You are on page 1of 9

Journal of Non-Crystalline Solids 546 (2020) 120284

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/locate/jnoncrysol

Strengthening and toughening mechanisms of metallic glass nanocomposites T


via graphene nanoplatelets
⁎,b
Zhuocheng Xiea,c, Wu-Rong Jian , Xiaochang Tanga, Xiaoqing Zhanga, Xiaohu Yaoa,c
a
Department of Engineering Mechanics, South China University of Technology, Guangzhou, Guangdong 510640, P. R. China
b
Department of Mechanical Engineering, University of California, Santa Barbara, Santa Barbara, California 93106, USA
c
State Key Laboratory of Subtropical Building Science, South China University of Technology, Guangzhou, Guangdong 510640, PR China

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

Keywords: The lack of ductility, mainly due to shear localization, limits the full exploitation of metallic glasses (MGs). Such
Strengthening and toughening a weakness can be mitigated via introducing graphene nanoplatelets (GNs), which helps strengthen and toughen
Metallic glass nanocomposites MG matrix. Using molecular dynamics simulations, we investigate the tensile deformation of Cu50Zr50 metallic
Graphene nanoplatelets glass-graphene nanoplatelet composites (MGGNCs) regarding the effects of graphene volume fraction and gra-
Molecular dynamics simulations
phene layer number. Increasing fraction of GN in MGGNC enhances both tensile strength and fracture strain. The
dominant component changes from MG matrix to GN, corresponding to a failure mode transition from shear
banding in MG to brittle fracture in graphene. For a given volume fraction, single-layered GN behaves better
than its double-layered counterpart, manifested as a higher fracture strain in MGGNCs. Our study illustrates the
strengthening and toughening mechanisms in MGGNCs, and why single-layered GN performs better. All of these
results can contribute to the synthesis of novel MG/graphene composites.

1. Introduction crystalline nanolaminates and introduced grain boundaries or glass-


glass interfaces into the laminates or the bulk MGs, which boosted the
Metallic glasses (MGs) have attracted great research interests over plasticity of the MG matrix, hence balancing the trade-off between
the past decades for superior strength and high elastic limit, but they ductility and strength as well as hardness.
generally lack tensile ductility that results in sudden and catastrophic Meanwhile, graphene was invented as an excellent material with
failure [1–4]. It is found that the rapid propagation of shear bands (SBs) high intrinsic strength and Young’s modulus [17–20], and three-di-
formed by localizing shear transformation zones (STZs), is the chief mensional graphene-reinforced metallic composites have become in-
culprit in most abrupt failures of MGs, especially under tensile loading creasingly interesting in recent years [21–24]. Kim et al. [25] conduted
[5–7]. Therefore, solutions are investigated to avoid shear banding of nanopillar compression experiment and molecular dynamics simula-
MGs and attain high ductility. Through all approaches to large ductility, tions to Cu (or Ni)-graphene nanolayered composites and reported that
one is reducing the external dimensions of MGs to the order of 100 nm dislocations were unable to penatrate graphene-metal interface, leading
[8–10], and the other common access is developing heterogeneity in to high strength of composites. Similar compression experiment for
MGs by the adoption of the second phase. For example, introducing graphene-Al nanolaminates indicated that graphene strengthened and
amorphous-crystalline interfaces can effectively impede the propaga- toughened the metal matrix as a result of the enhanced interface-dis-
tion of SBs in MGs, leading to a more homogeneous strain distribution location interactions and crack deflection by the robust graphene-Al
and thus better ductility [11,12]. Hofmann et al. claimed that utilizing interfaces in the composites [26–28]. Hwang et al. [29] found that
the inhomogeneous microstructure with isolated dendrites in a MG graphene blocked fatigue cracks generated during cyclic bending tests
matrix could stabilize the glass against the catastrophic failure asso- in Cu-graphene nanolayered composites, which greatly improved their
ciated with unlimited extension of a shear band and result in enhanced ductility. Liu et al. [30–32] systematically studied Cu-graphene nano-
global plasticity [13]. Besides, Zhou et al. [14] took advantage of layered composites under shear loading and the results showed that
amorphous-amorphous interfaces and soft amorphous component to graphene could amplify anti-fatigue properties by hindering disloca-
constrain SB propagation in MGs, which also enhanced ductility. tions propagation.
Moreover, Jian et al. [15,16] tuned the layer thickness of amorphous- Since graphene has been thought to be an ideal reinforcement


Corresponding author.
E-mail addresses: wurong@ucsb.edu (W.-R. Jian), tcqzhang@scut.edu.cn (X. Zhang).

https://doi.org/10.1016/j.jnoncrysol.2020.120284
Received 13 May 2020; Received in revised form 1 July 2020; Accepted 2 July 2020
0022-3093/ © 2020 Elsevier B.V. All rights reserved.
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

material, the combination of MGs and graphene for enhanced strength listed in Table 1. For MGDLGNCs, distance between two graphene layer
and ductility is taken into account. Park et al. [33] reported experi- is 3.4 Å [46]. The chirality of graphene in all samples along x axis is
mental evidence that MG-graphene nanolaminate exhibited high elastic zigzag.
modulus and yield strength. Homogeneous flow improving tensile Boundary conditions of all samples along three dimensions are set to
ductility of nanolaminates was realized through tailoring the thickness be periodic during equilibration under isothermal-isobaric (NPT) en-
of MG layers. To the best of our knowledge, previous studies of gra- semble with a constant temperature at 50 K and zero pressure. Here the
phene-reinforced nanocomposites mostly focus on their mechanical low temperature (50 K) is used to highlight the material responses upon
properties and the reinforcement effects of either single - layered gra- mechanical activation, since the simulated glass structure constructed
phene [34–37] or different distributions of single - layered graphene with high quenching rate is usually more prone to thermal activation
[38], neglecting the interaction between different layers of graphene than MGs used in experiments, which have much longer time to relax
due to the existence of interface [39,40] and their behavior during [47]. After relaxation, uniaxial tension along the x axis at a constant
deformation, which should be compared with their single - layered strain rate of 1 × 108 s−1 is imposed to all samples, during which
counterpart when chosing the ideal reinforcement. Moreover, while boundary conditions along x and z axes are remained periodic (zero
some computational work have been conducted to study the mechan- pressure along z direction) while that along y axis is adjusted to shrink -
ical behaviors of carbon nanotube (or Cu nanowire)-MG nanocompo- wrapped. Timestep in all MD simulations is 2 fs. Such a timestep is
sites [41,42], simulation investigations at atomistic scale about MG- widely used in MG uniaxial simulations at the strain rate of ~ 109 s−1,
graphene nanocomposites have been rarely reported [38,43]. Since it is which have been proved to be capable of fully capturing the accumu-
generally difficult to fully capture microscopic phenomena in-situ ex- lation of STZs and the formation of SBs [48,49]. Moreover, a strain
periments, molecular dynamics (MD) simulations serve as a powerful increment of about 5 ∼ 10 × 10−7 (corresponding to a timestep of
tool to explain atomistic mechanisms at nano scale [16,44]. 0.5 ~ 1 fs at the strain rate of 109 s−1) provides enough accuracy for
In the present work, we perform MD simulations to study the tensile describing the deformation and fracture processes of graphene under
behaviors of metallic glass-graphene nanoplatelet composites uniaxial loading [50,51]. Here, the step length of tensile strain is
(MGGNCs) with different graphene volume fractions and graphene 2 × 10−7, which is believed to be accurate enough for reflecting the
layer numbers. As shown by previous studies about metallic glass deformation details and illustrating the deformation mechanisms of
consisting of different composites, CuZr MG has been widely in- MGGNCs. Simulations are performed using the Large - scale Atomic/
vestigated with relatively outstanding properties such as high elastic Molecular Massively Parallel Simulator (LAMMPS) [52] and the cor-
modulus and ultimate strength [41,42]. Thus we select Cu50Zr50 MG as responding results are visualized by the Open Visualization Tool
a representative for further investigation. For MGs reinforced by (OVITO) [53].
single - layered graphene nanoplatelets (SLGNs), increasing graphene
volume fraction significantly improves ultimate tensile strength and
toughness of MGGNCs, and plastic deformation mode transits from 2.2. Interatomic potentials
shear localization to homogeneous flow as graphene volume fraction
rises. Moreover, less ductility is observed in composites with double - Since atomic interactions for MGGNCs include: (1) the interactions
layered graphene nanoplatelets (DLGNs) than those with SLGN at the between metallic atoms in the MG matrix, (2) the interactions between
approximate graphene volume fraction. The subsequent analysis re- carbon atoms in GN, and (3) the interactions between carbon atoms and
veals the poor behavior of MGGNCs with DLGN is ascribed to the in- metallic atoms, different potentials are considered to describe the in-
compatible deformation between two graphene layers. teractions above. In this study, embedded atom method (EAM) poten-
The paper is organized as follows. Details of atomistic simulations tial developed by Mendelev et al. [54] is used to simulate metallic
including modeling, loading and analysis methods, are presented in atomic interactions. Additionally, AIREBO potential [20] is adopted to
Section 2. In Section 3, strengthening and toughening mechanisms as define interactions among carbon atoms. Van der Waals forces between
well as failure modes of MGGNCs are discussed. The paper is concluded GN and MG matrix is described by Lennard-Jones potential, which has
in Section 4. been widely utilized to convey non-bonded interactions, as shown in
Eq. 1:
2. Methodology
σ σ
ELJ = 4ϵ[ ⎛ ⎞ 12 − ⎛ ⎞ 6],
2.1. Models and computations ⎝r ⎠ ⎝r ⎠ (1)

A small Cu50Zr50 MG model, composed of 10,000 atoms, is built by where σ is the interatomic separation when the value of potential is zero
melting the initial alloy configuration at 2000 K and then cooling it to and ϵ is the depth of the potential well. The related parameters are
50 K at the rate of 0.01 K/ps. During this process, constant-pressure- listed in Table 2, which are calculated according to methods proposed
temperature (NPT) ensemble and three-dimensional periodic boundary by Reza Rezaei using Lorentz–Berthelot mixing rule [55]. For Cu,
conditions are applied. The detailed procedures to produce the small σCu − Cu = 2.29726, ϵCu − Cu = 0.52031 were used as in Ref [56]. For Zr,
MG cube can be found in Ref. [45]. To construct the final Cu50Zr50 MG σZr − Zr = 2.9318, ϵZr − Zr = 0.73847 were adopted following Ref [57]. For
matrix, the small MG cube is replicated to reach a large sample with the non-bonded C atoms, σC − C = 3.4, ϵC − C = 0.00284 were taken from
required dimensions. In addition to one pure MG, eight different con- Ref [20]. σC-Cu and σC-Zr are obtained via σ12 = (σ11 + σ22)/2 while ϵC-Cu
figurations of MGGNCs are taken into account in this study and their and ϵC-Zr are calculated by ϵ12 = ϵ11ϵ22 . This fitting method and values
schematics are shown in Fig. 1. As Fig. 1 (a) shows, dimensions of all for LJ parameters of non - equal atoms have been validated by previous
configurations along x-, y-, z- directions are ~ 58 nm × 29 nm × studies to depict the interfacial potentials between C-Cu and C-Zr
5.8 nm, respectively. Sample I is pure MG, II to VII are metallic glass- [38,43,55]. The cutoff distance for Lennard - Jones potential is set as
single - layered graphene nanoplatelet composites (MGSLGNCs) with rc = 8.06 Å. The total potential energy for MGGNCs is calculated as
different graphene volume fractions (The length of graphene along x- Eq. 2:
axis is the same as that of MG). Sample VIII and IX are metallic glass- C − Cu C − Zr
Etotal = EEAM + EAIREBO + ELJ + ELJ . (2)
double - layered graphene nanoplatelet composites (MGDLGNCs).
Given a certain graphene volume fractions, V and VIII (or VI and IX) are
the control group, in which graphene layer number is the variable. The
layer number and volume fraction of graphene of all configurations are

2
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 1. (a) Three-dimensional sizes of each sample. Atomic configurations with different graphene volume fractions: (b) 0%, (c) 0.17%, (d) 0.52%, (e) 1.03%, (f)
1.63%, (g) 3.27%, (h) 5.76%, (i) 1.72%, (j) 3.27%. (b) denotes pure MG. The configurations in (c)–(h) and (i)–(j) have single-layered graphene nanoplatelets and
double-layered graphene nanoplatelets inserted, respectively.

Table 1 whole sample, corresponding to the sharp drop after peak stress in
The layer number (n) and graphene volume fraction (vGr%) in all samples stress-strain curve. In contrast, increasing graphene volume fraction in
(pure MG or MGGNCs). MGSLGNCs can significantly improve the ultimate tensile strength and
Sample n vGr% toughness of composites. When graphene volume fraction is at or under
1.63%, all curves of MGSLGNCs exhibit three stages. Here, we take
I ––– ––– sample II as an example (Fig. 3 (a)). Stage A represents the elastic
II one 0.17%
period and initial plastic stage, during which the stress continues to
III one 0.52%
IV one 1.03% rise. In stage B, notable plastic deformation occurs and the stress de-
V one 1.63% clines upon reaching the ultimate tensile strength, which corresponds to
VI one 3.27% the formation of a single, dominant shear band. In stage C, stress
VII one 5.76% fluctuates with the increase of strain, as SB continues to grow and the
VIII two 1.72%
IX two 3.27%
sample starts sliding along the SB. Finally, there is a rapid drop of stress.
Such a stress decrease indicates graphene has fractured and the sample
fails. As shown in last inset of Fig. 3 (a), SLGN is rolled seriously in the
Table 2 region where SB traverses the sample, leading to the fracture of SLGN.
The parameters of LJ potential for different types of atomic interactions. For MGSLGNCs with graphene volume fractions of 3.27% and 5.76%
(sample VI and VII), their tensile stresses increase continuously to the
Atomic interaction σ\Å ϵ\eV
maximum value, where the fracture of graphene begins. For example,
C-Cu 2.84863 0.03844 the stress-strain curve of sample VI is displayed in Fig. 3 (b). Without a
C-Zr 3.1659 0.0458 dominant SB in this situation, SLGN seems to be ripped directly for
overloading, resulting in the sudden failure of the sample. By com-
paring Fig. 3 (a) and (b), there exists a transition of dominant de-
3. Results and discussion formation mechanism from shear banding in MG matrix to brittle
fracture in graphene.
3.1. MGSLGNCs

3.1.1. The tensile behaviors of MGSLGNCs 3.1.2. Shear banding in MGSLGNCs


Fig. 2 depicts the tensile engineering stress - strain curves for pure To further understand shear banding in MGSLGNCs and the tran-
MG and MGSLGNCs with graphene volume fractions ranging from sition of the deformation mechanism, atomic shear strain analysis [58]
0.17% to 5.76%. Pure MG expires when SBs propagate through the is performed. Since the regions with large atomic shear strain also bear
severe localized plastic deformation, we only remain MG atoms with
atomic shear strain above 0.4, corresponding to the STZ clusters. For
samples with graphene volume fraction not exceeding 1.63% (sample I-
5), Fig. 4 shows snapshots of STZs and SLGNs during tensile loading. As
shown in Fig. 4 (a), STZs are triggered at the free surface of pure MG
sample and increase with further loading. After STZs converging into a
single SB, its propagation across the sample results in the subsequent
failure. Slightly different from those in pure MG, STZs in MGSLGNCs
are also generated at MG-SLGN interfaces where stress concentration
prevails (See the insert of Fig. 4 (d)). With higher graphene volume
fractions, the initiation of STZs is delayed by more pronounced re-
inforcement effects of graphene, characterized by later moment that
STZs appear (See the first columns of Fig. 4 (a)–(e)). Thus, the propa-
gation of SB is also postponed. Meanwhile, SLGNs remain undamaged
and stay in a mild atomic strain level during the activation and growth
of STZs. Therefore, the promotion in the tensile strength of MGSLGNCs
results from the load sharing of SLGN and the improvement of tough-
ness in MGSLGNCs is due to the obstruction to single dominant SB
generation and the succedent load-bearing capacity of SLGN after the
formation of SB.
Fig. 2. Tensile stress - strain curves of pure MG and MGSLGNCs with different To verify the influence of graphene volume fraction on the sup-
graphene volume fractions. pression to SB, the thickness of SB is measured after removing MG

3
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 3. Schematic illustrations of two representative types of stress-strain curves for MGSLGNCs with different graphene volume fractions, namely (a) sample II,
three-stage form and (b) sample VI, strain hardening form. The atoms in the insets are colored by atom shear strain.

atoms with atomic strain value under 0.4. Fig. 5 illustrates the evolu- These results confirm the transition of two governing deformation
tion of a single dominant SB during tensile loading, the profile of which mechanisms. Preeminent reinforcement effects of SLGN resist the for-
is outlined with dash lines. As illustrated in Fig. 5 (a), we denote the mation of a single dominant SB and multiple small STZs diffuse the
thickness of SBs as ‘T’ and it is estimated from the arithmetic average, strain distribution. Thus, plastic deformation pattern transits from shear
T
i.e., T = max
+
Tmin 2, where ‘Tmax ’ and ‘Tmin ’ represent the widest and the localization to homogeneous deformation. It is worth mentioning that
most narrow width of SB, respectively. As summarized in Fig. 6, the graphene controls the fracture strain of all MGSLGNCs, illustrating that
thickness of the dominant SB decreases with graphene volume fraction graphene volume fraction determines the toughness of MGSLGNCs.
increasing. Especially, for sample V, SLGN brings remarkable effect on With the increasing graphene volume fraction, tensile strength and
the suppression to the dominant SB. Obviously, the domination of the fracture strain of MG composite can be enhanced significantly
single SB vanishes with the increase in graphene volume fraction. By (Table 3).
contrast, the sample with graphene volume fraction above 1.63% (e.g.,
sample VI in Fig. 7), STZs distribute uniformly in the whole sample,
making the entire sample deform in a homogenous way. Hence, the
dominant SB is absent.

Fig. 4. The evolution from STZs to a single SB in (a) pure MG and (b)–(e) MGSLGNCs with different graphene volume fractions. Only MG atoms with atomic shear
strain above 0.4 and graphene atoms are shown. The former are all gray and the latter are colored by atomic shear strain.

4
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 5. The evolution of a single SB in (a)


pure MG and (b)–(e) MGSLGNCs with dif-
ferent graphene volume fractions. The pro-
files of SBs are hinted by dash lines and
their thicknesses are denoted as ’T’. Only
MG atoms with atomic shear strain above
0.4 are shown in purple. (For interpretation
of the references to colour in this figure le-
gend, the reader is referred to the web ver-
sion of this article.)

Table 3
Tensile behaviors of pure MG and MGSLGNCs with different graphene volume
fractions (vGr%).
Sample vGr% Tensile strength\GPa Fracture strain\%

I ––– 3.90 11.4


II 0.17% 3.94 14.04
III 0.52% 4.14 17.25
IV 1.03% 4.53 17.31
VI 3.27% 6.83 20.73
VII 5.76% 9.90 21.12

3.2. MGDLGNCs

3.2.1. The tensile behaviors of MGDLGNCs


After considering the effect of graphene volume fraction on the
strength and ductility of MGSLGNCs, we take into account the role of
graphene layer number in deformation behaviors. Herein, MGDLGNCs,
with graphene volume fractions of 1.72% and 3.27% (sample VIII in
Fig. 6. The thicknesses of SBs in pure MG and MGSLGNCs with different gra-
Fig. 1 (i) and IX in Fig. 1 (j)), are compared to their MGSLGNC coun-
phene volume fractions.
terparts with the approximating graphene volume fractions (1.63% for
sample V and 3.27% for sample VI), respectively. From the stress -
strain curves in Fig. 8 (a), it is seen that each pair of curves with similar
graphene volume fraction coincide with each other before stress drop or
further enhancement in plastic deformation, emphasizing that the
elastic and initial plastic regime prominently depends on graphene
volume fraction. This is because MG is sufficient to withstand external
loads and the number of GN rarely affects deformation behaviors
during these periods. Nevertheless, once plastic deformation proceeds,
the curves of MGDLGNCs fall below their MGSLGNC counterparts. This
suggests smaller single-layered graphene surface area in MGDLGNCs
gives rise to a limited enhancement in comparison to the corresponding
cases of MGSLGNCs at given graphene volume fraction. Consequently,
Fig. 7. The evolution of STZs in sample VI. Only MG atoms with atomic shear SLGN imposes more considerable strengthening effect than DLGN at a
strain above 0.4 and graphene atoms are shown. The former are all gray and the given graphene volume fraction. Akin to Fig. 3 (a) and (b), the curves in
latter are colored by atomic shear strain. Fig. 8 (b) and (c) also show three-stages mode and strain-hardening
mode, respectively. Such phenomena reveal that the transition in
plastic deformation mode mentioned above also exists in MGDLGNCs.

5
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 8. (a) Tensile Stress - strain curves of MGSLGNCs and MGDLGNCs (plotted in dash lines) with the approximating graphene volume fractions. Schematic
illustrations of two representative types of curves for MGDLGNCs with different graphene volume fractions, namely (b) sample VIII, three-stage form and (c) sample
IX, strain hardening form.

Interestingly, although graphene has an important impact on the frac- its movement have significant effect on the warp of DLGN, which leads
ture strain of all MG - GN composites, MGDLGNC shows a worse duc- to the deformation incompatibility in different neighboring regions of
tility than MGSLGNC at similar graphene volume fraction. What we DLGN (Fig. 10 (a)). As a result, DLGN prones to fracture.
observe indicates that the ductility of MG-GN composites is affected by In addition, the similar computation involving interface distances is
both graphene volume fraction and the number of GNs. applied to sample IX. The calculation results of upper GN-lower GN
interface, upper MG-upper GN interface and lower GN-lower MG in-
terface at the tensile strain of 16.07% are illustrated in Fig. 11 (a)–(c),
3.2.2. The interfaces of GN-GN and MG-GN respectively. Unlike the deformation mode of the previous MG - DLGN
To explain the unsatisfactory ductility of MGDLGNCs, we pay at- sample mentioned above, the interface distance distribution shown in
tention to GN-GN and MG-GN interfaces (Fig. 9). Here, only GN layer Fig. 11 are much more homogeneous in most areas, which results from
and the neighboring MG layers with the thickness of 2 Å are shown. At the absence of the dominant SB in plastic regime (See inserts of Fig. 8
first, we divide the whole x − y plane into numerous bins by binning (c)). However, the deformation incompatibility of DLGN still appears
analysis and calculate the average coordinate along z axis for atoms in during tension in plastic stage (denoted by dash line in Fig. 11 (a)). As a
each bin in all the GN and MG layers, respectively. Then, the distances result, crack in the GNs initiates at the boundary of the incompatible
of different interface are defined as the differences of these average area (See Fig. 11 (d)), which is in line with the fracture mode of the
coordinates we calculate. Finally, a contour for the distance between preceding MGDLGNC and gives rise to poor ductility compared with its
various interfaces are obtained. Fig. 10 (a)–(c) depict the distance dis- MGSLGNC counterpart with similar graphene volume fraction. Based
tributions of three interfaces, i.e., upper GN-lower GN interface, upper on the above-mentioned analysis, MGSLGNCs show more promising
MG-upper GN interface and lower GN-lower MG interface, in sample performance than MGDLGNCs. Moreover, the deformation incompat-
VIII at the tensile strain of 15.48%. As seen from Fig. 10 (a), the two ibility of DLGN is also demonstrated as interface instability or fluc-
ends of DLGN show much larger interface distance than its remaining tuation in previous studies [32,59]. However, its fracture mechanism
area, pointed by dash lines. In comparison to GN-GN interface, the that the damaged sites of DLGN lie at the boundary of the incompatible
distance distribution between MG-GN interface exhibit a com- area of interface distance, is first proposed and linked with the de-
plementary result (Fig. 10 (b) and (c)). Since GN-GN and MG-GN in- formation incompatibility. These results provide a valuable insight to
terfaces are constrained by Van der Waals force and the original dis- the design of GN-reinforced composites.
tance between DLGN is about 3.4 Å, much less than the maximum
distance values after intense deformation, interface debonding may
4. Conclusions
arise.
To better understand such a distribution of interface distance be-
In summary, we explore the strengthening and toughening effects of
tween DLGN, its three-dimensional schematic illustration and the dis-
SLGN and DLGN with different graphene volume fractions on the uni-
tributions of atomic shear strain in MG matrix and GNs as well as the
axial tensile behavior of MGGNC via MD simulations. The main con-
distorted shape of DLGN are presented in Fig. 10 (d)–(f). According to
clusions are:
the distance value shown in Fig. 10 (d), local regions with larger value
have distinct orientations, in accordance with the directions of the
dominant SBs lying at two ends of the sample, where GNs seriously • Significant dependence of tensile strength and ductility on graphene
volume fraction are found. With an increasing fraction of single-
deform (See arrows EF and GH in Fig. 10 (d) and arrows E ′F ′ and G ′H ′
layered GN in metallic glass composites, both tensile strength and
in Fig. 10 (e)). This reveals that the generation of the dominant SB and
fracture strain are enhanced.
• Rising graphene volume fraction effectively suppresses the forma-
tion of single dominant SB and multiple STZs prevail during the
plastic regime, which leads to a transition in the plastic deformation
mode of MGGNCs. Once the single dominant SB is impeded by GN,
plastic deformation mode transits from shear banding to homo-
geneous flow.
• Based on the comparison between MGSLGNCs and MGDLGNCs with
the similar graphene volume fraction, less ductility, due to the
earlier fracture of GN, appears in MGDLGNCs, which is rationalized
by the incompatibility of DLGN deformation. Thus, single-layered
Fig. 9. Schematic illustration of the measurement of distances between dif- graphene nanoplatelet behaves better than its double-layered
ferent interfaces in MGDLGNC. Only GN layer and the neighboring MG layers counterpart at a given volume fraction.
with the thickness of 2 Å are shown.

6
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 10. Snapshots of sample VIII at the tensile strain of 15.48%. The distance distribution of (a) upper GN-lower GN interface (dash lines imply the boundary
between regions with large value and small value), (b) upper MG-upper GN interface and (c) lower GN-lower MG interface. (d) Three - dimensional view of (a).
Arrows in (d) highlight the regions with large interfacial distance. (e) The distribution of atomic strain in MG matrix and the related configuration of DLGN.
Corresponding to those in (d), arrows in (e) show the orientations of SBs. (f) The distribution of atomic strain in DLGN, where dash lines show the boundary between
regions with large value and small value, corresponding to those in (a).

Our results provide an understanding about the strengthening and Declaration of Competing Interest
toughening mechanisms of GN-reinforced MG nanocomposites at
the nanoscale and offer a promising guideline for strong, ductile MG The authors whose names are listed immediately below certify that
composites. they have NO affiliations with or involvement in any organization or
entity with any financial interest (such as honoraria; educational
grants; participation in speakers’ bureaus; membership, employment,
CRediT authorship contribution statement consultancies, stock ownership, or other equity interest; and expert
testimony or patent-licensing arrangements), or non-financial interest
Zhuocheng Xie: Conceptualization, Methodology, Software, (such as personal or professional relationships, affiliations, knowledge
Writing - original draft. Wu-Rong Jian: Data curation, Formal analysis, or beliefs) in the subject matter or materials discussed in this manu-
Writing - review & editing. Xiaochang Tang: Formal analysis, Writing - script.
review & editing. Xiaoqing Zhang: Supervision, Writing - review &
editing. Xiaohu Yao: Supervision, Writing - review & editing.
Acknowledgments

The Acknowledgements session needs to be revised as: X. Yao would

7
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

Fig. 11. Snapshots of sample IX at the tensile strain of 16.07%. The distance distribution of (a) upper GN-lower GN interface (dash lines imply the boundary between
regions with large value and small value), (b) upper MG-upper GN interface and (c) lower GN-lower MG interface. (d) The distribution of atomic strain in DLGN,
where the dash line shows the boundary between regions with large value and small value, corresponding to that in (a).

like to express his sincere gratitude for the financial support by National intrinsic strength of monolayer graphene, Science 321 (5887) (2008) 385–388.
Science Foundation for Distinguished Young Scholars of China [18] A.K. Geim, Graphene: status and prospects, Science 324 (5934) (2009) 1530–1534.
[19] K.S. C.N.R. Rao A.K. Sood, Graphene: the new two-dimensional nanomaterial,
(11925203) and Natural Science Foundation of China (11672110). Angew Chem. Int. Ed. Engl. 48 (42) (2009) 7752–7777.
[20] W. Jian, X. Long, M. Tang, Y. Cai, X. Yao, S. Luo, Deformation and spallation of
References shock-loaded graphene: effects of orientation and grain boundary, Carbon 132
(2018) 520–528.
[21] D.-B. Xiong, M. Cao, Q. Guo, Z. Tan, G. Fan, Z. Li, et al., Graphene-and-copper
[1] A.L. Greer, Metallic glasses, Science 267 (5206) (1995) 1947–1953. artificial nacre fabricated by a preform impregnation process: bioinspired strategy
[2] W.-H. Wang, C. Dong, C. Shek, Bulk metallic glasses, Mat. Sci. Eng. R 44 (2–3) for strengthening-toughening of metal matrix composite, Acs Nano. 9 (7) (2015)
(2004) 45–89. 6934–6943.
[3] C.A. Schuh, T.C. Hufnagel, U. Ramamurty, Mechanical behavior of amorphous al- [22] M. Cao, D.-B. Xiong, Z. Tan, G. Ji, B. Amin-Ahmadi, Q. Guo, et al., Aligning gra-
loys, Acta Mater. 55 (12) (2007) 4067–4109. phene in bulk copper: nacre-inspired nanolaminated architecture coupled with in-
[4] P. Murali, T. Guo, Y. Zhang, R. Narasimhan, Y. Li, H. Gao, Atomic scale fluctuations situ processing for enhanced mechanical properties and high electrical con-
govern brittle fracture and cavitation behavior in metallic glasses, Phys. Rev. Lett. ductivity, Carbon 117 (2017) 65–74.
107 (21) (2011) 215501. [23] A. Nieto, A. Bisht, D. Lahiri, C. Zhang, A. Agarwal, Graphene reinforced metal and
[5] M. Jiang, L. Dai, On the origin of shear banding instability in metallic glasses, J. ceramic matrix composites: a review, Int. Mater. Rev. 62 (5) (2017) 241–302.
Mech. Phys. Solids 57 (8) (2009) 1267–1292. [24] J. Hwang, T. Yoon, S.H. Jin, J. Lee, T.-S. Kim, S.H. Hong, S. Jeon, Enhanced me-
[6] M.M. Trexler, N.N. Thadhani, Mechanical properties of bulk metallic glasses, Prog. chanical properties of graphene/copper nanocomposites using a molecular-level
Mater. Sci. 55 (8) (2010) 759–839. mixing process, Adv. Mater. 25 (46) (2013) 6724–6729.
[7] A. Greer, Y. Cheng, E. Ma, Shear bands in metallic glasses, Mat. Sci. Eng. R 74 (4) [25] Y. Kim, J. Lee, M.S. Yeom, J.W. Shin, H. Kim, Y. Cui, J.W. Kysar, J. Hone, Y. Jung,
(2013) 71–132. S. Jeon, et al., Strengthening effect of single-atomic-layer graphene in metal–-
[8] H. Guo, P. Yan, Y. Wang, J. Tan, Z. Zhang, M. Sui, E. Ma, Tensile ductility and graphene nanolayered composites, Nat. Commun. 4 (2013) 2114.
necking of metallic glass, Nat. Mater. 6 (10) (2007) 735. [26] S. Feng, Q. Guo, Z. Li, G. Fan, Z. Li, D.-B. Xiong, Y. Su, Z. Tan, J. Zhang, D. Zhang,
[9] D. Jang, J.R. Greer, Transition from a strong-yet-brittle to a stronger-and-ductile Strengthening and toughening mechanisms in graphene-al nanolaminated compo-
state by size reduction of metallic glasses, Nat. Mater. 9 (3) (2010) 215. site micro-pillars, Acta Mater. 125 (2017) 98–108.
[10] D. Jang, C.T. Gross, J.R. Greer, Effects of size on the strength and deformation [27] L. Zhao, Q. Guo, Z. Li, Z. Li, G. Fan, D.-B. Xiong, Y. Su, J. Zhang, Z. Tan, D. Zhang,
mechanism in zr-based metallic glasses, Int. J. Plasticity 27 (6) (2011) 858–867. Strain-rate dependent deformation mechanism of graphene-al nanolaminated
[11] Y. Wang, J. Li, A.V. Hamza, T.W. Barbee, Ductile crystalline–amorphous nanola- composites studied using micro-pillar compression, Int. J. Plasticity 105 (2018)
minates, Proc. Natl. Acad. Sci. 104 (27) (2007) 11155–11160. 128–140.
[12] W. Jian, L. Wang, B. Li, X. Yao, S. Luo, Improved ductility of cu64zr36 metallic [28] L. Zhao, Q. Guo, Z. Li, D.-B. Xiong, S. Osovski, Y. Su, D. Zhang, Strengthening and
glass/cu nanocomposites via phase and grain boundaries, Nanotechnology 27 (17) deformation mechanisms in nanolaminated graphene-al composite micro-pillars
(2016) 175701. affected by graphene in-plane sizes, Int. J. Plasticity 116 (2019) 265–279.
[13] D.C. Hofmann, J.-Y. Suh, A. Wiest, G. Duan, M.-L. Lind, M.D. Demetriou, [29] B. Hwang, W. Kim, J. Kim, S. Lee, S. Lim, S. Kim, S.H. Oh, S. Ryu, S.M. Han, Role of
W.L. Johnson, Designing metallic glass matrix composites with high toughness and graphene in reducing fatigue damage in cu/gr nanolayered composite, Nano Lett.
tensile ductility, Nature 451 (7182) (2008) 1085–1089. 17 (8) (2017) 4740–4745.
[14] X. Zhou, C. Chen, Strengthening and toughening mechanisms of amorphous/ [30] X. Liu, F. Wang, W. Wang, H. Wu, Interfacial strengthening and self-healing effect in
amorphous nanolaminates, Int. J. Plasticity 80 (2016) 75–85. graphene-copper nanolayered composites under shear deformation, Carbon 107
[15] W. Jian, L. Wang, X. Yao, S. Luo, Tensile and nanoindentation deformation of (2016) 680–688.
amorphous/crystalline nanolaminates: effects of layer thickness and interface type, [31] S. Zhang, Y. Xu, X. Liu, S.-N. Luo, Competing roles of interfaces and matrix grain
Comput. Mater. Sci. 154 (2018) 225–233. size in the deformation and failure of polycrystalline cu–graphene nanolayered
[16] W. Jian, L. Wang, X. Yao, S. Luo, Balancing strength, hardness and ductility of composites under shear loading, Phys. Chem. Chem. Phys. 20 (36) (2018)
cu64zr36 nanoglasses via embedded nanocrystals, Nanotechnology 29 (2) (2017) 23694–23701.
025701. [32] X. Liu, J. Cai, S.-N. Luo, Interfacial anti-fatigue effect in graphene–copper nano-
[17] C. Lee, X. Wei, J.W. Kysar, J. Hone, Measurement of the elastic properties and layered composites under cyclic shear loading, Phys. Chem. Chem. Phys. 20 (11)

8
Z. Xie, et al. Journal of Non-Crystalline Solids 546 (2020) 120284

(2018) 7875–7884. [45] A. Cao, Y. Cheng, E. Ma, Structural processes that initiate shear localization in
[33] S.-Y. Park, E.-J. Gwak, M. Huang, R.S. Ruoff, J.-Y. Kim, Nanolaminate of metallic metallic glass, Acta Mater. 57 (17) (2009) 5146–5155.
glass and graphene with enhanced elastic modulus, strength, and ductility in ten- [46] X. Tang, L. Meng, J. Zhan, W. Jian, W. Li, X. Yao, et al., Strengthening effects of
sion, Scr. Mater. 139 (2017) 63–66. encapsulating graphene in sic particle-reinforced al-matrix composites, Comput.
[34] R. Ansari, B. Motevalli, A. Montazeri, S. Ajori, Fracture analysis of monolayer Mater. Sci. 153 (2018) 275–281.
graphene sheets with double vacancy defects via md simulation, Solid State [47] Y. Cheng, A.J. Cao, H. Sheng, E. Ma, Local order influences initiation of plastic flow
Commun. 151 (17) (2011) 1141–1146. in metallic glass: effects of alloy composition and sample cooling history, Acta
[35] R. Ansari, S. Ajori, B. Motevalli, Mechanical properties of defective single-layered Mater. 56 (18) (2008) 5263–5275.
graphene sheets via molecular dynamics simulation, Superlattices Microstruct. 51 [48] Z. Sha, W.H. Wong, Q. Pei, P.S. Branicio, Z. Liu, T. Wang, T. Guo, H. Gao, Atomistic
(2) (2012) 274–289. origin of size effects in fatigue behavior of metallic glasses, J. Mech. Phys. Solids
[36] Z. Ni, H. Bu, M. Zou, H. Yi, K. Bi, Y. Chen, Anisotropic mechanical properties of 104 (2017) 84–95.
graphene sheets from molecular dynamics, Physica B Condens. Matter 405 (5) [49] Z. Sha, S. Qu, Z. Liu, T. Wang, H. Gao, Cyclic deformation in metallic glasses, Nano
(2010) 1301–1306. Lett. 15 (10) (2015) 7010–7015.
[37] S. Sharma, P. Kumar, R. Chandra, Mechanical and thermal properties of graphe- [50] F. Ma, Y. Sun, D. Ma, K. Xu, P.K. Chu, Reversible phase transformation in graphene
ne–carbon nanotube-reinforced metal matrix composites: a molecular dynamics nano-ribbons: lattice shearing based mechanism, Acta Mater. 59 (17) (2011)
study, J. Compos. Mater. 51 (23) (2017) 3299–3313. 6783–6789.
[38] R. Rezaei, C. Deng, M. Shariati, H. Tavakoli-Anbaran, The ductility and toughness [51] B. Mortazavi, Y. Rémond, S. Ahzi, V. Toniazzo, Thickness and chirality effects on
improvement in metallic glass through the dual effects of graphene interface, J. tensile behavior of few-layer graphene by molecular dynamics simulations,
Mater. Res. 32 (2) (2017) 392–403. Comput. Mater. Sci. 53 (1) (2012) 298–302.
[39] R. Ansari, S. Malakpour, S. Ajori, Structural and elastic properties of hybrid bilayer [52] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J.
graphene/h-bn with different interlayer distances using dft, Superlattices Comput. Phys. 117 (1) (1995) 1–19.
Microstruct. 72 (2014) 230–237. [53] A. Stukowski, Visualization and analysis of atomistic simulation data with ovito–the
[40] S. Haghighi, R. Ansari, S. Ajori, A molecular dynamics study on the interfacial open visualization tool, Model Simul. Mater. SC. 18 (1) (2009) 015012.
properties of carbene-functionalized graphene/polymer nanocomposites, Int. J. [54] M. Mendelev, D. Sordelet, M. Kramer, Using atomistic computer simulations to
Mech. Mater. Des. (2019) 1–14. analyze x-ray diffraction data from metallic glasses, J. Appl. Phys. 102 (4) (2007)
[41] S. Ajori, H. Parsapour, R. Ansari, A. Ameri, Buckling behavior of various metallic 043501.
glass nanocomposites reinforced by carbon nanotube and cu nanowire: a molecular [55] R. Rezaei, M. Shariati, H. Tavakoli-Anbaran, C. Deng, Mechanical characteristics of
dynamics simulation study, Mater. Res. Express. 6 (9) (2019) 095070. cnt-reinforced metallic glass nanocomposites by molecular dynamics simulations,
[42] S. Ajori, H. Parsapour, R. Ansari, A comprehensive analysis of the mechanical Comput. Mater. Sci. 119 (2016) 19–26.
properties and fracture analysis of metallic glass nanocomposites reinforced by [56] V. Filippova, S. Kunavin, M. Pugachev, Calculation of the parameters of the len-
carbon nanotubes and cu nanowires: a molecular dynamics study, Mech. Adv. nard-jones potential for pairs of identical atoms based on the properties of solid
Mater. Struct. (2020) 1–20. substances, Inorg. Mater. Appl. Res. 6 (1) (2015) 1–4.
[43] Z. Xie, W.-R. Jian, Z. Wang, X. Zhang, X. Yao, Layer thickness effects on the [57] S. Zhen, G. Davies, Calculation of the lennard-jones n–m potential energy para-
strengthening and toughening mechanisms in metallic glass-graphene nanolami- meters for metals, Phys. Status Solidi. 78 (2) (1983) 595–605.
nates, Comput. Mater. Sci. 177 (2020) 109536. [58] F. Shimizu, S. Ogata, J. Li, Theory of shear banding in metallic glasses and mole-
[44] W. Jian, X. Yao, L. Wang, X. Tang, S. Luo, Short-and medium-range orders in cular dynamics calculations, Mater. trans. (2007). 0710160231–0710160231
cu46zr54 metallic glasses under shock compression, J. Appl. Phys. 118 (1) (2015) [59] C. Wang, Y. Liu, L. Lan, H. Tan, Graphene wrinkling: formation, evolution and
015901. collapse, Nanoscale 5 (10) (2013) 4454–4461.

You might also like