10
Fatigue 2010 Microcrack initiation in low-cycle fatigue of an austenitic stainless steel P. Mu 1,2,3* , V. Aubin 1,2,3 1 Univ Lille Nord de France, F-59000 Lille, France 2 ECLille, LML, F-59650 Villeneuve d’Ascq, France 3 CNRS, UMR 8107, F-59650 Villeneuve d’Ascq, France Received 9 March 2010; revised 11 March 2010; accepted 15 March 2010 Abstract The present study aims at analyzing the crack initiation in an austenitic stainless steel in low-cycle fatigue. A fatigue test was carried out using a polished specimen. The surface of the specimen was observed in–situ during the fatigue test, in order to establish the time of slip activity or crack initiation. After a number of cycles sufficient to initiate small crack, the test was stopped and the surface observed in Scanning Electron Microscope. Electron backscattered diffraction (EBSD) was used to identify the orientations of surface grains in the central zone of the fatigue specimen. Crack-initiation sites and slip systems associated to the initiated microcracks have been identified. Keywords: Low-cycle fatigue, austenitic stainless steel, in-situ observation, EBSD, crack initiation 1. Introduction Austenitic stainless steels combine good mechanical properties to an excellent corrosion resistance. This combination made austenitic stainless steels the most widely used type of stainless steels. In these applications they are subjected to cyclic loadings which may induce the initiation of several cracks and their propagation. In the high- fatigue domain, plastic strain irreversibilities are localized in persistent slip bands and cracks initiate in the PSBs [1], [2]. In the low-cycle fatigue domain, the fatigue life is usually empirically described by a Manson-Coffin curve. However, the life of a component is the sum of the time needed for crack initiation and the time taken for these cracks to propagate until the failure of the component. In low-cycle fatigue, cracks initiate as early as 50 cycles, they afterwards propagate but their growth rate decreases rapidly because of interactions between crack and grain boundaries. Only few cracks will become sufficiently long to be dominant and lead to fracture. The present study aims at analyzing the crack initiation in an austenitic stainless steel in low-cycle fatigue. A specific device was used to observe the surface of the specimen in –situ during the fatigue test. Electron Back Scatter Diffraction was used after the test to establish grain orientation of the surface zone analyzed. The sites for * Corresponding author. E-mail address: [email protected] c 2010 Published by Elsevier Ltd. Procedia Engineering 2 (2010) 1951–1960 www.elsevier.com/locate/procedia 1877-7058 c 2010 Published by Elsevier Ltd. doi:10.1016/j.proeng.2010.03.210 Open access under CC BY-NC-ND license. Open access under CC BY-NC-ND license.

Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

Available online at www.sciencedirect.com

Procedia Engineering 00 (2009) 000–000

ProcediaEngineering

www.elsevier.com/locate/procedia

Fatigue 2010

Microcrack initiation in low-cycle fatigueof an austenitic stainless steel

P. Mu1,2,3*, V. Aubin1,2,3

1Univ Lille Nord de France, F-59000 Lille, France 2ECLille, LML, F-59650 Villeneuve d’Ascq, France

3CNRS, UMR 8107, F-59650 Villeneuve d’Ascq, France

Received 9 March 2010; revised 11 March 2010; accepted 15 March 2010

Abstract

The present study aims at analyzing the crack initiation in an austenitic stainless steel in low-cycle fatigue. A fatigue test wascarried out using a polished specimen. The surface of the specimen was observed in–situ during the fatigue test, in order to establish the time of slip activity or crack initiation. After a number of cycles sufficient to initiate small crack, the test wasstopped and the surface observed in Scanning Electron Microscope. Electron backscattered diffraction (EBSD) was used to identify the orientations of surface grains in the central zone of the fatigue specimen. Crack-initiation sites and slip systemsassociated to the initiated microcracks have been identified.

Keywords: Low-cycle fatigue, austenitic stainless steel, in-situ observation, EBSD, crack initiation

1. Introduction

Austenitic stainless steels combine good mechanical properties to an excellent corrosion resistance. This combination made austenitic stainless steels the most widely used type of stainless steels. In these applications they are subjected to cyclic loadings which may induce the initiation of several cracks and their propagation. In the high-fatigue domain, plastic strain irreversibilities are localized in persistent slip bands and cracks initiate in the PSBs [1], [2]. In the low-cycle fatigue domain, the fatigue life is usually empirically described by a Manson-Coffin curve. However, the life of a component is the sum of the time needed for crack initiation and the time taken for these cracks to propagate until the failure of the component. In low-cycle fatigue, cracks initiate as early as 50 cycles, they afterwards propagate but their growth rate decreases rapidly because of interactions between crack and grain boundaries. Only few cracks will become sufficiently long to be dominant and lead to fracture.

The present study aims at analyzing the crack initiation in an austenitic stainless steel in low-cycle fatigue. A specific device was used to observe the surface of the specimen in –situ during the fatigue test. Electron Back Scatter Diffraction was used after the test to establish grain orientation of the surface zone analyzed. The sites for

* Corresponding author. E-mail address: [email protected]

c© 2010 Published by Elsevier Ltd.

Procedia Engineering 2 (2010) 1951–1960

www.elsevier.com/locate/procedia

1877-7058 c© 2010 Published by Elsevier Ltd.doi:10.1016/j.proeng.2010.03.210

Open access under CC BY-NC-ND license.

Open access under CC BY-NC-ND license.

Page 2: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

2 P. Mu et al. / Procedia Engineering 00 (2010) 000–000

crack initiation have been identified, and the orientations of slip systems associated to microcracks have been analyzed.

2. Material and specimen preparation

The material studied in this work is an AISI 316L austenitic stainless steel. Its chemical composition is given in Tab. 1. The material was supplied in a rolled plate of 14 mm in thickness. The material was solution treated for one hour at 1050 °C and then water quenched. All alloying elements were thus put in solid solution. The studied material has an average grain size, determined by the linear intercept method, of approximately 60 m, see Fig. 1.

Tab. 1. Composition of the AISI 316L stainless steel used in this study (wt.%)

C Cr Ni Mo Mn Si P Fe

0.030 18.51 12.85 2.15 1.50 0.71 0.045 64.30

Fig. 1 Microstructure of AISI 316L stainless steel observed after chemical etching

3. Mechanical test

In order to analyze crack initiation in low-cycle fatigue, a fatigue test was carried out on a servohydraulic tension-compression testing machine (Fig. 2). A plate specimen was machined from the as received sheet, following the sketch given in Fig. 3. The specimen had a useful part of 23.72 mm in length, 6 mm in width, and 3.6 mm in depth. The specimen was mechanically polished down to 1 m grade diamond paste. In order to reveal grain boundaries, the specimen was then chemical etched in a solution composed of 3 vol HCl, 1 vol HNO3 for 40s. The time for chemical etching should be carefully controlled in order that the grain boundaries or twin boundaries are not too much attacked, which may be the preferential crack initiation sites.

The test was controlled in longitudinal strain measured by an extensometer with a gauge length of 20 mm. The cyclic test was carried out under total strain control. The waveform of the fatigue cycle was symmetrical triangular with a strain rate of 6.6×10-4 s-1 and strain amplitude of 0.5%. Load and strain were recorded simultaneously during the cycling. The specimen was loaded/unloaded until 2000 cycles. The test was paused every 100 cycles (200 cycles after 1400 cycles) in order to take images of the surface of the specimen. As it has been showed by [3] under the same strain amplitude, the crack density in austenitic stainless steels increases until about 1500 cycles and stabilizes. Observations after 2000 cycles should then allow to highlight all the crack nucleation sites.

1952 P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960

Page 3: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

P. Mu et al. / Procedia Engineering 00 (2010) 000–000 3

Fig. 2 (a) Experimental equipment; (b) Zoom on the specimen and the objective zoom.

Fig. 3 Specimen drawing

The cyclic hardening/ softening curve obtained during the test is shown in Fig. 4. The material firstly shows a cyclic hardening during the first cycles and then softens until a quasi-stabilization. In AISI 316L, at room temperature, the stabilization stage covers generally the main part of the fatigue life [4] [5] [6].

P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960 1953

Page 4: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

4 P. Mu et al. / Procedia Engineering 00 (2010) 000–000

Stress amplitude (MPa)

200220240260280300320340360380

0 500 1000 1500 2000

Number of cycles Fig. 4 Cyclic hardening/softening curve obtained during the test presented here.

4. Observation of the specimen surface

In order to follow the surface damage evolution during the low-cycle fatigue test in real time and to acquire images at the microstructural scale, an in-situ microscopy device was used. This device consists of a 10-bit charge-coupled device (CCD) camera (1040×518 pixels) equipped with coaxial objective and zoom allowing to observe a zone of 140 ×70 m2 at the objective magnification of 1.5 (see Fig. 2). This zone was enough small to reveal the fatigue damage at a microstructural scale, and enough large to see several grains in one image. This zone also ensured an acceptable quality of the images acquired, as the lighting is relatively homogeneous. The in-situ observation device was mounted on the frame of the testing machine (Fig. 2(a)). Two displacement tables permitted to move the camera on a plan parallel to the studied specimen surface. The working distance of the camera was 13 mm and the depth of field was about 1 m.

A domain of 8 × 6 (horizontal no. × vertical no.) images was observed during the test. The total dimensions of the observation domain were 1080 m × 400 m. Images of this domain were first acquired before the test. Then, during the test, the machine was stopped every 100 cycles and 8 × 6 images were taken. After 2000 cycles, the test was stopped and the same domain was observed in SEM. Every image of a given zone was taken with a lot of precautions in order to represent exactly the same zone of the specimen, which enables to observe the fatigue damage changes at the specimen surface directly. Images of a given zone are shown in Fig. 5 before the test, after 500 cycles, and after 2000 cycles.

1954 P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960

Page 5: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

P. Mu et al. / Procedia Engineering 00 (2010) 000–000 5

Fig. 5 Optical images of the same zone taken before the test (a), after 500 cycles (b) and 2000 cycles (c)

Images above show the fatigue changes in an observation zone during the cyclic test. Before the test, only grain boundaries and twin boundaries could be seen. After 500 cycles, several slip lines appeared in some grains. After 2000 cycles, slip lines intensified and could be seen in almost all grains. In some grains, several slip systems were activated. There seemed to be some cracks along slip lines and in grain boundaries. In order to verify the presence of initiated fatigue cracks, and also to identify the crack orientations, the same observation zone was then observed using a scanning electron microscope (SEM). Observations in this study were done in a scanning electron microscope HITACHI S3600, mainly on secondary electrons, in an accelerating voltage of 15 kV.

In order to have enough cracks for the analysis, the zone observed in SEM was bigger (2.1mm × 2.0mm) than the CCD observation zone (1.08mm × 0.4mm). 4 indentations were made at the surface of specimen to localize CCD and SEM zones.

The zone shown in Fig. 5 has been observed in SEM after the test (Fig. 7). SEM images are more precise than optical images and allow then to distinguish whether or not a crack initiated and to have more quantitative measurements. In particular, it allows to measure the orientations of slip lines and of cracks.

Fig. 7 SEM image of the zone shown in Fig. 4 taken after 2000 cycles

All cracks in the SEM observation zone were looked for and analysed. The crack density after 2000 cycles was 27 cracks / mm2. It depends on the loading conditions (strain amplitude, environment) [7] [8] and on the material (grain size, nitrogen content) [3]. 4 crack initiation origins were found: crack initiate in persistent slip bands (PSB)

P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960 1955

Page 6: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

6 P. Mu et al. / Procedia Engineering 00 (2010) 000–000

(Fig. 8a), in grain boundaries (GB) (Fig. 8b), in twin boundaries (TB) (Fig. 8c), and from inclusions (Fig. 8d). The distribution of the various crack origins is shown in Fig. 9. As only few cracks initiated from inclusions, this crack origin has not been considered. The distributions of PSB, GB and TB cracks are shown Fig. 9. When the type of a given crack was uncertain, it counted for 0.5 in each class. It can be seen that, at room temperature, under a medium strain amplitude ( a=5×10-3, 3.3×10-3) after 2000 cycles, cracks appear mainly in PSBs (51.3%) and in grain boundaries (42.5%). Cracks initiated in twin boundaries (6.2%) occupy only a very little proportion. These results can be compared to those of Mineur [7] and Kamaya [9], both obtained in cyclically strained type AISI 316 austenitic stainless steels under various strain amplitudes. It can be noted that the highest the strain amplitude, the less cracks initiate in persistant slip bands, which is consistent with observations showing that in the high cycle fatigue domain cracks initiate from PSBs [2].

pa

Fig. 8 Cracks initiated in a PSB (a), in grain boundaries (b), in twin boundaries (c), and from an inclusion (d). Cracks are indicated by arrows.

1956 P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960

Page 7: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

P. Mu et al. / Procedia Engineering 00 (2010) 000–000 7

0

10

20

30

40

50

60

PSB GB TBPe

rcen

tage

(%)

Crack Initiation Site

Fig. 9 Distribution of various crack initiation sites observed

Fig. 10 presents the crack orientation distribution of cracks initiated in PSBs. The angle between the load axis and each crack was measured. Quasi no crack initiated in the domain [0°, 40°]. About one crack in two initiates with an angle in domains [50°, 60°] or [70°, 80°]. The three other domains represent about 15% of the cracks.

0

5

10

15

20

25

30

30-40 40-50 50-60 60-70 70-80 80-90

Perc

enta

ge (%

)

Angle between load axis and crack: (°)

Fig. 10 Crack orientation distribution (for cracks initiated in PSBs)

For PSB cracks, it is interesting to know whether cracks initiated in first-appeared slip lines. 3 cracks initiated in PSB belonging to the zone observed with the CCD camera during the test (cracks called A, B and C) have been analysed. As the images of this zone were taken every 100 cycles, the appearance of persistent slip marks (PSMs) and the evolution of the specimen surface could be followed. The first slip marks corresponding to crack A (Fig. 11a), B (Fig. 11c) and C (Fig. 11e) appeared before 100 (Fig. 11b), 100 (Fig. 11d) and 400 cycles (Fig. 11f). Nevertheless, other slip marks were also already visible and did not lead to crack initiation. It seemed that PSB cracks initiated in PSBs that were activated at the very beginning of the cycling. As there were few (no more than 5) initiated cracks in the zone observed during the test, no statistical result could be obtained.

P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960 1957

Page 8: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

8 P. Mu et al. / Procedia Engineering 00 (2010) 000–000

a) b)

c) d)

Fig. 11. SEM images of cracks A (a) and B (c) at 2000 cycles and optical images of cracks A (b, at 100 cycles), B (d, at 400 cycles)

5. Analysis of the crystallographic orientations

EBSD measurements were performed on the specimen surface after the mechanical test to obtain crystallographic orientations of surface grains and the grain boundary arrangement. EBSD scans were performed in beam control mode with a spatial resolution of 3 m / step. The size of the zone scanned by EBSD (Fig. 12a) was 500 × 500 m2

and it contained 302 grains. Fig 12b shows the orientations of 302 grains of the analysed zone. Using high magnification SEM images obtained after 2000 cycles, the slip systems activated are identified in

each grain of the observed zone. From Euler angles measured, for each potential slip system (12 systems), the Schmid factor and the angle between the possible slip markings appearing at the free surface and the loading axis were calculated. A slip system is supposed to be activated when associated slip markings are observed at the surface. For each slip marking, the experimentally measured angle was compared to those calculated above. The slip plan with the Schmid factor closest to the measured value of was considered activated. The slip direction is determined making the assumption that the activated slip system is that with the highest Schmidt factor in the three possible slip systems.

Fig. 12. (a) Grain crystallographic orientation following the transverse direction, (b) grain locations in inverse pole figure [001], (c) SEM image of the same zone taken at the end of test (2000 cycles)

In most grains, only 1 or 2 slip systems are activated. The activated slip system usually has the highest Schmid factor, except in some rare cases. It is important to note that the grains in a polycrystal are in a triaxial state and their

1958 P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960

Page 9: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

P. Mu et al. / Procedia Engineering 00 (2010) 000–000 9

deformations are also influenced by the deformations of neighbour grains. This can explain the activation of a slip system of which the Schmid factor is not the highest.

Fig. 13 presents the cumulative frequency of the highest Schmid factors of all potential slip systems in the analyzed zone. It can be seen that for the grains of the analyzed zone, the highest Schmid factor is always higher than 0.30. This distribution was compared with that of the highest Schmid factors of the activated slip systems in the 75 grains showing plasticity activity. The two distributions are globally in good agreement confirming the validity of the criterion of the highest Schmit factor for slip system activation. The agreement is poorer for Schmid factors in the range [0.35, 0.41]. This may be explained by the size of the zone analyzed, Fig. 12b shows a slight texture of this zone. In particular, the direction [101] is over-represented. The cumulative frequency of the maximal Schmid factor in the analyzed zone is above the cumulative frequency of the maximal Schmid factor in a material perfectly non-textured in the range [0.36, 0.41] [8]. The analyzis of a wider zone is then necessary, in order to verify the texture of the material and the cumulative frequency of the maximal Schmid factor.

010 20 30 40 50 60 70 80 90

100

0.25 0.35 0.45

Percentage(%)

Schmid Factor μ

75 grains with activated slips systemsGrains of the analyzed zone

Fig. 13. Curves of cumulative frequencies of the maximal Schmid factors of the grains of the analyzed zone and of activated slip systems in the 75 analyzed grains

0.25-0.300.30-0.35

0.35-0.400.40-0.450.45-0.50

01234567

Fig. 14 Distribution of primary and secondary Schmid factors in grains showing multiple slip systems activated.

Some grains of the analyzed zone present multiple slip. Their Schmid factors have been specifically analyzed. It can be seen in Fig. 14 the primary Schmid factor was always higher than 0.40, and in 76% of the grains showing multiple slip, the secondary Schmid factor was also higher than 0.40. The two activated slip systems activate one after the other. As it can be seen in Fig. 8 and 11, the slip markings are straight and fine, parallel to each other. These characteristics are typical of the dislocation planar slip, which occurs in austenitic steels (FCC

P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960 1959

Page 10: Microcrack initiation in low-cycle fatigue of an ... · The material studied in th is work is an AISI 316L au stenitic stainless steel. Its chemical composition is given in Tab. 1

10 P. Mu et al. / Procedia Engineering 00 (2010) 000–000

crystallographic structures) all the more easily when the nitrogen content is high [5], [10], [11], [12]. Although the content in nitrogen of the studied alloy is not known, these observations indicate that it should be relatively high. Moreover, the fact that several slip systems with very high Schmid factors activate confirms this assumption. As the cross-slip is not easy, in certain grains, two slip systems are activated to accommodate the plastic deformation.

6. Conclusion and prospect

A fatigue test with a total strain amplitude of 0.5% was carried out on a polished specimen of AISI 316L stainless steel. During the test, images of the surface were regularly taken through an in-situ optical microscope mounted on the testing machine. The time of slip activity and crack initiation have been recorded during the test. After a number of cycles sufficient for small cracks to initiate, the test was stopped and the surface observed in Scanning Electron Microscope. EBSD measurements were made of the observed zone. Crack-initiation sites and slip systems associated to the initiated microcracks have been identified. About half of the cracks initiated in Persistent Slip Bands. Moreover, grain showing multiple slip have high Schmid factors, explained by a high nitrogen content which enhances planar slip.

A polycrystalline model is currently under progress in order to predict crack initiation occurrence observed in this type of material.

Acknowledgements

The authors are grateful to Iris Alvarez-Armas and Alberto Armas (Instituto de Fisica Rosario) for the mechanical test and to Colette Rey (MSSMat, Ecole Centrale Paris) for the EBSD measurements.

Reference

[1] Winter A. A model for the fatigue of copper at low plastic strain amplitudes. Philosophical Magazine 1974; 30, 719-738. [2] Mughrabi H., Wang R., Differt K. Fatigue crack initiation by cyclic slip irreversibilities in high-cycle fatigue. In: Fatigue mechanisms.

ASTM STP 811, American Society for Testing and Materials. 1983, 5-45. [3] Lindstedt U., Karlsson B., Nyström M. Small fatigue cracks in an austenitic stainless steel. Fatigue Fract. Engng. Mater. Struct. 1998;21,

85–98. [4] Zong C., Jin N.Y., Zhou X., Meng E. et Chen X.F. Cyclic deformation of AISI-310 stainless steel I. Cyclic stress-strain responses. Acta

Metallurgica Materiala 1990; 38:2135-2140.[5] Mateo A., Llanes L., Iturgoyen L. et Anglada M. Cyclic stress-strain response and dislocation substructure evolution of a ferrite-austenite

stainless steel. Acta Materiala 1996; 44:1143-1153. [6] Moussavi Toorshizi S.E. Ecrouissage cyclique d'aciers inoxydables austénitique, ferritique et austéno-ferritique : influence de l'histoire de

chargement, PhD thesis, University of Lille 1, 1997.[7] Mineur M., Villechaise P., Mendez J. Influence of the crystalline texture on the fatigue behavior of a 316L austenitic stainless steel.

Materials Science Engineering A, 2000; 286, 257–268. [8] Mineur M. Condition locales d’amorçage des fissures de fatigue dans un acier inoxydable de type 316L : aspects cristallographiques. PhD

thesis, University of Poitiers, 2000. [9] Kamaya M. Influence of bulk damage on crack initiation in low-cycle fatigue of 316 stainless steel. Fatigue Fract. Engng. Mater. Struct.

2009; 33, 94–104. [10] Degallaix, S., Taillard, R. and Foct, J. Role of nitrogen interstitials in plastic fatigue of austenitic stainless steels. In: Fatigue 84, Vol. 1,

(Edited by C. J. Beevers), 1984, 49-59. [11] Degallaix, S., Seddouki, A., Degallaix, G., Kruml, T. and Polàk J. Fatigue damage in austenitic-ferritic duplex stainless steels. Fatigue

Fract. Engng. Mater. Struct. 1995; 18, 65–77. [12] Kruml, T., Polàk J., Obrtlìk K. and Degallaix, S. Dislocation structures in the bands of localized cyclic plastic strain in austenitic 316L

and austenitic-ferritic duplex stainless steels. Acta Materiala 1997 45, 5145–5151.

1960 P. Mu, V. Aubin / Procedia Engineering 2 (2010) 1951–1960