You are on page 1of 9

Materials Technology

Some Strengthening Methods for Austenitic Stainless Steels


L.P. Karjalainen1), T. Taulavuori2), M. Sellman2), A. Kyrlinen3)
1) 2)

Department of Mechanical Engineering, University of Oulu, 90014 Oulu, Finland, pentti.karjalainen@oulu.fi Research Centre, Outokumpu Tornio Works, 95400 Tornio, Finland 3) Steelpolis, Raahe Region Technology Center Ltd, 92100, Raahe, Finland Austenitic stainless steels possessing good corrosion resistance have recently found growing applications as a constructional material. In this instance, increasing strength properties, which are typically quite low, is of great interest. Due to the low stacking fault energy, strain hardening of alloyed austenite is efficient for increasing tensile strength without impairing ductility seriously. In addition, certain grades are unstable, so that cold working creates strain-induced martensite that enhances strengthening. Grain size refinement to micrometer scale or even finer can also increase the yield strength, still providing good ductility. In the present paper dislocation and phase transformation strengthening and thereby properties achievable in temper rolled austenitic stainless steels are discussed. Strengthening by the reversion annealing is also described and excellent results achievable are shown. Finally, the effect of bake hardening through the static strain ageing is presented. Long-term research work in various projects indicates that the current knowledge of strengthening of austenitic stainless steels is close to the industrial utilisation. Keywords: austenitic stainless steels, strength, ductility, strain hardening, strain-induced martensite, grain size refinement, bake hardening, strain aging. DOI: 10.2374/SRI08SP040-79-2008-404; submitted on 24 January 2008, accepted on 26 February 2008.

Introduction Development of advanced high strength steels is a target set by the European Commission in its Steel Platform ESTEP for the year 2030. Accordingly, e.g. for automotive and transport applications, steels with high tensile strength and ductility should be developed, as shown in Figure 1 [1]. Stainless steels are among those new generation materials desired. Austenitic stainless steels are traditionally used in various applications in which good corrosion resistance is required. For structural applications, however, their demerit is that the yield strength is quite low, e.g., 230-260 MPa according to EN 1.4301 (AISI 304), and 350-380 MPa corresponding to EN 1.4318 (AISI 301LN). Tensile strength is in the range of 600 to 800 MPa and total elongation from 45 to 60 %. In regard to the ratio of strength and elongation (as illustrated in Figure 1), however, stainless steels are located well above the ranges of carbon steels. Nevertheless, there is a great interest in improving the mechanical properties of commercially available austenitic stainless steels, so that these materials could be better utilized in structural applications and automotive industry [2-4]. Nitrogen alloying and cold rolling are methods generally applied to increase the strength. However, in conventional Cr-Ni austenitic stainless steels, there is a low limit for nitrogen solubility. Problems with the hot ductility may also appear with increased nitrogen content. This topic will not be discussed here. In the following, the cold rolling strengthened austenitic stainless steels will be considered and a brief discussion is also given related to the physical metallurgical factors
404

affecting their strengthening behaviour. However, the intention of this paper is not to be any comprehensive literature survey of those factors. Furthermore, more uncommon strengthening mechanisms, reversion and bake annealing treatments, are described and typical enhanced properties are shown. Strengthening by Cold Rolling The yield strength of austenitic stainless steels, and particularly that of metastable alloys, can be drastically improved by cold rolling. Cold forming generates straininduced martensite in addition to ordinary strain hardening of high-alloyed austenite. In fact, it seems that the interaction of plasticity and phase transitions provides a route to develop materials with exceptional combinations of strength and ductility. The strength levels of commercial temper-rolled grades according to the design manual for structural stainless steel

Figure 1. The target range set by the European Commission for automotive steel development compared to present materials [1]. Typical property range of conventional annealed austenitic stainless steels is inserted.

steel research int. 79 (2008) No. 6

Materials Technology
Table 1. Strength levels of cold worked stainless steels for structural applications (applicable to the sheet thickness 6 mm) [5].

At Outokumpu, research has been conducted on structural cold strengthened stainless steels since the 1980s, mainly based on the utilisation of the Sendzimir process, but recently on processing using the integrated RAP (rolling-annealing-pickling) line. Temper-rolled grades have found applications in e.g. structural components of cargo-vehicles in light-weight structures and load-bearing constructional elements, in honeycomb structures [6] (Figure 2) and in automotive and transportation [7], e.g. tankcontainers. However, their use in pressure vessels is still restricted, even though the possibility to obtain Particular Material Appraisal (PMA) exists for a defined pressure equipment application according to the Pressure Equipment Directive (PED).

are shown in Table 1 [5]. In practice, in production recorded by Outokumpu, the mechanical properties are somewhat higher, for instance for the grade +CP500/+C850, which equals ASTM Hard, typically the yield strength Rp0.2 is 650 MPa, the tensile strength Rm 970 MPa and the total elongation A80 30 %. Besides high strength, the temper-rolled austenitic stainless steels possess good formability. In Table 2 it can be seen that especially the grade EN 1.4318 (AISI 301LN) has a low ratio of Rp0.2 to Rm (typically about 0.67 for +C850 grade) that gives advantages in subsequent forming processes. This is a result from unstable austenitic structure. Furthermore, the good ductility enables subsequent processing and intense work hardening during the forming process guarantees that the final product fulfils the mechanical requirements.

Dislocation Strengthening Inherently, austenitic stainless steels are characterized by low yield strength but very good uniform and total elongation. The strain-hardening rate of a single-phase metal or alloy in plastic straining is dependent on the type of dislocations in the lattice, perfect or extended, and their ability to move, slip, cross-slip or climb, to pass by the obstacles. Generally, in austenitic steels the stacking fault energy (SFE) is low, so that dislocations are dissociated to partial Shockley dislocations, while the two partial dislocations are bound together by the stacking fault, and they will therefore move together as a unit along the slip plane. In the instance of SFE 10 mJ/m2, the separation distance may be about 10 nm without external stress, but can become even infinite under applied stress [8]. The width of the stacking fault ribbon is of importance, since the two partial dislocations have to be brought together to form a perfect dislocation before intersection or cross-slip can occur. This constriction requests for higher applied stress, i.e. enhances flow stress and the strain-hardening rate and results in planar dislocation arrays. Some investigations on CrNi-based austenitic steels indicated that planar slip does not always need a low SFE, but another reason for planar glide in austenitic alloys might be the phenomenon of short range ordering [9,10]. SFE of austenitic steels is a function of alloy composition and temperature. It has been determined from experiments and semi-empirical simulations [11-14]. However, SFE is difficult to measure precisely, so that large inaccuracies may be associated with the experimental values quoted from the literature. Schramm and Reed [11] determined the SFE of seven commercial austenitic stainless steels by X-ray diffraction line profile analysis. They suggested a simple regression equation as follows: SFE = - 53 + 6.2 (Ni) + 0.7 (Cr) + 3.2 (Mn) + 9.3 (Mo) (1) where SFE is in (mJ/m2) and the elements are in mass-pct. Recently, Talonen and Hnninen [15] measured the SFE

Figure 2. Temper-rolled materials have been used successfully in lightweight sandwich structures.

Table 2. The ratio of Rp0.2 / Rm for some temper-rolled stainless steel grades calculated from the values in EN 10088-2.

steel research int. 79 (2008) No. 6

405

Materials Technology

Figure 3. Change of mechanical properties as a function of cold rolling reduction. (a) 18.5Cr-9.1Ni [24], (b) EN 1.4318 [25].

Figure 4. Martensite fraction formed in cold rolling in (a) various steel grades [26], (b) in EN 1.4318 at small or large rolling passes [27,28] and in tension at 0.1 s-1[15].

for EN 1.4318 (301LN) type steels by X-ray diffraction and obtained the average value of about 14 mJ/m2. This value is close to that calculated from the compositionbased equation of Brofman and Ansell [12] SFE (mJ/m2) = 16.7 + 2.1(Ni) - 0.9(Cr) + 26(C) (2)

Metastability and Strengthening In addition to strain hardening of austenite owing to obstructed movement of dislocations due to increasing dislocation density and obstacles formed by dislocation interactions, phase transformation can occur during plastic straining in metastable alloys as strain-induced. Numerous studies on the topic show that the transformation sequence is (fcc) to (hcp) to ' (bcc), e.g. [18-19]. However, the hcp-phase may not form as an intermediate phase if the composition is changed such that the SFE is increased. Talonen and Hnninen [15] detected the presence of martensite only in steels with SFE <13 mJ/m2. -martensite formation is also dependent on the crystallographic orientation with respect to the applied stress and the deformation mode (tension or compression) [20-22]. It has been found that the intersection of stacking faults or hcp-shear bands act as nuclei for '-martensite and the formation of the shear bands is a necessary precursor for the strain-induced '-martensite transformation [23]. Therefore, in addition to that SFE affects the deformation mechanism, SFE must also play an important role in determining the austenite stability, since it controls the formation of the shear bands, and thus the formation of nucleation sites for the '-martensite [15]. Martensite is acting as a reinforcing phase. However, as seen in Figure 3, the cold rolling reduction required for EN 1.4301 (AISI 304) to achieve the +CP500/+C850
steel research int. 79 (2008) No. 6

while Equation (1) of Schramm and Reed would result in distinctly lower values. These regression equations indicate certain relationships between SFE and alloy composition, so that, in principle, the SFE could be controlled by adjusting the concentrations of the alloying elements in order to get the properties such as strength and work hardening rate towards desired values. However, Vitos et al. [16] calculated the SFE of austenitic stainless steels using a quantum mechanical first-principles approach and they demonstrated that the same alloying element can cause totally opposite changes in the SFE of Cr-Ni alloys with different host composition. This means that no universal composition-based equations for the SFE can be established, and therefore the general value of regression equations may be questionable. It can also be mentioned that the effect of nitrogen, an important alloying element, on the SFE in austenitic stainless steels is still a controversial issue, as discussed by Gavriljuk et al. [17]. Both increasing and decreasing influences have been reported.
406

Materials Technology

grade level is about 10% and for EN 1.4318 (AISI 301LN) even less. This means that the '-martensite fraction created in this stage is still very low; about 2% in EN 1.4301 and about 5% in EN 1.4318 (Figure 4a); so that the marked increase in the yield strength must be mainly due to strain hardening: dislocations, twins and stacking faults, or the Suzuki locking. However, the bcc-phase formation in straining markedly affects the tensile strength, strain hardening and ductility of the material. There has been much previous work on this topic aimed at delineating compositions, which are most effective in producing high strength levels, and at gaining a fundamental understanding of martensite induced by concomitant plastic strain. The amount of -martensite is dependent on alloying and temperature, plastic strain, deformation method (applied stress state), grain size, strain rate and the characteristics of deformation mechanisms, e.g., Frehn et al. [29]. Most recently Ratte [30] has discussed the factors affecting the martensite formation in metastable austenitic stainless steels. A quantitative guideline regarding the influence of steel composition on the strain-induced martensite formation is the M d30 temperature, where 50% '-martensite is present after 30% tensile deformation. Some empirical equations have been proposed to estimate the martensite fraction including the influence of chemical composition. A commonly used equation is that of Nohara et al. [31]: Md30N (C) = 552 - 462(C+N) - 9.2Si - 8.1Mn - 13.7Cr 29(Ni+Cu) - 18.5Mo - 68Nb - 1.42(GS-8) (3) where the elements are in mass-pct and GS is the ASTM grain size. Sjberg [32] has suggested an equation taking the independent effects of carbon and nitrogen into account: Md30 (C) = 608 - 515(C) - 821(N) - 7.8(Si) - 12(Mn) 13(Cr) - 34(Ni) - 6.5(Mo) (4) There are a number of models proposed to describe the rate of martensite transformation per unit strain in austenitic stainless steel, e.g. [33,34]. The geometric model of Olson and Cohen [34] is useful as it is simple to interpret results, but it does not incorporate the transformation strain from fcc to bcc. Recently Spencer et al. [35] developed a transformation model based on assumptions that there is a critical strain required to generate hcp-shear bands from stacking faults, only the intersection volume of hcp-bands act as nucleation sites and the martensite will not grow beyond the intersection volume of two hcp-bands. However, even this model calls for further development. In steel manufacturing industry, cold rolling - or in some cases stretching of lower strength classes - is the processing stage to strengthen austenitic stainless steels. In Figure 4b it is demonstrated for EN 1.4318 steel in laboratory rolling that the '-fraction is highly dependent on the pass reduction, large passes resulting in much reduced '-fractions, obviously due to adiabatic heating occurring in cold rolling [27]. For comparison, data for
steel research int. 79 (2008) No. 6

martensite fraction formed in a tension test at the strain rate of 0.1 s-1 is also included [15]. If considering safety components in automotive applications, very high deformation rates may occur in crash accidents, for example. It has been observed that the tendency for -martensite formation is reduced for higher strain rates [15,36], as can be expected. This fact is also evident from tension tests, see e.g. [15,36]. This behaviour seems to be associated with the adiabatic heating generated during deformation at high velocities and its effects on the transformation temperature. Anisotropy and Asymmetry As a design problem, it has been observed that a pronounced directional anisotropy exists in Rp0.2 values in temper-rolled grades, while there is only negligible anisotropy in the soft annealed condition. For instance, it has been found that typically for EN 1.4318 (3 mm sheet), Rp0.2 values of specimens cut transverse to the rolling direction were only about 86% of those for the longitudinally cut specimens [26,37]. Under compression the opposite is prevailing, a decrease in Rp0.2 for longitudinally cut specimens compared to the specimens cut to the transverse direction. However, the decrease in the elastic modulus was not marked. Factors that affect the anisotropy need further investigations to be taken into account for structural design. In addition to the directional anisotropy being present in cold-strengthened grades, the ratio between the Rp0.2 values, measured for longitudinally cut specimens in compression and in tension, was measured to be 0.8, as an indication of pronounced asymmetry in the mechanical properties [26,37]. The opposite was valid for specimens cut transverse to the rolling direction. To take this asymmetry into account in structural applications where compression in the longitudinal direction is a relevant stress state (e.g., column behaviour in bending, where the cross-section is predominantly under compression), the characteristic value for the design strength should be taken as 0.8 x Rp0.2 given in Table 1. A higher value may be used if supported by appropriate experimental data. Softening of Temper-rolled Grades in Welding In welding of temper-rolled grades some softening in the heat-affected zone can be expected and consequently some loss of cold-deformation improved mechanical properties. As an example, Figure 5 shows the softening during annealing at various temperatures and times as revealed by the drop in hardness of EN 1.4318 C850 steel. However, in a short thermal cycle, marked softening by recovery and recrystallisation takes place only at high temperatures above 1100C, so that the softened zone in the heat affected zone of the base metal remains quite narrow. The influence of narrow zones with different strength and ductility properties, i.e. the mismatch effect, on the behaviour of the welded joint is a topic under current investigation and a model will be developed to predict the mechanical properties of a welded joint with the mismatching zones [38].
407

Materials Technology

Figure 5. Drop in hardness of EN 1.4318 C850 during annealing at various temperatures and holding times [38].

As expected, the degree of softening is dependent on the welding heat input, and a method such as laser welding can be recommended to minimize the degree of softening and the width of the soft zone. As shown in Figure 6, laser welding results in the equal strength of the joint with the strength class +CP500 base metal, while softening in TIG welding is detectable in all cases [39]. Welding tests for EN 1.4318 sheets of 3 mm thickness carried out with the tandem-MIG method using two passes reveal that he hardness of the soft zone in the HAZ is about 210 HV compared to 250 HV of the original hardness of the +C850 sheet, whereas the hardness of the weld metal is about 170 HV, i.e. much lower (Figure 7) [40]. Furthermore, the fusion boundary and the soft zone of the upper bead are not vertical, but very inclined as evident in Figure 7. It has been observed that the soft zone is not a critical zone in tension, except at excess heat inputs. The strength values fulfilling the requirements for the grade +C850 base metal can be achieved, for instance, using soft welding filler wire (such as 308LSi) and a proper heat input (in the range 2-2.5 kJ/cm), while the adequate height of the surface bead provides the extra strength required for the joint, see Figure 8. Grain Size Strengthening In ferritic structural steels, refining of the grain size is a well-known effective method to increase the yield strength. High attention has been paid to processing of steels with ultra-fine grain size, typically in the order of one micrometer, e.g. [41,42]. As a result, a yield strength of about 800 MPa can be achieved even in plain C-Mn grades, but unfortunately the strain-hardening rate becomes very low and the ductility/elongation remains low. In austenitic stainless steels, the grain size refinement by thermo-mechanical processing similarly as in ferritic steels is not possible, because there is no transformation of austenite to ferrite and the recrystallisation of austenite takes place at high temperatures and it does not refine the grain size markedly. However, grain refinement would be beneficial, for an established Hall-Petch type relationship suggests for 12.5Cr-9.5Ni-2Mo-0.1N steel that [43]: Rp0.2 (MPa) = o + k d-0.5 = 150 + 17 d -0.5 (5)

Figure 6. Yield strength of TIG or laser-welded joint of various temper-rolled EN 1.4301 grades compared to the values of base metal and standard requirements [39].

where o is the friction stress, k Hall-Petch slope and d grain size in mm. It follows that Rp0.2 is about 700 MPa and 1000 MPa corresponding to a grain size of 1 m and 0.5 m, respectively. Rajasekhara et al. [44] obtained a smaller k in another relationship Rp0.2 (MPa) = 252 + 8.7 d -0.5 (6)

Figure 7. An example of a cross section of a double-pass joint and its hardness distribution near root and upper surface, an undermatching filler wire was used [40]. (FL=fusion line).

for the reversion treated ultra-fine grained EN 1.4318 steel, but still fine grain size would mean higher strength. There are several Japanese publications describing the processing route to obtain sub-micron grain size in metastable austenitic stainless steels [43,45,46]. This approach consists of heavy cold rolling to induce the formation of stress-induced martensite and its deformation, followed by an annealing treatment to revert the deformed martensite into ultra fine-grained austenite. This process
steel research int. 79 (2008) No. 6

408

Materials Technology

was also reported by di Shino et al. [47] for EN 1.4310 (AISI 301) steel to obtain grain sizes around 2 m. Recently this processing route has been investigated extensively in a collaborative project at the University of Oulu and University of Texas [27,48-50]. In the research performed it has been found that in EN 1.4318 steel, even after cold rolling reductions as low as 45%, leaving a considerable amount of cold-deformed austenite in the strained structure, submicron sized austenite is readily formed in the diffusion-controlled reversion process. The reversion kinetics is dependent on the prior cold rolling reduction, annealing temperature and time, but it is very fast above 750C. The grain size after annealing at 800C for 1 s is slightly below 1 m (cold rolling reduction 62%). A typical fine-grained structure is shown in Figure 9a in a SEM-EBSD photo. After cold rolling reduction of 45% and annealing for 10 s at 800C, the grain structure is slightly coarser and it can be examined with optical microscopy (Figure 9b). In addition to fine grains, it contains some large austenite grains corresponding to retained austenite, untransformed in the rolling stage, which tends to recrystallise at 850-900C at longer times than the reversion occurs. The reversion at temperatures of 730-800C results in an excellent combination of yield strength and elongation, as demonstrated by typical stress-strain curves in Figure 10. Also tensile strength is high due to intensive strain hardening in tensile straining. Therefore, after optimum reversion treatments, the Rm - elongation combinations can even reach the range targeted by European Commission, as demonstrated in Figure 11. Rajasekhara et al. [44] have considered the role of fine grain size on Rp0.2 of the reversion annealed EN 1.4318 steel. They showed that the maximum contribution of the grain size can be substantial, about 250 MPa, in the instance of annealing at 800C for 1 s providing a micrometer scale austenite grain size. However, it may be concluded from the high strength-elongation combination of EN 1.4310 (AISI 301) steel obtained without any fine austenite grain size that (i) coarse shear-reversed austenite is present after annealing at 800C for a short period of time [27,48]; (ii) ultra-fine grain size is not absolutely crucial, whereas the equal contribution can be readily obtained also from dislocation strengthening, i.e. from dislocations inherited from the deformed martensite in the shear-type transformation. It seems that a more important factor regarding tensile strength and high elongation is a proper instability of the austenite. It should not be too high as evident when comparing the very different elongations between 17/7N (17Cr-7Ni-0.115N) and EN 1.4310 steels having practically the same Rm, and similarly between 17/7C (17Cr-7Ni-0.095C) and EN 1.4318 steels as annealed at 800C for 1s (Figure 12). It is obvious that the gradual formation of strain-induced martensite in tensile straining is a key factor for high ductility. In addition, it has to be kept in mind that a high degree of cold working may cause unfavourable internal stresses in the material but in reversion annealing the abovementioned high strength-elongation values can be achieved in a less-deformed structure, which may be of advantage for subsequent processing. This also seems to improve the corrosion resistance of the EN 1.4318 steel [51].
steel research int. 79 (2008) No. 6

Figure 8. Force-strain curves for a tandem-MIG welded joint and the 1.4318 C850 base metal.

Figure 9. Ultra fine grain size obtained in reversion annealing of EN 1.4318 at 800C (a) for 1 s after 62% cold rolling reduction (SEM-EBSD), (b) for 10 s after 45% reduction.

Figure 10. Stress-strain curves for EN 1.4318, as cold rolled 45% and reversion annealed. Typical curve for +C850 cold rolled grade is also plotted [27].

Figure 11. Tensile strength-elongation (A10) achieved for EN 1.4318, as cold rolled 45-72% and reversion annealed at 700900C for 1-10 s (data from [27,48-50]), compared to the target range set in ESTEP by the European Commission [1].

409

Materials Technology

It has been noticed that in reversion annealing of high carbon grades such as EN 1.4310, formation of carbides can also occur at grain boundaries, which may cause some degree of sensitization of the steel [52]. In production of longitudinally welded rectangular tubes from temper-rolled band, the tube corners become more deformed, reaching typically the deformation degree of about 40%, e.g. in the case of 40x40 mm tube of 2 mm wall thickness. Consequently, in the grade EN 1.4318 C850, the martensite content increases up to 35-40% near the inner surface at the corner region. A novel patented idea is to reversion anneal the martensite inside the corners to transform it to ultra-fine grained austenite, in order to improve the formability of the material of the corners that is needed in subsequent forming and shaping processes, e.g. in bending of the tube. Preliminary Gleeble annealing trials indicate that excellent Rp0.2/Rm/A10 combinations can be obtained by reversion.
Figure 12. Stress-strain curves of reversion annealed (at 800C) Cr-Ni steels with various preceding cold rolling reductions (CR) revealing very different elongation dependent on their strain hardening behaviour [49].

Bake Hardening Strengthening Bake hardening resulting from the static strain ageing is a typical method for increasing the yield strength of relatively low-strength mild steels used for automotive body sheets. Typically temperatures below 200C are sufficient and the combinations of temperature and time are well suited for the painting processes used in automotive industry. The R&D work performed in 1990s on advanced high strength steels showed that steels having dual-phase structure (DP), transformation induced plasticity (TRIP) and complex/multiphase (CP) also exhibit bake hardening behaviour as cold worked. A yield strength increase up to 200 MPa has been measured [53]. Recently, interest in the bake hardening treatment of austenitic stainless steels has been growing and more research has been targeted to this field. The frontier projects such as LIGHT&SAFE Weight reduction for safer, affordable passenger cars by using extra formable high strength austenitic steel funded by EU Framework Program 5 has increased the awareness for this strengthening method. Typically, a yield strength increase by over 100 MPa seems to be achievable with commercial austenitic stainless steels grades, as indicated in Figure 13, without impairing ductility and formability [54,55]. Some further studies performed using commercial austenitic stainless steel grades evidenced that at a cold rolling reduction of 30% bake annealing can result in an increase by 230 MPa [56], as shown in Figure 14 (even 300 MPa has been reported under optimal conditions [57]). Correlation between the -martensite fraction and the magnitude of strain ageing was demonstrated earlier in many papers, but hardening was remarkably even in materials having a martensite fraction as low as 1 to 5 %, for instance in the case of Mn-alloyed EN 1.4372 (AISI 201) steel, as illustrated in Figure 14. With higher cold working reduction, a yield strength of 1300 MPa (Rm ca. 1400 MPa) and a total elongation of 5 to10% can be reached with EN 1.4318 (AISI 301LN) by optimizing the martensite content of around 30-35% as well as the duration and temperature of the aging treatment [57,58]. Respectively, the maximum value reported for the grade
steel research int. 79 (2008) No. 6

Figure 13. Engineering stress-strain curves of steel EN 1.4318 prestrained in tension to the engineering strains of 0.05, 0.15, 0.25 and aged at 170C revealing the influence of strain aging [54].

Figure 14. Change in the yield strength of various 20 % and 30 % cold rolled steels in the rolling (RD) and transverse (TD) direction after ageing at 170C [56].

410

Materials Technology

EN 1.4310 is 1375 MPa (Rm ca. 1500 MPa) with a fracture elongation of 13 % [56]. The magnitude of strength increment depends on the direction of straining in tensile testing. Static strain ageing seems to be more pronounced when straining occurs transverse to the rolling direction, as shown in Figure 15. Typically, increasing the degree of cold working reduction increases the anisotropy, but according to the results in Figures 14 and 15, the strain aging seems to diminish this behaviour [56]. The mechanism of static strain ageing in austenitic steels has been studied but it is not fully understood yet. It has been suggested that the mechanism may involve redistribution of interstitial atoms in the strain-induced martensite or changes in the residual microstress state [59]. However, it seems that the strain ageing is quite significant even in the austenitic structure with very low martensite fractions as seen here in Figure 15. Furthermore, the activation energy of the ageing process is close to that of C and N diffusion in the austenite phase. These features may indicate that the static strain ageing is related to microstructural changes in the austenite phase, probably involving formation of atom pairs, irreversible effects of temperature on stacking faults and the Suzuki chemical interaction [59].

utilisation, to improve the mechanical properties and their uniformity in temper rolled steels.

Figure 15. Relationship between the martensite fraction and the increase in the yield strength both parallel to the rolling direction (RD) and transverse to rolling direction (TD) [55].

References Conclusions To substantially increase the strength of austenitic stainless steels without significantly impairing their ductility, cold rolling, reversion annealing and bake hardening are three potential processes to apply. In unstable grades, cold deformation, e.g. temper rolling, results in the formation of strain-induced martensite, which together with the effective strain hardening of Cr-Ni austenites increases the yield strength, depending on the reduction, up to 500 to 1300 MPa and the tensile strength to 850 to 1600 MPa with elongation from 50 to 5 %, respectively. Reversion annealing of cold rolled unstable grades (EN 1.4318 in particular) can change the martensite to micron-scale grained austenite that increases the yield strength by about 250 MPa. Excellent combinations of tensile strength and elongation (e.g., 1000 MPa/45%) are achievable, even reaching the target range set in the European Steel Platform. Under optimum conditions, bake hardening is very effective for increasing the yield strength by 230-300 MPa, and e.g. a combination of 1375 MPa yield strength with 13% elongation has been obtained. Research on cold rolling strengthening of austenitic stainless steels has taken place over two decades and the mechanism and phenomena are quite well understood and strength grades are standardized for industrial use and design. However, anisotropy and asymmetry in mechanical properties present in these grades still call for further research to be accounted for in design rules. Reversion annealing for grain size refinement of austenitic stainless steels, reported more than 15 years ago, is once more under increasing interest and may result in certain industrial applications. Research on bake hardening of austenitic stainless steels has shown its effectiveness and simplicity in processing, so that it is ready for industrial
steel research int. 79 (2008) No. 6
[1] European steel technology platform (ESTEP), Strategic research agenda; A vision for the future of the steel sector, European Commission, Belgium, 2005. [2] P-J. Cunat, T. Pauly: Stainless steel as a competitor to light material for building and automotive applications. Proc. of the 4th European Stainless Steel Science and Market Congress, 2002, Paris, France, p.10. [3] R. Andersson, E. Schedin, C. Magnusson, J. Ocklund, A. Persson: Stainless steel components in automotive vehicles. Proc. of the 4th European Stainless Steel Science and Market Congress, 2002, Paris, France, p.57. [4] C. Greisert: Strain-induced formation of martensite during forming and springback behaviour of annealed and hard cold-rolled stainless steel grades EN 1.4301 and EN 1.4318, PhD Thesis, RWTH Aachen, Helsinki University of Technology, 2004. [5] Design Manual For Structural Stainless Steel - Third Edition, Euro Inox and The Steel Construction Institute, 2006, p.19. [6] A. Gales, M. Sirn, J. Synjkangas, N. Akdut, D. Hoecke, R. Sanchez: Development of lightweight trains and metro cars by using ultra high strength stainless steels, Report EUR 22837 EN, European Communities, Luxembourg, 2007, p.15. [7] A. Kyrlinen, M. Vilpas, H. Hnninen: J. Materials Engineering and Performance, 9 (2000), No. 6, 669. [8] T.S. Byun: Acta Materialia, 51 (2003), 3063. [9] V. Gerold, H.P. Karnthaler, Acta Metallurgica, 37 (1989), 2177. [10] G. Saller, K. Spiradek-Hahn, C. Scheu, H. Clemens: Materials Science and Engineering A, 427 (2006) 246. [11] R.E. Schramm, R.P. Reed: Metall. Trans., 6A (1975), 1345. [12] P.J. Brofman, G.S. Ansell: Metall. Trans., 9A (1978), 879. [13] P.J. Ferreira, P.A. Mullner: Acta Materialia, 46 (1998), 4479. [14] I.A. Yakubtsov, A. Ariapour, D.D. Petrovic: Acta Materialia, 47 (1999), 1271. [15] J. Talonen, H. Hnninen: Acta Materialia, 55 (2007), 6108. [16] L. Vitos, J.-O. Nilsson, B. Johansson: Acta Materialia, 54 (2006), 3821. [17] V. Gavriljuk, Yu. Petrov, B. Shanina: Scripta Materialia, 55 (2006), 537. [18] K. Spencer, J.D. Embury, K.T. Conlon, M. Vron, Y. Brchet: Materials Science and Engineering A, 387389 (2004), 873. [19] P.L. Mangonon, G. Thomas: Metall. Trans., 1 (1970), 1577. [20] J.R. Patel and M. Cohen: Acta MetaIlurgica, lA (1953), 531. [21] G.B. Olson, M. Cohen: Metall. Trans., 7A (1976), 1897. [22] L. Bracke, G. Mertens, J. Penning, D. Cooman, M. Liebeherr, N. Akdut: Metall. Mater. Trans. A, 37A (2006), 307.

411

Materials Technology
[23] P. Hedstrm, U. Lienert, J. Almer, M. Odn: Scripta Materialia 56 (2007), 213. [24] A. Kyrlinen: High strength stainless steels temper rolled, International Symposium on Advanced in Stainless Steels 2007, ISAS 2007, Chennai, India, 9-11, April, 2007, The Indian Institute of Metals, CD. [25] R. Tarkiainen: Effect of cold rolling and carbon and nitrogen contents on the properties of a metastable stainless steel, Master thesis, Oulu University, Oulu, Finland, 1991, p.64 (in Finnish). [26] J. Salmn: The anisotropic behaviour of the mechanical properties in hard cold rolled austenitic stainless steels, Master Thesis (in Finnish), University of Oulu, Finland, 2004. [27] M.C. Somani, L.P. Karjalainen, M. Koljonen, P. Aspegren, T. Taulavuori, A. Kyrlinen: Microstructure and mechanical properties of reversion annealed cold-rolled 17Cr-7Ni type austenitic stainless steels, Stainless Steel 05, Proc. of the 5th European Congress, Stainless Steel Science and Market, (eds. J.A. Odriozola and A. Paul), Seville, Spain, Sept. 27-30, 2005, p.37. [28] M. Koljonen, Developing a fine-grained austenitic stainless steel by cold rolling and annealing, Master thesis, Oulu University, Oulu, Finland, 2005, 75 pp. [29] A. Frehn, E. Ratte, W. Bleck: Steel Grips, 2 (2004), 447. [30] E. Ratte: Sense and sensitivity of thermo-mechanical forming simulation of metastable austenitic stainless steels, Proc. of the 6th European Stainless Steel Conference - Science and Market, June 1013, 2008, Helsinki (in press). [31] K. Nohara, Y. Ono, N. Ohashi: J. Iron Steel Institute Japan, 63 (1977), No. 5, 772. [32] J. Sjberg: Wire, 23 (1973), 155. [33] R.G. Stringfellow, D.M. Parks, G.B. Olson: Acta Metallurgica, 40 (1992), 1703. [34] G.B. Olson, M. Cohen: Metallurgical Transactions, 6A (1975), 791. [35] K. Spencer, J.D. Embury, K.T. Conlon, M. Vron, Y. Brchet: Materials Science and Engineering A, 387389 (2004), 873. [36] J. Talonen, P. Nenonen, G. Pape, H. Hnninen: Metallurgical and Materials Transactions A, 36A (2005), 421. [37] T. Taulavuori, P. Aspegren, J. Synjkangas, J. Salmn and P. Karjalainen: The anisotropic behaviour of the nitrogen alloyed stainless steel grade 1.4318, Proc. Intern. Conf. on High Nitrogen Steels 2004, (eds. N. Akdut, B.C. DeCooman, J. Foct), 19-22 Sept. 2004, Oodense, Belgium, pp. 405. [38] Mis-Match project (Tekes 40031/07; No 2455/31/06 ) (unpublished results). [39] A. Kyrlinen, P. Yrjl, V.-P. Immonen, S. Sainio: Hitsaustekniikka, (2001), No. 1, 14 (in Finnish). [40] T. Oikarinen: Tandem MIG/MAG welding of EN 1.4318 C850 grade austenitic stainless steel, Master thesis, University of Oulu, 2008 (in Finnish). [41] A.K. Ibraheem, R. Priestner, J.R. Bowen, P.B. Prangnell, J.F. Humphreys: Iron and Steelmaking, 28 (2001), 203. [42] L.X. Pan, L.P. Karjalainen, M.C. Somani: Processing and stability of ultrafine grained structures in some microalloy steels, Materials Science Forum, 500-501 (2005), 363. [43] S. Takaki and Y. Tokunaga: Ultra Grain Refining in Metastable Austenitic Stainless Steels, Proc. Innovation Stainless Steels, 1993, Florence, Italy, Vol. 2, p. 327. [44] S. Rajasekhara, P.J. Ferreira, L.P. Karjalainen, and A. Kyrlinen: Metall. Mater. Transactions A, 38A, (2007), No. 6, 1207. [45] K. Tomimura, S. Takaki and Y. Tokunaga: ISIJ International, 31 (1991), No.12, 1431. [46] K. Tomimura, S. Takaki, S. Tanimoto and Y.Tokunaga: ISIJ International, 31, (1991), No. 7, 721. [47] A. di Schino, M. Barteri, J.M. Kenny: J. Materials Science Letters, 21 (2002), 751. [48] M. C. Somani, L.P. Karjalainen, A. Kyrlinen, T. Taulavuori: Processing of submicron grained microstructures and enhanced mechanical properties by cold-rolling and reversion annealing of metastable austenitic stainless steels, Materials Science Forum, 539543 (2007), 4875. [49] M. Somani, P. Karjalainen, P. Juntunen, S. Rajasekhara, P. Ferreira, A. Kyrlinen, T. Taulavuori, P. Aspegren: Submicron microstructure and mechanical properties achieved in a short annealing of cold-rolled austenitic stainless steels, Iron & Steel, Supplement, 40 (2005), 283. [50] P. Juntunen, M. Somani, P. Karjalainen, D. Misra, A. Kyrlinen: Property enhancement in metastable austenitic stainless steels through the formation of ultra-fine grained austenite and shear reversion, Intern. Symp. on Advances on Stainless Steels, ISAS 2007, 9-11, April, 2007, The Indian Institute of Metals, CD. [51] A.S. Hamada, L.P. Karjalainen, M.C. Somani: Materials Science and Engineering A, 431 (2006), 211. [52] D.L. Johannsen: Development of Nano/Submicron Grain Sizes in an AISI 301Austenitic Stainless Steel, The University of Texas at Austin, Master's thesis, 2004. [53] F. Behr: Material-Wissenschaft und Werkstofftechnik, 30 (1999), No. 11, 669. [54] J. Talonen, P. Nenonen, H. Hnninen: Static strain ageing of coldworked austenitic stainless steels, High Nitrogen Steels 2004, Ostend, Belgium, 2004, p.113. [55] B. Snen, N. Akdut: Strengthening of austenitic stainless steels by static strain ageing, High Nitrogen Steels 2004, Ostend, Belgium, 2004, p.141. [56] R. Ruoppa, T. Taulavuori: Bake hardening of some austenitic stainless steels, Proc. of the 6th European Stainless Steel Conference - Science and Market, June 10-13, 2008, Helsinki (in press). [57] P. Antoine, B. Soenen, N. Akdut: Static strain ageing in cold rolled metastable austenitic stainless steels, Materials Science Forum, 539543, (2007), No. 5, 4891. [58] R.L.K. Ruoppa, T.T. Taulavuori: Influence of cold rolling reduction and strain ageing on forming properties of some austenitic stainless steels, International Deep-drawing Research Group Conference, Gyr, Hungary, 2007, pp. 579-586. [59] J. Talonen, H. Hnninen: Static strain ageing of metastable austenitic stainless steel, Proc. of the 6th European Stainless Steel Conference Science and Market, June 10-13, 2008, Helsinki (in press).

412

steel research int. 79 (2008) No. 6

You might also like