Anda di halaman 1dari 8

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Materials Science 11 (2015) 24 31

5th International Biennial Conference on Ultrafine Grained and Nanostructured Materials,


UFGNSM15

The Effect of Strain Rate on Ultra-Fine Grained Structure of Cold


Drawn 304L Stainless Steel Wires
P. Parniana, M. H. Parsaa,b,c,*
a
School of Metallurgy and Materials Engineering, College of Engineering, University of Tehran, P.O. Box 11155-4563, Tehran, Iran
b
Center of Excellence for High Performance Materials, School of Metallurgy and Materials Engineering, University of Tehran, Tehran, Iran
c
Advanced Metal Forming and Thermomechanical Processing Laboratory, School of Metallurgy and Materials Engineering, University of
Tehran, Tehran, Iran.

Abstract

The effect of strain rate on nano structured and ultra-fine grained microstructure in austenitic stainless steel AISI 304L was
investigated during cold wire drawing process. This process was used to produce strain induced martensite transformation and
then samples were annealed to create reversion in examined material. Samples were subjected to two different strain rates of 0.41
and 1.032 s-1 with relatively heavy imposed cold drawing of 1.13 strain. Best condition of subsequent annealing were obtained
from experiments which eventuated to ultra-fine grained austenitic structure with different sizes. Metallographic examination, X-
ray diffraction and scanning electron microscopy have been used for microstructural characterizations. Finally a very fine
structures were produced with average grain sizes of 441 nm and 415 nm by applying strain rates of 1.032 s-1 and for 0.41 s-1
respectively.
2015
2015TheTheAuthors.
Authors. Published
Published by Elsevier
by Elsevier Ltd. is an open access article under the CC BY-NC-ND license
Ltd. This
Peer-review under responsibility of the organizing committee of UFGNSM15.
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of UFGNSM15
Keywords: Austenitic stainless steel; wire drawing; strain rate; ultrafine grained material; martensitic transformation.

* Corresponding author. Tel.: +9821-61114069; fax: +982188006076.


E-mail address: pouyan.parnian@ut.ac.ir (P. Parnian), mhparsa@ut.ac.ir (M. H. Parsa)

2211-8128 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review under responsibility of the organizing committee of UFGNSM15
doi:10.1016/j.mspro.2015.11.031
P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31 25

1. Introduction

Austenitic stainless steels are extensively used in different applications requiring good corrosion resistance and
formability. For some application, it is necessary that used material show high strength for weight reduction
propose. Therefore, development of high strength stainless steels characterized by a combination of high strength
and ductility continues to be a major field of study, Lo et al. (2009). Refinement of the grain size to improve
strength properties of engineering materials is one of the established methods, Hedstorm (2007). These steels do not
undergo a considerable grain refinement by recrystallization process and the achievable level of grain refining is
absolutely limited (~ 5-10 m), Lai et al. (2012). To achieve a combination of high strength and desired ductility,
advanced thermodynamic process based on the reversion of strain induced martensite to the austenitic stainless
steels should be carried out, Murata et al.. (1993) and Tomimura et al. (1991). At the primary stage, deformation at
sufficiently low temperatures causes strain induced martensite by transformation of austenite () to martensite. At
the next stage, carrying out of controlled annealing consequences, the reversion of martensite to austenite which
leads to refined austenite. It was argued by Tomimura, Tomimura and Tanimoto et al. (1991), that a nearly complete
martensite structure is essential for achievement of ultra-fine grained (UFG) structure in subsequent reversion
annealing. Two types of martensite phases can be formed during deformation: the HCP -martensite and BCT -
martensite, Maxwell et al. (1974), Olson and Cohen (1981). The mechanisms of martensitic transformation have
been studied extensively and it has been established that -martensite nucleates at the intersections of shear bands
that consist of -martensite, stacking fault bundles and mechanical twins, Lo et al. (2009). The amount of -
martensite decreases with increasing deformation rate due to the adiabatic heating that suppresses the strain-induced
transformation which was revealed by Talonen et al. (2005). The amount of -martensite increases with increasing
plastic strain at the expense of austenite and -martensite phases by the following mechanisms: and ,
Choi and Jin (1997). In the annealing stage, the heavily deformed martensite transforms back to fine-grained
austenite either through a martensitic shear or diffusional reversion mechanism, Smith and West (1971).
Due to the important application of high strength wire in various fields, aim of research was set to evaluate the
effect of various parameters on the deformation behavior of 304L wire during cold wire drawing process and
subsequent reversion annealing. In the following the effect of strain rate and subsequent annealing in formation of
ultrafine grain (UFG) austenite will be present.

2. Experimental procedure

Fig. 1 pictorially summarizes the applied thermo-mechanical processing to obtain UFG austenitic structure
through controlled reversion of strain induced martensie (SIM) in metastable 304L austenitic stainless steel (ASS).

Fig. 1. Schematic of advanced thermomechanical process to obtain Nano/ultrafine grained


austenitic stainless steel.
26 P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31

Commercial austenitic stainless steel grade AISI 304L with chemical composition given in Table 1 was used for
investigations.

Table 1. Chemical composition of the present stainless steel

Fe C Si Mn P S Cr Ni Mo Al W
Balanced 0.0228 0.288 1.76 0.0375 0.0282 18.38 8.37 0.401 0.0131 0.433
Co Cu Nb Ti V Pb Sn Ca Ta B
0.0956 0.0956 0.0164 0.0015 0.0784 0.00089 0.0257 0.0006 0.0022 0.00029

First 1.2 mm diameter annealed wires at 880 for 15 min were cold drawn to 0.91, 0.68, 0.49 mm diameters
without any intermediate annealing in 14 pass. These reductions are correspond to the true strains of 0.56, 1.13, and
1.7 respectively. Cold drawing was carried out at two different average strain rates of 0.41 and 1.032 s-1. The
drawing process was performed using an experimental laboratory drawing bench, characterized by a maximum force
of 9.73 KN. The designed dies had semi angle of 14 degree, bearing length of 0.5 mm and soap powder lubrication
was applied during drawing process.
X-ray diffraction (XRD) analysis was performed on the wires using a BRUKER D8-Advanced diffractometer
with Cu K radiation to examine the microstructure.
Standard metallographic techniques were employed for the preparation of samples for optical microscopy (OM)
and scanning electron microscopy (SEM). To characterize the microstructural details, electrolytic etching was
carried out using 60% HNO3 solution at 2 V by Shirdel et al. (2014) to reveal austenite grains.
For evaluation of mechanical properties, tension tests were performed at room temperature using a Tinius Olsen
universal materials testing machine operating under a strain rate of 10-3 s-1. The tensile test specimen was in
accordance with the ISO 6892-1 standard.

3. Results and discussion

Figure 2 shows the optical initial microstructure which exhibits equiaxed grains with average intercept length of 8.9
m or ASTM grain size of 9.5 and results of X-ray diffraction pattern that show complete austenitic grains.

Fig. 2. X-ray diffraction pattern of as-received sample along with the corresponding optical micrograph.
P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31 27

The austenite stability parameter of Md (30/50) was calculated using equation (1) by Angel (1954):

Md(30/50) (C)=413- [462 (%C+%N)+9. 2 (%Si)+8. 1 (%Mn) +13. 7 (%Cr)+9. 5 (%Ni)+18.5 (%Mo)] (1)

For the used material, the Md (30/50) temperature equal to 46.8 oC which is sufficiently above room temperature.
Calculated Ms temperature according to Eichelmann and Hull equation (equation (2)), Lo et al. (2009) is equal to -
81.6 oC, which is well below the room temperature.

Ms=1305- [1667 (%C+%N)+28 (%Si)+33 (%Mn)+42 (%Cr)+61 (%Ni)] (2)

Therefore, the studied material probably fulfills the basic pre-condition for austenitic metastability and formation
of SIM during cold drawing.
The volume fractions of -martensite, derived by the direct comparison method, Lo et al. (2009) as a function of
drawing strain at room temperature and two different strain rates of 0.41 and 1.032 s-1 are depicted in fig. 3. As
shown, the austenite phase () has not been completely transformed into -martensite phase by applying 1.13 strain.
Since achievement of nearly complete (~>90 vol %) martensitic structure is crucial in this stage, imposed strain was
increased to approximately 1.7. Therefore, applied strain of 1.7 at strain rate of ~ 0.41 s-1 is necessary to achieve
required amount of -martensite (more than 95%) for subsequent reversion annealing.

Fig. 3. Volume fraction of -martensite as a function of drawing strain in two different strain rates.
Fig 3. Volume fraction of -martensite as a function of drawing strain in two different straindc

A key factor to specify the dominant mechanism for the formation of SIM is stacking fault energy (SFE) which
can be estimated using equation (3), Angel (1954). The formation of SIM has been attributed to shear bands, which
are planar defects related to the overlapping of stacking faults on {111} plane in . It has been reported by Choi and
Jin (1997) that the lower stacking fault energies than 18 mJ.m-2 will lead to the transformation sequence of
, Angel (1954), Olson and Cohen (1975). The previous investigations revealed that the formation of twins and -
martensite during plastic deformation strongly depends on SFE and deformation temperature; the formation of the
mechanical twins increases with increasing temperature and SFE and while the formation of -martensite is
suppressed. The stacking fault energy (SFE) of the investigated material has been estimated as 21.11 mJ.m -2 using
the equation (3), Angel (1954):

SFE (mJ.m2 ) 53  6.2( Ni)  0.7(Cr )  3.2(Mn)  9.3(Mo)(Wt %) (3)


28 P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31

The estimated amount of SFE is higher than 18 mJ.m-2 and therefore, formation of -martensite is unexpected.
The change of deformation mechanism to mechanical twinning rather than -martensite formation would be also
attributed to the heat generation during high speed drawing. Considerable heat would be generated due to the
friction between the dies and the wire. This might have an effect equivalent to SFE increment. The absence of -
martensite in steel wires indicates that deformation mode is the sequence of mechanical twining ( ) -
martensite , Fujita and Katayama (1992) rather than the sequence of -martensite -martensite ,Choi and Jin
(1997), when wire drawing is adapted as the deformation method. As can be seen in Figure 3, by decreasing strain
rate, the amount of strain induced -martensite is increased, possibly due to the adiabatic heating phenomenon
Talonen et al. (2005), Litchenfeld et al. (2006). To evaluate this speculation, the XRD analysis was performed on
cold drawn sample to equivalent strain of 0.56 at two different strain rates of 0.41, 1.032 s-1 Fig. 4. A qualitative
inspection of this figure clearly demonstrates the absence of -martensite peaks.

Fig. 4. X-ray diffraction pattern of the cold drawn AISI 304L stainless steel at two different
strain rates.

The exposure of martensitic structure at high temperature leads to the formation of reverted austenite structure. In
this regard, the amount of reverted austenite directly relates to the soaking time and temperature. By increasing
annealing time or temperature the amount of reverted austenite will increase and the effect of annealing temperature
is more significant than that of annealing time, Shirdel et al. (2015). Accordingly, the growth rate was found to
dramatically increase at temperatures higher than 800 oC. Based on the reported results, for determination of nearly
full reverted austenitic microstructures, the specimen annealed at 650 oC for 240 min was selected as a best
candidate for further analysis.
The SEM image taken from the microstructure of mentioned samples is shown in Figure 5. Red arrows indicate
the nano sized austenite grains. Also, corresponding grain size distribution histogram has been plotted to obtain a
better quantitative view from the resultant microstructure. This interesting result shows that the thermomechanical
process of SIM reversion can be regarded as an important and applicable process for achievement of N/UFG
structure in AISI 304L ASS.
P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31 29

Fig. 5. SEM image taken from the microstructure of sample annealed at 650 oC for 240 min, and the corresponding grain
size histogram. (a), (b). SEM images from 0.41 and 1.032 s-1 strain rate. (c), (d). grain size histogram from 0.41 and 1.032
s-1 respectively.

As can be seen in figure. 5, decrease of strain rate caused increase in strain induced -martensite content and
accordingly a decrease in grain size after annealing. Also distribution of grain sizes in higher strain rate show more
pronounced bimodal grain size than lower strain rate, which may be due to adiabatic heating happened in higher
strain rate, Talonen et al. (2005), Litchenfeld et al. (2006). Grain refining is commonly known to raise the strength
and the hardness of polycrystalline materials, Murata et al. (1993), Olson and Cohen (1981). Table 2 summarizes the
comparative results of mechanical evaluations of studied material. Obviously, a significant improvement of the
mechanical properties was obtained; the yield strength and tensile strength increased to 840 and 1123 MPa from the
initial values of 273 and 543MPa, respectively, meanwhile the ductility is not deteriorated and reached an acceptable
value of 38%. Also, the heavily cold drawn sample is characterized by a severe increase of strength with consequent
lowering of percent elongation. The yield strength has been significantly increased to 1931 MPa, which is more than
7 times higher than that of the as received sample (273 MPa). Although in both strain rate a good composition of
large and small grains, caused a good ductility.
30 P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31

Table 2. The Phase percentage, grain size, and mechanical properties of the specimens before and after ultra-fine grain annealing.

Processing Elongation to
Structure MPa MPa Hardness
Condition fracture

As-received ~ 100 %
J ~10
Pm 273 543 46 158

Cold drawn with


0.41 s-1+annealed ~ 98 %
J ~ 415nm 840 1123 38 490
at 600C

Cold drawn with


~ 98 % - 1931 1950 8 680
0.41 s-1

4. Summary and conclusions

The results of the study on martensite reversion treatment of Type 304L ASS allow a detailed characterization of
the ultrafine grained matrix with average grain size of 415 and 440 nm for 0.41 and 1.032 s-1 strain rates
respectively. It was shown that the adiabatic heating produced at higher strain rate, caused lower strain induced -
martensite and therefore increasing grain size caused. Drawing to 1.13 strain with 1.032 s-1 strain rate could not
lead to production of required amount of martensite (~>90%). Thus, a reduction in strain rate to 0.41 s-1 or an
increase of strain to 1.7 showed to be sufficient for successful subsequent annealing to produce ultrafine grained
microstructure. The data reveal a considerable effect of grain refinement on the mechanical properties of 304L ASS.
In a way that a high strength of 1123 MPa was accompanied by high ductility of ~ 38%.

References

Angel, T., 1954. Formation of martensite in austenitic stainless steels-effects of deformation, temperature, and composition. Journal of the iron
and steel institute 177, 165.
Choi, J. Y., Jin, W., 1997. Strain induced martensite formation and its effect on strain hardening behavior in the cold drawn 304 austenitic
stainless steels. Scripta Materialia 36, 99-104.
Fujita, H., Katayama, T., 1992. In-situ Observation of Strain-Induced and Martensitic Transformations in Fe-Cr-Ni
Alloys. Materials Transactions JIM 33, 243-252.
Hedstrm, P., 2007. Deformation and Martensitic Phase Transformation in Stainless Steels.
Lai, J.K.L., Shek, C.H., Lo, K.H., 2012. Stainless steels: An introduction and their recent developments, Bentham Science Publishers.
Lichtenfeld, J.A., Van Tyne, C.J., Mataya, M.C., 2006. Effect of strain rate on stress-strain behavior of alloy 309 and 304L austenitic stainless
steel. Metallurgical and Materials Transactions A 37, 147-161.
Lo, K.H., Shek, C.H., 2009. Recent developments in stainless steels. Materials Science and Engineering: R 65, 39-104.
Maxwell, P.C., Goldberg, A., Shyne, J.C., 1974. Stress-assisted and strain-induced martensites in Fe-Ni-C alloys. Metallurgical Transactions 5,
1305 1318.
Murata, Y., Ohashi, S., Uematsu, Y., 1993. Recent Trends in the Production and Use of High Strength Stainless Steels. ISIJ international 33, 711-
720.
Olson, G.B., Cohen, M., 1975. Kinetics of strain-induced martensitic nucleation. Metallurgical transactions A 6, 791-795.
Olson, G.B., Cohen, M., 1981. A perspective on martensitic nucleation. Annual Review of Materials Science 11, 1-32.
Shirdel, M., Mirzadeh, H., Parsa, M.H., 2014. Microstructural evolution during normal/abnormal grain growth in austenitic stainless steel.
Metallurgical and Materials Transactions A 45, 5185-5193.
Shirdel, M., Mirzadeh, H., Parsa, M.H., 2015. Enhanced Mechanical Properties of Microalloyed Austenitic Stainless Steel Produced by
Martensite Treatment. Advanced Engineering Materials.
Smith, H., West, D.R.F., 1973. The reversion of martensite to austenite in certain stainless steels. Journal of Materials Science 8, 1413-1420.
Talonen, J., Hnninen, H., Nenonen, P., Pape, G., 2005. Effect of strain rate on the strain-induced -martensite transformation and
mechanical properties of austenitic stainless steels. Metallurgical and Materials Transactions A 36, 421-432.
Tomimura, K., Takaki, S., Tanimoto, S., Tokunaga, Y., 1991. Optimal Chemical Composition in Fe-Cr-Ni Alloys for Ultra Grain Refining by
Reversion from Deformation Induced Martensite. ISIJ International 31, 721-727.
P. Parnian and M.H. Parsa / Procedia Materials Science 11 (2015) 24 31 31

Tomimura, K., Takaki, S., Tokunaga, Y., 1991. Reversion Mechanism from Deformation Induced Martensite to Austenite in Metastable
Austenitic Stainless Steels. ISIJ international 31, 1431-1437.

Anda mungkin juga menyukai