Sie sind auf Seite 1von 27

Journal Pre-proof

Geometrical aspects of nanofillers influence the tribological performance of Al-based


nanocomposites

Soroosh Mohammadi, Abbas Montazeri, Herbert M. Urbassek

PII: S0043-1648(19)31255-4
DOI: https://doi.org/10.1016/j.wear.2019.203117
Reference: WEA 203117

To appear in: Wear

Received Date: 15 August 2019


Revised Date: 16 October 2019
Accepted Date: 6 November 2019

Please cite this article as: S. Mohammadi, A. Montazeri, H.M. Urbassek, Geometrical aspects of
nanofillers influence the tribological performance of Al-based nanocomposites, Wear (2019), doi: https://
doi.org/10.1016/j.wear.2019.203117.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier B.V.


Geometrical aspects of nanofillers influence the tribological
performance of Al-based nanocomposites

c,d,*
Soroosh Mohammadia, Abbas Montazeria,b, Herbert M. Urbassek
a
Computational Nanomaterials Lab (CNL), Faculty of Materials Science and Engineering, K. N. Toosi
University of Technology, Tehran, Iran
b
School of Nano Science, Institute for Research in Fundamental Sciences (IPM), P.O. Box 19395-5531, Tehran,
Iran
c
Physics Department, University Kaiserslautern, Erwin-Schrödinger-Straße, D-67663 Kaiserslautern, Germany
d
Research Center OPTIMAS, University Kaiserslautern, Erwin-Schrödinger-Straße, D-67663 Kaiserslautern,
Germany

Abstract
With respect to the inherent poor tribological behavior of aluminum (Al), reinforcement of this metal
with nanoscale fillers is a promising research area with significant industrial impact. Accordingly, in
the current research, we have probed the tribological characteristics of Al-based nanocomposites
embedded with nanofillers of different geometries. Targeting this purpose, a series of nanoscratching
tests based on molecular-dynamics simulations is carried out to explore the effect of spherical silicon
carbide nanoparticles and graphene monolayers on the coefficient of friction (COF) of the baseline Al
matrix. The results reveal that while the former inclusion reduces this parameter, COF is enhanced in
the presence of platelet graphene sheets. To find out the underlying mechanism, the microstructural
evolutions are thoroughly examined in the interfacial area by means of the dislocation extraction
analysis. Our findings are supported by an in-depth study on the pile-up features and more
pronounced plastic deformation zones in the samples. Furthermore, to overcome the deteriorating
influence of graphene on the COF of the Al matrix, various composite systems are designed which
differ in the depth where the nanofiller was positioned. Finally, a solution is presented to increase the
mechanical performance of the Al-based nanocomposites in the presence of graphene sheets without
sacrificing their tribological behavior.

Keywords:
Nanoscratch; Molecular dynamics; Al-based nanocomposite; Nanofiller geometry; Interfacial area

* Corresponding author at: Physics Department, University Kaiserslautern, Erwin-Schrödinger-Straße, D-67663


Kaiserslautern, Germany.
E-mail address: urbassek@rhrk.uni-kl.de (H.M. Urbassek).

1
1. Introduction

Considering the recent advances in the synthesis of metal-matrix nanocomposites (MMNCs)


[1-4], their tribological characteristics are a promising research area that is just beginning to
be explored. It should be pointed out that in nanoscale systems, surface-dependent
phenomena such as friction, wear, and adhesion are being intensified [5]. Accordingly,
tribological properties of nanodevices and their underlying mechanisms should be thoroughly
analyzed [6,7]. The nanoscratching test is one of the main approaches utilized for this
purpose, in which a hard tip is indented into the surface and then moved parallel to it. From
the lateral and normal mechanical response of the substrate, its tribological characteristics can
be obtained [8,9]. Since this procedure involves the deformation of only a small number of
atoms or atomic layers, there is a fundamental difference between atomic-scale friction and
what happens in macro- and even micro-scale experiments. Hence, it cannot be examined by
means of the conventional approaches in continuum mechanics. Additionally, experimental-
based techniques are unable to correctly probe the nanoscale friction due to their limitations.
Among them, we can enumerate the ones related to control and measurement of the atomic-
scale indenter movement and preparation the substrate. Furthermore, because of altered
orientation of the atomic planes in the different experimental setups, assessing the
reproducibility of the results is a hard process. Therefore, using numerical-based tools such as
molecular dynamics (MD) simulation would provide a deeper insight into the nanoscale
phenomena occurring during the mentioned tests [10,11]. This method has been successfully
employed to analyze various tribological features of metals and MMNCs including hardness
[12], adhesion [13], friction [14], wear [15], and machinability [16] at the atomic-scale.
Among conventional metals, aluminum (Al) presents a combination of thermal, mechanical,
electrical, and corrosion resistance, which have made it relevant for implementation in many
industrial applications [17,18]. However, its potential applications are limited by the poor
tribological properties ascribed to the inherent softness of this material. For example,
employing MD simulations, it has been well established that compared to other metals, Al
has higher friction coefficient [19]. Using this approach, similar findings have also been
reported by Junge and Molinari for mono- and polycrystalline Al substrates [20,21]. Hence,
to overcome this issue, addition of effective nanofillers within the Al matrix is of particular
importance and needs to be thoroughly analyzed [22]. It has been demonstrated that in the
presence of nanofillers having different geometries such as spherical SiC nanoparticles (NPs)
and graphene platelets, the abovementioned mechanical properties of the aluminum matrix

2
are substantially improved [23,24]. Accordingly, in the last decade, Al-based NCs have been
considered more than pure aluminum attributed to their higher strength and specific modulus,
better corrosion resistance, enhanced tribological features, and high-temperature mechanical
properties [25,26].
In this context, Reddy et al. [27] incorporated a small volume fraction (VF) of SiC NPs to
reinforce the Al matrix. The results showed that substantial improvement in the compressive
and tensile strength of aluminum was observed by adding only 1.5% VF of these nanofillers.
However, in the case of micro-scale fillers, more than 10% VF of this reinforcing agent was
required to achieve the same enhancement [27]. Similarly, Jiang et al. [28] reported a
noticeable increase in the ultimate tensile strength (about 18%) of aluminum by addition of
only 0.5%VF of graphene into the base material. In addition to the experimental observations,
MD simulations have also been performed on the aluminum-based NCs to explore the
influence of the mentioned nanofillers on the Al reinforcement. Applying this approach,
Tavakoli et al. [29] studied the effect of shock wave sintering on the mechanical properties of
SiC/aluminum nanocomposites [29]. Gu et al. [30] noticed that MD simulations could
elucidate the mechanical behavior of SiC/Al NCs better than the continuum-based finite
element approach and its results were closer to the experimental one. Utilizing this technique,
a considerable amount of literature has been devoted to explore the dependence of aluminum
mechanical response on the SiC morphology [31], Al/SiC interface [32], and the NP size
[33]. Also, more recently, Zhang and Urbassek [34] have addressed the effect of Si
geometrical shape on the strengthening mechanism of these nanocomposites via MD
simulations of nanoindentation. Moreover, using molecular dynamics simulations, the effect
of graphene sheets on the elastic constants of aluminum matrices has also been investigated
[35]. The results showed a considerable improvement in the mechanical properties of the base
matrix including the yield and ultimate tensile strength.
While the majority of the current literature on the numerical modeling aspects of the Al-based
NCs pays particular attention to their mechanical behavior under axial loading conditions,
their tribological features have been rarely investigated. Considering the soft nature of Al
[36], a noticeable gap in this research area has been the lack of a computational model to
explore its tribological properties such as the friction and wear behavior in the presence of
different nanoscale additives. An important preliminary question that should arise at this
point is the following: How do the geometrical characteristics of nanofillers affect the
coefficient of friction (COF) of the Al matrix? This paper aims to pursue this query using MD
simulations of a nanoscratching test. Correspondingly, at first, the friction coefficient of Al-
3
reinforced NCs embedded with graphene platelets and spherical silicon carbide NPs are
systematically examined and compared with that of the pristine aluminum. Due to the
different effect of these nanofillers on the results, we further analyze the interfacial area of
Al/graphene and Al/SiC samples in order to explore the governing mechanisms. This is
achieved through identification of the dislocation creation and movement in the samples
provided by the dislocation extraction algorithm (DXA). The results are supported by a
thorough analysis of the pile-up formation in these two case studies. Finally, a systematic
study is conducted to reveal how the aluminum COF is influenced by the depth, at which
graphene is positioned within the Al matrix. In the next sections, at first, we present the
simulation details associated with the nanoscratching test. Then, the numerical results and the
corresponding discussions are presented in Section 3 highlighting the role of interfacial
phenomena in the underlying mechanisms. Section 4 summarizes the most important findings
of the paper.

2. Details of MD simulation
In the present study, MD simulation was utilized to explore the role of the nanofiller
geometry on the wear properties of aluminum-matrix composites. In the following, the
nanoscratch procedure, the sample preparation method and the MD parameters are briefly
outlined. All simulations were performed using the open-source LAMMPS code [37]. The
samples have dimensions of 200 × 225 × 200 Ȧ and were constructed in each case to
resemble the pure aluminum matrix and its composite counterparts (See Fig. 1). The
geometrical characteristics of the spherical indenter and its location are also depicted in this
figure. It is noted that the basic computational cells were created in two steps. Fist, the
pristine aluminum sample was created using the built-in tools in LAMMPS with the specific
Al lattice parameter and its fcc crystallographic structure. Then a hole was introduced to
accommodate the nanofillers of either SiC NPs or a graphene platelet. To prevent any sample
movement during the test, the bottommost layers with the thickness of 15 Ȧ were kept fixed.
The next two layers were thermostatted to dissipate the generated heat away from the sample.
The remaining atoms, called Newtonian atoms in Fig. 1, were simulated in the canonical
ensemble (NVT) at room temperature (300 K) to prevent the activation of thermally-induced
mechanisms occurred at the higher temperatures. This was achieved through utilizing the
Nose-Hoover thermostat, which demonstrates fewer fluctuations compared to other
procedures [38]. In addition, periodic boundary conditions were applied to the front and back
sides to reduce the simulation box size, while the top surface was assumed to be free.
4
Fig. 1. 3D model of the substrate and indenter.

The initial velocities were sampled from the Maxwell–Boltzmann distribution at the desired
temperature. Then, setting the time-step equal to 1fs, the velocity-Verlet algorithm was
utilized to integrate the equations of motion [39]. It is of worthy to mention that during the
scratch, the produced dislocation loops could possibly propagate to long distances.
Accordingly, the computational cell was chosen sufficiently large to prevent the interaction
of dislocations with the boundaries of the simulation box [19,40]. The OVITO software was
employed for visualization and image analysis [41]. Moreover, dislocations were identified
using the DXA approach developed by Stukowski and Albe [42]. All scratch simulations
were performed on the Al (001) single crystal surface along the [010] direction, which helped
us to mainly analyze the role of nanofiller geometry on the tribological characteristics of the
Al matrix under identical conditions. This is the common scratch system treated in the
literature to analyze the tribological characteristics of FCC metals such as aluminum [17,43-
45].
To mimic the interactions within the Al/SiC sample, three potential functions were utilized.
In this regard, Mendelev [46] and Tersoff [47] potentials were employed to describe the
energetic bonds of aluminum matrix and SiC nanoparticles, respectively. Additionally, the
well-known Morse potential was used for interactions between the constituents of this
nanofiller and Al matrix [29,33]. The corresponding potential parameters are listed in Table
1. For the second case, i.e. Al/graphene nanocomposite sample, the AIREBO potential was
employed to simulate the covalent C-C bonding within the graphene sheet [48]. Moreover,
motivated by the previous researches [49-52], the Lennard-Jones (LJ) potential was adopted
to consider the van-der-Waals nature of the atomic interactions between the Al matrix and the

5
graphene nanofillers. Accordingly, here, the LJ values for the Al-graphene interface were
= 0.003457 eV and σ = 3.132 Ȧ based on the parameters proposed by Rong et al. [35].

Table 1. Parameters of the Morse potential for the Al/SiC system [29].

1
Cases ( ) ( ) ( °)
°

Al-C 0.4691 1.738 2.246


Al-Si 0.4824 1.322 2.920

Fig. 2. Geometrical characteristics of the Al/graphene and Al/SiC


nanocomposite samples from a top (a,c) and a side (b,d) view.

Fig. 2 shows the geometrical characteristics of the introduced nanocomposites in both a top
and a side view. Both nanofillers were positioned 50 Ȧ beneath the free surface. We will
further demonstrate that this parameter can noticeably affect the tribological behavior of the
analyzed samples. The interface between Al and the nanofiller was carefully relaxed at 300 K
for 80 ps. We obtained equilibrium distances of 3.14 Å and 3.0 Å for the Al/graphene and
Al/SiC samples, respectively. This corresponds well to the corresponding parameters for the
6
Al-C and Al-Si interactions that were reported experimentally and numerically in the range of
2.3 − 3.36 Ȧ [52] and 2.44 − 2.92 Ȧ [53], respectively.
After construction of the samples, each system was initially relaxed for 40 ps at 300 K before
being exposed to the scratching test. The scratch process was performed in an incremental
fashion, in which after each step of the indenter movement, the system is stopped to
equilibrate. This incremental-based MD procedure has led to better results in terms of
controlling the simulation temperature [54]. With respect to the indenter velocity, it should be
noted that in the conventional MD-based simulation of nanoscratching and nanoindentation,
this parameter has been reported in the broad interval of 1-200 ⁄ [55]. Accordingly, here,
the velocity of 20 ⁄ was considered to better examine the tribological behavior of the
samples in a reasonable computational cost. Similar to the previous works, a rigid spherical-
shaped indenter with the radius of R=50 Ȧ was implemented. Although this size is smaller
than that of the conventional indenters utilized in the experimental studies, currently,
production of indenters having a radius smaller than 10 nm has been achieved. The force
exerted on each substrate atom is computed through the following relation [19]:

"(#) = $−%(# − &) if r < R


'
(1)
0 if r > R
In this equation, R is the radius of the spherical indenter, r denotes the distance between the
substrate atom to the indenter center, and % is the force constant selected as 10 :;/ => , which
is in the range of 10 − 11.3 :;/ => reported in the previous nanoscratching works [12, 34]. As
depicted in Fig. 1, during the test, the scratch depth was chosen as 30 Ȧ. It is noteworthy to
mention that to report more accurate data in each case, individual tests have been repeated
four times and the average results have been presented here.

3. Results and Discussion

3.1. Nanoscratch of pure Al: Validation of the developed code

To validate the developed scratch code in LAMMPS, first, a pure aluminum sample with the
geometrical characteristics shown in Fig. 1 was analyzed under the designed nanoscratching
test. Fig. 3a shows the atomic configuration of the sample during the process, in which the
spherical indenter experiences the frictional (tangential) and normal forces. These forces are
denoted by "? and "@ , respectively, and are schematically shown in Fig. 3b. While the former
corresponds to the force required to move the indenter in the longitudinal y-direction, the
latter is the force necessary for maintaining the indenter at the scratch depth.

7
Fig. 4 shows the corresponding force-displacement curves of the sample. In the initial 40 Ȧ of
the indenter movement, the discussed force components increase monotonically after the
beginning of the process demonstrating the elastic deformation within the sample. Passing the
linear trend observed in the curves, dislocations were formed underneath and in front of the
indenter resulting in the formation of a plastic region. Consequently, the nucleation of
dislocations and their interactions would produce some fluctuations in the graphs.
Average values of the frictional and normal forces (averaged over the last 60 Å of scratching)
are utilized to determine the COF of the pure Al via the well-known Coulomb model of
friction:
BC
μ= (2)
BD

(a) (b)

Fig. 3. a) Atomic configuration of pure Al during the scratch test and


b) Schematic representation of the frictional and normal forces during the test.

Fig. 4. Variations of the frictional and normal forces during the scratch of the pure Al sample.

8
It is worthy of mention that although the scratch length was 150 Ȧ in the current study, more
accurate results would be achieved at the last 60 Ȧ of the curves as can be found from Fig. 4.
Considering this assumption, the average frictional and normal forces were approximately
found to be 0.110 and 0.179 EF, respectively. Substituting these values into Eq. (2), the
friction coefficient of the pristine sample was obtained as 0.61, which is in close agreement
with the available simulation results [13,19] verifying the developed code.

3.2. Analysis of the nanocomposite samples: The role of nanofiller geometry

After examining the pristine aluminum, the next step involved analysis of the effect of
nanofiller geometry on the wear characteristics of the Al matrix. In this regard, using the
MD-based procedure defined in the previous section, the friction coefficient of Al-based NCs
embedded with graphene platelets and silicon carbide NPs were examined. We focused on
highlighting the role of interfacial interactions on the results. Accordingly, DXA analysis was
utilized to identify the dislocation creation and movement in this area. This was followed by
inspection of the underlying mechanisms as discussed in detail in the rest of this section. Fig.
5 shows the variation of the frictional and normal forces for the introduced nanocomposite
samples, from which the friction coefficient of the Al/graphene and Al/SiC samples were
calculated to be 0.68 and 0.55, respectively. At this point, we would like to draw the reader’s
attention to the different effects of these nanofillers on the results. As seen, addition of SiC
NPs would lead to a 9% reduction of the COF parameter, while this parameter shows an 11%
enhancement for Al matrix embedded with a graphene platelet. To provide a route for more
investigation of this difference, Fig. 6 shows the average values of the frictional and normal
forces for the NC samples along with those of the pure aluminum sample. It is noted that to
remove the effect of perturbations on the results, they have been averaged over the last 60 Ȧ
of the scratch course. As seen, in the case of the Al/graphene sample, a simultaneous increase
of the frictional force and a reduction in its normal counterpart led to the observed increase in
the COF. In contrast, in the Al/SiC composite sample, the normal force has increased while
the frictional force stayed constant. Accordingly, a reduction in the COF was observed. Such
a behavior might be attributed to the occurrence of work hardening at the different zones of
these samples [56]. Based on the presented results, it would be anticipated that for Al/SiC and
Al/graphene nanocomposites, this phenomenon has occurred underneath and in front of the
indenter, respectively. Employing the dislocation-based analysis tool, this issue is inspected
in the next section.

9
(a)

(b)
Fig. 5. Variations of a) Frictional and b) Normal forces for the Al/SiC and Al/graphene samples
compared to those of pure Al.

Fig. 6. Average values of the frictional and normal forces of the NC samples in comparison with the pure Al.

10
3.3. Towards the interface-dependent underlying mechanism

To explore the mechanism supporting the different behavior of the studied NC samples, we
further proceeded to explore their interfacial features provided by the OVITO software. For
single-crystalline materials embedded with inclusions, it has been demonstrated that when the
lattice mismatch exceeds more than 5%, a semi-coherent interface is expected to form
between them [57]. This was the case occurring in the Al/SiC sample, in which this
parameter amounts to about 7%. Accordingly, as depicted in Fig. 7, the interfacial energy is
increased producing local stress concentration around the SiC nanoparticles, which in turn
promotes the nucleation of dislocations. This leads to disorder of the Al matrix extending
several atomic layers in the vicinity of the nanoparticles as reported in the literature [24,26].
In contrast, for the Al/graphene sample, owing to the formation of a coherent-like interface,
the stress distribution is only limited to the surrounding layer of the embedded graphene
sheet.

Fig. 7. Von-Mises stress distribution in the nanocomposite samples: (a) Al/SiC and (b) Al/graphene.

To provide a complete picture of how the results can be influenced by the semi-coherent
interface in the Al/SiC sample, we take a closer look at the dislocation creation and emission
during the nanoscratching process of this sample. For this purpose, the DXA algorithm was
employed to extract the dislocation type and their movements. As mentioned before, due to
the stress concentration in the interface zone, new dislocations were created around the
nanoparticle surface. From Fig. 8, we may deduce that SiC NPs act as obstacles for the
dislocations movement. Cross-slip and obstacle cutting are the most common mechanisms,
through which dislocations could pass an obstacle [58]. However, owing to the existence of
the strong stress field around the inclusions, providing the activation energy needed to cross-

11
slip would not be possible. Additionally, the obstacle-cutting mechanism has rarely been
observed for hard barriers such as SiC particles. Therefore, as seen in this figure, dislocations
accumulate near the interface under the indenter tip. This is followed by the creation of new
dislocations produced by the enduring loading imposed during the test. Subsequently, some
reactions occur between the new and the immobilized dislocations at the Al/SiC interface
leading to the creation of locked dislocations (e.g., stair-rod) as stated in Eq. (3) [59]:
G G G
H
I2 1 1J + H I121J = H [1 1 0] (3)
Accordingly, the interface would be transformed into a zone containing a high dislocation
density. This phenomenon caused this area to be strengthened – denoted as Orowan
strengthening [33] – that led to increase of the normal forces and as a result, reduction of the
friction coefficient in this case study.

Fig. 8. Nucleation, movement and interaction of dislocations in the Al/SiC nanocomposite sample: a) Nucleation
of a dislocation loop beneath the indenter, b-e) Emission of the nucleated loop in its glide system, f) SiC NPs act
as obstacles preventing the dislocation movement, g) Nucleation of new dislocations under the indenter, and h)
Generation of locked dislocations caused by the interaction between new dislocations and the old immobilized
ones.

12
However, as previously discussed, a reverse trend was observed for the aluminum COF in the
presence of graphene sheets. To reveal the governing mechanism, one can compare the role
of the nanofiller geometry on the dislocation movement in the interfacial area. To this aim,
the DXA analysis was implemented to identify the type and extent of dislocations for the
sample reinforced with a 2D graphene platelet. As illustrated in Fig. 9, the nucleation of
dislocations beneath the indenter forms a dislocation loop moving along the graphene
surface. Contrary to the former case, no significant perturbation happens during the loop
movement; this behavior may be attributed to the preserved atomic arrangement of the Al
matrix in the coherent-like interface region. A similar behavior has also been observed
experimentally by Jiang et al. [28] for graphene-nanosheet/aluminum composites. This
smooth movement of the dislocations on the graphene surface would lead to an increase of
the dislocation density ahead of the indenter tip causing more plastic deformation in this area
and to a concomitant rise of the frictional force required for moving forward the indenter: As
Fig. 6 shows, its value rises from 0.110 µN for pure Al to 0.118 µN for the Al/graphene
nanocomposite.

Fig. 9. Interaction between dislocations and graphene in the Al/graphene nanocomposite sample: a) Formation
of a dislocation loop under the indenter, b) Interaction of the dislocation loop and graphene, and c) Smooth
movement of the dislocation in the interfacial area.

Considering Eq. (2), it is also important to analyze the change in the normal force component
in the presence of graphene sheet to better examine the COF of the pure Al embedded with
this 2D nanostructure. As graphically shown in Fig. 6, the corresponding results were found

13
to be 0.179 and 0.172 μN for the case of pure Al and Al/graphene samples, respectively. This
softening behavior has been previously addressed for Ni-based NCs reinforced with graphene
sheets oriented perpendicular to the loading direction [12,60]. In this regard, Montazeri and
Mobarghei [12] demonstrated that depending on the alignment of the graphene layer within
the host nickel, the NC could become softer than the base metal. This was attributed to the
very low out-of-plane stiffness of graphene when subjected to the vertical forces. To find out
the underlying mechanism, the interaction of the generated dislocations with the graphene
interface was recently addressed by Vardanyan and Urbassek [60]. For this purpose, a series
of MD simulation was carried out to systematically analyze the role of the graphene position
below the nickel substrate. It was shown that placing the graphene flake near the metal
surface would lead to a weakening of the composite system as observed in the present study.
The induced height depression of the graphene due to the dislocation absorption (Fig. 5 of
Ref. [60]) and the creation of a step structure in the Ni/graphene interface (Fig. 6a of Ref.
[60]) were enumerated as the main reasons for the softening behavior of the discussed metal-
based composite systems. The latter has also been observed for the Al/graphene sample
analyzed in the present work. It is noted that with increasing the applied normal force,
Shockley partial dislocations are created within the sample. Subsequently, as seen in Fig. 10,
they move towards the interface and as a result, stacking fault regions appear between these
dislocations leading to the formation of a surface step in the top Al interface layer. This
phenomenon would manifest itself by bending the graphene flake slightly below this step,
which leads to a drop in the force acting on the indenter. For a more comprehensive coverage
of the underlying mechanisms, the reader is referred to Ref. [60]. Therefore, the presence of a
graphene sheet near the Al surface results in the simultaneous increase of the frictional force
and the reduction of the normal force; both these effects lead to an increase of the COF of the
Al/graphene sample compared to that of the pure aluminum.

3.4. Pile-up investigation: More in-depth study on the proposed mechanism

Considering the discussion in the previous section, depending on the nanofiller utilized to
reinforce the matrix, Al-based nanocomposites could have different tribological behavior
compared to the base aluminum matrix. The proposed mechanisms highlighted the role of the
nanofiller geometry on the nucleation of dislocations in the interfacial area and the control of
their movements. To provide a qualitative tool supporting this idea, we analyzed the plastic
deformation zones within the studied samples. Fig. 11 shows frontal and lateral views of the

14
deformed nanocomposites after the scratch test. The results reveal that, as expected, the
presence of a graphene platelet caused the plastic deformation to occur more intensely in
front of the indenter. This happens because the dislocation movement was facilitated in the
coherent-like interfacial zone (Figs 7e and 7f).

Fig. 10. Step generation by stacking-fault movement: a) Perfect interface, b) Appearance of the stacking fault
region between Shockley partial dislocations, and c) Formation of a surface step in the interface.

Fig. 11. Frontal (a, c, e) and lateral (b, d, f) representations of the plastic deformation occurring in the samples
after the scratching test: Left: Al/SiC, Middle: Pure Aluminum, Right: Al/graphene.

15
To shed more light on this issue, the pile-up generated upon scratching of the introduced
samples was also studied as depicted in Fig. 12. It has been well established that the presence
of the slip systems in front of the indenter results in the materials pile-up in this area, while
the activation of the lateral glide systems leads to the material accumulation on the indenter
and groove sides [19]. During the scratching process, all slip systems are naturally activated
causing the dislocations to appear in all slip systems [61]. The front views of the pile-up
presented in this figure demonstrate a considerable variety of shapes reflected by the different
interfacial characteristics of the aformentioned samples. Additionally, to facilitate a better
comparison, top views of the scratch groove and the surrounding pile-up have also been
illustrated. As seen, in the case of the Al/graphene sample, material accumulation against the
indenter is larger than for the other samples and increase the force required for the indenter
movement. This was ascribed to the enhanced dislocation mobility on the graphene surface as
discussed before. However, for the Al/SiC nanocomposite, the presence of the reinforcement
was accompanied by the creation of a stress field at the interface zone (Fig. 7a). This created
an area with high dislocation density manifesting itself by intensifying the pile-up
concentration underneath the indenter, which is in line with our previous observations
depicted in Figs 8g and 8h.

Fig. 12. Pile-up features from the frontal (a, c, e) and top (b, d, f) views for the introduced samples:
Left: Al/SiC, Middle: Pure Aluminum, and Right: Al/graphene.

3.5. Effect of the depth of graphene below the surface

As discussed in the introduction part, there are several experimental-based studies


demonstrating that mechanical characteristics of pristine Al could be enhanced significantly
in the presence of small amounts of graphene sheets. For example, Wang et al. [62] showed
16
that adding only 0.3 wt% of this 2D nanofiller to the Al matrix causes an increase of 62% in
the tensile strength of the pure matrix. This is ascribed to the higher specific surface area and
rippled topography of these 2D nanostructures providing a route for enhanced load transfer in
the interfacial area and thus leading to better mechanical performance. However, as analyzed
in the current work, the presence of graphene layers near the free surface of the Al matrix
would lead to a deterioration of the wear resistance of the metal matrix by enhancing the
friction coefficient. With respect to the fact that increasing this parameter is unfavorable in
most mechanical systems, we attempt to reduce the COF of Al/graphene nanocomposites
through new designs. For this purpose, the scratch test of this sample was carried out by
positioning the graphene reinforcement at the depths of 40, 50, 60, 80, and 160 Ȧ beneath the
Al surface. Repeating the procedure described in the previous sections, the force-
displacement graphs for these samples were obtained, from which their friction coefficient
was determined. The results listed in Table 2 reveal that increasing the depth of the graphene
causes a rapid decrease in the COF of the Al/graphene sample.
In order to gain more insight into this behavior, we analyze the trends observed in the
variation of the frictional and normal forces with the graphene depth in Table 2. As discussed
in Section 3.3, the increase of the frictional force in the presence of graphene near the
indenter is ascribed to the smooth passage of the dislocation loops along the flake surface,
which leads to a stronger pile-up of dislocations in front of the indenter and eventually
increases the frictional force. As this mechanism is also working for the samples having
graphene at larger depth, this parameter has minor effects on the frictional force component,
as seen in Table 2. In contrast, there is a direct correlation between the graphene depth and
the magnitude of the normal force, which pushes up the indenter. Based on the mechanisms
introduced in Section 3.3, placing the graphene near the indenter weakens the nanocomposite
sample under normal loading conditions. This was attributed to the attractive interactions
between dislocations and the graphene interface. However, as thoroughly analyzed in Ref.
[60], when the graphene is far from the indented surface, the interface depression mechanism
does not work effectively. As a result, the softening behavior disappears and the normal force
component increases monotonically, see 3rd column of Table 2. Therefore, even though the
frictional force is almost unchanged, this increase in the normal force results in the reduction
of the COF. However, beyond the highlighted critical depth (i.e., 80 Ȧ), the presence of
graphene does not have any extra effect on the COF and the results reach a plateau. In
summary, although the presence of graphene sheets within the aluminum is destructive in

17
terms of wear properties, due to their vital role in improving the mechanical properties of the
base matrix, they should be arranged in an optimum way.

Table 2. Average values of the normal and frictional forces along with the friction coefficient for
the Al-based nanocomposites embedded with graphene positioned at different depths. The value of
80 Ȧ is the critical depth, beyond which the effect of graphene disappears.

Graphene depth ( > ) Frictional force (EF) Normal force (EF) Friction coefficient

40 0.114 0.166 0.69


50 0.118 0.172 0.68
60 0.119 0.192 0.62
80 0.121 0.204 0.59
100 0.120 0.203 0.59

4. Conclusion
In this paper, a series of MD simulations was carried out to analyze the role of the nanofiller
geometry on the tribological features of Al-based nanocomposites utilizing a nanoscratching
test. This was achieved through exploring the friction coefficient of an Al matrix embedded
with spherical SiC nanoparticles and graphene platelets under identical conditions. Different
trends were observed in the variation of the COF compared to that of the baseline matrix in
the presence of these reinforcements. It was deduced that SiC NPs cause a decrease in the
friction coefficient of aluminum, while this parameter is enhanced in the Al/graphene sample.
The governing mechanism of this behavior was investigated by exploring the interfacial
interactions of the Al dislocations. The results revealed that the existence of a semi-coherent
interface in the Al/SiC nanocomposite sample leads to local stress concentrations around the
nanoparticles promoting the nucleation of dislocations. Moreover, it was demonstrated that
SiC NPs act as obstacles for the dislocation movement, which normally proceeds via the
well-known cross-slip and obstacle-cutting mechanisms. So, dislocations were accumulated
under the indenter tip creating some locked dislocations at the Al/SiC interface. This
phenomenon was demonstrated to be the main factor leading to an increase of the normal
forces experienced by the indenter in this strengthened zone and thus to a reduction of the
COF. In contrast, for the Al/graphene sample, no significant perturbation was detected during
the dislocation movement; this behavior was attributed to the coherent-like interface region in
this case. Accordingly, more pronounced plastic deformation was observed in front of the
indenter, which led to an increase of the tangential force and, as a result, of the friction

18
coefficient. The mentioned severe plastic deformation for the aluminum/graphene sample
was also addressed employing the pile-up investigation.
Finally, to reduce the discussed weakening effect of graphene, additional simulations were
carried out to design new systems, in which the depth of the graphene layer was varied. It
was demonstrated that there is a critical value for this parameter, beyond which the
detrimental role of graphene on the wear behavior of the aluminum matrix disappears. We
believe that the framework developed here could be useful not only to predict the wear
characteristics of MMNCs, but also to enhance their engineered design with respect to an
improved tribological and mechanical performance.

Acknowledgements
The authors thank Arash Kardani for his help during the DXA-based analysis of the results
and for the corresponding discussions. HMU is grateful for support by the Deutsche
Forschungsgemeinschaft via the International Research and Training Group 2057 (project
number 252408385).

19
References:

[1] S. Achanta and J.P. Celis, Nanotribology of MEMS/NEMS. In: E. Gnecco, E. Meyer,
(eds) Fundamentals of Friction and Wear, NanoScience and Technology, Springer,
Berlin, Heidelberg (2007).
[2] X. Li, B. Bhushan, K. Takashima, C.W. Baek, Y.K. Kim, Mechanical characterization
of micro/nanoscale structures for MEMS/NEMS applications using nanoindentation
techniques, Ultramicroscopy 97 (2003) 481–494.
[3] P. Zhang, H. Zhao, C. Shi, L. Zhang, H. Huang, L. Ren, Influence of double-tip
scratch and single-tip scratch on nano-scratching process via molecular dynamics
simulation, Applied Surface Science 280 (2013) 751– 756.
[4] H. Dai, S. Li, G. Chen, Molecular dynamics simulation of subsurface damage
mechanism during nanoscratching of single crystal silicon, Proceedings of the
Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology 233
(2018) 61-73.
[5] P.X.L. Feng, D.J. Young, C.A. Zorman, MEMS/NEMS Devices and Applications. In:
B. Bhushan, (eds) Springer Handbook of Nanotechnology, Springer Handbooks,
Springer, Berlin, Heidelberg (2017).
[6] J. Li, Q. Fang, Y. Liu, L. Zhang, Scratching of copper with rough surfaces conducted
by diamond tip simulated using molecular dynamics, International Journal of
Advanced Manufacturing Technology 77 (2015) 1057-1070.
[7] X. Zhang, X. Zhong, X. Meng, G. Yi, J. Jia, Adhesion and friction studies of nano-
textured surfaces produced by self-assembling Au nanoparticles on silicon wafers,
Tribology Letters 46 (2012) 65-73.
[8] I.A. Alhafez, A. Brodyanski, M. Kopnarski, H.M. Urbassek, Influence of tip
geometry on nanoscratching, Tribology Letters 65 (2017) 26-39.
[9] A.I. Dimitriev, A.Yu. Nikonov, A.R. Shugurov, A.V. Panin, Numerical study of
atomic scale deformation mechanisms of Ti grains with different crystallographic
orientation subjected to scratch testing, Applied Surface Science 471 (2019) 318-327.
[10] Q. Fang, Y. Tian, J. Li, Q. Wang, H. Wu, Interface-governed nanometric machining
behaviour of Cu/Ag bilayers using molecular dynamics simulation, Royal Society of
Chemistry Advances 9 (2019) 1341–1353.
[11] J. Ren, M. Hao, M. Lv, S. Wang, B. Zhu, Molecular dynamics research on ultra-high-
speed grinding mechanism of monocrystalline nickel, Applied Surface Science 455
(2018) 629–634.
[12] A. Montazeri, A. Mobarghei, Nanotribological behavior analysis of graphene/metal
nanocomposites via MD simulations: New concepts and underlying mechanisms,
Journal of Physics and Chemistry of Solids 115 (2018) 49–58.
[13] L. Si, X. Wang, G. Xie, N. Sun, Nano-adhesion and friction of multi-asperity contact:
a molecular dynamics simulation study, Surface and Interface Analysis 47 (2015)
919–925.
[14] P. Zhu, Y. Z. Hu, T.B. Ma, H. Wang, Molecular dynamics study on friction due to
ploughing and adhesion in nanometric scratching process, Tribology Letters 41
(2011) 41–46.

20
[15] I.A. Alhafez, H.M. Urbassek, Scratching of hcp metals: A molecular-dynamics study,
Computational Materials Science 113 (2016) 187–197.
[16] J. Zhang, H. Zheng, M. Shuai, Y. Li, Y. Yang, T. Sun, Molecular Dynamics Modeling
and Simulation of Diamond Cutting of Cerium, Nanoscale Research Letters 12 (2017)
464-474.
[17] R. Komanduri, N. Chandrasekaran, L.M. Raff, MD simulation of indentation and
scratching of single crystal aluminum, Wear 240 (2000) 113–143.
[18] T. Otieno, K. Abou-El-Hossein, Molecular dynamics analysis of nanomachining of
rapidly solidified aluminium, International Journal of Advanced Manufacturing
Technology 94 (2018) 121–131.
[19] I.A. Alhafez, C.J. Ruestes, H.M. Urbassek, Size of the plastic zone produced by
nanoscratching, Tribology Letters 66 (2018) 20–32.
[20] T. Junge, J.F. Molinari, Molecular dynamics nano-scratching of aluminium: a novel
quantitative energy-based analysis method, Procedia IUTAM 3 ( 2012 ) 192 – 204.
[21] T. Junge, J.F. Molinari, Plastic activity in nanoscratch molecular dynamics simulations
of pure aluminium, International Journal of Plasticity 53 (2014) 90–106.
[22] L. Liu, M. Zhou, L. Jin, L. Li, Y. Mo, G. Su, X. Li, H. Zhu, Y. Tian, Recent advances
in friction and lubrication of graphene and other 2D materials: Mechanisms and
applications, Friction 7 (2019) 1-18.
[23] O. El-Kady, A. Fathy, Effect of SiC particle size on the physical and mechanical
properties of extruded Al matrix nanocomposites, Materials and Design 54 (2014)
348-353.
[24] T. Varol, A. Canakci, S. Ozsahin, Prediction of effect of reinforcement content, flake
size and flake time on the density and hardness of flake AA2024-SiC nanocomposites
using neural networks, Journal of Alloys and Compounds 739 (2018) 1005-1014.
[25] S. Chen, M.K. Hassanzadeh-Aghdam, R. Ansari, An analytical model for elastic
modulus calculation of SiC whisker-reinforced hybrid metal matrix nanocomposite
containing SiC nanoparticles, Journal of Alloys and Compounds 767 (2018) 632–641.
[26] S.A. Khadem, S. Nategh, H. Yoozbashizadeh, Structural and morphological
evaluation of Al–5 vol.%SiC nanocomposite powder produced by mechanical milling,
Journal of Alloys and Compounds 509 (2011) 2221–2226.
[27] M.P. Reddy, R.A. Shakoora, G. Parandeb, V. Manakarib, F. Ubaida, A.M.A.
Mohamedc, M. Gupta, Enhanced performance of nano-sized SiC reinforced Al metal
matrix nanocomposites synthesized through microwave sintering and hot extrusion
techniques, Progress in Natural Science: Materials International 27 (2017) 606–614.
[28] Y. Jiang, R. Xu, Z. Tan, G. Ji, G. Fan, Z. Li, D.B. Xiong, Q. Guo, Z. Li, D. Zhang,
Interface-induced strain hardening of graphene nanosheet/aluminum composites,
Carbon 146 (2019) 17-27.
[29] M. Tavakol, M. Mahnama, R. Naghdabadi, Shock wave sintering of Al/SiC metal
matrix nano-composites: A molecular dynamics study, Computational Materials
Science 125 (2016) 255–262.
[30] H. Gu, X.L. Gao, and X.C. Li, Molecular dynamics study on mechanical properties
and interfacial morphology of an aluminum matrix nanocomposite reinforced by β-
silicon carbide nanoparticles, Computational and Theoretical Nanoscience 6 (2009)
61–72.
[31] Y. Yan, S. Zhou, S. Liu, Atomistic simulation on mechanical behaviors of Al/SiC
nanocomposites, International Conference on Electronic Packaging Technology 18
(2017) 357-362.

21
[32] H. Gu, J.J. Wang, Z. Li, Molecular dynamics simulation of tensile behavior on
ceramic particles reinforced aluminum matrix nanocomposites, International Journal
of Materials Science and Applications 5 (2016) 151-159.
[33] S. Huo, L. Xie, J. Xiang, S. Pang, F. Hu, U. Umer, Atomic-level study on mechanical
properties and strengthening mechanisms of Al/SiC nano-composites, Applied
Physics A 124 (2018) 209–221.
[34] Z. Zhang, H.M. Urbassek, Dislocation-based strengthening mechanisms in metal-
matrix nanocomposites: a molecular dynamics study of the influence of reinforcement
shape in the Al-Si system, Computational Materials Science 145 (2018) 109–115.
[35] Y. Rong, H.P. He, L. Zhang, N. Li, Y.C. Zhu, Molecular dynamics studies on the
strengthening mechanism of Al matrix composites reinforced by graphene
nanoplatelets, Computational Materials Science 153 (2018) 48–56.
[36] J. Zhong, R. Shakiba, J. B. Adams, Molecular dynamics simulation of severe adhesive
wear on a rough aluminum substrate, Journal of Physics D: Applied Physics 46 (2013)
055307.
[37] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, Journal of
Computational Physics 117 (1995) 1–19.
[38] W.G. Hoover, Canonical dynamics: Equilibrium phase-space distributions, Physical
Review A 31 (1985) 1695-1697.
[39] M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Second edition, Oxford
University Press, 2017.
[40] A.T. AlMotasem, J. Bergström, A. Gåård, P. Krakhmalev, L.J. Holleboom, Tool
microstructure impact on the wear behavior of ferrite iron during nanoscratching: An
atomic level simulation, Wear 370-371 (2017) 39-45.
[41] A. Stukowski, Visualization and analysis of atomistic simulation data with OVITO–
the Open Visualization Tool, Modelling and Simulation in Materials Science and
Engineering 18 (2010) 015012.
[42] A. Stukowski, K. Albe, Extracting dislocations and non-dislocation crystal defects
from atomistic simulation data, Modelling and Simulation in Materials Science and
Engineering 18 (2010) 085001.
[43] R. Komanduri, N. Chandrasekaran, L.M. Raff, Molecular dynamics simulation of
atomic-scale friction, Physical Review B 61 (2000) 14007– 14019.
[44] T. Tsuru, Y. Shibutani, Atomistic simulations of elastic deformation and dislocation
nucleation in Al under indentation-induced stress distribution, Modelling and
Simulation in Materials Science and Engineering 14 (2006) 55–62.
[45] P. Peng, G. Liao, T. Shi, Z. Tang, Y. Gao, Molecular dynamic simulations of
nanoindentation in aluminum thin film on silicon substrate, Applied Surface Science
256 (2010) 6284–6290.
[46] M.I. Mendelev, M.J. Kramer, C.A.Becker, M. Asta, Analysis of semi-empirical
interatomic potentials appropriate for simulation of crystalline and liquid Al and Cu,
Philosophical Magazine 88 (2008) 1723–1750.
[47] J. Tersoff, Empirical interatomic potential for silicon with improved elastic properties,
Physical Review B 38 (1988) 9902-9905.
[48] D.W. Brenner, O.A. Shenderova, J.A. Harrison, S.J. Stuart, B. Ni, S.B. Sinnott, A
second-generation reactive empirical bond order (REBO) potential energy expression
for hydrocarbons, Journal of Physics: Condens Matter 14 (2002) 783–802.
[49] A.Mokhalingam, D. Kumara, A. Srivastava, Mechanical behaviour of graphene
reinforced aluminum nano composites, Materials Today: Proceedings 4 (2017) 3952–
3958.

22
[50] B.K. Choi, G.H. Yoon, S. Lee, Molecular dynamics studies of CNT-reinforced
aluminum composites under uniaxial tensile loading, Composites Part B: Engineering
91 (2015) 119-125 .
[51] N. Silvestre, B. Faria, J.N.C. Lopes, Compressive behavior of CNT-reinforced
aluminum composites using molecular dynamics, Composites Science and
Technology 90 (2014) 16–24.
[52] W.G. Jiang, Y. Wu, Q.H. Qin, D.S. Li, X.B. Liu, M.F. Fu, A molecular dynamics
based cohesive zone model for predicting interfacial properties between graphene
coating and aluminum, Computational Materials Science 151 (2018) 117–123.
[53] C.R. Dandekar, Y.C. Shin, Molecular dynamics based cohesive zone law for
describing Al–SiC interface mechanics, Composites: Part A 42 (2011) 355–363.
[54] G. Ziegenhain, A. Hartmaier, H.M. Urbassek, Pair vs many-body potentials: influence
on elastic and plastic behavior in nanoindentation of FCC metals, Journal of the
Mechanics and Physics of Solids 57 (2009) 1514–1526.
[55] P. Zhu, F. Fang, Study of the minimum depth of material removal in nanoscale
mechanical machining of single crystalline copper, Computational Materials Science
118 (2016) 192–202.
[56] A. Waghmare and P. Sahoo, Friction analysis at elastic-plastic contact of rough
surfaces using n-point asperity model, Proceedings of the Institution of Mechanical
Engineers, Part J: Engineering Tribology 230 (2016) 1258-1272.
[57] D.A. Porter, K.E. Easteling, Phase Transformations in Metals and Alloys, second
edition, Chapman Hall, 1981.
[58] D. Chen, L.L. Costello, C.B.Geller, T. Zhu, D.L. McDowell, Atomistic modeling of
dislocation cross-slip in nickel using free-end nudged elastic band method,
Computational Materials Science 168 (2019) 436–447.
[59] A. Kardani, A. Montazeri, MD-based characterization of plastic deformation in
Cu/Ag nanocomposites via dislocation extraction analysis: Effects of nanosized
surface porosities and voids, Computational Materials Science 152 (2018) 381–392.
[60] V. H. Vardanyan, H. M. Urbassek, Dislocation interactions during nanoindentation of
nickel-graphene nanocomposites, Computational Materials Science 170 (2019)
109158.
[61] Y. Gao, A. Brodyanski, M. Kopnarski, H.M. Urbassek, Nanoscratching of iron: A
molecular dynamics study of the influence of surface orientation and scratching
direction, Computational Materials Science 103 (2015) 77–89.
[62] J. Wang, Z. Li, G. Fan, H. Pang, Z. Chen, D. Zhang, Reinforcement with graphene
nanosheets in aluminum matrix composites, Scripta Materialia 66 (2012) 594-597.

23
- Tribological features of Al-based nanocomposites are analyzed by MD simulation
- Influence of nanofiller geometry on the COF of the Al matrix is thoroughly studied
- Different trends are observed in the presence of spherical and platelet nanofillers
- Underlying mechanisms are investigated via DXA analysis of the interfacial area
- Dependence of the friction coefficient on the graphene beneath depth is explored
- Results are applicable toward the engineered design of metal-matrix nanocomposites
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

Das könnte Ihnen auch gefallen