Next Article in Journal
Evaluation of Wear Resistance of AISI L6 and 5140 Steels after Surface Hardening with Boron and Copper
Next Article in Special Issue
Experimental Investigation on the Wear Performance of Nano-Additives on Degraded Gear Lubricant
Previous Article in Journal
Study on Sealing Characteristics of Compliant Foil Face Gas Seal under Typical Hypervelocity Gas Effects
Previous Article in Special Issue
Impact of Thermal and Activation Energies on Glauert Wall Jet (WJ) Heat and Mass Transfer Flows Induced by ZnO-SAE50 Nano Lubricants with Chemical Reaction: The Case of Brinkman-Extended Darcy Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Tribological Behaviour of Hybrid MoS2@Ti3C2 MXene as an Effective Anti-Friction Additive in Gasoline Engine Oil

by
Kalaimani Markandan
1,*,
Thachnatharen Nagarajan
2,
Rashmi Walvekar
3,
Vishal Chaudhary
4,5 and
Mohammad Khalid
6,7,*
1
Department of Chemical & Petroleum Engineering, Faculty of Engineering, Technology & Built Environment, UCSI University, Kuala Lumpur 56000, Malaysia
2
Faculty of Defence Science and Technology, National Defence University of Malaysia, Kuala Lumpur 57000, Malaysia
3
Department of Chemical Engineering, School of Energy and Chemical Engineering, Xiamen University Malaysia, Sepang 43900, Malaysia
4
Research Cell & Department of Physics, Bhagini Nivedita College, University of Delhi, Delhi 110043, India
5
SUMAN Laboratory (Sustainable Materials and Advanced Nanotechnology), New Delhi 110072, India
6
Graphene & Advanced 2D Materials Research Group (GAMRG), School of Engineering and Technology, Sunway University, Petaling Jaya 47500, Malaysia
7
Sunway Materials Smart Science & Engineering (SMS2E) Research Cluster, School of Engineering and Technology, Sunway University, Petaling Jaya 47500, Malaysia
*
Authors to whom correspondence should be addressed.
Lubricants 2023, 11(2), 47; https://doi.org/10.3390/lubricants11020047
Submission received: 27 December 2022 / Revised: 24 January 2023 / Accepted: 27 January 2023 / Published: 29 January 2023
(This article belongs to the Special Issue Nanolubrication and Superlubrication)

Abstract

:
Hybrid molybdenum disulfide (MoS2)-MXene (Ti3C2) was added as an additive in SAE 5W-40-based engine oil in an attempt to reduce interfacial friction between contact surfaces. It was found that the coefficient of friction (COF) and wear scar diameter (WSD) were reduced by 13.9% and 23.8%, respectively, with the addition of 0.05 wt.% MoS2-Ti3C2 compared to base engine oil due to the interlaminar shear susceptibility of MXene. However, we postulate that the high surface energy and presence of -OH, -O and -F functional groups on the surfaces limited the dispersibility and stability of MXene in base oil, while high activity of MoS2 nanoparticles due to large surface area and vigorous Brownian motion prompted fast settling of nanoparticles due to gravitational force. As such, in the present study, hybrid MoS2-Ti3C2 were amine-functionalized to attain stability in SAE 5W-40-based engine oil. Experimental findings indicate that amine-functionalized 0.05 wt.% MoS2-Ti3C2 exhibited higher COF and WSD, i.e., 12.8% and 12.3%, respectively, compared to base oil added with 0.05 wt.% unfunctionalized MoS2-Ti3C2. Similarly, Noack oil volatility was reduced by 24.6% compared to base oil, indicating reduced oil consumption rate, maximal fuel efficiency and enhanced engine performance for a longer duration.

1. Introduction

In automotive engines, parasitic energy loss due to friction and wear of worn surfaces remains a major concern. According to studies, the overall power produced by the engine can be reduced by 17–19% due to frictional losses, whereas 33% of fuel energy in passenger automobiles is squandered to reduce frictional losses [1,2]. The high frictional losses result in higher engine wear, failures of steady running components, higher energy consumption and fuel energy loss [3]. As such, employing lubricants can effectively reduce friction and wear, fulfilling the low carbon emission and fuel economy requirements in car engines. Generally, lubrication can be categorized into four regimes—(i) boundary lubrication, (ii) elastohydrodynamic lubrication, (iii) hydrodynamic lubrication and (iv) mixed lubrication. Engine and transmission systems in cars operate under all the aforementioned lubrication regimes.
Recent studies have reported that adding nanomaterials as lubricant additive can potentially improve tribological behaviour, such as the wear and friction properties of commercial lubricants. This is because, besides affecting the physiochemical properties of engine oil, nanoparticles could minimize possibilities of asperities direct contact (i.e., valleys between asperities of frictional surfaces filled) to form a tribo-boundary film on worn surfaces to enhance tribological behaviour of an engine, i.e., via boundary lubrication conditions [4]. For instance, in a study by Ali et al., coefficient of friction (COF) of SAE 5W-30 oil was reduced by 50% and 45% with the addition of TiO2 and Al2O3 nanoparticles, respectively [5]. In other studies, Li et al., reported that the addition of 0.1 wt.% of ZrO2-SiO2 nanoparticles reduced COF of base oil by 16.24% [6], while Ali et al., [1] reported that the addition of graphene improved anti-friction and antiwear properties by 29% and 22%, respectively.
Today, the family of 2D materials has been augmented by MXene made of 2-dimensional transition metal carbides, nitrides or carbon-nitrides with a general formula Mn+1XnTx where M represents the transition metal (n = 1,2 or 3), X denotes C and/or N while Tx represents the termination group (e.g., OH, -O or -F) [7,8,9,10]. MXene is produced from the MAX phases, where A layers (i.e., group IIIA or IVA element) of Mn+1Xn are removed by selective etching. In addition to energetic and catalytic applications, the lamellar Ti3C2, as the most representative member of the MXene family, are particularly interesting for tribological purposes owing to its graphite-like structure with low shear strength and self-lubricating properties [11,12]; all of which were verified by density functional calculation and molecular dynamic simulations [13]. In a study by Liu et al., it was reported that the addition of 0.8 wt.% Ti3C2 reduced the COF of PAO8 oil by 9.6% [14]. In another study by Gao et al., it was reported that Ti3C2Tx-based nano additives reduced the wear volume of base oil by 87% with a high load carrying capacity (i.e., 500 N) [15]. Despite the tremendous potential of MXene, studies on tribological behaviour of Ti3C2, particularly as an additive (or hybrid additive) in gasoline engine oil, remain limited.
In view of these limitations, the present study aims to investigate the tribological behaviour of hybrid molybdenum disulfide (MoS2)-Ti3C2 as an additive in SAE 5W-40-based engine oil. The hexagonal crystal structured MoS2 is a very stable compound made of Mo atoms sandwiched between double layers of sulphur atoms and linked via van der Waals forces [16,17,18]. MoS2 exhibits low COF owing to the ease of its basal plane shearing, particularly under vacuum conditions. Studies have reported that (MoS2) is capable of imparting excellent lubrication at high contact stress by the mechanism of settling a solid lubricant layer on the contact surfaces [19]. As such, the present study aims to investigate the effect of hybrid MoS2-Ti3C2 additive on the coefficient of friction, wear scar diameter, viscosity index, oxidative induction time (OIT) and Noack volatility of SAE 5W-40-based engine oil. To explore the stability of nanoparticle dispersion in base oil (i.e., sufficient repulsive interaction to counter agglomeration) and the subsequent increase in tribological behaviour, the effect of amine functionalization of MoS2-Ti3C2 was also studied in detail.

2. Materials and Method

2.1. Materials

All chemicals used in this experiment were analytical grade and used as received. Ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O) purchased from Fisher Chemicals (Chicago, IL, USA), titanium aluminium carbide (Ti3AlC2) purchased from Xiamen TOB New Energy Technology (Xiamen, China), thiourea (SC(NH2)2) purchased from R&M Chemicals (Dundee, UK), octadecylamine (C18H37NH2), ethanol (C2H5OH), lithium fluoride (LiF), hydrochloric acid (HCl) and ethanol purchased from Sigma-Aldrich, St. Louis, MO, USA were used for MoS2-Ti3C2 hybrid nanoparticle synthesis. The lubricant oil used was gasoline engine oil, SAE5W40.

2.2. Synthesis of MXene (Ti3C2)

The microwave-hydrothermal synthesis method was used to synthesize Ti3C2 since it is advantageous compared to conventional etching or delamination process due to remarkably reduced synthesis time, energy requirement and high quality of MXene that can be produced in bulk quantities. The procedure has been described in detail in our previous study [20,21]. A total of 1.6 g of LiF was dissolved in 10 mL of 9M HCl by stirring for 15 min at room temperature. Later, 1 g of Ti3AlC2 was slowly added to the mixture and left to stir at room temperature for another 30 min. The mixture was later sonicated for 30 min and transferred into the advanced microwave synthesis platform (Milestone, flexiWAVE, Bergamo, Italy). The mixture was heated for 30 min at 40 °C and left to cool naturally at room temperature once the synthesis was completed. The synthesized sample was centrifuged and washed with distilled water and ethanol to attain pH of ˃5.0, then sonicated in Argon (Ar) atmosphere for an hour and later centrifuged to separate the precipitate from the supernatant where finally, the supernatant will be freeze-dried to extract Ti3C2 powder.

2.3. Preparation of MoS2-Ti3C2 Hybrid Nanoparticle

The hybrid MoS2-Ti3C2 nanoparticles were also synthesized by advanced microwave synthesis method. A total of 3.70758 g of (NH4)6Mo7O24·4H2O and 6.8508 g thiourea were slowly added in 105 mL of deionized water and continuously stirred for 30 min to obtain a homogeneous solution. Later, 1 g of Ti3C2 powder was dissolved in the mixture by sonicating for 30 min. The mixture was then transferred into the advanced microwave synthesis platform (Milestone, flexiWAVE, Italy) and heated for 15 min at 200 °C. Upon completion of synthesis, the reaction mixture was left to cool naturally to room temperature and later centrifuged and washed with distilled water and ethanol and finally freeze-dried to obtain MoS2-Ti3C2 particles.

2.4. Synthesis of Amine-Functionalized MoS2-Ti3C2 Hybrid Nanoparticle

Amine-functionalized MoS2-Ti3C2 hybrid nanoparticles were synthesized by adding 1 g of MoS2-Ti3C2 in 100 mL of DI water and stirred continuously for 30 min. Later, 2 g of octadecylamine was added to 100 mL ethanol and stirred until all solutes were completely dissolved. Both these solutions were slowly mixed and heated at 40 °C under constant stirring for 24 h. The as-formed amine-functionalized MoS2-Ti3C2 hybrid nanoparticles were washed with DI water to remove unreacted octadecylamine and freeze-dried to obtain the functionalized MoS2-Ti3C2 hybrid nanoparticle.

2.5. Characterizations

The morphology of MoS2, Ti3C2 and hybrid MoS2-Ti3C2 particles was obtained using FEI Quanta 400F, Fremont, CA, USA, at 20 kV. XRD analysis was performed using a PANalytical X-ray Diffractometer using copper material to generate X-rays with wavelength (K alpha) of 1.54 Å and filtered through Ni using an operational voltage of 45 kV and current of 27 mA. The samples were scanned from 20 to 80 degrees at a step size of 1 degree/min and divergence slit of 0.9570 degrees. Raman spectroscopy analysis was performed using a stellar-Pro ML150 laser coupled to an optical microscope (Leica DM 2500 M) with Renishaw He-Ne laser with 633 nm wavelength as excitation source while zeta potential measurements to evaluate the stability of lubricants were determined using Zetasizer Nano (Malvern, Worcestershire, UK).
The Ducom four-ball tribotester TR-30L was used to analyze the tribological behaviour of nanolubricants, such as coefficient of friction (COF) and wear scar diameter (WSD). Before performing each tribological experiment, the steel balls were washed in ethanol and dried to avoid contamination. The hardness, density and diameter of the steel balls were 1 HRC, 7.79 g/cm3 and 12.7 mm, respectively. The test was conducted according to ASTM 4172-94, where the rotational speed, applied load, time, and temperature were 12,000 rpm, 392.5 N, 3600 s and 75 °C, respectively. The coefficient of friction was calculated using Equation (1)
μ = 2.22707 τ ρ
where μ represents the coefficient of friction, τ represents the average frictional torque in kg-cm, and ρ is the load exerted while performing the test. The viscosity index was calculated according to ASTM D2270 based on kinematic viscosity at 40 °C and 100 °C. Kinematic viscosity in the present study was calculated according to Equation (2).
v = μ ρ
where v , μ and ρ represent kinematic viscosity, dynamic viscosity and density, respectively. Densities of the lubricant were determined using Anton Paar (DMA 4500 M) at 40 °C and 100 °C. Oxidation induction time (OIT) were determined using TA Instrument’s High Pressure Differential Scanning Calorimeter (HP-DSC) 25P with ultrahigh purity (UHP) grade oxygen gas in an airtight sample chamber with flow rate of 50 mL/min, pressure of 500 psi and isothermal temperature of 200 °C with ramping rate of 10 °C/min. Approximately 3.2 mg of oil samples equilibrated at 50 °C were used for the testing. The relationship between kinetic rate constant and temperature in kinetic expressions, which governs OIT measurements, was based on the Arrhenius equation. Sample volatility was determined according to ASTM D6375 using Noack analysis. Approximately 40 mg of samples were heated to 249 °C at 65 °C/min with a continuous air purge at 150 mL/min using the NETZSCH TGA (TG 209 F3 Tarsus). The samples were held isothermally at 249 °C for 15 min, and percentage mass loss was determined at Noack reference time (t = 11.7 min), as specified by analyzing a Noack reference oil (RL-N).

3. Results and Discussion

3.1. Morphology and Characterization

The structure and composition of bulk MAX phase precursor, MXene sheets and the synthesized hybrid MoS2-Ti3C2 were characterized by SEM and EDX, as shown in Figure 1. From Figure 1A, the microstructure of bulk MAX phase precursor Ti3AlC2 shows solid and stacked multilayer nanosheets, while Figure 1B shows the SEM image of Ti3AlC2 after etching and exfoliation with HCl-LiF via microwave-assisted hydrothermal synthesis. The MXene sheets in Figure 1B showed crumpled structures with thinner and few layered structures, indicating the successful formation of thin layers of 2D MXene nanosheets by excluding Al element from packed MAX layers during the etching and exfoliation process. Additionally, it can be seen in Figure 1C,D that the hybrid MoS2-Ti3C2 was properly formed where MoS2 and a densely aligned enclosed surface of Ti3C2 MXene. The MoS2 nanoparticles were uniformly distributed, well faceted, densely grown, semi-vertically and interleaving lamellar nanosheets with rough edges, which ascertains the nanosheet morphology of MoS2. The edge-oriented structure of MoS2 is highly desirable to expose more active sites to enhance tribological behaviour [22]. The EDX spectrum and elemental mapping are shown in Figure 1E,F, ascertaining the existence of sulphur, molybdenum and titanium and the successful formation of hybrid MoS2-Ti3C2 via microwave-assisted hydrothermal synthesis method.
Figure 2 shows the XRD spectrum of MoS2, Ti3C2 and MoS2-Ti3C2 composites, which is used to characterize the phase and crystal structure of the samples. From the XRD spectrum (Figure 2A), the diffraction peaks located at 14.1, 33 and 59.1 degrees can be attributed to the (002), (100) and (110) peaks of pure hexagonal MoS2 in accordance with JCPDS 371492 [21,23]. The diffraction pattern of Ti3C2 MXene shows peaks at 9, 19.4, 27 and 61 degrees which correspond to (002), (006), (008) and (110) crystal planes, respectively. Previous studies have reported the (002) peak shift of Ti3C2Tx composites when deposited with MoS2 (i.e., downshifted from 8.96 to 7.33 degrees after deposition with MoS2) [24]. This was attributed to the interlayer spacing of both MoS2 and Ti3C2 MXene, which was expanded simultaneously by the formation of MoS2-Ti3C2 composite due to ion intercalation and integrated 2D layer stackings [24,25]. However, in the present study, the peak shift was minimal (i.e., from 9 to 8.6 degrees), indicating that MoS2 did not significantly impact the interlayer distance of MoS2 and Ti3C2 MXene.
Figure 2B shows the Raman spectrum of MoS2, Ti3C2 and MoS2-Ti3C2 composites. The MoS2 samples have shown characteristics at 377 cm−1 and 403 cm−1, corresponding to E2g and A1g vibration peaks, respectively. The A1g vibration mode is favoured for edge-terminated structures owing to polarization dependence [25]. The relative intensity of A1g to E2g can be used to describe the MoS2 planes where in the present study, the A1g/E2g ratio was approximately 1.7 indicating an edge-enriched MoS2 structure. On the other hand, characteristic peaks of Ti3C2 were observed, such as at 208 cm−1 and 719 cm−1, which correspond to the A1g symmetry out-of-plane vibrations of Ti and C atoms, while the peak at 272 cm−1 and broad peak at 622 cm−1 corresponds to the Eg group vibrations including the in plane shear modes of Ti, C and surface functional group atoms [26,27]. Notably, Ti3C2 related peak (e.g., 622 cm−1) diminished in the MoS2-Ti3C2 due to the high surface coverage of MoS2 and the intrinsically weak signal nature of Ti3C2 [25]. Nevertheless, the co-existence of peaks of MoS2 and Ti3C2 in the MoS2-Ti3C2 ascertains the successful preparation of the composite.

3.2. Tribological Behaviour

Figure 3 shows the coefficient of friction (COF), average wear scar diameter as well as SEM images of the wear surfaces of MoS2-Ti3C2-based nanolubricant. From Figure 3A,B, it can be seen that COF of the base oil was 0.108. The addition of MoS2-Ti3C2 gradually reduced the COF of nanolubricant by 13.9% when 0.05 wt.% MoS2-Ti3C2 was added. Sliding off the nanosheets at the asperities and deformed surfaces of individual nanosheets produces a protective layer (i.e., tribofilm) which effectively reduces COF while providing an excellent ability to withstand shear failure [28,29]. The hybrid MoS2-Ti3C2 also covers the plateau region between rough asperities, which assists in forming robust tribolfilm on the steel surface. The formed tribofilm effectively prevents the occurrence of tribo-oxidation, which in turn reduces the COF.
It should be noted that reduction in COF of the nanolubricant when added with 0.005 wt.% MoS2-Ti3C2 was insignificant (i.e., 5.1%), which can be due to insufficient concentration of nano-additive to support the formation of a uniform and dense friction film between the sliding pair [30]. It is also noteworthy that functionalized MoS2-Ti3C2-based nanolubricant exhibited significantly lower COF compared to the unfunctionalized counterpart. For instance, adding 0.05 wt.% of functionalized MoS2-Ti3C2 decreased the coefficient of friction by 24.9% when compared to base oil (without additives). However, the COF for lubricants added with both functionalized and unfunctionalized additives showed a declining tendency above 0.05 wt.% MoS2-Ti3C2. This phenomenon can be related to the possible agglomeration of nanoparticles due to strong van der Waals or Coulombic attractive interactions at contact compulsion, which inhibits their performance at the friction interface.
Figure 3C shows the wear scar diameters of MoS2-Ti3C2-based nanolubricant monitored using an optical profilometer, while Figure 3D–F shows the SEM images of wear scar created on the ball bearing during tribological testing. From Figure 3B, it can be seen that the wear scar diameter of the base oil was reduced when added with MoS2-Ti3C2 nanoparticles. The reduction in wear scar diameters was more significant for the functionalized MoS2-Ti3C2-based nanolubricant. For instance, the wear scar diameter of base oil was reduced by 23.8% and 33.2% when added with 0.05 wt.% of unfunctionalized and functionalized MoS2-Ti3C2 nanoparticles, respectively. In our previous study, it was reported that the addition of 0.05 wt.% functionalized MoS2 reduced the COF and WSD of base oil by 10.3% and 10.6%, respectively [21], significantly lower than COF and WSD reduced in the present study (i.e., 24.9% and 33.2%, respectively) at a similar concentration of functionalized hybrid MoS2-Ti3C2. The excellent abrasive resistance can be attributed to the synergistic role of MoS2 and Ti3C2 nanoparticles, which form a thin lubricating layer between the contact surfaces to reduce contact pressure and frictional torque. These findings were ascertained by the SEM images of the wear scar where surface of ball bearing with base oil without additives exhibited darker concentric grooves (darker furrow is deeper while brighter furrow is shallower) (Figure 3D) while the addition of unfunctionalized MoS2-Ti3C2 nanoparticles (Figure 3E) did not show significant appearance of the dark concentric grooves indicating lesser abrasive wear. When the lubricant was combined with functionalized MoS2-Ti3C2 nanoparticles (Figure 3F), bright and smooth wear tracks were visible, indicating a decrease in the contact surfaces between the steel balls. It is postulated that the planar geometry of MoS2 nanosheets allows them to penetrate easily and readily slide between the oil surface. Moreover, the nanosheets provide a continuous layer on the sliding surfaces due to their excellent adherence to contacts. This phenomenon is related to the mending effect of MoS2-Ti3C2, where the nanoparticles occupy the scratched, worn surface to avoid direct contact between the two surfaces, reducing the wear scar diameter.
Figure 4 shows the zeta potential and visual observation (for sedimentation) of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant after 14 days. Zeta potential measurements which measure the potential difference between the surface of nanoparticles and a stagnant layer of lubricant attached to the particles, are important quantitative indicators to measure the dispersion stability of nanofluids. It accurately represents the magnitude of electrostatic charge between nanoparticles dispersed in the lubricant. In the present study, the zeta potential of both unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant after 14 days was higher than 60 mV. Previous research has shown that dispersion with a greater zeta potential magnitude (whether negative or positive) is electrically stable, whereas lower zeta potential magnitudes signal probable agglomeration or flocculation in the fluid. Dispersion with zeta potential >60 mV was considered extremely stable, while dispersion with a zeta potential of 40–60 mV was considered to exhibit physical stability and dispersion with zeta potential <20 mV has limited stability with high possibilities of nanoparticle agglomeration in the lubricant [31,32]. As such, zeta potential values obtained in the present study (i.e., >60 mV) indicates the excellent stability of MoS2-Ti3C2 in the lubricant.
It is also evident from the visual observations of nanolubricants where MoS2 nanoparticles were stable in the nanolubricant after 14 days. However, nanolubricant reinforced with 0.05 wt.% and 0.1 wt.% MoS2-Ti3C2 showed visible sedimentation after 14 days. This observation is ascertained by the zeta potential measurements, which were reduced by 78.8% and 69.6% when added with 0.1 wt.% of unfunctionalized and functionalized MoS2-Ti3C2, respectively. This indicates that at higher concentrations (>0.05 wt.%), there is limited nanoparticle movement which hinders the growth of energy barrier and leads to particle agglomeration and sedimentation.

3.3. Viscosity Behaviour and Oxidative Induction Time (OIT)

Figure 5 shows the kinematic viscosity of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant analyzed at 40 °C and 100 °C according to ASTM D445. Figure 5 shows that kinematic viscosity of all synthesized nanolubricant in the present study conforms to viscosity class SAE 5W40 lubricant, which indicates synthetic, multigrade lubricant with high shear (HS) stable polymer from group III. The decreased kinematic viscosity with an increase in temperature is due to the reduced attractive binding energy between the molecules when temperature rises [33]. As such, from the kinematic viscosity results, it can be ascertained that MoS2-Ti3C2 nanoparticles are effective as viscosity modifiers in the nanolubricant when temperature varies.
The importance of reducing CO2 emissions from automobiles has urged the need to develop engine oils that exhibit improved fuel-saving performance, which can be envisaged by developing engine oil with high viscosity index (VI). Figure 6 shows the variation in viscosity index with nanoparticle concentration for unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant. Figure 6 shows that the VI of base oil increased by 3.1% and 4% when added with 0.01 wt.% of unfunctionalized and functionalized MoS2-Ti3C2, respectively. Functionalization has modified the surface of the nanoparticles, so they remain well dispersed in the lubricant. Moreover, the higher VI of all MoS2-Ti3C2-based nanolubricant compared to base oil indicates the lower viscosity change in lubricant at fluctuating temperatures where MoS2-Ti3C2 particles ensured consistent and stable lubricative properties across a wide temperature range. The MoS2-Ti3C2 particles provided excellent resistance to thinning of lubricant film for optimized fuel consumption in the engine. Generally, engine oils tend to be more viscous at a lower temperature, raising drag inside the engine. Engine oil with high VI will have lower viscosity at lower temperatures, which translates into enhanced fuel-saving performance. As such, MoS2-Ti3C2-based nanolubricant synthesized in the present study show great potential for use within the automobile industry.
Figure 6C shows the oxidation induction time (OIT) of the MoS2-Ti3C2-based nanolubricant synthesized in the present study, where the induction time accurately represents their oxidative stability. In the automobile industry, oxidation of lubricants is caused by intense temperature, high load and continuous contact with air. The oxidation process leads to accelerated engine oil degradation, affecting the performance, innate high efficiency and lifespan. Based on Figure 6C, it can be seen that OIT improved with the addition of MoS2-Ti3C2 as an additive in the engine oil, where maximum OIT improved by 17.8% with the addition of 0.01 wt.% functionalized MoS2-Ti3C2. The synergistic interaction between molybdenum dialkyldithiocarbamate (MoDTC) and zinc dialkyldithiophosphate (ZDDP) due to MoS2 production, alongside the antiwear/antioxidant (i.e., ability of phosphate to digest oxides) behaviour and extreme pressure characteristics of ZDDP has been reported previously [34,35]. Generally, lubricants undergo a three-step oxidation process, i.e., (i) initiation to produce free radicals, (ii) propagation, where free radicals combine with oxygen to produce peroxide radicals and (iii) termination stage, where two radicals combine to form a stable molecule. It is postulated that the synergistic effect of MoS2 with ZDDP promotes hydrogen donation, which stops the radical propagation stage, and increases the OIT [21].
Figure 6. (A) Variation of viscosity index with nanoparticle concentration for unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant (B) Schematic illustration describing relationship between engine oil viscosity and temperature, adapted with permission from ref. [36] (C) OIT analysis of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant.
Figure 6. (A) Variation of viscosity index with nanoparticle concentration for unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant (B) Schematic illustration describing relationship between engine oil viscosity and temperature, adapted with permission from ref. [36] (C) OIT analysis of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant.
Lubricants 11 00047 g006

3.4. Noack Volatility

Figure 7 shows the Noack volatility analysis of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant synthesized in the present study. Noack volatility (evaporative loss) of engine oils is important to determine the rate of oil evaporation at very high temperatures and additional oil consumption as a result. For instance, when volatility is low, more oil will stay in the engine with its ideal viscosity for a longer duration. However, when volatility is high, oil loss occurs easily, resulting in a viscosity increase, which leads to reduced fuel economy due to an increase in parasitic load. From Figure 7A,B, it can be seen that adding MoS2-Ti3C2 nanoparticles up to 0.05 wt.% reduced the Noack volatility of lubricants. The decrease in volatility was more significant for lubricants added with 0.05 wt.% functionalized MoS2-Ti3C2 nanoparticles, where Noack volatility was reduced by 24.6% compared to the base oil. The reduced oil volatility indicates the reduced oil consumption rate, maximal fuel efficiency and enhanced engine performance for a longer duration.
Figure 7D shows the Noack volatility of various commercially available SAE 5W-40 lubricants compared to 0.05 wt.% functionalized MoS2-Ti3C2-based nanolubricant synthesized in the present study. Regardless of the testing standard, it is noteworthy that all the lubricants exhibited Noack volatility of <15%; since any volatility >15% is considered high and will not pass the crucial oxidation test, IIG engine test and does not adhere to the requirements of API oil guidelines [37]. On the other hand, General Motors and European automobile manufacturer’s association (ACEA) have established specifications to limit maximum Noack volatility at 13%. In the present study, we report Noack volatility of the synthesized MoS2-Ti3C2-based nanolubricant to be 10.7, which is lower than many commercial SAE 5W-40 engine oils such as Duragard® Diamond Synthetic, Wakefield Diesel Engine Oil or Castrol EDGE among others. Although the Noack volatility can be further improved to meet the standards of several commercial engine oils such as Havoline® ProDS® or RAVENOL VDL, it is nonetheless plausible to claim that proper use and further research on functionalized MoS2-Ti3C2-based nanolubricant can allow tremendous synergistic benefits between reduced volatility, excellent abrasive resistance and friction modifier to reduce micro-pitting by minimizing local stresses near the surfaces (i.e., lowered friction).

4. Conclusions

In conclusion, the present study has successfully reported hybrid MoS2@Ti3C2 MXene as an effective anti-friction additive with enhanced tribological performance in SAE 5W-40-based engine oil. Amine functionalization showed better dispersibility, stability and tribological behaviour in the engine oi. For instance, the coefficient of friction was reduced by 24.9% compared to base oil with the addition of 0.05 wt.% functionalized MoS2-Ti3C2. Similarly, wear scar diameter of base oil was reduced by 23.8% and 33.2% when added with 0.05 wt.% of unfunctionalized and functionalized MoS2-Ti3C2 nanoparticles, respectively. Moreover, OIT improved by 17.8% with the addition of 0.01 wt.% functionalized MoS2-Ti3C2. Based on Noack’s analysis, the volatility of MoS2-Ti3C2-based nanolubricant was 10.7, which is lower compared to many commercial SAE 5W-40-based engine oil. The reduced oil volatility indicates the reduced oil consumption rate, maximal fuel efficiency and enhanced engine performance for a longer duration.

Author Contributions

Conceptualization, R.W., K.M., V.C. and M.K.; methodology, T.N. and M.K.; formal analysis, K.M. and T.N., writing—original draft preparation, K.M. and T.N., writing—review and editing, K.M., M.K. and R.W.; visualization, K.M. and T.N.; supervision, M.K. and K.M.; project administration, K.M. and M.K.; funding acquisition, K.M. and M.K. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge partial funding for this project by UCSI Research Excellence & Innovation Grant (REIG-FETBE-2022/019) and Sunway University’s International Research Network Grant Scheme (STR-IRNGS-SET-GAMRG-01-2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ali, M.; Xianjun, H.; Abdelkareem, M.; Gulzar, M.; Elsheikh, A. Novel approach of the graphene nanolubricant for energy saving via anti-friction/wear in automobile engines. Tribol. Int. 2018, 124, 209–229. [Google Scholar] [CrossRef]
  2. Holmberg, K.; Andersson, P.; Erdemir, A. Global energy consumption due to friction in passenger cars. Tribol. Int. 2012, 47, 221–234. [Google Scholar] [CrossRef]
  3. Ali, M.; Xianjun, H.; Turkson, R.; Peng, Z.; Chen, X. Enhancing the thermophysical properties and tribological behaviour of engine oils using nano-lubricant additives. RSC Adv. 2016, 6, 77913–77924. [Google Scholar] [CrossRef]
  4. Yu, R.; Liu, J.; Zhou, Y. Experimental study on tribological property of MoS2 nanoparticle in castor oil. J. Tribol. 2019, 141, 3–7. [Google Scholar] [CrossRef]
  5. Ali, M.; Xianjun, H.; Mai, L.; Qingping, C.; Turkson, R.; Bicheng, C. Improving the tribological characteristics of piston ring assembly in automotive engines using Al2O3 and TiO2 nanomaterials as nano-lubricant additives. Tribol. Int. 2016, 103, 540–554. [Google Scholar] [CrossRef]
  6. Li, W.; Zheng, S.; Cao, B.; Ma, S. Friction and wear properties of ZrO2/SiO2 composite nanoparticles. J. Nanoparticle Res. 2011, 13, 2129–2137. [Google Scholar] [CrossRef]
  7. Rasheed, A.; Khalid, M.; Nor, A.B.M.; Wong, W.; Duolikun, T.; Natu, V.; Barsoum, M.; Leo, B.; Zaharin, H.; Ghazali, M. MXene-graphene hybrid nanoflakes as friction modifiers for outboard engine oil. IOP Conf. Ser. Mater. Sci. Eng. 2020, 834, 012039. [Google Scholar] [CrossRef]
  8. Naguib, M.; Mochalin, V.; Barsoum, M.; Gogotsi, Y. 25th anniversary article: MXenes: A new family of two-dimensional materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef] [PubMed]
  9. Gogotsi, Y.; Anasori, B. The Rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Gogotsi, Y.; Huang, Q. MXenes: Two-Dimensional Building Blocks for Future Materials and Devices. ACS Nano 2021, 15, 5775–5780. [Google Scholar] [CrossRef] [PubMed]
  11. Yan, H.; Zhang, L.; Li, H.; Fan, X.; Zhu, M. Towards high-performance additive of Ti3C2/graphene hybrid with a novel wrapping structure in epoxy coating. Carbon 2020, 157, 217–233. [Google Scholar] [CrossRef]
  12. Nguyen, H.; Chung, K. Assessment of tribological properties of ti3c2 as a water-based lubricant additive. Materials 2020, 13, 5545. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, D.; Ashton, M.; Ostadhossein, A.; Van Duin, A.; Hennig, R.; Sinnott, S. Computational study of low interlayer friction in Tin+1Cn (n = 1, 2, and 3) MXene. ACS Appl. Mater. Interfaces 2017, 9, 34467–34479. [Google Scholar] [CrossRef]
  14. Liu, Y.; Zhang, X.; Dong, S.; Ye, Z.; Wei, Y. Synthesis and tribological property of Ti3C2TX nanosheets. J. Mater. Sci. 2017, 52, 2200–2209. [Google Scholar] [CrossRef]
  15. Gao, J.; Du, C.-F.; Zhang, T.; Zhang, X.; Ye, Q.; Liu, S.; Liu, W. Dialkyl Dithiophosphate-Functionalized Ti3C2TX MXene Nanosheets as Effective Lubricant Additives for Antiwear and Friction Reduction. ACS Appl. Nano Mater. 2021, 4, 11080–11087. [Google Scholar] [CrossRef]
  16. Vazirisereshk, M.; Martini, A.; Strubbe, D.; Baykara, M. Solid lubrication with MoS2: A review. Lubricants 2019, 7, 57. [Google Scholar] [CrossRef] [Green Version]
  17. Serles, P.; Gaber, K.; Pajovic, S.; Colas, G.; Filleter, T. High Temperature Microtribological Studies of MoS2 Lubrication for Low Earth Orbit. Lubricants 2020, 8, 49. [Google Scholar] [CrossRef]
  18. Thachnatharen, N.; Khalid, M.; Shahabuddin, S.; Anwar, A.; Sridewi, N. Tribological analysis of advanced microwave synthesized Molybdenum disulfide (MoS2) as anti-friction additives in diesel engine oil for military vehicles. Mater. Today Proc. 2022, 62, 7243–7247. [Google Scholar] [CrossRef]
  19. Gradt, T.; Schneider, T. Tribological performance of MoS2 coatings in various environments. Lubricants 2016, 4, 32. [Google Scholar] [CrossRef] [Green Version]
  20. Numan, A.; Rafique, S.; Khalid, M.; Zaharin, H.A.; Radwan, A.; Mokri, N.A.; Ching, O.P.; Walvekar, R. Microwave-assisted rapid MAX phase etching and delamination: A paradigm shift in MXene synthesis. Mater. Chem. Phys. 2022, 288, 126429. [Google Scholar] [CrossRef]
  21. Nagarajan, T.; Khalid, M.; Sridewi, N.; Jagadish, P.; Shahabuddin, S.; Muthoosamy, K.; Walvekar, R. Tribological, oxidation and thermal conductivity studies of microwave synthesized molybdenum disulfide (MoS2) nanoparticles as nano-additives in diesel based engine oil. Sci. Rep. 2022, 12, 14108. [Google Scholar] [CrossRef] [PubMed]
  22. Li, X.; Lv, X.; Sun, X.; Yang, C.; Zheng, Y.; Yang, L.; Li, S.; Tao, X. Edge-oriented, high-percentage 1T′-phase MoS2 nanosheets stabilize Ti3C2 MXene for efficient electrocatalytic hydrogen evolution. Appl. Catal. B Environ. 2021, 284, 119708. [Google Scholar] [CrossRef]
  23. Wang, X.; Li, H.; Li, H.; Lin, S.; Ding, W.; Zhu, X.; Sheng, Z.; Wang, H.; Zhu, X.; Sun, Y. 2D/2D 1T-MoS2/Ti3C2 MXene Heterostructure with Excellent Supercapacitor Performance. Adv. Funct. Mater. 2020, 30, 0190302. [Google Scholar] [CrossRef]
  24. Wang, H.; Ma, H. The electromagnetic and microwave absorbing properties of MoS2 modified Ti3C2Tx nanocomposites. J. Mater. Sci. Mater. Electron. 2019, 30, 15250–15256. [Google Scholar] [CrossRef]
  25. Wei, S.; Fu, Y.; Liu, M.; Yue, H.; Park, S.; Lee, Y.; Li, H.; Yao, F. Dual-phase MoS2/MXene/CNT ternary nanohybrids for efficient electrocatalytic hydrogen evolution, Npj 2D. Mater. Appl. 2022, 6, 25. [Google Scholar] [CrossRef]
  26. Liu, Y.; He, Y.; Vargun, E.; Plachy, T.; Saha, P.; Cheng, Q. 3D Porous Ti3C2 MXene/NiCo-MOF Composites for Enhanced Lithium Storage. Nanomaterials 2020, 10, 10040695. [Google Scholar] [CrossRef] [Green Version]
  27. Mao, X.; Zou, Y.; Xu, F.; Sun, L.; Chu, H.; Zhang, H.; Zhang, J.; Xiang, C. Three-Dimensional Self-Supporting Ti3C2 with MoS2 and Cu2O Nanocrystals for High-Performance Flexible Supercapacitors. ACS Appl. Mater. Interfaces 2021, 13, 22664–22675. [Google Scholar] [CrossRef]
  28. Radhika, P.; Sobhan, C.; Chakravorti, S. Improved tribological behavior of lubricating oil dispersed with hybrid nanoparticles of functionalized carbon spheres and graphene nano platelets. Appl. Surf. Sci. 2021, 540, 148402. [Google Scholar] [CrossRef]
  29. Bustillos, J.; Montero, D.; Nautiyal, P.; Loganathan, A.; Boesl, B.; Agarwal, A. Integration of graphene in poly(lactic) acid by 3D printing to develop creep and wear-resistant hierarchical nanocomposites. Polym. Compos. 2018, 39, 3877–3888. [Google Scholar] [CrossRef]
  30. Feng, P.; Ren, Y.; Li, Y.; He, J.; Zhao, Z.; Ma, X.; Fan, X.; Zhu, M. Synergistic lubrication of few-layer Ti3C2Tx/MoS2 heterojunction as a lubricant additive. Friction 2022, 10, 2018–2032. [Google Scholar] [CrossRef]
  31. Mukherjee, S. Preparation and Stability of Nanofluids—A Review. IOSR J. Mech. Civ. Eng. 2013, 9, 63–69. [Google Scholar] [CrossRef]
  32. Zainon, S.; Azmi, W. Recent progress on stability and thermo-physical properties of mono and hybrid towards green nanofluids. Micromachines 2021, 12, 12020176. [Google Scholar] [CrossRef] [PubMed]
  33. Salameh, T.; Kumar, P.; Sayed, E.; Abdelkareem, M.; Rezk, H.; Olabi, A. Fuzzy modeling and particle swarm optimization of Al2O3/SiO2 nanofluid. Int. J. Thermofluids 2021, 10, 100084. [Google Scholar] [CrossRef]
  34. Morina, A.; Neville, A.; Priest, M.; Green, J. ZDDP and MoDTC interactions and their effect on tribological performance—Tribofilm characteristics and its evolution. Tribol. Lett. 2006, 24, 243–256. [Google Scholar] [CrossRef]
  35. Thornley, A.; Wang, Y.; Wang, C.; Chen, J.; Huang, H.; Liu, H.; Neville, A.; Morina, A. Optimizing the Mo concentration in low viscosity fully formulated oils. Tribol. Int. 2022, 168, 107437. [Google Scholar] [CrossRef]
  36. ENEOS Corporation. Development of Next Generation Engine Oils. Lubricants (n.d.). Available online: https://www.eneos.co.jp/english/company/rd/intro/lubricants/hairyo.html (accessed on 14 December 2022).
  37. Shah, R.; Hope, K.; Aragon, N. Fuel economy considerations: Effect of lube oils and their volatility. Lube Magazine, November 2020; pp. 1–4. [Google Scholar]
Figure 1. SEM images of (A) Ti3AlC2 MAX phase (B) Ti3C2 MXene (C,D) Low and high magnification images of hybrid MoS2-Ti3C2 (E,F) EDX spectra of hybrid MoS2-Ti3C2.
Figure 1. SEM images of (A) Ti3AlC2 MAX phase (B) Ti3C2 MXene (C,D) Low and high magnification images of hybrid MoS2-Ti3C2 (E,F) EDX spectra of hybrid MoS2-Ti3C2.
Lubricants 11 00047 g001
Figure 2. (A) XRD spectrum and (B) Raman analysis of MoS2, Ti3C2 and MoS2-Ti3C2.
Figure 2. (A) XRD spectrum and (B) Raman analysis of MoS2, Ti3C2 and MoS2-Ti3C2.
Lubricants 11 00047 g002
Figure 3. (A) Coefficient of friction (COF) against time for base oil with unfunctionalized MoS2-Ti3C2 and functionalized MoS2-Ti3C2 (indicated by letter F). (B) COF and (C) average wear scar diameter of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant. (DF) SEM images of the wear surfaces of base oil without additive, base oil with unfunctionalized MoS2-Ti3C2 and base oil with functionalized MoS2-Ti3C2, respectively.
Figure 3. (A) Coefficient of friction (COF) against time for base oil with unfunctionalized MoS2-Ti3C2 and functionalized MoS2-Ti3C2 (indicated by letter F). (B) COF and (C) average wear scar diameter of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant. (DF) SEM images of the wear surfaces of base oil without additive, base oil with unfunctionalized MoS2-Ti3C2 and base oil with functionalized MoS2-Ti3C2, respectively.
Lubricants 11 00047 g003
Figure 4. (A) Zeta potential of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant (B) Schematic representation illustrating stability and zeta potential ranges (C,D) Visual observation of unfunctionalized MoS2-Ti3C2-based nanolubricant before and after 14 days (E,F) Visual observation of functionalized MoS2-Ti3C2-based nanolubricant before and after 14 days.
Figure 4. (A) Zeta potential of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant (B) Schematic representation illustrating stability and zeta potential ranges (C,D) Visual observation of unfunctionalized MoS2-Ti3C2-based nanolubricant before and after 14 days (E,F) Visual observation of functionalized MoS2-Ti3C2-based nanolubricant before and after 14 days.
Lubricants 11 00047 g004
Figure 5. Variation of kinematic viscosity with temperature for (A) unfunctionalized and (B) functionalized MoS2-Ti3C2-based nanolubricant.
Figure 5. Variation of kinematic viscosity with temperature for (A) unfunctionalized and (B) functionalized MoS2-Ti3C2-based nanolubricant.
Lubricants 11 00047 g005
Figure 7. (A) TGA Noack analysis showing mass loss across time MoS2-Ti3C2-based nanolubricant (B) Effect of varying concentrations of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant on Noack volatility (C) Schematic illustration showing low, medium and high volatility of lubricants at specified Noack reference time (D) Noack volatility of various commercially available SAE 5W-40 lubricants in comparison to MoS2-Ti3C2-based nanolubricant in the present study (i.e., 0.05 wt.% functionalized MoS2-Ti3C2-based nanolubricant).
Figure 7. (A) TGA Noack analysis showing mass loss across time MoS2-Ti3C2-based nanolubricant (B) Effect of varying concentrations of unfunctionalized and functionalized MoS2-Ti3C2-based nanolubricant on Noack volatility (C) Schematic illustration showing low, medium and high volatility of lubricants at specified Noack reference time (D) Noack volatility of various commercially available SAE 5W-40 lubricants in comparison to MoS2-Ti3C2-based nanolubricant in the present study (i.e., 0.05 wt.% functionalized MoS2-Ti3C2-based nanolubricant).
Lubricants 11 00047 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Markandan, K.; Nagarajan, T.; Walvekar, R.; Chaudhary, V.; Khalid, M. Enhanced Tribological Behaviour of Hybrid MoS2@Ti3C2 MXene as an Effective Anti-Friction Additive in Gasoline Engine Oil. Lubricants 2023, 11, 47. https://doi.org/10.3390/lubricants11020047

AMA Style

Markandan K, Nagarajan T, Walvekar R, Chaudhary V, Khalid M. Enhanced Tribological Behaviour of Hybrid MoS2@Ti3C2 MXene as an Effective Anti-Friction Additive in Gasoline Engine Oil. Lubricants. 2023; 11(2):47. https://doi.org/10.3390/lubricants11020047

Chicago/Turabian Style

Markandan, Kalaimani, Thachnatharen Nagarajan, Rashmi Walvekar, Vishal Chaudhary, and Mohammad Khalid. 2023. "Enhanced Tribological Behaviour of Hybrid MoS2@Ti3C2 MXene as an Effective Anti-Friction Additive in Gasoline Engine Oil" Lubricants 11, no. 2: 47. https://doi.org/10.3390/lubricants11020047

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop