You are on page 1of 15

Wear 418–419 (2019) 290–304

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Wear behavior of silica and alumina-based nanocomposites reinforced with T


multi walled carbon nanotubes and graphene nanoplatelets
Nidhi Sharmaa, Syed Nasimul Alama, , Bankim Chandra Raya, Surekha Yadavb, Krishanu Biswasb

a
Metallurgical and Materials Engineering Department, National Institute of Technology Rourkela, Rourkela, Orissa 769008, India
b
Material Science and Engineering Department, Indian Institute of Technology Kanpur, Kanpur, Uttar Pradesh 208016, India

ARTICLE INFO ABSTRACT

Keywords: Spark plasma sintering (SPS) technique has been employed to fabricate silica (SiO2) and alumina (Al2O3) based
Silica nanocomposites having multiwalled carbon nanotubes (MWCNTs) and exfoliated graphite nanoplatelets (xGnPs)
Alumina as reinforcement phase. The influence of MWCNTs and xGnPs on the wear behavior of the nanocomposites is
MWCNTs analyzed. SiO2-MWCNT, SiO2-xGnP nanocomposites having 0.5, 3 and 5 vol% and Al2O3-MWCNT, Al2O3-xGnP
xGnPs
nanocomposites having 0.2, 0.8, 3 and 5 vol% of the respective nanofiller were developed. SiO2-based nano-
Wear resistance
Abrasive wear
composites were fabricated at 1350 °C under a pressure of 40 MPa whereas Al2O3-based nanocomposites were
Adhesive wear prepared at 1450 °C under a pressure of 50 MPa. Experimental results indicate that the wear behavior of various
SiO2 and Al2O3-based nanocomposites improves with the introduction of carbonaceous nanofillers like MWCNTs
and xGnPs. With the addition of these nanofillers, a continuous supply of graphitic reinforcement phase at the
ceramic interface provided a better lubrication to the reinforced nanocomposites, which was not found in the
case of monolithic SiO2 or Al2O3 samples. Both MWCNTs and xGnPs effectively withstand the applied stress
during the wear test owing to their structural morphology once subjected to the wear strain. A remarkably varied
wear behavior of both SiO2 and Al2O3 based nanocomposites was observed. SiO2 based nanocomposites depict
narrower wear tracks with high wear rates whereas Al2O3 based nanocomposites were found to depict broader
wear tracks with low wear rates.

1. Introduction the wear loss in ceramic-based materials under varied operating con-
ditions [4]. However, despite the rapid progress, finding solid lu-
Ceramics like silica (SiO2) and alumina (Al2O3) serve as the po- bricants which could efficiently enhance the wear resistance of ceramic-
tential candidates for several structural and mechanical applications based materials is still a challenge. In the last decade, the addition of
owing to their high-temperature stability, high hardness, good com- nanoparticles as reinforcing agents into ceramic matrices has emerged
pressive strength, exceptional oxidation resistance and relatively low as the latest trend for the production of tougher and wear resistant
density [1]. Owing to these remarkable properties, both SiO2 and Al2O3 ceramic matrix nanocomposites (CMNCs). Nowadays, new functional
based materials are used in many applications like valves, cutting tools CMNCs showing drastic improvement in the dry sliding wear perfor-
and bearings, where friction and wear processes dominate [2]. How- mance against metallic and polymeric counterparts are constantly
ever, due to their extreme brittleness, monolithic ceramics are me- emerging [5]. It has been found that the addition of nanoparticles into
chanically unstable, due to which their use is highly limited in the monolithic ceramics can enhance their load-bearing capacity during
applications requiring friction tolerance. This makes wear a crucial sliding motion and thus decrease the tribological problems like friction,
concern for ceramic-based materials due to high energy and material abrasion and wear. The extremely large surface area of the nanofillers
loss. To reduce the frictional and wear loss in brittle ceramics, various effectively improves the wear resistance of CMNCs [6]. The sliding
new lubricants have been developed in recent years. Lubricants mini- friction coefficient of carbonaceous nanofillers like graphene, graphene
mize the friction between the components in contact and thus sig- oxide, nano-graphite, carbon fibers and carbon nanotubes (CNTs) is
nificantly reduce the energy loss in mechanical and structural assem- reported lowest among all existing solids [7]. Graphene and carbon
blies [3]. Recently, researchers have made several attempts in nanotubes (CNTs) owing to their unrivaled amalgam of mechanical,
developing new lubricants and coatings which can effectively decrease thermal, chemical and electrical properties are the ideal choices as


Corresponding author.
E-mail address: syedn@nitrkl.ac.in (S.N. Alam).

https://doi.org/10.1016/j.wear.2018.10.008
Received 23 July 2018; Received in revised form 12 September 2018; Accepted 9 October 2018
Available online 15 October 2018
0043-1648/ © 2018 Elsevier B.V. All rights reserved.
N. Sharma et al. Wear 418–419 (2019) 290–304

nanostructured solid lubricants for a variety of materials [8]. As a re- tribofilm whereas CNTs are associated with the improvement in the
sult, numerous studies have been conducted on these unique materials fracture toughness of the nanocomposite by preventing the pull-out of
and impressive improvements in wear, friction, strength and toughness alumina grains during the wear test.
have been observed [9,10]. A significantly small amount of these solid The present work focuses on the study of wear mechanisms in SiO2
lubricants can explicitly affect the friction stability, sliding distance, and Al2O3-based nanocomposites developed using MWCNTs and xGnPs
wear resistance and torque variation of the CMNCs. A small loading as nanofillers and the effect of both nanofillers on the wear behavior of
level of graphene and CNTs could dramatically reduce the wear loss. It various SPSed SiO2-MWCNT, SiO2-xGnP, Al2O3-MWCNT and Al2O3-
has also been demonstrated that these nanofillers work equally well in xGnP composites has been compared. The results suggest that xGnPs
varied environmental conditions (dry as well as humid), which is not serve as a novel substitute over MWCNTs in improving the mechanical
possible for other commonly used solid lubricants [11]. The wear me- properties such as hardness and fracture toughness. It has been ob-
chanism of MWCNTs and graphene is ascribed to the debonding of in- served that the Al2O3-based nanocomposites show lower wear rate as
plane C-C bonds and shear forces acting at the matrix-nanofiller inter- compared to the SiO2-based nanocomposites.
face [12]. For CMCs which are frequently subjected to high loads and
stresses, self-lubricating solid nanofillers are preferred due to their 2. Experimental details
ability of automatic self-release during the wear process to reduce the
frictional loss. For tribological components, it is a key necessity to 2.1. Nanofiller synthesis and preparation of nanocomposite powder mixture
provide low friction, minimal wear, least deformation and high stress
bearing ability. To achieve these properties in CMNCs, CNTs and gra- For the fabrication of various SiO2 and Al2O3-based nanocompo-
phene effectively work as novel solid lubricants in various operating sites, xGnPs were prepared by acid intercalation method and later ul-
conditions. The wear behavior of CMNCs reinforced with these nano- trasonicated for 20 h to avoid the problem of agglomeration during
fillers is mainly effected by two different mechanisms: the enhancement their dispersion in the respective ceramic matrix. The schematic dia-
in mechanical attributes caused by the inclusion of nanofiller in the gram in Fig. 1 describes the steps followed during the synthesis of
ceramic matrices and the emergence of a lubricating tribo-film over the xGnPs.
contact area due to the exfoliation of the nanofillers [13]. For the synthesis of MWCNTs, low-pressure chemical vapor de-
Many authors have stated the formation of a tribo-layer during the position (LPCVD) technique was used. Owing to their high van der
wear experiments in CNT and graphene reinforced CMNCs. Belmonte Waals forces and high aspect ratio, an adequate surface modification of
et al. [14] studied the formed tribo-film using micro Raman and re- MWCNTs is necessary to overcome the problem of agglomeration and
ported it to be comprised of debris consisting of graphitized and de- enhance their homogeneous dispersion in the ceramic matrices. To
formed CNTs pulled out from the Si3N4 based composites. Several achieve this property, MWCNTs were acid functionalized with H2SO4
works have documented the high wear performance in graphene and and HNO3 in the ratio of 3:1. The route followed to synthesize MWCNTs
CNT reinforced CMNCs due to the toughening processes like crack is illustrated in the schematic diagram in Fig. 2(a) and the process of
branching, crack deflection and crack bridging combined with the acid functionalization of MWCNTs is shown in Fig. 2(b).
evolution of a protective lubricating film on the worn surface [15,16]. For the fabrication of various nanocomposites, SiO2 powder
At the sliding speed of 100 mm s−1 and load of 25 N, a decrease of (100–200 mesh size) was procured from Thermo Fischer Scientific Pvt.
~91.5% in the wear rate of Al2O3-exfoliated graphene composites Limited, India and alumina (α-Al2O3) powder was procured from RFCL
having 1.0 vol% of graphene was reported by Kim et al. [17]. A de- Limited, India. Powder metallurgy technique was opted to fabricate
crement of ~50% in the wear rate and ~10% in the coefficient of various SiO2-MWCNT, SiO2-xGnP, Al2O3-xGnP and Al2O3-MWCNT na-
friction of SPSed Al2O3-0.22 wt% graphene composite was reported by nocomposites. For the preparation of nanocomposite powder mixtures,
Gonzalez et al. [18]. Balani et al. [19] observed a reduction of ~49 a Fritsch Pulverisette 5 planetary ball mill was used. The ball milling of
times in the sliding wear volume of Al2O3–8 wt% CNT coating against a various powder mixtures was carried out at the rotational speed of
ZrO2 pin as compared to pure Al2O3 coating. Similarly, Keshri et al. 2.5 s−1 for 30 min with zirconia (ZrO2) milling media.
[20] also reported a reduction of ~72% in the wear rate of Al2O3–8 wt
% CNT coating against a WC ball as compared to pure Al2O3 coating. 2.1.1. Preparation of nanocomposite powder mixtures
Yazdani et al. [21] examined the tribological behavior of Al2O3 hybrid Before the introduction into SiO2 or Al2O3 matrix, xGnPs were ul-
CMNCs reinforced with graphene and carbon and reported that gra- trasonicated for 20 h and MWCNTs were acid functionalized as pre-
phene nanoplatelets mainly contribute in the formation of lubricating sented in Fig. 1 and Fig. 2(b) respectively. The steps followed for the

Fig. 1. Synthesis of xGnPs by acid intercalation method.

291
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 2. (a) The synthesis of MWCNTs by LPCVD method, (b) Process of acid functionalization of MWCNTs.

preparation of various SiO2-MWCNT, SiO2-xGnP, Al2O3xGnP and


Al2O3-MWCNT nanocomposite powder mixtures are as follows:

2.2. Sintering and densification


Al2O3-MWCNT nanocomposites was carried out at 1450 °C for 10 min
SiO2-MWCNT, SiO2-xGnP, Al2O3-MWCNT and Al2O3-xGnP nano- whereas for Al2O3-xGnP nanocomposites, SPS was performed at
composites were fabricated from their respective milled powder mix- 1450 °C for the duration of 5 min under a pressure of 50 MPa. The other
tures by spark plasma sintering (SPS). SPS was done using Dr. Sinter sintering parameters maintained during SPS of various SiO2 and Al2O3
515S apparatus (SPS Syntex Inc., Kanagawa, Japan). Pure SiO2, SiO2- based nanocomposites are shown in Table 1.
0.5, 3, 5 vol% MWCNT, SiO2-0.5, 3, 5 vol% xGnP and pure Al2O3, At the end of the SPS process, the pressure was removed and the
Al2O3-0.2, 0.5, 0.8, 3, 5 vol% MWCNT, Al2O3-0.2, 0.5, 0.8, 3, 5 vol% sintered samples were naturally cooled inside the furnace. The SPS
xGnP nanocomposites were fabricated. For the fabrication of SiO2- profiles for the fabrication of pure Al2O3 and pure SiO2 samples along
MWCNT and SiO2-xGnP nanocomposites, SPS was carried out at with various SiO2-MWCNT, SiO2-xGnP, Al2O3-MWCNT and Al2O3-xGnP
1350 °C for the duration of 10 min under a pressure of 40 MPa. SPS of nanocomposites are shown in Fig. 3.

292
N. Sharma et al. Wear 418–419 (2019) 290–304

Table 1 Table 2
Sintering parameters during SPS. Process parameters during wear test.
Sintering parameter Specification Sintering parameter Specification

Heating Rate 100 °C/min Diameter of sliding indenter 2 mm


Pulse on-off ratio 12:2 Normal load 9.8 N
Diameter of graphite die 15 mm Sliding time 10 min
Vacuum in the chamber 6 Pa Sliding speed 0.33 s−1
Ambience Ar (flow rate ~2 L/min) Diameter of wear track 6 mm
Current flow ~1200 A
Voltage 20 V
Fig. 4(b). The average particle size of SiO2 particles was observed in the
range of ~70–220 µm. From the SEM images, it can be stated that SiO2
The Archimedes’ method was used to determine the bulk density of particles show a much smoother surface as compared to the Al2O3
the sintered nanocomposites and the rule of mixture was followed to particles. The structural morphology of MWCNTs and xGnPs used in the
obtain their theoretical density. The density of Al2O3, SiO2, xGnP and fabrication of various SiO2 and Al2O3 based nanocomposites is shown in
MWCNT was assumed as 3.95 g/cm3, 2.65 g/cm3, 2.26 g/cm3 and the HRTEM images in Fig. 4(c, d) and Fig. 4(e, f) respectively. The
2.6 g/cm3 respectively. The morphological analysis of the polished HRTEM micrographs in Fig. 4(c, d) confirm the large aspect ratio of
SPSed nanocomposites was carried out using SEM. The nanofillers, MWCNTs having an outer diameter in the range of ~20–50 nm. The
milled powder mixtures and various fabricated nanocomposites were prominent black dots in Fig. 4(c) correspond to the agglomerates of
characterized by several techniques like scanning electron microscopy MWCNTs adhered to the tubular sidewalls [23]. During HRTEM,
(SEM), X-ray diffraction (XRD) and high-resolution transmission elec- MWCNTs were found stable under the irradiation of the electron beam.
tron microscopy (HRTEM) as described elsewhere [22]. The hardness of Fig. 4(e) shows a high magnification HRTEM image of xGnPs comprised
SPSed nanocomposites was measured on a polished cross-section using of several graphene layers stacked together. The overlapped and su-
the Vickers microhardness tester at a load of 500 gf for a period of 10 s. perimposed graphite nanoplatelets are clearly evident in Fig. 4(f). The
An average of five indents was considered for each sample and the size nanoplatelets were found to be electron transparent [24].
of the samples was 5 mm × 15 mm. Fig. 5 depicts the SEM micrographs of various SiO2-MWCNT, SiO2-
Dry sliding wear test of the fabricated pure SiO2, pure Al2O3 and xGnP, Al2O3-MWCNT and Al2O3-xGnP milled powder mixtures. Ob-
various SiO2 and Al2O3-based nanocomposites was performed using a taining near full density, without deteriorating the morphology and
DUCOM TR208-M1 ball-on-plate type tribometer equipped with a structure of the nanofillers is a key challenge during the fabrication of
hardened steel ball (SAE 52100) indenter of 2 mm diameter. The har- CMNCs as most of the mechanical properties of the nanocomposites
dened steel ball slided unidirectionally on the fixed samples and the depend upon the distribution of the nanofillers in the host matrix.
wear test was carried out at the room temperature (~27 °C). The var- Homogeneous mixing and uniform distribution of the nanofillers in the
iation of wear depth and wear rate with respect to the sliding time was matrix is a prime objective during the processing of CMNCs to ensure
investigated for the sintered nanocomposites. The various process optimum densification [25]. The short milling duration of 30 min re-
parameters during wear test are shown in Table 2. sulted in the homogeneous mixing of the nanofillers with SiO2 and
Al2O3 powders. The SEM images in Fig. 5(a, b) and Fig. 5(c, d) show the
3. Results and discussion adherence of MWCNTs to SiO2 and Al2O3 particles respectively. Due to
their large aspect ratio, MWCNTs tend to entangle and form complex
3.1. Microstructural analysis agglomerates [26]. Dense bundles of MWCNTs deposited on the surface
of milled SiO2 and Al2O3 particles could be easily viewed. The SEM
Fig. 4(a) shows the SEM image of pure unmilled Al2O3 particles used micrographs of SiO2-xGnP and Al2O3-xGnP powder mixtures in Fig. 5(c,
in the fabrication of Al2O3-based nanocomposites. The inset image in d) and Fig. 5(g, h) reveal the smooth planar morphology of xGnPs.
Fig. 4(a) clearly reveals the distinct porous surface of monolithic Al2O3 Several graphene nanoplatelets stacked together as a sheet-like struc-
particles. The average diameter of Al2O3 particles was found in the ture are evident in Fig. 5(h). This could be ascribed to the adhesion of
range of ~90–150 µm. The SEM micrograph of pure unmilled SiO2 thin xGnPs due to stresses caused by ZrO2 balls during milling [27].
particles used in the fabrication of various SiO2-based nanocomposites The microstructures of various SPSed SiO2 and Al2O3-based com-
is shown in Fig. 4(b). The irregular shape of SiO2 particles with smooth posites were analyzed to determine the dispersion of the nanofillers in
surfaces and sharp edges can be confirmed by the inset image in the SPSed nanocomposites. Formation of carbides in the form of SiC and

Fig. 3. SPS profile for the preparation of (a) pure SiO2, SiO2-MWCNT and SiO2-xGnP, (b) pure Al2O3, Al2O3-MWCNT, (c) pure Al2O3, Al2O3-xGnP nanocomposites.

293
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 4. SEM images of (a) pure Al2O3 and (b) pure SiO2. HRTEM images of (c, d) MWCNTs and (e, f) xGnPs used in the fabrication of various SiO2 and Al2O3-based
nanocomposites.

Al4C3 nanorods after the high temperature-pressure assisted SPS of toughness and the wear behavior of the nanocomposites [29]. In
SiO2-MWCNT and Al2O3-MWCNT nanocomposites could be observed in Fig. 6(b), well sintered SiO2-MWCNT conglomerate represents the dif-
the SEM images in Fig. 6(a) and Fig. 6(f). The diameter of SiC rods in fusion of MWCNTs with SiO2 grains. In Fig. 6(c, d) xGnPs were found
SiO2-MWCNT nanocomposites and Al4C3 rods in Al2O3-MWCNT nano- well dispersed in between the grain boundaries of the SiO2 particles. In
composites were found to be in the range of ~90–230 nm and Fig. 6(g), diffused xGnPs with Al2O3 grains, forming well sintered
~50–200 nm respectively. The formation of carbide nanorods could be Al2O3-xGnP conglomerates could be easily seen. Several xGnPs stacked
associated with the carbo-thermal reactions of MWCNTs and the together to form thick graphene sheets could be observed in Fig. 6(h). It
ceramic matrix. MWCNTs react with the matrix and form carbides in should be noted that the structure of both the nanofillers was found
the vapor phase, depending upon the atmosphere and pressure during well preserved in the SPSed nanocomposites.
sintering [28]. The extent of carbide formation increased with the in- XRD analysis was carried out for the SPSed nanocomposites to
crease in the amount of MWCNTs from 0.5 vol% to 5 vol% in both the analyze the formation of any new phases during sintering. Peaks cor-
SiO2-MWCNT and Al2O3-MWCNT nanocomposites. Irregular cracks and responding to (002) plane of xGnP at ~26.43° in SiO2-xGnP and Al2O3-
surface damages can be seen in the SEM micrographs. These surface xGnP nanocomposites and (002) plane of MWCNT at ~26.38° in SiO2-
defects and cracks deteriorate the mechanical attributes, fracture MWCNT and Al2O3-MWCNT nanocomposites could be seen in the XRD

Fig. 5. SEM images of milled (a, b) SiO2-3 vol% MWCNT, (c, d) SiO2-3 vol% xGnP, (e, f) Al2O3-3 vol% MWCNT and (g, h) Al2O3-3 vol% xGnP powder mixtures.

294
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 6. SEM of SPSed (a, b) SiO2-3 vol% MWCNT, (c, d) SiO2-3 vol% xGnP, (e, f) Al2O3-3 vol% MWCNT and (g, h) Al2O3-3 vol% xGnP nanocomposites.

plots in Fig. 7(a, b) which confirm that the crystalline structure of the 3.2. Characteristic properties
nanofillers was well preserved in the ceramic matrix during SPS. The
intensity of (002) peak of xGnPs and MWCNTs increases on increasing Fig. 8 shows the variation of relative density with respect to the
the concentration of the nanofillers in the nanocomposites. From volume fraction of MWCNTs and xGnPs in SPSed SiO2 and Al2O3-based
Fig. 7(a), it is evident that sharper peaks of MWCNTs were found in nanocomposites. Fig. 8(a) indicates that the Al2O3-MWCNT nano-
SiO2-based nanocomposites as compared to the Al2O3-based nano- composites developed at 1450 °C for a holding time of 10 min are
composites. The (002) peak of MWCNT at 2θ ~26.38° in the case of denser as compared to SiO2-MWCNT nanocomposites developed at
Al2O3-MWCNT nanocomposites was not as sharp as the peak observed 1350 °C for a holding time of 5 min. Addition of MWCNTs up to the
in the case of SiO2-MWCNT nanocomposites. This could be associated concentration of 3 vol% led to the improvement in the relative density
with the higher thermal conductivity of Al2O3 (12-38.5 W/m K) as of Al2O3-MWCNT nanocomposites. The maximum relative density of
compared to SiO2 (1.3–1.5 W/m K) as a result of which SiO2 could ef- ~99.62% was obtained for Al2O3-3 vol% MWCNT nanocomposites
fectively reduce the MWCNT degradation and preserve its crystallinity whereas the density of pure Al2O3 sample fabricated under similar
[30]. A slight shift in the XRD peak of Al2O3 was seen only in the case of sintering conditions was found to be ∼99.24%. For SiO2-MWCNT na-
the Al2O3-xGnP system. However, any prominent shift in the peaks of nocomposites, a similar trend in the variation of relative density was
Al2O3 or SiO2 was not observed for other systems due to short sintering observed. The minimum relative density of ~99.21% was observed for
time during SPS, which prevented the carbon diffusion from nanofillers SiO2-0.5 vol% MWCNT nanocomposite whereas the maximum relative
into the ceramic matrix [31]. Formation of any new phase could not be density of ~99.47% was observed for SiO2-3 vol% MWCNT nano-
traced from the XRD plots of both SiO2 and Al2O3-based nanocompo- composite. The higher relative densities of Al2O3-MWCNT nano-
sites when the amount of nanofiller was increased up to 5 vol% in the composites as compared to SiO2-MWCNT nanocomposites suggest that
ceramic matrices. It should also be noted that several small peaks ob- the addition of MWCNTs is more effective in densification of the Al2O3
served for SiO2-0.5 vol% MWCNT and SiO2-0.5 vol% xGnP nano- matrix as compared to the SiO2 matrix. Further addition of MWCNTs in
composites diminish on the increment of the loading level of both the Al2O3 and SiO2 matrix beyond 3 vol% led to a decrease in the density of
nanofillers from 0.5 vol% to 5 vol% in the SiO2 matrix. the Al2O3-MWCNT and SiO2-MWCNT nanocomposites. However, the

Fig. 7. XRD plots of (a) SiO2-MWCNT, Al2O3-MWCNT and (b) SiO2-xGnP, Al2O3-xGnP nanocomposites.

295
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 8. Variation in the relative density of (a) SiO2-MWCNT, Al2O3-MWCNT and (b) SiO2-xGnP, Al2O3-xGnP composites.

Fig. 9. Variation in the hardness of (a) SiO2-MWCNT, Al2O3-MWCNT and (b) SiO2-xGnP, Al2O3-xGnP composites.

density of all the SiO2-MWCNT and Al2O3-MWCNT nanocomposites CMNCs as beyond the loading level of 3 vol% of both MWCNTs and
was found much higher as compared to the density of pure SiO2 and xGnPs, they tend to agglomerate and form weak interfaces with the
Al2O3 samples. Fig. 8(b) represents the variation in the relative density matrix [33]. However, the hardness value of all the SPSed nano-
of SPSed SiO2-xGnP and Al2O3-xGnP nanocomposites. It is evident from composites was found higher than that of pure sintered SiO2 and Al2O3
the plot that the relative density of the nanocomposites increases with samples. From Fig. 9(a), the hardness of Al2O3-3 vol% MWCNT nano-
the increase in the xGnP content in the ceramic matrices. For SiO2-xGnP composite was found ∼8.32 GPa which was ~8.06% lower than the
nanocomposites, the maximum relative density of ~99.77% was ob- hardness of SiO2-3 vol% MWCNT nanocomposite (∼9.05 GPa). An
served for SiO2-3 vol% xGnP nanocomposite whereas for Al2O3-xGnP overall improvement of ~35.2% in the hardness value of pure Al2O3
nanocomposites an optimal loading level of 0.8 vol% xGnP in the Al2O3 sample (∼5.43 GPa) was observed when 3 vol% of MWCNT was added
matrix resulted in the maximum density of ~99.64%. Any further ad- to it. Similarly, on optimum loading level of MWCNTs (3 vol%), an
dition of xGnPs beyond the optimum loading level results in their ag- increase of ~28.83% in the hardness value of SiO2-3 vol% MWCNT
glomeration in the ceramic matrix which deteriorates the densification nanocomposite (~9.05 GPa) was noticed as compared to the pure SiO2
of the nanocomposites [32]. sample (~6.44 GPa). In the case of xGnP reinforced nanocomposites
Fig. 9 shows the variation in the hardness of various SiO2 and Al2O3- (Fig. 9(b)), the highest hardness value was obtained for Al2O3-3 vol%
based nanocomposites with respect to the loading level of MWCNTs and xGnP nanocomposite and SiO2-3 vol% xGnP nanocomposite. The
xGnPs. For MWCNT reinforced nanocomposites (Fig. 9(a)), SiO2- hardness of Al2O3-3 vol% xGnP nanocomposite and SiO23- vol% xGnP
MWCNT nanocomposites possess better hardness as compared to Al2O3- nanocomposite were observed ∼10.91 GPa and ∼9.74 GPa respec-
MWCNT nanocomposites whereas for xGnP reinforced composites tively. An increase in the concentration of xGnPs beyond 3 vol% led to
(Fig. 9(b)) the hardness value of Al2O3-xGnP nanocomposites was found the decrease in the hardness of the nanocomposites. Nanofillers restrict
higher as compared to SiO2-xGnP nanocomposites. However, all the the grain growth in the ceramic matrix by enhancing the interfacial
four nanocomposite systems were found to follow the same pattern in bonding at the matrix-nanofiller interface. Nanofillers like MWCNTs
the increment of the hardness value. The major challenge in the de- and xGnPs bridge the grains and pin the grain boundaries which leads
velopment of SiO2-MWCNT and Al2O3-MWCNT nanocomposites is the to higher hardness of MWCNT and xGnP reinforced composites caused
homogenous dispersion of MWCNTs in the ceramic matrices due to by grain boundary pinning. Both xGnPs and MWCNTs prevent the grain
their high agglomeration tendency owing to their very high aspect boundary motion by exerting the pinning pressure which impedes the
ratio. Introduction of MWCNTs and xGnPs to monolithic SiO2 and Al2O3 driving force pushing the grain boundaries [34].
led to a notable improvement in their hardness. However, on the ad-
dition of the nanofillers beyond an optimal loading level of 3 vol%, a 3.3. Wear behavior
decrease in the value of hardness was observed for all the four nano-
composite systems. It implies that an optimum loading level of nano- Wear in CMNCs is mainly caused due to abrasion, cracks and re-
fillers is essential for achieving a significant increase in the hardness of sidual stresses. The various signs indicating wear in the nanocomposites

296
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 10. SEM micrographs of the wear tracks of (a-c) SiO2-0.5, 3, 5 vol% MWCNT, (d-f) SiO2-0.5, 3, 5 vol% xGnP, (g-i) Al2O3- 0.8, 3, 5 vol% MWCNT and (j-l) Al2O3-
0.8, 3, 5 vol% xGnP nanocomposites.

include increased surface roughness, microscopic changes in surface analyzed. It has been found that the addition of both MWCNTs and
morphology, fatigue and micro-cracks. The major factors contributing xGnPs led to the decrement in the diameter of the wear tracks. The
towards the wear in CMNCs are material structure, interaction condi- diameter of the wear tracks was found to be decreasing up to the ad-
tions like load, stress, time and surface texture [35]. dition of 3 vol% of both MWCNTs and xGnPs irrespective of the host
The sliding wear test of various SiO2 and Al2O3-based nano- matrix. Although, on increasing the concentration of the nanofillers up
composites was conducted. The wear rate, wear depth, mechanism of to 5 vol%, the diameter of the wear tracks shows an instantaneous in-
nanofiller pull out during wear test, the morphology of wear tracks and crement. Both MWCNTs and xGnPs significantly enhance the wear re-
the effect of the nanofillers on the wear behavior of composites were sistance of the nanocomposites by the formation of a lubricating

Fig. 11. Wear tracks of (a) SiO2-3 vol% MWCNT, (b) SiO2-3 vol% xGnP nanocomposites.

297
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 12. Schematic diagrams showing (a) two-body, (b) three-body abrasive wear mechanisms.

tribofilm on the wear tracks. On addition beyond an optimum value, was found dominant where the surface layer of SiO2 based composites
these graphitic nanofillers tend to agglomerate owing to their high was gradually peeled layer-by-layer. The debris obtained from the worn
surface area. This deteriorates the uniformity of the lubricating tribo- surface after the sliding wear test acted as a third body collapsed in
film and ultimately reduces the wear resistance of nanocomposites between the contact surfaces and changed the wear mechanism from
containing higher volume fractions of the nanofillers [14,21]. the two-body abrasion to the three-body abrasive wear. A schematic
By the SEM micrographs in Fig. 10, it is clear that the width of the diagram illustrating the abrasive wear mechanism is shown in Fig. 12.
wear tracks for MWCNT reinforced composites is much lower as com- This resulted in the galling of the wear track surface. The galling
pared to the width of the wear tracks of CMNCs reinforced with xGnPs. parallel to the sliding direction created thin deformed flat sheets which
The wear tracks observed for SiO2-MWCNT (Fig. 10(a-c)) and Al2O3- later caused delamination of the wear track. With an increase in the
MWCNT (Fig. 10(g-i)) nanocomposites depict much lower width than sliding time compacted thin flat sheets were created which finally led to
the wear tracks of SiO2-xGnP nanocomposites (Fig. 10(d-f)) and Al2O3- the delamination of the track surface [37]. Therefore, galling served as
xGnP nanocomposites (Fig. 10(j-l)). This variation can be attributed to the mode of material removal in SiO2 based nanocomposites. A sche-
the coiling and rolling behavior of MWCNTs when subjected to the wear matic diagram illustrating the galling during wear is shown in Fig. 13.
stresses. As the sliding duration increased, the wear stresses caused due to
When MWCNTs come in the contact of sliding indenter, they coil up the abrasion of matrix-nanofiller interface increased [38]. The in-
and roll over to a different location on the wear tracks without de- creased stresses reduced the shear forces caused by the frictional heat
formation and fragmentation of their tubular structure. On the other induced during the sliding motion. However, as the sliding continued,
hand, xGnPs due to their flake-like morphology, crumple and scroll the abrasion exceeded the shear and led to the deformation of the
along the sliding direction on the application of the wear stresses. This contact surfaces. Due to the applied stresses, the surface layers com-
results in the degradation of xGnPs and decreases their efficiency in pressed and deformed, resulting in the emergence of thin plate-like
improving the wear resistance of CMCs. Due to the morphological dif- sheets. With the increase in the accumulated strain, the deformed layers
ference between the two nanofillers, MWCNTs provide a uniform lu- flattened and ultimately delaminated [39].
bricity by the formation of a continuous tribofilm and serve more ef- The wear mechanism of Al2O3 based nanocomposites have been
ficiently in the wear reduction of the CMNCs as compared to the analyzed in Fig. 14. The wear tracks of Al2O3-3 vol% MWCNT and
protective film formed by the xGnPs [36]. Al2O3-3 vol% xGnP nanocomposites shown in Fig. 14(a, b) depict lower
In order to examine the lubricating effects of MWCNTs and xGnPs, delamination as compared to the SiO2-based nanocomposites. The wear
the wear tracks of SiO2-3 vol% MWCNT, SiO2-3 vol% xGnP and Al2O3- tracks having several micro-pits and deep grooves along the sliding
3 vol% MWCNT, Al2O3-3 vol% xGnP nanocomposites were analyzed direction could be easily seen. Due to continuous sliding, the composite
using SEM. It is evident from Fig. 11(a, b) and Fig. 14(a, b) that the surface breaks into smaller abrasive particles which remain entrapped
width of wear tracks of SiO2-based nanocomposites was much lower as at the contact junction of the indenter and the sliding surface and thus
compared to the width of wear tracks of Al2O3-based nanocomposites, promotes the abrasive wear. The entrapment of abrasive Al2O3 particles
although the loading level of both MWCNTs and xGnPs was kept same on the wear track surface transforms the two-body abrasion into three-
as 3 vol% for both CMCs. The variation in the morphology of wear body abrasion, similarly as observed in the case of SiO2-based nano-
tracks in both matrices could be attributed to the difference in the composites. The entrapped Al2O3 particles elevate the abrasion and
alignment of the nanofillers in the host matrix. lead to the formation of deep grooves and pits on the wear track. The
For the SiO2 based nanocomposites, delamination was observed as wear mechanism of Al2O3 based nanocomposites was characterized by
the major mode of wear. A similar wear phenomenon was observed for two distinct mechanisms: abrasive deformation and adhesive de-
both SiO2-MWCNT and SiO2-xGnP nanocomposites. The abrasive wear formation. The abrasive wear was caused due to the rolling or sliding of

Fig. 13. Schematic diagram of various wear mechanisms.

298
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 14. Wear tracks of (a) Al2O3-3 vol% MWCNT, (b) Al2O3-3 vol% xGnP nanocomposites.

abrasive Al2O3 particles on the surface of lubricating nanofillers. During surface topography of the wear track of Al2O3-3 vol% MWCNT nano-
sliding, the wear track was continuously ploughed by the abrasive composite. The deformed wear track is evident in Fig. 16(e). The worn
matrix particles and the worn material dislocated towards the sides of surface exhibits continuous shallow grooves suggesting that the applied
the formed grooves. On further sliding, the similar mechanism repeated load was accumulated at the contact surfaces. Densely packed Al2O3
and led to the formation of deep grooves at the edges of the wear tracks. particles are visible on the worn surface in Fig. 16(f). Fine debris ac-
The formed grooves were continuous, indicating that the tip of the cumulated on the edges of the wear track could be observed on the
sliding indenter was firmly embedded in the counterface [40]. worn surface of Al2O3-3 vol% xGnP nanocomposite in Fig. 16(g). This
Dry sliding in the Al2O3-based nanocomposites also led to the for- suggests that the xGnP pull-out was caused during the sliding wear.
mation of a high-pressure zone between the contacting asperities, Formation of lateral cracks due to material removal and fracture during
which resulted in their adhesion. These adhesive zones formed due to plastic deformation is evident in Fig. 16(h). As the sliding distance in-
the combination of both normal and tangential stresses were later creased, the majority of xGnPs were pulled out from the wear track, as
sheared by the abrasive motion during sliding. Fig. 14(b) shows the seen in Fig. 16(h) [42].
surface of the wear track of Al2O3-3 vol% xGnP nanocomposite, which Fig. 17 illustrates the SEM micrographs of the wear debris obtained
confirms that the worn surface was severely deformed due to the in- from the wear tracks of SiO2-3 vol% MWCNT, SiO2-3 vol% xGnP, Al2O3-
tensive adhesion of the contact surface and sliding indenter which re- 3 vol% MWCNT and Al2O3-3 vol% xGnP nanocomposites. MWCNTs
sulted in the ploughing and ultimately led to the formation of deep adhered to the SiO2 particles are clearly visible in Fig. 17(a). A con-
grooves and micro-pits. The subsequent removal of the excess material siderably large volume of MWCNTs in the wear debris suggests that the
after prolonged sliding caused the visibility of the subsurface on the interfacial bonding between the MWCNTs and SiO2 matrix was weak
wear tracks [41]. A schematic diagram describing the adhesive wear which promoted the easy pull out of the nanofiller from the host matrix
mechanism is presented in Fig. 15. during the wear test. Flower-like SiC nanostructures were noticed in the
Fig. 16 shows the microstructural characteristics of the worn sur- wear debris collected from the SiO2-3 vol% MWCNT nanocomposite.
faces observed on the wear tracks of various SiO2 and Al2O3-based SiC nanorods were also observed in the SEM micrographs of the SiO2-
nanocomposites. Fig. 16(a) shows the surface deformation of the MWCNT nanocomposite in Fig. 6(a). The SEM image in Fig. 17 (b)
smooth delaminated wear track of SiO2-3 vol% MWCNT nanocompo- shows the wear debris obtained from the wear track of SiO2-3 vol%
site. A deep irregular groove having MWCNTs on its edges, formed by xGnP nanocomposite. The agglomerate of diffused SiO2 and fine xGnPs
galling could be evidently seen in Fig. 16(b). This suggests the occur- adhered to the SiO2 particles could be observed. It is well known that
rence of MWCNT pull-out during the galling (Fig. 13). The morphology the grain boundaries in the monolithic ceramics remain under tensile
of the wear track of SiO2-3 vol% xGnP nanocomposite is shown in stresses which are responsible for the crack propagation between the
Fig. 16(c, d). The worn surface was comprised of two distinct regions: grains and promote their pull out. These tensions thereby are the major
(i) areas revealing the scratches caused by the deformation during cause for the generation of the wear debris. The incorporation of the
ploughing and (ii) the areas depicting a large amount of removed ma- nanofillers like xGnPs and MWCNTs reduce the stresses between the
terial due to xGnP pull-out. Continuous surface scratches could be ea- grains and thus reduce the pullout phenomenon [43]. Fig. 17(c) is the
sily observed in Fig. 16(d) which indicate that the tip of the sliding SEM image of the wear debris collected from the wear track of Al2O3-
indenter was firmly embedded in the counterface. Fig. 16(e, f) show the 3 vol% MWCNT nanocomposite. It could be observed that the wear

Fig. 15. Schematic diagram illustrating the adhesive wear mechanism.

299
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 16. Surface topography of the worn surfaces on the wear tracks of (a, b) SiO2-3 vol% MWCNT, (c, d) SiO2-3 vol% xGnP, (e, f) Al2O3-3 vol% MWCNT, (g, h) Al2O3-
3 vol% xGnP nanocomposites.

debris consists of both MWCNTs and Al2O3 particles. Agglomerates of Fig. 18 depicts the variation in the wear rate of various SiO2 and
sintered Al2O3 with MWCNTs adhered to its surface could be easily Al2O3-based nanocomposites. It is clear that the Al2O3-based nano-
seen. The weakly bonded grains in the Al2O3 matrix chipped off during composites show better wear resistance than the SiO2-based nano-
the wear test. In Fig. 17(d), the sintered aggregate of Al2O3 and xGnPs composites. It implies that the Al2O3 matrix serves as a better host for
in the form of a lump pulled out from the wear track of Al2O3-3 vol% both MWCNTs and xGnPs in the formation of lubricating tribofilm as
xGnP nanocomposite could be seen. These adhered xGnPs act as self- compared to the SiO2 matrix. For SiO2 based nanocomposites, the ad-
lubricating layers between the nanocomposite surface and the sliding dition of both MWCNTs and xGnPs up to 3 vol% reduced the wear rate
indenter. xGnPs form a lubricating tribofilm between the indenter and of both SiO2-MWCNT and SiO2-xGnP nanocomposites. However, the
the composite surface which not only lowers the wear rate but also wear rate again increased on the addition of both the nanofillers up to
reduces the sliding abrasion during the initial wear stage [18]. How- 5 vol% loading level. The minimum wear rate observed for SiO2-3 vol%
ever, when the sliding distance increased, the majority of xGnPs were xGnP composites was ~0.265 × 10−4 g/Nm which was ~33.4% times
worn-out from the wear track (Fig. 14(g, h)). lower than the wear rate observed for SiO2-3 vol% MWCNT composite

Fig. 17. Wear debris obtained from the wear tracks of (a) SiO2-3 vol% MWCNT (b) SiO2-3 vol% xGnP (c) Al2O3-3 vol% MWCNT and (d) Al2O3-3 vol% xGnP
nanocomposite.

300
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 18. Variation of wear rate in (a) SiO2-MWCNT, Al2O3-MWCNT and (b) SiO2-xGnP, Al2O3-xGnP nanocomposites.

(~0.398 × 10−4 g/Nm) when developed under similar sintering con- On further addition of xGnPs up to 0.8 vol%, the wear rate slightly
ditions. In the case of Al2O3-based composites, the effect of both increased and attained a value ~0.195 × 10−4 g/Nm. However, on the
MWCNTs and xGnPs on the variation of wear rate was found different further increment of xGnPs up to 3 vol%, the wear rate showed a de-
than that observed for SiO2-based composites. On the introduction of crease and was observed ~0.143 × 10−4 g/Nm. This confirms that
MWCNTs in the Al2O3 matrix, wear rate continuously decreased up to there was an enhancement of ~26.6% in the wear rate of Al2O3-3 vol%
the 3 vol% concentration of the nanofiller. The wear rate for pure sin- xGnP nanocomposite over Al2O3-0.8 vol% xGnP nanocomposite. Fur-
tered Al2O3 was found ~0.437 × 10−4 g/Nm whereas the wear rate of ther addition of xGnPs up to 5 vol% in the Al2O3 matrix deteriorated the
Al2O3-3 vol% MWCNT composite was found ~0.238 × 10−4 g/Nm. It wear resistance of the Al2O3-xGnP composites. The wear rate of
implies that an optimal loading level of 3 vol% MWCNT provided an Al2O3–5 vol% xGnP nanocomposite was observed ~0.16 × 10−4 g/Nm
improvement of ~45.5% in the wear rate of the composites. However, which was ~10.6% higher than that observed for Al2O3-3 vol% xGnP
the addition of 5 vol% MWCNT in the Al2O3 matrix increased the wear nanocomposite. However, it is noteworthy that the wear rate of all SiO2
rate (~0.318 × 10−4g/Nm). On the other hand, the nature of variation and Al2O3-based nanocomposites was found lower as compared to pure
in wear rate of Al2O3-xGnP composites was quite different. The wear sintered SiO2 and Al2O3 samples. The improvement in the wear re-
rate of Al2O3-xGnP composite first decreased on a small loading level of sistance of the nanocomposites could be associated with the emergence
0.2 vol% xGnP in the host matrix and was found ~0.065 × 10−4 g/Nm. of a protective tribofilm on the wear tracks by the nanofillers, which

Fig. 19. Variation in the wear depth of (a) SiO2-MWCNT, (b) Al2O3-MWCNT, (c) SiO2-xGnP and (d) Al2O3-xGnP nanocomposites.

301
N. Sharma et al. Wear 418–419 (2019) 290–304

Fig. 20. Variation in the fracture toughness of (a) SiO2-MWCNT, Al2O3-MWCNT and (b) SiO2-xGnP, Al2O3-xGnP composites.

Fig. 21. Toughening mechanisms in (a) SiO2-3 vol% MWCNT, (b) Al2O3-3 vol% MWCNT, (c) SiO2-3 vol% xGnP and (d) Al2O3-3 vol% xGnP nanocomposites.

provided the lubricating effect to the CMNCs [21]. nanocomposites (Fig. 3). The fluctuations in the wear depth of SiO2-
The variation in the wear depth of various SiO2 and Al2O3-based MWCNT nanocomposites can be associated with their non-uniform
nanocomposites with respect to time is shown in Fig. 19. It is evident densification. The wear depth of SiO2-0.5 vol% MWCNT and SiO2-3 vol
from Fig. 19(a) that on increasing the concentration of MWCNTs from % MWCNT nanocomposites in Fig. 19(a) were found to be stable up to
0.5 vol% to 3 vol%, the wear resistance of the SiO2-MWCNT nano- the initial 500 s of the sliding wear test but depicted a sudden fluc-
composites improves. However, a decrease in the wear resistance of tuation after that. The outer surface having higher wear resistance was
SiO2-MWCNT nanocomposites was observed when the MWCNT content removed first during the wear test followed by the material removal
was increased up to 5 vol%. A similar phenomenon was observed for from the inner subsurface having comparatively lesser density and
Al2O3-MWCNT nanocomposites in Fig. 19(b). The wear depth of Al2O3- lower wear resistance [44]. The increase in the wear depth of Al2O3-
based composites was found to be decreasing when MWCNT loading based nanocomposites noticed during the earlier stage of the wear test
level in the Al2O3-MWCNT nanocomposites was increased from 0.2 vol could be associated with the eviction of the rough irregular surface.
% to 3 vol%. However, on further addition of up to 5 vol% MWCNT, the Further, due to the lubricating behavior of the nanofillers, the friction
wear depth increased. The decrement in the wear resistance of the reduced and the wear tracks became much smoother. It is noteworthy,
nanocomposites beyond 3 vol% loading level of MWCNTs can be as- that in the case of various SiO2-xGnP and Al2O3-xGnP nanocomposites,
sociated with the agglomeration of the nanofiller in the ceramic matrix. a comparatively smoother variation of wear depth was observed as
It should be noted that the pattern of wear depth variation in the case of xGnPs show a much lesser tendency to agglomerate than MWCNTs. The
Al2O3-MWCNT nanocomposites was much smoother than the pattern efficacy of the nanofillers in enhancing the wear resistance of the na-
observed for SiO2-MWCNT nanocomposites. This suggests that more nocomposites depends upon the relative tangential stresses applied on
uniform densification was achieved for Al2O3-MWCNT nanocomposites the matrix grains which leads to the decrement of the grain pull out in
than the SiO2-MWCNT nanocomposites. This could be ascribed to the the wear test [17]. The wear depth of SiO2-based nanocomposites was
fact that SPS of Al2O3-MWCNT nanocomposites was done at elevated found much lower as compared to the wear depth observed in the
sintering temperature for longer holding time unlike the SiO2-MWCNT Al2O3-based nanocomposites.

302
N. Sharma et al. Wear 418–419 (2019) 290–304

3.4. Matrix-nanofiller interface and toughening mechanisms MWCNTs and xGnPs are highly efficient in hindering the path of the
crack propagation. Crack deflection, crack bridging, crack branching
The fracture toughness of various SiO2 and Al2O3-based nano- and crack blunting were found to be as the major toughening me-
composites was investigated using single-edge notched beam method chanisms in all the nanocomposite systems.
and KIC values were calculated by the Shetty equation, using the The improvement in the fracture toughness of both SiO2 and Al2O3-
Palmqvist crack model as described elsewhere by Alam et al. [45]. The based nanocomposites is credited to the drag provided by the nano-
variation in the value of fracture toughness of various SiO2 and Al2O3- fillers. A high energy is required to pull out the nanofillers from the host
based nanocomposites with respect to the concentration of the nano- matrix due to their large surface area. Close contact and better inter-
fillers is shown in Fig. 20. The fracture toughness of Al2O3-0.2 vol% facial bonding between the nanofillers and matrix enable firm an-
MWCNT and Al2O3-0.2 vol% xGnP composites was found to be choring of the nanofillers to the matrix grains [50]. In the case of
~3.17 MPa m1/2 and ~3.91 MPa m1/2 respectively, corresponding to an MWCNT reinforced nanocomposites, a significant number of nanorods
increase of ∼8.39% and ∼25.72% respectively to the fracture tough- in the form of SiC and Al4C3 for SiO2-MWCNT and Al2O3-MWCNT
ness value of pure Al2O3 sample (~2.9 MPa m1/2) developed under the composites could be evidently seen in Fig. 21(a, b). The carbide rods
similar sintering conditions. Results suggest that due to its two-di- remain embedded in the ceramic matrix and deflect the cracks, hence
mensional geometry, higher surface area and better interfacial bonding refine the fracture toughness of the nanocomposites [51,52].
with the matrix, xGnPs are more efficient than MWCNTs in improving
the fracture toughness of the ceramic-based nanocomposites. Contra- 4. Conclusions
rily, due to their tendency to form clustered agglomerates, MWCNTs
deteriorates the fracture toughness of the nanocomposites [46]. The Various SiO2 and Al2O3 based nanocomposites reinforced with na-
SiO2-MWCNT nanocomposites show the maximum fracture toughness nofillers like xGnPs and MWCNTs were fabricated through the SPS
value of ∼0.68 MPa m1/2 for SiO2-3 vol% MWCNT nanocomposite technique. Mechanical properties such as hardness and fracture
which is ~22.72% lower than that observed for SiO2-3 vol% xGnP toughness of the developed nanocomposites were investigated. The
composite (∼0.88 MPa m1/2) in Fig. 20(a). The decrement in the frac- sliding wear test was conducted on the SPSed composites to determine
ture toughness of SiO2-MWCNT and Al2O3-MWCNT nanocomposites their wear resistance and involved wear mechanisms. The prime con-
can be ascribed to the emergence of brittle SiC and Al4C3 nanorods at clusions acquired from this work can be outlined as-
the matrix-nanofiller [47]. The formation of carbide nanorods was not
observed in the xGnP reinforced nanocomposites. Although, the addi- 1. A uniform dispersion of the nanofillers like xGnPs and MWCNTs is a
tion of both MWCNTs and xGnPs, beyond their optimum loading level, key necessity to obtain the desired properties in the sintered com-
deteriorated the fracture toughness of both SiO2 and Al2O3-based na- posites. The structure of both nanofillers were maintained after SPS
nocomposites. This can be associated with the non-uniform distribution as confirmed by the SEM and XRD analysis.
of MWCNTs and xGnPs in the respective matrices, resulting in their 2. The emergence of carbide nanorods was observed in MWCNT re-
agglomeration. With the increase in the concentration of the nanofillers inforced nanocomposites. SiC nanorods in SiO2-MWCNT and Al4C3
in the ceramic matrix, they overlap and form an interlinked mesh of the nanorods in Al2O3-MWCNT nanocomposites were observed in the
agglomerates, which acts as a frail interface at the grain boundaries. SEM micrographs of the sintered nanocomposites.
The fracture toughness of various SiO2-based nanocomposites shows a 3. The relative density of the SPSed nanocomposites improved on the
remarkable improvement up to the addition of 3 vol% of both nano- addition of the nanofillers up to 3 vol%. Further addition of nano-
fillers while for Al2O3-based nanocomposites, the same could be seen up fillers beyond 3 vol% up to 5 vol% causes the agglomeration, which
to the addition of 0.8 vol% of both nanofillers. Various toughening deteriorates the relative density of CMNCs.
mechanisms such as crack deflection, crack bridging and crack 4. At an optimum loading level of 3 vol%, SiO2-MWCNT nanocompo-
branching contribute in increasing the fracture toughness of the nano- sites were found to possess better hardness as compared to the
composites (Fig. 21) [48]. The highest value of fracture toughness in Al2O3-MWCNT nanocomposites whereas in the case of xGnP re-
the case of Al2O3-based composites was observed as ~4.15 MPa m1/2 inforced nanocomposites, the hardness of Al2O3-xGnP nanocompo-
for Al2O3-0.8 vol% xGnP nanocomposite which corresponds to an in- sites was found higher than the SiO2-xGnP nanocomposites.
crement of ∼7.22% when compared to the value of fracture toughness 5. The fracture toughness of both SiO2 and Al2O3-based nanocompo-
observed for Al2O3-0.8 vol% MWCNT nanocomposite (~3.85 MPa m1/ sites increased with the addition of the nanofillers in the respective
2
). Similarly, the value of fracture toughness of SiO2-MWCNT nano- matrix only up to an optimum value. In the case of Al2O3-based
composites was observed much lower than that of SiO2-xGnP nano- nanocomposites, the loading level of 0.8 vol% of both MWCNTs and
composites. xGnPs provided the maximum fracture toughness values whereas in
The matrix-nanofiller interface plays a vital role in the toughening the case of SiO2-based nanocomposites, the maximum value of
of CMNCs. The applied stress during sliding wear test is transferred to fracture toughness was observed on the addition of 3 vol% of both
the high strength nano-sized reinforcement at the interface. Thus, a the nanofillers in the host matrix.
strong interface imparts better toughness to the composites whereas a 6. Crack deflection, crack blunting, crack branching and crack bridging
weak interface leads to inefficient utilization of the nanofiller properties were observed as the major toughening mechanisms in all the
due to their pull-out. Microstructural flaws in CMNCs act as a point of composite systems.
concentration for the wear stresses which leads to the crack propaga- 7. Both MWCNTs and xGnPs improved the wear resistance of the na-
tion resulting in the material eviction during wear. The damaged sur- nocomposites due to the formation of a protective lubricating tri-
faces observe higher loads due to the small area in contact and thus bofilm on the wear tracks. The addition of MWCNTs and xGnPs led
create fractured zones in the composites. This implies that the wear to the reduction in the width of the wear tracks. A decrease in the
resistance of the nanocomposites could only be enhanced if the inter- width of the wear track was observed up to the addition of 3 vol% of
facial bonding between the host matrix and the nanofiller is improved both the nanofillers irrespective of the host matrix.
[49]. Microcracks on the surface of various SiO2 and Al2O3-based na- 8. In the case of SiO2 based nanocomposites, galling was found to be
nocomposites were observed to determine the mechanism by which the major mode of wear whereas Al2O3 based composites exhibited
nanofillers arrest the crack propagation in CMCs. The SEM images in both abrasion and adhesion wear mechanisms.
Fig. 21 show the various toughening mechanisms which contribute in 9. The wear rate of Al2O3-based nanocomposites was found much
improving the fracture toughness of the nanocomposites. The SEM lower as compared to the wear rate of SiO2-based nanocomposites.
images imply that at an optimum concentration of 3 vol%, both

303
N. Sharma et al. Wear 418–419 (2019) 290–304

Acknowledgments [23] K.S. Munir, M.Q. Yuncang, L. Daniel, T. Oldfield, P. Kingshott, D.M. Zhu, C. Wen,
Quantitative analyses of MWCNT‐Ti powder mixtures using raman spectroscopy:
the influence of milling parameters on nanostructural evolution, Adv. Eng. Mater.
We sincerely appreciate the assistance provided by the SEM la- 17 (11) (2015) 1660–1669.
boratory of Metallurgical and Materials Engineering Department, [24] S.N. Alam, L. Kumar, N. Sharma, B.C. Ray, Effect of sonication on the synthesis of
FESEM laboratory of Ceramic Engineering Department, HRTEM la- exfoliated graphite nanoplatelets by thermal exfoliation process, Graphene 2 (2)
(2014) 75–87.
boratory of Chemical Engineering Department and XRD laboratory of [25] M. Estili, A. Kawasaki, H. Sakamoto, Y. Mekuchi, M. Kuno, T. Tsukada, The
Physics Department, NIT Rourkela. We are grateful to IIT Kanpur, for homogeneous dispersion of surfactantless, slightly disordered, crystalline, multi-
providing the SPS facility, procured with partial funding from CARE walled carbon nanotubes in α-alumina ceramics for structural reinforcement, Acta
Mater. 56 (15) (2008) 4070–4079.
Funding, IIT Kanpur and Department of Science and Technology, [26] J. Cho, A.R. Boccaccini, M.S.P. Shaffer, Ceramic matrix composites containing
Government of India. We are also thankful to the Materials Science carbon nanotubes, J. Mater. Sci. 44 (8) (2009) 1934–1951.
Laboratory, Physics Department, Jamia Millia Islamia for the help [27] C.F. Burmeister, A. Kwade, Process engineering with planetary ball mills, Chem.
Soc. Rev. 18 (42) (2013) 7660–7667.
provided in the synthesis of MWCNTs.
[28] A. Azarniya, M.S. Safavi, S. Sovizi, A. Azarniya, B. Chen, H.R.M. Hosseini,
S. Ramakrishna, Metallurgical challenges in carbon nanotube-reinforced metal
References matrix nanocomposites, Metals 7 (10) (2017) 1–44.
[29] A.A. Elguezabal, M.H.B. Bernal, Fracture behaviour of α-Al2O3 ceramics reinforced
with a mixture of single-wall and multi-wall carbon nanotubes, Comp. B: Eng. 60
[1] J. Silvestre, N. Silvestre, J. De Brito, An overview on the improvement of me- (2014) 463–470.
chanical properties of ceramics nanocomposites, J. Nanomater. 106494 (2015) [30] Y. Okazaki, M. Kozako, M. Hikita, T. Tanaka, Effects of addition of nano-scale
(2015) 1–13. alumina and silica fillers on thermal conductivity and dielectric strength of epoxy/
[2] P.C. Milak, F.D. Minatto, A. De Noni Jr., O.R.K. Montedo, Wear performance of alumina microcomposites, in: Proceedings of Int. Conf. Solid Diel. Potsdam,
alumina-based ceramics - a review of the influence of microstructure on erosive Germany, 2010, pp. 279–282.
wear, Ceramica 61 (2015) 88–103. [31] P. Palmero, Structural ceramic nanocomposites: a review of properties and powders'
[3] N. De Laurentis, A. Kadiric, P. Lugt, P. Cann, The influence of bearing grease synthesis methods, Nanomaterials. 5 (2015) 656–696.
composition on friction in rolling/sliding concentrated contacts, Tribol. Int. 94 [32] H. Porwal, S. Grasso, M.J. Reece, Review of graphene–ceramic matrix composites,
(2016) 624–632. Adv. Appl. Ceram. 112 (8) (2013) 443–454.
[4] W. Zhang, G.J. Ma, C.W. Wu, Anti-friction, wear-proof and self-lubrication appli- [33] W.C. Tu, F.F. Lange, A.G. Evans, Concept for a damage-tolerant ceramic composite
cation of carbon nanotubes, Rev. Adv. Mater. Sci. 36 (2014) 75–88. with ‘strong’ interfaces, J. Am. Ceram. Soc. 79 (20) (1996) 417–424.
[5] G.B. Choon, C.I. Sik, Microstructural analysis and wear performance of carbon fiber [34] P.R. Rios, F.S. Jr, H.R.Z. Sandim, R.L. Plaut, A.F. Padilha, Nucleation and growth
reinforced SiC composite for brake pads, Materials 701 (10) (2017) 1–14. during recrystallization, Mater. Res. 8 (3) (2005) 225–238.
[6] P. Kun, O. Tapaszto, F. Weber, C. Balazsi, Determination of structural and me- [35] Y. Wang, S.M. Hsu, Wear and wear transition mechanisms of ceramics, Wear 195
chanical properties of multilayer graphene added silicon nitride-based composites, (1–2) (1996) 112–122.
Ceram. Int. 38 (1) (2012) 211–216. [36] J. Krim, Friction and energy dissipation mechanisms in adsorbed molecules and
[7] I. Kholmanov, J. Kim, E. Ou, R.S. Ruoff, L. Shi, Continuous carbon nanotube-ul- molecularly thin films, Adv. Phys. 61 (3) (2012) 155–323.
trathin graphite hybrid foams for increased thermal conductivity and suppressed [37] S. Jahanmir, N.P. Suh, E.P. Abrahamson, The delamination theory of wear and the
subcooling in composite phase change materials, ACS Nano 9 (12) (2015) wear of a composite surface, Wear 32 (1975) 33–49.
11699–11707. [38] B. Shivamurthy, K. Murthy, S. Anandhan, Tribology and mechanical properties of
[8] J.F. Palacio, I.A. Garcia, S.J. Bull, Novel carbons in tribology, Tribol. Int. 37 (11–12) carbon fabric/MWCNT/epoxy composites, Adv. Tribol. 2018 (1508145) (2018)
(2004) 929–940. 1–10.
[9] P. Wu, X. Li, C. Zhang, X. Chen, S. Lin, H. Sun, C.T. Lin, H. Zhu, J. Luo, Self- [39] K.G. Thirugnanasambantham, A.G.G. Kumar, Mechanistic studies on degradation in
assembled graphene film as low friction solid lubricant in macroscale contact, ACS sliding wear behavior of carburized AISI 8620 steel at 100 °C under unlubricated
Appl. Mater. Interfaces 9 (25) (2017) 21554–21562. conditions, Mater. Today Proc. 5 (2) (2018) 6258–6267.
[10] Z.H. Xia, J. Lou, W.A. Curtin, A multiscale experiment on the tribological behavior [40] L. Vilhena, M. Sedlacek, B. Podgornik, Z. Rek, I. Zun, CFD modeling of the effect of
of aligned carbon nanotube/ceramic composites, Scr. Mater. 58 (3) (2008) different surface texturing geometries on the frictional behavior, Lubricants 6 (15)
223–226. (2018) 1–25.
[11] Y. Huang, Q. Yao, Y. Qi, Y. Cheng, H. Wang, Q. Li, Y. Meng, Wear evolution of [41] N.L. McCook, D.L. Burris, N.H. Kim, W.G. Sawyer, Cumulative damage modeling of
monolayer graphene at the macroscale, Carbon 115 (2017) 600–607. solid lubricant coatings that experience wear and interfacial fatigue, Wear 262
[12] Z. Baig, O. Mamat, M. Mustapha, Recent progress on the dispersion and the (2007) 1490–1495.
strengthening effect of carbon nanotubes and graphene-reinforced metal nano- [42] D. Berman, A. Erdemir, A.V. Sumant, Graphene: a new emerging lubricant, Mater.
composites: a review, Crit. Rev. Sol. State Mater. Sci. 43 (1) (2018) 1–46. Today 17 (1) (2014) 31–42.
[13] L. Zhang, J. Pu, L. Wang, Q. Xue, Frictional dependence of graphene and carbon [43] D. Galusek, D. Galuskova, Alumina matrix composites with non-oxide nanoparticle
nanotube in diamond-like carbon/ionic liquids hybrid films in vacuum, Carbon 80 addition and enhanced functionalities, Nanomaterials 5 (1) (2015) 115–143.
(2014) 734–745. [44] K.J. Kubiak, T.W. Liskiewicz, T.G. Mathia, Surface morphology in engineering ap-
[14] M. Belmonte, J. Gonzalez-Julian, P. Miranzo, M.I. Osendi, Spark plasma sintering: a plications: influence of roughness on sliding and wear in dry fretting, Tribol. Int. 44
powerful tool to develop new silicon nitride-based materials, J. Eur. Ceram. Soc. 30 (2011) 1427–1432.
(2010) 2937–2946. [45] S.N. Alam, N. Sharma, B.C. Ray, S. Yadav, K. Biswas, Effect of graphite nanopla-
[15] A. Nieto, A. Bisht, D. Lahiri, C. Zhang, A. Agarwal, Graphene reinforced metal and telets on the mechanical properties of alumina-based composites, Ceram. Int. 43
ceramic matrix composites: a review, Int. Mater. Rev. 62 (5) (2017) 241–302. (14) (2017) 11376–11389.
[16] I. Ahmad, A. Kennedy, Y.Q. Zhu, Wear resistant properties of multi-walled carbon [46] G.D. Zhan, J.D. Kuntz, J. Wan, A.K. Mukherjee, Single-wall carbon nanotubes as
nanotubes reinforced Al2O3 nanocomposites, Wear 269 (1–2) (2010) 71–78. attractive toughening agents in alumina-based nanocomposites, Nat. Mater. 2 (1)
[17] H.J. Kim, S.M. Lee, Y.S. Oh, Y.H. Yang, Y.S. Lim, D.H. Yoon, C. Lee, J.Y. Kim, (2003) 38–42.
R.S. Ruoff, Unoxidized graphene/alumina nanocomposite: fracture- and wear-re- [47] K. Fukaura, Y. Yokoyama, D. Yokoi, N. Tsujii, K. Ono, Fatigue of cold-work tool
sistance effects of graphene on alumina matrix, Sci. Rep. 4 (05176) (2015) 1–10. steels: effect of heat treatment and carbide morphology on fatigue crack formation,
[18] C.F.G. Gonzalez, A. Smirnov, A. Centeno, A. Fernandez, B. Alonso, V.G. Rocha, life, and fracture surface observations, Met. Mater. Trans. A 35 (4) (2004)
R. Torrecillas, A. Zurutuza, J.F. Bartolome, Wear behavior of graphene/alumina 1289–1300.
composite, Ceram. Int. 41 (6) (2015) 7434–7438. [48] Y.F. Chen, J.Q. Bi, C.L. Yin, G.L. You, Microstructure and fracture toughness of
[19] K. Balani, S.P. Harimkar, A. Keshri, Y. Chen, N.B. Dahotre, A. Agarwal, Multiscale graphene nanosheets/alumina composites, Ceram. Int. 40 ((9)A) (2014)
wear of plasma-sprayed carbon-nanotube-reinforced aluminum oxide nanocompo- 13883–13889.
site coating, Acta Mater. 56 (2008) 5984–5994. [49] G.D. Zhang, R. Chen, Effect of the interfacial bonding strength on the mechanical
[20] A.K. Keshri, A. Agarwal, Wear behavior of plasma-sprayed carbon nanotube-re- properties of metal matrix composites, Comp. Interface 1 (4) (1993) 337–355.
inforced aluminum oxide coating in marine and high-temperature environments, J. [50] I.A. Ovidko, A.G. Sheinerman, Toughening due to crack deflection in ceramic- and
Therm. Spray Technol. 20 (6) (2011) 1217–1230. metal-graphene nanocomposites, Rev. Adv. Mater. Sci. 43 (2015) 52–60.
[21] B. Yazdani, F. Xu, I. Ahmad, X. Hou, Y. Xia, Y. Zhu, Tribological performance of [51] J.A. Arsecularatne, L.C. Zhang, Carbon nanotube reinforced ceramic composites
Graphene/Carbon nanotube hybrid reinforced Al2O3 composites, Sci. Rep. 5 and their performance, Rec. Pat. Nanotechnol. 1 (2007) 176–185.
(11579) (2015) 1–11. [52] K.T. Faber, A.G. Evans, Intergranular crack defection toughening in silicon carbide,
[22] N. Sharma, S.N. Alam, B.C. Ray, S. Yadav, K. Biswas, Silica-graphene nanoplatelets J. Am. Ceram. Soc. 66 (6) (1983) C-94–C-96.
and silica-MWCNT composites: microstructure and mechanical properties, Diam.
Relat. Mater. 87 (2018) 186–201.

304

You might also like