Journal of Materials Science & Technology, 2020, 50(0): 192-203 DOI: 10.1016/j.jmst.2020.02.036

Research Article

Achieving work hardening by forming boundaries on the nanoscale in a Ti-based metallic glass matrix composite

Jing Fana,b, Wei Raoc, Junwei Qiao,a,b,*, P.K. Liaw4, Daniel Şopu5,6, Daniel Kiener7, Jürgen Eckert6,7, Guozheng Kangc, Yucheng Wu,a,**

aCollege of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan, 030024, China

bKey Laboratory of Interface Science and Engineering in Advanced Materials, Ministry of Education, Taiyuan University of Technology, Taiyuan, 030024, China

cKey Laboratory of Advanced Technologies of Materials, School of Mechanics and Engineering, Southwest Jiaotong University, Ministry of Education of China, Chengdu, 610031, China

dDepartment of Materials Science and Engineering, The University of Tennessee, Knoxville, IN, 37996 2200, USA

eTechnische Universitat Darmstadt, Institut für Materialwissenschaft, Fachgebiet Materialmodellierung, Otto- Berndt- Strasse 3, D-64287, Darmstadt,Germany;

fErich Schmid Institute of Materials Science, Austrian Academy of Sciences, JahnstraBe 12, A- 8700, Leoben, Austria

gDepartment of Materials Science, Montanuniversitat Leoben, JahnstraJSe 12, A-8700, Leoben, Austria

Corresponding authors: * College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan, 030024, China.E-mail addresses:qiaojunwei@gmail.com(J. Qiao),**wyc@tyut.edu.cn(Y. Wu).

Received: 2019-12-21   Accepted: 2020-02-20   Online: 2020-08-1

Abstract

Achieving work hardening in metallic glass matrix composites (MGMCs) is the key to the extensive use of these attractive materials in structural and functional applications. In this study, we investigated the formation of nanoscale boundaries resulted from the interaction between matrix and dendrites, which favors the work-hardening deformation in an in-situ Ti41Zr32Ni6Ta7Be14 MGMC with β-Ti dendrites in a glassy matrix at room temperature. The microstructures of samples after tension were observed by high-resolution transmission electron microscopy (HRTEM) and X-ray diffraction (XRD). The work-hardening mechanism of the present composites involves: (1) appearance of dense dislocation walls (DDWs), (2) proliferation of shear bands, (3) formation of boundaries on the nanoscale, and (4) interactions between hard and soft phases. A theoretical model combined with experimental data reveals the deformation mechanisms in the present work, proving that the in-situ dendrites with outstanding hardening ability in the glass matrix can provide the homogeneous deformation under tensile loading at room temperature.

Keywords: Metallic glass matrix composites ; Plastic deformation ; Work hardening ; Dense dislocation walls ; Nanoscale boundaries

PDF (5852KB) Metadata Metrics Related articles Export EndNote| Ris| Bibtex  Favorite

Cite this article

Jing Fan, Wei Rao, Junwei Qiao, P.K. Liaw, Daniel Şopu, Daniel Kiener, Jürgen Eckert, Guozheng Kang, Yucheng Wu. Achieving work hardening by forming boundaries on the nanoscale in a Ti-based metallic glass matrix composite. Journal of Materials Science & Technology[J], 2020, 50(0): 192-203 DOI:10.1016/j.jmst.2020.02.036

1. Introduction

Bulk metallic glasses (BMGs) possess the potential to be used as structural materials for a variety of applications due to their outstanding mechanical, chemical, and physical properties [1,2]. Unfortunately, they usually exhibit a brittle-fracture behavior under freestanding loading conditions at room temperature, such as upon the uniaxial tension [3]. Significant efforts have been made to overcome this shortcoming in recent years with the development of metallic glass matrix composites (MGMCs), consisting of ductile crystals precipitated in-situ in a glass-forming matrix [4]. The existence of crystals within the amorphous matrix hinders the rapid propagation of shear bands, and favors shear band blocking and multiplication. The formation of a highly organized pattern of multiple shear bands which, in turn, interact in a very complex way with each other and the crystals retards the early failure of MGMCs and ensures improved ductility.

Previous studies have introduced various MGMCs with excellent room-temperature tensile ductility [[5], [6], [7]]. Unfortunately, most of these composites exhibit macroscopic-strain softening with early onset of necking ascribed to a lack of work-hardening capacity, which restricts the structural applications for such kinds of in-situ MGMCs [8,13]. Because of this challenge, transformation-induced plasticity (TRIP) [7,9] and twining-induced plasticity (TWIP) [14] mechanisms have been employed in MGMCs through controlling their compositions, and concomitant work hardening has been realized in MGMCs. For example, B2-type MGMCs containing spherical polymorphous B2 CuZr/Ti precipitates [7,9] and β-type primary β-Zr/Ti solid solution dendrites embedded in a glassy matrix [[10], [11], [12]] are two common in-situ MGMCs. The metastable B2 CuZr phase undergoes martensitic phase transformation during plastic straining, which produces excellent room-temperature plasticity and work-hardening capacity through TRIP effect [7].

In addition to the work-hardening mechanisms of MGMCs under tension referred before, the novel in-situ Ti-based MGMCs, presented in this study, show distinct tensile ductility with significant work-hardening capability, which is ascribed to the formation of nanoscale boundaries at interface between the dendrites and the matrix. The generation of boundaries on the nanoscale is based on the formation of DDWs within the dendrites, and develops with the interplay between shear bands in the matrix and DDWs. With the help of finite element modeling (FEM), digital-image correlation (DIC), nanoindentation tests, as well as high-resolution transmission electron microscopy (HRTEM), it is well proved that the ductile dendrites with remarkable work-hardening capacity act as a prerequisite for achieving uniform elongation of the MGMCs. The deformation mechanisms are explored by accounting for the competition between the matrix and the dendrites upon tensile loading. The generation of dense dislocation walls (DDWs) and the development of boundaries on the nanoscale in dendrites during tension induce remarkable work- hardening behavior. On the other side, the nucleation and propagation of shear bands within matrix lead to softening. Moreover, a model has been developed to study the deformation behavior of the composites. The work-hardening coefficients “n” and “n'” obtained through the model indicate that the hardening originated from boundaries on the nanoscale in the dendrites and softening induced by shear banding within the glass matrix, respectively. The experimental results together with the developed constitutive relations clearly unveil the work-hardening mechanisms of such kinds of composites. Hence, the current findings extend the fundamental understanding of the deformation mechanisms of MGMCs upon tension, and pave a way to design MGMCs with significant uniform elongation at room temperature.

2. Experimental

The in-situ Ti41Zr32Ni6Ta7Be14 [atomic percent (at.%)] MGMCs specimens were prepared by arc melting the mixture of pure elements [purity >99.9 wt percent (wt.%)], Ti, Zr, Ni, Ta, and Be, under a Ti-gettered argon atmosphere. The ingots were re-melted at least four times in a high-purity argon atmosphere to ensure chemical homogeneity. Copper-mold-casting was used to fabricate plate-shape samples. The size of the samples was 2 mm in thickness and 10 mm in width. The structures of phases in the as-cast ingots were characterized by X-ray diffraction (XRD; Ultimalv) with Cu-Kα radiation. Microstructures and chemical compositions were analyzed by scanning electronic microscopy (SEM; TESCAN, LYRA 3 XMH) combined with energy-dispersive spectrometry (EDS). The tensile samples, which have been cut into a shape of dog -bone-like (10 mm (length) ×2 mm (width) ×2 mm (thickness)). The tensile experiments at room temperature were conducted with an Instron 5969 testing machine at a strain rate of 5 × 10-4 s-1. To confirm the reproducibility of the tests, at least three specimens were tested. The evolution of local strains during the experiments was detected by digital image correlation (DIC) using the MatchID system. For these tests, white paint was sprayed on one side of the samples and then coated with black spots in order to increase the contrast of the images taken during the DIC experiments. XRD, SEM, transmission electron microscopy (TEM), and HRTEM in a JEM-2010 microscope were performed to check the failed samples. Mechanical grinding was the first step when preparing the TEM specimens, followed by ion milling (ion energy: 4.5-5 keV) with Gatan 691 Polisher. The hardness and Young’s modulus required in the present work were calculated from nanoindentation tests (MTS Systems, USA). Computational micromechanics investigation on the formation and evolution of deformation bands have been developed in the current work. The deformation behavior of monolithic BMGs is generally described by the free-volume model and the mechanism-based strain-gradient plasticity model [[15], [16], [17]], which are described in a model with a finite element code (ABAQUS) as user material subroutines. The shear bands in the glass matrix and the slip bands in the dendrites, which are induced by the concentration of free volumes and pile-ups of dislocation, respectively, are internal state variable SDV1 in the code. In the current work, a two-dimensional (2-D) plane-strain finite element model (FEM) was chosen. As in a similar model in previous work by Rao et al. [15,17], the loading boundary in this model was perpendicular to the loading direction. The normal constraints appear opposite to the loading boundary, and the full constraints appear at a vertex of the opposite boundary.

3. Results

Fig. 1 shows the result from SEM observations for the as-cast Ti41Zr32Ni6Ta7Be14 composites. In-situ dendrites are continuously distributed within the glassy matrix, as denoted by the arrows. The volume fraction of the dendrites and the average width of the dendrite arm were evaluated by the Image pro software from the SEM images, and the values are 49 ± 3 vol.% and 1-2 μm, respectively. In order to clarify the chemical compositions of both phases, the content of the elements Ti, Zr, Ni, and Ta was acquired by EDS analysis. Be is hard to detect precisely, since it is very light and completely dissolved in the amorphous matrix [18]. Table 1 presents the chemical compositions of both phases. The values given for the composition in the first line correspond to the nominal composition. The glass matrix is enriched in Zr and Ni and the dendrites in Ti and Ta, similar to others developed in-situ Ti-based MGMCs [10].

Fig. 1.

Fig. 1.   SEM image of the as-cast composite.


Table 1   Contents of different elements in the present Ti-based MGMCs (in at.%).

ZoneTiZrNiTaBe
Composite41326714
Matrix34.8 ± 0.247.2 ± 2.112.4 ± 1.85.6 ± 0.2-
Dendrites47.2 ± 0.534.8 ± 2.04.0 ± 1.714.0 ± 0.4-

New window| CSV


The TEM images of an as-cast sample are shown in Fig. 2 to reveal the details of the microstructure. As can be found from Fig. 2(a), bright elliptical particles and gray regions correspond to the dendrites and matrix, respectively. The selected-area electron diffraction (SAED) patterns displayed in Fig. 2(b) and (c) correspond to the yellow and green regions in Fig. 2(a), respectively, and convincingly demonstrate the formation of the body-centered-cubic (bcc) β-Ti dendrites and the amorphous matrix. Besides the diffraction spots from the β-Ti solid solution, additional weak diffraction spots indicated by red circles, corresponding to the red circles in Fig. 2(a), at 1/3 {112}β and 2/3 {112}β are visible [19]. Among several binary and multi-component Ti and Zr alloys, the ω phase may appears when the stability of the β phase is enhanced by adding β stability elements, such as Mo, Co, and Ta [20]. Consequently, these spots can be identified as ω-Ti, which is formed by a displacive shuffle dominated mechanism during rapid cooling [20]. However, the structure of ω-Ti cannot be observed in XRD and SEM, which means the amount of ω-Ti is very small or can be almost neglected. Fig. 2(d) shows the HRTEM image of the as-cast sample, which displays the interface between the dendrites and the matrix. It can be seen that the atomic bonding, shown by the red line, between the two phases are very good, no obvious defects and stress concentration happened in this place. The deformation mechanisms are closely related to the intrinsic composition of the composites. In particular, special attention has been directed towards the Ti-based alloys developed by adding β-phase or α-phase stabilizing elements [20]. In view of MGMCs, thermos-mechanical processing of Ti-based alloys (the dendrites) involves deformation in the different phases depending on the compositions and the change of the β-phase stabilizer (Ta) in the dendrites affects the phase stability, which influences the structure and deformation mechanisms of MGMCs upon loading [20,21].

Fig. 2.

Fig. 2.   TEM images of the as-cast sample: (a) bright-field TEM image, selected-area electron diffraction (SAED) patterns of the dendrites and the matrix are presented in (b) and (c), respectively, (d) the HRTEM image of the interface.


The XRD patterns of as-cast and deformed samples are shown in Fig. 3. Sharp peaks, denoting crystalline phases, can be found to be superimposed on a broad diffuse diffraction background corresponding to the amorphous phases, indicating the concomitant existence of bcc β-Ti and amorphous phases, respectively. The XRD patterns after tensile deformation exhibit the same set of diffraction peaks and no additional peaks, confirming that there is no phase transformation upon plastic deformation.

Fig. 3.

Fig. 3.   XRD patterns of the composite before and after tensile deformation.


Fig. 4(a) displays a characteristic tensile true stress-true strain curve of the composite at ambient temperature. Apparently, the MGMC exhibits a distinct work-hardening behavior. It yields at 1200 ± 45 MPa, followed by a region with increasing flow stress, and finally reaches an ultimate tensile strength (UTS) of 1640 ± 37 MPa and a total uniform strain of 5%. The plastic deformation is susceptible to localized deformation (necking), and the Considére criterion can be used to detect necking [22].(1)dσdε≤σwhere, σ, is the true stress, and, ε, is the true strain. The work-hardening rate (dσ/dε) calculated from the true stress-true strain curve is shown in the inset of Fig. 4(a). Comparing these two curves in Fig. 4(a) reveals that necking does not happen until final fracture, and the value of the work-hardening rate exceeds the tensile stress during the whole plastic deformation stage. In other words, the significant work hardening without macroscopic necking dominates when the composite is subjected to tensile straining. The different slopes shown in the C-J plot (dσ/dε vs. ε) provide insight into possible changes in the deformation mechanisms. Multi-stage work-hardening behavior, which is common in crystalline alloys with remarkable work-hardening, is evident in the present composite [Fig. 4(a)] [23,24]. The fracture morphology of the samples after tension is displayed in Fig. 4(b). Profuse shear bands indicate the large plastic deformation for the present MGMCs. The inset in Fig. 4(b) presents an image of a fractured sample. Apparently, the tensile failure of the current composite occurs in a shear fracture mode. Three fracture angles (57.7°, 77.6°, and 90°) can be detected on the fracture surface. Zhang et al. have reported that the tensile fracture angle is 54° in monolithic Zr-based BMGs [25]. The tensile fracture angles are larger in the present MGMC, indicating that both normal and shear stresses act on the fracture plane.

Fig. 4.

Fig. 4.   True stress-true strain curve of the composite (a), the work-hardening rate curve is shown in the inset. (b) SEM images of a sample after fracture.


The detailed deformation structures in the fractured samples after tension are presented in Fig. 5. Fig. 5(a) and (b) shows the TEM images of a fractured specimen. Compared with as-cast sample shown in Fig. 2(a), DDWs are frequently found within the dendrites along the interface. Fig. 5(c) presents a HRTEM image taken from the region near the DDWs, where located at the boundary between two phases, represented by red lines in the figure. The red line close to the part of the matrix is the boundary, and the region between two lines belong to the dendrites. It indicates that an irregular nanoscale boundary takes place of the original regular interface (Fig. 2(d)) between two phases during tension. The fast Fourier transform (FFT) patterns inserted at the left-hand bottom side of Fig. 5(c) correspond to the areas of the glass matrix and dendrites, denoted by the squares A and B, respectively. Fig. 5(d) displays a HRTEM image of the glass matrix taken from an area, where a shear band (the light region bordered by red lines) was coincidently caught. The shear band is blocked by a crystal (cf. the ordered lattice within the yellow ellipse).

Fig. 5.

Fig. 5.   TEM images of a sample after fracture: (a) bright-field image, (b) dark-field image, (c) HRTEM image of the interface, (d) a shear band in the matrix, (e) and (f) HRTEM images of the dendrites.


In crystalline alloys, slips and dislocations are the major carriers of plasticity. In contrast, the plastic deformation in a glassy phase is accommodated by shear bands. Shao et al. have reported that two-zone shear bands (one is a solid-like zone, and the other a liquid-like zone) control the plasticity of metallic glasses [26]. This trend can be inferred from the structure in Fig. 5(d). The generation of a liquid-like zone (the area between the red lines), surrounded by a solid-like zone, is prevented by the crystalline dendrites (the area within the yellow ellipse), which retards the plastic instability. The crystals in MGMCs can hinder or distribute the expansion of shear bands within the matrix. Fig. 5(e) and 5(f) displays magnified HRTEM images of the dendrites. The boundary on nanoscale formed close to the interface is indicated in Fig. 5(e), corresponding to the area between red lines in Fig. 5(c). Dislocations, denoted by “T”, severe lattice distortion, represented by yellow circles, and diverse lattice orientations are simultaneously present. The SAED pattern in Fig. 5(e) shows clear diffraction lattice of nanocrystals, which states the structure of the boundary on nanoscale have changed from ordered crystalline lattice to sub-grains and nanocrystal structure. Comparing the HRTEM images of as-cast and fractured samples, as shown in Fig. 2(d) and 5(c), respectively, significant changes on the interface between the dendrites and the matrix have taken place. It indicates that the nanoboundary, consisting of sub-grains and nanocrystal, is formed during tension. The microstructure observed by TEM shown in Fig. 5(e) demonstrates the existence of sub-grains and nanocrystals near boundaries in fractured samples. It is noted that the nanoscale boundary, corresponding the area between red lines in Fig. 5(c), is generated upon straining. Fig. 5(f) depicts the area far away from the boundary. Lattice distortions and dislocations can be observed, corresponding to the inverse FFT image, shown at the left bottom of the figure. The inset in the upper right corner displays the corresponding SAED pattern, which demonstrates that the interior of the dendrites retains the bcc structure for β-Ti solutions, and no phase transformation happens. Consequently, boundaries on the nanoscale are generated within the dendrites close to the interface. The boundaries possess disparate crystalline structure, i.e., neither structure of the glass matrix nor bcc β-Ti of dendrites.

Fig. 6 illustrates the strain increment distribution curve upon tensile loading, as obtained by DIC. Fig. 6(a) shows the strain-distribution maps. And the location of fracture, marked by blue rectangle in Fig. 6(a). The inset shown in Fig. 6(b), was selected for analysis of the temporal strain. The instantaneous strain distribution along loading direction, extracted from the strain-distribution map near the location of fracture, is presented in Fig. 6(b). Obvious necking cannot be found along the gauge length, and the strain increases uniformly during tension except for two inflection points on the instantaneous strain curve at local strains of 0.7% and 2.7%. The above results signify that shear bands generated locally within the glass matrix are arrested by the ductile dendrites, resulting in macroscopically homogenous elongation for the current MGMC.

Fig. 6.

Fig. 6.   DIC images of the deformed composite: (a) the strain distribution with time, (b) the corresponding temporal strain curve.


In order to better understand the deformation behavior, finite element modeling (FEM) was performed. The materials parameters of the glassy and dendrite phases used in the simulation are listed in Table 2. Em and Ed are elastic modulus of the matrix and dendrites, respectively, νm and νd are Poisson ratio of the matrix and dendrites, respectively, α, ξ0, χ, v*, T, t0-1, and τ0 are the geometrical factor, initial free volume concentration within the matrix, geometrical parameter of the matrix, critical volume for the matrix, absolute temperature, characteristic time, and the reference stress of the matrix, respectively. M, b, α, dp, ψ, ε˙0, k20, n, NB, ξ, m0, and σy are Taylor factor, Burgers vector, empirical constant, effective particle size, proportional factor, reference strain rate, dynamic recovery constant, power exponent relating to the dislocation annihilation, maximum number of dislocation loops, viscosity coefficient, and yield strength of the dendrite, respectively.

Table 2   Material parameters used for the FEM model of the current MGMCs.

MatrixEm (GPa)νmαξ0χv* (ÅÅ)T (k)t0-1 (s-1)τ0 (MPa)
109.30.330.80.04651.2520300324414
DendritesEd (GPa)νdM1,3M2b (nm)αdp (μm)ψε˙0 (s-1)k20nNBξ (nm)m0σy1,2 (MPa)σy3 (MPa)
670.361.02.40.2560.332.50.2118.512.54535109001200

New window| CSV


The inhomogeneous distribution of the stress field is ascribed to the different interactions between two phases when choosing different dendrites. The propagation of shear bands and cracks strongly depends on the mechanical property of the dendrites. In the present FEM model, as shown in the Table 2, the same parameters for the glass matrix were developed and three varied dendrites were chosen as the reinforce phase in the FEM, named by dendrites 1, 2, and 3, respectively. Among three dendrites, materials parameters remain constant except for the M and σy. Larger Taylor factor, M2, was set for the dendrites 2 comparing with the dendrites 1 and 3 (M1,3), which ensures work hardening achieved in dendrites 2 upon tension. Higher yield strength, σy3, was developed for the dendrites 3 comparing with the dendrites 1 and 2 (σy1,2), which intends to observe the deformation behavior of the kind of stronger dendrites. The results indicate that work hardening is not easily achieved even when the reinforce dendrites possess very high strength. And the work hardening results only when the dendrites have remarkable work-hardening capacity. Consequently, the simulation results obtained using dendrites 2 present a similar deformation behavior as the current MGMCs.

The uniaxial tensile stress-strain curves of the in-situ MGMC are predicted and compared with the experiments in Fig. 7. The comparison shows a good agreement, indicating that the adopted finite element model is suitable to numerically depict the tensile deformation behavior of the current MGMC reasonably well. Fig. 8(a), (b), and (c) illustrates the evolution of free volume in the glassy matrix, the dislocation density in the dendrites, and the local stress distribution at a strain of 3%. The corresponding values for the late deformation stage, 7%, are shown in Fig. 8(d), (e), and (f), respectively. It is suggested that homogeneous deformation can be found in the present MGMC at the first time when yielding occurs, as displayed by Fig. 8(a), (b), and (c). In contrast, the free volume and the dislocation density rapidly increase at 7% strain, which reveals work softening and hardening in the matrix and dendrites, respectively. What can be observed from Fig. 8(d) and (e) is that the region containing a high concentration of free volume in the matrix [Fig. 8(d)] is adjacent to a region, where the dislocation density is relatively high within the dendrites [Fig. 8(e)]. This trend confirms that the stress concentration at the interface has an effect on the generation of slip bands in the dendrites and on the initiation of shear bands in the amorphous matrix [5]. Fig. 8(f) depicts a map of the strain distribution with respect to the loading direction. Shear localization is not significant. The plastic deformation of the current MGMC under tension is relatively homogenous, as demonstrated in Fig. 8(f). These findings are close to the experimental results.

Fig. 7.

Fig. 7.   True stress-true strain curves obtained by the experiment and FEM, respectively.


Fig. 8.

Fig. 8.   FEM results obtained at the strain of 3% and 7%: (a) and (d) the evolution of free volumes within the glassy matrix, (b) and (e) the dislocation density in the dendrites, (c) and (f) the local stress distribution.


In addition, another two dendrites were selected to make a comparison with dendrites 2 through the FEM analysis, as shown in Fig. 9. Fig. 9 (a) and (b) demonstrates the results, (a) displaying the stress-strain curves for varied dendrites, and (b) presenting the stress-strain curves of MGMCs with the corresponding dendrites in (a). It is suggested that dendrites 2 exhibit excellent work hardening during deformation, and no hardening behavior can be found in the other two types of dendrites. Consequently, only MGMC 2, with in-situ formed dendrites 2 in the amorphous matrix, demonstrates work hardening. For the other two soften dendrites, dendrites 1 with lower yield strength and dendrites 3 with higher one, cannot assist appearance of work hardening in MGMCs, which means that the work-hardening behavior of the dendrites during tension is the necessary factor for achieving work hardening in the corresponding MGMCs. Fig. 10 illustrates the evolution of the free volume within the glassy matrix, the dislocation density in the dendrites, and the local stress distribution at 5% strain for three the MGMCs. Fig. 10(a), (b), and (c) are results of MGMC 1, (d), (e), (f) represent MGMC 2, and (g), (h), and (i) corresponds to MGMC 3. The evolution of free volumes is similar among the three MGMCs [(a), (d), and (g)], which means that the matrix undergoes softening. Obviously, a higher dislocation density forms in MGMC 2 [(b), (e), and (h)]. It is noted that the local stress distribution is more homogenous in MGMCs 2 and 3 when compared to MGMC 1.

Fig. 9.

Fig. 9.   FEM analysis for another two dendrites: (a) stress-strain curves of different dendrites, (b) stress-strain curves of MGMCs with the corresponding dendrites in (a).


Fig. 10.

Fig. 10.   Evolution of free volumes in the glassy matrix, dislocation density in the dendrites, and local stress distribution at the strain of 5% for three MGMCs: (a), (b), and (c) MGMC 1, (d), (e) and (f) MGMC 2, (g), (h), and (i) MGMC 3.


4. Discussion

It is well known that the introduction of dendrites into the matrix effectively improve the ductility of BMGs, and the volume friction and size of the dendrites influence the deformation behavior of MGMCs [16]. For the same system MGMCs, the reinforced dendrites act as an important role in deformation mechanism of MGMCs. In other words, the work hardening of MGMCs can be achieved when the reinforced dendrites possess excellent work-hardening ability. In the present MGMC, the dendrites show significant work hardening upon tension, which is induced by the formation of DDWs. The element of Ta works as stabilizer of a β phase in titanium alloys, and TRIP and TWIP effects on the work hardening upon tension among Ti-based MGMCs have been studied [10,14]. Based on the previous studies, fine adjustment of the proportion of Ta and other elements has been conducted. The sub-grain and nanocrystal generated at the interface between two phases is an intermediate state. The composition of the dendrites within MGMCs may determine the deformation mechanism, i.e., the formation of boundary on nano scale stems from the intrinsic property of the present MGMC.

The results described above indicate that boundaries on the nanoscale observed during tension plays an important role in achieving work hardening. Different from TRIP and TWIP deformation mechanisms, the work hardening observed in the present MGMCs is induced by the formation of special nanocrystalline structure. Zhai et al. [27] designed Ti-based MGMCs through adding Sn to tailor the intrinsic properties. Significantly ductile deformation and excellent work hardening can be achieved with Sn addition, totally due to dislocation multiplication, similar to that usually found among MGMCs with work softening [4,5,42]. Actually, for the MGMCs, asynchronized deformation might happen for both phases. Previously, it has been found that a disparity of the elastic limits, Young’s moduli, and yield strengths exist for both phases in such kinds of composites [12,28]. Consequently, the dendrites yield preferentially, and plastic deformation starts first within the dendrites, when the matrix is still upon elasticity. Meantime, many DDWs generated in the dendrites begin to develop. Simultaneously, shear bands are poised for the activation and propagation within the matrix. Instead, multiple shear bands with limited length given by the constraints of the crystalline phase form. After both phases enter the plastic deformation stage, the interaction between shear bands and DDWs assists the development of boundaries on the nanoscale at the interface.

Different from the studied by Zhai et al. [27], the moving and pile-up of dislocations in the dendrites induce the formation of DDWs at boundary of the dendrites in the present MGMCs. Furthermore, the interaction between the shear bands and DDWs causes stress concentration at the interface. Since the nanocrystalline structure stems from the interaction between two phases, and is located at the interface. It is clear that the present MGMCs could be divided into three different structures due to the formation of new boundaries after tension. During tension, the work hardening occurred in the dendrites and softening happen in the matrix control the tensile deformation behavior of the dendrites and the matrix, respectively. It results in the intrinsically soft dendrites becoming hard ones, and the relatively hard glass matrix turning into a soft phase. The dendrites show significant work hardening which effectively offsets the work softening of the matrix, and ultimately achieved the excellent work hardening.

4.1. Dislocation pile-ups

Plastic deformation firstly occurs in soft dendrites as soon as multiple dislocations operate during tension. The DDWs in the plastically-deformed dendrites indicate a continuous pile-up of dislocations upon successive straining. Only when dislocation pile-ups develop, i.e., dislocation movement is effectively impeded by grain boundaries [29], twins [14], or interfaces (phase boundaries) [30], successive work-hardening will be available.

To better understand the deformation mechanisms, FEM has been employed to study different parameters corresponding to various dislocation densities in the dendrites. Three defect densities (1 × 107, 5 × 1014, and 1.5 × 1015/m2) were chosen for the FEM simulations.

Fig. 11 shows the calculated stress-strain curves for different dislocation densities, and the inset shows the corresponding work-hardening rate curve. The results indicate that increasing dislocation densities leads to pronounced work hardening. Fig. 12 displays maps of the dislocation density (left column), the free volume density (center) and strain contours (right) for MGMCs with different dislocation densities in the dendrites at a strain of 5%. The top row corresponds to a dislocation density of 1 × 107/m2 in the dendrites, while the middle row displays the results for a dislocation density of 5 × 1014/m2. The bottom images in Fig. 12(g-i) represent the findings for a defect density of 1.5 × 1015/m2. Consistent with our expectations, successive dislocation multiplication and pile-up contribute to the distinct work-hardening in the current composite. Usually, pile-up of dislocations is not achieved and shear banding prevails, accompanied by softening after yielding in the other in-situ dendrite-reinforced MGMCs [8,18]. In other words, increasing and pile-up of dislocation are important reasons of the work hardening in such kinds of composites, which is ascribed to the significant work hardening of the dendrites.

Fig. 11.

Fig. 11.   Calculated true stress-true strain curves of MGMCs with different dislocation densities in the dendrites as obtained by FEM, the inset corresponds to work-hardening curves.


Fig. 12.

Fig. 12.   Free volume density and maps of strain contours of the MGMCs with different defect densities in the dendrites at a true strain of 5%: (a), (b) and (c) 1 × 107/m2, (d), (e) and (f) 5 × 1014/m2, (g), (h) and (i) 1.5 × 1015/m2.


4.2. Work-hardening behavior during tension

As reported before, three stages can be found in the deformation stage of MGMCs, i.e., elastic-elastic, elastic-plastic, and plastic-plastic stages [31]. Coincidentally, the tensile strain hardening rate of the present MGMC can be accordingly divided into three stages, as displayed in the inset of Fig. 4(a) and marked as A, B, and C stages. A similar situation can be often found in a variety of crystalline alloys, including Mg-alloys [32], alloys exhibiting transformation-induced plasticity (TRIP) [33], twining-induced plasticity (TWIP) steels [34], and high-entropy alloys (HEAs) [35]. The work-hardening rate first decreases, then remains constant (stage A), and subsequently decreases again (stages B and C) with increasing strain.

At stage A, the dendrites in the present composites have entered the plastic range when the glass matrix still is in the elastic range. The first decrease of the work-hardening rate occurs at the initial state when the dendrites in the composite yield (i.e., the work-hardening rate enters the non-linear deformation stage), which is governed by dislocation glide within the crystalline dendrites. During this stage, the dislocations generated in the dendrites move towards the interface, and pile-up near the interface, as demonstrated in Fig. 5(a) and (b). Meanwhile, dislocations within dendrites can diffuse into the matrix and become absorbed by the matrix. The work-hardening rate in stage A is controlled by the successively increasing dislocation density, effectively restricting dislocation slip. It is noted that the start of stage A at the strain of 0.7%, corresponds to the first inflection point shown in Fig. 6(b), which indicates the yield point of the dendrites. The other point in Fig. 6(b) is regarded as the stage when profuse formation of shear bands in the glass matrix sets in, i.e., the beginning of work softening. At the same time, the multiple dislocations generated within the dendrites move towards the phase boundary, forming DDWs. Loosely bonded free volume regions, which act as flow defects in the glass matrix, can continuously accumulate before the occurrence of plastic deformation. It is noted that softening (i.e., the generation of shear bands) in the glass matrix can be ignored during the elastic-plastic stage. Hence, the plateau of the work-hardening rate in stage A corresponds to stage Ⅱ in traditional crystalline alloys [22]. It provides an extended work-hardening capacity stretching over about 22% of the whole plastic deformation stage, in a good agreement with crystalline alloys [36]. DDWs and free volume regions appear synchronous near the phase boundary, and cause the formation of the heterogeneous structure at the interface.

After stage A, the propagation of shear bands becomes prevailing at the whole stage B. It has been speculated that the glass matrix generally exhibits softening due to localized structural and thermal shearing [1,4]. Accordingly, the work-hardening rate begins to decrease. Nevertheless, the overall deformation behavior remains hardening at this stage. It should be emphasized that the blocking of work-hardened dendrites to the prompt propagation of shear bands is significant, as demonstrated in Fig. 4(b) and 5(d).

With even more plastic straining the dislocation density reaches a dynamic equilibrium state. Dislocation rearrangement becomes active, which is one of the sources for grain boundary migration and the generation of nanocrystals [29,37]. As a result, the interface between the two phases forms a unique boundary structure. The nanoscale boundary in the current work originates from a hierarchical microstructure, not a chemical composition heterogeneous structure, as investigated in Huang’s work [38]. The HRTEM and SAED images in Fig. 5(e) show the existence of nanocrystals, which suggests that the interface evolves into boundaries on the nanoscale, and the inner part within dendrites retains the original structure. The heterogeneous structure induced by the formation of special boundaries is particularly beneficial and has an overarching effect on maintaining work hardening in stage B [39]. On one hand, plasticity gradients are built up, which contributes to blocking of the dislocation movement, accompanied by continuous work hardening in dendrites. On the other hand, as reported by Wang et al. [40], the interface between both phases exhibits unique inelastic shear-transfer characteristics, where the glass matrix acts as high-capacity sink enabling the absorption of plasticity. As overall effect, both the successive dislocation pile-up within dendrites and the profuse shear bands in the glass matrix accommodate the plasticity.

Stage C is the final stage, where the work-hardening rate is lower than the true stress [Fig. 4(a)]. Plastic instability occurs at this stage. On the basis of the results in stage B, the glass matrix undergoes pronounced softening upon plastic deformation, which is attributed to shear-induced local dilatation [41]. In contrast, the dendrites are strongly hardened. The boundaries on the nanoscale separate the matrix and the dendrites, and trilaminar heterogeneous structure forms. Meanwhile, the dislocation density generated in the dendrites saturates, which means that defects, such as cracks, have formed in the dendrites at this stage. Hence, multiple mature shear bands appear in the matrix and the dendrites cannot stop them. At this stage, the work-softening mechanisms generated in the glass matrix become more pronounced, while the work-hardening mechanisms generated in the dendrites weaken.

The hardness of the matrix and the dendrites before and after tension have been measured by nanoindentation tests. The Berkovich diamond indenter tip penetrates into the matrix and dendrites, respectively, when conducting nanoindentation tests, and the maximum indentation depth was 1000 μm. At least ten points were selected during the nanoindentation experiments. The results are summarized in Table 3, which verifies the distinctly different deformation behavior of the two phases. The initial hard (glass matrix)/soft (dendrites) model of the MGMC changes into a soft (glass matrix)/hard (dendrites) model after tension, which is caused by the particular deformation mechanisms described above. The excellent work hardening in the dendrites and obvious work softening in the matrix induce the change of hard and soft phases. The nanoscale boundary generates during the process of competition between the dendrites and matrix, which induces the work-hardening behavior occurring in the present MGMC. No macroscopic necking is found, since the composite fails immediately when the plastic instability is reached. On the whole, homogenous plastic elongation is dominant upon tensile loading.

Table 3   Hardness of the glass matrix and the dendrites before and after tensile testing measured by nanoindentation.

Type of specimenMatrix (GPa)Dendrites (GPa)
As-cast6.96 ± 0.214.58 ± 0.45
Deformed6.18 ± 0.136.22 ± 0.27

New window| CSV


4.3. Constitutive relationship

On the basis of the present and previous findings [14], some implications can be drawn. A transition of different tensile deformation mechanisms corresponding to the various stages upon deformation occurs in the current composites.

At the beginning, both phases of the current composite are in the elastic state, and attempts to derive the Young’s moduli for each phase and the composite have been made. Generally, the average Poisson ratio of most BMGs can be assumed as a similar value to crystalline phases [4,42,43]. Hence, for the further considerations the Poisson ratio here is set for 0.33 in both phases [44]. One of the attempts can be established according to Hashin and Shtrikman, as follows [45]:(2)Ec=Em1+fvEd-Em1-fvβEd-Em+Em,where Ec, Em, and Ed are the Young’s moduli of the composite, the matrix and the dendrites, respectively, the value of fv is 0.49, which represents the volume fraction of dendrites, β is constant and β=8-10νm151-νm, and νm (0.33) is the Poisson ratio of the matrix. Em and Ed are equal to 109 and 83 GPa, respectively. Therefore, Ec can be calculated as 95.5 GPa by inserting these parameters into Eq. (2). Another approach considers simple rule of mixtures (ROM),(3)Ec=Em1-fv+Edfv.

The value of Ec is 96.3 GPa, similar to the results obtained by Eq. (2).

With increasing strain, the soft dendrites yield first. On the basis of the ROM, the stress of the composite can be established as:(4)σc=fvσd+1-fvσm,where σc, σd, and σm are the tensile stresses of the composite, dendrites, and glass matrix, respectively. An overall strain (ε) of 1.5%, a median of this stage, is chosen as an input into Eq. (4), which equates to: 1-fvEmε=1-0.49×109×0.015=833 (MPa). Referring to Fig. 4(a), the stress at the strain of 1.5% is about 1200 MPa. As a result, σd can be calculated as 749 MPa.

The constitutive equation of the composites at this stage can be expressed according to [46]:(5)σ=13cd3σd+3Em1-β1-fνfvεp3,where cd and cm are the average stress-concentration factors of the dendrites and the matrix, respectively (cm=βEd-Em+Emfv+1-fvβEd-Em+Em, cd=Edfv+1-fvβEd-Em+Em). εp is the plastic strain of the composites, which is given as εp=fνcdεdp, and εdp is the plastic strain for dendrites. Consequently, Eq. (5) can be written as:σ=σd+0.28Emεdp. As presented in Fig. 6(b), the yield strain of the dendrites is about 0.7%, and this yield strain of the dendrites in the current MGMC is highly consistent with other Ti-based MGMCs [4,14]. Consequently, the plastic strain of the dendrites, εdp, at the overall strain of 1.5% is about 0.8%. According to Eq. (5), the stress acting on the dendrites at the overall strain of 1.5% is calculated as 950 MPa, very close to the stress in dendrites in other Ti-based MGMCs and Ti-based crystalline alloys [4,22]. The model in Eq. (5) is more suitable for interpreting the stress conditions at this stage. The yield stress for the dendrites is about 580 MPa, according to Hooke’s law. The work-hardening exponent,n, is estimated to be 0.66.

Both phases undergo plastic deformation at this stage when the glass phase begins to yield. This stage corresponds to the decrease of the work-hardening rate, as shown in Fig. (4 b). For a dislocation-free crystal, it has been postulated by Frenkel that the theoretical shear strength can be assumed to be τy≅G/5. Therefore, the yield strength of the glass matrix can be calculated as 1640 MPa (σym=2τy) [47]. The tensile stress of the current composite in this stage is given by [46]:(6)σ=13cm3σym-3Em(1-β)εp3,and Eq. (6) can be simplified as:σ=σym-0.524Emεp. In another way, the tensile stress-strain relation of the matrix in the stage C can be represented as [46]:(7)σ=σref(σym/Em+εp)n',where σref is the reference stress of the matrix during uniaxial tension, and σref=Emn'σymn'-1, n' is the work-hardening coefficient of the matrix. The work-softening strain of the stage C can be inferred at a strain of 3.0% [Fig. 4(a)], which is close to the second inflection point displayed in Fig. 6(b). Combining Eqs. (6) and (7), the work-hardening coefficient of the matrix at this stage can be calculated as - 0.069. This trend indicates that the work-hardening in the dendrites is more pronounced than the softening induced in the glass matrix, which results in the excellent work hardening in the current composite. Such a situation is in line with the experimental data (cf. Table 3). The introduction of ductile dendrites with significant work hardening is an effective route to achieve work hardening upon tension of the MGMCs, the hardening of the dendrites can surpass work softening of the matrix. During tension, the hardness of the dendrites and the matrix show opposite trends, the hardness of the dendrites increase while the hardness of the matrix decrease. In the present MGMCs, Table 3 reveals that the hardness of the matrix decreases while that of the dendrites increases upon tensile deformation. The dendrite hardness becomes higher than that of the glass matrix, suggesting that the work-hardening effect overcomes the softening effect. In the present MGMCs, the hardness of the dendrites is higher than that of the matrix after tension. Hence, the ratio of the hardness values of the dendrites to the matrix for a variety of MGMCs with work-hardening ability (Hd/Hm≥1), is exhibited in Fig. 13 (the literature data were taken from Refs. [[48], [49], [50], [51]]). It means that dendrites with a higher hardness than the matrix are beneficial for work hardening in composites. However, the excellent work-hardening ability of the MGMCs depend on the work hardening of the dendrites. It is note that work hardening can be also achieved among the MGMCs even when the hardness of the dendrites lower than that of the matrix after tension [27].

Fig. 13.

Fig. 13.   Ratios of the hardness of the dendrites (Hd) and the matrix (Hm) for different MGMCs.


Furthermore, the FEM results obtained above also confirm the conclusions. The deformation behavior of dendrite 1-3 can be regarded as different deformation stages of MGMC 2 during tension. Dendrite 1 presents the initial state, and dendrite 3 correspond to the final state induced by work hardening (dendrite 2). The MGMC 1 with soft dendrite 1 displays work softening and a shear-localization distribution occurred during deformation [Fig. 10 (c)]. On the other hand, the MGMC 3 with the hard dendrite 3 shows similar local stress distributions [Fig. 10 (f) and (i)], which indicates that the hard dendrite phase within the matrix makes MGMCs hardening. The MGMC 2 in FEM is based on the present composite, showing dendrites change from soft to hard phase during tension. And the corresponding work hardening will be achieved at the same time. Hence, it is reasonable to rationalize that the work-hardening capacity in the present composite is a result of the pronounced hardening of the dendrites.

5. Conclusion

In the previous work, the deformation mechanisms of the Ti41Zr32Ni6Ta7Be14 MGMC have been revealed and characterized by experiments and theoretical calculations. The work-hardening behavior in the dendrites is ascribed to the formation of boundaries on the nanoscale, which are induced by generating DDWs during tensile deformation. On the other hand, softening in the matrix is induced by the formation of shear bands. Combining these factors, the stress condition of the current composites has been studied step by step according to the three deformation stages, thus revealing an overall homogeneous plastic deformation mechanism. Obviously, the work-hardening ability of the dendrites overcomes the softening of the glass matrix and leads to a significantly homogeneous tensile elongation.

Acknowledgements

J.W.Q. acknowledges the financial supports of the National Natural Science Foundation of China (No. 51371122), the Natural Science Foundation of Shanxi Province, China (No. 201901D111105), and the Transformation of Scientific and Technological Achievements Programs of Higher Education Institutions in Shanxi (2019). P.K.L. is very grateful for the support from the National Science Foundation (DMR-1611180 and 1809640) with the program directors, Drs. G. Shiflet and D. Farkas. D.Ş. acknowledges the financial support by the German Science Foundation (DFG) through the grant SO 1518/1-1. J.E. and D.Ş. would like to acknowledge the support through the ERC Advanced Grant INTELHYB (grant ERC-2013-ADG-340025). D.K. would like to acknowledge the support by the European Research Council under the ERC Grant Agreement (No. 771146 TOUGHIT).

Reference

W.L. Johnson, MRS Bull. 24 1999 42-56.

DOI      URL     [Cited within: 2]

W.H. Wang, C. Dong, C.H. Shek, Mater. Sci. Eng. R 44 (2004) 45-89.

DOI      URL     [Cited within: 1]

C. Schuh, T. Hufnagel, U. Ramamurty, Acta Mater. 55 2007 4067-4109.

DOI      URL     [Cited within: 1]

AbstractThe mechanical properties of amorphous alloys have proven both scientifically unique and of potential practical interest, although the underlying deformation physics of these materials remain less firmly established as compared with crystalline alloys. In this article, we review recent advances in understanding the mechanical behavior of metallic glasses, with particular emphasis on the deformation and fracture mechanisms. Atomistic as well as continuum modeling and experimental work on elasticity, plastic flow and localization, fracture and fatigue are all discussed, and theoretical developments are connected, where possible, with macroscopic experimental responses. The role of glass structure on mechanical properties, and conversely, the effect of deformation upon glass structure, are also described. The mechanical properties of metallic glass-derivative materials – including in situ and ex situ composites, foams and nanocrystal-reinforced glasses – are reviewed as well. Finally, we identify a number of important unresolved issues for the field.]]>

J.W. Qiao, H. Jia, P.K. Liaw, Mater. Sci. Eng. R 100 (2016) 1-69.

DOI      URL     [Cited within: 6]

Y.Y. Liu, P.Z. Liu, J.J. Li, P.K. Liaw, F. Spieckermann, D. Kiener, J.W. Qiao, Int. J. Plast. 105 2018 225-238.

DOI      URL     [Cited within: 3]

Z. Liu, R. Li, G. Liu, W. Su, H. Wang, Y. Li, M. Shi, X. Luo, G. Wu, T. Zhang, Acta Mater. 60 2012 3128-3139.

DOI      URL     [Cited within: 1]

Y. Wu, Y. Xiao, G. Chen, C.T. Liu, Z. Lu, Adv. Mater. 22 2010 2770-2773.

DOI      URL     PMID      [Cited within: 4]

Y.Y. Liu, J.J. Li, Z. Wang, X.H. Shi, J.W. Qiao, Y.C. Wu, Intermetallics 108 (2019) 72-76.

DOI      URL     [Cited within: 2]

K.K. Song, S. Pauly, Y. Zhang, R. Li, S. Gorantla, N. Narayanan, U. Kühn, T. Gemming, J. Eckert, Acta Mater. 60 2012 6000-6012.

DOI      URL     [Cited within: 2]

A triple yielding phenomenon resulting in high fracture strength (up to 2360 +/- 10 MPa) and large ductility (up to 13.5 +/- 0.5%) during compressive deformation is first discovered and systematically investigated in metastable Cu47.5Zr47.5Al5 composites. Based on electron microscopy and X-ray diffraction studies, the major deformation mechanisms at the different stages of compression are revealed to be: (1) the constraint effect of the amorphous phase; (2) the pre-existence of a small number of microtwins and a small amount of martensitic phases; (3) the initiation and development of martensitic transformation; (4) the interplay between CuZr crystals and shear bands; and (5) the activation of partial detwinning, together with a high density of dislocations. The yield strengths of the composites are modeled as a function of the volume fraction of the constituent phases and a transition from the rule-of-mixtures to a load-bearing model is observed. The underlying ductile nature of the composites is further correlated with the unique electronic structure of the B2 CuZr intermetallics. The present results provide an understanding of the deformation mechanism of metastable CuZr-based composites and give guidance on how to improve the ductility/toughness of bulk metallic glasses. (C) 2012 Acta Materialia Inc. Published by Elsevier Ltd.

Y.S. Oh, C.P. Kim, S. Lee, N.J. Kim, Acta Mater. 59 2011 7277-7286.

DOI      URL     [Cited within: 3]

In the present study, two Ti-based amorphous matrix composites containing ductile dendrites dispersed in an amorphous matrix were fabricated by a vacuum arc melting method, and deformation mechanisms related to the improvement of strength and ductility were investigated by focusing on how ductile dendrites affected the initiation and propagation of deformation bands, shear bands or twins. Ti-based amorphous matrix composites contained 70-73 vol.% coarse dendrites of size 90-180 mu m, and had excellent tensile properties of the yield strength (1.2-1.3 GPa) and elongation (8-9%). The Ta-containing composite showed strain hardening after yielding, and reached fracture without showing necking, whereas necking occurred straight after yielding without strain hardening in the Nb-containing composite. The improved tensile elongation and strain hardening behavior was explained by the homogeneous distribution of dendrites large enough to form deformation bands or twins, the role of beta phases surrounding alpha phases to prevent the formation of twins, and deformation mechanisms such as strain-induced beta to alpha transformation. (C) 2011 Acta Materialia Inc. Published by Elsevier Ltd.

L. Zhang, W.Q. Li, Z.W. Zhu, H.M. Fu, H. Li, Z.K. Li, H.W. Zhang, A.M. Wang, H.F. Zhang, J. Mater. Sci. Technol. 33 2017 708-711.

DOI      URL     [Cited within: 1]

L. Zhang, R.L. Narayan, H.M. Fu, U. Ramamurty, W.R. Li, Y.D. Li, H.F. Zhang, Acta Mater. 168 2019 24-36.

DOI      URL     [Cited within: 2]

F. Szuess, C.P. Kim, W.L. Johnson, Acta Mater. 49 2001 1507-1513.

DOI      URL     [Cited within: 1]

AbstractDuctile phase containing bulk metallic glass composites are prepared via an in situ method by rapid quenching of a homogenous Zr56.2Ti13.8Nb5.0Cu6.9Ni5.6Be12.5 melt. The microstructure of the resulting two phase material is investigated by SEM, X-ray diffraction, and microprobe analysis. The composite material, as well as single phase materials with the corresponding matrix and second phase compositions, are tested in uniaxial tension and compression. Young's Modulus, shear modulus and Poisson ratios are analyzed by ultrasonic sound velocity measurements. The composite material demonstrates strongly improved Charpy impact toughness (by a factor of 2.5 compared to Vitreloy 1) and ductility (average fracture strain up to 8.3% in compression and 5.5% in tension). These remarkable improvements are explained by the effect of the mechanically soft and ductile second phase, which acts stabilizing against shear localization and critical crack propagation.]]>

J. Fan, J.W. Qiao, Z.H. Wang, W. Rao, G.Z. Kang, Sci. Rep. 7 2017 1877.

DOI      URL     PMID      [Cited within: 5]

The present study demonstrates that Ti-based metallic glass matrix composites (MGMCs) with a normal composition of Ti43Zr32Ni6Ta5Be14 containing ductile dendrites dispersed in the glass matrix has been developed, and deformation mechanisms about the tensile property have been investigated by focusing on twinning-induced plasticity (TWIP) effect. The Ti-based MGMC has excellent tensile properties and pronounced tensile work-hardening capacity, with a yield strength of 1100 MPa and homogeneous elongation of 4%. The distinguished strain hardening is ascribed to the formation of deformation twinning within the dendrites. Twinning generated in the dendrites works as an obstacle for the rapid propagation of shear bands, and then, the localized necking is avoided, which ensures the ductility of such kinds of composites. Besides, a finite-element model (FEM) has been established to explain the TWIP effect which brings out a work-hardening behavior in the present MGMC instead of a localized strain concentration. According to the plasticity theory of traditional crystal materials and some new alloys, TWIP effect is mainly controlled by stacking fault energy (SFE), which has been analyzed intensively in the present MGMC.

W. Rao, J. Zhang, G.Z. Kang, C. Yu, H. Jiang, Int. J. Plast. 115 2019 238-267.

DOI      URL     [Cited within: 2]

Y. Jiang, X. Shi, K. Qiu, Mater. Des. 77 2015 32-40.

DOI      URL     [Cited within: 2]

W. Rao, J. Zhang, H. Jiang, G.Z. Kang, Mech. Mater. 103 2016 68-77.

DOI      URL     [Cited within: 2]

C.C. Hays, C.P. Kim, W.L. Johnson, Phys. Rev. Lett. 84 1999, 2901-1904.

DOI      URL     PMID      [Cited within: 2]

Results are presented for a ductile metal reinforced bulk metallic glass matrix composite based on glass forming compositions in the Zr-Ti-Cu-Ni-Be system. Primary dendrite growth and solute partitioning in the molten state yields a microstructure consisting of a ductile crystalline Ti-Zr-Nb beta phase, with bcc structure, in a Zr-Ti-Nb-Cu-Ni-Be bulk metallic glass matrix. Under unconstrained mechanical loading organized shear band patterns develop throughout the sample. This results in a dramatic increase in the plastic strain to failure, impact resistance, and toughness of the metallic glass.

L. Zhang, Z. Zhu, H. Fu, H. Li, H. Zhang, Mater. Sci. Eng. A 689 (2017) 404-410.

DOI      URL     [Cited within: 1]

D. Banerjee, J.C. Williams, Acta Mater. 61 2013 844-879.

DOI      URL     [Cited within: 4]

The basic framework and - conceptual understanding of the metallurgy of Ti alloys is strong and this has enabled the use of titanium and its alloys in safety-critical structures such as those in aircraft and aircraft engines. Nevertheless, a focus on cost-effectiveness and the compression of product development time by effectively integrating design with manufacturing in these applications, as well as those emerging in bioengineering, has driven research in recent decades towards a greater predictive capability through the use of computational materials engineering tools. Therefore this paper focuses on the complexity and variety of fundamental phenomena in this material system with a focus on phase transformations and mechanical behaviour in order to delineate the challenges that lie ahead in achieving these goals.

M. Marteleur, F. Sun, T. Gloriant, P. Vermaut, P.J. Jacques, F. Prima, Scripta Mater. 66 2012 749-752.

DOI      URL     [Cited within: 1]

G.E. Dieter, in: F.M. Robert (Ed.), Mechanical Metallurgy, McGraw-Hill, London, 1961, p. 290.

[Cited within: 3]

X. Li, R. Song, N. Zhou, J. Li, Scripta Mater. 154 2018 30-33.

DOI      URL     [Cited within: 1]

C.L. Yang, Z.J. Zhang, P. Zhang, Z.F. Zhang, Acta Mater. 136 2017 1-10.

DOI      URL     [Cited within: 1]

Z.F. Zhang, J. Eckert, L. Schultz, Acta Mater. 51 2003 1167-1179.

DOI      URL     [Cited within: 1]

AbstractThe compressive and tensile deformation, as well as the fracture behavior of a Zr59Cu20Al10Ni8Ti3 bulk metallic glass were investigated. It was found that under compressive loading, the metallic glass displays some plasticity before fracture. The fracture is mainly localized on one major shear band and the compressive fracture angle, θC, between the stress axis and the fracture plane is 43°. Under tensile loading, the material always displays brittle fracture without yielding. The tensile fracture stress, σFT, is about 1.58 GPa, which is lower than the compressive fracture stress, σFC(=1.69 GPa). The tensile fracture angle, θT, between the stress axis and the fracture plane is equal to 54°. Therefore, both θC and θT deviate from the maximum shear stress plane (45°), indicating that the fracture behavior of the metallic glass under compressive and tensile load does not follow the von Mises criterion. Scanning electron microscope observations reveal that the compressive fracture surfaces of the metallic glass mainly consist of a vein-like structure. A combined feature of veins and some radiate cores was observed on the tensile fracture surfaces. Based on these results, the fracture mechanisms of metallic glass are discussed by taking the effect of normal stress on the fracture process into account. It is proposed that tensile fracture first originates from the radiate cores induced by the normal stress, then propagates mainly driven by shear stress, leading to the formation of the combined fracture feature. In contrast, the compressive fracture of metallic glass is mainly controlled by the shear stress. It is suggested that the deviation of θC and θT from 45° can be attributed to a combined effect of the normal and shear stresses on the fracture plane.]]>

Y. Shao, K. Yao, M. Li, X. Liu, Appl. Phys. Lett. 103 2013, 171901.

DOI      URL     [Cited within: 1]

H.M. Zhai, H.F. Wang, F. Liu, J. Alloys. Compd. 685 2016 322-330.

DOI      URL     [Cited within: 3]

H. Zhai, Y. Xu, F. Zhang, Y. Ren, H. Wang, F. Liu, J. Alloys. Compd. 694 2017 1-9.

DOI      URL     [Cited within: 1]

L.S. Tóth, Y. Estrin, R. Lapovok, C. Gu, Acta Mater. 58 2010 1782-1794.

DOI      URL     [Cited within: 2]

AbstractA new model is proposed that aims to capture within a single modelling frame all the main microstructural features of a severe plastic deformation process. These are: evolution of the grain size distribution, misorientation distribution, crystallographic texture and the strain-hardening of the material. The model is based on the lattice curvature that develops in all deformed grains. The basic assumption is that lattice rotation within an individual grain is impeded near the grain boundaries by the constraining effects of the neighbouring grains, which gives rise to lattice curvature. On that basis, a fragmentation scheme is developed which is integrated in the Taylor viscoplastic polycrystal model. Dislocation density evolution is traced for each grain, which includes the contribution of geometrically necessary dislocations associated with lattice curvature. The model is applied to equal-channel angular pressing. The role of texture development is shown to be an important element in the grain fragmentation process. Results of this modelling give fairly precise predictions of grain size and grain misorientation distribution. The crystallographic textures are well reproduced and the strength of the material is also reliably predicted based on the modelling of dislocation density evolution coupled with texture development.]]>

W. Guo, E. Jägle, J. Yao, V. Maier, S. Korte-Kerzel, J.M. Schneider, D. Raabe, Acta Mater. 80 2014 94-106.

DOI      URL     [Cited within: 1]

Introducing a soft crystalline phase into an amorphous alloy can promote the compound's ductility. Here we synthesized multilayered nanolaminates consisting of alternating amorphous Cu54Zr46 and nanocrystalline Cu layers. The Cu layer thickness was systematically varied in different samples. Mechanical loading was imposed by nanoindentation and micropillar compression. Increasing the Cu layer thickness from 10 to 100 am led to a transition from sharp, cross-phase shear banding to gradual bending and co-deformation of the two layer types (amorphous/nanocrystalline). Specimens with a sequence of 100 nm amorphous Cu54Zr46 and 50 nm Cu layers show a compressive flow stress of 2.57 +/- 0.21 GPa, matching the strength of pure CuZr metallic glass, hence exceeding the linear rule of mixtures. In pillar compression, 40% strain without fracture was achieved by the suppression of percolative shear band propagation. The results show that inserting a ductile nanocrystalline phase into a metallic glass prevents catastrophic shear banding. The mechanical response of such nanolaminates can be tuned by adjusting the layer thickness. (C) 2014 Acta Materialia Inc. Published by Elsevier Ltd.

J.W. Qiao, J. Mater. Sci. Technol. 29 2013 685-701.

DOI      URL     [Cited within: 1]

S.B. Yi, C.H.J. Davies, H.G. Brokmeier, R.E. Bolmaro, K.U. Kainer, J. Homeyer, Acta Mater. 54 2006 549-562.

DOI      URL     [Cited within: 1]

AbstractThe anisotropic mechanical behavior of extruded AZ31 magnesium alloy is described in relation to the crystallographic texture. Specimens taken at 0°, 45° and 90° to the extrusion direction were uniaxially loaded in tension and compression, and the texture evolution was measured under load, using synchrotron radiation. The correlation between the initial texture, the mechanical anisotropy and the activation of different deformation modes was interpreted using the in situ texture measurements and viscoplastic self-consistent simulation results. The activity of the basal 〈a〉 slip and the tensile twinning exert a significant effect on the mechanical anisotropy during tension, while the importance of the 〈c + a〉 slip increases during compression.]]>

O. Grässel, L. Krüger, G. Frommeyer, L.W. Meyer, Int. J. Plast. 16 2000 1391-1409.

DOI      URL     [Cited within: 1]

B.C. De Cooman, Y. Estrin, S.K. Kim, Acta Mater. 142 2018 283-362.

DOI      URL     [Cited within: 1]

L. Ma, L. Wang, Z. Nie, F. Wang, Y. Xue, J. Zhou, T. Cao, Y. Wang, Y. Ren, Acta Mater. 128 2017 12-21.

DOI      URL     [Cited within: 1]

V.S.A. Challa, X.L. Wan, M.C. Somani, L.P. Karjalainen, R.D.K. Misra, Scripta Mater. 86 2014 60-63.

DOI      URL     [Cited within: 1]

In the context of obtaining high strength-high ductility combination in nanograined/ultrafine-grained austenitic stainless steels, we underscore the dependence of grain structure and deformation mechanism. In high strength nanograined/ultrafine-grained steel, deformation twinning contributed to the excellent ductility and high strain hardening rate, while in the low strength coarse-grained steel, ductility and strain hardening ability was also good, but due to strain-induced martensite. The change in deformation mechanism with grain size is attributed to austenite stability-strain energy relationship. (C) 2014 Acta Materialia Inc. Published by Elsevier Ltd.

K.Y. Zhu, A. Vassel, F. Brisset, K. Lu, J. Lu, Acta Mater. 52 2004 4101-4110.

DOI      URL     [Cited within: 1]

AbstractA nanostructured surface layer up to 50 μm thick was produced on commercially pure titanium using surface mechanical attrition treatment (SMAT). The microstructural features of the surface layer produced by SMAT were systematically characterized by cross-sectional optical microscopy observations, transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) investigations. The grain refinement process, accompanied by an increase in strain in the surface layer, involves: (1) the onset of twins and the intersection of twin systems, (2) the formation of dislocation walls, (3) the nucleation of microbands associated with the splitting of dislocation walls, (4) the subdivision of microbands into low angle disoriented blocks and then highly disoriented polygonal submicronic grains, and (5) further breakdown of submicronic polygonal grains into randomly oriented nanograins. The final grain refinement step to form nanograins has been discussed on the basis of a recrystallization process.]]>

E.W. Huang, J. Qiao, B. Winiarski, W.J. Lee, M. Scheel, C.P. Chuang, P.K. Liaw, Y.C. Lo, Y. Zhang, M.D. Michiel, Sci. Rep. 4 2014 4394.

DOI      URL     PMID      [Cited within: 1]

In-situ synchrotron x-ray experiments have been used to follow the evolution of the diffraction peaks for crystalline dendrites embedded in a bulk metallic glass matrix subjected to a compressive loading-unloading cycle. We observe irreversible diffraction-peak splitting even though the load does not go beyond half of the bulk yield strength. The chemical analysis coupled with the transmission electron microscopy mapping suggests that the observed peak splitting originates from the chemical heterogeneity between the core (major peak) and the stiffer shell (minor peak) of the dendrites. A molecular dynamics model has been developed to compare the hkl-dependent microyielding of the bulk metallic-glass matrix composite. The complementary diffraction measurements and the simulation results suggest that the interface, as Maxwell damper, between the amorphous matrix and the (211) crystalline planes relax under prolonged load that causes a delay in the reload curve which ultimately catches up with the original path.

F. Roters, D. Raabe, G. Gottstein, Acta Mater. 48 2000 4181-4189.

DOI      URL     [Cited within: 1]

Y. Wang, J. Li, A.V. Hamza, T.W. Barbee, Proc. Natl. Acad. Sci. U. S. A. 104 2007 11155-11160.

DOI      URL     PMID      [Cited within: 1]

H. Bei, S. Xie, E.P. George, Phys. Rev. Lett. 96 2006, 105503.

DOI      URL     PMID      [Cited within: 1]

By controlling the specimen aspect ratio and strain rate, compressive strains as high as 80% were obtained in an otherwise brittle metallic glass. Physical and mechanical properties were measured after deformation, and a systematic strain-induced softening was observed which contrasts sharply with the hardening typically observed in crystalline metals. If the deformed glass is treated as a composite of hard amorphous grains surrounded by soft shear-band boundaries, analogous to nanocrystalline materials that exhibit inverse Hall-Petch behavior, the correct functional form for the dependence of hardness on shear-band spacing is obtained. Deformation-induced softening leads naturally to shear localization and brittle fracture.

X.H. Sun, J.W. Qiao, Z.M. Jiao, Z.H. Wang, H.J. Yang, B.S. Xu, Sci. Rep. 5 2015 13964.

DOI      URL     PMID      [Cited within: 2]

With regard to previous tensile deformation models simulating the tensile behavior of in-situ dendrite-reinforced metallic glass matrix composites (MGMCs) [Qiao et al., Acta Mater. 59 (2011) 4126; Sci. Rep. 3 (2013) 2816], some parameters, such as yielding strength of the dendrites and glass matrix, and the strain-hardening exponent of the dendrites, are estimated based on literatures. Here, Ti48Zr18V12Cu5Be17 MGMCs are investigated in order to improve the tensile deformation model and reveal the tensile deformation mechanisms. The tensile behavior of dendrites is obtained experimentally combining nano-indentation measurements and finite-element-method analysis for the first time, and those of the glass matrix and composites are obtained by tension. Besides, the tensile behavior of the MGMCs is divided into four stages: (1) elastic-elastic, (2) elastic-plastic, (3) plastic-plastic (work-hardening), and (4) plastic-plastic (softening). The respective constitutive relationships at different deformation stages are quantified. The calculated results coincide well with the experimental results. Thus, the improved model can be applied to clarify and predict the tensile behavior of the MGMCs.

D.P. Wang, S.L. Wang, J.Q. Wang, Corros. Sci. 59 2012 88-95.

DOI      URL     [Cited within: 1]

W.H. Wang, Prog. Mater. Sci. 57 2012 487-656.

DOI      URL     [Cited within: 1]

Z. Hashin, J. Mech. Phys. Solids. 11 1963 127-140.

DOI      URL     [Cited within: 1]

S.H. Xia, J.T. Wang, Int. J. Plast. 26 2010 1442-1460.

DOI      URL     [Cited within: 3]

W.L. Johnson, K. Samwer, Phys. Rev. Lett. 95 2005, 195501.

DOI      URL     PMID      [Cited within: 1]

Room temperature (TR) elastic constants and compressive yield strengths of approximately 30 metallic glasses reveal an average shear limit gammaC=0.0267+/-0.0020, where tauY=gamma CG is the maximum resolved shear stress at yielding, and G the shear modulus. The gammaC values for individual glasses are correlated with t=TR/Tg , and gamma C for a single glass follows the same correlation (vs t=T/Tg). A cooperative shear model, inspired by Frenkel's analysis of the shear strength of solids, is proposed. Using a scaling analysis leads to a universal law tauCT/G=gammaC0-gammaC1(t)2/3 for the flow stress at finite T where gammaC0=(0.036+/-0.002) and gammaC1=(0.016+/-0.002).

J. Joseph, N. Stanford, P. Hodgson, D.M. Fabijanic, J. Alloys. Compd. 726 2017 885-895.

DOI      URL     [Cited within: 1]

J.A. Kolodziejska, H. Kozachkov, K. Kranjc, A. Hunter, E. Marquis, W.L. Johnson, K.M. Flores, D.C. Hofmann, Sci. Rep. 6 2016 22563.

DOI      URL     PMID      [Cited within: 1]

The microstructure and tension ductility of a series of Ti-based bulk metallic glass matrix composite (BMGMC) is investigated by changing content of the beta stabilizing element vanadium while holding the volume fraction of dendritic phase constant. The ability to change only one variable in these novel composites has previously been difficult, leading to uninvestigated areas regarding how composition affects properties. It is shown that the tension ductility can range from near zero percent to over ten percent simply by changing the amount of vanadium in the dendritic phase. This approach may prove useful for the future development of these alloys, which have largely been developed experimentally using trial and error.

J.C. Lee, Y.C. Kim, J.P. Ahn, H.S. Kim, S.H. Lee, B.J. Lee, Acta Mater. 52 2004 1525-1533.

DOI      URL     [Cited within: 1]

AbstractWith the development of various processes to produce bulk amorphous composites with enhanced plasticity, investigations of the mechanical behaviors of the amorphous alloys in the plastic regime have now become feasible. In addition to dramatically enhanced plasticity, some bulk amorphous composites have exhibited a work hardening behavior during plastic deformation. Considering that most strengthening mechanisms, such as solid solution hardening, martensitic hardening, etc., operative in crystalline metals are associated with dislocations, the work hardening behavior observed from amorphous composites, where dislocations do not exist, is of a special scientific interest. We have observed that quasistatic compression imposed to the amorphous composite induces the homogeneous precipitation of nanocrystallites from the amorphous matrix of the composite, which, in turn, leads to strengthening the amorphous composite. The strengthening mechanism of the amorphous matrix composite is investigated as well.]]>

L. Zhang, H.F. Zhang, W.Q. Li, T. Gemming, Z.W. Zhu, H.M. Fu, J. Eckert, S. Pauly, Scripta Mater. 125 2016 19-23.

DOI      URL     [Cited within: 1]

/