Sunteți pe pagina 1din 4

Available online at www.sciencedirect.

com

Scripta Materialia 63 (2010) 1057–1060


www.elsevier.com/locate/scriptamat

Probing deformation processes in near-defect free volume in high


strength–high ductility nanograined/ultrafine-grained (NG/UFG)
metastable austenitic stainless steels
R.D.K. Misra,a,* Z. Zhang,a Z. Jia,a M.C. Somanib and L.P. Karjalainenb
a
Center for Structural and Functional Materials, University of Louisiana at Lafayette, P.O. Box 44130, Lafayette, LA 70504, USA
b
Department of Mechanical Engineering, The University of Oulu, P.O. Box 4200, 90014 Oulu, Finland
Received 24 July 2010; accepted 29 July 2010
Available online 2 August 2010

The deformation behavior of a nanograined/ultrafine-grained (NG/UFG) austenitic stainless steel characterized by high
strength–high ductility combination is investigated via nanoindentation and electron microscopy, and the behavior compared with
its coarse-grained (CG) counterpart. In NG/UFG steel, mechanical twinning was an active deformation mechanism, while in CG
steel, nucleation of strain-induced martensite at the shear bands occurred. The differences in deformation mechanisms of NG/UFG
and CG steels are reflected in the discrete burst in the force–displacement plots and attributed to austenite stability.
Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Metastable austenitic stainless steel; Ultrafine/nanograined; Deformation; Structure

There is currently a strong need to pioneer a new In austenitic stainless steels, the mechanical stability
frontier of metallic materials with high strength–high of the austenite phase and strain-induced transforma-
ductility combination for lightweight constructions, tion to martensite govern the ductility. The underlying
including stainless steels. In this regard, there is a signifi- reason is that the transformation of austenite to mar-
cant interest to understand the deformation mechanisms tensite increases the strain-hardening rate and delays
in nanograined/ultrafine-grained (NG/UFG) materials the onset of localized necking. However, it is known that
in relation to the coarse-grained (CG) counterparts. the grain size of austenite affects its stability [5–7] and,
Recently, we have developed a novel processing route of for instance, in high-Mn austenitic steels the grain size
producing NG/UFG structure in metastable austenitic variation affects the types of deformation mechanism
stainless steels involving controlled phase reversion [5]. In order to better understand the factors controlling
annealing of the cold deformed austenite [1–4]. In this ap- ductility in NG/UFG materials, it is intriguing to exam-
proach, in certain commercial steel grades cold rolling ine the deformation behavior using a sensitive nano/
(thickness reduction 45–77%) of austenite at room tem- micromechanical technique such as nanoindentation,
perature leads to predominantly dislocation-cell-type where we can probe a small volume of the material that
martensite [2]. Upon annealing, the severely deformed can be presumed to be initially defect-free.
strain-induced martensite reverts to austenite either The starting material was a commercial Type 301LN
through a martensitic shear or diffusional reversion mech- austenitic stainless steel of 1.5 mm thickness and having
anism, depending on the chemistry of the steel [1,3]. NG/ nominal composition (in wt.%) of Fe–0.017C–0.52Si–
UFG stainless steel of Type 301LN was characterized by 1.3Mn–17.3Cr–6.5 Ni–0.15Mo–0.15N. The strips were
high yield strength and elongation of 900–1000 MPa and cold rolled in a laboratory rolling mill to 62% thickness
30–40%, respectively [1–4], which greatly exceeds the reduction and subsequently annealed at 800 °C for 10 s
yield strength of 350 MPa and elongation of 40% for in a Gleeble 1500 thermo-mechanical simulator. The
the annealed CG counterpart. annealing experiments were carried out on strips of
120 mm  10 mm (thickness 0.34–0.6 mm). Given that
the nanoindents were to be subsequently examined by
* Corresponding author. Tel.: +1 337 482 6430; fax: +1 337 482 transmission electron microscopy (TEM) for microstruc-
1220; e-mail: dmisra@louisiana.edu tural evolution, the following procedure was adopted.

1359-6462/$ - see front matter Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.scriptamat.2010.07.041
1058 R. D. K. Misra et al. / Scripta Materialia 63 (2010) 1057–1060

First, 3 mm disks were punched from NG/UFG and CG 4 pffiffiffi


P ¼ E Rh1:5 ð1Þ
steels. To ensure that the nanoindents are distributed 3
along the thin area of the disk for examination by where P is the applied load, E* is the effective indenta-
TEM, a modification of twin-jet electropolishing and a tion modulus, R is the radius of the Berkovich indenter
new design of sample mounting for nanoindentation were tip and h is the corresponding indentation depth. The
developed. The disks were partially jet electropolished in effective indentation modulus is also related to the
a refrigerated electrolyte of 10% perchloric acid in acetic Young’s modulus, E, and Poisson’s ratios, m, of the sam-
acid at 25 V for 30 s to obtain a shining surface in the ple and the indenter via [10]:
center part of the 3 mm disk. These partially electropo-  1
lished disks were horizontally placed inside 4 mm diam-  1  m2s 1  m2i
E ¼ þ ð2Þ
eter pits that were drilled on the polished basal surface of Es Ei
an aluminum block of diameter 25 mm using invisible
tape and padding with solder drops, for nanoindentation The subscripts i and s refer to the indenter and sample,
experiments. Nanoindentation experiments were carried respectively.
out in load-controlled mode with a constant loading rate On utilizing the isotropic elastic constant for stainless
of 2 lN s–1. The maximum load was set to 0.5 mN using steel (Es = 193 GPa and ms = 0.25 [8]) and Ei = 1141 GPa
an MTS Nanoindenter system with a Berkovich three- and mi = 0.07 for the diamond indenter tip [10], a value
sided pyramidal diamond indenter with a nominal angle of effective indentation modulus of E* = 185 GPa is
of 65.3° and indenter diameter of 20 nm. This peak load obtained. Hertzian analysis of the data from the elastic
ensured that the displacement was confined within the portion of CG and NG/UFG in Figure 2 gives similar
NG/UFG. An array of indents of a 12  12 matrix was values for E*, of 171 and 162 GPa, respectively. The theo-
defined with an indent gap of 10 lm. After the indenta- retical Hertzian elastic behavior is shown with the broken-
tion experiments, the disk were removed from the mount line in Figure 2 and is consistent with the experimental
and final electropolishing was carried out only from the plot. It also implies that the region before the first discon-
side opposite to the indented surface. Using this proce- tinuity or pop-in is elastic and after is plastic.
dure, the area surrounding the indents, which is present The maximum shear stress at the first pop-in was
along the final jet-polished hole, was electron transparent determined using smax = 0.31p0, where [10]
to study the deformation behavior using TEM. 6PE2 1=3
Light and transmission electron micrographs illustrat- p0 ¼ ð Þ ð3Þ
p3 R2
ing the CG and NG/UFG structures are presented in
Figure 1. The average grain size of starting CG steel in accordance with the standard Hertzian analysis. The
was 52 ± 5 lm, while the grain structure of NG/UFG smax was determined to be 15 GPa for NG/UFG and
steel (cold formed to 62% reduction and annealed at 12 GPa for CG using Eq. (3) and the load at the first
800 °C for 10 s) predominantly consisted of nanograins pop-in observed in the load–displacement plot. The
(dia. 100 nm) and ultrafine grains (dia. 100–500 nm), measured smax is 1/7th of the theoretical shear modu-
as shown previously [1,3]. lus of stainless steel at 78 GPa [11] and is within the
Representative force–displacement plots for NG/ range for the theoretical strength of polycrystalline met-
UFG and CG specimens acquired via nanoindentation als (G/30 to G/5) [12]. Considering the diameter of
experiments are presented in Figure 2. In the case of the indenter tip (20 nm) and the maximum displacement
NG/UFG steel, a small discontinuity is observed at a of 38 nm (Fig. 2b), we are examining here a small vol-
displacement of 12 nm and an applied load of ume of the material that is initially essentially defect-
0.125 mN. In contrast, the CG steel shows six distinct free. Thus, the first pop-in event in both the NG/UFG
displacement bursts (pop-ins), with the first one occur- and CG is related to nucleation of dislocations and is ex-
ring at a displacement of 10 nm and an applied load pected to occur when the smax approaches the theoretical
of 0.06 mN. The loading data prior to the first value, stheo. In nanoindentation, stresses can reach the
pop-in can be described by a power law relationship theoretical strength of the material [10,13].
(Ph1.5) consistent with the Hertzian elastic contact We now describe the electron microscopy results of the
solution [9,10]: deformed region. Representative results of the deforma-
tion processes observed in NG/UFG and CG and steels
are presented in Figures 3 and 4, respectively. From
Figure 3, we can notice that in the NG/UFG steel twin-
ning has been an active deformation mechanism. Based
on the observation of nanotwins on repeated experimen-
tation, the twins in NG/UFG steel do not thicken as the
strain increases; instead, the high density of twins is a con-
sequence of nucleation of new deformation twins. Also,
shown in Figure 3b is a high-angle twin boundary that
acts as a barrier for gliding dislocation and enhancing
strain-hardening rate. Occasionally, extended disloca-
Figure 1. (a) Light micrograph of starting CG steel (average grain size
tions and short stacking faults were observed. The stack-
of 52 ± 5 lm) and (b) transmission electron micrograph of phase ing faults result from the formation of Shockley partial
reversion and annealed NG/UFG austenitic stainless steel (NG: dislocations, emitted from the grain boundaries of
100 nm; UFG: 100–500 nm). nano-sized grains, and were recently discussed by us [1,3].
R. D. K. Misra et al. / Scripta Materialia 63 (2010) 1057–1060 1059

0.5 0.5
NG/UFG steel CG steel
0.4 Hertzian solution 0.4
Hertzian solution
6

Load (mN)

Load (mN)
0.3 0.3

5
0.2 0.2
4
3
0.1 1 0.1 2
1
0.0 0.0
0 10 20 30 40 0 10 20 30 40
Displacement (nm) Displacement (nm)

Figure 2. Representative force–displacement plots for NG/UFG and CG stainless steel acquired via nanoindentation experiments in the load-
controlled mode with a Berkovich indenter of radius 20 nm (maximum load of 0.5 mN and loading rate of 2 lN s1).

Figure 3. Representative bright-field TEM micrographs of the NG/UFG steel showing (a and b) twinning and (c) with occasional presence of
extended dislocations and stacking faults. The arrow in (b) shows dislocation being stopped at the twin boundary.

of strain-induced martensite at the intersection of shear


bands and a fully developed martensitic structure were
observed in CG steel (Fig. 4a–c), besides heavy disloca-
tion tangles (Fig. 4d). The strain-induced nucleation of
martensite prevents localization of strain and thereby
enhances ductility. Talonen and Hanninen [14] observed
that the intersections of shear bands or stacking faults
act as nuclei for martensite and the formation of shear
bands is a precursor for the strain-induced martensite
formation. Based on observations of Figures 3 and 4,
there is a clear distinction between the deformation
structures in NG/UFG and CG steels that points to-
ward the grain size effect such that twinning contributes
to the excellent ductility of NG/UFG steels, while in CG
steel ductility is also good, but due to martensite forma-
tion. Thus, the strain hardening behavior in NG/UFG
and CG austenitic stainless steel with a stacking fault en-
ergy (SFE) of 15 mJ m2 [14] can be controlled by
twinning and strain-induced martensite formation,
respectively, irrespective of the dislocation glide in both
the systems. The above outlined difference between the
two steels is a new and significant finding.
Figure 4. Representative bright field TEM micrographs of the CG The second and subsequent pop-ins in the CG steel is
steel showing evolution of strain-induced martensite. (a, b) at the a consequence of indentation-induced martensite, where
intersection of shear bands, (c) well-developed martensite laths, and (d) the slope of second and subsequent pop-ins is greater
heavy dislocation tangles. The data is from samples with displacement
than the slope prior to the first pop-in. These slopes
of 25 nm and 35 nm.
are similar to those recorded for the formation of mar-
tensite phase, and the pop-in event is consistent with
Figure 4 summarizes the evolution of microstructure the recent observations of plateau in the load–displace-
for the CG steel for selected displacements to correlate ment plot observed during transformation of austenite
the second and subsequent pop-ins to the deformed to the martensite phase in stainless [15] and TRIP-as-
microstructure. In contrast to NG/UFG steel, nucleation sisted multiphase steels [12]. It is pertinent to state here
1060 R. D. K. Misra et al. / Scripta Materialia 63 (2010) 1057–1060

that a detailed TEM analysis of at least 20 samples in microscopy. In NG/UFG steel, mechanical twinning con-
close proximity to the indentation indicated martensite tributes to excellent ductility, whereas in CG steel, ductil-
formation, while regions some distance away from the ity is governed by gradual nucleation of strain-induced
indenter remained austenite. Furthermore, for the pop- martensite. Thus, ductility in austenitic steels depends
in to be associated with nucleation-controlled martensite on the metastability of the austenite and is affected by
formation, which occurs swiftly, the indenter tip should the grain size. Moreover, the differences in the deforma-
react at an exceedingly higher rate than the experimental tion mechanisms of NG/UFG and CG steels are dis-
loading rate of 2 lN s1 during the phase transforma- tinctly reflected in the force–displacement plots and
tion to maintain a constant loading rate, which produces attributed to differences in austenitic stability associated
pop-in [13]. Thus, based on the above discussion, the ini- with the grain size effect.
tial displacement burst in CG and NG/UFG steels is re-
lated to first generation of dislocations, while the latter The study presented here was supported by
pop-ins in CG are associated with austenite-to-martens- National Science Foundation through Grant No.
ite formation. Another viewpoint is that pop-in associ- CMMI 0940402 and Center and Functional Materials,
ated with martensite formation represents geometrical University of Louisiana at Lafayette. M.C.S. and
softening induced by the martensite variant selection L.P.K. acknowledge with gratitude the funding of The
that minimizes the total energy during austenite-to-mar- Finnish Funding Agency for Technology and Innova-
tensite formation [13]. Mechanical twinning can also tion (Tekes) via FIMECC Oy in the LIGHT-program.
cause twin variant selection, though, unlike martensite Discussion with Professor H.N. Han of Seoul National
formation, the nanotwins observed in Figure 3 lack sig- University, Korea is gratefully acknowledged.
nificant distortion of the lattice such that the concept of
geometrical softening is not applicable.
In regard to the mechanistic change in the deforma- [1] R.D.K. Misra, S. Nayak, S. Mali, J. Shah, M. Somani,
tion behavior from strain-induced martensite and dislo- L.P. Karjalainen, Met. Mater. Trans. A 41 (2010) 1543.
cation tangles in CG steel to intense mechanical [2] R.D.K. Misra, S. Nayak, S. Mali, J. Shah, M. Somani,
twinning in NG/UFG steel, this can be described in L.P. Karjalainen, Met. Mater. Trans. A 41 (2010) 3.
terms of the interplay between austenite stability and [3] R.D.K. Misra, S. Nayak, S. Mali, J. Shah, M. Somani,
SFE via the grain size effect. Deformation twinning in L.P. Karjalainen, Met. Mater. Trans. A 40 (2009) 2498.
[4] M.C. Somani, P. Juntunen, L.P. Karjalainen, R.D.K.
NG/UFG steel and strain-induced martensite in CG
Misra, A. Kyrolainen, Met. Mater. Trans. A 40 (2009)
steel are microstructurally similar, i.e. both of them in- 729.
volve diffusionless shear of a constrained plate-shaped [5] K. Nohara, Y. Ono, N. Ohashi, J. Iron Steel Inst. Japan
region in the parent crystal but differ from each other 63 (1977) 772.
such that the latter is driven by a chemical free energy [6] A. Frehn, E. Ratte, W. Bleck, Steel Grips 2 (2004) 447.
change. First, it is known that austenite stability in- [7] S. Turtletaub, A.S.J. Suiker, Inter. J. Solids Mech. 43
creases with decrease in grain size [5–7]. Second, there (2006) 7322.
are a few reports in the literature suggesting that the [8] F. de las Cuevas, M. Reis, A. Ferraiulo, G. Pratolongo,
SFE increases with decrease in austenite grain size from L.P. Karjalainen, J. Alkorta, J. Gill Sevillano, Key Eng.
a few micrometers to 50 lm [16–18]. The SFE includ- Mater. 423 (2010) 147.
[9] K.L. Johnson, Contact Mechanics, Cambridge University
ing the grain size effect was referred as “effective SFE”.
Press, Cambridge, 1985.
The plastic deformation of austenitic alloys is related to [10] S. Shim, H. Bei, E.P. George, G.M. Pharr, Scripta Mater.
the SFE of the austenite matrix and the austenite stabil- 59 (2008) 1095.
ity is greater for alloys with comparatively higher SFE. [11] http://asm.matweb.com/search/SpecificMaterial.=MQ302AL
Thus, the enhanced contribution of grain boundaries [12] R.W. Honeycombe, Plastic Deformation of Metals, Sec-
and their adjacent zones (slip systems and availability) ond ed., Edward Arnold, London, 1984.
resulting in the increased stability of the austenite in [13] T. -H Ahn, C.-S. Oh, D.H. Kim, H. Bei, E.P. George,
the NG/UFG steel, with a SFE of about 15 mJ m2, de- H.N. Han, Scripta Mater. 63 (2010) 5240.
lays the phase transformation and twinning is observed [14] J. Talonen, H. Hanninen, Acta Mater. 55 (2007) 6108.
instead of martensite formation. [15] Q. Furnemont, M. Kempf, P.J. Jacques, F. Delannay,
Mater. Sci. Eng. A 328 (2002) 26.
In summary, we have provided evidence for the first
[16] P. Yu Volosevich, V.N. Gindnev, Y.N. Petrov, PhI. Met.
time that the deformation mechanisms of NG/UFG Metallog. 42 (1976) 126.
austenitic stainless steels with high strength–high ductility [17] A. Rohatgi, K.S. Vecchio, G.T. Gray, Met. Mater. Trans.
combination is distinctly different from the CG counter- A 32 (2001) 135.
part, as documented via nanoindentation and electron [18] J.-H. Jun, C.-S. Choi, Mater. Sci. Eng. A 257 (1998) 353.

S-ar putea să vă placă și