Vous êtes sur la page 1sur 8

Materials Science and Engineering A 472 (2008) 179186

Microstructure and tensile properties of friction stir welded AZ31B magnesium alloy
N. Afrin a , D.L. Chen a, , X. Cao b , M. Jahazi b
a

Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario M5B 2K3, Canada b Aerospace Manufacturing Technology Centre, Institute for Aerospace Research, National Research Council Canada, 5145 Decelles Avenue, Montreal, Quebec H3T 2B2, Canada Received 28 January 2007; received in revised form 5 March 2007; accepted 6 March 2007

Abstract The microstructural change in AZ31B-H24 magnesium (Mg) alloy after friction stir welding (FSW) was examined. The effects of tool rotational speed and welding speed on the microstructure and tensile properties were evaluated. The grain size was observed to increase after FSW, resulting in a drop of microhardness across the welded region from about 70 HV in the base metal to about 50 HV at the center of the stir zone. The obtained HallPetch type relationship showed a strong grain size dependence of the hardness. The aspect ratio and fractal dimension of the grains decreased towards the center of the stir zone. The welding speed had a signicant effect on the microstructure, with larger grains at a lower welding speed. The yield strength and ultimate tensile strength increased with increasing welding speed due to a lower heat input. A lower rotational speed of 500 rpm led to higher yield strength than a higher rotational speed of 1000 rpm. The friction stir welded joints were observed to fail mostly at the boundary between the weld nugget and thermomechanically affected zone at the advancing side. Fracture surfaces showed a mixture of cleavage-like and dimple-like characteristics. Crown Copyright 2007 Published by Elsevier B.V. All rights reserved.
Keywords: Friction stir welding; Magnesium alloy; Microstructural characterization; Grain size; Aspect ratio; Fractal dimension; Microhardness; HallPetch relationship; Tensile properties

1. Introduction Friction stir welding (FSW) is a solid state metal joining technique which was developed and patented by The Weld Institute of Cambridge, UK, in 1991 [1]. This technique is termed as green technology by many researchers due to its energy efciency and environment friendliness [2]. This joining technique was rst used to join aluminum and its alloys. Recently FSW is being used to weld magnesium alloys and other alloys. Structural applications of magnesium alloys are rapidly increasing in automotive and aerospace equipment due to their low density, and ease of castability. Joining of magnesium alloys by conventional techniques is very difcult due to the several problems such as, cracking, expulsion and void in the weld zone [35]. FSW is capable of joining magnesium alloy without melting it and thus can eliminate problems related to the

Corresponding author. Tel.: +1 416 979 5000x6487; fax: +1 416 979 5265. E-mail address: dchen@ryerson.ca (D.L. Chen).

solidication. As FSW does not requite any ller material in the weld zone, the metallurgical problems associated with it can be eliminated and good quality weld can be obtained. Grain renement and higher microhardness in the weld zone of FSW AZ31 magnesium alloy were reported by Wang and Wang [6], whereas Satoshi et al. [7] reported a lower hardness and smaller grain size in the weld zone of AZ31 magnesium alloy compared to the base metal. Nagasawa et al. [8] found smaller grain size in the stir zone of hot rolled AZ31B plate after friction stir welding with insignicant difference in the hardness between the weld zone and the base metal. Separate research carried out by Lee et al. [9] showed signicant reduction in the weld hardness due to the evolution of bigger grain size in the weld zone of FSW AZ31B-H24 hot rolled alloy. Lee et al. [10] studied the effect of different friction stir welding parameters on the tensile strength of extruded AZ31B-H24 magnesium alloy joints and reported that tensile strength decreases with increasing welding speed, whereas Lim et al. [11] found no signicant effect of processing parameters on the tensile strength of friction stir welded AZ31B-H24 alloy. Furthermore,

0921-5093/$ see front matter. Crown Copyright 2007 Published by Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2007.03.018

180

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

investigations carried out by two separate researchers Wang and Wang [6] and Pareek et al. [12] on AZ31B-H24 magnesium alloy showed that the tensile strength increases with increasing rotational speed. These studies showed signicant effects of the welding parameters (rotational speed and welding speed) on the friction stir welded magnesium alloy. It is possible to produce good quality joints with high strength using optimum welding conditions. The purpose of this study is, therefore, to examine the microstructural evolution, hardness and tensile properties of friction stir welded joints of AZ31B-H24 magnesium alloy in different welding conditions. 2. Experimental procedures Friction stir butt welded joints of AZ31B-H24 magnesium alloy, with a thickness of 4.95 mm and chemical composition of 2.53.5% Al, 0.71.3% Zn, 0.201.0% Mn and the balance Mg, were selected in the present study. Friction stir welding was carried out using MTS-FSW machine and in the direction perpendicular to the rolling direction. An adjustable pin tool with a nominal pin diameter of 6.35 mm and shoulder diameter of 19.05 mm was used in the present investigation. Details of friction stir welding conditions are shown in Table 1. Six different welding conditions were applied to these samples. Welding speeds were varied from 1 to 4 mm/s, and rotational speeds were varied from 500 to 1000 rpm. Welding parameters of 500 rpm, 1 mm/s; 500 rpm, 2 mm/s; 750 rpm, 2 mm/s; 1000 rpm, 2 mm/s; 1000 rpm, 3 mm/s; 1000 rpm, 4 mm/s were indicated by samples I, II, III, IV, V and VI, respectively. Samples for microstructural characterization were taken perpendicular to the welding direction (i.e., parallel to the rolling direction in this work). All samples were cut approximately 44 mm in length and were cold-mounted using Lecoset 7007 resin powder and liquid. The samples were then manually ground and polished. The polished samples were etched using 5% nital to show general ow structure of the alloy. A standard reagent made of 4.2 g picric acid, 10 ml acetic acid, 10 ml H2 O and 70 ml ethanol (95% concentration) was used to reveal the microstructure of the welded joints. Microscopic images were taken by light microscope at a magnication of 400 and image analyses were subsequently performed using Clemex software to obtain the grain size, aspect ratio and fractal dimension of the friction stir welded joints. Based on the fractal concept proposed by Manderbrot [13], Manderbrot et al. [14], for irregular objects
Table 1 Welding parameters selected in the present study for the friction stir welding of AZ31B-H24 magnesium alloy butt joints Number I II III IV V VI Rotational speed (rpm) 500 500 750 1000 1000 1000 Welding speed (mm/s) 1.0 2.0 2.0 2.0 3.0 4.0 Fig. 1. A typical example of perimeter (P) vs. area (A) of grains for sample VI showing a linear variation of log(P) with log(A), where the slope represents a half of the fractal dimension.

such as microstructures and fracture surfaces in materials the following relationship holds among the perimeter (P), area (A) and fractal dimension (D), P 1/D A1/2 . This equation can be rewritten as, log P = constant + D 2 log A. (2) (1)

It can be seen that D/2 is the slope of a straight line in a loglog diagram of the perimeter versus area. A typical plot of the grain perimeters (P) as a function of the grain areas (A) in the double-log scales is shown in Fig. 1. A straight line can be seen from the plot, where the fractal dimension is twice the slope [1316]. A computerized Buehler microhardness testing machine was used for the microindentation hardness tests where a load of 100 g and 15 s duration time were used. The test results were recorded in the computer using a Hyper Terminal. Microindentation tests were performed along a path of 40 mm across the sample with an interval of 0.5 mm between two successive dents. The tests were carried out at different distances (e.g., 1 and 4 mm) from the bottom surface and at the top surface of the samples, as shown in Fig. 2. The samples for tensile tests were prepared according to the ASTM E8 standard for sheet type material (i.e., 50 mm gage length and 12.5 mm gauge width). Tensile tests were performed at room temperature and at a strain rate of 6.0 104 s1 . Base metal specimens and samples IV and VI were further tested at a strain rate of 6.7 105 and 7.0 106 s1 to observe the effect of strain rate. After tensile tests fracture surfaces were examined using a scanning electron microscope (SEM) equipped with an energy dispersive X-ray spectroscopy (EDS) system. 3. Results and discussion Fig. 2 shows an overall macroscopic cross sectional image of the friction stir weld zone in sample IV of AZ31B-H24 alloy.

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

181

Fig. 2. A typical macroscopic image of the welded joint of sample IV after friction stir welding.

Fig. 3 shows typical optical micrographs of: (a) stir zone, (b) thermomechanically affected zone (TMAZ), (c) heat-affected (HAZ) and (d) base metal corresponding to the zones marked in Fig. 2. The elongated grains in the base metal have become equiaxed and recrystallized in the stir zone and transition zone between thermomechanically affected zone and stir zone after friction stir welding. The evolution of recrystallized grain structure in the stir zone is due to the severe plastic deformation and frictional heat introduced by the rotating tool pin and its shoulder in the stir zone during welding [9,1719]. The grains in the TMAZ in the present study have also become equiaxed and recrystallized which is different from some magnesium alloys [10] and most of the Al alloys [20] where the TMAZ was characterized by deformed and elongated grains. Fig. 4a shows the variation of grain size across the FSW specimen at different depths (at 1 and 4 mm from the bottom surface, and on the top surface). The grain size of approximately 5 m of the base metal

became considerably bigger (8.59.3 m) in the center of the stir zone after FSW. Slightly larger grains are observed on the top surface compared to the bottom surface in the stir zone. The formation of larger grains in the stir zone was also reported by Lim et al. [11] and Pareek et al. [12] for AZ31-H24 alloy. Similarly, Chang et al. [19] observed a smaller grain size at the bottom than at the top of friction stir processed AZ31 magnesium alloy and the heat generated by the tool shoulder is considered to be responsible for the formation of larger grains close to the top surface. Fig. 4b refers to the variation of grain size across the welded specimen having two different welding conditions 1000 rpm, 2 mm/s (sample IV) and 1000 rpm, 4 mm/s (sample VI), respectively. The maximum grain size near the center of the weld zone is slightly bigger in sample IV (about 10.2 m) than in sample VI (about 9.3 m). The smaller grains produced in the stir zone induced by the faster welding speed is attributed to the relative lower heat input in the weld. Another reason for

Fig. 3. Optical microscope images of: (a) weld nugget, (b) thermomechanically affected zone (TMAZ), (c) heat affected zone (HAZ) and (d) base metal as indicated in Fig. 2.

182

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

Fig. 4. Grain size distribution across the weld (a) at different depths from the bottom surface for sample VI (1000 rpm, 4 mm/s), (b) at a distance of 1 mm from the bottom surface for sample VI (1000 rpm, 4 mm/s) and sample IV (1000 rpm, 2 mm/s).

Fig. 5. (a) Aspect ratio and (b) fractal dimension as a function of the distance from the center of stir zone for sample VI (1000 rpm, 4 mm/s) at different depths from the bottom surface.

the grain size reduction with increasing welding speed could be the greater straining in the material which activates more strain free nucleation sites as reported by Pareek et al. [12]. Fig. 5a shows typical proles of aspect ratio of a specimen where the aspect ratio is smaller in the stir zone due to the equiaxed grains and it increases with increasing distance from the weld centerline towards the base metal. Fig. 5b shows the fractal dimension of grains in the AZ31B-H24 alloy across the weld zone and base metal. Likewise, the fractal dimension close to the center of the weld zone has smaller values than those close to the base metal. This is due to the less irregularity in the stir zone exhibited by the equiaxed grains. Fig. 6 shows a typical hardness prole of a specimen (welding parameters 1000 rpm, 4 mm/s) across the weld zone at the top surface, 1 and 4 mm from the bottom surface. The hardness prole shows a lower hardness in the stir zone than in the base metal due to larger grain sizes in the stir zone. Similar results were also observed in [9]. Vilaca et al. [21] reported that if the ratio of the rotational speed (rpm) to the welding speed (mm/min) becomes more than 4, it can be considered as a hot welding condition and the resulting hardness exhibits a lower value between TMAZ and HAZ. Fig. 7 shows a comparison of hardness proles between two samples at 1 mm from the bottom surface and 4 mm from the bottom sur-

face, respectively. It is seen that sample VI with a higher welding speed of 4 mm/s shows slightly higher hardness than sample IV with 2 mm/s welding speed. The formation of smaller grain sizes due to the lower heat input in sample VI (1000 rpm, 4 mm/s) rendered slightly higher hardness in the weld zone compared to sample IV (1000 rpm, 2 mm/s). The obtained hardness values are plotted as a function of the reciprocal of the square root of

Fig. 6. Typical microhardness prole across the friction stir welded specimen VI (1000 rpm, 4 mm/s) at different depths (top, 1 mm and 4 mm from the bottom surface).

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

183

Fig. 7. A comparison of the hardness values between samples IV (1000 rpm, 2 mm/s) and VI (1000 rpm, 4 mm/s) at: (a) 1 mm from the bottom surface and (b) 4 mm from the bottom surface.

Fig. 9. Yield strength (YS) and ultimate tensile strength (UTS) of FSW AZ31BH24 alloy at: (a) different rotational speeds (rpm) and (b) different welding speeds (mm/s).

the grain sizes in Fig. 8. The hardness values and grain sizes of sample VI (1000 rpm, 4 mm/s) are taken at the center and both sides of the weld nugget at a distance of 5, 10, 15 and 20 mm. It is seen that the HallPetch type linear relationship is followed and could be written as HV = 16.4 + 119.5d1/2 , where d is the grain size. The relationship shows a strong grain size dependence of

Fig. 8. A typical plot of the HallPetch type relationship for the friction stir welded sample VI (1000 rpm, 4 mm/s).

the hardness in the FSWed AZ31B-H24 alloy. The grain boundaries thus become the main obstacle to the slip of dislocations and the materials with a smaller grain size would have higher hardness or strength as it would impose more restriction to the dislocation movement. Fig. 9 shows the tensile test results for specimens having: (a) different rotational speeds and (b) different welding speeds. The yield strength (YS) decreases with increasing rotational speed, whereas the ultimate tensile strength (UTS) decreases when the rotational speed changes from 500 to 750 rpm and then increases form 750 rpm to 1000 rpm as seen in Fig. 9a. Fig. 9b shows both YS and UTS increase with increasing welding speed. Specimens with a lower rotational speed (500 rpm) show a higher yield strength than the specimens with a higher rotational speed (1000 rpm) regardless of the welding speed. The highest UTS of about 201 MPa in the present study was obtained for sample VI (1000 rpm, 4 mm/s). An average elongation of about 2.5% was observed for FSW joints, which is approximately 21% of the elongation of the base metal. The base metal has also a higher yield strength (about 208 MPa), ultimate tensile strength (about 309 MPa) and elongation (12%) compared to those of friction stir welded joints of AZ31B-H24 alloy, as indicated by the dashed lines in Fig. 9a. The reason for such a decrease in the tensile properties could be explained by the grain growth in the stir zone and TMAZ, and the presence of signicant amount

184

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

Fig. 10. A comparison of the yield strength (YS) and ultimate tensile strength (UTS) obtained in the present study with those reported in the literature.

Fig. 11. Yield strength and ultimate tensile strength as a function of the strain rate for the base metal and friction stir welded samples IV (1000 rpm, 2 mm/s) and VI (1000 rpm, 4 mm/s) of AZ31B-H24 alloy.

of oxide layer in the boundary between the TMAZ and the stir zone at AS, as reported by Lim et al. [11] and Gharacheh et al. [22]. Another reason would be that magnesium alloys having an hcp crystal structure found themselves responsible for this reduction in the tensile properties due to the formation of new crystallographic texture in the weld zone as reported by Park et al. [17]. Fig. 10 presents a comparison of the tensile test results obtained in the present study with those reported in the literature at varying welding speeds. It can be seen that the obtained YS and UTS are in good agreement with the results reported by other researchers [11,12]. Fig. 11 shows the tensile test results of both base metal and friction stir welded joints under conditions of 1000 rpm, 4 mm/s and 1000 rpm, 2 mm/s as a function of the average strain rates. The change of the strain rates seems to exhibit little or no effect on the YS and UTS of AZ31B-H24 magnesium alloy and its FSWed joints. However, decreasing the tensile test strain rate from 6.0 104 to 6.7 105 s1 increases the elongation from 12 to16.4% for the base metal, and from 2.3 to 3% for sample IV (1000 rpm, 2 mm/s), and slightly decreases the elongation from 3 to 2.8% for sample VI (1000 rpm, 4 mm/s). Similar results on the elongation were observed when the strain rate was further decreased to 7.0 106 s1 . Fig. 12 shows a typical tensile fracture location of the friction stir welded joints of AZ31B-H24 alloy. Among 15 tensile

samples tested, 14 of them failed at the boundary between the stir zone and TMAZ at the advancing side. The failure of the samples was basically 45 shear fracture, which could be due to the formation of texture by the shear deformation resulted from the rotation of the pin and tool shoulder in that region [12]. Some authors reported the fracture at the boundary between the TMAZ and the stir zone due to the fact that the accumulation of basal slip plane (0 0 0 1) having their normal parallel to the transverse direction (TD) is maximum in that location leading to the minimum Schmids factor [23,24]. Fracture locations are observed to be about 58 mm from the weld centerline measured at the top surface of the specimens. At a low rotational speed of 500 rpm, failure occurs closer (56 mm from the weld centerline) to the weld nugget than at a higher rotational speed of 1000 rpm (68 mm from weld centerline). The lower rotational speed creates smaller weld zone due to the lower heat input, and the boundary between the TMAZ and the SZ (i.e., the fracture location in the present study) thus shifts closer to the weld centerline [22]. Fig. 13 shows some typical SEM micrographs of fracture surfaces for: (a) the base metal, (b and c) the friction stir welded joints. The base metal exhibits mainly features of elongated dimples together with some tear ridges, as shown in Fig. 13a. Both dimple-like characteristics (Fig. 13b) and cleavage-like features (Fig. 13c) can be seen in different areas on the fracture surfaces

Fig. 12. A macroscopic image showing typical fracture location of a friction stir welded specimen of AZ31B-H24 alloy.

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

185

Fig. 14. (a) An SEM micrograph of tensile fracture surface of a friction stir welded specimen of AZ31B-H24 alloy and (b) EDS line scan proles across the particle showing the distribution of magnesium, oxygen and aluminum.

Fig. 13. Typical SEM micrographs of fracture surfaces of AZ31B-H24 alloy after tensile testing: (a) base metal, (b and c) friction stir welded specimens III (750 rpm, 2 mm/s) and VI (1000 rpm, 4 mm/s), respectively.

4. Conclusions 1. Microstructural examinations of AZ31B-H24 alloy after friction stir welding (FSW) revealed that the grains in the stir zone and thermomechanically affected zone (TMAZ) exhibited recrystallization and growth. The grain shape became equiaxed, giving rise to smaller values of both aspect ratio and fractal dimension. Smaller grain sizes were observed in the stir zone at a higher welding speed due to a lower heat input. 2. Lower hardness values in the stir zone and the TMAZ were observed. Faster welding speed produced slightly higher hardness in the stir zone and the TMAZ. The obtained HallPetch type relationship showed a strong grain size dependence of the hardness of AZ31B-H24 alloy after FSW.

after FSW. Tensile fracture initiation could have started from a cleavage area between the weld nugget and the TMAZ as also reported by Lim et al. [11]. Fig. 14a shows an SEM image taken from the tensile fracture surface of an FSW specimen containing some particles. An EDS line scan analysis across the particle reveals signicant oxygen content as seen from Fig. 14b. These oxide particles present on the fracture surface could be the trapped oxide during friction stir welding and would be partially responsible for the reduction in the tensile strength and elongation of the joints [10,11].

186

N. Afrin et al. / Materials Science and Engineering A 472 (2008) 179186

3. The tensile test results showed that the yield strength and ultimate tensile strength increased with increasing welding speed due to the lower heat input in the weld region. A lower rotational speed of 500 rpm gave rise to a higher yield strength than a higher rotational speed of 1000 rpm. 4. The change in the strain rate from 6.0 104 to 6.7 105 s1 showed little or insignicant effect on the yield strength and ultimate tensile strength of both base metal and FSWed AZ31B-H24 alloy. However, the elongation of the base metal increased from 12 to 16% when the strain rate changed from 6 104 to 6.7 105 s1 . 5. The majority of the FSWed joints were observed to fail at the boundary between the stir zone and the thermomechanically affected zone (TMAZ) at the advancing side. The failure of the tensile samples suggested a 45 shear fracture. The failure in the TMAZ could be explained by the grain growth and the presence of oxides on the fracture surface. 6. Fracture surfaces exhibited both cleavage-like and dimplelike fracture characteristics for the FSWed AZ31B-H24 alloy after tensile tests. Acknowledgements The authors would like to thank the National Sciences and Engineering Research Council (NSERC) of Canada for providing nancial support. This investigation involves a part of multi-national CanadaChinaUSA Collaborative Research Project on the Magnesium Front End Research and Development (MFERD). One of the authors (D.L. Chen) is also grateful for the nancial support by the Premiers Research Excellence Award (PREA), Canada Foundation for Innovation (CFI) and Ryerson Research Chair (RRC) program. The authors would also like to thank Messrs. A. Machin, Q. Li, J. Amankrah and R. Churaman for easy access in the laboratory of Ryerson University and their assistance in the experiments during the study. Thanks are also due to Mr. M. Guerin for the preparation of FSWed samples using FSW system, and to Professor S.D. Bhole for his helpful discussion.

References
[1] W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Church, P. Templesmith, C.J. Dawes, GB Patent Application No. 9125978.9 (December 1991). [2] R.S. Mishra, Z.Y. Ma, Mater. Sci. Eng. R 50 (2005) 178. [3] Y.R. Wang, Z.D. Zhang, Trans. China Weld. Inst. 27 (2006) 912. [4] L. Liu, C. Dong, Mater. Lett. 60 (2006) 21942197. [5] T. Kim, J. Kim, Y. Hasegawa, Y. Suga, Proceedings of the Third International Symposium on Designing, Processing and Properties of Advanced Engineering Materials, Jeju Island, South Korea, November 58, 2003, pp. 417420. [6] X.H. Wang, K.S. Wang, Mater. Sci. Eng. A 431 (2006) 114117. [7] H. Satoshi, O. Kazutaka, D. Masayuki, O. Hisanori, I. Masahisa, A. Yasuhisa, Q. J. Jpn. Weld. Soc. 21 (2003) 539545. [8] T. Nagasawa, M. Otsuka, T. Yokota, T. Ueki, in: H.I. Kaplan, J. Hryn, B. Clow (Eds.), Magnesium Technology 2000, TMS, 2000, pp. 383387. [9] W.B. Lee, Y.M. Yeon, S.B. Jung, Mater. Sci. Technol. 19 (2003) 785790. [10] W.B. Lee, Y.M. Yeon, S.K. Kim, Y.J. Kim, S.B. Jung, in: H.I. Kaplan (Ed.), Magnesium Technology 2002, TMS, 2002, pp. 309312. [11] S. Lim, S. Kim, C.-G. Lee, C.D. Yim, S.J. Kim, Metall. Mater. Trans. A 36 (2005) 16091612. [12] M. Pareek, A. Polar, F. Rumiche, J.E. Indacochea, Proceedings of the Seventh International Conference on Trends in Welding Research, Pine Mountain, GA, United States, May 1620, 2005, ASM International, 2006, pp. 421426. [13] B.B. Manderbrot, Int. J. Fract. 138 (2006) 1317. [14] B.B. Manderbrot, D.E. Passoja, A.J. Paullay, Nature 308 (1984) 721722. [15] D.L. Chen, D.X. Pang, Z.J. Yang, S. Kong, L.T. Wang, K. Yang, G.W. Quao, J. Phys. C: Solid State Phys. 21 (1988) 271276. [16] Z.G. Wang, D.L. Chen, X.X. Jiang, S.H. Ai, C.H. Shih, Scripta Mater. 22 (1988) 827832. [17] S.H.C. Park, Y.S. Sato, H. Kokawa, Metall. Mater. Trans. A 34 (2003) 987994. [18] S.H.C. Park, Y.S. Sato, H. Kokawa, Scripta Mater. 49 (2003) 161166. [19] C.I. Chang, C.J. Lee, J.C. Huang, Scripta Mater. 51 (2004) 509514. [20] G. Bussu, P.E. Irving, Int. J. Fatigue 25 (2003) 7788. [21] P. Vilaca, L. Quintino, J.F.D. Santos, J. Mater. Proc. Technol. 169 (2005) 452465. [22] M.A. Gharacheh, A.H. Kokabi, G.H. Daneshi, B. Shalchi, R. Sarra, Int. J. Mach. Tools Manuf. 46 (2006) 19831987. [23] W.D. Callister Jr., Materials Science and EngineeringAn Introduction, seventh ed., John Wiley & Sons, Inc., New York, 2007. [24] W. Woo, H. Choo, D.W. Brown, P.K. Liaw, Z. Feng, Scripta Mater. 54 (2006) 18591864.

Vous aimerez peut-être aussi