You are on page 1of 5

Materials Science and Engineering A 387389 (2004) 456460

Effect of duplex ferrite grain size distribution on local fracture stresses of Nb-microalloyed steels
S.J. Wu a,b, , C.L. Davis a
b a Department of Metallurgy and Materials, The University of Birmingham, Edgbaston, Birmingham, UK School of Materials Science and Engineering, Beijing University of Aeronautics and Astronautics, Beijing 100083, PR China

Received 25 August 2003; received in revised form 15 January 2004

Abstract Two commercial thermomechanically controlled rolled (TMCR) microalloyed steel plates have been used to investigate the relationship between the duplex ferrite grain size distribution and local fracture stresses. Statistical analyses of the grain size distributions were performed for the ne and coarse ferrite grains in the two steel plates. Microhardness values were measured for each grain size region and it was found that the ne grain areas have signicantly higher microhardness values than the coarse grain areas. Tensile and blunt-notch slow bend tests were carried out over a range of temperatures on samples from the two commercial TMCR steel plates. The local fracture stress ( F ) values were calculated and the results show that the F values are almost independent of temperature. The presence of a mixed grain size distribution results in signicant scatter in the local fracture stresses of the steels. The distribution of fracture stress values can be correlated to the coarse grain size distribution in the steels examined. 2004 Elsevier B.V. All rights reserved.
Keywords: Duplex grain size; TMCR; Fracture stress; Microalloyed steel

1. Introduction Thermomechanical controlled rolling (TMCR) is used to improve the mechanical properties of microalloyed steels, predominantly through grain size renement to give high toughness and strength levels. However, duplex grain size microstructures (i.e., mixed coarse and ne grain regions) have been observed for steels subjected to the TMCR process [14]. The presence of the mixed grain size distribution may affect the fracture mechanisms in the material (i.e., crack initiation and propagation) particularly for cleavage fracture, and result in signicant scatter in the mechanical properties of the steels. In order to design steels with optimum strength and toughness levels it is essential that the effects of the coarse grain regions on toughness are understood. Numerous experimental results have demonstrated that cleavage microcracks nucleate at grain boundary carbide particles due to the stress intensication caused by plastic deformation (dislocation pile-up) in the neighbouring ferrite
Corresponding author. Tel.: +44-121-4143265; fax: +44-121-4145232. E-mail address: s.wu.1@bham.ac.uk (S.J. Wu). 0921-5093/$ see front matter 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2004.01.144

grain [58]. However, whilst it is necessary, the nucleation of a microcrack is not sufcient for cleavage failure to occur. The cleavage microcrack must also be able to grow into the adjacent ferrite grain and then to propagate across the ferrite grain boundaries. Therefore, the dominant parameter for cleavage fracture of steels is the ferrite grain size [4]. Over the past 40 years, much research work has been carried out on the quantitative relationships between the microstructures and mechanical properties of steels. It is well-established that the yield strength ( y ), the critical local fracture stress ( F ) and the Charpy impact transition temperature (ITT) of 1/2 , where d is the mean steels are a linear function of (d) grain size of polygonal ferrite [911]. However, in the case of a duplex ferrite grain size distribution, such as those seen in some TMCR steels, the average grain size value does not properly represent the microstructure [1] and therefore will have little relevance to cleavage fracture. It has been reported that cleavage microcracks form preferentially in the largest polygonal ferrite grains [4], suggesting that the F and the ITT should correlate with some measure of the large grains rather than the mean value of the grain size distribution [4,1214]. The present paper examines the relationship between the local fracture stress and the coarse ferrite grain

S.J. Wu, C.L. Davis / Materials Science and Engineering A 387389 (2004) 456460 Table 1 The nominal chemical compositions (wt.%) of the commercial steel plates Materials Com-A Com-B C 0.11 0.11 Si 0.31 0.30 Mn 1.39 1.43 P 0.010 0.011 S 0.003 0.003 Cr 0.03 0.023 Ni 0.32 0.30 Al 0.037 0.039 Cu 0.033 0.013 Nb 0.024 0.040 N 0.006 0.005 Ti 0.002 0.003 V

457

0.045 0.063

bands as well as the grain size distribution within the coarse grain bands in two commercial TMCR steels.

2. Experimental The present work focused on two commercial microalloyed TMCR steel plates (Com-A and Com-B). The nominal chemical compositions of the steel plates are listed in Table 1. The original/nal gauges of the TMCR steel plates are 230 mm/40 mm for Com-A and 230 mm/50 mm for Com-B steels. Full details of the TMCR processing schedule are given elsewhere [15]. The steel plates were characterised on the section normal to the transverse direction (TD). Statistical analyses of the ferrite grain size distributions were carried out, using optical microscopy and image analysis software (ZEISS KS400 3.0), on a minimum of 1200 grains on samples ground, polished and etched with 2% Nital. The grain size is represented by the equivalent circular diameter (ECD) converted from the grain area. Vickers microhardness values (25 g) were measured, using a MVK-H1 hardness testing machine, within both ne grain bands and coarse grain regions on the whole section vertical to the TD of the two steel plates. Hounseld No.13 tensile specimens were tested between 80 and 196 C, using an Instron testing machine with a loading speed of 1 mm/min, to determine the tensile properties of the two steels. Blunt-notch four-point-bend specimens were chosen to determine the local fracture stress F . The dimensions of the blunt-notch bend specimens were 10 mm 10 mm 60 mm with a 45 notch of 3.3 mm depth and 0.2 mm root radius. The notches were cut along the section normal to TD. The blunt-notch four-point-bend tests were carried out between 160 and 196 C to ensure brittle fracture, using a 50 kN DMG testing machine with a loading rate of 0.5 mm/min. The stressstrain distribution ahead of the notch root has been analysed using the nite element method (FEM) [16] together with appropriate values of yield strength and strain-hardening rate of the steels.

studded in a matrix of desirable ne grains resulting in a duplex ferrite grain distribution. The duplex grain structure may stem from segregation of microalloying elements [1], especially niobium [17,18] during solidication, resulting in an inhomogeneous distribution of microalloy precipitates and consequently partial recrystallisation during the TMCR process. Statistical analysis of the ECD grain sizes measured from the rolling surface to the centre of the two commercial steel plates showed that the distributions are skewed towards larger grain sizes which indicates the existence of a significant number of coarse grains. The average grain size of Com-B steel (9.7 m) is greater than that of Com-A steel (8.3 m), which is due to differences in the TMCR processes, since Com-B steel contains more microalloying elements (Nb, Ti, V) and would therefore be expected to have a higher number density of grain boundary pinning precipitates and a smaller mean grain size [15]. Ferrite grains within the coarse grain patches are also analysed by setting a lower grain size threshold limit of 6 m. The grain size distributions are shown in Fig. 2 with a mean coarse grain size of 12.2 m for Com-A and 13.2 m for Com-B. The average areal proportion of the coarse grain patches is 38.7 and 47.1% for Com-A and Com-B, respectively. Therefore, there is a signicant statistical chance of a crack tip sampling a coarse grain patch. Vickers microhardness (25 g) values were measured in the coarse grain patches and ne grain bands and are displayed statistically in Fig. 3. It is obvious that the microhardness of the ne grains is much higher than that of the coarse grains. The signicant difference in microhardness of both the ne and coarse grain regions between Com-A and

3. Results and discussion 3.1. Microstructural characterisation The two TMCR steel plates exhibit a microstructure of ferrite/pearlite banded in the rolling direction (Fig. 1). Variations in ferrite grain sizes were frequently observed through the thickness, as shown in Fig. 1, with coarse grain patches

Fig. 1. Microstructure of the TMCR microalloyed steels showing coarse and ne grain sizes.

458

S.J. Wu, C.L. Davis / Materials Science and Engineering A 387389 (2004) 456460

Fig. 2. Equivalent ferrite grain size distributions of the grains measured within the coarse grain bands of the two commercial TMCR microalloyed steels.

Com-B steels is due to their different TMCR procedure [15], Com-A (nish rolling at 702 C) was rolled at lower temperatures than Com-B (nish rolling at 776 C) which may result in more dislocations being retained in the nal structure (i.e., not lost by recovery/recrystallisation) and/or the presence of ner precipitates, which will strengthen the material. 3.2. Local fracture stress ( F ) Critical local fracture stress ( F ) values for both commercial TMCR steels were determined using values of fracture load in combination with a 2D nite element analysis (FEA) of the stressstrain distribution ahead of the notch root under plane strain conditions [16]. The stressstrain distributions were calculated using the ABAQUS software code, based upon the true stresstrue strain curves of the steels. The results of the analyses were expressed in terms of the ratio of maximum principal stress to yield strength, mp / y , corresponding to the position ahead of the notch tip and the largest value of mp / y at each applied load normalised by the general yield load (Papp /PGY ). PGY is calculated based on the following formula [19]: y PGY = 1.155cF B(W a)2 (1) 2S
Fig. 4. Variation of the local fracture stresses with temperature (a) Com-A, and (b) Com-B steel.

where B is the specimen breadth, (W a) the specimen depth below the notch and S is the bend span, cF is the constraint factor taken as 1.175 for the two steels. Assuming that the largest maximum principal stress corresponding to the failure load of the specimen is the critical local fracture stress F , the F values can then be obtained from the failure loads of the blunt-notch four-point-bend specimens and the yield stress values determined through tensile tests at different temperatures. Fig. 4 shows the F values for the two commercial steels tested at different temperatures. It can be seen that the F values are almost independent of temperature and that large scatter in F values exists for both steels due to the duplexity of the ferrite grain sizes. The mean F value of 1749 MPa for Com-A is greater than that of Com-B (1666 MPa). This may result from the consistently smaller grain size (overall and coarse) and the smaller area fraction of the coarse grain bands in Com-A despite its higher hardness. 3.3. Effect of duplex grain structure on F The duplex ferrite grain structure and the resultant hardness duplexity are of great signicance in structural design and integrity assessment since some mechanical properties (such as Charpy impact energy [4] and fracture stress [12]) depend upon the area fraction of larger grains. The well-established empirical expressions for normalised steels may still be applicable to predict the tensile strengths of the

Fig. 3. Statistic analysis of the microhardness values obtained from both coarse and ne grain bands in the TMCR commercial steels.

S.J. Wu, C.L. Davis / Materials Science and Engineering A 387389 (2004) 456460

459

TMCR steels using its mean grain size because tensile tests average the microstructure (for ductile failure). The mean grain size for the duplex grains is, however, less meaningful to the prediction of the Charpy impact energy and the fracture stress due to the highly concentrated stress and strain ahead of the notch or fatigue-precrack tip. Failure of the notched or precracked specimens is dominated by the intensication of stress and strain within a small area, usually the plastic zone, in front of the notch or crack, which strongly limits the sampling of the microstructure. It has been shown [8,10,12] that the critical event for cleavage failure in a notched steel specimen is the propagation of a microcrack across the adjacent ferrite/ferrite grain boundary. This will occur when the stress level at the microcrack exceeds the ferrite grain strength which is given in terms of the grain diameter d by the equation [10]: F = Ep (1 2 )d
1/2

(2)

A value of p of 52 J/m2 is obtained for both TMCR steels which is in the range reported in the literature [20]. Using Eq. (2) and the obtained p value, the coarse ferrite grain size distributions in Fig. 2 can then be converted into local fracture stress distributions. The probability of cleavage failure occurring at a certain fracture stress value calculated from Eq. (2) corresponding to a particular grain size are shown in Fig. 5 together with the experimental results of the F values. The reason that there are no test data at the lower value side of the F distributions may be attributed to the limited number of tested specimens and the very low probability of sampling the extremely large ferrite grains due to their rarity. It can be seen that the experimental results t very well to the predicted F distributions obtained from Eq. (2) based upon the coarse ferrite grain distributions for the two TMCR steels. This implies that the fracture stress F of a TMCR microalloyed steel can be predicted based on the coarse grain size distributions within the coarse grain bands.

where E is Youngs modulus, is Poissons ratio, and p is the effective surface energy. Assuming that the mean local fracture stress of the TMCR steels is related to the mean ferrite grain size of the grains within the coarse grain patches (since cleavage microcracks will form preferentially in the large polygonal ferrite grains), the p value can then be determined, using Eq. (2) and = 0.3, E = 208 103 MPa.

4. Conclusions (1) Metallographic observation and measurement showed that TMCR steels have a duplex ferrite grain structure with coarse grain patches surrounded by ne grain bands. (2) The microhardness of the coarse grain patches is lower than that of the ne grain bands, the values are dependent upon the TMCR schedules for the two steels examined. (3) The local fracture stresses of the TMCR steels are almost independent of temperature within the range examined (196 to 160 C) and show a large scatter due to the wide ferrite grain size distributions. (4) The local fracture stress is related to the large ferrite grains within the coarse grain patches and can be predicted from the coarse ferrite grain distribution of the steels using the model described in this paper.

Acknowledgements The authors would like to thank the Engineering and Physical Sciences Research Council (EPSRC) for nancial support and Corus UK Ltd., for nancial support, provision of test material and data. Thanks are also due to Professor P Bowen for the provision of research facilities at the University of Birmingham.

References
[1] C.L. Davis, M. Strangwood, J. Mater. Sci. 37 (2002) 10831090. [2] M.J.W. Green, F.A.M. Maas, A.J. Rose, Y. Yang, in: Proceedings of the Challenges in Reheating Furnaces, IOM Communications, London, 2002, pp. 5564.

Fig. 5. Comparison of the local fracture stress distribution predicted from the coarse grain size distribution with the experimental data (a) Com-A, and (b) Com-B steel.

460

S.J. Wu, C.L. Davis / Materials Science and Engineering A 387389 (2004) 456460 [12] J.H. Chen, L. Zhu, H. Ma, Acta Metall. Mater. 38 (1990) 25272535. [13] J.H. Chen, G.Z. Wang, Z. Wang, L. Zhu, Y.Y. Gao, Met. Trans. 22 (1991) 22872296. [14] A. From, R. Sandstrom, Mater. Charact. 42 (1999) 111122. [15] S.J. Wu, C.L. Davis, J. Microsc. 213 (2004). [16] S.J. Wu, J.F. Knott, in: J. Edwards, et al. (Eds.), Proceedings of the Conference on Structural Integrity in the 21st Century, EMAS Publishing, 2000, pp. 247255. [17] B. Dutta, C.M. Sellars, Mater. Sci. Technol. 3 (1987) 197206. [18] A.C. Knessi, G. Posch, C.I. Garcia, A.J. Deardo, in: R. Asfahani, G. Tither (Eds.), Proceedings of the International Symposium on Low-carbon Steels for the 90s, TMS, Pittsburgh, USA, 1993, pp. 113119. [19] J.F. Knott, Fundamentals of Fracture Mechanics, Butterworths, London, 1973, 34 pp. [20] W.W. Gerberich, E. Kurman, Scripta Metall. 19 (1985) 295298.

[3] T. Tanaka, Int. Met. Rev. 4 (1981) 185212. [4] M.T. Shehata, J.D. Boyd, in: Proceedings of the Conference on Advances in Physical Metallurgy and Applications of Steels, University of Liverpool, TMS, 1982, pp. 229236. [5] E. Smith, in: Proceedings of the Conference on Physical Basis of Yield and Fracture, Institute of Physics, Oxford, UK, 1966, pp. 3645. [6] D. Curry, J.F. Knott, Met. Sci. 12 (1978) 511514. [7] D. Curry, J.F. Knott, Met. Sci. 13 (1979) 341345. [8] R.O. Ritchie, J.F. Knott, J.R. Rice, J. Mech. Phys. Solids 21 (1973) 395410. [9] B. Mintz, G. Peterson, A. Nassar, Ironmak. Steelmak. 21 (1994) 215222. [10] T. Lin, A.G. Evans, R.O. Ritchie, Met. Trans. 18 (1987) 641651. [11] T. Gladman, D. Dulieu, I.D. McIvor, Microalloying, vol. 75, Union Carbide Corp., New York, 1977, pp. 3255.

You might also like