Sie sind auf Seite 1von 12

Additive Manufacturing 25 (2019) 499–510

Contents lists available at ScienceDirect

Additive Manufacturing
journal homepage: www.elsevier.com/locate/addma

Full Length Article

On the microstructure, mechanical properties and wear resistance of an T


additively manufactured Ti64/metallic glass composite
⁎ ⁎
Xiao-Jun Shena, Cheng Zhanga, , Yan-Ge Yangb, Lin Liua,
a
State Key Laboratory of Material Processing and Die & Mould Technology and School of Materials Science and Engineering, Huazhong University of Science and
Technology, 430074, Wuhan, China
b
Laboratory for Corrosion and Protection, Institute of Metal Research, Chinese Academy of Sciences, Shenyang, 110016, China

A R T I C LE I N FO A B S T R A C T

Keywords: Selective laser melting (SLM) provides flexibility in creating novel metal-matrix composites (MMCs) with unique
Selective laser melting microstructures and enhanced mechanical properties over conventionally manufactured MMGs. In this study, a
Titanium alloys Zr-based metallic glass (MG) decorated Ti6Al4V (Ti64) composite with a unique hybrid nanostructure and en-
Metallic glass hanced mechanical properties and wear resistance was fabricated using SLM. The results revealed that a near-full
Mechanical properties
dense and crack-free Ti-based composite was produced, with its reinforcements consisting of ultrafine β den-
Wear resistance
drites set with partially crystallized MG nanobands uniformly distributed along the boundaries of the melt pool.
The addition of MG significantly affected the solidification behavior of the Ti-liquid because of its higher dy-
namic viscosity and density as well as compositional effect on the phase stability. With such a unique nanos-
tructured reinforcement, the Ti64/MG composite exhibited an enhanced yield strength (> 1 GPa) with rea-
sonable ductility and fracture toughness. On the basis of the result of a theoretical analysis, we attributed the
main strengthening mechanism to Orowan strengthening. The wear resistance was also much improved in the
Ti64/MG composite, arising from the higher hardness of the nanostructured reinforcement and the formation of
a more protective tribo-oxide layer during sliding. The confinement of the 3D distributed reinforcement phase
played a crucial role in preventing the delamination of the tribo-layer on the matrix. This work opens a pathway
to the design of novel additively manufactured MMCs with outstanding mechanical properties.

1. Introduction Selective laser melting (SLM), one of the emerging additive manu-
facturing (AM) technologies, has become a promising route for manu-
Ti-alloys are a workhorse in the modern aerospace and biomedical facturing Ti-MMC parts with complex or customized geometries. During
industries due to their low density, excellent corrosion resistance, high SLM, near-net shape components are built by selective laser melting
fracture toughness and good biocompatibility [1–3]. However, their directly from digital files [9,10]. It was reported that the advantages
low hardness and poor wear resistance usually result in galling failure stemming from AM (especially SLM) can decrease the production costs
and fretting fatigue, which are challenges for the application of Ti-al- for Ti-based components (e.g., engine brackets) by 50% by reducing
loys [4]. To improve the strength and wear resistance, ceramic particles exorbitant machining costs and material loss [11]. Extensive works
have been commonly introduced into Ti and Ti-alloys, resulting in Ti- have been carried out to understand the relationship between the mi-
based metal-matrix composites (MMC). Low-density and hard ceramic crostructure and mechanical properties of Ti-alloys and Ti-MMCs pro-
particle reinforcements [5], such as TiB, TiB2, TiC, TiN, B4C and Ti5Si3, cessed by SLM [8–18]. Typically, the high temperature gradient and
demonstrate the potential to improve the strength and wear perfor- fast cooling rate inherent to the SLM process produce fine and me-
mance, but severely degrade the plasticity and toughness of Ti and Ti- tastable microstructures (e.g., the martensitic α’ phase [18]) in the
alloys because of their poor wettability with the Ti matrix [5–7] Fur- resultant Ti-alloys. Attar et al. [19] compared the wear properties of
thermore, the electrical insulativity and high melting points of the commercially pure Ti processed by SLM and casting, and concluded
ceramic reinforcements give rise to challenges in the machining and that the SLMed Ti exhibited better wear resistance than cast CP-Ti due
processing of Ti-MMCs by using conventional manufacturing techni- to its martensitic microstructure and finer grain size. Zhu et al. [20]
ques, such as casting and powder metallurgy [8]. reported that the wear of SLMed Ti6Al4V (Ti64 hereafter) alloy in


Corresponding authors.
E-mail addresses: czhang@hust.edu.cn (C. Zhang), lliu2000@mail.hust.edu.cn (L. Liu).

https://doi.org/10.1016/j.addma.2018.12.006
Received 10 September 2018; Received in revised form 9 November 2018; Accepted 8 December 2018
Available online 10 December 2018
2214-8604/ © 2018 Elsevier B.V. All rights reserved.
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

contact with hardened 38CrMoAl alloy is severe, dominated by oxida- wavelength of 1.06 μm, a maximum output power of 500 W and a laser
tive, abrasion, and delamination wear mechanisms. Recently, there spot diameter of 80 μm. The chamber was first purged to vacuum and
have been many attempts to enhance the mechanical and wear prop- then filled with argon gas, maintaining the oxygen content below
erties of SLMed Ti-alloys by making different Ti-MMC structures 100 ppm during the entire process. All samples were produced using
[8,21–25]. Yet, the common problem persisting through these attempts pre-optimized processing parameters: a laser power P of 175 W, a
is the high residual porosity (1.7–15%) of the obtained MMC compo- scanning speed v of 1000 mm s−1, a 100 μm hatch spacing h and a
nents, as a result of the poor wettability between the metal and ceramic 40 μm layer thickness t. The layers were scanned in a continuous laser
particles, as well as insufficient time for pore filling between large mode according to a zigzag pattern, rotating 90° after each layer, as
ceramic particles due to the extremely short interval (< 1 ms) of the illustrated in Fig. 1h. The appearance of a representative composite
liquid phase during SLM [21,26]. As a consequence of the very large component prepared by SLM is presented in Fig. 1i.
temperature gradient, thermal stress concentrations always emerge at
the ceramic/metal interface, which initiate cracks and debonding upon 2.2. Microstructural characterization
loading [23,27].
Metallic glasses (MGs) have attracted increasing interest as a new The phase identification of the as-printed samples was performed by
type of advanced reinforcement because of their superior mechanical X-ray diffraction (XRD, 7000SX, Shimadzu) with Cu Kα radiation. For
properties compared to their crystalline counterparts [28–32], as well microscopic studies, longitudinal and transverse section of the as-
as better ductility and wettability with the metal matrix compared to printed parts were prepared by conventional grinding, polishing and
ceramic reinforcements [33]. In particular, Zhang et al. [32] found that chemical etching with Kroll’s reagent (1 ml HF, 2 ml HNO3 and 50 ml
the refinement of the MG micro-particles to nano-particles results in deionized H2O). The metallurgical defects and microstructures were
MMCs with high strength and plasticity. To date, the majority of MG- investigated by optical microscopy (OM, Leica DFC 450), three-di-
reinforced MMCs have been fabricated based on powder metallurgy, mensional X-ray tomography (XRT, Xradia versa XRM 500) and scan-
and few studies have attempted to manufacture MG-reinforced MMCs ning electron microscopy (SEM, Zeiss Gemini 300) coupled with en-
by AM technologies. Very recently, Zhang et al. [34] carried out pre- ergy-dispersive X-ray spectroscopy (EDX). The interfacial structures
liminary work on the fabrication of a Fe-based MG-strengthened between the matrix and reinforcement phase were characterized
stainless steel composite by SLM, and found that the tensile strength of through transmission electron microscopy (TEM, FEI Titan Themis
the as-printed composite increased from 819 MPa to 1090 MPa while its 200). The TEM thin foil was prepared by a focused ion beam (FIB, FEI
elongation decreased from 24.7% to 9.8%. However, the underlying Helios 450S).
strengthening mechanism associated with the addition of MG particles
remained elusive. Our recent works have verified that fully densified 2.3. Mechanical property measurements
Zr-based MG components with a high amorphous phase content
(> 90%) and complex geometries can be readily 3D-printed by SLM Compressive mechanical testing was carried out on cylindrical parts
[35], providing us the possibility to fabricate Zr-based MG-reinforced (Φ 3 mm × 6 mm) along the SLM building direction (Z) using a uni-
Ti-alloy composites by SLM and investigate the relation between the versal testing machine (Zwick/Roell 020) at a constant strain rate of
microstructure and mechanical properties. 1.0 × 10−4 s-1. The two ends of the specimens were polished carefully
With this aim, this study investigates the microstructural evolution, to ensure their parallelism. Three-point bending tests were performed at
mechanical properties and wear performance of an MG reinforced Ti64 a displacement rate of 0.5 mm s-1, and the tested samples had dimen-
composite prepared by SLM. Here, a Zr-based MG system with a nom- sions of 37 × 4.8 × 2 mm3. The span was set to be 30 mm. Afterwards,
inal composition of Zr55Cu30Ni5Al10 is selected as the reinforcement the surfaces of the fractured samples were observed by SEM. At least 3
due to its high glass-forming ability, good printability, and comparable tests were repeated for each sample to ensure the repeatability of the
Young’s modulus and coefficient of thermal expansion (CTE) with the data. The notch fracture toughness (KQ) of the samples was measured
Ti64 alloy [35,36]. It is shown that the addition of MG particles led to via a three-point bending test at a displacement rate of 0.1 mm min-1, in
the formation of a unique nanostructured reinforcement, which re- which specimens with dimensions of 4.5 mm × 9 mm × 36 mm and a
sulted in the enhancement of the strength and wear resistance of the notch with a depth of 4.5 mm and a root radius of approximately
Ti64 alloy while maintaining good plasticity and fracture toughness. 144 μm (cut by a low-speed diamond saw) were used. The fracture
The strengthening mechanisms and origins of the enhancement of the toughness (KQ) values were calculated by using standard references for
wear resistance of the as-printed composite are discussed in detail. this geometry [37].

2. Experimental procedures 2.4. Hardness and wear tests

2.1. Powder materials and SLM processing Vickers hardness testing was performed on polished samples using a
microhardness tester with a 300 g load and 10 s dwelling time. To probe
Commercial Ti6Al4V (Ti64) alloy powder (< 53 μm, provided by the possible difference in the mechanical properties of the matrix and
Falcon Tech Co, Ltd.) and a gas-atomized Zr55Cu30Ni5Al10 (at%) MG reinforcement phase, a nanoindentation test was conducted using a
powder (< 53 μm, provided by Nanjing MeiRui Wear-resistant Triboindenter (Hysitron TI750, USA) equipped with a Berkovich geo-
Materials Tech Co. Ltd.) were used in this study. The SEM morphologies metry diamond tip. The indentations were performed to a maximum
of the two types of powder are shown in Fig. 1a and b, revealing good load of 5 mN with a loading rate of 1 mN s−1, and the measurements
spherical shapes. The glassy nature of the MG powders prepared was were performed at least 25 times to ensure data reliability. The wear
confirmed by X-ray diffraction measurement (see inset in Fig. 1b). The and friction performance of the Ti64 and Ti64/MG composite was
two powders were mixed together through low-energy ball-milling evaluated at room temperature using a UMT-3 Instruments ball-on-
using a planetary mill to obtain a Ti64/MG mixture powder. SEM and plate wear tester in air and 0.1 M NaCl solution. Tribo tests were con-
EDX mapping (Fig. 1c–f) verified the homogeneous distribution of MG ducted against a Si3N4 ceramic ball with a diameter of 6.35 mm using
powder in the mixture. Here, the optimal addition of MG powder was the following parameters: a normal load (N) of 10 N, a sliding velocity
determined to be 10 wt% since any further increase resulted in a sig- of 10 mm s−1, a 5 mm oscillating stroke and a 30 min total sliding
nificant decrease in the ductility of the resultant composites. The SLM duration (S). New balls were used for each test. The wear tracks were
experiments were performed with a commercial SLM machine (For- analyzed using an optical interferometer (VK-X200) to determine the
wedo LM-120, China, Fig. 1g) containing a fiber laser with a wear volume loss (V). The wear rate, Q, was then calculated by the

500
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 1. (a, b) The SEM morphologies of the Ti64 and metallic glass (MG) powder. Inset in (b) is the XRD pattern of the MG powder. (c–f) The SEM and EDX mapping
of Ti64/MG powder mixture. (g, h) The SLM processing and scanning strategy used for the SLM experiments. (i) The appearance of an as-polished Ti64/MG
composite component.

following equation: Q = V/NS. The surfaces of the worn tracks were Ni, and Al as the solutes) into the β phase and α′ martensite. In addi-
observed by SEM. Microscratching was performed with a 200 mN load tion, no diffraction halo for the amorphous phase was observed in the
with a sliding speed of 20 μm s−1 by using a diamond probe with a pattern of the Ti64/MG composite, probably due to the small volume
radius of 100 μm to evaluate the phase-specific coefficient of friction fraction of the amorphous phase and/or overlap between the amor-
and wear resistance. phous halo and diffraction peaks at 38-39°.
Fig. 3a–d show SEM micrographs of the microstructural features in
3. Results the as-printed Ti64/MG composite and Ti64 alloy. The microstructure
of Ti64 alloy consists of predominantly fine α′ acicular martensite with
3.1. Phase identification and microstructures a width of ∼1 μm and a length of ∼5 μm (Fig. 3c), similar to the
findings in previous work [10,38]. The chemical composition of the
The XRD pattern of the as-printed Ti64/MG composite is depicted in Ti64 alloy determined by EDX analysis is 86.41at% Ti, 9.94 at% Al and
Fig. 2, compared with that of the as-printed Ti64 alloy. The samples 3.65 at% V, which agrees well with its nominal composition. No oxygen
exhibit similar diffraction peaks corresponding to primary α′ martensite content was detected in the as-printed Ti64 part. On the other hand, the
and minor β phases, which are typical phase constituents in SLMed Ti- microstructure of the fabricated Ti64/MG composite has an elliptical
based alloys. The formation of α′ martensite originates from the high melt pool, with white-contrast regions mainly distributed along the
cooling rate associated with the SLM process [18]. Compared to the boundaries of the melt pool (Fig. 3b). A higher-magnification SEM
Ti64 alloy, the peaks of the β and α′ phases in the Ti64/MG composite image (Fig. 3d) further revealed the presence of ultrafine white-contrast
specimen are broadened and have a lower intensity. More importantly, dendrite-like regions with intercellular spacing. The EDX mapping in
the peaks of the β phase and α′ martensite in the composite shifted to a Fig. 3e–g confirmed that those white-contrast regions are enriched in
lower 2θ regime compared to the Ti64 alloy. The insets in Fig. 2 show the MG components (e.g., Zr, Cu, Ni and Al) and Ti. Inside the melt
comparisons of a single peak of the β phase and an α′ phase peak in the pool, some convection swirls are clearly observed, probably resulting
Ti64 and Ti64/MG composite as an example. The results indicate that from the flow of the mixture of MG and the Ti64 melt during SLM. A
the Ti64/MG composite had a finer grain size with an increased lattice similar swirling morphology has been observed during the laser
parameter in both the β and α′ phases. Given that Cu and Ni are β welding of Zr-based MG and Ti64 alloy [39]. No oxygen trace was
stabilizers while Al is α stabilizer, the slightly increased lattice para- presented on the surface of the as-printed Ti64/MG composite, as
meter mostly arises from the solution of MG constituents (i.e., with Cu, verified by EDX analysis.

501
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 2. The XRD patterns of the as-printed Ti64 alloy and Ti64/MG composite. Insets are XRD patterns of the two peaks recorded with a scan rate of 0.4°/min.

To probe the microstructural evolution of the Ti64/MG composite in for the Ti64 alloy, indicative of near-full dense parts being prepared.
more detail, especially in the interfacial regions between the matrix and The density of the present as-printed composite is higher than that of
reinforcement, TEM observation was carried out. Herein, the interfacial the SLMed Ti-MMCs reported previously [21]. According to the statis-
region was carefully extracted by FIB. Fig. 4a shows a representative tical results, the majority of the pores have a size in the range of
high-angle annular dark-field (HAADF) TEM image of the interfacial 5–20 μm, and there are fewer pores in the Ti64/MG composite than in
structure between the matrix and reinforcement. The matrix region the Ti64 alloy.
(right side in Fig. 4a) is characteristic of fine acicular α′ martensite,
while the reinforcement region (left side in Fig. 4a) manifests dendrites
3.2. Mechanical properties
and nanobands. Since the contrast in HAADF mode is due to changes in
the average atomic weight, the nanobands in bright contrast might
Fig. 6a shows representative true stress-strain curves for the as-
thereby correspond to MG regions because of the higher atomic num-
printed Ti64/MG composite and Ti64 alloy under compression. As ex-
bers of Zr, Cu and Ni compared to those of Ti and Al. The square-shaped
pected, the yield strength (σY = 1041 ± 22 MPa) and ultimate strength
dendrites were confirmed to be β phase according to the corresponding
(σU = 1455 ± 28 MPa) of the Ti64/MG composite are significantly
selective area electron diffraction (SAED) pattern (see inset in Fig. 4b).
higher than those of the Ti64 alloy (σY = 903 ± 16 MPa). The strain
The nanobands were confirmed to be the amorphous phase with minor
hardening rates obtained from compressive loading experiments are
nanocrystals, as evidenced by the high-resolution TEM observations
illustrated in Fig. 6b, which clearly demonstrates a higher strain
and SAED patterns shown in Figs. 4c and 5 d. The nanocrystal was
hardenability and strengthening effect in the Ti64/MG composite. The
indexed to the CuZr2 phase (PDF#652647, d112 = 0.2109 nm) ac-
as-printed Ti64/MG composite also remains a pronounced compressive
cording to the IFFT pattern. Fig. 4e schematically shows the distribution
plasticity (20%), indicative of the good ductility of the composite pre-
of the various phases formed in the SLMed Ti64/MG composite. α′
pared.
martensite constitutes the matrix region, while ultrafine β phase den-
In addition to compression tests, three-point bending tests were also
drites and partially crystallized MG nanobands constitute the nanos-
performed on the Ti64/MG composite and Ti64 alloy. Fig. 7 shows the
tructured reinforcement in the resultant composite. Owing to changes
bending strength-displacement curves of the two samples. The flexural
in the thermodynamics and kinetics of the solidification process upon
yield strength of the composite exceeds 1000 MPa, a much higher value
MG addition, the solidification of the mixture melt might follow a
than that of the Ti64 alloy (880 MPa). The Ti64/MG composite also
dendrite-solidification mode, wherein wherein the β phase stabilized by
displays appropriate bendability (see inset in Fig. 7), albeit lower than
Cu and Ni solidifies first due to its higher melting point, and serves as
that of the Ti64 alloy. Fractographs of the bending fractured samples
the nucleators and dendrite centers. The dendritic liquation increases
were observed by SEM, as shown in Fig. 8. The Ti64 alloy exhibits a
the intercellular concentration of MG elements, benefiting the forma-
macroscopically rough fracture surface (Fig. 8a) and microscopic
tion of an amorphous phase as a result of concentration overcooling.
dimple ruptures (Fig. 8b), which are characteristics of ductile fracture.
Partial crystallization occurs in MG nanobands due to a reheating effect
In contrast, the Ti64/MG composite displays a mixture of the features of
upon the layer-by-layer deposition [35,36]. A more detailed discussion
ductile and brittle fracture including dimple zone, vein-like pattern and
on the formation of such unique microstructure will be presented in
decohesive pattern, as depicted in Fig. 8c and d. The inhomogeneous
later section.
fracture surface is attributed to the interaction between the reinforce-
Fig. 5 shows the three-dimensional X-ray tomography images of the
ment and Ti64 matrix. The vein-like pattern, a typical fracture char-
as-printed parts, in which no cracks are observed except for a small
acteristic of ductile MGs, formed due to the instantaneous increase in
number of micropores. Based on the micro-CT data, the relative density
the temperature at the final fracture (adiabatic heating) [40,41]. The
was calculated to be 99.83% for the Ti64/MG composite and 99.79%
decohesive pattern arose from the docohesion fracture of the matrix/

502
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 3. The SEM micrographs showing the typical microstructural features in (a, c) the Ti64 alloy and (b, d) Ti64/MG composite. (e–g) The EDX composition mapping
of the reinforcement near the melt pool boundary.

reinforcement interfaces [42]. Generally, the β-phase is relatively soft composite is high when compared with those of other additively
and ductile when compared to the MG and α′ martensite, whose plastic manufactured alloys and MMCs, e.g., TiAl alloys (18.6-26.7 MPa m1/2)
deformation characteristics involve shear localization [43]. Upon [45], AlSi alloys (19.3-47.0 MPa m1/2) [46], MoSiB alloys (12-18 MPa
plastic straining, the strain incompatibility that develops at the matrix/ m1/2) [47] and Ti/HA composites (1.0-3.5 MPa m1/2) [48].
reinforcement and MG nanobands/β grains could weaken the inter-
faces, which consequently result in the pull-off of the reinforcements 3.3. Wear resistance
from the matrix.
Since the ductility per se is not of major concern in the context of Prior to the wear tests, the hardness of the Ti64/MG composite was
additively manufactured materials, we further measured the notch measured. The mean Vickers microhardness of the Ti64/MG composite
fracture toughness (KQ) of the Ti64 alloy and Ti64/MG composite. The is 382 ± 13 HV0.3, which is 11% larger than that of the Ti64 alloy
Ti64 alloy shows a high KQ value of 70 ± 3 MPa m1/2, consistent with (344 ± 11 HV0.3). In addition, an instrumented nanoindentation test
previously reported values for additively manufactured Ti64 alloy [44]. was carried out to probe the hardness of the reinforcement and matrix
By contrast, the Ti64/MG composite exhibits a slightly reduced fracture regions. The load-depth curves shown in Fig. 9 indicate a much lower
toughness of 42 ± 1 MPa m1/2, which is likely due to the existence of maximum penetration depth of the MG-rich reinforcement, indicating
brittle intermetallic compounds and/or refined α′ martensite. Never- that the reinforcement is harder and stronger than the matrix. The as-
theless, the fracture toughness for the current 3D printed Ti64/MG measured nanohardness is 6.3 ± 0.24 GPa and 5.1 ± 0.24 GPa for the

503
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 4. (a, b) The HAADF-STEM micrograph shows the interface between matrix and MG-rich reinforcement lamellae, the inset is the SAED pattern from as-boxed
region “B”. HRTEM images of region “A” that contains (c) amorphous phase and (d) nanocrystal. Inset is the IFFT pattern of the CuZr2 nanocrystal. (e) Schematic
microstructure of the as-printed Ti64/MG composite.

reinforcement and matrix, respectively. Accordingly, the reinforcement (5.0 ± 0.23 GPa), arising from a grain refinement effect and solid so-
could act as a stress-bearing site to enhance the mechanical perfor- lution strengthening in the composite.
mance. Notably, the nanohardness of the matrix region of the Ti64/MG Fig. 10 compares the wear rates of the Ti64/MG composite and Ti64
composite is slightly higher than that of the Ti64 alloy alloy in different environment. Overall, the Ti64/MG composite

504
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 5. The pore size and distribution in the Ti64 alloy and Ti64/MG composite. Insets are the micro-CT reconstructed images showing the defect characteristics
inside the two samples.

exhibits considerably lower wear rates when compared with the Ti64 sliding interface between the ball and disk. Compared to the wear
alloy, indicative of improved wear resistance. For example, under dry- surface of the Ti64 alloy, less wear debris (caused by delamination) and
wear condition, the wear rate was measured to be 2.33 × 10−4 mm3 N- mixed oxides appeared on the surface of the Ti64/MG composite, cor-
1
m-1, approximately 24% lower than that of the monolithic Ti64 alloy responding to a better wear resistance. In the case of corrosion-wear,
and 69–82% lower than that of SLMed Ti-TiB MMC [24]. In 0.1 M NaCl the wear tracks are manifested by the presence of an adherent tribo-
solution, the Ti64/MG composite also displays a lower wear rate layer enriched with oxygen and slight ploughing grooves, which re-
(1.71 × 10−4 mm3 N-1 m-1) compared to that of the Ti64 alloy present a coupled mechanism of adhesion wear, oxidation wear and
(2.25 × 10−4 mm3 N-1 m-1), indicating that the Ti64/MG composite abrasion wear. The tribolayer formed on the Ti64/MG composite is
performs better even under a corrosion-wear environment. more smooth and has a lower oxygen content compared with that on
The wear tracks of the Ti64/MG composite and Ti64 alloy observed the Ti64 alloy, and is expected to be more protective against wear,
by SEM are shown in Fig. 11. In the case of dry-wear, the wear tracks of which accounts for the lower wear rate in the 0.1 M NaCl solution.
both materials present typical ploughing grooves and flake-shaped wear
debris (see insets in Fig. 11a and b), representing abrasive wear and
delamination wear mechanisms, respectively. In addition, an EDX 4. Discussion
analysis shows oxygen content along the wear surfaces. Since no oxygen
trace is detected on the surface of the as-printed samples, the appear- 4.1. On the microstructure of the as-printed Ti64/MG composite
ance of oxidation is most likely due to the temperature rise at the
In the realm of the SLM of Ti-alloys and Ti-based MMCs, substantial

Fig. 6. (a) True stress-strain curves for the Ti64 alloy and Ti64/MG composite under compression. (b) Comparison of the strain hardening rate between the two
materials.

505
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 7. Bending strength-deformation curves for the Ti64 alloy and Ti64/MG Fig. 9. Load-depth curves of the matrix and reinforcement regions in the Ti64/
composite. The inset shows the picture of as-bended Ti64/MG composite during MG composite and Ti64 alloy.
loading.
nanobands and ultrafine β grains) could play a crucial role in the im-
efforts have been invested in tailoring the types, morphology, grain size proved mechanical properties and wear resistance of the resultant
and characteristic length scales of the constituent phases (α, β, α′) to Ti64/MG composite.
mitigate the strength-ductility trade-off [10,14,17,18]. This work has The MG nanobands, having a distinctly different morphology from
shown that the introduction of a multicomponent Zr-based MG can not the starting spherical morphology of the powder, is direct evidence for
only result in a novel nanostructured reinforcement but also provide a the occurrence of the complete melting of the MG during SLM.
new route for tailoring the microstructure of the Ti alloy (i.e., stabili- Consequently, the formation of MG nanobands might be processed via a
zation of β phase, refinement and solution of β and α′ phases), which dissolution/re-solidification mechanism. Typically, the solidification of
can provide high strength and good ductility to the as-fabricated MMC the Ti64 liquid occurs via planar solidification as a result of the small
part. As mentioned earlier, it is recognized that the refinement of the variation between the liquidus and solidus temperatures (TL-TS = 5°
MG from micro-particles to nano-particles is beneficial to obtain high [49]). However, adding 10% MG to the Ti64 could disturb this equili-
strength and plasticity simultaneously [28–32]. We therefore supposed brium, triggering a transition from planar mode to dendritic solidifi-
that the formation of a unique nanostructured reinforcement (i.e., MG cation mode. Such a transition occurs when the thermal gradient in the

Fig. 8. The SEM images showing the fractographs of the bending fractured samples. (a, b) Ti64 alloy and (c, d) Ti64/MG composite.

506
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

affect the first term in Eq. (1), and the enlarged TL-TS range due to the
addition of Zr or Al could affect the second term in Eq. (1) [49], both of
which can lead to an increase in the critical thermal gradient. This
ultimately results in a dendritic solidification of the mixed liquids of
Ti64 and the MG.
It is noted that the variations of the density and viscosity between
the Ti64 and MG liquids have a consequence on the resultant solidified
microstructure. First, given a substantial variation in the density be-
tween the Ti64 and MG melts, i.e., ∼4.0 g cm−3 for the Ti64 liquid [51]
and ∼6.8 g cm−3 for the MG [52], a gravity effect will cause the ma-
jority of the MG melt to settle to the bottom of the melt pool. Second,
the dynamic viscosity μ of a liquid is given by Eq. (2) [13]:

16 m
μ= γ
15 KB T (2)

Fig. 10. Comparison of the wear rates between the Ti64 alloy and Ti64/MG where m is the atomic mass, γ is the surface tension (γ), T is the tem-
composite samples in air and 0.1 M NaCl solution. perature and KB is the Boltzmann constant. The Ti64 and MG liquids
exhibit almost the same surface tension (γ), with a value of ∼1.5 N m−1
liquid phase at the solidification front is below a critical value, δT
, [53,54], while the much larger m value of the MG gives rise to a higher
δXcrit μ with respect to that of the Ti64 liquid according to Eq. (2). In prin-
expressed by the following Eq. (1) [49]:
ciple, the higher density and viscosity of the MG melt allow in its se-
δT C 1 − k δTL paration from the Ti64 melt. The β phase stabilized by Cu and Ni so-
=− 0 lidifies first due to its higher melting point. The dendrites reject the
δXcrit DL / R k δC (1)
solutes (Cu, Ni and Al) across the solid/liquid interface into the su-
Wherein C0 is the overall solute concentration, DL is the solute diffusion percooled liquid, and thus the liquid becomes enriched with MG ele-
coefficient in Ti liquid, R is the solidification rate, and k is the solute ments benefiting the formation of an amorphous phase. In our recent
partition coefficient related to the solidification range between the li- work [35,36], it was proven that the local rapid cooling rate (103-106 K
quidus and solidus temperatures. The much slower diffusion coeffi- s-1) in association with the SLM process is sufficiently high for the
cients of Cu and Ni in liquid Ti [50] with respect to Al and V could formation of an amorphous structure of the current Zr-based system.

Fig. 11. The worn surface SEM morphologies and EDX analysis of the (a, c) Ti64 alloy and (b, d) Ti64/MG composite samples after wear tests in air and 0.1 M NaCl
solution. Insets in (a, b) show the SEM images of wear debris.

507
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

The rapid cooling limits the length scale of the amorphous phase reinforcements (e.g., MG micro-particles and ceramic phases) in terms
formed, yielding the MG nanobands observed in Figs. 3d and 4 b. Yet, of the ductility and toughness. The internal confinement between na-
the MG nanobands could be partially crystallized due to repeated noscale constituent phases could effectively prevent the brittle inter-
heating (> Tx temperature) upon the layer-by-layer deposition. As a metallic component, MG and martensite from premature fracture. This
result, unique “hard/soft” nanostructured reinforcements consisting of provides an opportunity for plastic deformation, which is evident from
partially crystallized MG nanobands (hard phase) entrapped in β-grains the fracture morphology observations (Fig. 8d). In fact, previous works
(soft phase) formed in the Ti64/MG composite. have readily confirmed that the confinement effect and nanoscale di-
mensions could make the brittle intermetallic and amorphous phases
4.2. Strengthening mechanisms in the as-printed Ti64/MG composite deformable [59–61]. Furthermore, the unique hybrid microstructures,
including grain boundaries, β-phase/MG nanoband interface and melt
The strength increment in the as-printed Ti64/MG composite can be zone boundaries, could make the crack propagation rather difficult
rationalized by several contributions: the mismatch of the coefficients [18], thus improving the fracture toughness. It is expected that the
of thermal expansion (CTE) between Ti64 and MG, Hall-Petch mechanical properties of the present composites could be further im-
strengthening, the load transfer effect, an Orowan strengthening proved via tailoring these structural features through processing opti-
[32,55]. mization. In a preliminary attempt, we have performed a post heat
First, based on the TEM results, no apparent dislocation was ob- treatment (annealing at 500 ℃ for 2 h in vacuum) on the composite via
served in the as-printed Ti64/MG composite around the reinforcements. the fully crystallization of the MG. However, the results of compressive
Therefore, the increased dislocation density strengthening due to the tests indicated that such heat-treatment did not increase the strength
CTE mismatch can be neglected. In fact, the CTE values of the Ti64 much; and instead, remarkably decreased the ductility (εp = 12%),
alloy (∼9 × 10−6 K-1) and Zr-based MG (∼10 × 10−6 K-1) are quite which is probably due to the formation of brittle intermetallic com-
small. Second, nanoindentation measurements revealed only a slight pounds in the Ti64/MG composite. Therefore, other post-processing
increase in the nanohardness of the matrix of the Ti64/MG composite, routes should be explored in future study.
suggesting that the Hall-Petch strengthening owing to grain refinement
might not be so significant. Third, for fiber- or whisker-reinforced 4.3. On the enhanced wear resistance of the as-printed Ti64/MG composite
MMCs, load transfer is considered one of the most important
strengthening mechanisms, wherein load transfer from a relatively soft According to the Archard equation [62], which predicts that the
matrix to a stiff second phase under an applied external load con- wear resistance of a material is proportional to its hardness, it is natural
tributes to the strengthening of the base material. The strengthening to consider that the improved wear resistance (by 24–35%) of the Ti64/
effect due to load transfer in an MMC can be calculated as below [56]: MG composite can be ascribed to its higher hardness then that of the
Ti64 alloy (increase by 11%). However, the macroscopic worn surface
l
ΔσLT = σm Vm + σf Vf ⎛1 − c ⎞ − σm (l > lc ) observations cannot provide direct evidence regarding the role of the
⎝ 2l ⎠ (3)
hard reinforcement in reducing wear loss. To this end, a micro-scratch
l test was performed on Ti64/MG composite and Ti64 alloy to obtain
ΔσLT = σm Vm + σf Vf ⎛ c ⎞ − σm (l < lc ) phase-specific friction/wear properties. The results of the scratch test
⎝ 2l ⎠ (4)
are presented in Fig. 12. The Ti64/MG composite displays a much lower
σf df average COF value (∼0.45) coupled with a narrower scratch width (see
lc =
2τm (5) inset in Fig. 12a) compared to the Ti64 alloy (COF∼0.75), again de-
monstrating the superior friction and wear resistance of the as-printed
where σm and σf are the yield strengths of the matrix (903 MPa) and Ti64/MG composite. The SEM observation (Fig. 12b) of the micro-
reinforcements (1500 MPa [36]), respectively. Vm and Vp are the vo- scratches on the Ti64 matrix and reinforcement in the composite
lume fraction of the matrix and reinforcement (5.56% in this study). l sample revealed a non-uniform wear track, wherein the scratch was
and lc represent the actual and critical lengths of the reinforcements. df wider on the matrix regions accompanied by severe plastic deformation
is the average diameter of the reinforcements (∼50 nm for an MG na- on the edges; however, the scratch became narrowed when en-
noband). τm is the ultimate shear strength of the matrix (0.65 σm countering hard reinforcement phases coupled with many fewer plastic
≈587 MPa). The calculated lc is 64 nm by Eq. (5), which is much lower deformation traces. The findings here thus directly point to the positive
than the actual length of MG nanoband (∼500 nm). As such, the cal- role of the strong, nanostructured reinforcement in the enhanced wear
culated value of Δσload is 28 MPa according to Eq. (3). Finally, Orowan resistance of the Ti64/MG composite. On the other hand, because of the
strengthening results from the interaction between dislocation and the high oxygen affinity of Ti or Zr, a passive oxide film always exists on
reinforcements. The contribution of the Orowan strengthening me- their surface. Since the Ti64 matrix is relatively soft and deformable,
chanism can be estimated by the following equation [57]: the strain mismatch between the subsurface and brittle and less-de-
0.13Gm b dp formable oxidized layer under a high load rubbing would lead to crack
ΔσOrowan = ln initiation and delamination (formed as flake-like debris), as observed in
λ 2b (6)
Fig. 11a. Differently, due to the confinement effect of the hard re-
where Gm is the shear modulus (38 GPa for Ti64 [58]), dp is the dia- inforcement that three-dimensionally distributed near the melt
meter of the reinforcement, b is the Burgers vector (0.3 nm for Ti64 boundaries against the subsurface deformation and delamination of
[58]), and λ is the interparticle distance (λ = dp [ 3 1 − 1]). The cal- tribo-oxide layer, the wear loss of the Ti64/MG composite is much less
2Vp
culated ΔσOrowan due to the nanostructured reinforcements is than that of the Ti64 alloy.
∼101 MPa. Since the total increment of the yield strength is approxi- When the as-printed parts slid in 0.1 M NaCl solution, an adhesion
mately 138 MPa, the Orowan strengthening represents most of the and strain hardening tribo-layer was formed on both materials.
contribution to the increased yield strength. Generally, the tribo-oxides fulfilled a protective role and reduced wear
In addition to the high strength, the present Ti64/MG composite under dry-sliding and lubrication conditions, as reported for many
also exhibits good ductility and fracture toughness, although being in- other materials [19]. Compared with that formed on the Ti64 alloy, the
ferior to those of the as-printed Ti64 alloy due to the existence of grain- tribo-layer on the composite appeared to be more smooth and con-
refined martensite and a brittle intermetallic phase (CuZr2). tained less oxygen, indicating a higher protectiveness and lower brit-
Nevertheless, the unique composite structure of MG nanobands coupled tleness, which, in turn, delivered a superior wet-wear resistance for the
with a ductile β phase shows advantages to those of strong yet brittle Ti64/MG composite in the 0.1 M NaCl solution. No significant corrosion

508
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

Fig. 12. (a) Coefficient of the friction (COF) as a function of the scratch distance during the micro-scratch experiment. Insets show the OM images of the scratches on
the Ti64 alloy and Ti64/MG composite. (b) SEM image of a scratched region composing of Ti64 matrix and MG-rich region in the composite.

pits could be observed on the wear tracks of the Ti64/MG composite, more protective tribo-oxide layers against corrosion wear, thus
demonstrating that the addition of MGs of good corrosion resistance providing superior wear resistance compared to the unreinforced
does not significantly degrade the anticorrosion performance of Ti-al- Ti64 alloy.
loys.
Acknowledgements
5. Conclusions
This work was financially supported by the National Key R&D
In this study, a Ti64 composite reinforced with Zr-based metallic Program of China (No. 2016YFB1100101); National Nature Science
glass (MG) was fabricated via selective laser melting (SLM). The mi- Foundation of China (Grant Nos. 51531003 and 51771077); and the
crostructure, mechanical properties and wear behaviors of the as- Hubei Provincial Natural Science Foundation of China (No.
printed Ti64/MG composite were investigated and compared to the 2018CFA003). The authors are grateful to the Analytical and Testing
Ti64 alloy, suggesting the following findings: Center, Huazhong University of Science and Technology for technical
assistance.
1 The addition of MG (10 wt%) densified the as-printed Ti64 alloy,
resulting in a near-full Ti64/MG composite with an absence of References
cracks. A unique microstructure with reinforcement consisting of
partially crystallized MG nanobands and fine β grains formed and [1] X.Y. Liu, P.K. Chu, C.X. Ding, Surface modification of titanium, titanium alloys, and
distributed along the melt pool boundaries in the Ti64/MG com- related materials for biomedical applications, Mater. Sci. Eng. R. 47 (2004) 49–121.
[2] D. Banerjee, J.C. Williams, Perspectives on titanium science and technology, Acta
posite. Mater 61 (2013) 844–978.
2 The Ti64/MG composite exhibited a high yield strength of larger [3] A. Devaraj, V.V. Joshi, A. Srivastava, S. Manandhar, V. Moxson, V.A. Duz,
than 1 GPa under compression and bending, excellent compressive C. Lavender, A low-cost hierarchical nanostructured beta-titanium alloy with high
strength, Nat. Commun. 7 (2016) 11176.
ductility (20%) and fairly good fracture toughness (Kq = 42 MPa [4] H. Mohseni, P. Nandwana, A. Tsoi, R. Banerjee, T.W. Scharf, In situ nitride titanium
m1/2). Based on a theoretical analysis, the increased strength is alloys: microstructure evolution during solidification and wear, Acta Mater. 83
mainly attributed to the Orowan strengthening mechanism, while (2015) 61–74.
[5] Y. Jiao, L.J. Huang, L. Geng, Progress on discontinuously reinforced titanium matrix
the good ductility and fracture toughness result from the ductile β composites, J. Alloys Compd. 767 (2018) 1196–1215.
phase and the confinement effects between the nanoscale con- [6] T. Christman, A. Needleman, S. Nutt, S. Suresh, On microstructural evolution and
stituent phases and various types of interfaces in the composite. micromechanical modelling of deformation of a whisker-reinforced metal-matrix
composite, Mater. Sci. Eng. A 107 (1989) 49–61.
3 The hardness and wear performance was also improved in the re-
[7] S.C. Tjong, Z.Y. Ma, Microstructural and mechanical characteristics of in situ metal
sultant Ti64/MG composite. The hard nanostructured reinforce- matrix composites, Mater. Sci. Eng. R. 29 (2000) 49–113.
ments had been directly proven through micro-scratch tests to re- [8] L.C. Zhang, H. Attar, Selective laser melting of titanium alloys and titanium matrix
duce the COF and wear volume. 3D uniformly distributed composites for biomedical applications: a review, Adv. Eng. Mater. 18 (2016)
463–475.
reinforcements are able to reduce the delamination wear caused by [9] T. DebRoy, H.L. Wei, J.S. Zuback, T. Mukherjee, J.W. Elmer, J.O. Milewski,
the subsurface plastic deformation of the matrix and also generate A.M. Beese, A. Wilson-Heid, A. De, W. Zhang, Additive manufacturing of metallic

509
X.-J. Shen et al. Additive Manufacturing 25 (2019) 499–510

components-process, structure and properties, Prog. Mater. Sci. 92 (2018) 112–224. (2017) 128–134.
[10] D. Herzog, V. Seyda, E. Wycisk, C. Emmelmann, Additive manufacturing of metals, [37] E.W. Qin, L. Lu, T.R. Tao, J. Tan, K. Lu, Enhanced fracture toughness and strength in
Acta Mater. 117 (2016) 371–392. bulk nanocrystalline Cu with nanoscale twin bundles, Acta Mater. 57 (2009)
[11] P. Barriobero-Vila, J. Gussone, A. Stark, N. Schell, J. Haubrich, G. Requena, 6215–6225.
Peritectic titanium alloys for 3D printing, Nat. Commun. 9 (2018) 3426. [38] M. Simonelli, Y.Y. Yse, C. Tuck, Effect of the build orientation on the mechanical
[12] L. Thijs, F. Verhaeghe, T. Craeghs, J.V. Humbeeck, J.P. Kruth, A study of the mi- properties and fracture modes of SLM Ti-6Al-4V, Mater. Sci. Eng. A 616 (2014)
crostructure evolution during selective laser melting of Ti-6Al-4V, Acta Mater. 58 1–11.
(2010) 3303–3312. [39] Y.Q. Li, Y.Y. Shen, M.C. Leu, H.L. Tsai, Building Zr-based metallic glass part on Ti-
[13] D.D. Gu, Y.C. Hagedorn, W. Meiners, G.B. Meng, R.J.S. Batista, K. Wissenbach, 6Al-4V substrate by laser-foil-printing additive manufacturing, Acta Mater. 144
R. Poprawe, Densification behavior, microstructure evolution, and wear perfor- (2018) 810–821.
mance of selective laser melting processed commercially pure titanium, Acta Mater. [40] J.W. Qiao, A.C. Sun, E.W. Huang, Y. Zhang, P.K. Liaw, C.P. Chuang, Tensile de-
60 (2012) 3849–3860. formation micromechanisms for bulk metallic glass matrix composites: from work-
[14] W. Xu, M. Brandt, S. Sun, J. Elambasseril, Q. Liu, K. Latham, K. Xia, M. Qian, hardening to softening, Acta Mater. 59 (2011) 4126–4137.
Additive manufacturing of strong and ductile Ti-6Al-4V alloy by selective laser [41] H.L. Jia, G.Y. Wang, S.Y. Chen, Y.F. Gao, W.D. Li, P.K. Liaw, Fatigue and fracture
melting via in situ martensite decomposition, Acta Mater. 85 (2015) 74–84. behavior of bulk metallic glasses and their composites, Prog. Mater. Sci. 98 (2018)
[15] C.L. Qiu, C. Panwisawas, M. Ward, H.C. Basoalto, J.W. Brooks, M.M. Attallah, On 168–248.
the role of melt flow into the surface structure and porosity development during [42] S. Yang, X.C. Li, R. Wei, M. Meng, L. He, S.F. Zhang, Deformation characteristic of a
selective laser melting, Acta Mater. 96 (2015) 72–79. Ti-based bulk metallic glass composite with fine microstructure, Mater. Sci. Eng. A
[16] Y.J. Liu, S.J. Li, H.L. Wang, W.T. Hou, Y.L. Hao, R. Yang, T.B. Sercombe, 733 (2018) 224–231.
Microstructure, defects and mechanical behavior of beta-type titanium porous [43] R.L. Narayan, P.S. Singh, D.C. Hofmann, N. Hutchinson, K.M. Flores,
structures manufactured by electron beam melting and selective laser melting, Acta U. Ramamurty, On the microstructure-tensile property correlations in bulk metallic
Mater. 113 (2016) 56–67. glass matrix composites with crystalline dendrites, Acta Mater. 60 (2012)
[17] W. Xu, E.W. Lui, A. Pateras, M. Qian, M. Brandt, In situ tailoring microstructure in 5089–5100.
additively manufactured Ti-6Al-4V for superior mechanical performance, Acta [44] X. Zhang, F. Martina, J. Ding, X. Wang, S.W. Williams, Fracture toughness and
Mater. 125 (2017) 390–400. fatigue crack growth rate properties in wire arc additive manufactured Ti-6Al-4V,
[18] P. Kumar, O. Prakash, U. Ramamurty, Micro- and meso-structures and their influ- Fat. Fract. Eng. Mater. Struct. 40 (2017) 790–803.
ence on mechanical properties of selectively laser melted Ti-6Al-4V, Acta Mater. [45] J.J. Lewandowski, M. Seifi, Metal additive manufacturing: a review of mechanical
154 (2018) 246–260. properties, Annu. Rev. Mater. Res. 46 (2016) 151–186.
[19] A. Attar, K.G. Prashanth, A.K. Chaubey, M. Calin, L.C. Zhang, S. Scudino, J. Eckert, [46] J. Suryawanshi, K.G. Prashanth, S. Scudino, J. Eckert, Prakash Om, U. Ramamurty,
Comparison of wear properties of commercially pure titanium prepared by selective Simultaneous enhancement of strength and toughness in an Al-12Si alloy synthe-
laser melting and casting processes, Mater. Lett. 142 (2015) 38–41. sized using selective laser melting, Acta Mater. 115 (2016) 285–294.
[20] Y. Zhu, X. Chen, J. Zou, H. Yang, Sliding wear of selective laser melting processed [47] S.K. Makineni, A.R. Kini, J.H. Springer, D. Raabe, B. Gault, Synthesis and stabili-
Ti6Al4V under boundary lubrication conditions, Wear 368-369 (2016) 485–495. zation of a new phase regime in a Mo-Si-B based alloy by laser-based additive
[21] H. Attar, S. Ehtemam-Haghighi, D. Kent, M.S. Dargusch, Recent developments and manufacturing, Acta Mater. 151 (2018) 31–40.
opportunities in additive manufacturing of titanium-based matrix composites: a [48] C.J. Han, Y. Li, Q. Qang, D.S. Cai, Q.S. Wei, L. Yang, S.F. Wen, J. Liu, Y.S. Shi,
review, Inter. J. Mach. Tools Manuf. 133 (2018) 85–102. Titanium/hydroxuapatite (Ti/HA) gradient materials with quasi-continuous ratios
[22] S. Kumar, J.P. Kruth, Composites by rapid prototyping technology, Mater. Des. 31 fabricated by SLM: materials interface and fracture toughness, Mater. Des. 141
(2010) 850–856. (2018) 256–266.
[23] H. Attar, M. Bonisch, M. Calin, L.C. Zhang, S. Scudino, J. Eckert, Selective laser [49] B. Vrancken, L. Thijs, J.P. Kruth, J. Van Humbeeck, Microstructure and mechanical
melting of in situ titanium-titanium boride composites: processing, microstructure properties of a novel beta titanium metallic composite by selective laser melting,
and mechanical properties, Acta Mater. 76 (2014) 13–22. Acta Mater. 68 (2014) 150–158.
[24] N. Kang, P. Coddet, Q. Liu, H.L. Liao, C. Coddet, In-situ TiB/near α Ti matrix [50] H. Nakajima, M. Koiwa, Diffusion in titanium, ISIS Int. 31 (1991) 757–766.
composites manufactured by selective laser melting, Add. Manuf. 11 (2016) 1–6. [51] A. Schmon, K. Aziz, G. Pottlacher, Density of liquid Ti-6Al-4V, EPJ Web Conf. 151
[25] D.D. Gu, H.Q. Wang, G.Q. Zhang, Selective laser melting additive manufacturing of (2017) 04003.
Ti-based nanocomposites: the role of nanopowder, Matal. Mater. Trans. A 45A [52] T. Nasu, S. Kanazawa, S. Hayashizaki, S.X. Zhao, S. Takahashi, T. Usuki, T. Kameda,
(2014) 464–476. Work softening behavior of zirconium-aluminum-nickel-copper bulk metallic glass
[26] A.L. Maximenko, E.A. Olevsky, Pore filling during selective laser melting-assisted by rolling, Mater. Trans. 56 (2015) 249–252.
additive manufacturing of composites, Scr. Mater. 149 (2018) 75–78. [53] R. Aune, L. Battezzati, R. Brook, I. Egry, H.J. Fecht, J.P. Garandet, K.C. Mills,
[27] C.L. Ma, D.D. Gu, D.H. Dai, H.M. Zhang, L. Du, H. Zhang, Development of interfacial A. Passerone, P.N. Quested, E. Ricci, S. Schneider, S. Seetharaman,
stress during selective laser melting of TiC-reinforced TiAl composites: influence of R.K. Wunderlich, B. Vinet, Surface tension and viscosity of industrial alloys from
geometric feature of reinforcement, Mater. Des. 157 (2018) 1–11. parabolic flight experiments-results of the Thermllab project, Microgravity-Sci.
[28] C.A. Schuh, T.C. Hufnagel, U. Ramamurty, Mechanical behavior of amorphous al- Technol. XVI-1 (2005) 11–14.
loys, Acta Mater. 55 (2007) 4067–4109. [54] S. Mukherjee, W.L. Johnson, W.K. Rhim, Noncontact measurement of high-tem-
[29] D.V. Dudina, K. Georgarakis, Y. Li, M. Aljerf, A. LeMoulec, A.R. Yavari, A. Inoue, A perature surface tension and viscosity of bulk metallic glass-forming alloys using
magnesium alloy matrix composite reinforced with metallic glass, Compos. Sci. the drop oscillation technique, Appl. Phys. Lett. 86 (2005) 014104.
Technol. 69 (2009) 2734–2736. [55] X.P. Li, G. Ji, Z. Chen, A. Addad, Y. Wu, H.W. Wang, Vleugels, J. Van Humbeeck,
[30] S. Scudino, G. Liu, K.G. Prashanth, B. Bartusch, K.B. Surreddi, B.S. Murty, J. Eckert, J.P. Kruth, Selective laser melting of nano-TiB2 decorated AlSi10Mg alloy with high
Mechanical properties of Al-based metal matrix composites reinforced with Zr- fracture strength and ductility, Acta Mater. 129 (2017) 183–193.
based glassy particles produced by powder metallurgy, Acta Mater. 57 (2009) [56] X. Zhang, S.F. Li, B. Pan, D. Pan, S.Y. Zhou, S.H. Yang, L. Jia, K. Kondoh, A novel
2029–2039. strengthening effect of in-situ nano Al2O3w on CNTs reinforced aluminum matrix
[31] M. Aljerf, K. Georgarakis, D. Louzguine-Luzgin, A. Le Moulec, A. Inoue, A.R. Yavari, nanocomposites and the matched strengthening mechanisms, J. Alloys Compd. 764
Strong and light metal matrix composites with metallic glass particulate re- (2018) 279–288.
inforcement, Mater. Sci. Eng. A 532 (2012) 325–330. [57] Z. Zhang, D.L. Chen, Contribution of Orowan strengthening effect in particulate-
[32] W.W. Zhang, Y. Hu, Z. Wang, C. Yang, G.Q. Zhang, K.G. Prashanth, reinforced metal matrix nanocomposites, Mater. Sci. Eng. A 483-484 (2008)
C. Suryanarayana, A novel high-strength Al-based nanocomposite reinforced with 148–152.
Ti-based metallic glass nanoparticles produced by powder metallurgy, Mater. Sci. [58] X.P. Tan, Y.H. KoK, Y.J. Tan, M. Descoins, D. Mangelinck, S.B. Tor, K.F. Leong,
Eng. A 734 (2018) 34–41. C.K. Chua, Graded microstructure and mechanical properties of additive manu-
[33] C. Zhang, H. Zhou, L. Liu, Laminar Fe-based amorphous composite coatings with factured Ti-6Al-4V via electron beam melting, Acta Mater. 97 (2015) 1–16.
enhanced bonding strength and impact resistance, Acta Mater. 72 (2014) 239–251. [59] Z. Wang, R.T. Qu, S. Scudino, B.A. Sun, K.G. Prashanth, D.V. Louzguine-Luzgin,
[34] Y.J. Zhang, J.L. Zhang, Q. Yan, L. Zhang, M. Wang, B. Song, Y.S. Shi, Amorphous M.W. Chen, Z.F. Zhang, J. Eckert, Hybrid nanostructured aluminum alloy with
alloy strengthened stainless steel manufactured by selective laser melting: enhanced super-high strength, NPG Asia Mater 7 (2015) 229.
strength and improved corrosion resistance, Scr. Mater. 148 (2018) 20–23. [60] J.R. Greer, J.ThM. De Hosson, Plasticity in small-sized metallic systems: intrinsic
[35] C. Yang, C. Zhang, W. Xing, L. Liu, 3D printing of Zr-based bulk metallic glasses versus extrinsic size effect, Prog. Mater. Sci 56 (2011) 654–724.
with complex geometries and enhanced catalytic properties, Intermetallics 94 [61] R. Liontas, J.R. Greer, 3D nano-architechted metallic glass: size effect suppresses
(2018) 22–28. catastrophic failure, Acta Mater. 133 (2017) 393–407.
[36] D. Ouyang, N. Li, W. Xing, J.J. Zhang, L. Liu, 3D printing of crack-free high strength [62] J.F. Archard, Contact and rubbing of flat surfaces, J. Appl. Phys. 24 (1953) 981.
Zr-based bulk metallic glass composite by selective laser melting, Intermetallics 90

510

Das könnte Ihnen auch gefallen