Next Article in Journal
Effect of Austenite-to-Ferrite Phase Transformation at Grain Boundaries on PWHT Cracking Susceptibility in CGHAZ of T23 Steel
Next Article in Special Issue
Microstructure Evolution in Super Duplex Stainless Steels Containing σ-Phase Investigated at Low-Temperature Using In Situ SEM/EBSD Tensile Testing
Previous Article in Journal
Extraction of Vanadium from Ammonia Slag under Near-Atmospheric Conditions
Previous Article in Special Issue
Effect of the Martensitic Transformation on the Stamping Force and Cycle Time of Hot Stamping Parts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deformation Behavior of High-Mn TWIP Steels Processed by Warm-to-Hot Working

by
Vladimir Torganchuk
1,
Aleksandr M. Glezer
2,*,
Andrey Belyakov
1 and
Rustam Kaibyshev
1
1
Belgorod State University, Belgorod 308015, Russia
2
National University of Science & Technology (MISIS), Moscow 119049, Russia
*
Author to whom correspondence should be addressed.
Metals 2018, 8(6), 415; https://doi.org/10.3390/met8060415
Submission received: 11 May 2018 / Revised: 30 May 2018 / Accepted: 1 June 2018 / Published: 3 June 2018

Abstract

:
The deformation behavior of 18%Mn TWIP steels (upon tensile tests) subjected to warm-to-hot rolling was analyzed in terms of Ludwigson-type relationship, i.e., σ = K1·εn1 + exp(K2n2·ε). Parameters of Ki and ni depend on material and processing conditions and can be expressed by unique functions of inverse temperature. A decrease in the rolling temperature from 1373 K to 773 K results in a decrease in K1 concurrently with n1. Correspondingly, true stress approached a level of about 1750 MPa during tensile tests, irrespective of the previous warm-to-hot rolling conditions. On the other hand, an increase in both K2 and n2 with a decrease in the rolling temperature corresponds to an almost threefold increase in the yield strength and threefold shortening of the stage of transient plastic flow, which governs the duration of strain hardening and, therefore, manages plasticity. The change in deformation behavior with variation in the rolling temperature is associated with the effect of the processing conditions on the dislocation substructure, which, in turn, depends on the development of dynamic recovery and recrystallization during warm-to-hot rolling.

1. Introduction

High-manganese austenitic steels with low stacking fault energy (SFE) are currently considered as promising materials for various structural/engineering applications because of their outstanding mechanical properties [1,2,3]. Owing to their low SFE, these steels are highly susceptible to deformation twinning, which results in the twinning-induced plasticity (TWIP) effect. Austenitic TWIP steels are characterized by pronounced strain hardening, which retards the strain localization and cracking during plastic deformation and, therefore, provides a beneficial combination of high strength with high ductility [1]. Deformation twinning, therefore, is the most crucial deformation mechanism governing the mechanical properties of high-Mn TWIP austenitic steels [4]. The deformation twins appear as bundles of closely spaced twins with thickness of tens nanometers, crossing over the original grains [5,6]. The deformation twins prevent the dislocation motion and promote an increase in dislocation density, leading to strain hardening. Frequent deformation twinning develops in steels with SFE in the range of approx. 20 mJ/m2 to 50 mJ/m2, which can be adjusted by manganese and carbon content [4].
The exact values of mechanical properties of austenitic steels, e.g., yield strength, ultimate tensile strength, total elongation, etc., depends on processing conditions. Hot rolling is frequently used as a processing technology for various structural steels and alloys. Final mechanical properties of processed steels and alloys depend on their microstructures that develop during hot working. Metallic materials with low SFE like high-Mn austenitic steels experience discontinuous dynamic recrystallization (DRX) during hot plastic deformation [7]. The developed microstructures depend on the deformation temperature and/or strain rate. Namely, the DRX grain size decreases with a decrease in temperature and/or an increase in strain rate and can be expressed by a power law function of temperature compensated strain rate (Z) [8]. A decrease in the deformation temperature to warm deformation conditions results in a change in the DRX mechanism from discontinuous to continuous, leading to a decrease in the grain size exponent in the relationship between the grain size and the deformation conditions, although this relationship remains qualitatively the same as that for hot working conditions [9]. The grain refinement with an increase in Z is accompanied with an increase in the dislocation density in DRX microstructures, irrespective of the DRX mechanisms [10]. Thus, the yield strength of the warm to hot worked semi-products can be evaluated by using various structural parameters. This approach has been successfully applied for strength evaluation of a range of structural steels and alloys, including high-Mn TWIP steels subjected to various thermo-mechanical treatments [5,6,10,11,12]. On the other hand, the effect of processing conditions on the deformation behavior of high-Mn TWIP steels has not been qualitatively evaluated, although, particularly for these steels, the deformation behavior is one of the most important properties, which manages the practical applications of the steels. It should be noted that the stress-strain behavior of austenitic steels with low SFE cannot be described by any well elaborated models like Hollomon or Swift equations, especially, at relatively small strains because of exceptional strain hardening [13]. Ludwigson modified the Hollomon-type relationship with an additional term to compensate the large difference between experimental and predicted flow stresses at small strains for such metals and alloys [14]. In spite of certain achievements in the application of the Ludwigson-type equation for the stress-strain behavior prediction, the selection of suitable parameters in this equation is still arbitrary in many ways. The aim of the present study, therefore, is to obtain the relationships between the processing conditions, the developed microstructures, and the stress-strain equation parameters in order to predict the tensile deformation behavior of advanced high-Mn TWIP steels processed by warm-to-hot rolling.

2. Materials and Methods

Two steels, Fe-18%Mn-0.4%C and Fe-18%Mn-0.6%C, have been selected in the present study as typical representatives of high-Mn TWIP steels. The steel melts were annealed at 1423 K, followed by hot rolling with about 60% reduction. The steels were characterized by uniform microstructures consisting of equiaxed grains with average sizes of 60 μm (18Mn-0.4C) and 50 μm (18Mn-0.6C). The steels were subjected to plate rolling at various temperatures from 773 K to 1373 K to a total rolling reduction of 60%. After each 10% rolling reductions, the samples were re-heated to the designated rolling temperature. Structural investigations were carried out using a Quanta 600 scanning electron microscope (SEM), equipped with an electron backscattering diffraction (EBSD) analyzer incorporating orientation imaging microscopy (OIM). The SEM samples were electro-polished at a voltage of 20 V at room temperature using an electrolyte containing 10% perchloric acid and 90% acetic acid. The OIM images were subjected to a clean-up procedure, setting the minimal confidence index of 0.1. The tensile tests were carried out using Instron 5882 testing machine with tensile specimens with a gauge length of 12 mm and a cross section of 3 × 1.5 mm2 at an initial strain rate of 10−3 s−1. The tensile axis was parallel to the rolling axis.

3. Results and Discussion

3.1. Developed Microstructures

Typical deformation microstructures that develop in the high-Mn steels during warm to hot rolling are shown in Figure 1. The mechanisms of microstructure evolution operating in austenitic steels during warm to hot working and the developed microstructures have been considered in detail in previous studies [10]. The deformation microstructures in the present high-Mn steels subjected to warm to hot rolling at temperatures of 773–1323 K can be briefly characterized here as follows. The temperature range above 1073 K corresponds to hot deformation conditions. Therefore, the deformation microstructures evolved during deformation in this temperature range result from the development of discontinuous DRX. The uniform microstructures consisting of almost equiaxed grains with numerous annealing twins are clearly seen in the samples hot rolled at temperatures above 1073 K (Figure 1a,b). The transverse DRX grain size decreases from 50–80 μm to 5–10 μm with a decrease in the rolling temperature from 1323 K to 1073 K.
In contrast, DRX hardly develops during warm rolling at temperatures below 1073 K. The deformation microstructures composed of flattened original grains evolve during warm rolling (Figure 1c,d). It is worth noting that the transverse grain size in the deformation microstructures developed during warm rolling does not remarkably depend on the rolling temperature. Relatively low deformation temperature suppresses discontinuous DRX. In this case, the structural changes are controlled by dynamic recovery. Under conditions of warm working, continuous DRX, which is assisted by dynamic recovery, can be expected after sufficiently large strains [7]. The present steels, however, are characterized by low SFE of 20–30 mJ/m2 [4]. Such a low SFE makes the dislocation rearrangements during plastic deformation difficult and slows down the recovery kinetics. Therefore, 60% rolling reduction as applied in the present study is not enough for continuous DRX development in high-Mn steels. The final grain size, therefore, seems to be dependent on the original grain size and the total rolling reduction.

3.2. Mechanical Properties

The stress-elongation curves obtained by tensile tests of the high-Mn steels processed by warm-to-hot rolling at different temperatures in the range of 773–1373 K are shown in Figure 2. A decrease in the rolling temperature results commonly in an increase in the strength and a decrease in the plasticity. The effect of the rolling temperature on the tensile tests properties is more pronounced for the warm working domain, i.e., rolling at temperatures below 1073 K, than that for hot working conditions, i.e., rolling at temperatures above 1073 K. A decrease in the temperature from 1373 K to 1073 K results in an increase in the yield strength (σ0.2) by about 130 MPa while the ultimate tensile strength (UTS) does not change remarkably. In contrast, further decrease in the rolling temperature from 1073 K to 773 K leads to almost twofold increase in σ0.2, which approaches about 900 MPa, and increases UTS by about 200 MPa. Correspondingly, the strengthening by warm to hot rolling is accompanied by a degradation of plasticity. It is interesting to note that total elongation gradually decreases with a decrease in the rolling temperature for Fe-18%Mn-0.6%C steel, where as that in Fe-18%Mn-0.4%C steel exhibits a kind of bimodal temperature dependence. The total elongation in the Fe-18%Mn-0.4%C steel tends to saturate at a level of 60–65% as the rolling temperature increases above 1073 K, following a rapid increase from 30% to 55% with an increase in the rolling temperature from 773 K to 1073 K.
An apparent saturation for the total elongation of the Fe-18%Mn-0.4%C steel with an increase in the rolling temperature above 1073 K can be associated with a variation in the deformation mechanisms operating during tensile tests. The steel with lower carbon content has somewhat lower SFE [4] and, thus, may involve the strain-induced martensite upon tensile tests at room temperature. This difference in deformation mechanisms has been considered as a reason for the difference in plasticity [10]. The Fe-18%Mn-0.4%C steel subjected to hot rolling at temperatures above 1073 K exhibits the maximal plasticity, which can be obtained in the case of partial ε-martensitic transformation, whereas the Fe-18%Mn-0.6%C steel demonstrates increasing plasticity, which is improved by deformation twinning, with an increase in the rolling temperature. On the other hand, the strength properties, which depend on the grain size and dislocation density, are certainly affected by the rolling temperature, even in the range of hot working.
The Fe-18%Mn-0.6%C steel exhibits higher strength and elongation than the Fe-18%Mn-0.4%C steel after rolling warm to hot rolling in the studied temperature range. This additional strengthening of the Fe-18%Mn-0.6%C steel can be attributed to the difference in carbon content, which has been considered as the contributor to the yield strength of high-Mn TWIP steels [15]. The difference in 0.2 wt % carbon should result in about 85 MPa difference in the yield strength [15].

3.3. Tensile Behavior

Generally, the strength and plasticity during tensile tests depends on strain hardening, which, in turn, depends on the operating deformation mechanisms [1]. The plastic deformation of austenitic steels with low SFE at an ambient temperature is commonly expressed by the Ludwigson relation [14]:
σ = K1·εn1 + exp(K2n2·ε),
where the first term with the strength factor of K1 and strain hardening exponent n1 represents Hollomon equation and the second term has been introduced by Ludwigson to incorporate the transient deformation stage, which differentiates the deformation behavior of fcc-metals/alloys with low-to medium SFE from other materials at relatively small strains. The stress of σ = exp(K2) is close to the stress of plastic deformation onset and an inverse value of n2 corresponds to the transient stage duration.
The parameters of K1, K2, n1, n2 providing the best correspondence between Equation (1) and experimental stress-strain curves are listed in Table 1. Note here, similar values for parameters of Ludwigson relation have been reported in other studies on low SFE austenitic steels [14,16,17,18]. The larger values of K2 and K1 for the present 0.6%C steel as compared to those for the 0.4%C steel reflect the higher stress levels of the former at early deformations and at large tensile strains, respectively. On the other hand, the n1 values for both steels are close, suggesting similar strain hardening at large tensile strains, irrespective of some differences in the carbon content and SFE. The n2 values are also close for both steels, indicating the same effect of the rolling temperature on the transient deformation stage during subsequent tensile tests.
The monotonous changes of obtained parameters with rolling temperature suggest unique relationships between all parameters and deformation conditions. The effects of processing temperature on the parameters of Equation (1) are represented in Figure 3. Except for K1, all parameters can be expressed by unique linear functions of the inverse rolling temperatures (Figure 3). The bimodal temperature dependencies obtained for K1 with inflection points at 1073 K in Figure 3 are associated with the transition from warm to hot rolling conditions at this temperature, which reflects clearly on the deformation microstructures (see Figure 1). Using the indicated (Figure 3) linear relationships between the parameters of Ludwigson relation and the inverse rolling temperatures, the true stress vs strain curves calculated by Equation (1) are shown in Figure 4, along with the experimental curves obtained by tensile tests. Figure 4a,c show a general view of the stress-strain curves to validate the first term of Equation (1), whereas Figure 4b,d are plotted in double logarithmic scale to display the deformation behaviors at relatively small strains, which are described by the second term of Equation (1). The clear correspondence between the calculated and experimental plots testifies to the proposed treatments above.
The tensile deformation behavior of the steel samples should indeed be closely related to the steel microstructures, which were evolved by previous thermo-mechanical treatment. In turn, the developed microstructures depend on the processing conditions, i.e., rolling temperature, as the main processing variable in the present study. Generally, the deformation microstructures including the mean grain size and dislocation density that develop in metallic materials during warm-to-hot working can be expressed by power law functions of Zener-Hollomon parameter (temperature-compensated strain rate); Z = ε·exp(Q/RT), where Q and R are the activation energy and universal gas constant, respectively [7,12]. Such microstructural changes are associated with thermally activated mechanisms of microstructure evolution in metallic materials. Therefore, the unique linear relationships in Figure 3 between the parameters of the flow stress predicting equation and the inverse rolling temperature suggest exponential relationships between the flow stress and the microstructures developed by warm-to-hot rolling. The second (exponential) term in Equation (1) predicts the flow stresses at relatively small strains (transient deformation), when the deformation behavior is associated with the dislocation ability to planar glide [14]. Thus, the stress-strain relationship of the high-Mn TWIP steels depends on their microstructures, namely, dislocation densities, evolved by previous thermo-mechanical treatments. Similar conclusions about a dominant role of dislocation density in the yield strength [10] and the work-hardening rate [19] were drawn in other studies on TWIP steels.
It is worth noting in Figure 4 that maximal true stresses during tensile tests comprise about 1750 MPa for all steel samples, irrespective of the previous rolling conditions. Such gradual change in the altitude and slope of the true stress-strain curves can be represented by a gradual decrease in K1 concurrently with n1. An apparent saturation for the true stresses can be attributed to the strain-hardening ability owing to dislocation accumulation.
The strain, at which the transient deformation stage decays (εL) can be evaluated from the following relation [14]:
exp(K2n2·εL)/(K1εLn1) = r,
setting an arbitrary small value for r. According to the original Ludwigson treatment [14], r = 0.02 is selected in the present study. The values of εL calculated by Equation (2) for the present steels subjected to warm-to-hot rolling are shown in Figure 5 as functions of the rolling temperature. Formally, this strain (εL) limiting the transient deformation duration can be considered as a critical point below which, the plastic flow cannot be adequately described by Hollomon-type relation, i.e., the first term in Equation (1). The flow stresses during the transient deformation can be calculated taking into account the second term in Equation (1). The strain of εL can be roughly related to a strain when cross slip and dislocation rearrangements, which are closely connected with dynamic recovery, impair the strain hardening [14]. Therefore, an increase in εL should promote plasticity, including both uniform and total elongations.
It is clearly seen in Figure 5 that εL increases from about 0.08 to 0.24 with an increase in the rolling temperature from 773 K to 1373 K; this suggests an improvement in plasticity with increase in rolling temperature. A decrease in SFE promotes planar slip and, thus, should increase εL. Indeed, the 0.4%C steel is characterized by a larger εL than the 0.6%C steel after hot rolling at temperatures above 1300 K (Figure 5), although the hot rolled steel with higher carbon content exhibits larger total elongation. This relatively low plasticity of the Fe-18%Mn-0.4%C steel is associated with an ε-martensitic transformation [10]. Commonly, transformation-induced plasticity (TRIP) steels demonstrate lower plasticity than TWIP steels [20,21]. After rolling at temperatures below 1300 K, the values of εL for the Fe-18%Mn-0.4%C steel are smaller than those for the Fe-18%Mn-0.6%C steel processed under the same conditions (Figure 5). This can be attributed to the effect of warm-to-hot rolling at the evolved dislocation density. The latter has been shown to increase as rolling temperature decreases [10]. Therefore, the transient deformation stage upon the tensile tests is shortened because of the previous plastic deformation during warm-to-hot rolling, which partially consumed the dislocation ability to planar glide.

4. Conclusions

The deformation behavior during tensile tests of Fe-18%Mn-0.4%C and Fe-18%-0.6%C steels subjected to warm to hot rolling was studied. The main results can be summarized as follows.
  • The hot rolling at temperatures above 1073 K was accompanied by the development of discontinuous dynamic recrystallization, leading to a decrease in the transverse grain size with a decrease in the rolling temperature. On the other hand, microstructure evolution during warm rolling at temperatures below 1073 K was controlled by the rate of dynamic recovery, which slowed down with a decrease in the rolling temperature.
  • The true stress-strain curves obtained by tensile tests at ambient temperature can be correctly represented by the Ludwigson-type relationship—σ = K1·εn1 + exp(K2n2·ε)—where parameters of Ki and ni depended on material and processing conditions and can be expressed by unique functions of inverse temperature of previous warm-to-hot rolling. A decrease in the rolling temperature from 1373 K to 773 K resulted in a decrease in K1 (from approx. 2600 MPa to 2300 MPa) concurrently with n1 (from approx. 0.7 to 0.25). Correspondingly, the true stress approached a level of about 1750 MPa during tensile tests, irrespective of the previous warm-to-hot rolling conditions. On the other hand, an increase in both K2 and n2 with decrease in the rolling temperature corresponded to an almost threefold increase in the yield strength and analogous degradation of plasticity.
  • The stage of transient plastic flow providing initial strain hardening and, therefore, controlling the total plasticity increases from about 0.08 to 0.24 with an increase in the rolling temperature from 773 K to 1373 K. The Fe-18%Mn-0.4%C steel is characterized by smaller values of εL than the Fe-18%Mn-0.6%C steel subjected to warm-to-hot rolling at the same temperatures below 1300 K, although the former should possess lower stacking fault energy. The shortening of the transient deformation stage upon tensile tests of the steels subjected to warm-to-hot rolling can be attributed to previous deformation, which partially consumed the dislocation ability to planar glide.

Author Contributions

Conceptualization, A.M.G. and R.K.; Methodology, V.T. and A.B.; Investigation, V.T.; Writing-Original Draft Preparation, A.B.; Writing-Review & Editing, A.M.G. and R.K.; Visualization, V.T.; Supervision, R.K.

Funding

This work was performed partly under the State Order of the Ministry of Education and Science of the Russian Federation No. 20`7/113 (2097).

Acknowledgments

The authors are grateful to the personnel of the Joint Research Centre of “Technology and Materials”, Belgorod State University, for their assistance with instrumental analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bouaziz, O.; Allain, S.; Scott, C.P.; Cugy, P.; Barbier, D. High manganese austenitic twinning induced plasticity steels: A review of the microstructure properties relationships. Curr. Opin. Solid State Mater. Sci. 2011, 15, 141–168. [Google Scholar] [CrossRef]
  2. De Cooman, B.C.; Estrin, Y.; Kim, S.K. Twinning-induced plasticity (TWIP) steels. Acta Mater. 2018, 142, 283–362. [Google Scholar] [CrossRef]
  3. Kusakin, P.S.; Kaibyshev, R.O. High-Mn twinning-induced plasticity steels: Microstructure and mechanical properties. Rev. Adv. Mater. Sci. 2016, 44, 326–360. [Google Scholar]
  4. Saeed-Akbari, A.; Mosecker, L.; Schwedt, A.; Bleck, W. Characterization and prediction of flow behavior in high-manganese twinning induced plasticity steels: Part I. mechanism maps and work-hardening behavior. Metall. Mater. Trans. A 2012, 43, 1688–1704. [Google Scholar] [CrossRef]
  5. Kusakin, P.; Belyakov, A.; Haase, C.; Kaibyshev, R.; Molodov, D.A. Microstructure evolution and strengthening mechanisms of Fe–23Mn–0.3C–1.5Al TWIP steel during cold rolling. Mater. Sci. Eng. A 2014, 617, 52–60. [Google Scholar] [CrossRef] [Green Version]
  6. Yanushkevich, Z.; Belyakov, A.; Haase, C.; Molodov, D.A.; Kaibyshev, R. Structural/textural changes and strengthening of an advanced high-Mn steel subjected to cold rolling. Mater. Sci. Eng. A 2016, 651, 763–773. [Google Scholar] [CrossRef]
  7. Sakai, T.; Belyakov, A.; Kaibyshev, R.; Miura, H.; Jonas, J.J. Dynamic and post-dynamic recrystallization under hot, cold and severe plastic deformation conditions. Prog. Mater. Sci. 2014, 60, 130–207. [Google Scholar] [CrossRef]
  8. Tikhonova, M.; Belyakov, A.; Kaibyshev, R. Strain-induced grain evolution in an austenitic stainless steel under warm multiple forging. Mater. Sci. Eng. A 2013, 564, 413–422. [Google Scholar] [CrossRef]
  9. Tikhonova, M.; Enikeev, N.; Valiev, R.Z.; Belyakov, A.; Kaibyshev, R. Submicrocrystalline austenitic stainless steel processed by cold or warm high pressure torsion. Mater. Sci. Forum 2016, 838, 398–403. [Google Scholar] [CrossRef]
  10. Torganchuk, V.; Belyakov, A.; Kaibyshev, R. Effect of rolling temperature on microstructure and mechanical properties of 18%Mn TWIP/TRIP steels. Mater. Sci. Eng. A 2017, 708, 110–117. [Google Scholar] [CrossRef]
  11. Morozova, A.; Kaibyshev, R. Grain refinement and strengthening of a Cu–0.1Cr–0.06Zr alloy subjected to equal channel angular pressing. Philos. Mag. 2017, 97, 2053–2076. [Google Scholar] [CrossRef]
  12. Yanushkevich, Z.; Dobatkin, S.V.; Belyakov, A.; Kaibyshev, R. Hall-Petch relationship for austenitic stainless steels processed by large strain warm rolling. Acta Mater. 2017, 136, 39–48. [Google Scholar] [CrossRef]
  13. Choudhary, B.K.; Isaac Samuel, E.; Bhanu Sankara Rao, K.; Mannan, S.L. Tensile stress-strain and work hardening behaviour of 316LN austenitic stainless steel. Mater. Sci. Technol. 2001, 17, 223–231. [Google Scholar] [CrossRef]
  14. Ludwigson, D.C. Modified stress-strain relation for fcc metals and alloys. Metal. Trans. 1972, 2, 2825–2828. [Google Scholar] [CrossRef]
  15. Kusakin, P.; Belyakov, A.; Molodov, D.A.; Kaibyshev, R. On the effect of chemical composition on yield strength of TWIP steels. Mater. Sci. Eng. A 2017, 687, 82–84. [Google Scholar] [CrossRef]
  16. Mannan, S.L.; Samuel, K.G.; Rodriguez, P. Stress-strain relation for 316 stainless steel at 300 K. Scr. Metall. 1982, 16, 255–257. [Google Scholar] [CrossRef]
  17. Satyanarayana, D.V.V.; Malakondaiah, G.; Sarma, D.S. Analysis of flow behaviour of an aluminium containing austenitic steel. Mater. Sci. Eng. A 2007, 452, 244–253. [Google Scholar] [CrossRef]
  18. Milititsky, M.; De Wispelaere, N.; Petrov, R.; Ramos, J.E.; Reguly, A.; Hanninen, H. Characterization of the mechanical properties of low-nickel austenitic stainless steels. Mater. Sci. Eng. A 2008, 498, 289–295. [Google Scholar] [CrossRef]
  19. Liang, Z.Y.; Li, Y.Z.; Huang, M.X. The respective hardening contributions of dislocations and twins to the flow stress of a twinning-induced plasticity steel. Scr. Mater. 2016, 112, 28–31. [Google Scholar] [CrossRef]
  20. Song, W.; Ingendahl, T.; Bleck, W. Control of strain hardening behavior in high-Mn austenitic steels. Acta Metall. Sin. (Engl. Lett.) 2014, 27, 546–555. [Google Scholar] [CrossRef]
  21. Lee, Y.-K.; Han, J. Current opinion in medium manganese steel. Mater. Sci. Technol. 2015, 31, 843–856. [Google Scholar] [CrossRef]
Figure 1. Typical OIM (orientation imaging microscopy) micrographs for deformation microstructures evolved in the Fe-18%Mn-0.4%C steel during hot-to-warm rolling at 1273 K (a), 1173 K (b), 1073 K (c) and 973 K (d). Colored orientations are shown for the transverse direction (TD).
Figure 1. Typical OIM (orientation imaging microscopy) micrographs for deformation microstructures evolved in the Fe-18%Mn-0.4%C steel during hot-to-warm rolling at 1273 K (a), 1173 K (b), 1073 K (c) and 973 K (d). Colored orientations are shown for the transverse direction (TD).
Metals 08 00415 g001
Figure 2. Engineering stress vs elongation curves of the Fe-18%Mn-0.6%C (a) and Fe-18%Mn-0.4%C (b) steels subjected to rolling at the indicated temperatures.
Figure 2. Engineering stress vs elongation curves of the Fe-18%Mn-0.6%C (a) and Fe-18%Mn-0.4%C (b) steels subjected to rolling at the indicated temperatures.
Metals 08 00415 g002
Figure 3. Effect of the rolling temperatures on the parameters of Ludwigson equation, K1 (a), n1 (b), K2 (c), and n2 (d).
Figure 3. Effect of the rolling temperatures on the parameters of Ludwigson equation, K1 (a), n1 (b), K2 (c), and n2 (d).
Metals 08 00415 g003
Figure 4. True tensile stress vs strain plots for Fe-18%Mn-0.4%C steel (a,b) and Fe-18%Mn-0.6%C steel (c,d) subjected to warm-to-hot rolling at indicated temperatures. The stress-strain curves obtained by tensile tests are shown by thick gray-scaled lines and those calculated by Equation (1) are shown by dashed lines.
Figure 4. True tensile stress vs strain plots for Fe-18%Mn-0.4%C steel (a,b) and Fe-18%Mn-0.6%C steel (c,d) subjected to warm-to-hot rolling at indicated temperatures. The stress-strain curves obtained by tensile tests are shown by thick gray-scaled lines and those calculated by Equation (1) are shown by dashed lines.
Metals 08 00415 g004
Figure 5. Effect of the rolling temperature on the strain for transient deformation during tensile tests of the Fe-18%Mn-0.4%C and Fe-18%-0.6%C steels.
Figure 5. Effect of the rolling temperature on the strain for transient deformation during tensile tests of the Fe-18%Mn-0.4%C and Fe-18%-0.6%C steels.
Metals 08 00415 g005
Table 1. Parameters of the Ludwigson equation for the Fe-18%Mn-0.4%C and Fe-18%Mn-0.6%C steels processed by warm-to-hot rolling.
Table 1. Parameters of the Ludwigson equation for the Fe-18%Mn-0.4%C and Fe-18%Mn-0.6%C steels processed by warm-to-hot rolling.
SteelRolling Temperature (K)K1 (MPa)n1K2n2
Fe-18%Mn-0.4%C77322480.246.037.8
Fe-18%Mn-0.4%C87322600.356.230.8
Fe-18%Mn-0.4%C97323420.466.020.3
Fe-18%Mn-0.4%C107324100.485.825.2
Fe-18%Mn-0.4%C117325560.585.712.8
Fe-18%Mn-0.4%C127325750.635.511.4
Fe-18%Mn-0.4%C137325980.675.512.3
Fe-18%Mn-0.6%C77323680.256.233.5
Fe-18%Mn-0.6%C87324100.386.324.5
Fe-18%Mn-0.6%C97325540.496.220.7
Fe-18%Mn-0.6%C107326000.556.120.7
Fe-18%Mn-0.6%C117326080.645.917.5
Fe-18%Mn-0.6%C127326080.685.916.0
Fe-18%Mn-0.6%C137326170.735.812.5

Share and Cite

MDPI and ACS Style

Torganchuk, V.; Glezer, A.M.; Belyakov, A.; Kaibyshev, R. Deformation Behavior of High-Mn TWIP Steels Processed by Warm-to-Hot Working. Metals 2018, 8, 415. https://doi.org/10.3390/met8060415

AMA Style

Torganchuk V, Glezer AM, Belyakov A, Kaibyshev R. Deformation Behavior of High-Mn TWIP Steels Processed by Warm-to-Hot Working. Metals. 2018; 8(6):415. https://doi.org/10.3390/met8060415

Chicago/Turabian Style

Torganchuk, Vladimir, Aleksandr M. Glezer, Andrey Belyakov, and Rustam Kaibyshev. 2018. "Deformation Behavior of High-Mn TWIP Steels Processed by Warm-to-Hot Working" Metals 8, no. 6: 415. https://doi.org/10.3390/met8060415

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop