Você está na página 1de 7

This article appeared in a journal published by Elsevier.

The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elseviers archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Author's personal copy

Intermetallics 19 (2011) 738e743

Contents lists available at ScienceDirect

Intermetallics
journal homepage: www.elsevier.com/locate/intermet

Effects of characteristics of Mo dispersions on the plasticity of Mg-based bulk metallic glass composites
J.S.C. Jang a, *, T.H. Li b, S.R. Jian b, J.C. Huang c, T.G. Nieh d
a

Department of Mechanical Engineering, Institute of Materials Science and Engineering, National Central University, Chung-Li 32001, Taiwan, ROC Department of Materials Science and Engineering, I-Shou University, Kaohsiung 840, Taiwan, ROC c Department of Materials and Optoelectronic Science, Center for Nanoscience and Nanotechnology, National Sun Yat-Sen University, Kaohsiung 804, Taiwan, ROC d Department of Materials Science and Engineering, The University of Tennessee, Knoxville, TN 37996, USA
b

a r t i c l e i n f o
Article history: Received 31 August 2010 Received in revised form 24 January 2011 Accepted 26 January 2011 Available online 15 February 2011 Keywords: A. Composites B. Dispersion strengthening B. Metallic glasses B. Mechanical properties at ambient temperature C. Casting

a b s t r a c t
The Mg58Cu28.5Gd11Ag2.5 bulk metallic glass composites (BMGCs) dispersion strengthened by porous Mo particles with different volume fractions and particle sizes were synthesized by casting and characterized. The presence of porous Mo particles could restrict the shear band propagation. It was found for a given volume fraction of Mo particles, smaller particles would lead to more interfacial areas, shorter inter-particle free spacings, smaller connement zone sizes than the larger particles, and consequently the improvement of compression plasticity from 10% up to 27%. Also, for a given Mo particle size, higher volume fraction would lead to larger compression plasticity. The inter-particle free spacing, as well as the connement zone size (mean free path of shear bands), appears to be the controlling factor in limiting the propagation of shear bands which in turn affecting the plasticity of BMGCs. 2011 Elsevier Ltd. All rights reserved.

1. Introduction After Mg-based bulk metallic glasses (BMGs) were discovered in the early 1990s [1,2], they are attractive as engineering materials because of their high specic strength/density ratio as compared to other BMGs [1e6]. However, the monolithic Mg-based BMGs are known to be extremely brittle and tend to fracture into pieces before yielding [7,8]. To amend the weakness, the high m/B ratio (where the m and B are shear and bulk modulus, respectively) concept was proposed to predict the compositions of monolithic Mg-based BMGs by Zhang et al. [9]. However, the resulting BMGs still could not attain the ductility level seen in Pt- or Zr-based BMGs. Recently, many efforts have devoted to develop Mg-based and Zr-based metallic glasses composites (BMGCs) for plasticity improvement. These include the incorporation of in-situ precipitated secondary nano- or micro-sized phases, ex-situ added refractory ceramics or ductile metal particles in the Mg-based BMGCs [9e24]. Among those Mg-based BMGCs, the current authors have demonstrated a highly tough Mg-based BMGC rods strengthened by porous Mo particles [15,18,25]. These porous Mo particles are

uniformly distributed in the amorphous matrix of BMG and arranges as the micron-scaled connement zones. The small region can undergo great deformation without failure and limit the propensity of forming mature shear bands [26e30]. These porous Mo particles also act as ductilizers to signicantly improve the plasticity. Moreover, the residual porosity content in the cast Mgbased BMGC can be drastically decreased by increasing the cooling rate [18], which results in further enhancement of plasticity. To understand the dispersion toughening of this Mg-based BMGC, a series of Mg58Cu28.5Gd11Ag2.5 BMGCs reinforced with different sizes and volume fractions of porous Mo particles were systematically investigated. 2. Experimental details Mg-based BMGCs were prepared by selecting the composition of Mg58Cu28.5Gd11Ag2.5 as the raw alloy and adding different volume fractions (Vf 5e25 vol%) of porous Mo particles. High purity Cu and Gd (>99.9%) were pre-alloyed into Cu-Gd master alloy ingot by arc melting in a Ti-gettered argon atmosphere. Then the Cu-Gd master alloy was melted together with high purity Mg and Ag pieces to obtain the target composition by induction melting under the argon atmosphere. While melting, high purity porous Mo particles (99.9wt% pure and fabricated by sintering

* Corresponding author. Tel.: 886 4267379; fax: 886 4254501. E-mail address: jscjang@ncu.edu.tw (J.S.C. Jang). 0966-9795/$ e see front matter 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.intermet.2011.01.013

Author's personal copy

J.S.C. Jang et al. / Intermetallics 19 (2011) 738e743

739

method as shown in Fig. 1) with three different particle sizes, namely, 20 4 mm (ranging within 10e25 mm), 30 4 mm (25e44 mm), and 58 8 mm (56e65 mm) in diameter, were added into the matrix alloy under the argon atmosphere, respectively. Mechanical stirring was exerted to enhance the homogeneous mixing of the particles within the melt. The composite alloy ingot was further remelted by induction melting in a quartz tube and injected into a water-cooled Cu mold with temperature-controlled at 8  C by argon pressure to obtain the BMGC rods with a size of 2 mm in diameter. The temperature of Cu mold were kept at 8  C to reach a cooling rate of 63 K/s for 2 mm f rods, and to obtain a BMGC rod with residual porosity less than 0.15 vol% [18]. The conrmation of amorphous state in the Mg-based BMGCs was veried by TA Instruments DSC 2920 differential scanning calorimeter (DSC) under owing puried argon with a heating rate of 20 K/min, and was examined by Scintag X-400 X-ray diffractometer with monochromatic Cu-Ka radiation. The compression tests were performed at a strain rate of 5 104 s1 by an MTS 810 mechanical test system. The compression samples with the height (4 mm) to diameter (2 mm) ratio of 2:1 were cut in parallel and carefully polished to insure the end atness. Multiple compression tests were conducted to ensure the reproducible trend and the scattering of the stress and strain was less than 5%. The distribution of Mo particles in the amorphous matrix of BMGCs and the morphology of fracture surfaces were examined by a Hitachi S4700 and S3400 scanning electron microscope (SEM) with energy dispersive spectrometry (EDS) capability. 3. Results 3.1. Microstructure The representative XRD patterns of the Mg-based BMGC rods with 20 4 mm particle size of Mo are shown in Fig. 2, including an amorphous phase with the broadening diffused humps. Similar patterns are obtained from other Mg-BMGCs with larger sized Mo dispersoids. No apparent crystalline peaks are detected in the pattern except the high intensity crystalline peaks of pure Mo particles. This indicates that no obvious chemical reaction occurs to deteriorate the glass forming ability. This is also demonstrated by the TEM examination on the amorphous matrix-Mo interface in our previous report [18]. As a result, the GFA index, e.g. g [31] and gm

(110)

Intensity (a. u.)

(211) (200) 15 vol%

20 vol% 25 vol% 20 30 40 50 2 (deg.)


Fig. 2. XRD pattern of the Mg58Cu28.5Gd11Ag2.5 BMGCs contained 20 4 mm Mo particles with different vol.% Mo particles. (Jason S.-C. Jang et al.)

60

70

80

[32], of all these Mg-based BMGCs exhibits similar values as their matrix counterparts, about 0.422 and 0.752, respectively. SEM micrographs of Mg58Cu28.5Gd11Ag2.5 BMG contained different particle sizes of Mo are shown in Fig. 3. These Mo particles, in the size (d) range of 20 4, 30 4, and 58 8 mm, are observed to have a relatively homogeneously dispersion in the amorphous matrix. These Mo particles are porous and during casting they are inltrated by Mg-based metal liquid, resulting in completely lled spheres. The nal volume fraction Vf of the Mo particles in the composite estimated by image analysis was found very close to the initial addition. The average mean free distance in the current Mgbased composites are summarized in Table 1. 3.2. Mechanical performance Fig. 4 shows the compression stressestrain curves for the monolithic Mg-based BMG and its composites (Vf 0e25%) tested under an initial strain rate of 5 104 s1 at room temperature. Emphasis in this report is made on the failure strain. It is evident in Fig. 4 that the failure strain exhibits an increase with increasing Mo content and decreasing Mo particle size. It was also noted that the ow stress increases with increasing Mo volume fraction. The ow stress increases from w800 MPa for the monolithic BMG up to 1050 MPa; and the failure strain increases from zero for the based BMG up to 27% (the highest failure strain reported in the literature for the Mg-based metallic glass system). Note the largest failure strain of all BMGCs occurs at the highest volume fractions, e.g., 12% strain for the BMGC with Vf 25% and d 58 8 mm, 24% strain for the BMGC with Vf 25% and d 30 4 mm, and 27% strain for the BMGC with Vf 25% and d 20 4 mm. After yielding, these BMGC specimens continuously deform and show slight work hardening accompanied by a remarkable improvement in plastic strain. Plastic deformation in both monolithic BMG and the current composites is mainly by shear banding in the amorphous matrix. Particularly in the composites, shear bands cannot travel freely or run away but are often arrested by the amorphous matrix-Mo interfaces. The traveling distance, i.e. mean free path, of shear bands is affected by the presence of Mo particles and, specically, limited by the inter-particle spacing of Mo particles. For the composites, the mean free path is determined by both the size and

Fig. 1. Appearance of the porous Mo particles with particle size of 58 8 mm. (J.S.C. Jang et al.)

Author's personal copy

740

J.S.C. Jang et al. / Intermetallics 19 (2011) 738e743 Table 1 Inter-particle spacings (in mm) of Mo particles in the Mg-based BMGCs with different Mo particle volume fraction (vol.%) and sizes. Vf of Mo particles (vol%) Mo particle size 20 4 mm 30 4 mm 58 8 mm 5 10 15 20 25 Porosity of single porous Mo particle (vol%) e e 53 40 31 13 93 61 55 42 34 18 23 14 10 6 6 3 e 75 59 44 39 35 33 11 13 8 10

11 18 6 1

a
True stress (MPa)

1200 1000 800 600 400 200

Mo particle Size: 58/8 m

25 vol.% 20 vol.% 15 vol.% 10 vol.% Base

4% 0.20 0.24 0.28

0.00

0.04

0.08

0.12

0.16

Ture strain

b
True stress (MPa)

1200 1000 800 600 400 200

Mo particle Size: 30/4 m

25 vol.% 20 vol.% 15 vol.% 10 vol.% 5 vol.%

4% 0.24 0.28

0.00

0.04

0.08

0.12

0.16

0.20

True strain

c
True stress (MPa)

1200 1000 800 600 400 200


25 vol.% 20 vol.% 15 vol.%

Mo particle size: 20/4 m

Fig. 3. SEM images of the cross section for the Mg58Cu28.5Gd11Ag2.5 BMGC rods contain 25 vol.% of Mo particles with different Mo particle size which fabricated by a cooling rate of 63 K/s: (a) 20 4 mm, (b) 30 4 mm, and (c) 58 8 mm Mo particle size. (Jason S.-C. Jang et al.)

4% 0.16 0.20 0.24 0.28

volume fraction of Mo particles; the values are listed in Table 1. For a given Vf of Mo particles, the smaller particles would lead to more interfacial areas, shorter inter-particle spacings, and smaller mean free path. This inter-particle free spacing is an effective index for the mean free path of shear banding. The smaller mean free path results in higher plasticity.

0.00

0.04

0.08

0.12

True strain
Fig. 4. Compressive true stressestrain curves of the Mg58Cu28.5Gd11Ag2.5 BMGC rods with different vol.% and particle size of Mo particles: (a) 58 8 mm, (b) 30 4 mm, and (c) 20 4 mm Mo particle size. (Jason S.-C. Jang et al.)

Author's personal copy

J.S.C. Jang et al. / Intermetallics 19 (2011) 738e743

741

Fracture Toughness (MPa m0.5)

50 40 30 20 10 0

Fracture Toughness Strain

40 35

fraction. It is suggested that the current porous Mo particles with nano grain size exhibits stronger strengthening capability.

25 20 15 10 5 0.10 0.12 0.14 0.16 0.18 0

Failure strain (%)

30

3.3. SEM observation on the fractured specimen Multiple shear bands can be observed clearly on the specimen surface. Near the fracture surface and around the Mo particles, as shown in Fig. 7(a), lots of stress concentration was released by the plastic deformation of ductile Mo particles. This indicates that there exists a strong interaction between ductile Mo particle and shear banding from the amorphous matrix. The entire fracture surface of these Mg-based BMGC rods after compression test exhibits two kinds of morphology, one wavy fracture surface mixed with some area with a vein pattern (Fig. 7(b)) and rough shear-off cleavage fracture surface (Fig. 7(c)). Fig. 7(b) gives direct evidence that the shear bands in the matrix are obstructed by the Mo particles, the propagation of shear banding was deected in avoiding a catastrophic cracking and fracture. On the other hand, Fig. 7(c) reveals that several porous Mo particles dispersed in the glassy matrix were cut off by the shear-banding process. Here, many micro-cracking traces (similar to the river patterns in fatigue tests, denoted by white arrows) were terminated at the boundaries of porous Mo regions, and they appeared to be absorbed by the Mo particles, as shown in Fig. 7(d). It directly shows that the micro-cracks cannot easily propagate across the porous Mo particles, but are conned within the inter-particle regions. Within the ne-scale compartments (w1 mm) inside one porous Mo particle (w20 4 mm for this sample), deformation and fracture traces can also be observed in the amorphous matrix, as shown in Fig. 7(e). The sub-micron vein pattern is evidently a result of local shear banding within the amorphous matrix inside that porous Mo particle. 4. Discussion SEM micrographs in Fig. 7 show that the deformation occurs mainly through a shear-banding mechanism in the amorphous matrix. It is known that new shear bands are preferentially initiated at structure inhomogeneous sites where there is local stress concentration [34]. Accordingly, the large elastic modulus difference between the Mo particles (E 327 GPa) and the Mg base amorphous matrix (E 40 GPa) would cause a large stress concentration at the particleematrix interface, and will induce the initiation of shear bands. Therefore, we can nd numerous shear bands originated or terminated at the Mo particle-amorphous matrix interface as shown in Fig. 7(a). Also, a vein pattern around the Mo particles can often be observed on the fracture surface, as shown in Fig. 7(b). Vein pattern is a result of sudden release of strain energy on fracture. This means that a substantial increase in temperature in these localized shear bands has been achieved, thus the shear-induced temperature rise can adiabatically heat the porous Mo particles to a high level and can result in softening of Mo considerably [28]. This would be benecial in arresting shear band propagation by the softened Mo at the shear band front. For a given amount of particles, the porous particles will generate more interfaces between the reinforcements and matrix, and thus, can conne lots of micro-sized compartments of the Mgbased glassy phase even within one porous Mo particle. This will result in the formation of multiple shear bands within or around the porous Mo particles, promoting the deformation to distribute more uniformly across the specimens and making difcult for a shear band to propagate through the entire sample. As a result, the mechanical instability is retarded and the ductility is improved. Xi et al. [35] has demonstrated a correlation between the fracture toughness, K1C, and the length scale of plastic process zone, w, for various brittle and tough BMGs and associates the local plastic

d-0.5 (m-0.5)
Fig. 5. (a) Kc (stress concentration factor) and compressive plastic strain as a function of interspacing of Mo particles of the Mg-based BMGC rods with 2 mm in diameter. (Jason S.-C. Jang et al.)

The fracture toughness (KC, stress concentration factor) for these Mg-based BMGC rods was measured by an indentation method [33]. Fig. 5 shows the dependency of the fracture toughness (KC) and the compression failure strain as a function of the mean free path. Both the failure strain and fracture toughness exhibit a similar trend, namely, they increase with decreasing mean free path. It is apparent that the mean free path plays an important and controlling role on the toughening of the Mg-based metallic glasses. The dependence trend of d0.5 (d is mean free path) on the toughness as well as compression failure strain in Fig. 5 presents a linear relationship, which is similar to the strengthening behavior of grain renement on the crystalline alloys. When the mean free path is smaller than about 40 mm, the fracture toughness KC can be greater p than 20 MPa m and the failure strain can be higher than 10%, favorable for application. Mean free path also has a minor positive effect on the Vickers hardness (Hv) and yield stress (YS), as depicted in Fig. 6. It is of interest to note that even though the polycrystalline Mo dispersoid has lower YS than the Mg-based amorphous matrix, the hardness and YS of the Mg-BMGCs still presents small increasing trend with decreasing inter-particle spacing: Hv increases from 250 up to 280 and YS increases from 800 up to 900 MPa. This differs from the other BMGCs with ductile dispersoids [17,19,20], showing decreasing hardness with increasing ductile dispersoid volume

340
Hardenss

1100 1000 900 800 700 600

Hardness (Hv)

300 280 260 240 220 200 0.10 0.12 0.14 0.16 0.18

d-0.5 (m)-0.5
Fig. 6. Hardness and compressive yield strength as a function of interspacing of Mo particles of the Mg-based BMGC rods with 2 mm in diameter. (Jason S.-C. Jang et al.)

Yield Strength (MPa)

320

Yield Strength

Author's personal copy

742

J.S.C. Jang et al. / Intermetallics 19 (2011) 738e743

Fig. 7. SEM micrographs of (a) specimen surface near the fracture area after compression test for the Mg-based BMGC containing 25 vol.% Mo particles with size of 20 4 mm, (b) some area appears vein pattern mixed with wavy fracture surface, (c) some area appears shear-off Mo particles which block the moving of plastic shear bands, (d) enlarged image from the circle of Fig. 7(c), and (e) enlarged image from the circle of Fig. 7(d). (Jason S.-C. Jang et al.)

zone with the crack tip opening displacement (CTOD), using the equation

2 w 1=6p KC =sy

(1)

Where sy is the yield stress. For brittle Mg65Cu25Tb10 BMG, with KC 2 and sy 660 MPa, only gave a plastic zone size about 0.2 mm. On the contrary, tougher metallic glasses exhibit a much larger scale of local plasticity (e.g., Zr-based BMG having a 57 mm plastic zone size) as well as drastic crack bifurcation and branching, and then leading to a much higher CTOD. For the present monolithic Mg58Cu28.5Gd11Ag2.5 BMG (without porous Mo), fracture toughness KC and yield strength sy are about 5 MPa m0.5 and 840 MPa, respectively [18]. The plastic zone size w calculated by the above equation is only about 2 mm and obviously too small in comparison with that in the more ductile Zr-based p BMGs. However, when we insert KC of w35 MPa m and sy of 1050 MPa for the Mg-based BMGC, the plastic zone size w is about 59 mm, which is comparable or even greater than that in the ductile Zr-based BMGs. This plastic zone size is also noted to be greater than or comparable to the mean free path for shear bands (Table 1). Thus, shear bands can stably propagate in the material without causing catastrophic failure.

It is pointed out that for the current Mg-based BMGC, there are two mean free paths. The rst one is the dimension of free space within the porous Mo particle, measuring about 1 mm or less, smaller than the plastic zone size (w2 mm for monolithic Mg-based BMG). Thus, it is conceivable that the crack would not be fully developed inside the porous Mo particle. On the other hand, the mean free path for the Mg-based BMGCs with 25vol% 20, 30 or 58 mm porous Mo particles are about 30e40 mm, which is much larger than that within a porous Mo particle. Therefore, fracture is primary controlled by shear bands moving between Mo particles, the crack propagation will be deected into a wavy path and results in remarkable plasticity increase. 5. Conclusions In summary, the Mg-based BMG composites with different volume fractions and particle sizes of porous Mo particles have been successfully developed in this study. It was found that shear bands are often arrested by the amorphous matrix-Mo interfaces in these Mg-based BMGCs. The traveling distance, i.e. mean free path, of shear bands is drastically limited by the inter-particle free spacing of Mo particles. For a given Vf of Mo particles, the smaller particles would lead to more interfacial areas, shorter inter-particle

Author's personal copy

J.S.C. Jang et al. / Intermetallics 19 (2011) 738e743

743

free spacings, and smaller mean free path of shear banding than the larger particles. This will result in forming denser multiple shear bands within or around the porous Mo particles, making more difcult for a shear band to propagate through the entire sample. As a result, the compression plasticity can be improved from 10% strain (for the BMGC with Vf 25% and d 58 8 mm) up to 27% strain (for the BMGC with Vf 25% and d 20 4 mm). Additionally, both the failure strain and fracture toughness exhibit a similar increasing trend with decreasing mean free path. It is apparent that the mean free path plays an important and controlling role on the toughening of the Mg-based metallic glasses. Therefore, we consider that the ductile porous Mo dispersoids provide a good toughening effect in promoting the practical use of amorphous materials. Acknowledgment It is gratefully to acknowledge the sponsorship by National Science Council of Taiwan, ROC, under the Project No. NSC95e2210E-214e015-MY3 and NSC98-2221-E-008-116- MY3. In addition, the authors also like to acknowledge the help on SEM and TEM analysis by the Micro and Nano Analysis Laboratory of I-Shou University. References
[1] Inoue A, Masumoto T. Mater Sci Eng A 1991;133:6. [2] Inoue A, Nakamura T, Nishiyama N, Masumoto T. Mater Trans JIM 1992; 33:937. [3] Wang WH, Dong C, Shek CH. Mater Sci Eng. R 2004;44:45. [4] Zeng Q, Nishiyama N, Inoue A. Scripta Mater 2006;55:711. [5] Ponnambalam V, Poon SJ, Shiet GJ. J Mater Res 2004;19:1320.

[6] Kawashima A, Zeng Y, Fukuhara M, Kurishita H, Nishiyama N, Miki H, et al. Mater Sci Eng A 2008;498:475. [7] Amiya K, Inoue A. Mater Trans JIM 2000;41:1460. [8] Men H, Kim DH. J Mater Res 2003;18:641. [9] Zheng Q, Ma H, Ma E, Xu J. Scr Mater 2006;55:541. [10] Ma H, Xu J, Ma E. Appl Phys Lett 2003;83:2793. [11] Pan DG, Zhang HF, Wang AM, Hu ZQ. Appl Phys Lett 2006;89:261904. [12] Xu YK, Ma H, Xu J, Ma E. Acta Mater 2005;53:1857. [13] Ma H, Shi LL, Xu J, Li Y, Ma E. Appl Phys Lett 2005;87:181915. [14] Brothers AH, Dunand DC, Zheng Q, Xu J. J Appl Phys 2007;102. 023508. [15] Jang JSC, Ciou JY, Hung TH, Huang JC, Du XH. Appl Phy Lett 2008;92. 011930. [16] Huang JC, Chu JP, Jang JSC. Intermetallics 2009;17:973. [17] Jang Jason SC, Jian SR, Li TH, Huang JC, Tsao Chi YA, Liu CT. J Alloys Compd 2009;485:290. [18] Jang Jason SC, Ciou JY, Li TH, Huang JC, Nieh TG. Intermetallics 2010;18:451. [19] Kinaka M, Kato H, Hasegawa M, Inoue A. Mater Sci Eng. A 2008;494:29. [20] Jang J SC, Chang YS, Li TH, Hsieh PJ, Huang JC, Tsao Chi YA. J Alloy Compd 2010;504S:S102. [21] Du XH, Huang JC, Hsieh KC, Lai YH, Chen HM, Jang JSC, et al. Appl Phys Lett 2007;91:131901. [22] Hofmann DC, Suh JY, Wiest A, Duan G, Lind ML, Demetriou MD, et al. Nat Lett 2008;451:1085. [23] Wu Y, Xiao YH, Chen GL, Liu CT, Lu ZP. Adv Mater 2010;22:2770. [24] Zhu Z, Zhang H, Hu Z, Zhang W, Inoue A. Scripta Mater 2010;62:278. [25] Chen HM, Lee CJ, Huang JC, Li TH, Jang JSC. Intermetallics 2010;18:1240. [26] Donohue A, Spaepen F, Hoagland RG, Misra A. Appl Phys Lett 2007;91:241905. [27] Guo H, Yan PF, Wang YB, Tan J, Zhang ZF, Sui ML, et al. Nat Mater 2007;6:735. [28] Wang YM, Li J, Hamza AV, Barbee TW. Proc Natl Acad Sci U.S.A 2007;104:11155. [29] Chen HM, Du XH, Huang JC, Jang JSC, Nieh TG. Intermetallics 2009;17:330e5. [30] Nieh TG, Wadsworth J. Intermetallics 2008;16:1156e9. [31] Lu ZP, Liu CT. Intermetallics 2004;12:1035. [32] Du XH, Huang JC, Liu CT, Lu ZP. J Appl Phys 2007;101:086108. [33] Meyers MA, Chawla KK. Mechanical behavior of materials. Cambridge, UK: Cambridge University Press; 2009. p.549. [34] He G, Eckert J, Loser W. Acta Mater 2003;51:1621. [35] Xi XK, Zhao DQ, Pan MX, Wang WH, Wu Y, Lewandowski JJ. Phys Rev Lett 2005;94:125510.

Você também pode gostar