springer Verlag, (1992), pp 33-35
16. J. Gil Sevillano, C.E.I.T., San Sebastirin,private commuuication.
17 H.J. Kleemolaaud E.A Kuusisto, ScandinavianJoumaI ofMetallurgy, 5,151, (1976)
18. H.B. Aaron and G.R. Kotler, MetaIl. Tram, A, 2,393, (1971)
19. C. Zener, J. Appl. Phys.,21,5, (1950)
20. AH. Cottrell and B.A. Bilby, Proc. Phys. Sot., 62,49, (1949)
21. S. Harper, Phys. Rev, 83,709, (1951)
22. B.S. Lemeut and M. Cohen, Acta Metall., 4,469, (1956)
23. S. JaiswaI and I.D. McIvor, Ironmaking and Steelmaking.16,1,49, (1989)
25. w. wepner, Acta Metall., 5,703, (1957)

You might also like