Next Article in Journal
Traction Properties of Aerospace Lubricant under Low-Temperature Micro-Oil Droplet Supply Conditions
Next Article in Special Issue
Effect of Speed and Number of Sliding Cycles on the Film Formation Behavior of Fluorine-Based Grease
Previous Article in Journal
Dynamometric Investigation on Airborne Particulate Matter from Automobile Brake: Impact of Disc Materials on Brake Emission Factor
Previous Article in Special Issue
On the Difference in the Action of Anti-Wear Additives in Hydrocarbon Oils and Vegetable Triglycerides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MoS2 Nanomaterials as Lubricant Additives: A Review

by
Ziyan Lu
1,
Qingqing Lin
2,
Zhaotao Cao
2,
Wanyuan Li
2,
Junjie Gong
2,
Yan Wang
2,
Kunhong Hu
2,* and
Xianguo Hu
1,*
1
Institute of Tribology, Hefei University of Technology, Hefei 230009, China
2
School of Energy Materials and Chemical Engineering, Hefei University, Hefei 230601, China
*
Authors to whom correspondence should be addressed.
Lubricants 2023, 11(12), 527; https://doi.org/10.3390/lubricants11120527
Submission received: 29 October 2023 / Revised: 30 November 2023 / Accepted: 9 December 2023 / Published: 11 December 2023

Abstract

:
Improving the lubricating properties of base oils through additives is a crucial objective of tribological research, as it helps to reduce friction and wear of materials. Molybdenum disulfide (MoS2) is a 2D nanomaterial with excellent tribological properties that is often used as a lubricant additive. Several studies have been conducted on the preparation and utilization of MoS2 and its nanocomposites as lubricant additives. This paper reviews the research progress on MoS2 nanomaterials as lubricant additives. It firstly introduces various synthesis methods of MoS2 nanomaterials while focusing on the preparation of nano-MoS2 as lubricant additives. It then summarizes the dispersion stability of nano-MoS2 in lubricating oils which has been paid extensive attention. Moreover, this paper reviews and discusses the tribological properties of nano-MoS2 and its various composites as lubricant additives. The possible anti-wear and friction reduction mechanisms of nano-MoS2 and its composites are also discussed. Finally, this paper presents the challenges faced by nano-MoS2 additives in the field of lubrication and the prospects for future research in view of previous studies.

1. Introduction

Friction and wear have always been the top issues in different industrial fields. While modern industry is developing rapidly, global energy consumption also maintains rapid development. In mechanical systems, energy waste due to friction and wear accounts for about 1/3–1/2 [1,2], and friction and wear of machineries and equipment will seriously affect the normal operation and life cycle of its components [3] and may even produce more dangerous consequences under the process of operation [3]. Mechanical systems in industrial production are usually lubricated with lubricants, which can not only significantly reduce friction and wear and improve lubrication efficiency, but also have important significance for energy saving and environmental protection [4]. Currently, lubricants commonly used in various industries and research experiments are composed of base oils and various lubricant additives [5]. Traditional anti-wear and extreme pressure additives such as molybdenum dialkyl dithiocarbamate (MoDTC) and zinc dialkyl dithiophosphate (ZDDP) can greatly reduce friction and wear by generating beneficial products in situ from the reaction between the additive and the metal friction pair surface [6,7]. However, traditional lubrication additives are facing serious challenges with the development of the industry. In order to fulfill the needs of different fields such as ultra-low friction, wear, and extreme conditions, new requirements are placed on lubricants such as resistance to high temperature and pressure [8]. Lubricant additives can give new properties to the lubricant and make up for the shortcomings of the base oil, thus greatly improving the tribological properties of the lubricant.
The utilization of nanomaterials as additives in lubricants presents a promising avenue to enhance the performance of mechanical systems, surpassing that of the traditional lubricant additives. Nanomaterials are tiny particles, with at least one dimension ranging between 1–100 nanometers, which allows them to easily penetrate the contact zone [9]. Due to their high surface activity, they can strongly adsorb onto the friction surface and form a stable protective film that prevents direct contact between solid surfaces, effectively reducing friction. These nanomaterials include metals [10], metal oxides [11], metal sulfides [12], carbon-based materials [13], nitrides [14], and their composites. During friction processes, nanomaterials can be physically or chemically adsorbed onto the contact surfaces of friction pairs, thus improving lubricant tribological properties, a subject that has been extensively researched in the field of tribology [15,16,17]. Transition metal dichalcogenides (TMDs), due to their structural similarity to graphene, have attracted significant attention from global scholars in recent years for their ability to significantly enhance lubrication and anti-wear properties in lubricants.
The study of using transition metal disulfides as nano-lubricating additives is still under research and development processes. MoS2 is a typical representative of transition metal disulfides and is used as an anti-wear and friction-reducing additive in solid lubricants, lubricants, and greases, or as a lubricating component in coatings [18]. MoS2 is an excellent solid lubricant and has gained the reputation of “king of lubrication” because of its good lubrication performance even under special working conditions like ultra-high temperature and ultra-vacuum. MoS2 also has certain problems, such as humidity and oxygen enrichment that can degrade its performance [19,20,21,22]. Low humidity and low oxygen environments need to be constructed for them to perform. However, MoS2 can overcome the effects of moisture and oxygen by developing specific structures or morphologies. Although some progress has been made in improving MoS2’s resistance to moisture and oxidation [20], there is still much work to be undertaken.
MoS2 was initially used in the mid-19th century during the California Gold Rush as a lubricant for horse-drawn carriage bearings. MoS2 as a lubricant has been the subject of a great deal of scientific research, with reviews and books on the subject published as early as the 1960s [23,24]. Although its use as a dry film lubricant was established in the mid-20th century, its effective use in oils took longer to perfect. Research to understand and improve the lubricating properties of MoS2 is still active. Meanwhile, compounding various nanomaterials provides an alternative approach to creating new materials with excellent tribological properties. Recent studies have focused on integrating MoS2 with other nanomaterials (e.g., graphene) into nanocomposites and improving their lubrication properties through controlled doping [25,26,27,28].
This paper aims to provide guidance for researching on lubricant additives, with specific focus on MoS2 as a lubricant additive. It reviews the progress of tribological research on MoS2 nanomaterials in recent years and summarizes the current understanding of lubricant additives. This study does not address applications of MoS2 as a solid lubricant or in coatings. Section 2 provides details on the structure and synthesis of MoS2, while Section 3 focuses on methods to improve the dispersion stability of MoS2 in base oils by surface modification and preparation of composites. Section 4 discusses the tribological properties and lubrication mechanisms of MoS2 and its complexes. Finally, Section 5 summarizes the challenges that may arise with the use of MoS2 materials as lubricant additives, as well as new directions for future development.

2. Structure and Synthesis

2.1. Structure of MoS2

A single layer of MoS2 is composed of three atomic layers. It consists of a molybdenum layer sandwiched between two sulfur atom layers, forming a sandwich-like structure. MoS2 belongs to the hexagonal or rhombohedral crystal system, with six sulfur atoms distributed around each molybdenum atom and three molybdenum atoms oiled around each sulfur atom, Mo and S are bonded to each other by covalent bonding, whereas the triatomic layer of S-Mo-S is bonded to each other by van der Waals forces, with a layer-to-layer spacing of about 0.615 nm (PDF#37-1492). In addition, due to the difference in the relative positions between Mo and S atoms, three crystal structures are formed as shown in Figure 1 [29]: the 1T (lattice parameters a = 5.60 Å, c = 5.99 Å9), 2H (lattice parameters a = 3.15 Å, c = 12.30 Å), and 3R (lattice parameters a = 3.17 Å, c = 18.38 Å) crystal structures [30]. The 1T-MoS2 structure has an octahedral coordination, which is metallic, and belongs to the substable structure; the 2H-MoS2 crystal form contains two Mo-S units, which belongs to the stable state structure; and the 3R-MoS2 crystal form has one Mo-S unit more than the 2H-MoS2 crystal form which has one more Mo-S unit, i.e., it contains three Mo-S units and belongs to the substable structure. Two of these substrates, 1T-MoS2 and 3R-MoS2, can be transformed into the stable 2H-MoS2 form at high temperatures [31,32]. Based on the fact that MoS2 layers are connected by van der Waals forces, the layers slide very easily, which confers excellent lubrication properties to MoS2.

2.2. Synthesis Method of Nano-MoS2 Nanomaterials

MoS2 is a substance with excellent lubricating properties and has gained widespread attention in the field of tribology. The last few years have seen rapid advancements in the preparation technology and processes of MoS2, which have laid a solid foundation for its basic research and application and significantly promoted its development. The synthesis method of MoS2 is the cornerstone for its development and application and holds significant scientific and economic significance. Similar to graphene, the preparation method for two-dimensional layered MoS2 starts with the mechanical stripping method, and several other preparation methods have been developed through continuous research and exploration. The preparation methods of MoS2 include mechanical exfoliation, hydrothermal, solvothermal, liquid-phase precipitation method, and more.

2.2.1. Mechanical Exfoliation

The earliest method used to prepare MoS2 was micromechanical exfoliation, i.e., the use of adhesive tape to strip thin sheets of MoS2 from a block of MoS2. This method was first proposed by Frindt et al. [33] in 1965, and several tens of layers of MoS2 were successfully obtained by this method. Although the MoS2 obtained by this method has a high degree of crystallinity, the preparation is inefficient and reproducible, which makes it difficult to realize large-scale applications. For this reason, researchers have explored the ultrasound-assisted exfoliation method. Liu et al. [34] prepared graphene-like MoS2 materials by ultrasound exfoliation with the assistance of 1-Dodecanethiol (Figure 2A). Dual-solvent ultrasonic exfoliation was achieved by introducing chloroform and the highest concentration of graphene-like MoS2 was prepared when the volume ratio of 1-dodecanethiol to trichloromethane was 1:1 and the ultrasonication time was 12 h. Yao et al. [35] used chitosan/acetic acid (CS/HAc) solution to sonicate the native phase MoS2 to obtain oligomerized MoS2, the (002) crystalline surface stacking of the treated MoS2 samples was reduced, and the crystalline phase was transformed from a polycrystalline to a monocrystalline structure.

2.2.2. Hydrothermal

The hydrothermal is a method to complete the material synthesis and preparation by heating and pressurizing the reaction system (or self-generated steam pressure) in a specially designed closed reactor (e.g., autoclave) with water as the reaction medium to create a relatively high-temperature and high-pressure reaction condition. The advantage of this method is that the synthesized products are characterized by complete grain development, small and uniform particle size distribution, and light particle agglomeration, and it is one of the most commonly used methods for the preparation of two-dimensional materials, including MoS2 and its nanocomposites [39,40]. Tang et al. [41] successfully obtained flower-like microspheres consisting of MoS2 nanosheets by the surfactant-assisted hydrothermal method at 180 °C for 24 h using a certain proportion of (NH4)2MoO4, NH2OH-HCl and CH4N2S as the initial reactants. Li et al. [42] obtained MoS3 precursors by adding an appropriate amount of [EMIM]Br to an acidic aqueous solution of sodium molybdate and thioacetamide and reacting at 200 °C for 24 h, followed by high-temperature desulfurization to obtain hollow core/shell MoS2 microspheres. Vijaya et al. [43] mixed ammonium hexamolybdate tetrahydrate ((NH4)6Mo7O24·4H2O) and H2NCSNH2 with 30 mL of deionized water and hydrothermally reacted at a temperature of 200 °C for 24 h to obtain MoS2 nanosheets. Xie et al. [41] dissolved ammonium molybdate, thiourea, and glucose in deionized water and placed the reaction at 200 °C for 24 h to obtain CS@1T-MoS2 (Figure 2B).
It can be seen that the hydrothermal method usually involves pre-preparing an aqueous solution containing the raw material, placing it in a sealed autoclave reactor reacting it for some time at a temperature of about 200 °C, and then cooling it naturally to room temperature. The precipitate after the reaction is collected for some additional post-treatment (e.g., calcination) to optimize the product. However, the hydrothermal method requires high-temperature and high-pressure reaction conditions, high dependence on equipment, expensive autoclaves for fabrication, safety issues during the reaction process, and the inability to observe the reaction while synthesizing the nanomaterials, which makes it difficult to realistically control the parameters [44], thus limiting the development of the method to some extent.

2.2.3. Solvothermal

Solvothermal is a high-pressure growth method that does not require any catalyst. It is developed from the hydrothermal method, which is similar but differs in using organic solvents or non-aqueous solvents (such as N, N-dimethylformamide (DMF) [37,45], ethylene glycol [46]) instead of water as the reaction medium. This method is used to prepare materials that cannot grow in aqueous solution and are easily oxidized, hydrolyzed, or sensitive to water, such as III-V semiconductor compounds, nitrides, and sulfur compounds.
Li et al. [45] used ammonium tetrathiomolybdate ((NH4)2MoS4) as a precursor for Mo and S and reacted it in DMF at 210 °C for 36 h to obtain MoS2 nanosheets. In some cases it is also necessary to use DMF and water as a mixed solvent (Figure 2C). Zhao et al. [37] reported that the proper DMF/water ratio is very important for the construction of MoS2 nanocomposites, because the use of DMF as a solvent does not allow for the self-assembly of the hybrid material into the desired structure, and the use of pure deionized water does not result in homogeneous MoS2 nanoparticles. Li et al. [46] conducted an experiment where CoC4H6O4·4H2O and Na2MoO4·2H2O were dissolved in ethylene glycol and stirred with sublimated sulfur and GO to form a homogeneous solution, which was placed in an autoclave and kept at 200 °C for 24 h to obtain the target product. In addition, various other solvents such as N-methyl pyrrolidone (NMP) [47], ethanol [48,49], polyethylene glycol [50], and isopropanol [51] are favorable for solvothermal methods.
Solvothermal has some limitations. Although they are usually less demanding than hydrothermal, they still require high temperature and pressure conditions in most cases. Additionally, the synthesis process usually lasts longer [52,53,54] and consumes more energy. Comparing the two methods, the simple hydrothermal method usually has no negative impact on the material properties, whereas solvothermal methods usually have a greater impact on material properties due to differences in solvents. However, the solvothermal method is more likely to result in a better morphology than the hydrothermal method.

2.2.4. Liquid-Phase Precipitation

Liquid-phase precipitation is a process in which a solution containing the desired reactants is generated by controlling the acidity and temperature of the reaction. By adjusting these factors, a large number of particles can be generated, and the use of surfactants prevents agglomeration of the particles, resulting in uniformly dispersed nanoparticles. This method is low-cost, easy to operate, and has a mild reaction process, making it suitable for industrial production.
Hou et al. [55] prepared MoS2 nanoparticles using a simple method. They dissolved Na2MoO4·H2O and CH3CSNH2 in deionized water, added a surfactant, and adjusted the pH to 1.0 using H2SO4. The mixture was stirred for 5 min at 90 °C, resulting in the formation of MoS3. The product was washed, dried, and desulphurized at 800 °C under a hydrogen atmosphere, which led to the formation of MoS2 nanoparticles with a size range of 150–300 nm. Hu et al. [56,57] utilized thioacetamide (CH3CSNH2, TAA) as the sulfur source and dissolved it in distilled water along with Na2MoO4·2H2O. The resulting solution was heated to 82 °C for 5 min, then HCl or H2SO4 was rapidly poured into the reaction system and the reaction was continued for 5 min. The product was then washed and dried to give the desired MoS3 precursor. Hu and colleagues [58,59] utilized Na2S·9H2O as a source of sulfur, which was then dissolved in deionized water along with Na2MoO4·2H2O. While stirring the mixture at room temperature, they added HCl or H2SO4 dropwise to the reaction system. The obtained precipitate was washed and dried to obtain the desired MoS3 precursor. The MoS3 precursors obtained under both conditions were subjected to calcination under a high-purity hydrogen atmosphere at a selected temperature for 50 min to obtain MoS2 nanoparticles. It was also suggested that the sulfur source has a significant effect on the morphology and size of MoS3 precursors [38], and the MoS3 precursors prepared by the TAA sulfur source consisted of hollow nanorods and solid nanoparticles, while those prepared by the Na2S sulfur source consisted of micrometer-sized particles (Figure 2D). The MoS3 hollow nanorods and solid nanoparticles obtained by calcination at 780 °C under H2 yielded ~150 nm MoS2 hollow nanospheres and ~40 nm solid nanoparticles, respectively, and the MoS3 micrometer-sized particles prepared by calcination at 780 °C produced ~30 nm MoS2 nanosheets.
Xu et al. [60] were able to successfully synthesize spherical nano-MoS2 particles using an improved method based on the original synthesis method in the literature [55]. The improvements made included optimizing the ratios of reactants, lowering the calcination temperature to 480 °C, and substituting H2 with N2. This resulted in the formation of spherical nano-MoS2 particles with a solid structure, as opposed to hollow ones.

3. Stability of MoS2 Nanomaterial Dispersion

In recent years, MoS2 nanoparticles have received increasing attention as lubricant additives for their excellent tribological properties. In practice, poor dispersion of the additive causes agglomeration and precipitation, which can adversely affect the friction system. However, due to the small particle size of MoS2 nanoparticles with high surface energy, such nanoparticles have a strong tendency to agglomerate and are prone to aggregate and form agglomerates with larger sizes, thus making it difficult to form a stable lubricant suspension, which has become a challenging problem for the application of nano-MoS2 as a lubricant additive [61,62,63], seriously hindering the application of nano-MoS2 materials in practice. Therefore, optimizing the dispersion stability of MoS2 in oil solutions and slowing the tendency of particle agglomeration in suspensions can be regarded as prerequisites for studying other properties. Currently, methods for improving the dispersion stability of MoS2 nanomaterials in base oils are divided into two main categories: physical and chemical methods.

3.1. Physical Methods

Physical dispersion of nanomaterials in the base lubrication medium can be achieved by mechanical techniques such as mechanical agitation, ultrasonic treatment [64,65,66,67], ball milling [68,69], and high-pressure homogenization [70], with ultrasonic treatment being one of the most commonly used physical methods.
Ultrasonic dispersion is a relatively mature technology, which is used in textile, printing and dyeing, chemical, biological, pharmaceutical, and many other industries and fields. The mechanism of ultrasonic dispersion occurs when ultrasonic waves propagate in a dispersing medium, wherein the pressure by the speed of sound oscillatory changes, which produces a series of rapid compressions of the liquid so that the suspended particles are instantly broken and dispersed. However, overheated ultrasonic mixing should also be avoided, as the chance of particle collision increases with the increase in mechanical energy in the thermal phase, leading instead to further agglomeration. It was mentioned earlier that ultrasonic treatment helps in stripping the bulk material into nanoscale products without any other reactants [33,34] and also helps in homogeneous dispersion of MoS2 material in solution or solvent [71,72].
Alazemi et al. [73] dispersed CS-MoS2 particles in reference oil at different weight percent concentrations using a 750 W piezoelectric ultrasonic probe at a power amplitude of 20% for 120 s. The particles were then dispersed in a reference oil. Visual inspection of the results showed that using the ultrasound-mediated technique, CS-MoS2 particles could remain suspended in the reference oil for a long period (about 2 weeks) even without the addition of surfactant. Xu et al. [74] dispersed graphene and MoS2 in esterified bio-oil (EBO) with different mass ratios and treated them with ultrasonic oscillation at room temperature for 20 min. The dispersion stability of EBO containing 0.5 wt% of the copolymer was better and remained stable for at least 240 h after ultrasonic dispersion. Usually, the combination of the two methods gives better results. Wang et al. [75] homogeneously dispersed modified reduced graphene oxide and modified MoS2 in base oil (PAO6) by ultrasonic treatment and microwave-assisted ball milling (Figure 3).
Ultrasonic dispersion technology is a simple and efficient dispersion method with easy operation, fast speed, and no pollution, but insufficient ultrasonic time cannot completely break the inter-particle agglomeration, while too long ultrasonic time will produce re-agglomeration among nanoparticles, thus weakening the dispersion performance of the solution, and the high cost of ultrasonic equipment is prone to cause a drastic increase in temperature in the process of use, which in turn leads to the dispersion of the particles for further agglomeration. In addition, the difficulty of acoustic energy transfer and heat dissipation in large-scale operations affects its industrial application.

3.2. Chemical Methods

Graphene-like topic nanomaterials cannot be dissolved in most solvents due to the strong π-π bonds, and neither direct dispersion nor simple ultrasonic dispersion can make MoS2 stably dispersed in lubricating oils. Currently, surface modification or composite materials with other materials are commonly used to improve the dispersion of MoS2 in lubricating oils.
Adding surfactants as dispersants is a relatively easy method. Organic dispersants are adsorbed by these particles suitably to create a state of repulsion between the particles, thus counteracting the natural attraction between the particles [76]. There are many common dispersants such as Span 80, tween 80, oleic acid (OA), sodium dodecyl sulfate sodium dodecyl benzene sulfonate, etc. Hou et al. [77] investigated the effect of different dispersants (OA and Span 80) on the dispersion stability of MoS2 nanomaterials in PAO6 to provide a theoretical basis for the stability of the MoS2 nano-lubricant. Oleic acid and Span 80 successfully modified the MoS2 nanomaterials, and the MoS2 nano-lubricant was able to remain somewhat stable at room temperature for a period of 7 days. The paper also points out that concentration, sonication time, and morphology (spherical and sheet) all have a certain effect on the dispersion stability of MoS2 in base oils as well. Gulzar et al. [78] examined the effect of MoS2 particles on the dispersion stability and anti-wear properties of PAO10 and bio-based base oils and found that poorly dispersed nanoparticles led to a partial loss of wear protection. Surfactants contributed to the suspension of MoS2 particles in both base oils and the corresponding suspensions were found to give more favorable anti-wear. Physical adsorption of surfactants on the particle surface reduces the surface tension of the particles, thus preventing the formation of aggregates [79]. In addition, surfactant-coated additives can overcome the van der Waals gravity between them through electrostatic/spatial repulsion [80]. However, surfactant molecules are subject to shedding, decomposition, and degradation during friction due to shear stresses and localized high temperatures at the contact interface, resulting in unstable friction properties [81].
Surface modification is a commonly used method for MoS2 surface modification. Through surface grafting modification, specific functional groups in surfactants are grafted on the surface of nanoparticles to improve their oil solubility, so that the nanoparticles can be stably dispersed in lubricating oils, which has become an important direction in the research on the practicalization of nano-lubricant additives [82,83,84]. OA is an unsaturated fatty acid, which is often used as a chemical modifier for MoS2 [85,86]. Wu et al. [85] successfully obtained OA-MoS2 by one-step modification of homemade five-layer-thickness MoS2 nanosheets with oleic acid (Figure 4A). The OA molecules chemically confined the MoS2 nanosheets through C-S bonding, which significantly improved the dispersion and stability of the MoS2 nanosheets in the base oil. The results of four-sphere machine tests showed that the addition of OA-MoS2 nanosheets significantly improved the friction reduction, anti-wear, and extreme pressure capabilities of the base oil. Kumari et al. [87] chemically functionalized MoS2 using octadecanethiol (ODT). Structural defects and empty sulfur sites on the Mo atoms of MoS2 nanosheets were the targets for the grafting of ODT via Mo-S coordination bonds (Figure 4B). Van der Waals interactions between the octadecyl chains of MoS2-ODT and the oleate chains of polyol ester lubricants facilitated MoS2-ODT nanosheet dispersion (Figure 4C).
Meng et al. [88], in order to mitigate the agglomeration and deposition of MoS2 nanofluid, conducted an experiment wherein surface modification of MoS2 nanosheets was carried out by using homemade surface modifiers based on triethanolamine (TEA) and stearic acid (SA), and modified MoS2 nanosheets with better dispersion stability were obtained (Figure 4D). The effects of the surface modification on the tribological properties of the modified MoS2 nanofluids were investigated using a four-ball friction and wear tester and molecular dynamics simulation. The diffusion rate, interlayer spacing, interlayer adsorption energy, and shear stress of MoS2 nanosheets before and after modification were calculated. The interlayer spacing of the modified MoS2 nanosheets increased due to the interaction between the MoS2 nanosheets and the surface modifier. Especially at atmospheric pressure, the interlayer adsorption energy and interlayer shear stress were decreased. The modified MoS2 nanosheets not only have better dispersion stability but also have excellent tribological properties.
Due to the strong attraction between metal atoms and polar functional groups (-OH, -O-, -NH2, etc.) in MoS2, it helps to graft surfactant molecules containing such groups onto the nanoparticles [89,90,91], which not only reduces the agglomeration of nanoparticles but also avoids the separation of the nanoparticles and grafted functional groups due to frictional stress, which effectively improves the dispersive stabilization and lubrication properties of the nanoparticles. With the deepening of the research, a single decentralized method can no longer meet the demand, and the combination of methods has become the choice of many researchers.
The composite material realizes the result of “1 + 1 > 2” for two or more materials. The composite of MoS2 with a dibasic material to improve its dispersion in the base oil and then improve its lubrication performance is also one of the research hotspots of MoS2 in recent years.
The hexagonal boron nitride (h-BN) nanosheets prepared by Kumari et al. [92] via alkali-assisted hydrothermal stripping were used for the growth of a MoS2 sheet by chemical reduction in the presence of cetyltrimethylammonium bromide (CTAB). The CTA molecules on the surface of the composite nanomaterials avoided their re-stacking through the spatial site resistance of the long alkyl chains, and furthermore, the alkyl chains of the CTA molecules in the h-BN-MoS2-CTAB had van der Waals interactions with the hydrocarbons in the 5W30 engine oil, resulting in the formation of a heterogeneous structure of h-BN-MoS2-CTAB. Thus, the dispersion of h-BN-MoS2-CTAB in 5W30 engine oil was stable for one month (Figure 5A). Farsadi et al. [93] complexed functionalized reduced graphene oxide (FrGO) with MoS2 to improve its dispersion stability in base oils. MoS2-loaded FrGO was better dispersed compared to other materials and remained uniformly distributed over a month. Guan et al. [94] synthesized magnesium silicate hydroxide (MSH)-enhanced MoS2 hybrid nanomaterials using a hydrothermal method, and MSH-MoS2 showed good dispersion stability due to its abundant active anchor sites. Chouhan et al. [95] synthesized ZnO-modified reduced graphene oxide/MoS2 (Gr-MS-Zn) nanosheets and the obtained Gr-MS-Zn was formed by three to seven molecular layers of MoS2 nanosheets fully distributed on the graphene backbone through weak interfacial interactions. The curved and bent structure of MoS2 nanosheets grown on graphene sheets subsidized the cohesive interactions, and these MoS2 nanosheets wrinkled on graphene nanosheets to promote dispersion by minimizing the contact or adhesion interactions between the nanocomposites, thus reducing the interactions between the Gr-MS-Zn components, which was important for inhibiting the formation of agglomerates and maintaining their good dispersion in the fully formulated 10W40 engine oil, which is in turn important to obtain a good dispersion of Gr-MS-Zn in the final stable dispersion of the fully formulated engine oil (Figure 5B).
Therefore, although physical methods are one of the simplest and most economical ways to disperse MoS2 into lubricants, the nanoparticles are prone to re-agglomerate due to degradation problems under frictional conditions. Surfactant-modified MoS2 has better dispersion properties, but the presence of other atoms or functional groups may lead to significant degradation of its intrinsic properties. The composites obtained by compositing with other two-dimensional nanomaterials not only obtain good stability of the products, but also can choose the composite two-dimensional materials according to the need for their own defects to make up for the corresponding defects, and at the same time, the composite nanoparticles are able to realize self-dispersion in lubricating oils, which is a very promising method.

4. Tribological Behavior and Lubrication Mechanism of MoS2 Nanomaterials

MoS2 has excellent mechanical and physical properties, remarkable chemical stability, and a unique layered structure, showing its great potential for application as a lubricant additive [96,97,98]. MoS2 nanoparticles and their composites were prepared and dispersed into lubricating oils, and their tribological behavior was investigated on a friction tester. In the course of this study, the lubrication mechanism was illustrated by a number of modern surface analysis tools.

4.1. Tribological Properties of Nano-MoS2

It is widely recognized that the tribological properties of 2D nanomaterials are closely related to their size and morphology [99,100,101,102]. In recent years, with the continuous development of the synthesis process, researchers have prepared MoS2 particles with various morphologies, such as nanoflower [103], nanospherical [104], hollow core-shell [105], nanorod [106], and nanowire [107].
Hu et al. [99] investigated the effect of nanospheres, nanosheets, and bulk 2H-MoS2 additives on the tribological properties of liquid paraffin (LP) using a four-ball friction and wear tester. The results showed that all the MoS2 additives used could improve the tribological properties of LP, and the lubrication effect of nano-MoS2 particles in LP was better than that of micrometer MoS2 particles. The LP containing nanospheres had the best friction reduction and anti-wear properties when the MoS2 content was 1.5 wt%. This was attributed to the chemical stability of the layer-closed spherical structure of the nanospheres. This study further deepens the understanding of the relationship between the tribological properties and morphology of MoS2. Tontini et al. [103] synthesized nanoflower-like structured MoS2 particles with an average diameter of 250 nm and evaluated their suitability as lubricant additives. In order to ensure good stability of the nano-oils, the particles were dehydrated and lipophilized by solvent heat treatment with anhydrous ethanol. Tribological characterization was carried out by using a pin-disc tribometer with a reciprocating motion under impregnation. The results showed that the coefficient of friction of polyolester nano oil, on the other hand, decreased by 86%. Luo et al. [104] developed a new green laser-assisted solution growth of ideally spherical fullerene-like MoS2, where the extreme non-equilibrium conditions generated by laser irradiation can reconfigure MoS2 nanosheets into perfect solid nanorods with layer-closed structures to release the high surface tension energy of the nanosheets. Such MoS2 nanospheres are effective as additives to paraffinic fluids to reduce the coefficient of friction (~47% reduction) and enhance the extreme pressure performance (>2.24 GPa). This excellent lubrication performance may be attributed to the molecular bearing-like rolling effect of the fullerene-like MoS2 nanospheres and the formation of friction film between mechanical contact surfaces (Figure 6A). This study is important for the advancement of nano microspheres as additives for internal combustion engine oils in preserving mechanical energy and reducing mechanical failures due to wear. Xu et al. [105] prepared hollow core-shell MoS2 nanoparticles with flower-like surfaces using a two-step solvent-thermal method (Figure 6B), and the friction and wear performance of the prepared hollow core-shell MoS2 nanoparticles in oil was investigated by a ball-and-disk friction and wear tester. The results show that the prepared MoS2 nanoparticles can significantly improve the friction reduction and anti-wear performance of lubricating oils. The friction coefficient was reduced by 43.80% and the wear was reduced by eight times after the addition of hollow core MoS2 to the lubricating oil.

4.2. Tribological Properties of MoS2 Composites

When single MoS2 particles are used as lubrication additives, the MoS2-based friction film adsorbed on the contact surface of the friction pair due to the experimental process is easily destroyed under certain extreme friction conditions and easily removed from the friction interface. A feasible way to improve the tribological performance of pure MoS2 additives is to develop MoS2 nanocomposites, which, as mentioned in the previous section, can not only enhance the dispersion stability of the material in the base oil but also realize the synergistic enhancement effect during the friction process.
Carbon-based solid lubrication materials have been one of the focuses of research in the field of new materials for friction reduction and anti-wear, with broad application value and prospects. Scholars at home and abroad have shown strong interest in the research of carbon materials and MoS2 composites, especially graphene. Xu et al. [108] outlined the development of graphene/MoS2 nanocomposites in recent years, and discussed the synthesis method, dispersion behavior, tribological properties, and lubrication mechanism of the composites, and put forward the challenges that graphene/MoS2 nanocomposites will face in the field of tribology. It is also pointed out that modified graphene/MoS2 nanocomposites have good prospects in the field of tribology. Hu et al. [109] prepared rice husk charcoal/MoS2 (RHC/MoS2) nanoparticles by the precipitation method and modified the synthesized products with surface modifiers. The tribological properties of the nanoparticles in polyethylene glycol (PEG) before and after modification were tested by a ball-and-disc friction and wear tester. When the experimental tests were conducted at 80 N and 300 rpm for 30 min, the nanoparticle additions before and after modification were 0.5 wt%, and the wear rates were reduced by 88.6% and 97.8%, respectively, compared with pure PEG. The modified RHC/MoS2 particles are more effective in improving the anti-wear and friction reduction performance of PEG. Gong et al. [110] synthesized MoS2 nanoparticles grown on three carbon materials, i.e., carbon nanotubes (MoS2@CNT), graphene (MoS2@Gr), and fullerene C60 (MoS2@C60), using the solvothermal method. Incorporation of the three composites as lubricant additives into polyalkylene glycol (PAG) base oils showed significantly improved friction reduction and anti-wear behavior compared to PAG containing CNT, Gr, C60, and MoS2 as well as pure PAG (Figure 7). Xie et al. [41] prepared a novel carbon sphere (CS)@metallographic MoS2 (1 T-MoS2) hybrid using a one-step hydrothermal method. They investigated the effect of this complex on the friction and wear performance of PEG 200 in the MM-W1A friction and wear tester to investigate the effect of the complex on the tribological properties of PEG 200. When 0.3125 wt% CS@ 1T-MoS2 was added, the coefficient of friction was reduced by 43% compared to pure PEG 200. The wear diameter was significantly reduced from 1860 µm in pure base oil to 880 µm.
Combinations with other nanomaterials are still remarkable, such as h-BN [92,111], graphitic phase carbon nitride (g-C3N4) [112], and MSH [94,113]. Kumari et al. [92] used the design idea of generating ultra-low friction at the interface of two asymmetrically stacked 2D/2D heterostructures with different lattices to prepare h-BN-MoS2-CTAB by alkali-assisted hydrothermal stripping. h-BN-MoS2-CTAB, which was added to the engine oil (5W30) as a 30 ppm compound, resulted in a reduction of friction and wear of a steel friction pair by 44% and 96%, respectively (Figure 8A). The 2D/2D heterostructure in this study could be a revolutionary material for the development of a new generation of lubricants. Min et al. [112] successfully synthesized covalently bonded g-C3N4/MoS2 nanocomposites using a one-step hydrothermal method. g-C3N4/MoS2 nanocomposites showed an average coefficient of friction and wear rate of 27.86% and 70.87% lower than that of the pure oil at a concentration of 0.20 wt%. Guan et al. [94] synthesized MSH-reinforced MoS2 hybrid nanomaterials into PAO by a hydrothermal method at 220 °C. Tribological properties were tested by a four-ball machine, and the anti-friction performance, wear resistance, and final non-seize load of the friction pair under this lubricant were increased by 44.8%, 41.2%, and 116.6%, respectively. Guan et al. [113] also evaluated the tribological properties and wear surface repair performance of MSH-MoS2 nanocomposites as lubricant additives in PAO by a reciprocating ball-and-disk tester. The results of the tribological tests showed that the anti-friction performance and wear resistance of the friction pair were improved by 27.7% and 37.4%, respectively, with the addition of 1.0 wt% MSH-MoS2 in PAO (Figure 8C,D). What is more, the wear of the friction disk could be repaired by 33.2% after 5 h of sliding under the lubrication of PAO and MSH-MoS2.

4.3. Lubrication Mechanism

Understanding the lubrication mechanism of MoS2 additives is crucial for enhancing the preparation of these additives and for their application in the field of lubrication. There has been a considerable amount of research conducted by scholars from all over the world on the lubrication mechanism of MoS2 particles, but a unified conclusion has not yet been reached. Based on the previous studies of the tribological properties of MoS2 additives, the most widely accepted lubrication mechanism can be summarized into four aspects: rolling mechanism, shear slip, formation of friction film, and synergistic lubrication.

4.3.1. Rolling Mechanism

According to researchers, spherical nanoparticles can roll between two sliding surfaces and help reduce friction and wear. Stabilized spherical nanoparticles can also improve extreme pressure performance and lubricant load-carrying capacity [114,115,116]. The rolling effect of spherical nanoparticles depends on the thickness of the film. When the nanoparticle diameter is close to the thickness of the membrane, the shape of the nanoparticles will remain unchanged and the ball-bearing mechanism will dominate [117,118].
When the nanoparticles are spherical in shape, they can act as a kind of ball-bearing during friction [119,120,121], thus improving the lubrication performance. Alazemi et al. [73] synthesized CS-MoS2 particles and added them to engine oil (SAE 5W30) and studied the tribological properties of the lubricant blend using a friction and wear meter. When 1 wt % CS-MoS2 particles were kindly added to the engine, there was a significant reduction in friction and wear (15–35%) compared to the original reference oil at various disk speeds. Raman spectroscopic investigations of wear scars after tribological tests showed that the CS-MoS2 particles were highly chemically stable. The improved tribological performance of the CS-MoS2 and engine oil blend lubricant was attributed to the fact that the CS-MoS2 particles prevented direct contact between the sliding surfaces and acted as particulate ball bearings on the nanoscale. Meng et al. [122] prepared MoS2 quantum dots stably dispersed in paraffin oil, and the spherical MoS2 quantum dots have a proper ball-bearing lubrication effect during friction. Luo et al. [123] prepared a lamellar composite structure consisting of ultra-smooth MoS2 submicroscopic spheres embedded in a multilayer graphene. The ultra-smooth MoS2 spheres within this layered structure can significantly reduce friction by transforming sliding friction into rolling friction under the strong shear force generated by the moving contact surface.

4.3.2. Shear Slip

MoS2 has good lubricating properties due to its unique crystal structure. Within each molecular layer of MoS2, sulfur atoms are strongly bonded to molybdenum atoms through covalent bonds, while sulfur atoms between the layers are connected to molybdenum atoms through weak van der Waals forces. As a result, a low-shear plane is formed. When a small amount of shear force is applied between the molecules, the molecular layer breaks easily and a slip plane is created. For instance, a MoS2 surface film with a thickness of 0.5 μm contains 800 molecular layers and 799 slip planes. The numerous slip surfaces transform the original relative slip of the two metal surfaces in direct contact into the relative slip of the MoS2 molecular layer, which reduces the friction factor and wear.
Xie et al. [124] pointed out that the tribological mechanism of lamellar nanocomposites is different from that of spherical nanocomposites. They concluded that the lamellar nanocomposites were able to provide lower COF and lower wear rate due to the shearing effect of weak van der Waals bonds between molecular layers. Compared to spherical nanocomposites, sheet nanocomposites are less likely to roll on the friction surface. Meng et al. [125] analyzed the lubrication mechanism using molecular dynamics simulations, which showed that an increase in normal pressure promotes interlayer slip and reduces the friction coefficient between the MoS2 nanosheets. Wang et al. [126] used alkylamine-grafted MoS2 oxide nanosheets (MoS2-O-OLA) dispersed in 15# industrial white oil for friction experiments, confirming that MoS2-O-OLA nanosheets exhibited better friction and wear resistance, resulting in a significant reduction of friction (36.2%) and wear (22.4%) by 0.02 wt% MoS2-O-OLA nanosheets as compared to pure 15# industrial white oil. MoS2-O-OLA nanosheets played an important role in improving tribological properties by generating interlayer slip at the steel ball contact interface, resulting in surface protection and a uniform oil film.
For rGO/MoS2 nanocomposites [127], when their concentration is low and their size is smaller than the oil film thickness, they can form interlayer slips between the friction interfaces and provide friction reduction. However, once the concentration is too high, the excess nanocomposites will agglomerate in the base oil, and this interlayer slip to reduce friction is more difficult to realize.

4.3.3. Tribofilm Formation

Numerous studies have shown that nanoparticles tend to migrate toward the friction surface during the friction process due to two factors. One is that the localized high temperature caused by friction may cause the undulation of nanoparticles on the friction surface to be greater than that of nanoparticles in the nanofluid. Due to this undulation, random migration is more active and ultimately increases the chances of nanoparticles migrating to the friction surface. Secondly, the acceleration of escaping electrons generated during friction enhances the electric field at the friction surface, the presence of which may lead to the aggregation of nanoparticles at the friction surface [128]. Nanoparticles that flow to the friction contact surface during oil lubrication are highly susceptible to friction chemical reactions under stress and friction heat, forming a friction-responsive film in the friction region [67,129,130].
Wu et al. [131] investigated the effect of different percentages of polyisobutylene amine succinimide (PIBS) on the tribological properties of MoS2 nanosheets (Figure 9A). At low percentages of PIBS and without any PIBS, the nanosheet aggregates gathered in front of the contact area tend to enter the contact area and form a uniformly distributed friction film in the contact area, effectively reducing the coefficient of friction (COF) and the amount of wear. When lubricated with a high percentage of PIBS, the discrete particles tend to move and spread around/away from the contact region due to lateral flow, and thus the particles cannot enter the contact region resulting in poor lubrication performance. To investigate the friction film source of MoS2 nanoparticles as lubricant additives, Wu et al. [132] investigated the effect of MoS2 nanoparticles on the oil film thickness of PAO4 and 0W20 engine oils by examining the effect of MoS2 nanoparticles on the oil film thickness of PAO4 and 0W20 engine oils as well as by observing the particle flow pattern (Figure 9B). The expected increase in oil film thickness, from 30 nm to 60 nm in 15 min, was obtained with the addition of 0.1 wt% MoS2 to PAO4 base oil, whereas during the test period using 0.1 wt% MoS2 blended with 3 wt% PIBS as a dispersant in PAO4 base oil and 0.75 wt% MoS2 in 0W20 engine oil, no increase in oil film thickness and particle aggregation were observed. The results suggest that the nanoparticles responsible for friction film formation originate from particle aggregation rather than well-dispersed nanoparticles in point contact. This understanding should contribute to advances in the design of novel lubricant additives.
Xu et al. [133] used Raman microscopy (Raman) and atomic force microscopy (AFM) to characterize the friction behavior and friction film formation and removal of MoDTC/ZDDP (Figure 9C). Friction tests were combined with a collection of non-in situ Raman intensity maps to analyze MoS2 friction film accumulation. The rates of MoS2 removal from friction films obtained at different temperatures indicate that MoS2 friction films are more readily removed from friction contacts compared to anti-wear ZDDP friction films. In addition, new insights into the link between MoS2 film formation and friction properties are provided: the formation of MoS2 films in localized friction contacts leads to an immediate reduction of microscopic friction, which is terminated as soon as the amount of MoS2 reaches a so-called “saturation value”. In the “saturation zone”, a balance between the formation and removal of MoS2 is achieved. This is the first study to relate MoS2 content and coverage to friction behavior and provides a basis for the development of numerical models that can predict friction by taking into account friction chemical processes.
Tribofilm formation is currently the most popular theory to explain the anti-wear and friction reduction properties of nanocomposites. Most researchers attribute the excellent lubrication properties of nanoparticle lubricants to the fact that a friction film is formed on the friction surface during the friction process.

4.3.4. Cooperative Lubrication

The lubrication ability of nano MoS2 can be enhanced by utilizing the synergistic effect of nano MoS2 with other materials [124,134,135]. MSH-reinforced MoS2 hybrid nanocomposites synthesized by Guan et al. [94] as lubricant additives produced a strong two-layer friction film on the friction surface during sliding due to the synergistic effect of MSH and MoS2 (shown in Figure 10); the upper layer was rich in MoS2, which had an excellent friction-reducing effect, and the neighboring layer, which consisted of the reaction products of the MSH and the lubricant additives with the substrate, strongly ensured the stability of the first layer, which resulted in the reduction of friction and wear. Hu et al. [136] found that composite MoS2 additives with different morphologies (MoS2 nanorods and nanosheets) improved the wear resistance and friction reduction of liquid paraffin waxes more than those used individually due to the presence of synergistic lubrication between the two different morphologies of MoS2 particles. Hu’s team also investigated the synergistic lubrication of a series of material-loaded MoS2 complexes during friction, including sericite (DHSM/SM) [137], montmorillonite (MoS2/K10) [138], and fly ash (MoS2/FA) [139], in which MoS2/FA lubricant, in addition to offering strong lubrication performance, featured complexes that contain about 71% of low-cost FA, and it is of great environmental significance to obtain high-value-added products from FA and thus to turn waste into treasure.

4.3.5. Other Mechanisms

Apart from the aforementioned mechanisms, other mechanisms have also been demonstrated such as repair and inlay, elastic deformation, and stripping. During the friction process, nanoparticles fill and repair the microcracks on the surface by desorption or deposition, which eventually results in a smoother and more even friction surface. It is worth noting that the “self-healing” mechanism is not just about the accumulation of nanoparticles on the friction surface. As the size of the nanoparticles reduces, their melting point also decreases drastically. Under the high temperature of the friction surface, these nanoparticles can easily melt or sinter in the microcracks in the contact zone, forming ordered nanoparticles and closely bonding with the friction surface [128]. Yi et al. [140] used a molecular dynamics approach to construct rough surfaces and multilevel dimpled structures to simulate oil-poor lubrication before oil film rupture, and the simulation results showed that multilayered MoS2 acted as a load-bearing agent at light loads or low speeds and went into the grooves to repair the surfaces at heavy loads or high speeds. Yi et al. [141] used a solvent/hydrothermal method to prepare MoS2 particles in three morphologies: flower-like, microspheres, and nanosheets. The tribological behavior of the synthesized MoS2 particles as liquid paraffin lubrication additives was investigated using a ball and disc friction tester. It is noted in the paper that under heavy loads, MoS2 nanosheets exhibit superior friction and wear reduction behavior compared to the other two types of MoS2 particles. The reason for the optimal lubrication performance of nanosheet MoS2 is discussed based on their exfoliation into ultrathin nanosheets and nanofragments in friction tests.

4.3.6. Diversification of Lubrication Mechanism

It is worth noting that in most practical friction processes, nanoparticles do not act through a single lubrication mechanism, but rather two or more mechanisms act simultaneously [128]. The combination of different mechanisms leads to some extent to the complexity of mechanism studies. Kumari et al. [87] showed that when fully dispersed MoS2-ODT was used as an additive in polyol lubricant base oils, the uninterrupted supply of MoS2-ODT nanosheets on the friction contact surfaces resulted in the formation of MoS2 films at the friction interface, and the easy shear driven by weak van der Waals interactions between the MoS2 lamellae and the high mechanical strength of MoS2 combined to improve friction performance. Zhang et al. [142] prepared carbon ball surface-coated MoS2 (C@MoS2) by a hydrothermal method to realize the unification of the lubrication characteristics of carbon balls (CS) with the high load carrying capacity and low shear friction performance of MoS2. Then, 0.5 wt% of C@MoS2 particles were dispersed into PAO40 to enhance the anti-wear and friction reduction capability. The excellent tribological properties of C@MoS2 particles were attributed to the filling effect of CS and the synergistic lubrication between bearings as well as the friction reduction capability of the MoS2 shell layer. In this study, the synergistic lubrication mechanism of C@MoS2 can be categorized into different stages (shown in Figure 11): (1) shell layer fragmentation stage, which corresponds to the rolling bearing effect of C@MoS2 particles at the friction interface; (2) core/shell separation stage, which corresponds to the rolling bearing effect and the self-repairing effect of the C@MoS2 particles; and (3) core/shell composite stage, which corresponds to the formation of the protective film and the enhancement in which the friction and wear performance is substantially improved.

5. Summary and Outlook

This paper reviews the notable research progress in the field of MoS2 nanomaterials as lubricant additives in recent years, focusing on the synthesis strategy, structure, dispersion stability, tribological performance evaluation, and lubrication mechanism of MoS2 nanomaterials. Dispersion stability is the key to the research and application of graphene lubricant additives. There are physical and chemical methods to improve the dispersion stability of MoS2 nanomaterials in lubricants, in which the dispersion, stability, and tribological properties can be easily adjusted and improved by combining MoS2 with other materials, further expanding the application of MoS2 nanomaterials in tribology. Various lubrication mechanisms such as interlayer sliding, ball action, formation of friction film, and synergistic action, which are currently recognized, are outlined, and the excellent tribological properties of MoS2 nanomaterials are explained in depth.
Although significant breakthroughs and developments have been made in the research on MoS2 nanomaterials as lubricant additives, there are still some difficulties and challenges to overcome as follows.
(1)
The dispersion stability of MoS2 nanomaterials in lubricating oils has not been adequately addressed. During the friction process, the organic modifier is prone to degradation due to the heat generated by friction, which leads to the re-colonization of MoS2 nanoparticles in the lubricating oil. Therefore, research on the long-term dispersion stability of MoS2 nanomaterials is still necessary.
(2)
In terms of lubrication mechanism, there are fewer systematic descriptions of the lubrication mechanism of MoS2 nanomaterials with different morphologies in lubricants. The influence of each component on the tribological performance is still unclear, such as the interaction between additive materials and lubricants, the synergistic lubrication of each component in composite nanomaterials, and the interaction between nanomaterials and modifiers. It is necessary to further elucidate the relevant mechanisms by combining advanced characterization techniques, molecular dynamics simulations, and theoretical calculations.
(3)
Most reported additives for MoS2-based nanomaterials have been studied under laboratory conditions. In the future, there is a need to develop new additives with good tribological properties under extreme conditions or multiple environments. In addition, low-cost, large-scale preparation routes and evaluation of tribological properties in practical applications are essential for the practical application of these additives.
To promote the use of nanomaterials as lubricant additives, the following aspects should be considered for future research, taking into account future developments and trends in MoS2 lubricant additives:
First, it is important to prepare high-quality MoS2 through a green, economical, and scaled-up production method. Combined with modern computer simulation methods, it can not only reduce the unnecessary waste of resources but also optimize the design of MoS2 nanoparticles with good structure and obtain better performance. Secondly, dispersion stability is very important as a prerequisite for practical applications. However, there are still difficulties in improving the dispersion stability of nanomaterials for practical industrial applications. For example, simple physics is susceptible to reaggregation. Chemical modifications may mask the original properties and chemical modifiers tend to degrade during friction. Therefore, further research is needed to improve dispersion methods. The self-dispersion method does not require any modifier and has good dispersion stability. The stability of the prepared MoS2 and MoS2-based composite nanomaterials dispersed in the lubricant is crucial for long-term enhancement of tribological properties. Finally, with the rapid development of advanced industrial equipment, many advanced mechanical devices need to meet extreme working conditions such as high load, high temperature, and high speed. It is important to study new lubricant additives for extreme friction conditions. Many highly efficient lubricant additives decompose at 150 °C and have a flash point not exceeding 100 °C. The application of lubricant additives is still limited to temperatures of 1000 °C and loads of 3000 N. Therefore, it is of great significance to further improve the performance of MoS2 and MoS2-based composites under extreme conditions to overcome the bottleneck of friction and wear inefficiency.
In future research, it is believed that with the efforts of a large number of researchers, MoS2 nanomaterials will make surprising breakthroughs and progress and realize the expected concepts.

Author Contributions

Conceptualization, Z.L., K.H. and X.H.; methodology, Z.L., Z.C. and J.G.; investigation, Z.C., J.G. and W.L.; resources, X.H.; data curation, Z.L. and Y.W. writing—original draft preparation, Z.L.; writing—review and editing, Z.L., Q.L., K.H. and X.H.; supervision, K.H. and X.H.; project administration, K.H. and X.H.; funding acquisition, K.H. and X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Nature Science Foundation of China, grant number 52075141, 52075144, 52075145 and the Natural Science Foundation for Colleges and Universities in Anhui Province, grant number 2022AH010096.

Data Availability Statement

Not applicable.

Acknowledgments

Thanks to Enzhu Hu for providing corrections to the English grammar of this article.

Conflicts of Interest

The authors declare no conflict of interest.

List of Abbreviations

MoDTCMolybdenum dialkyl dithiocarbamate
ZDDPZinc dialkyl dithiophosphate
TMDsTransition metal dichalcogenides
CS/HAcchitosan/acetic acid
DMFN, N-dimethylformamide
NMPN-methylpyrrolidone
TAAThioacetamide
OAOleic acid
ODTOctadecanethiol
TEATriethanolamine
SAStearic acid
h-BNhexagonal boron nitride
CTABCetyltrimethylammonium bromide
FrGOFunctionalized reduced graphene oxide
MSHMagnesium silicate hydroxide
Gr-MS-ZnZnO-modified reduced graphene oxide/MoS2
PAOPolyalphaolefin
LPLiquid paraffin
SEMScanning electron microscope
TEMTransmission electron microscopy
RHC/MoS2Rice husk charcoal/MoS2
PEGPolyethylene glycol
CNTCarbon nanotubes
PAGPolyalkylene glycol
CS @ 1 T-MoS2carbon sphere @ metallographic MoS2
g-C3N4graphitic phase carbon nitride
MoS2-O-OLAAlkylamine-grafted MoS2 oxide
PIBSPolyisobutylene amine succinimide
COFCoefficient of friction
RamanRaman microscopy
AFMAtomic force microscopy
K10Montmorillonite K10
FAFly ash

References

  1. Holmberg, K.; Andersson, P.; Erdemir, A. Global energy consumption due to friction in passenger cars. Tribol. Int. 2012, 47, 221–234. [Google Scholar] [CrossRef]
  2. Holmberg, K.; Andersson, P.; Nylund, N.-O.; Mäkelä, K.; Erdemir, A. Global energy consumption due to friction in trucks and buses. Tribol. Int. 2014, 78, 94–114. [Google Scholar] [CrossRef]
  3. Zhao, J.; Huang, Y.; He, Y.; Shi, Y. Nanolubricant additives: A review. Friction 2020, 9, 891–917. [Google Scholar] [CrossRef]
  4. Mang, T.; Bobzin, K.; Bartels, T. Industrial Tribology: Tribosystems, Friction, Wear and Surface Engineering, Lubrication; John Wiley and Sons, Ltd.: New York, NY, USA, 2011. [Google Scholar]
  5. Mang, T.; Dresel, W. Lubricants and Lubrication; John Wiley and Sons, Ltd.: New York, NY, USA, 2007. [Google Scholar]
  6. Neville, A.; Morina, A.; Haque, T.; Voong, M. Compatibility between tribological surfaces and lubricant additives—How friction and wear reduction can be controlled by surface/lube synergies. Tribol. Int. 2007, 40, 1680–1695. [Google Scholar] [CrossRef]
  7. Umer, J.; Morris, N.; Rahmani, R.; Balakrishnan, S.; Rahnejat, H. Nanoscale frictional characterisation of base and fully formulated lubricants based on activation energy components. Tribol. Int. 2020, 144, 106115. [Google Scholar] [CrossRef]
  8. Rudnick, L.R. Lubricant Additives: Chemistry and Applications, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  9. Guo, D.; Xie, G.; Luo, J. Mechanical properties of nanoparticles: Basics and applications. J. Phys. D Appl. Phys. 2014, 47, 013001. [Google Scholar] [CrossRef]
  10. Han, Y.; Pan, L.; Zhang, H.; Zeng, Y.; Yin, Z. Effect of lubricant additives of Cu, Fe and bimetallic CuFe nanoparticles on tribological properties. Wear 2022, 508–509, 204485. [Google Scholar] [CrossRef]
  11. Zhu, Y.; Zhang, H.; Li, N.; Jiang, Z. Friction and Wear Characteristics of Fe3O4 Nano-Additive Lubricant in Micro-Rolling. Lubricants 2023, 11, 434. [Google Scholar] [CrossRef]
  12. Liu, C.; Friedman, O.; Meng, Y.; Tian, Y.; Golan, Y. CuS Nanoparticle Additives for Enhanced Ester Lubricant Performance. ACS Appl. Nano Mater. 2018, 1, 7060–7065. [Google Scholar] [CrossRef]
  13. Liu, Y.; Ge, X.; Li, J. Graphene lubrication. Appl. Mater. 2020, 20, 100662. [Google Scholar] [CrossRef]
  14. Liu, W.; Li, W.; Li, R.; Lu, Z.; Li, D.; Zhang, G.; Wu, Z. Green oil additive g-C3N4: A feasible strategy to enhance the tribological properties of DLC film. Mater. Res. Express 2019, 6, 115036. [Google Scholar] [CrossRef]
  15. Zhang, S.; Hu, L.; Feng, D.; Wang, H. Anti-wear and friction-reduction mechanism of Sn and Fe nanoparticles as additives of multialkylated cyclopentanes under vacuum condition. Vacuum 2013, 87, 75–80. [Google Scholar] [CrossRef]
  16. Xiao, H.; Liu, S. 2D nanomaterials as lubricant additive: A review. Mater. Des. 2017, 135, 319–332. [Google Scholar] [CrossRef]
  17. Uflyand, I.E.; Zhinzhilo, V.A.; Burlakova, V.E. Metal-containing nanomaterials as lubricant additives: State-of-the-art and future development. Friction 2019, 7, 93–116. [Google Scholar] [CrossRef]
  18. Bhushan, B. Introduction to Tribology; John Wiley and Sons, Ltd.: New York, NY, USA, 2013. [Google Scholar]
  19. Zhao, X.; Perry, S.S. The role of water in modifying friction within MoS2 sliding interfaces. ACS Appl. Mater. Interfaces 2010, 2, 1444–1448. [Google Scholar] [CrossRef] [PubMed]
  20. Chhowalla, M.; Amaratunga, G. Thin films of fullerene-like MoS2 nanoparticles with ultra-low friction and wear. Nature 2000, 407, 164–167. [Google Scholar] [CrossRef] [PubMed]
  21. Donnet, C.; Martin, J.M.; Le Mogne, T.; Belin, M. Super-low friction of MoS2 coatings in various environments. Tribol. Int. 1996, 29, 123–128. [Google Scholar] [CrossRef]
  22. Khare, H.; Burris, D. Surface and subsurface contributions of oxidation and moisture to room temperature friction of molybdenum disulfide. Tribol. Lett. 2014, 53, 329–336. [Google Scholar] [CrossRef]
  23. Winer, W.O. Molybdenum disulfide as a lubricant: A review of the fundamental knowledge. Wear 1967, 10, 422–452. [Google Scholar] [CrossRef]
  24. Lansdown, A.R. Molybdenum Disulphide Lubrication; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  25. Acikgoz, O.; Guerrero, E.; Yanilmaz, A.; Dagdeviren, O.; Çelebi, C.; Strubbe, D.A.; Baykara, M.Z. Intercalation leads to inverse layer dependence of friction on chemically doped MoS2. Nanotechnology 2023, 34, 015706. [Google Scholar] [CrossRef]
  26. Bojarska, Z.; Kopytowski, J.; Mazurkiewicz-Pawlicka, M.; Bazarnik, P.; Gierlotka, S.; Rożeń, A.; Makowski, Ł. Molybdenum disulfide-based hybrid materials as new types of oil additives with enhanced tribological and rheological properties. Tribol. Int. 2021, 160, 106999. [Google Scholar] [CrossRef]
  27. Singh, K.K.; Prabhu, B.R.; Choudhary, S.; Pramanik, C.; John, N.S. Effect of graphene and MoS2 flakes in industrial oils to enhance lubrication. ACS Omega 2019, 4, 14569–14578. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, Y.; Fan, X.; Yue, Z.; Zhu, M. Mechanism on heterogeneous transfer film formed by diamond-like carbon film under molybdenum disulfide hybrid polyethylene glycol lubrication. Carbon 2023, 210, 118030. [Google Scholar] [CrossRef]
  29. Toh, R.J.; Sofer, Z.; Luxa, J.; Sedmidubský, D.; Pumera, M. 3R phase of MoS2 and WS2 outperforms the corresponding 2H phase for hydrogen evolution. Chem. Commun. 2017, 53, 3054–3057. [Google Scholar] [CrossRef] [PubMed]
  30. Krishnan, U.; Kaur, M.; Singh, K.; Kumar, M.; Kumar, A. A synoptic review of MoS2: Synthesis to applications. Superlattices Microstruct. 2019, 128, 274–297. [Google Scholar] [CrossRef]
  31. Wang, Q.H.; Kourosh-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef]
  32. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.W.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef]
  33. Frindt, R.F. Single crystals of MoS2 several molecular layers thick. J. Appl. Phys. 1966, 37, 1928–1929. [Google Scholar] [CrossRef]
  34. Liu, Y.; Li, R. Study on ultrasound-assisted liquid-phase exfoliation for preparing graphene-like molybdenum disulfide nanosheets. Ultrason. Sonochem. 2020, 63, 104923. [Google Scholar] [CrossRef]
  35. Yao, Q.; Guo, J.; Guan, F.; Zhao, M.; Zhang, S.; Tuo, X.; Yang, Q. Ultrasonic-Assisted Exfoliation Bulk-Phase of MoS2 with Chitosan/Acetic Acid Solution. JOM 2023, 75, 701–707. [Google Scholar] [CrossRef]
  36. Xie, M.; Pan, B.; Liu, H.; Li, N.; Chen, Z.; Yan, J.; Fu, Z.; Guo, S.; Wang, H. One-step synthesis of carbon sphere@1T-MoS2 towards superior antiwear and lubricity. Tribol. Int. 2022, 176, 107927. [Google Scholar] [CrossRef]
  37. Zhao, L.; Hong, C.; Lin, L.; Wu, H.; Su, Y.; Zhang, X.; Liu, A. Controllable nanoscale engineering of vertically aligned MoS2 ultrathin nanosheets by nitrogen doping of 3D graphene hydrogel for improved electrocatalytic hydrogen evolution. Carbon 2017, 116, 223–231. [Google Scholar] [CrossRef]
  38. Hu, K.; Hu, X.; Jiang, P. Large-scale and morphology-controlled synthesis of nano-sized molybdenum disulphide particles by different sulphur sources. Int. J. Mater. Prod. Technol. 2010, 39, 378–387. [Google Scholar] [CrossRef]
  39. Pallikkarathodi Mani, N.; Cyriac, J. Green approach to synthesize various MoS2 nanoparticles via hydrothermal process. Bull. Mater. Sci. 2022, 45, 184. [Google Scholar] [CrossRef]
  40. Park, S.Y.; Lee, J.E.; Kim, Y.H.; Kim, J.J.; Shim, Y.-S.; Kim, S.Y.; Lee, M.H.; Jang, H.W. Room temperature humidity sensors based on rGO/MoS2 hybrid composites synthesized by hydrothermal method. Sens. Actuators B Chem. 2018, 258, 775–782. [Google Scholar] [CrossRef]
  41. Tang, G.; Zhang, J.; Liu, C.; Zhang, D.; Wang, Y.; Tang, H.; Li, C. Synthesis and tribological properties of flower-like MoS2 microspheres. Ceram. Int. 2014, 40, 11575–11580. [Google Scholar] [CrossRef]
  42. Li, J.; Wang, D.; Ma, H.; Pan, Z.; Jiang, Y.; Li, M.; Tian, Z. Ionic liquid assisted hydrothermal synthesis of hollow core/shell MoS2 microspheres. Mater. Lett. 2015, 160, 550–554. [Google Scholar] [CrossRef]
  43. Vijaya, S.; Landi, G.; Wu, J.J.; Anandan, S. MoS2 nanosheets based counter electrodes: An alternative for Pt-free dye-sensitized solar cells. Electrochim. Acta 2019, 294, 134–141. [Google Scholar] [CrossRef]
  44. Gan, Y.X.; Jayatissa, A.H.; Yu, Z.; Chen, X.; Li, M. Hydrothermal synthesis of nanomaterials. J. Nanomater. 2020, 2020, 8917013. [Google Scholar] [CrossRef]
  45. Li, L.; Qin, Z.; Ries, L.; Hong, S.; Michel, T.; Yang, J.; Salameh, C.; Bechelany, M.; Miele, P.; Kaplan, D.; et al. Role of sulfur vacancies and undercoordinated Mo regions in MoS2 nanosheets toward the evolution of hydrogen. ACS Nano 2019, 13, 6824–6834. [Google Scholar] [CrossRef]
  46. Li, X.; Feng, Z.; Zai, J.; Ma, Z.-F.; Qian, X. Incorporation of Co into MoS2/graphene nanocomposites: One effective way to enhance the cycling stability of Li/Na storage. J. Power Sources 2018, 373, 103–109. [Google Scholar] [CrossRef]
  47. He, D.; Yang, Y.; Liu, Z.; Shao, J.; Wu, J.; Wang, S.; Shen, L.; Bao, N. Solvothermal-assisted assembly of MoS2 nanocages on graphene sheets to enhance the electrochemical performance of lithium-ion battery. Nano Res. 2020, 13, 1029–1034. [Google Scholar] [CrossRef]
  48. Wang, C.; Jiang, J.; Ruan, Y.; Ao, X.; Ostrikov, K.; Zhang, W.; Lu, J.; Li, Y.Y. Construction of MoO2 quantum dot–graphene and MoS2 nanoparticle–graphene nanoarchitectures toward ultrahigh lithium storage capability. ACS Appl. Mater. Interfaces 2017, 9, 28441–28450. [Google Scholar] [CrossRef]
  49. Cao, H.; Bai, Z.; Li, Y.; Xiao, Z.; Zhang, X.; Li, G. Solvothermal Synthesis of Defect-Rich Mixed 1T-2MoS2 Nanoflowers for Enhanced Hydrodesulfurization. ACS Sustain. Chem. Eng. 2020, 8, 7343–7352. [Google Scholar] [CrossRef]
  50. Chen, J.; Xu, Z.; Hu, Y.; Yi, M. PEG-assisted solvothermal synthesis of MoS2 nanosheets with enhanced tribological property. Lubr. Sci. 2020, 32, 273–282. [Google Scholar] [CrossRef]
  51. Najafi, L.; Bellani, S.; Martín-García, B.; Oropesa-Nuñez, R.; Del Rio Castillo, A.E.; Prato, M.; Moreels, I.; Bonaccorso, F. Solution-processed hybrid graphene flake/2H-MoS2 quantum dot heterostructures for efficient electrochemical hydrogen evolution. Chem. Mater. 2017, 29, 5782–5786. [Google Scholar] [CrossRef]
  52. Liu, Y.; Zhu, Y.; Fan, X.; Wang, S.; Li, Y.; Zhang, F.; Zhang, G.; Peng, W. (0D/3D) MoS2 on porous graphene as catalysts for enhanced electrochemical hydrogen evolution. Carbon 2017, 121, 163–169. [Google Scholar] [CrossRef]
  53. Huang, H.; Huang, J.; Liu, W.; Fang, Y.; Liu, Y. Ultradispersed and single-layered MoS2 nanoflakes strongly coupled with graphene: An optimized structure with high kinetics for the hydrogen evolution reaction. ACS Appl. Mater. Interfaces 2017, 9, 39380–39390. [Google Scholar] [CrossRef]
  54. Tan, L.; Li, X.; Wang, Z.; Guo, H.; Wang, J. Lightweight Reduced Graphene Oxide@MoS2 Interlayer as Polysulfide Barrier for High-Performance Lithium–Sulfur Batteries. ACS Appl. Mater. Interfaces 2018, 10, 3707–3713. [Google Scholar] [CrossRef]
  55. Hou, S.; Li, Y.; Huo, Y.; Wu, C.; Zhang, D.; Zhang, H. Preparation of nano-scale molybdenum disulfide by liquid phase precipitation method and its lubricating properties. Ferroelectrics 2018, 524, 79–85. [Google Scholar] [CrossRef]
  56. Hu, K.H.; Wang, Y.R.; Hu, X.G.; Wo, H.Z. Preparation and characterisation of ball-like MoS2 nanoparticles. Mater. Sci. Technol. 2007, 23, 242–246. [Google Scholar] [CrossRef]
  57. Hu, K.H.; Hu, X.G.; Sun, X.J.; Jing, H.F.; Zhan, S. Synthesis and characterization of nanosize molybdenum disulfide particles by quick homogeneous precipitation method. Key Eng. Mater. 2007, 353–358, 2107–2110. [Google Scholar] [CrossRef]
  58. Hu, X.; Sun, X.; Hu, K. Preparation Parameters Optimization of Noncrystalline MoS3–The Precursor of Nano-MoS2. Solid State Phenom. 2007, 121–123, 1309–1312. [Google Scholar] [CrossRef]
  59. Hu, K.; Hu, X. Formation, exfoliation and restacking of MoS2 nanostructures. Mater. Sci. Technol. 2009, 25, 407–414. [Google Scholar] [CrossRef]
  60. Xu, Y.; Hu, E.; Hu, K.; Xu, Y.; Hu, X. Formation of an adsorption film of MoS2 nanoparticles and dioctyl sebacate on a steel surface for alleviating friction and wear. Tribol. Int. 2015, 92, 172–183. [Google Scholar] [CrossRef]
  61. Kim, D.; Archer, L.A. Nanoscale organic-inorganic hybrid lubricants. Langmuir 2011, 27, 3083–3094. [Google Scholar] [CrossRef]
  62. Bakunin, V.N.; Suslov, A.Y.; Kuzmina, G.N.; Parenago, O.P. Recent achievements in the synthesis and application of inorganic nanoparticles as lubricant components. Lubr. Sci. 2005, 17, 127–145. [Google Scholar] [CrossRef]
  63. Devendiran, D.K.; Amirtham, V.A. A review on preparation, characterization, properties and applications of nanofluids. Renew. Sustain. Energy Rev. 2016, 60, 21–40. [Google Scholar] [CrossRef]
  64. Huang, W.; Liu, W.; Wu, D.H. Investigations into lubrication in grinding processes using MWCNTs nanofluids with ultrasonic-assisted dispersion. J. Clean. Prod. 2016, 137, 1553–1559. [Google Scholar] [CrossRef]
  65. Maheswaran, R.; Sunil, J. Experimental analysis of tribological properties of ultrasonically dispersed garnet nanoparticles in SN500 grade lubricating oil. Ind. Lubr. Tribol. 2018, 70, 250–255. [Google Scholar] [CrossRef]
  66. Ashour, M.; Mohamed, A.; Elshalakany, A.B.; Osman, T.; Khatab, A. Rheological behavior of lithium grease with CNTs/GNPs hybrid nanocomposite as an additive. Ind. Lubr. Tribol. 2018, 70, 331–338. [Google Scholar] [CrossRef]
  67. Mosleh, M.; Atnafu, N.D.; Belk, J.H.; Nobles, O.M. Modification of sheet metal forming fluids with dispersed nanoparticles for improved lubrication. Wear 2009, 267, 1220–1225. [Google Scholar] [CrossRef]
  68. Li, W.; Cheng, Z.-L.; Liu, Z. Novel Preparation of Calcium Borate/Graphene Oxide Nanocomposites and Their Tribological Properties in Oil. J. Mater. Eng. Perform. 2017, 26, 285–291. [Google Scholar] [CrossRef]
  69. Shi, D.; Yang, M.; Chang, B.; Ai, Z.; Zhang, K.; Shao, Y.; Wang, S.; Wu, Y.; Hao, X. Ultrasonic-ball milling: A novel strategy to prepare large-size ultrathin 2D materials. Small 2020, 16, 1906734. [Google Scholar] [CrossRef] [PubMed]
  70. Guerra, V.; Wan, C.; Degirmenci, V.; Sloan, J.; Presvytis, D.; Watson, M.; McNally, T. Characterisation of graphite nanoplatelets (GNP) prepared at scale by highpressure homogenization. J. Mater. Chem. C 2019, 7, 6383–6390. [Google Scholar] [CrossRef]
  71. Kumar, D.P.; Hong, S.; Reddy, D.A.; Kim, T.K. Ultrathin MoS2 layers anchored exfoliated reduced graphene oxide nanosheet hybrid as a highly efficient cocatalyst for CdS nanorods towards enhanced photocatalytic hydrogen production. Appl. Catal. B 2017, 212, 7–14. [Google Scholar] [CrossRef]
  72. Upadhyay, R.K.; Kumar, A. Effect of humidity on the synergy of friction and wear properties in ternary epoxy-graphene-MoS2 composites. Carbon 2019, 146, 717–727. [Google Scholar] [CrossRef]
  73. Alazemi, A.A.; Dysart, A.D.; Phuah, X.L.; Pol, V.G.; Sadeghi, F. MoS2 nanolayer coated carbon spheres as an oil additive for enhanced tribological performance. Carbon 2016, 110, 367–377. [Google Scholar] [CrossRef]
  74. Xu, Y.; Peng, Y.; Dearn, K.D.; Zheng, X.; Yao, L.; Hu, X. Synergistic lubricating behaviors of graphene and MoS2 dispersed in esterified bio-oil for steel/steel contact. Wear 2015, 342–343, 297–309. [Google Scholar] [CrossRef]
  75. Wang, S.; Chen, D.; Chen, Y.; Zhu, K. Dispersion stability and tribological properties of additives introduced by ultrasonic and microwave assisted ball milling in oil. RSC Adv. 2020, 10, 25177–25185. [Google Scholar] [CrossRef]
  76. Sahoo, R.R.; Biswas, S.K. Deformation and friction of MoS2 particles in liquid suspensions used to lubricate sliding contact. Thin Solid Films 2010, 518, 5995–6005. [Google Scholar] [CrossRef]
  77. Hou, X.; Jiang, H.; Ali, M.K.A.; Liu, H.; Su, D.; Tian, Z. Dispersion behavior assessment of the molybdenum disulfide nanomaterials dispersed into poly alpha olefin. J. Mol. Liq. 2020, 311, 113303. [Google Scholar] [CrossRef]
  78. Gulzar, M.; Mahmood, K.; Zahid, R.; Alabdulkarem, A.; Masjuki, H.H.; Kalam, M.A.; Varman, M.; Zulkifli, N.W.M.; Ahmad, P.; Malik, M.S.S. The effect of particle size on the dispersion and wear protection ability of MoS2 particles in polyalphaolefin and trimethylolpropane ester. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 2018, 232, 987–998. [Google Scholar] [CrossRef]
  79. Paramashivaiah, B.M.; Rajashekhar, C.R. Studies on effect of various surfactants on stable dispersion of graphene nano particles in simarouba biodiesel. IOP Conf. Ser. Mater. Sci. Eng. 2016, 149, 012083. [Google Scholar] [CrossRef]
  80. Large, M.J.; Ogilvie, S.P.; Graf, A.A.; Lynch, P.J.; O’Mara, M.A.; Waters, T.; Jurewicz, I.; Salvage, J.P.; Dalton, A.B. Large-scale surfactant exfoliation of graphene and conductivity-optimized graphite enabling wireless connectivity. Adv. Mater. Technol. 2020, 5, 2000284. [Google Scholar] [CrossRef]
  81. Zhao, J.; Gao, T.; Li, Y.; He, Y.; Shi, Y. Two-dimensional (2D) graphene nanosheets as advanced lubricant additives: A critical review and prospect. Mater. Today Commun. 2021, 29, 102755. [Google Scholar] [CrossRef]
  82. Fu, Z.; Gu, X.; Hu, L.; Li, Y.; Li, J. Radiation Induced Surface Modification of Nanoparticles and Their Dispersion in the Polymer Matrix. Nanomaterials 2020, 10, 2237. [Google Scholar] [CrossRef] [PubMed]
  83. Khazaei, M.A.; Bastani, D.; Mohammadi, A.; Kordzadeh, A. Adsorption Dynamics of Surface-Modified Silica Nanoparticles at Solid–Liquid Interfaces. Langmuir 2022, 38, 12421–12431. [Google Scholar] [CrossRef] [PubMed]
  84. Li, W.; Zheng, S.; Chen, Q.; Cao, B. A new method for surface modification of TiO2/Al2O3 nanocomposites with enhanced anti-friction properties. Mater. Chem. Phys. 2012, 134, 38–42. [Google Scholar] [CrossRef]
  85. Wu, P.; Li, W.; Liu, Z.; Cheng, Z. Preparation and tribological properties of oleic acid-decorated MoS2 nanosheets with good oil dispersion. J. Dispers. Sci. Technol. 2018, 39, 1742–1751. [Google Scholar] [CrossRef]
  86. Jiang, H.; Hou, X.; Qian, Y.; Liu, H.; Ali, M.K.A.; Dearn, K.D. A tribological behavior assessment of steel contacting interface lubricated by engine oil introducing layered structural nanomaterials functionalized by oleic acid. Wear 2023, 524–525, 204675. [Google Scholar] [CrossRef]
  87. Kumari, S.; Mungse, H.P.; Gusain, R.; Kumar, N.; Sugimura, H.; Khatri, O.P. Octadecanethiol-grafted molybdenum disulfide nanosheets as oil-dispersible additive for reduction of friction and wear. FlatChem 2017, 3, 16–25. [Google Scholar] [CrossRef]
  88. Meng, Y.; Sun, J.; He, J.; Yang, F.; Wu, P. Surface modification induced improvement of dispersion stability and tribological properties of MoS2 nanosheets. J. Dispers. Sci. Technol. 2023, 44, 1010–1020. [Google Scholar] [CrossRef]
  89. Aralihalli, S.; Biswas, S.K. Crafting of dispersants on MoS2 nanoparticles in base oil lubrication of steel. Tribol. Lett. 2013, 49, 61–76. [Google Scholar] [CrossRef]
  90. Lee, J.D. Concise of Ingoranic Chemistry, 5th ed.; Oxford University Press: Oxford, UK, 2013. [Google Scholar]
  91. Sahoo, R.R.; Math, S.; Biswas, S.K. Mechanics of deformation under traction and friction of a micrometric monolithic MoS2 particle in comparison with those of an agglomerate of nanometric MoS2 particles. Tribol. Lett. 2010, 37, 239–249. [Google Scholar] [CrossRef]
  92. Kumari, S.; Chouhan, A.; Kumar Konathala, L.N.S.; Sharma, O.P.; Ray, S.S.; Ray, A.; Khatri, O.P. Chemically functionalized 2D/2D hexagonal boron Nitride/Molybdenum disulfide heterostructure for enhancement of lubrication properties. Appl. Surf. Sci. 2022, 579, 152157. [Google Scholar] [CrossRef]
  93. Farsadi, M.; Bagheri, S.; Ismail, N.A. Nanocomposite of functionalized graphene and molybdenum disulfide as friction modifier additive for lubricant. J. Mol. Liq. 2017, 244, 304–308. [Google Scholar] [CrossRef]
  94. Guan, Z.; Wu, Z.; Liu, J.; Tu, X.; Li, S. Controllable fabrication of magnesium silicate hydroxide reinforced MoS2 hybrid nanomaterials as effective lubricant additives in PAO. Appl. Surf. Sci. 2022, 597, 153777. [Google Scholar] [CrossRef]
  95. Chouhan, A.; Sarkar, T.K.; Kumari, S.; Vemuluri, S.; Khatri, O.P. Synergistic lubrication performance by incommensurately stacked ZnO-decorated reduced graphene oxide/MoS2 heterostructure. J. Colloid Interface Sci. 2020, 580, 730–739. [Google Scholar] [CrossRef]
  96. Rabaso, P.; Ville, F.; Dassenoy, F.; Diaby, M.; Afanasiev, P.; Cavoret, J.; Vacher, B.; Le Mogne, T. Boundary lubrication: Influence of the size and structure of inorganic fullerene-like MoS2 nanoparticles on friction and wear reduction. Wear 2014, 360, 161–178. [Google Scholar] [CrossRef]
  97. Jia, X.; Huang, J.; Li, Y.; Yang, J.; Song, H. Monodisperse Cu nanoparticles @ MoS2 nanosheets as a lubricant additive for improved tribological properties. Appl. Surf. Sci. 2019, 494, 430–439. [Google Scholar] [CrossRef]
  98. Song, W.; Yan, J.; Ji, H. Fabrication of GNS/MoS2 composite with different morphology and its tribological performance as a lubricant additive. Appl. Surf. Sci. 2019, 469, 226–235. [Google Scholar] [CrossRef]
  99. Hu, K.H.; Hu, X.G.; Xu, Y.F.; Huang, F.; Liu, J.S. The effect of morphology on the tribological properties of MoS2 in liquid paraffin. Tribol. Lett. 2010, 40, 155–165. [Google Scholar] [CrossRef]
  100. Uzoma, P.C.; Hu, H.; Khadem, M.; Penkov, O.V. Tribology of 2D Nanomaterials: A Review. Coatings 2020, 10, 897. [Google Scholar] [CrossRef]
  101. Wang, R.; Zhang, F.; Yang, K.; Xiong, Y.; Tang, J.; Chen, H.; Duan, M.; Li, Z.; Zhang, H.; Xiong, B. Review of two-dimensional nanomaterials in tribology: Recent developments, challenges and prospects. Adv. Colloid Interface Sci. 2023, 321, 103004. [Google Scholar] [CrossRef] [PubMed]
  102. Cho, D.H.; Wang, L.; Kim, J.S.; Lee, G.H.; Kim, E.S.; Lee, S.; Lee, S.Y.; Hone, J.; Lee, C. Effect of surface morphology on friction of graphene on various substrates. Nanoscale 2013, 5, 3063–3069. [Google Scholar] [CrossRef] [PubMed]
  103. Tontini, G.; Semione, G.D.L.; Bernardi, C.; Binder, R.; de Mello, J.D.B.; Drago, V. Synthesis of nanostructured flower-like MoS2 and its friction properties as additive in lubricating oils. Ind. Lubr. Tribol. 2016, 68, 658–664. [Google Scholar] [CrossRef]
  104. Luo, T.; Chen, X.; Wang, L.; Wang, P.; Li, C.; Zeng, H.; Cao, B. Green laser irradiation-stimulated fullerene-like MoS2 nanospheres for tribological applications. Tribol. Int. 2018, 122, 119–124. [Google Scholar] [CrossRef]
  105. Xu, W.; Fu, C.; Hu, Y.; Chen, J.; Yang, Y.; Yi, M. Synthesis of hollow core-shell MoS2 nanoparticles with enhanced lubrication performance as oil additives. Bull. Mater. Sci. 2021, 44, 88. [Google Scholar] [CrossRef]
  106. Zhang, C.; Wu, H.B.; Guo, Z.; Lou, X.W.D. Facile synthesis of carbon-coated MoS2 nanorods with enhanced lithium storage properties. Electrochem. Commun. 2012, 20, 7–10. [Google Scholar] [CrossRef]
  107. Li, W.J.; Shi, E.W.; Ko, J.M.; Chen, Z.Z.; Ogino, H.; Fukuda, T. Hydrothermal synthesis of MoS2 nanowires. J. Cryst. Growth 2003, 250, 418–422. [Google Scholar] [CrossRef]
  108. Xu, Y.; Fu, K.; Liu, K.; Sun, K.; Dong, Y.; Yao, L. A State of the Art Review of the Tribology of Graphene/MoS2 Nanocomposites. Mater. Today Commun. 2022, 34, 105108. [Google Scholar] [CrossRef]
  109. Hu, E.; Su, E.; Chen, Y.; Subedi, A.; Wang, J.; Hu, K.; Tang, L. Preparation and Tribological Behaviors of Modified Rice Husk Carbon/MoS2 Composite Particles as a Functional Additive in Polyethylene Glycol. Tribol. Trans. 2022, 65, 564–577. [Google Scholar] [CrossRef]
  110. Gong, K.; Lou, W.; Zhao, G.; Wu, X.; Wang, X. MoS2 nanoparticles grown on carbon nanomaterials for lubricating oil additives. Friction 2021, 9, 747–757. [Google Scholar] [CrossRef]
  111. Jiang, H.; Hou, X.; Ma, Y.; Su, D.; Qian, Y.; Ahmed Ali, M.K.; Dearn, K.D. The tribological performance evaluation of steel-steel contact surface lubricated by polyalphaolefins containing surfactant-modified hybrid MoS2/h-BN nano-additives. Wear 2022, 504–505, 204426. [Google Scholar] [CrossRef]
  112. Min, C.; Yang, Y.; Liang, H.; He, Z.; Zhu, J.; Li, Q. Covalently Bonded 2D/0D g-C3N4/MoS2 Nanocomposites for Enhanced Tribological Properties in Oil. ChemistrySelect 2021, 6, 1661–1668. [Google Scholar] [CrossRef]
  113. Guan, Z.; Zhang, P.; Florian, V.; Wu, Z.; Zeng, D.; Liu, J.; Wang, B.; Tu, X.; Li, S.; Li, W. Preparation and tribological behaviors of magnesium silicate hydroxide-MoS2 nanoparticles as lubricant additive. Wear 2022, 492–493, 204237. [Google Scholar] [CrossRef]
  114. Luo, T.; Wei, X.; Huang, X.; Huang, L.; Yang, F. Tribological properties of Al2O3 nanoparticles as lubricating oil additives. Ceram. Int. 2014, 40, 7143–7149. [Google Scholar] [CrossRef]
  115. Aldana, P.U.; Dassenoy, F.; Vacher, B.; Le Mogne, T.; Thiebaut, B. WS2 nanoparticles anti-wear and friction reducing properties on rough surfaces in the presence of ZDDP additive. Tribol. Int. 2016, 102, 213–221. [Google Scholar] [CrossRef]
  116. Ku, B.C.; Han, Y.C.; Lee, J.E.; Lee, J.K.; Park, S.H.; Hwang, Y.J. Tribological effects of fullerene (C60) nanoparticles added in mineral lubricants according to its viscosity. Int. J. Precis. Eng. Manuf. 2010, 11, 607–611. [Google Scholar] [CrossRef]
  117. Bao, Y.Y.; Sun, J.L.; Kong, L.H. Tribological properties and lubricating mechanism of SiO2 nanoparticles in waterbased fluid. IOP Conf. Ser. Mater. Sci. Eng. 2017, 182, 12025. [Google Scholar] [CrossRef]
  118. Azman, N.F.; Syahrullail, S.; Sot, M.N.H.M. Investigation of tribological properties of CuO/palm oil nanolubricant using pin-on-disc tribotester. Green Mater. 2018, 6, 30–37. [Google Scholar] [CrossRef]
  119. Xie, H.; Wang, Y.; Wang, P.; Liu, S.; Ye, Q.; Liu, W. Poly (tannic acid) functionalized onion-like carbon nanoparticles derived from candle soot serving as potent lubricant additives. J. Mol. Liq. 2023, 379, 121697. [Google Scholar] [CrossRef]
  120. Kotia, A.; Ghosh, G.K.; Srivastava, I.; Deval, P.; Ghosh, S.K. Mechanism for improvement of friction wear by using Al2O3 and SiO2/Gear oil nanolubricants. J. Alloys Compd. 2019, 782, 592–599. [Google Scholar] [CrossRef]
  121. Dai, W.; Kheireddin, B.; Gao, H.; Liang, H. Roles of nanoparticles in oil lubrication. Tribol. Int. 2016, 102, 88–98. [Google Scholar] [CrossRef]
  122. Guo, J.; Peng, R.; Du, H.; Shen, Y.; Li, Y.; Li, J.; Dong, G. The Application of Nano-MoS2 Quantum Dots as Liquid Lubricant Additive for Tribological Behavior Improvement. Nanomaterials 2020, 10, 200. [Google Scholar] [CrossRef] [PubMed]
  123. Luo, T.; Chen, X.; Li, P.; Wang, P.; Li, C.; Cao, B.; Luo, J.; Yang, S. Laser irradiation-induced laminated graphene/MoS2 composites with synergistically improved tribological properties. Nanotechnology 2018, 29, 265704. [Google Scholar] [CrossRef] [PubMed]
  124. Xie, H.; Jiang, B.; He, J.; Xia, X.; Pan, F. Lubrication performance of MoS2 and SiO2 nanoparticles as lubricant additives in magnesium alloy-steel contacts. Tribol. Int. 2016, 93, 63–70. [Google Scholar] [CrossRef]
  125. Meng, Y.; Sun, J.; He, J.; Yang, F.; Wu, F. Interfacial interaction induced synergistic lubricating performance of MoS2 and SiO2 composite nanofluid. Colloids Surf. A Physicochem. Eng. Asp. 2021, 626, 126999. [Google Scholar] [CrossRef]
  126. Wang, G.; Yang, C.; Bai, Q.; Guan, K.; Shang, K.; Li, D.; Guo, S. Alkylamine-Grafted Molybdenum Disulfide Nanosheets for Enhanced Dispersion Stability and Lubricity Properties. Langmuir 2023, 39, 12476–12487. [Google Scholar] [CrossRef]
  127. Liu, T.; Qin, J.; Wang, J.; Li, J. On the tribological properties of RGO–MoS2 composites surface modified by oleic acid. Tribol. Lett. 2022, 70, 14. [Google Scholar] [CrossRef]
  128. Kong, L.; Sun, J.; Bao, Y. Preparation, characterization and tribological mechanism of nanofluids. RSC Adv. 2017, 7, 12599. [Google Scholar] [CrossRef]
  129. Hu, K.H.; Liu, M.; Wang, Q.J.; Xu, Y.F.; Schraube, S.; Hu, X.G. Tribological properties of molybdenum disulfide nanosheets by monolayer restacking process as additive in liquid paraffin. Tribol. Int. 2009, 42, 33–39. [Google Scholar] [CrossRef]
  130. Gong, K.; Wu, X.; Zhao, G.; Wang, X. Nanosized MoS2 deposited on graphene as lubricant additive in polyalkylene glycol for steel/steel contact at elevated temperature. Tribol. Int. 2017, 110, 1–7. [Google Scholar] [CrossRef]
  131. Wu, H.; Qin, L.; Dong, G.; Hua, M.; Yang, S.; Zhang, J. An investigation on the lubrication mechanism of MoS2 nano sheet in point contact: The manner of particle entering the contact area. Tribol. Int. 2017, 107, 48–55. [Google Scholar] [CrossRef]
  132. Wu, H.; Wang, L.; Dong, G. Origin of the tribofilm from MoS2 nanoparticle oil additives: Dependence of oil film thickness on particle aggregation in rolling point contact. Friction 2021, 9, 1436–1449. [Google Scholar] [CrossRef]
  133. Xu, D.; Wang, C.; Espejo, C.; Wang, J.; Neville, A.; Morina, A. Understanding the friction reduction mechanism based on molybdenum disulfide tribofilm formation and removal. Langmuir 2018, 34, 13523–13533. [Google Scholar] [CrossRef]
  134. Song, W.; Yan, J.; Ji, H. Tribological Study of the SOCNTs@MoS2 Composite as a Lubricant Additive: Synergistic Effect. Ind. Eng. Chem. Res. 2018, 57, 6878–6887. [Google Scholar] [CrossRef]
  135. Zhang, Q.; Hu, X.; Duan, F.; Meng, Y. An efficient lubrication approach to mitigate soot-induced wear: Synergistic repair effect of magnetic MoS2 composites and magnetic field. Wear 2022, 488–489, 204182. [Google Scholar] [CrossRef]
  136. Hu, K.; Cai, Y.; Hu, X.; Xu, Y. Synergistic lubrication of MoS2 particles with different morphologies in liquid paraffin. Ind. Lubr. Tribol. 2013, 65, 143–149. [Google Scholar] [CrossRef]
  137. Liu, Y.; Hu, K.; Hu, E.; Guo, J.; Han, C.; Hu, X. Double hollow MoS2 nano-spheres: Synthesis, tribological properties, and functional conversion from lubrication to photocatalysis. Appl. Surf. Sci. 2017, 392, 1144–1152. [Google Scholar] [CrossRef]
  138. Cheng, L.; Hu, E.; Chao, X.; Zhu, R.; Hu, K.; Hu, X. MoS2/montmorillonite nanocomposite: Preparation, tribological properties and inner synergistic lubrication. Nano 2018, 13, 1850144. [Google Scholar] [CrossRef]
  139. Hu, K.; Miao, Y.; Lu, Z. Preparation and tribological performance of MoS2 nanoparticles supported on fly ash microparticles. Ind. Lubr. Tribol. 2023, 75, 51–59. [Google Scholar] [CrossRef]
  140. Yi, C.; Hu, C.; Shi, L.; Bai, M.; Li, Y.; Tang, D. Frictional properties of MoS2 on a multi-level rough wall under starved lubrication. Phys. Chem. Chem. Phys. 2023, 25, 14348–14358. [Google Scholar] [CrossRef] [PubMed]
  141. Yi, M.; Zhang, C. The synthesis of MoS2 particles with different morphologies for tribological application. Tribol. Int. 2017, 116, 285–294. [Google Scholar] [CrossRef]
  142. Zhang, Z.; Fan, X.; Ma, X.; Zhu, M. Synergistic lubrication mechanism of core/shell C@MoS2 particles as lubricant additives. Appl. Surf. Sci. 2023, 639, 158234. [Google Scholar] [CrossRef]
Figure 1. 1T, 2H, and 3R phases of MoS2 (Reproduced from Ref. [29] with permission from the Royal Society of Chemistry).
Figure 1. 1T, 2H, and 3R phases of MoS2 (Reproduced from Ref. [29] with permission from the Royal Society of Chemistry).
Lubricants 11 00527 g001
Figure 2. Schematic diagram of MoS2 preparation. (A) Ultrasonic stripping of MoS2 nanosheets (Reproduced from Ref. [34]), (B) Hydrothermal synthesis of CS@1T−MoS2 nanocomposites (Reproduced from Ref. [36]), (C) Solvothermal preparation of MoS2/N−rGO nanocomposites (Reproduced from Ref. [37]), and (D) Liquid−phase precipitation method for the preparation of MoS2 nanoparticles (Reproduced from Ref. [38]).
Figure 2. Schematic diagram of MoS2 preparation. (A) Ultrasonic stripping of MoS2 nanosheets (Reproduced from Ref. [34]), (B) Hydrothermal synthesis of CS@1T−MoS2 nanocomposites (Reproduced from Ref. [36]), (C) Solvothermal preparation of MoS2/N−rGO nanocomposites (Reproduced from Ref. [37]), and (D) Liquid−phase precipitation method for the preparation of MoS2 nanoparticles (Reproduced from Ref. [38]).
Lubricants 11 00527 g002
Figure 3. (a) Microwave ball milling device and (b) ultrasonic ball milling device (Reproduced from Ref. [75]).
Figure 3. (a) Microwave ball milling device and (b) ultrasonic ball milling device (Reproduced from Ref. [75]).
Lubricants 11 00527 g003
Figure 4. Chemical modification and dispersion stability of MoS2 (A) Synthesis route of OA-MoS2 nanosheets (Reproduced from Ref. [85]), (B) Schematic diagram of grafting ODT molecules onto sulfur vacancies in MoS2 nanosheets (Reproduced from Ref. [87]), (C) Dispersion stability of pristine MoS2 and MoS2-ODT nanosheets in polyol ester lubricant base oils (Reproduced from Ref. [87]), (D) Surface modifiers for the synthesis and flow chart of the modification of MoS2 nanosheets (Reproduced from Ref. [88]).
Figure 4. Chemical modification and dispersion stability of MoS2 (A) Synthesis route of OA-MoS2 nanosheets (Reproduced from Ref. [85]), (B) Schematic diagram of grafting ODT molecules onto sulfur vacancies in MoS2 nanosheets (Reproduced from Ref. [87]), (C) Dispersion stability of pristine MoS2 and MoS2-ODT nanosheets in polyol ester lubricant base oils (Reproduced from Ref. [87]), (D) Surface modifiers for the synthesis and flow chart of the modification of MoS2 nanosheets (Reproduced from Ref. [88]).
Lubricants 11 00527 g004
Figure 5. (A) Dispersions of 5W30 engine oil, h-BN-MoS2, and h-BN-MoS2-CTAB in 5W30 engine oil (Reproduced from Ref. [92]), (B) Dispersion of Gr-MS-Zn in 10W40 engine oil with different resting times (Reproduced from Ref. [95]).
Figure 5. (A) Dispersions of 5W30 engine oil, h-BN-MoS2, and h-BN-MoS2-CTAB in 5W30 engine oil (Reproduced from Ref. [92]), (B) Dispersion of Gr-MS-Zn in 10W40 engine oil with different resting times (Reproduced from Ref. [95]).
Lubricants 11 00527 g005
Figure 6. (A) Schematic of the growth of fullerene-like MoS2 nanospheres and their lubrication mechanism as lubricant additives: Under laser irradiation-induced instantaneous ultrahigh temperature and ultrahigh pressure, MoS2 nanoflakes will bend and melt in water at ambient conditions. The following quenching process of surrounding water solidifies the liquid MoS2 droplet into ideally fullerene-like nanospheres. (Reproduced from Ref. [104]), (B) SEM and TEM images of hollow core-shell MoS2 nanoparticles: (a,b) SEM and (ck) TEM images of the hollow core-shell MoS2 nanoparticles. Moreover, the broken hole in Figure 2B (directed by arrows) illustrates a hollow structure (Reproduced from Ref. [105]).
Figure 6. (A) Schematic of the growth of fullerene-like MoS2 nanospheres and their lubrication mechanism as lubricant additives: Under laser irradiation-induced instantaneous ultrahigh temperature and ultrahigh pressure, MoS2 nanoflakes will bend and melt in water at ambient conditions. The following quenching process of surrounding water solidifies the liquid MoS2 droplet into ideally fullerene-like nanospheres. (Reproduced from Ref. [104]), (B) SEM and TEM images of hollow core-shell MoS2 nanoparticles: (a,b) SEM and (ck) TEM images of the hollow core-shell MoS2 nanoparticles. Moreover, the broken hole in Figure 2B (directed by arrows) illustrates a hollow structure (Reproduced from Ref. [105]).
Lubricants 11 00527 g006
Figure 7. (A) Electron microscopy characterizations of MoS2@CNMs. FE-SEM images of (a) MoS2@CNT, (b) MoS2@Gr, and (c) MoS2@C60. Inset in (c) is the SEM image of pure C60. TEM images of (d) MoS2@CNT, (e) MoS2@Gr, and (f) MoS2@C60; (B) (a) COF and (b) wear volumes of steel discs lubricated by PAG base oil and PAG addicted with 1 wt% CNT, 1 wt% Gr, 1 wt% C60, 1 wt% MoS2, 1 wt% MoS2@CNT, 1 wt% MoS2@Gr, and 1 wt% MoS2@C60 at 100 °C (SRV conditions: load, 100 N; stroke, 1 mm; frequency, 25 Hz; duration, 30 min) (Reproduced from Ref. [110]).
Figure 7. (A) Electron microscopy characterizations of MoS2@CNMs. FE-SEM images of (a) MoS2@CNT, (b) MoS2@Gr, and (c) MoS2@C60. Inset in (c) is the SEM image of pure C60. TEM images of (d) MoS2@CNT, (e) MoS2@Gr, and (f) MoS2@C60; (B) (a) COF and (b) wear volumes of steel discs lubricated by PAG base oil and PAG addicted with 1 wt% CNT, 1 wt% Gr, 1 wt% C60, 1 wt% MoS2, 1 wt% MoS2@CNT, 1 wt% MoS2@Gr, and 1 wt% MoS2@C60 at 100 °C (SRV conditions: load, 100 N; stroke, 1 mm; frequency, 25 Hz; duration, 30 min) (Reproduced from Ref. [110]).
Lubricants 11 00527 g007
Figure 8. (A) (a) Average friction coefficient and WSD for steel tribe-pair using 5W30 fully formulated engine oil and its blend with variable doses of the h-BN-MoS2-CTAB heterostructure, (b) Friction profiles of steel tribe-pair using 5W30 fully formulated engine oil and its blend with h-BN-MoS2-CTAB (Reproduced from Ref. [92]). (B) (a) The curves in regard to friction coefficient of oil with g-C3N4 nanosheets, spherical MoS2, and g-C3N4/MoS2 nanocomposites at 0.20% concentration or without additive; (b) the parallel average friction coefficient and wear rate of steel sheets after being lubricated by these samples in sequence (Reproduced from Ref. [112]). (C) (a) COF curves as a function of sliding time, (b) average values for steel/steel under the lubrication of four kinds of oil samples, and (c) COF curves with the lubrication of PAO and PAO-MSH-MoS2 under different loads. (D) 3D topographies, wear scars profiles perpendicular to sliding directions, and surface height profiles along with the sliding directions in GCr15 discs under the lubrication of PAO (ac), PAO-MSH (df), PAO-MoS2 (gi), and PAO-MSH-MoS2 (jl) (Reproduced from Ref. [113]).
Figure 8. (A) (a) Average friction coefficient and WSD for steel tribe-pair using 5W30 fully formulated engine oil and its blend with variable doses of the h-BN-MoS2-CTAB heterostructure, (b) Friction profiles of steel tribe-pair using 5W30 fully formulated engine oil and its blend with h-BN-MoS2-CTAB (Reproduced from Ref. [92]). (B) (a) The curves in regard to friction coefficient of oil with g-C3N4 nanosheets, spherical MoS2, and g-C3N4/MoS2 nanocomposites at 0.20% concentration or without additive; (b) the parallel average friction coefficient and wear rate of steel sheets after being lubricated by these samples in sequence (Reproduced from Ref. [112]). (C) (a) COF curves as a function of sliding time, (b) average values for steel/steel under the lubrication of four kinds of oil samples, and (c) COF curves with the lubrication of PAO and PAO-MSH-MoS2 under different loads. (D) 3D topographies, wear scars profiles perpendicular to sliding directions, and surface height profiles along with the sliding directions in GCr15 discs under the lubrication of PAO (ac), PAO-MSH (df), PAO-MoS2 (gi), and PAO-MSH-MoS2 (jl) (Reproduced from Ref. [113]).
Lubricants 11 00527 g008
Figure 9. (A) Proposed lubrication mechanism for influence of dispersants on oil film thickness (Reproduced from Ref. [131]). (B) Proposed lubrication mechanism for influence of dispersants on oil film thickness. (a) In PAO4 base oil without proper dispersants and (b) in the presence of rich dispersants (Reproduced from Ref. [132]). (C) Schematic diagram of the MoDTC-derived tribofilm evolution (Reproduced from Ref. [133]).
Figure 9. (A) Proposed lubrication mechanism for influence of dispersants on oil film thickness (Reproduced from Ref. [131]). (B) Proposed lubrication mechanism for influence of dispersants on oil film thickness. (a) In PAO4 base oil without proper dispersants and (b) in the presence of rich dispersants (Reproduced from Ref. [132]). (C) Schematic diagram of the MoDTC-derived tribofilm evolution (Reproduced from Ref. [133]).
Lubricants 11 00527 g009
Figure 10. Schematic diagrams of the lubrication mechanism of MSH-MoS2 nanocomposites (a) four-ball tester contact model, (b) MSH-MoS2′s crystal structure, (c) an enlarged contact area shows the entry and adhesion of MSH-MoS2 nanocomposites on the sliding surfaces, (d) the formation of pad-like tribofilms on the worn interface, (e) the formation of pad-like tribofilms on the worn surface (Reproduced from Ref. [94]).
Figure 10. Schematic diagrams of the lubrication mechanism of MSH-MoS2 nanocomposites (a) four-ball tester contact model, (b) MSH-MoS2′s crystal structure, (c) an enlarged contact area shows the entry and adhesion of MSH-MoS2 nanocomposites on the sliding surfaces, (d) the formation of pad-like tribofilms on the worn interface, (e) the formation of pad-like tribofilms on the worn surface (Reproduced from Ref. [94]).
Lubricants 11 00527 g010
Figure 11. The staged action mechanism of C@MoS2 particles in the lubrication process (Reproduced from Ref. [142]).
Figure 11. The staged action mechanism of C@MoS2 particles in the lubrication process (Reproduced from Ref. [142]).
Lubricants 11 00527 g011
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

Lu, Z.; Lin, Q.; Cao, Z.; Li, W.; Gong, J.; Wang, Y.; Hu, K.; Hu, X. MoS2 Nanomaterials as Lubricant Additives: A Review. Lubricants 2023, 11, 527. https://doi.org/10.3390/lubricants11120527

AMA Style

Lu Z, Lin Q, Cao Z, Li W, Gong J, Wang Y, Hu K, Hu X. MoS2 Nanomaterials as Lubricant Additives: A Review. Lubricants. 2023; 11(12):527. https://doi.org/10.3390/lubricants11120527

Chicago/Turabian Style

Lu, Ziyan, Qingqing Lin, Zhaotao Cao, Wanyuan Li, Junjie Gong, Yan Wang, Kunhong Hu, and Xianguo Hu. 2023. "MoS2 Nanomaterials as Lubricant Additives: A Review" Lubricants 11, no. 12: 527. https://doi.org/10.3390/lubricants11120527

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