You are on page 1of 5

International Journal of Fatigue 27 (2005) 14201424 www.elsevier.

com/locate/ijfatigue

Dynamic strain aging effect on the fatigue resistance of type 316L stainless steel
Seong-Gu Hong, Keum-Oh Lee, Soon-Bok Lee*
Department of Mechanical Engineering, Korea Advanced Institute of Science and Technology, 373-1 Guseong-dong, Yuseong-gu, Daejeon 305-701, South Korea Available online 18 August 2005

Abstract Mechanism of dynamic strain aging (DSA) and its effect on the high-temperature low-cycle fatigue resistance in type 316L stainless steel were investigated by carrying out low-cycle fatigue tests in a wide temperature range from 20 to 650 8C with strain rates of 3.2!10K51! 10K2/s. The regime of DSA was evaluated using the anomalous features of material behavior associated with DSA. The activation energies for each type of serration were about 0.570.74 times those for lattice diffusion indicating that a mechanism other than lattice diffusion is involved. It is reasonably concluded that the pipe diffusion of solute atoms along the dislocation core is responsible for DSA. Dynamic strain aging reduced the fatigue resistance by ways of multiple crack initiation, which comes from the DSA-induced inhomogeneity of deformation, and rapid crack propagation due to the DSA-induced hardening. q 2005 Elsevier Ltd. All rights reserved.
Keywords: Dynamic strain aging (DSA); Pipe diffusion; Planar slip; Low-cycle fatigue (LCF); 316L stainless steel

1. Introduction Type 316L stainless steel is a prospective material for the reactor vessels and piping systems in liquid metal cooled fast breeder reactor (LMFBR) whose integrity governs the safety of nuclear power plants. In these applications, the components are exposed to severe conditions, such as high temperature ranging from 300 to 600 8C, and undergo temperature-gradient induced cyclic thermal stresses as a result of start-ups and shut-downs. Therefore, low-cycle fatigue (LCF) is recognized as one of the main degradation mechanisms affecting the reactor vessels and piping systems integrity of LMFBR [14]. It has been reported that, in the temperature range of 300600 8C where LMFBR operates, type 316L stainless steel is susceptible to dynamic strain aging (DSA) which may induce a signicant change in material properties such as strength and ductility [3,4]. Dynamic strain aging is the phenomenon of interactions between diffusing solute atoms
* Corresponding author. Tel.: C82 42 869 3029; fax: C82 42 869 3210. E-mail address: sblee@kaist.ac.kr (S.-B. Lee).

0142-1123/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijfatigue.2005.06.037

and mobile dislocations during plastic deformation and depends on the deformation rate and temperature, which govern the velocities of mobile dislocations and diffusing solute atoms, respectively. At present, some research has been conducted on DSA, and it is well established for monotonic tensile loading. However, few systematic investigations have been reported about the inuence of DSA on LCF behavior, and the results are contradictory [13,5,6]. Abdel-Raouf et al. [5] examined the elevated temperature fatigue deformation characterization of Ferrovac E iron containing 0.007% carbon and reported that DSA causes a reduction in fatigue life. In contrast, in 2 1/4Cr1Mo steel [6], an increase in fatigue life was observed in the temperature range of 427600 8C due to an interactive solid solution hardening resulting from cyclic stress enhanced precipitation of Mo2C. Therefore, to apply type 316L stainless steel to LMFBR applications, more detailed research on the inuence of DSA on LCF behavior is required. In this study, the mechanism of DSA was identied by evaluating the activation energy of serrated yielding and the role of DSA on the change of fatigue resistance at elevated temperature was investigated regarding the effect of DSA on crack initiation and propagation.

S.-G. Hong et al. / International Journal of Fatigue 27 (2005) 14201424

1421

2. Experiment 2.1. Material and specimen The material used in this study was cold-worked (CW) 316L stainless steel having the following chemical composition in wt%: C-0.025, Si-0.41, Mn-1.41, P-0.025, S-0.025, Ni-10.22, Cr-16.16, Mo-2.09, N-0.043, Fe-balance. The asreceived material was manufactured by the following processes: the material was solution-treated at 1100 8C for 40 min, then water-quenched, and, nally, a round bar with a diameter of 16 mm was made through the cold-drawing, which introduced a tensile pre-strain of 17%. This treatment yielded an average intercept grain size of 44.2 mm. The asreceived material was fabricated into dog-bone type specimens with a gauge length of 36 mm and a gauge diameter of 8 mm in accordance with ASTM standard E606-92. 2.2. Test equipment and procedure A closed-loop servo-hydraulic test system with 5-ton capacity was used to perform LCF tests, and a 3-zone resistance type furnace which can control temperature within a variation of G1 8C at steady state was used for temperature control. A high temperature extensometer (MTS model no.: 632-13F-20) was used to measure a strain and control a strain signal. LCF tests were carried out in air under a fully reversed total axial strain control mode with a triangular waveform at room temperature (RT), 200, 400, 550, 600, and 650 8C. A strain amplitude was xed at 0.5% and strain rates varied from 3.2!10K5 to 1!10K2/s.

dragged by dislocations moving at less than a critical velocity. This model gives the relationship between temperature and strain rate for the onset of serrated yielding as following; _Z 3   4brC v D 0 K Q exp kT l (1)

3. Results and discussion 3.1. The regime of DSA The anomalies associated with DSA during LCF deformation in type 316L stainless steel, the same material in the present study, have been investigated in our previous study [4]. The results revealed that DSA under LCF loading is manifested in the forms of the negative temperature dependence of cyclic peak stress, the negative temperature dependence of plastic strain amplitude or softening ratio, the negative strain rate sensitivity, and the negative strain rate dependence of plastic strain amplitude or softening ratio. Based on those results, the regime of DSA was in the temperature range of 250550 8C at a strain rate of 1!10K4/s, in 250600 8C at 1!10K3/s, and in 250650 8C at 1!10K2/s. 3.2. The mechanism of DSA Cottrell [7] suggested the DSA model for the onset of serrated yielding assuming that solute atmospheres are

where b, r, Cv, D0, l, Q, and k represent the Burgers vector, the dislocation density, the vacancy concentration, the diffusion frequency factor, the effective radius of the atmosphere, the activation energy for solute migration, and the Boltzmann constant, respectively. It was reported that the regime of DSA under LCF loading is consistent with that in tensile loading [4]. Hence, it is thought to be reasonable to use the results of tensile tests in order to evaluate the activation energy for DSA under LCF loading. Typical segments of the stressstrain curves from tensile tests are shown in Fig. 1. The different types of serrations, which follows the generally accepted nomenclature in Ref. [8], appeared in the stressstrain curves with temperature. At a given strain rate, type D serration was observed at low temperatures, type ACBCE serrations, which indicate that type A, B, and E serrations appear simultaneously in the stressstrain curve, at intermediate temperatures, and type ACBCCCE serrations at high temperatures. The condition for the occurrence of serrated yielding is presented in Fig. 2. Following Eq. (1), the activation energy for the onset of serrated yielding can be obtained from the slope of the boundaries delineating _ K1=T plot (Fig. 2), different serrated ow regimes in a ln 3 as QZslope!k. The calculated values are 75.3 kJ/mol for type D serration (QD), 182.2 kJ/mol for type ACBCE serrations (QABE), and 205.3 kJ/mol for type ACBCCCE serrations (QABCE). Three different activation energy values in the temperature range of serrated ow indicate that serrated yielding in CW 316L stainless steel is attributed to more than one mechanism. Samuel et al. [9] have reported Q values, obtained using Eq. (1), in type 316 stainless steel. Q values are 133 kJ/mol
550 Engineering stress (MPa)
CW 316L SS Strain rate=1*104/s

200oC D

500 400 C 450 C A B


o

300o C o 250oC 450oC 500 C 550oC 600oC E

400

350 0 2 4 6 Engineering strain (%) 8

650 C 10

Fig. 1. Segments of the stress-strain curves from tensile tests.

1422

S.-G. Hong et al. / International Journal of Fatigue 27 (2005) 14201424

CW 316L SS 0.01 Strain rate (s1)


No Serration Type D Serration Type A+B+E Serrations Type A+B+C+E Serrations

1E-3

Qdisappear QABE 1E-4 QABCE QD

QD = 75.3 kJ/mol QABE = 182.2 kJ/mol QABCE = 205.3 kJ/mol Qdisappear = 227.3 kJ/mol

0.0010 0.0015 0.0020 0.0025 0.0030 0.0035 (12.7oC) (727oC) 1/Temperature, K1


Fig. 2. Regimes of occurrence of serrated yielding under tensile loading.

at low temperatures (250350 8C) and 278 kJ/mol at high temperatures (400650 8C). Diffusion of interstitial solutes (C or N) to dislocations and diffusion of substitutional solutes like Cr are considered as the mechanisms responsible for serrated ow in the low and high temperature regimes, respectively. A similar result has been also reported in type 316 stainless steel [10]. However, it is noticed that those values from the literature are much larger than the observed values in the present study. The lower observed values of activation energy for DSA have been attributed to strain-induced diffusion [11] or pipe diffusion along the dislocation core [12]. Cuddy and Lesile [12] have suggested that substitutional solutes can diffuse rapidly along the dislocation core with the activation energy which may be 0.40.7 times the activation energy for bulk diffusion in a-Fe. The calculated values of activation energy in the present study are about 0.570.74 times the activation energy for lattice QABE z 0:66Qbulk and diffusion: QD z 0:57Qbulk C or N , Cr , bulk QABCE z 0:74QCr . Therefore, it is thought to be reasonable to infer that serrated yielding in CW 316L stainless steel is caused by the pipe diffusion of solute atoms along the dislocation core: type D serration results from the pipe diffusion of interstitial solutes, such as C or N, and the pipe diffusion of substitutional Cr is responsible for type ACBC E or ACBCCCE serrations. 3.3. The mechanism of plastic deformation As shown in Fig. 3, the dislocation structure changed from a cellular structure somewhat elongated at temperatures below 250 8C to a planar structure in the temperature range of 250600 8C (DSA regime), and back to a (equiaxial) cell at high strain rates or a subgrain structure at low strain rates beyond 600 8C. This fact indicates that the mechanism of plastic deformation changes from wavy slip mode in the non-DSA regime to planar slip mode in the regime of DSA.

Fig. 3. TEM micrographs depicting the substructures at D3t/2Z0.5%: (a) _ Z 1 ! 10K3 =s, (b) 400 8C, 3 _ Z 1 ! 10K4 =s, and (c) 650 8C, RT, 3 _ Z 3:2 ! 10K5 =s. 3

3.4. The effect of DSA on the fatigue resistance Fatigue life is deed as the number of cycles at which the peak tensile stress drops by 30% from the half-life stress. The variations of fatigue life with temperature and strain rate are presented in Fig. 4. At a given strain rate, a factor of 24 reduction of fatigue life was observed with an increase in temperature, and this tendency became more pronounced with an increase in strain rate (Fig. 4(a)). As shown in Fig. 4(b), fatigue life reduced with a decrease in strain rate at a given temperature, and the strain rate dependence of

S.-G. Hong et al. / International Journal of Fatigue 27 (2005) 14201424

1423

(a) 10000 Total strain amplitude = 0.5% Number of cycles to failure, Nf 8000
1*10 /s 1*10 /s
4 2

1*10 /s 3.2*10 /s
5

6000

4000

2000 X : DSA regime 0 100 0 100 200 300 400 500 Temperature (oC) 600 700

(b) 3500 Number of cycles to failure, Nf 3000 2500 2000 1500 1000 500 X : DSA regime 0 10
4 3

Total strain amplitude = 0.5% 400oC 550oC 600oC 650oC

10

102

1 Strain rate (s )

Fig. 4. The variations of fatigue life with (a) temperature and (b) strain rate.

fatigue life was accelerated as temperature decreased in the temperature range of 400650 8C. In austenitic stainless steels [13], a reduction of fatigue resistance has been reported at elevated temperature, and it is attributed to creep, oxidation, and DSA effects or interaction among these factors. However, it should be noted that the effects of creep and oxidation become more pronounced as temperature increases or strain rate decreases. This will provide the opposite prediction, compared with the present results. Hence, creep and oxidation effects were not enough to account for the variation of the fatigue resistance with temperature and strain rate, observed in the present study. This fact could be also conrmed by observing the fracture surfaces of LCF failed specimens with scanning electron microscope. In the regime of DSA, there were no remarkable evidences of creep or oxidation (Fig. 5(b) and (c)): fatigue cracks initiated at the slip bands connected to the surface. Initial propagation occurred along slip planes oriented at 458 to the applied stress axis designated as stage I. The initial propagation was conned to the active slip planes as evidenced by the cleavage facet in Fig. 5(c). This type of cracking

Fig. 5. SEM micrographs showing the fracture surfaces at D3t/2Z0.5%: (a) _ Z 1 ! 10K3 =s, (b) and (c) 550 8C, 3 _ Z 1 ! 10K2 =s [3]. RT, 3

continued through only a few grains before the transition to stage II transgranular cracking, evidenced by striations on the fracture surface, occurred. That is, there was no evidence of intergranular damage or cracking which could be induced by creep. Oxidation became important only at high temperatures above 600 8C with low strain rates below 1!10K4/s where DSA disappeared. Therefore, the temperature and strain rate dependence of fatigue resistance could be explained by the DSA effect. When the dynamic strain aging operates, the planar dislocation structure is developed, as shown in Fig. 3(b), and, thus, plastic deformation is localized on the slip bands. This localized deformation on slip bands, if connected to the surface, could serve as crack initiation sites, which were

1424

S.-G. Hong et al. / International Journal of Fatigue 27 (2005) 14201424

evidenced by the cleavage facet shown in Fig. 5(c), and resulted in the multiplication of crack initiation site. This mechanism reduces the crack initiation life. The crack propagation rate can be described by Eq. (2) [13,14]: dl f Dsa lb dN (2)

Acknowledgements This work was supported by Computer Aided Reliability Evaluation (CARE) National Research Laboratory in Korea Advanced Institute of Science and Technology (KAIST) and the Korea Research Foundation Grant funded by Korea Government (MOEHRD, Basic Research Promotion Fund) (No. M01-2005-000-10267-0).

where l, a, and b are the crack length and material constants, respectively. For a given strain amplitude, the crack propagation rate will increase with an increase in the developed stress. As mentioned in Section 3.1, DSA induces the hardening of the material, characterized by the negative temperature dependence of ow stress and the negative strain rate sensitivity. This results in the development of a higher stress with an increase in temperature or a decrease in strain rate, at a given strain. Therefore, a higher stress concentration at the crack tip, in the regime of DSA, will enhance the crack propagation rate and lead to the reduction in crack propagation life.

References
[1] Srinivasan VS, Sandhya R, Bhanu Sankara Rao K, Mannan SL, Raghavan KS. Effect of temperature on the low cycle fatigue behavior of nitrogen alloyed type 316L stainless steel. Int J Fatigue 1991;13(6): 4718. [2] Srinivasan VS, Valsan M, Sandhya R, Bhanu Sankara Rao K, Mannan SL, Sastry DH. High temperature time-dependent low cycle fatigue behavior of a type 316L(N) stainless steel. Int J Fatigue 1999; 21:1121. [3] Hong SG, Lee SB. The tensile and low-cycle fatigue behavior of cold worked 316L stainless steel: inuence of dynamic strain aging. Int J Fatigue 2004;26(8):889910. [4] Hong SG, Lee SB. Dynamic strain aging under tensile and LCF loading conditions, and their comparison in cold worked 316L stainless steel. J Nucl Mater 2004;328(23):23242. [5] Abdel-Raouf H, Plumtree A, Topper TH. Effects of temperature and deformation rate on cyclic strength and fracture of low carbon steel. ASTM STP 1973;519:2857. [6] Challenger KD, Miller AK, Brinkman CR. An explanation for the effects of hold periods on the elevated temperature fatigue behavior of 2 1/4 Cr1Mo steel. Trans ASME J Eng Mater Tech 1981;103:714. [7] Cottrell AH. Dislocations and plastic ow in crystals. London: Oxford University Press; 1953. [8] Rodriguez P. In: Encyclopedia of materials science and engineering, vol. 1. New York: Pergamon; 1988. p. 504508, Suppl. 5. [9] Samuel KG, Mannan SL, Rodriguez P. Serrated yielding in AISI 316 stainless steel. Acta Metall 1988;36(8):23237. [10] Mannan SL, Samuel KG, Rodriguez P. Trans Ind Inst Metals 1983;36: 313. [11] Baluf RW. Phys Stat Solidi 1970;42:11. [12] Cuddy LJ, Leslie WC. Some aspects of serrated yielding in substitutional solid solutions of iron. Acta Metall 1972;20:115767. [13] Yamaguchi K, Kanazawa K, Yoshida S. Crack propagation in lowcycle fatigue of type 316 stainless steel at temperatures below 600 8C observed by scanning electron microscopy. Mater Sci Eng 1978;33: 17581. [14] Tomkins B. Fatigue crack propagationan analysis. Philos Mag 1968;18:104166.

4. Conclusions (1) The regime of DSA under LCF loading was in the temperature range of 250550 8C at a strain rate of 1! 10K4/s, in 250600 8C at 1!10K3/s, and in 250650 8C at 1!10K2/s. (2) It seems to be that, at low temperatures where type D serration appears, DSA occurs by the pinning of dislocations through the pipe diffusion of interstitial atoms, such as C or N, along the dislocation core, and, at high temperatures where type ACBCE or type AC BCCCE serrations appear, the pipe diffusion of substitutional Cr along the dislocation core is responsible for DSA. (3) Dynamic strain aging reduced the crack initiation and propagation life by ways of multiple crack initiation, which comes from the DSA-induced inhomogeneity of deformation, and rapid crack propagation due to the DSA-induced hardening, respectively.

You might also like