Dry Sliding Wear

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Metallurgical and Materials Engineering Research paper

Association of Metallurgical Engineers of Serbia AMES https://doi.org/10.30544/824

DRY SLIDING WEAR BEHAVIOR AND ITS RELATION TO


MICROSTRUCTURE OF ARTIFICIALLY AGED Al-Si-Mg/TiB2
IN SITU COMPOSITES

Nishith Rathod1*, Jyoti Menghani2


1
Department of Mechanical Engineering, C K Pithawala College of Engineering and
Tech., Surat, India.
2
Department of Mechanical Engineering, Sardar Vallabhbhai National Institute of
Technology, Surat, India.

Received 10.11.2021
Accepted 17.06.2022

Abstract
Mechanical stir casting is utilized to produce an artificially aged Al-Si-Mg alloy,
whereas halide slat (K2TiF6 and KBF4) synthesis is utilized to produce Al-Si-Mg/TiB2
aluminum matrix composites. The dry sliding pin on disc wear test was conducted using
a DUCOM/TR-20LE-PHM-200 machine to simulate an automobile application (Piston-
Ring assembly). Where pistons are made of aluminum alloy (for the Pin) and rings are
made of grey cast iron (for the disc material). At room temperature, a wear test was
conducted by altering the ageing time (3, 6, 9, 12), sliding speed (2, 2.5 m/s), and applied
load (14.71, 19.62, 24.52 N) with the disc speed (500 rpm) held constant (10 min). The
results indicate that the aluminum matrix composite (AMC) wear rate is reduced by 37
percent at higher sliding speeds (2.5 m/s) and by 4 percent at lower sliding speeds (2.0
m/s) compared to the base alloy. Field emission scanning electron microscope-energy
dispersive spectroscopy (FESEM-EDS) and X-ray photoelectron spectroscopy (XPS)
analysis revealed that the formation of the mechanically mixed layer (MML) or oxidative
layers on the worn surfaces reduces the wear rate under conditions of longer ageing time,
higher sliding speed, and lower applied load. The research demonstrates that composite
wear is a function of sliding velocity, aging period, and applied force. As sliding speed
rose from 2 m/s to 2.5 m/s, the wear rate of composites dropped reasonably, yet
composites are softer than basic alloys. It is conceivable due to the presence of a
considerable amount of MML and the formation of oxidative layers between pins and
their equivalents.

Keywords: TiB2; artificially aged; dry sliding wear; mild wear; MML.

* Corresponding Author: Nishith Rathod, [email protected]


270 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

1. Introduction
The huge growth in demand for fuel-efficient and lightweight materials in the
automotive and aviation industries has led to a surge in the development and design of
new, custom-made materials. Due to their remarkable corrosion resistance, high ductility,
high strength-to-weight ratio, and low cost, Al-Si-Mg master alloys are one of the most
widely used alloys in the vehicle industry. These materials, however, have poor
tribological properties [1]. Therefore, alloy modification and improved mechanical and
tribological responses are required, which will eventually lead to the development of
aluminum metal matrix composites (AMC). Metal matrix composites (MMC) exhibit
exceptional rigidity, abrasion resistance, and strength enhancement. Aluminum metal
matrix composites have numerous applications in the engineering, marine, automotive,
and aerospace sectors. Oxides [2], borides [3], nitrides [4], and carbides [3] are the most
frequently used reinforcements for application-specific MMC.
In-situ or ex-situ particle reinforcement could significantly improve the
tribological properties of the Al-Si system [5]. Conventional composite fabrication is
accomplished by reinforcing hard particles in pure metals and/or metal alloys, termed ex-
situ MMCs. SiC, TiC, TiB2, ZrB2, B4C, and Al2O3 are some of the most prevalent
reinforcing elements that, in addition to enhancing wear resistance and mechanical
properties, also serve as heterogeneous nucleation sites, resulting in grain refinement.
These reinforcing particles are lightweight and possess high hardness and melting points.
They function as dispersion strengthening agents, impede dislocation motion, and
increase the strength of AMC. These AMCs are ideal for aircraft and automotive
components such as pistons, gears, cylinder liners, clutches, and engine block housing
[6]. TiB2 is one of the most widely used reinforcements due to its exceptional mechanical
and physical properties, including its exceptional hardness, high melting point, high
elastic modulus, and exceptional wear resistance. TiB2 particles added to an Al matrix
reportedly stimulate heterogeneous nucleation of α-Al and have a favorable
crystallographic orientation relationship that results in high coherency and low solid
particle interfacial energy [7]. An approach was to develop reinforcement particulates via
chemical reaction within a liquid or solid at a higher temperature, known as the salt
synthesis route (in situ route). The in situ route has significant advantages against the ex-
situ route like outstanding bonding strength, clean matrix-reinforcement interface,
thermodynamically stable and fine grain structure, good wettability, and minimum
processing cost. A significant parameter that influences the properties of AMC is the
chemical reaction between the reactants of the solution and the holding time of response.
TiB2, TiC, and Al2O3 are the most common reinforcements fabricated by the salt synthesis
route because they are not reacting with the matrix [8].
The researchers devoted a substantial amount of effort to the Al-Si-Mg system
strengthened with SiC particles [5]. However, few studies on wear of AMCs reinforced
via salt synthesis route (in-situ route) with TiB2 particles have been published [2, 3].
Rathod N. R., & Menghani J. (2019) and P. Samal et al. (2020) advocated the
advantageous effect of in-situ TiB2 particles [2, 3]. In addition, no aluminum-based
intermetallics or reaction products are formed near the interface of the matrix and
reinforcement [9, 10]. Moreover, Al/TiB2 AMCs prepared by the salt synthesis route have
recently been cited with improved mechanical and tribological properties [11].
Precipitation hardening is an additional technique for fortifying Al-Si-Mg alloy. In
recent years, researchers have devoted a substantial amount of time and energy to
N. Rathod et al. - Dry sliding wear behavior and its relation to … 271

studying the effects of precipitation hardening on aluminum-based systems. The


precipitation-hardening process is attained by heating a supersaturated solid solution at
elevated temperature followed by quenching in different media and natural ageing
processes (ageing occurs at room temperature) or artificial ageing process (ageing takes
place at high temperature). The solution heat treatment followed by aging is crucial for
increasing the mechanical properties and durability of metallic components [12]. The
precipitation sequence in the Al-Mg-Si system is represented as follows;

Clusters of Si atoms GP-I zones (primary monoclinic) GP-II- zones (β")


(needle shape) β' (hexagonal) β' (Mg 2Si)

First, the decomposition of the system starts and passes through different unstable
phases (called GP zones), and in the last stage, the system ultimately losses coherency,
and the equilibrium β (Mg2Si) phase is formed [13]. The storage time (aging time),
temperature, and alloy composition are crucial factors that influence the performance of
aged aluminium-based systems [14]. The peak aged AMC demonstrates the greatest
resistance to wear under both over-aged and under-aged conditions [15].
The wear behavior of AMC is determined by the type and size of reinforcements,
the microstructure of the system, and the wear governing parameters. This study examines
the wear of four types of samples: (a) base alloy (Al-Si-Mg system), (b) artificially aged
Al-Si-Mg system, (c) Al-Si-Mg/TiB2 in situ composites fabricated by salt synthesis route,
and (d) artificially aged Al-Si-Mg/TiB2 in situ composites fabricated by salt synthesis
route. Two steps of artificial ageing treatment; (a.) Solutionizing treatment (b.)
Artificially ageing (varying times 3, 6, 9, 12 hours) incorporates the test samples. ASTM
- G99 standard is used to perform the pin on disc wear test to record dry sliding wear
responses. The wear test used different sliding speeds (2, 2.5 m/s) and normal applied
load (14.71, 19.62, 24.52 N). To incorporate wear response and hardness of both
categorical samples (base alloy and composites) BHN hardness test is performed as per
ASTM E10-14 standards. The microstructure and surface morphology of worn surfaces
incorporate by X-ray diffraction (XRD), Field emission scanning electron microscope-
energy dispersive spectroscopy (FESEM-EDS) and X-ray photoelectron spectroscopy
(XPS) analysis.

2. Materials and methods


2.1 Processing of in situ composites
The commercial Al-7Si-0.3Mg master alloy was heated up to 850 ºC, and two
halide salts (KBF4 and K2TiF6) were added to the molten Al-7Si-0.3Mg system in the
atomic ratio of Ti:2B by the mechanical stirring method. A ceramic stirrer stirs the melt.
The exothermic reaction between the two halide salts and the molten Al-7Si-0.3Mg
system resulted in situ TiB2 particles in the Al-7Si-0.3Mg system. The reaction holding
time and the exothermic reaction temperature are 45 minutes and 850 ºC, respectively.
Fig. 1, flow process chart represents fabrication of in situ composites via salt synthesis
route [16].
272 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

Fig. 1. Flow process chart of the in situ process used for synthesizing Al-7Si-
0.3Mg/TiB2 composite.

2.2 Heat treatments


The Al-Si-Mg system with and without TiB2 was artificially aged. Solutionizing
heat treatment was attempted at 501 ºC for 1 h [17], followed by water quenching and
single-stage artificial ageing at 175 °C for different ageing times (3, 6, 9, 12 h).
N. Rathod et al. - Dry sliding wear behavior and its relation to … 273

Motor

Reactive
Support salt feeder

Electric
Furnace
Mechanical
stirrer
Graphite
Crucible

Bottom Control
Pouring Panel

Fig. 2. Schematic diagram of bottom pouring mechanical stir casting furnace.

2.3 Pin on disc wear test


DUCOM/TR-20LE-PHM-200 machine incorporate to perform the dry sliding pin
on disc wear test. The fabrication of test sample is as per ASTM - G99, cylindrical pin
(30 mm height and 10 mm diameter) of the Al-Si-Mg system (with and without ageing
treatment and with and without TiB2). This paper aims to replicate the piston ring
assembly application, where the fabrication process of a commercial piston from the Al-
Si-Mg system and its AMC (Al-Si-Mg/TiB2) and the disc from grey cast iron. In context,
an artificially aged Al-Si-Mg master alloy and AMC (Al-Si-Mg/TiB2) are used to
fabricate cylindrical pin and grey cast iron (Brinell HB 224) is used to manufacture the
counterpart (the disc). Pin on disc wear test is carried out at different sliding speeds (2,
2.5 m/s) and normal applied load (14.71, 19.62, 24.52 N). The counterpart (disc) rotation
speed and time, 500 rpm and 10 minutes, remain constant for all samples. In the absence
of lubrication conditions, an ambient environment is used for all wear tests. FESEM-EDS
and XPS studies of the Al-Si-Mg system and Al-Si-Mg/TiB2 AMCs carried out for worn
samples. Weight loss assessment technique used for calculating the amount of wear rate
for given samples.

Wear rate, (mm3/m) = Volume loss/ sliding distance 1

Volume loss, (mm3) = Mass loss/ density 2


Where density in g/cm3.
274 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

Load
(a) (b)
Pin

Disc

(c)

Fig. 3. (a) Pin on Disc apparatus, (b) Data acquisition system and (c) Disc (made of
gray cast-iron).

Result and discussion

3.1 Formation and examination of TiB2 in situ composites


The following chemical reaction represents the formation of TiB2 particles through
in situ route.
K2TiF6 (1) + KBF4 (1) + Al (1) TiB2 (in alloy) + KFAlF3 (1) + KF (1) 3
+AlF3 (1)
To confirm the presence of TiB2, a Panalytical Xpert X-ray diffraction machine
was used to conduct an XRD test. Fig. 4 depicts the XRD (X-ray diffraction) analysis
plot. Along with Al and Si, the corresponding peaks of TiB 2, AlB2, and Al3Ti are also
reported. At high temperatures (850 °C), both halide salts (KBF4 and K2TiF6) decompose,
and the decomposed elements K and F form slag on the surface of the aluminum melt.
Given a molten state, the ternary system Al-Ti-B is theorized to exist. The research
revealed that the formation of TiB2 occurs at 1023 K (750 °C). However, as the reaction
temperature rises, the Gibbs' free energy of TiB2 rises and the system's TiB2 formation
rate increases.
N. Rathod et al. - Dry sliding wear behavior and its relation to … 275

*
* Al
∆ TiB2
Intensity
(counts)

¤ Si
# AlB2
∆*
⁍ Al3Ti


⁍ *
*
# #⁍ ¤ ∆ ∆ ∆ * ∆
10 20 30 40 50 60 70 80 90
2θ, (degree)
Fig. 4. XRD plot of Al-Si-Mg/TiB2 in situ composites fabricated by salt synthesis route.
Theoretically, as the reaction temperature rises, the decrease in Gibbs free energy
of Al3Ti decreases, indicating that Al3Ti is unstable [18]. Consequently, TiB2 increased,
Al3Ti decreased, and more Ti was converted to the TiB 2 phase [19].
Figure 5 illustrates the FESEM-EDS image and its mapping of composite (Al-Si-
Mg/TiB2) in order to comprehend and validate XRD peaks. In Fig. 4, the intensity of the
AlB2 and Al3Ti peaks is weak. At some point, the Al3Ti peak overlaps the Al peak.
FESEM-EDS mapping verifies whether or not secondary phases AlB2 and Al3Ti formed
during the salt reaction. Nonetheless, this observation does not indicate any secondary
phase. This analysis makes it simple to exclude elements of the secondary phase and their
detrimental effect on the material to exempt the secondary phase elements and its
detrimental effect on the material.
276 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

TiB2 Cluster

Fig. 5. FESEM-EDS image and its mapping of composites (Al-Si-Mg/TiB2).

3.2 Quantification of BHN (hardness)

The ASTM E10-14 standard is utilized for the BHN hardness test. It measures the
average value of five indentation centers to surface under a 500kgf load with a 10 mm
indenter.
N. Rathod et al. - Dry sliding wear behavior and its relation to … 277

98
96
94
92
BHN

90
88
86
84
Al-Si-Mg Al-Si-Mg/TiB2
82
80
0 3 6 9 12 15
Aging time, h

Fig. 6. Effect of aging time on BHN, base alloy (Al-Si-Mg) and Composite (Al-Si-
Mg/TiB2).
The BHN test revealed that the composites and base alloys have identical hardness
under non-aging conditions. However, as ageing time increases, composites become
softer than base alloy in the same condition up to 9 h ageing time, but thereafter, grain
refinement during 12 h ageing improves hardness. The particle size of TiB 2 is crucial to
the hardness of composites. According to reports, the grain growth of TiB 2 particles is
affected by reaction holding time. Up to 30 minutes of reaction holding time reveals the
optimal particle size of TiB2 via the salt synthesis route. However, grain size increased as
reaction holding time increased beyond 40 minutes [3]. However, after 40 minutes, the
brittle secondary element Al3Ti is completely dissolved and converted into TiB2 particles
(In this study, the reaction holding time is 45 minutes).

3.3 Quantification of wear rate of base alloy and composites


3.3.1 Effect of sliding speed, aging time and normal applied load on wear
behaviour
The wear rate of base alloy (Al-Si-Mg - A1) and AMC (Al-Si-Mg/TiB2 -A2) as a
function of ageing time (3, 6, 9, 12 h) for sliding speed and normal applied load is
illustrated in Figure 7. This study examined sliding speeds of 2, 2.5 m/s and normal
applied loads of 14,71, 19,62, and 24,52 N. As shown in Fig. 7 (a), (b), and (c), as sliding
speed increased from 2 to 2.5 m/s, wear rate decreased at 14.71 N and 24.52 N normal
load as ageing time increased. With the exception of 19.62 N, the wear rate at 2 m/s
sliding speed is less than 2.5 m/s. Al-Si-Mg/TiB2 AMC, on the other hand, exhibits a
steady decrease in wear as ageing time increases for normal loads of 14.72 N and 19.62
N (Fig. 7(d) and (e)). There is an initial decrease followed by an increase in wear for
24.52 N. At longer ages, the rate of deterioration is observed to decrease steadily in all
278 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

instances. The sliding speed increased to 2.5 m/s and the amount of wear decreased
compared to 2 m/s for all loading conditions as a result of aging. According to the
discussion, it is evident that as ageing time increases, the achieved wear rate decreases at
higher sliding speeds.
(a) A1-2 m/s-14.71 N (b) A1-2 m/s-19.62 N
35 A1-2.5 m/s-14.71 N A1-2.5 m/s-19.62 N
35

Wear rate, 10-4 (mm3/m)


Wear rate, 10-4 (mm3/m)

30 30
25 25
20 20
15 15
10 10
5 5
0 0
0 3 6 9 12 0 3 6 9 12
Aging time, h Agign time, h
(c) A1-2 m/s-24.52 N (d)
A2-2 m/s-14.71 N
A1-2.5 m/s-24.52 N
35 35 A2-2.5 m/s-14.71 N
Wear rate, 10-4 (mm3/m)

Wear rate, 10-4 (mm3/m)

30 30
25 25
20 20
15 15
10 10
5 5
0 0
0 3 6 9 12 0 3 6 9 12
Agign time, h Agign time, h

(e) A2-2 m/s-19.62 N (f) A2-2 m/s-24.52 N


A2-2.5 m/s-19.62 N A2-2.5 m/s-24.52 N
Wear rate, 10-4 (mm3/m)

35 35
Wear rate, 10-4 (mm3/m)

30 30
25 25
20 20
15 15
10 10
5 5
0 0
0 3 6 9 12 0 3 6 9 12
Agign time, h Agign time, h
Fig. 7. (a), (b), (c), (d), (e), and (f) are the wear rate of base alloy (Al-Si-Mg – A1)
and Composites (Al-Si-Mg/TiB2 – A2) as a function of ageing time by varying sliding
speed
(2, 2.5 m/s) and normal applied load (14.71, 19.62, 24.52 N).

At 9 h ageing time, 2 m/s sliding speed, 14.71N load and 3 h ageing time, 2.5 m/s sliding
speed, 24.52N load, and the wear rate of the base alloy increases significantly in Figures
N. Rathod et al. - Dry sliding wear behavior and its relation to … 279

7 (a) and (c), respectively. BHN of wear samples is performed to comprehend the
phenomenon of uneven wear increment. According to reports, the hardness (average
value of five indentations at a different location) of respective samples decreases from 96
to 56 (9 h ageing time) and 89 to 47 (3 h ageing time) BHN after 9 h and 3 h of aging,
respectively. Consequently, a significant hardness reduction or work softening
phenomenon is responsible for this condition's increased wear rate.

3.3.2 Effect of composition on wear behaviour


Figure 8 compares the wear rates of the Al-Si-Mg system and the Al-Si-Mg/TiB2
AMC at sliding speeds of 2, 2.5 m/s and applied loads of 14.71, 19.62, and 24.52 N. Fig.
8 (a), (b), and (c) depict the wear behavior of base alloy and composites at 2 m/s as a
function of ageing time under a normal applied load of 14.71, 19.62, and 24.52 N. At
14.71 and 24.52 N load conditions, the wear rate of composites decreases with increasing
ageing time (greater than 9 hours) when sliding at 2 m/s. Nonetheless, in Fig. 7 (b) at a
normal load condition of 19.62 N, the wear rate of the base alloy is less than that of
composites. Figure 8 (d), (e), and (f) indicate that as the sliding speed increased by 2.5
m/s, the wear rate of composites with a longer ageing time (greater than 9 h) decreased
steadily compared to the base alloy.
(a) (b)
35 A1-2 m/s-14.71 N 35 A1-2 m/s-19.62 N
Wear rate, 10-4 (mm3/m)
Wear rate, 10-4 (mm3/m)

30 A2-2 m/s-14.71 N 30 A2-2 m/s-19.62 N


25 25
20 20
15 15
10 10
5 5
0 0
0 3 6 9 12 0 3 6 9 12
Agign time, h
Agign time, h
(c) A1-2 m/s-24.52 N (d) A1-2.5 m/s-14.71 N
A2-2 m/s-24.52 N 35 A2-2.5 m/s-14.71 N
35
Wear rate, 10-4 (mm3/m)
Wear rate, 10-4 (mm3/m)

30 30
25 25
20 20
15 15
10 10
5 5
0 0
0 3 6 9 12 0 3 6 9 12
Agign time, h
Aging time, h
280 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

(e) (f) A1-2.5 m/s-24.52 N


A1-2.5 m/s-19.62 N
35 A2-2.5 m/s-19.62 N
35 A2-2.5 m/s-24.52 N

Wear rate, 10-4 (mm3/m)


Wear rate, 10-4 (mm3/m)

30 30
25 25
20 20
15 15
10 10
5 5
0
0
0 3 6 9 12
Agign time, h 0 3 6 9 12
Agign time, h
Fig. 8. (a), (b), (c), (d), (e), and (f) are a comparison of base alloy (Al-Si-Mg – A1)
and Composites (Al-Si-Mg/TiB2 – A2) wear rate as a function of aging time by
varying sliding speed (2, 2.5 m/s) and normal applied load (14.71, 19.62, 24.52 N).
At the higher aging time (greater than 9 hours) reported in this discussion, wear
rate is a function of sliding speed. In most cases, as the sliding speed increased by 2.5
m/s, the wear rate decreased significantly with increased ageing time, with the exception
of the 19.62 N normal load condition in the base alloy.

3.4 XPS studies


X-ray photoelectron spectroscopy (XPS) explains the surface's stoichiometry and
chemical composition due to its exceptional high surface sensitivity, quantitative nature,
and element selectivity [20]. A lower wear rate in composites with a longer ageing time
and higher sliding velocity than the base alloy [21]. In order to comprehend the worn
surface morphology of Al-Si-Mg/TiB2 AMC, two representative samples were chosen at
different sliding speeds (2 and 2.5 m/s) and a longer ageing time (9 and 12 h). As the
valance state of an element increases, the respective element's binding energy also
increases. The binding energy of Fe 2p3/2 was about 709 EV for Fe+2 and 711 EV for Fe+3.
Fig. 9 (d) shows the XPS peaks of the Fe 2p regions and Fe 2p 3/2 at 709 EV and Fe 2p1/2
at 720 EV, indicating that Fe+2 and Fe+3 should be guaranteed, resulting in the
development of Fe3O4. The satellite peak at about 717 EV was a distinctive peak of
saturated Fe+3 in γ-Fe2O3 (maghemite) [21, 22]. It proposes that Fe3O4 nanoparticles were
partially oxides, and the increased intensity indicates that Fe +2 and Fe+3 predominate in
the system. When Fe content is transferred from the disc (counterpart) surface to the
composite pin, shared Fe atoms and the Fe-containing disc will react to form a
mechanically mixed layer on the composite pin surface, thereby reducing the system's
wear rate [23].
The Mg 2p peak was located at 54 EV and the Al 2p peak was located at 75 EV.
Fig. 9 (a) and (c) demonstrate that as the number of oxidizing layers increases, the number
of oxide peaks on the high binding energy side of the primary Mg and Al photoelectron
peaks increases significantly.
N. Rathod et al. - Dry sliding wear behavior and its relation to … 281

Fig. 9. XPS plots of composites at 9 h artificially aged, sliding speed 2 m/s, load
24.52 N (29-2-24.52) and 12 h artificially aged, sliding speed 2.5 m/s , 14.71 N load
(212-2-14.71) (a) Al 2p, (b) Si 2p, (c) Mg 2p, (d) Fe 2p, (e) C 1s, (f) O 1s.
The XPS results detected the peaks allotted to SiO2, TiO2, and Al2O3. The existence
of Si-O implies that Al-Mg-Si has been decomposed and formed oxides during the
abrasion process [10]. The XPS results that tribo film formation on the worm surface of
Al/TiB2 mainly due to Al-Mg-Si oxides which reduces the contact of two counterparts
282 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

and serves low shear strength junction at the interface [24]. That significantly reduces
ploughing between two counterparts, thus reducing the wear rate of the system. That must
be responsible for the lower wear rate at higher speed (2.5 m/s), higher load (24.52 N),
and ageing time at 9 and 12 h, respectively. Thus the variation in surface composition of
worn samples resulted in a change in the wear pattern of pieces. The wear samples'
behaviour was incorporated by FESEM-EDS mapping.

3.5 FESEM analysis of worn surfaces


Fig. 10, 11, 12, 13 shows the worn surfaces of AMCs and base alloy at the different
ageing times, sliding speed, and applied load. Al-Si-Mg alloy's dry sliding wear rate was
reasonably higher than Al-Si-Mg/TiB2 AMC at ambient temperature in most of the
conditions. The overall decrease in AMC's wear rate at lower sliding speed (2.5 m/s) was
37%, and the higher sliding speed (2 m/s) was 4% than the base alloy reported from the
results.

Fig. 10. FESEM images of worn surfaces of the base alloy at 3 h aging time, 2.5 m/s
sliding speed and 24.52 N applied load. (a) at lower magnification (b) at higher
magnification.
Characteristics of worn surface morphology of AMCs are more complex than those
of metals or alloys due to the influence of two major factors: extrinsic factors (normally
applied load, sliding speed/distance, temperature, surface finish and hardness of
counterpart, nominal contact area) and intrinsic factors (reinforcement size and shape,
types, volume/weight fraction, interfacial bonding, porosity, wettability). Therefore, a
comprehensive understanding of the wear mechanism is relatively difficult. FESEM-EDS
analysis of worn surfaces facilitates the comprehension of AMCs' wear mechanism. Base
alloy at 3 h ageing time, 2.5 m/s sliding speed, normal applied load 24.52 N (29.739 x 10 -
4
mm3/m), AMCs at 9 h ageing, 2 and 2.5 m/s sliding speed, applied load 24.52 N (9.987
x 10-4, 14. 231 x 10-4 mm3/m respectively) and 12 h ageing, 2.5 m/s, applied load 14.71
N (3.245 x 10-4 mm3/m) selected to understand sever, and mild wear mechanism.

3.5.1 Effect of normal applied load


Low and high magnification FESEM micrograph of Al-Si-Mg system at 3 h ageing
time, 2.5 m/s sliding speed, and 24.52 N applied load is shown in Fig. 10 (a.) (b). Figure
10 (a) depicts parallel soft ridges and grooves running in the sliding direction; under
conditions of high sliding speed and high applied load, there were few delamination
N. Rathod et al. - Dry sliding wear behavior and its relation to … 283

zones. The wear rate of particulate MMCs under varying conditions of normal applied
load revealed three wear regimes. In Regime I, under low load conditions, reinforcement
particles support the load for which MMCs have a lower wear rate than aluminum alloys.
In Regime II, the aluminum alloy and MMC wear rates are identical. Regime III is
characterized by the transformation of mild wear to severe wear under high load
conditions as the surface temperature exceeds a critical value [3]. As the load applied
increased, the wear resistance decreased, and the wear mechanism revealed oxidative
layer formation under low load conditions and delamination and adhesion under high load
conditions [25]. This conditioning system enables the formation of more stable oxidative
layers on the surfaces because the pin on the disc was subjected to wear testing at ambient
air temperature [26].
Due to the formation of a sound oxide film on the pin, the amount of abrasion that
forms on its surfaces is restricted. Similarly, Fig. 10 (a) depicts deeper observed grooves
and ridges in the context of a reported high wear rate. Due to low sliding speed and in-
situ TiB2 particles, AMCs exhibited gross plastic deformation. At a low sliding speed,
flakier-shaped particles are ploughed and delamination zones appear, resulting in a higher
wear rate.

3.5.2 Effect of sliding speed


Al-Si-Mg and Al-Si-Mg/TiB2 AMC appear to have relatively high deformation
and exposed cracks in the given zone under higher magnification and slower sliding speed
(Figs. 10, 11). However, at high sliding speeds and long aging times, AMCs appear to
have more minor cracks than base alloy (Fig. 11, 12). AMCs have a low wear rate at
higher sliding speeds. It is a result of the formation of a dense oxidative transfer layer on
the pin surfaces. In addition, Al-Si-Mg and Al-Si-Mg/TiB2 composites with a slower
siding speed show extreme wear [27]. At a slower sliding speed, Fig. 10 (d) depicts a
reduced oxide and Fe transfer. Consequently, at lower sliding speeds, oxidative layers
were not stable due to insufficient Fe content transfer from the counterpart to the pin; this
circumstance does not permit the formation of sound inter-reactive components. This is
responsible for a significant amount of delamination, loose oxide layers, and increased
stresses in the disc and pin's contact region.

3.5.3 Effect of ageing time


However, Al-Si-Mg/TiB2 AMC exhibits a smooth surface morphology at a longer
ageing time and higher sliding velocity (Fig. 12, 13). On the surface of specimens,
plowing and material flow cause the formation of small globular-shaped particles, small
pits, and soft ridges. These phenomena indicate that abrasive wear has occurred under
conditions of high sliding speed and long ageing [28]. Transformation of mild wear to
severe wear depends on sliding speed and ageing time, but also on the combined effect
of other parameters such as particle size and shape, volume/weight fraction of
reinforcements, temperature, and formation of new phases at interfaces.
284 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

Fig. 11. (a), (b), (c) FESEM images of Al-Si-Mg/TiB2 composites at 9 h ageing time, 2
m/s sliding speed and 24.52 N, (d) EDS mapping of Al-Si-Mg/TiB2 composites at 9 h
ageing time, 2 m/s sliding speed and 24.52 N.

3.5.4 Effect of hardness


The response of the base alloy and the hardness of AMC is depicted in Figure 6.
The test samples' hardness is a function of their aging time. In comparison to AMC, the
hardness of the base alloy increases as ageing time increases. However, increasing
hardness diminishes the rate of wear. In addition, experimental evidence suggests that the
adhesive process slowed the scuffing and seizing of mating counterparts with greater
hardness [10]. Fig. 6 depicts the effect of composition on wear behaviour; under lower
sliding speed conditions, base alloys exhibit a lower wear rate than composites up to 6 h
of ageing time. Moreover, as ageing time and sliding speed increase, the base alloy's wear
rate increases in contrast to composites. In addition, XPS results confirm that at higher
sliding speeds (2.5 m/s) and longer aging times (12 h), the amount of MML (mechanically
mixed layer) significantly increases between counterparts and reduces wear intensity.

3.6.4 Effect of MML (mechanically mixed layer)


Mandal A. et al. (2009) investigated the wear mechanism and achieved wear of Al-
Si-Mg/TiB2 in situ composites through the following steps [26];
N. Rathod et al. - Dry sliding wear behavior and its relation to … 285

Delamination ⟶ Disintegration ⟶ Simultaneous interaction ⟶


⟶ Formation of complex (Al − Si − Fe − Ti) phase (MML)

Woydt M. (2000), Huang Z. et al. (2007), and Xu Z. et al. (2014) reported that
disintegration and simultaneous interaction result in the formation of a complex oxidative
Al-Si-Fe-Ti phase (MML). At low and high loads, mild and severe wear mechanisms
were reported, respectively. The development of MML at higher sliding speeds and
longer aging times restrains material loss by depressing the contact surfaces of the disc
and the pin, which is susceptible to mild wear.

Fig. 12. (a), (b), (c) FESEM images of Al-Si-Mg/TiB2 composites at 9 h ageing time, 2.5
m/s sliding speed and 24.52 N, (d) EDS mapping of Al-Si-Mg/TiB2 composites at 9 h
ageing time, 2.5 m/s sliding speed and 24.52 N.

Al-Mg-Si oxides reduce contact between two counterparts and serve as a low shear
strength junction at the interface, according to an XPS analysis of two composites
subjected to a higher load and a higher sliding speed over a period of 12 hours. This
significantly reduces the amount of plowing between the two halves, thereby decreasing
the system's wear rate.
Fig. 9 reported that as ageing time increased peak of oxides or MML on the
surfaces significantly increased. It was a reasonably reduced wear rate.
286 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

Fig. 13. (a), (b), (c) FESEM images of Al-Si-Mg/TiB2 composites at 12 h ageing time,
2.5 m/s sliding speed and 14.71 N, (d) EDS mapping of Al-Si-Mg/TiB2 composites at 12
h ageing time, 2.5 m/s sliding speed and 14.71 N.
During dry sliding, the hard ceramic (TiB2) particles easily abraded graphite flakes
and Fe particles from the disc surface, resulting in the oxidation of the pin surface and the
formation of an even thin layer of Al2O3, TiO2 [22, 23], SiO2, Fe2O3, Fe3O4 [21, 22] base
compound on the pin surface. The transfer rate of graphite flakes and Fe increased with
the dispersion AMCs of TiB2 particles; extended aging time and high sliding velocity-
based AMCs were observed in this type of phenomenon [29, 30].
Particle formation and reinforcement dispersion in the system substantially
improved the dimensional stability of AMCs. Peak aging conditions altered the
morphology of α-Al and Si. As α-Al dendrites become smaller or modified, the peak
strain and hardening rate for similar alloys increase dramatically [31]. Consequently,
ageing and TiB2 enhance work hardening and wear resistance at higher loads and speeds.
In addition, the formation of iron-rich oxides or MML reduced the rate of wear in AMCs
[32].
In addition, the coefficient of thermal expansion of TiB2 (7 x 10-6 K-1) is three-time
lesser than aluminium (24 x 10-6 K-1), it is leading to a considerable amount of strain at
the matrix-particle interfaces, and that is relieved by the development of dislocations [15].
This is one reason to increase the dislocation density in the AMCs, which significantly
boosts the strength and aging kinetics. Simultaneously increasing load and sliding speed
increases the formation of MML in the system. As a portion of the particle formation
N. Rathod et al. - Dry sliding wear behavior and its relation to … 287

increases, the uniform distribution of reinforcements and formation of MML improves


wear resistance by reducing the disc and pin's contact surfaces [33].

Fig. 14. EDS mapping of Al-Si-Mg/TiB2 composites at 9 h ageing time, 2.5 m/s sliding
speed and 24.52 N.
According to the results, the transfer layer (MML) acted as a protective layer on
the pin surface and reduced the wear rate. Transitions in sliding distance, load, and sliding
speed induce mild and severe wear. TiB2 serves as a wear-resistant load-carrying material
at higher loads and sliding speeds because MML is compromised [34, 35]. The wear rate
decreases as the ageing time, sliding speed, and applied load conditions increase.
Literature reports that the base alloy attained primary hardening phases after 3 to 4 hours
of aging. After 6-12 hours of aging, a few more precipitate phases developed in AMCs,
which dominated the significant increase in hardness and were responsible for the lower
wear rate in composites with longer aging times [34]. The EDS analysis revealed a dense
layer of MML on the worn surfaces of composites with longer aging times and faster
sliding speeds.
288 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

Fig. 15. EDS mapping of Al-Si-Mg/TiB2 composites at 12 h ageing time, 2.5 m/s sliding
speed and 14.71 N.
Variable process parameters can be used to attribute fluctuations in the wear rate
of composites compared to the base alloy to a complex reaction system during dry sliding
conditions. This leads to a comprehensive examination and study of the microstructure
formed during the wear process. However, the following explains the Al-Si-Mg/TiB2
AMCs' wear mechanism.
Table 1. Mode of wear and wear mechanism.
Mode of wear Mild wear Severe wear
FESEM
images of
worn surfaces

Wear  Abrasive wear and  The metal flow was


mechanism MML have drawn plough and responsible for gross plastic
and condition groove, mild delamination with deformation, and as a result, it
MML on worn surfaces. removed MML layers from the
 As discussed above, worn surfaces.
mild wear appears due to low  Severe wear appears due
applied load, high sliding speed, to high applied load, low sliding
and high ageing time. speed, and low ageing time, as
discussed above.
N. Rathod et al. - Dry sliding wear behavior and its relation to … 289

Conclusion
An artificially aged Al-Si-Mg alloy and Al-Si-Mg/TiB2 AMCs selected for the wear
studies by varying sliding speed (2, 2.5 m/s) and normal applied load (14.71, 19.62, 24.52
N). The investigation draws the following conclusions.
a.) XRD results reported forming in situ TiB2 particles in the alloy system and
forming Al3Ti and AlB4 intermetallics. Micrographs results show uniform particle
distribution of TiB2 in the system.
b.) The overall decrease in AMCs wear rate at higher sliding speed (2.5 m/s) was
37%, and lower sliding speed (2 m/s) was 4% than the base alloy reported from the
results. Work hardening due to in situ ceramic (TiB 2) reinforcement and ageing
substantially decreases wear rate. However, in base alloy at some point work
softening is reported (Fig. 7 (a), and (c)).
c.) The FESEM fractography of worn surfaces of AMCs reported mild ridges,
grooves and ploughed at mild wear (2.5 m/s, 14.71 N) and severe wear (2 m/s and
24.52 N) observed gross plastic deformation due to high metal flow.
d.) Combined FESEM-EDS and XPS analysis reveal that the formation of MML or
oxidative layers on the worn surfaces reduces the wear rate at the higher ageing time,
high sliding speed, and low applied load condition.
e.) However, hardness of base alloy is higher than composites but reasonable wear
rate is decreased due to ceramic reinforcements (TiB2), aging and significant amount
of MML and oxidative layers formation and it is increased with sliding speed
increased.
f.) The wear rate of composites is a function of ageing time, sliding speed, and
applied load.
At the higher ageing time (12 h), high sliding speed (2.5 m/s) and low applied load
(14.71 N) attributed to the lowest wear rate.

Acknowledgements
The author appreciatively acknowledges the support of MRC, MNIT, Jaipur,
ACMC Department of IIT-Kanpur, and CIF Department, IIT Gandhinagar for
metallurgical characterization and TEQIP sponsored High-Pressure High-Temperature
Press program at S V National Institute of Technology.

References:
[1] J. Xu, W. Liu: Wear, 260(4-5) (2006) 486-492.
[2] N. R. Rathod, J. Menghani: Metall Mat Eng, 25 (3) (2019) 195-208.
[3] P. Samal, P. R. Vundavilli, A. Meher, & M. M. Mahapatra: J Manu Pro 59 (2021)
131-152.
[4] R. K. Arya, & A. Telang: Int J Eng Adv Technol, 9 (2020) 3366-3374.
[5] A. D. Sarkar, J. Clarke: Wear, 75(1) (1982) 71-85.
[6] S. L. Pramod, S. R. Bakshi, B. S. Murty: J Mat Eng Perform, 24 (6) (2015) 2185-
2207.
[7] F. Chen, F. Mao, Z. Chen, J. Han, G. Yan, T. Wang, Z. Cao: J Alloy Compd, 622
(2015) 831-836.
[8] B. M. Viswanatha, M. P. Kumar, S. Basavarajappa, T. S. Kiran: Tribol Ind, 36 (1)
(2014) 40-48.
290 Metall. Mater. Eng. Vol 28 (2) 2022 p. 269-290

[9] C. A. Caracostas, W. A. Chiou, M. E. Fine, H. S. Cheng: Metall Mat Trans A, 28


(2) (1997) 491-502.
[10] C. S. Ramesh, A. Ahamed: Wear, 271(9-10) (2011) 1928-1939.
[11] S. M. Ma, P. Zhang, G. Ji, Z. Chen, G. A. Sun, S. Y. Zhong, V. Ji, H. W. Wang: J
Alloys Compd, 616 (2014) 128-136.
[12] H. Möller, G. Govender, W. E. Stumpf: Int J Cast Met Res, 20(6) (2007) 340-346.
[13] G. A. Edwards, K. Stiller, G. L. Dunlop: M. J. Couper: Act Mat, 46 (11) (1998)
3893-3904.
[14] M. Werinos, H. Antrekowitsch, T. Ebner, R. Prillhofer, P. J. Uggowitzer, S.
Pogatscher: Mat Des 107 (2016) 257-268.
[15] K. Pavitra, R. Mitra: Mat Sci Eng A, (2012) 557 84-91.
[16] S. Lakshmi, L. Lu, M. Gupta: J Mat Process Tech, 73(1-3) (1998) 160-166.
[17] S. K. Shaha, F. Czerwinski, W. Kasprzak, J. Friedman, & D. L. Chen: Mat Sci
Eng A, 652 (2016) 353-364.
[18] N. L. Yue, L. Lu, M. O. Lai: Struct, 47(1-4) (1999) 691-694.
[19] K. Niranjan, P. R. Lakshminarayanan: Mat Des, 47 (2013) 167-173.
[20] I. Bertoti: Surf Coat Tech, 151 (2002) 194-203.
[21] S. Gota, E. Guiot, M. Henriot, M. Gautier-Soyer: Phy Rev B, 60(20) (1999) 14387.
[22] P. Li, E. Y. Jiang, H. L. Bai: J Phy D: App Phy, 44 (7) (2011) 075003.
[23] P. L. Menezes, S. V. Kailas: Wear, 265 (11-12) (2008) 1655-1669.
[24] Z. Xu, X. Shi, Q. Zhang, W. Zhai, J. Yao, L. Chen, Q. Zhu, Y. Xiao: J Mat Eng
Perform, 23(6) (2014) 2255-2264.
[25] F. Gul, M. Acilar: Comp Sci Tech, 64(13-14) (2004) 1959-1970.
[26] A. Mandal, B. S. Murty, M. Chakraborty: Wear, 266(7-8) (2009) 865-872.
[27] K. M. Shorowordi, A. S. M. A. Haseeb, J. P. Celis: Wear, 256 (11-12) (2004)
1176-1181.
[28] I. Kakaravada, A. Mahamani, V. Pandurangadu: Int J Mat Eng Innov, 11(2) (2020)
145-162.
[29] J. Singh, A. T. Alpas: Metall Mat Trans, A 27(10) (1996) 3135-3148.
[30] B. Venkataraman, G. Sundararajan: Wear, 245(1-2) (2000) 22-38.
[31] Q. G. Wang, C. H. Caceres: Mat Sci Eng: A, 234 (1997) 106-109.
[32] A. S. Vivekananda, S. B. Prabu: Tribo Lett, 66 (1) (2018) 1-14.
[33] S. Suresh, N. S. V. Moorthi, S. C. Vettivel, N. Selvakumar: Mat Des, 59 (2014)
383-396.
[34] Y. Shen, T. Hong, J. Geng, G. Han, D. Chen, X. Li, H. Wang: Mat Chart, 124
(2017) 25-30.
[35] S. D. Kumar, A. Mandal, M. Chakraborty: Mat Sci Eng: A, 636 (2015) 254-262.

Creative Commons License

This work is licensed under a Creative Commons Attribution 4.0 International License.

You might also like