Next Article in Journal
Numerical and Experimental Investigations of Oil Return Efficiency in Tapered Roller Bearings Under Oil Jet Lubrication
Previous Article in Journal
A Possibility of Tribological Investigation of Physicochemical Processes in a Friction Pair Operating Under Selective Transfer Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of B Element Doping on High-Temperature Tribological Properties of WS2-Based Composite Coatings

1
School of Software, Henan University of Science and Technology, Luoyang 471003, China
2
High-End Bearing Henan Collaborative Innovation Center, Luoyang 471003, China
3
School of Mechatronics Engineering, Henan University of Science and Technology, Luoyang 471003, China
4
Henan Key Laboratory for Machinery Design and Transmission System, Henan University of Science and Technology, Luoyang 471003, China
*
Author to whom correspondence should be addressed.
Lubricants 2025, 13(8), 332; https://doi.org/10.3390/lubricants13080332
Submission received: 22 June 2025 / Revised: 21 July 2025 / Accepted: 25 July 2025 / Published: 30 July 2025

Abstract

WS2 coating, as a solid lubricating material, plays a significant role in the lubrication of rotating components in spacecraft. During the launch process, however, spacecraft are exposed to high-temperature and humid atmospheric environments, which can lead to oxidative failure in the coating, thereby limiting its engineering applications. By doping with B elements, B/WS2 was successfully prepared as a composite coating. The results demonstrate that the fabricated coating exhibits excellent high-temperature tribological performance in atmospheric environments. The mechanism through which B doping improves the high-temperature friction and wear properties of the WS2 composite coating was revealed through high-temperature friction and wear tests. With the incorporation of B elements, the average friction coefficient of the coating was 0.071, and the wear rate was 7.63 × 10−7 mm3·N−1·m−1, with the wear mechanisms identified as abrasive wear and spalling. Due to high-temperature oxidation, thermal decomposition effects, and the formation of WB4 during sputtering, the wear resistance and anti-plastic deformation capability of the coating were further improved. Compared to room-temperature test conditions, the B/WS2 composite coating at different high temperatures exhibited superior friction coefficients and wear rates. Notably, at 150 °C, the average friction coefficient was as low as 0.015, and the wear forms were abrasive wear and adhesive wear.

1. Introduction

With continuous improvements in the thrust-to-weight ratios of aeroengines, the operating temperature continues to rise, posing increasingly severe high-temperature challenges in bearing systems. The main bearing of a specific large passenger aircraft engine reaches a working temperature of nearly 300 °C during cruise conditions, far exceeding the service limits of traditional bearings [1]. High temperatures can lead to a series of issues, such as reduced material strength, the degradation of lubricant performance, and the thermal expansion of components, significantly affecting the service life and reliability of bearings. It is therefore necessary to develop a new generation of lubricating coatings for bearings that demonstrate self-lubricating properties, high-temperature resistance, and low friction [2]. Compared with grease lubrication and oil lubrication, solid lubricating coatings can avoid dependency on working environments and overcome replenishment difficulties.
As a solid lubricating material, WS2 coating is widely applied in the aerospace field [3,4] due to its excellent tribological properties, which arise from the interlayer shear slippage between S-W atoms under weak shear forces coupled with the formation of WO3 (a high-temperature lubricant) through oxidation [5]. The porous microstructure [6] and reactive dangling bonds at crystal edges [7], however, make WS2 coatings prone to oxidation through interactions with atmospheric oxygen in humid environments, resulting in limited wear resistance, which hinders their application in high-load and long-service-life conditions.
In recent years, researchers worldwide have significantly enhanced the performance of WS2 coatings through the doping of metallic elements (e.g., Cr [8], Ti [9], Ag [10], Ni [11], Ta [12], and Al [13]) or non-metallic elements (e.g., C [14] and N [15]), as well as various compounds. The results indicate that, on the one hand, the doped elements improve the microhardness and elastic modulus of the coating, enhancing its wear resistance and load-bearing capacity. On the other hand, they alter the microstructure and crystallographic orientation of the coating, leading to a denser surface structure for the WS2 coating and thus improving its tribological properties. When doped with reactive rare-earth metal elements (e.g., La [16] and Ce [17]), these elements act as oxygen scavengers, simultaneously enhancing the coating’s tribological and oxidation resistance properties. Ma [18] investigated the fabrication and high-temperature tribological performance of Ag-doped WS2-based composite coatings. By precisely controlling the Ag content during the deposition of WS2-Ag composite coatings, they systematically examined the influence of the Ag concentration on the friction and wear characteristics across a broad temperature range (RT~800 °C). Furthermore, their study elucidated the underlying wear mechanisms governing the WS2-Ag composite coatings under elevated temperatures. Xu et al. [19,20,21] incorporated Al, Cu, and Ni elements into WS2 coatings via radio-frequency co-sputtering. Their systematic investigation revealed distinct roles for different dopants: Cu doping promoted coating amorphization and enhanced structural densification; Ni doping effectively suppressed the columnar grain growth of WS2, with this inhibitory effect becoming more pronounced at higher Ni concentrations; and Al doping significantly improved environmental resistance by protecting the coating against water vapor and oxygen penetration, thereby enhancing oxidation resistance in humid environments. Dai et al. [22]. fabricated WS2-doped diamond-like carbon coatings using mid-frequency magnetron sputtering with ion source-assisted deposition technology. Their results demonstrated that the incorporation of WS2 into DLC coatings led to the formation of WC1-x and WS2 nanocrystalline clusters, which were uniformly dispersed within the carbon-based network. This unique microstructure endowed the W-S-C composite coatings with both the high hardness and excellent wear resistance that are characteristic of DLC coatings, as well as the superior self-lubricating properties inherent in WS2.
The rapid development of aviation equipment has led to higher demands in the performance of WS2 coatings, requiring enhanced friction and wear resistance, as well as high-temperature durability. However, the research on incorporating high-temperature-resistant materials into WS2 solid lubricant coatings remains relatively limited. Few studies have systematically investigated the doping of thermally stable additives in WS2-based solid lubricant coatings. As a non-metallic element, boron (B) exhibits distinctive physical properties such as a low density, metalloid characteristics, a high hardness, and excellent thermal resistance [23,24]. The incorporation of B into WS2 coatings to form B/WS2 composite coatings through structural modulation enables the strategic combination of soft-phase WS2 components with hard B-phase constituents. This synergistic design effectively harnesses the intrinsic advantages of both materials, maintaining superior lubricating properties while simultaneously enhancing the mechanical strength and tribological performance.

2. Materials and Methods

2.1. Fabrication of B/WS2 Composite Coatings

This study employed a JGP045CA magnetron sputtering deposition system (Shenyang Scientific Instrument Co., Ltd., Shenyang, China) to fabricate B/WS2 composite coatings. High-purity WS2 and B targets (99.99% purity) with dimensions of 50 mm in diameter and 3 mm in thickness were selected for the deposition process. The WS2 target was sputtered using radio-frequency power, while the B target was deposited via a direct current power supply. During the sputtering process, 9Cr18 bearing steel substrates with a diameter of 30 mm were selected for the evaluation of the coating’s tribological properties, while single-crystal silicon wafers (8 mm × 8 mm × 1 mm) were used to characterize the mechanical properties and cross-sectional morphology of the coatings. The pre-deposition treatment process was conducted as follows: the substrate surface was sequentially ground using silicon carbide abrasive papers with varying grit sizes (800, 1000, 1500, and 2000), followed by precision mechanical polishing using a diamond polishing compound (Henan Xinhong Grinding Materials Co., Ltd., Xuchang, China) (1.5 μm and 0.5 μm particle sizes). Subsequently, the substrates were ultrasonically cleaned in anhydrous ethanol and acetone for 10 min each, dried with nitrogen gas, and then transferred to a vacuum chamber. When the base pressure reached 5 × 10−4 Pa, argon gas was introduced for plasma sputtering. Prior to deposition, the substrates were subjected to 10 min of plasma cleaning, followed by the deposition of a 15 nm thick Cr interlayer. During the deposition process, dual magnetron sputtering targets were operated simultaneously while the substrate holder rotated at a constant speed of 20 revolutions per minute to ensure uniform film thickness and composition.
B/WS2 composite coatings were deposited under varying sputtering power conditions. The as-deposited B/WS2 composite coatings were subjected to high-temperature tribological tests at 150 °C, 300 °C, and 450 °C to evaluate their thermal stability and wear resistance. For comparative analysis, three types of coatings were selected, pure WS2 coatings, B/WS2 composite coatings, and B/WS2 composite coatings after high-temperature tribological testing, to systematically investigate the influence of different elevated-temperature conditions on their tribological performance. The detailed processing parameters are summarized in Table 1.

2.2. Microstructure and Property Characterization of B/WS2 Composite Coatings

The surface morphology, cross-sectional microstructure, and wear track characteristics of the B/WS2 composite coatings were systematically investigated using scanning electron microscopy(SEM; Zeiss Sigma 300, Carl Zeiss AG, Oberkochen, Germany). The phase composition and elemental distribution of the B/WS2 composite coatings were characterized using energy-dispersive X-ray spectroscopy. The crystal structure of the B/WS2 composite coatings was analyzed using X-ray diffraction, and the test conditions were as follows: an X-ray wavelength (λ) of 0.15405 nm; an operating voltage and current of 40 kV and 30 mA, respectively; a scanning range of 10~80°; and a scanning rate of 1°/min. The chemical valence states of the B/WS2 composite coating were analyzed using X-ray photoelectron spectroscopy. The interplanar spacing of the B/WS2 composite coating was characterized via high-resolution transmission electron microscopy.
The hardness and elastic modulus of the B/WS2 composite coating were analyzed using a nanoindentation tester manufactured by Nano Mechanics, San Jose, CA, USA. A Berkovich indenter was employed to perform single-point hardness tests on a monocrystalline silicon wafer. To minimize experimental errors, measurements were conducted at nine different locations, and the average value was calculated as the final result. The tests were performed under a load of 50 mN, with the maximum indentation depth set to be less than one-tenth of the coating thickness to avoid substrate effects.
The high-temperature tribological properties of the coating under dry friction conditions were evaluated using an HT-1000 rotary friction wear tester (HT-1000, Lanzhou Zhongke Kaihua Technology Development Co., Ltd., Lanzhou, China). Tests were conducted in four different environments, an ambient atmosphere, 150 °C, 300 °C, and 450 °C, employing a circular motion mode. A GCr15 steel ball (GCr15 steel ball, Shenzhen Dingte Steel Ball Co., Ltd., Liaocheng, China) with a diameter of 6 mm served as the counterpart. The experimental parameters included a normal load of 10 N, a rotational speed of 330 rpm, and a test duration of 15 min. To ensure reliability, each friction wear test was repeated at least three times.
The wear profile of the B/WS2 composite coating after tribological testing was characterized using a precision profilometer. The average wear cross-sectional area (S) was determined by integrating the profile curves at multiple locations along the wear track. Subsequently, the wear volume (V) was calculated by multiplying S by the circumferential length of the wear track. The wear rate was then computed using the standard wear rate formula [17].
W = V F L
Here, W represents the wear rate, with the unit mm3/(N·m); V denotes the worn volume of the wear scar, with the unit mm3; F stands for the applied load force, with the unit N (Newtons); and L indicates the total friction distance, with the unit m (meters).
The wear rate, determined by averaging the results of three repeated friction tests to reduce experimental deviations, was adopted as the evaluation metric for the coating’s wear performance in this study.

3. Results

3.1. Morphology Analysis of B/WS2 Composite Coating

Figure 1 shows the surface morphology of the B/WS2 composite coating. As can be seen in Figure 1, the pure WS2 coating exhibits large granular clusters with clearly defined grain boundaries, a relatively large grain size, numerous surface defects, and poor surface compactness. After the introduction of B, the surface morphology undergoes a significant transformation, shifting from coarse granular clusters to a “cellular-like” protrusion structure. The coating exhibits finer surface particles, improved flatness, and reduced surface roughness. Analysis suggests that the enhanced coating densification induced via B incorporation can be attributed to multiple effects, including significant room-temperature strengthening, surface hardening [25], and the formation of WB4 compounds. The surface morphology analysis of the B/WS2 composite coating after heating revealed that the cellular-like protrusion structure became more pronounced at 150 °C, accompanied by improved coating densification. With an increasing temperature, enhanced grain stacking was observed, leading to the further optimization of surface flatness. When the temperature reached 450 °C, a lamellar stacking phenomenon emerged in the coating. This can be attributed to high-temperature oxidation effects, which progressively enhanced the coating densification. Remarkably, the B/WS2 composite coating maintained its structural integrity even at 450 °C without undergoing morphological transformation compared to its room-temperature state.

3.2. Phase Analysis of B/WS2 Composite Coatings

Figure 2 presents the XRD diffraction patterns of the B/WS2 composite coatings. As shown in Figure 2, the pure WS2 coating exhibits characteristic diffraction peaks corresponding to the (002) and (101) crystal planes of WS2, with a predominant (002) basal plane orientation. The (002) basal plane orientation is primarily responsible for the coating’s lubricating properties, effectively reducing the sliding resistance and enhancing performance. Upon the incorporation of boron (B), the B/WS2 composite coating exhibited a noticeable reduction in peak intensity and broadening of the (002) diffraction peak, indicative of grain refinement and enhanced grain boundary strengthening. Concurrently, the WS2 (101) diffraction peak disappeared, while new diffraction peaks corresponding to the (101) and (112) crystallographic planes of WB4 emerged, confirming the formation of a WB4 compound with a preferred basal orientation. The results indicate that doping with B element significantly refines the grain size, resulting in grain boundary strengthening and solid solution strengthening effects. Additionally, the formation of WB4 compounds as a hard phase enhances both the densification and lubricating properties of the WS2-based solid lubricating coating. The B/WS2 composite coating was subjected to XRD analysis after high-temperature heating. At 150 °C, the peak width of the WS2 (002) diffraction plane showed no significant change, while the preferred orientation of the WB4 (101) and WB4 (112) diffraction planes intensified slightly, though the overall variation remained minor. When the temperature increased to 300 °C, the WS2 (002) peak broadened with a reduced intensity, and the preferred orientation of the WB4 (101) and WB4 (112) planes weakened. This trend became more pronounced at 450 °C. The analysis suggests that the heating temperature significantly alters the peak intensity and width of the WS2 and WB4 diffraction planes due to high-temperature oxidation and thermal decomposition, thereby affecting the lubricating properties and surface densification of the coating.

3.3. Analysis of Mechanical Properties of B/WS2 Composite Coatings

Figure 3 presents the microhardness and elastic modulus of the B/WS2 composite coatings. As evident from the figure, the pure WS2 coating exhibited a relatively low microhardness of only 0.83 GPa. In contrast, the B-doped B/WS2 composite coating demonstrated a significantly enhanced microhardness of 3.604 GPa, representing a 4.34-fold increase. Similarly, while the elastic modulus of pure WS2 measured 32.9 GPa, the B/WS2 composite coating achieved a substantially higher value of 54.7 GPa. Remarkably, after exposure to 450 °C, the B/WS2 coating showed further improvement in mechanical properties, with its microhardness increasing from 3.604 GPa to 4.17 GPa and its elastic modulus rising from 54.7 GPa to 55.84 GPa.
Figure 4 presents the H/E and H3/E2 ratios of the B/WS2 composite coatings. Our analysis reveals that the pure WS2 coating exhibited an H/E ratio of 0.0131, while B doping significantly enhanced this value to 0.065 for the B/WS2 composite coating, corresponding to a 4.99-fold improvement in wear resistance. After high-temperature exposure at 450 °C, the H/E ratio further increased to 0.076, representing a 5.84-fold enhancement compared to the undoped coating. More remarkably, the H3/E2 ratio, which reflects resistance to plastic deformation, showed even more dramatic improvements. The pure WS2 coating displayed an H3/E2 value of 5.842 × 10−4, whereas the B-doped composite coating achieved a value of 0.015: a substantial 25.6-fold increase. This enhancement became even more pronounced after thermal treatment at 450 °C, with the H3/E2 ratio reaching 0.0241 (41.26 times higher than the undoped coating). These results clearly demonstrate that boron doping can dramatically improve the mechanical properties of WS2 coatings. Notably, the B/WS2 composite coating maintains excellent mechanical performance even under elevated temperatures up to 450 °C, suggesting its potential for high-temperature tribological applications.
Analysis suggests that, due to the smaller atomic radius of B, its doping enhances the bond energy between adjacent atoms in the coating, contributing to grain boundary strengthening and solid solution strengthening. Additionally, the magnetron sputtering process generates WB4 compounds, the high hardness of which improves dislocation defects between unit cells, thereby enhancing the coating’s performance. At elevated temperatures, the oxidation of the coating affects the hardness and elastic modulus of the fibers. The primary composition of the B/WS2 composite coating, however, remains largely unchanged with an increasing temperature. The doping and higher content of B lead to improved mechanical properties of the coating. Furthermore, the thermal decomposition effect of the B/WS2 composite coating is pronounced, providing favorable conditions for high-temperature applications. These factors collectively contribute to the enhancement of the microhardness, elastic modulus, wear resistance, and resistance to plastic deformation of the B/WS2 composite coating.

3.4. Analysis of Tribological Properties of B/WS2 Composite Coatings

In the high-temperature friction and wear test, the coating was subjected to three cyclic wear tests, with a track radius of 2 mm, 2.5 mm, and 3 mm. The wear time was 15 min, and the speed was 330 r/min. The GCr15 steel ball was replaced after each test to ensure the accuracy of the test data. During the test, the experimental equipment was in a closed space, and a dehumidifier and refrigerator were used to ensure that the equipment was in an environment with a temperature of 25 °C and less than 30% humidity.
Figure 5 displays the friction coefficient curves of the B/WS2 composite coatings. As shown in Figure 5, the pure WS2 coating exhibits significant fluctuations in its friction curve, with intensified oscillations occurring around 10 min, yielding an average friction coefficient of 0.11. With the incorporation of B element, the fluctuation of the coating’s friction curve markedly decreases, achieving an average friction coefficient of 0.071, which is substantially lower than that of the pure WS2 coating. This result demonstrates that B doping can effectively enhance the tribological performance of the coating. The high-temperature tribological tests of the B/WS2 composite coating deposited at 70 W sputtering power reveal that its average friction coefficient initially decreases and then increases with a rising temperature. The coating exhibits the most favorable friction performance at 150 °C, achieving an ultralow average friction coefficient of 0.015. At 300 °C, the friction coefficient slightly increases to 0.024. Further temperature elevation to 450 °C leads to a more pronounced rise in the average friction coefficient, reaching 0.052. According to a study by Yan et al. [26] on Ag-doped WS2, when the Ag content is 6.24 at.%, the coating achieves optimal tribological performance, with an average friction coefficient of 0.023. Rengifo et al. [27] conducted research on Al-doped WS2, finding that, when the composition is Al-2 vol% WS2, the coating’s friction coefficient reaches a minimum of 0.55. Pei et al. [28] investigated Cu-doped WS2, noting that, when the Cu target sputtering power is 15 W, the coating’s friction coefficient reaches a minimum of 0.10. Additionally, according to the research on WS2 coatings conducted by Cai et al. [29], a pure WS2 coating exhibits an average friction coefficient of 0.147 at 500 °C, with relatively significant fluctuations in its friction coefficient curve. In our study, the B/WS2 composite coating shows a friction coefficient of 0.015 at 150 °C, and its friction coefficient at 450 °C is lower than that of the pure WS2 coating. It can therefore be concluded that B doping effectively enhances the high-temperature tribological properties of WS2 coatings.
Table 2 presents the wear rates of the B/WS2 composite coatings. As shown, the pure WS2 coating exhibits a relatively high wear rate in the order of 10−6 mm3·N−1·m−1 (3.63 × 10−6 mm3·N−1·m−1). However, with the incorporation of B, the coating demonstrates enhanced microhardness and an enhanced elastic modulus, leading to improved wear resistance and a significantly reduced wear rate (7.63 × 10−7 mm3·N−1·m−1). High-temperature tribological tests conducted on the B/WS2 coating deposited at 70 W sputtering power reveal that the wear rate initially decreases and then increases with a rising temperature. The lowest wear rate (1.89 × 10−7 mm3·N−1·m−1) is achieved at 150 °C, corresponding to the minimum average friction coefficient. At 300 °C, the wear rate slightly increases to 2.26 × 10−7 mm3·N−1·m−1. Further elevating the temperature to 450 °C results in a more substantial increase in the wear rate (8.96 × 10−7 mm3·N−1·m−1), which, although higher than that of the B/WS2 coating at room temperature, remains lower than that of pure WS2. These findings confirm that B doping significantly enhances the wear performance of the coating.
Figure 6 presents the wear scar morphology of the B/WS2 composite coatings. As observed in Figure 6a, the pure WS2 coating exhibits distinct grooves and wide wear tracks, indicating poor wear resistance with the characteristics of typical adhesive wear. This behavior originates from the coating’s low hardness and inferior wear resistance, making its surface susceptible to shear-induced sliding under cyclic loading, with wear debris accumulating along the groove edges. After doping with B element, the width of the wear track decreases slightly, and minor scratches appear at the edges of the wear track, while brittle spalling occurs inside the track [30]. The wear mechanisms of the coating involve both abrasive wear and spalling (Figure 6b). This is because external forces generate microcracks in the wear track during friction, and when these cracks propagate to a critical length, the coating delaminates from the substrate. Compared to the pure WS2 coating, the B/WS2 composite coating formed through B doping exhibits a more compact microstructure, improved wear resistance, and enhanced resistance to plastic deformation. Consequently, the wear mechanism of the B/WS2 composite coating shifts from adhesive wear to spalling and abrasive wear.
When the friction and wear test conditions shift from room temperature to elevated temperatures, the wear mechanism of the B/WS2 composite coating undergoes corresponding changes. At 150 °C, only slight wear traces are observed on the wear track of the B/WS2 composite coating, with abrasive wear being the dominant mechanism (Figure 6c). A similar trend is observed at 300 °C. At 450 °C, however, debris accumulation appears along both sides of the wear track, and the coating exhibits a combination of abrasive wear and partial adhesive wear (Figure 6e). Analysis suggests that the B/WS2 composite coating possesses internal residual stress. Under high-temperature testing conditions, this stress is released, enhancing the coating’s resistance to cyclic external loading and preventing premature failure. Due to its dense microstructure and superior wear resistance, the dominant wear mechanism at elevated temperatures is still abrasive wear [31]. At 450 °C, however, the high-temperature environment facilitates rapid reactions between the coating and oxygen/other gaseous impurities, generating additional abrasive particles. Consequently, the combined abrasive action of the steel counterpart and these wear particles intensifies. As the wear debris accumulates beyond a critical threshold, partial debris forms piled-up ridges along the wear track edges, while the remainder transforms into a transfer film, contributing to lubrication. This transition alters the wear mechanism to adhesive wear [32].
Figure 7 shows the wear scar morphology and EDS spectra of steel balls with different coatings. It can be observed that transfer films are formed on the pure WS2 coatings, B-doped B/WS2 composite coatings, and B/WS2 composite coatings under various test temperatures. The generation of transfer films contributes to the excellent friction and wear performance of the coatings. The pure WS2 coating produces a relatively large amount of transfer film, accompanied by considerable wear debris. This is attributed to the low hardness and loose porous structure of the pure WS2 coating, which leads to severe wear under cyclic loading, resulting in increased transfer film formation in the wear scar. In contrast, the B-doped WS2 coating exhibits higher hardness and resistance to plastic deformation. On the one hand, the superior mechanical properties of the coating enable it to withstand the alternating stresses caused by cyclic loading, minimizing damage to the coating structure and generating less wear debris. Some of the debris form transfer films, while the rest participate in the wear process as fine abrasive particles [33]. On the other hand, the transfer films present participate in the friction and wear process, resulting in lower transfer film formation for the B-doped B/WS2 composite coating during sliding. Under high-temperature testing conditions, the oxidation and thermal decomposition effects of the B/WS2 composite coating further enhance its hardness, wear resistance, and resistance to plastic deformation, thereby improving its tribological performance. Consequently, smaller wear scars and fewer transfer films are observed, which can be attributed to the superior wear resistance of the coating. Excessively high temperatures, however, degrade the original mechanical properties of the B/WS2 composite coating, leading to larger wear scar areas and increased transfer film formation.
EDS analysis of the oxygen distribution in the wear scars of the steel balls revealed the presence of oxygen across all coatings. In the pure WS2 coating, oxygen accumulation was observed at the edges of the wear scar, whereas the B-doped WS2 composite coating exhibited no significant oxygen enrichment, indicating that boron doping effectively enhanced the oxidation resistance of the WS2 coating. Under high-temperature testing at 150 °C, the B/WS2 composite coating still showed no oxygen aggregation. As the test temperature increased to 300 °C and 450 °C, however, distinct oxygen enrichment became apparent. This phenomenon can be attributed to frictional heating and plastic deformation during wear, which generate localized heat and surface activation, promoting oxidation. At elevated temperatures, tungsten (W) oxidizes to form WO3, while the combined effects of frictional heat and environmental heating further accelerate the oxidative degradation of the coating.
To gain deeper insights into the chemical composition of the B/WS2 composite coating, XPS analysis was conducted to examine the wear morphology and scar composition at both room and elevated temperatures. Figure 5, Figure 6, Figure 7 and Figure 8 present the XPS fitting results of the coatings. As shown in Figure 8, W, S, and B elements were detected on the surface of the B-doped WS2 coating at room temperature. The deconvolution of the W4f peak for the pure WS2 coating revealed minor metallic W (W4f7/2 31.6 eV, W4f5/2 33.7 eV), a moderate amount of WB4 (W4f7/2 36.1 eV, W4f5/2 38.2 eV), and a dominant WS2 phase (W4f7/2 35.6 eV, W4f5/2 36.1 eV) [34,35]. Notably, XPS analysis confirmed the presence of W-B bonds, with the binding energies consistent with the standard values for tungsten tetraboride (WB4). The high hardness and superior wear resistance of WB4 significantly enhanced the mechanical properties of the coating, reducing the average friction coefficient to 0.069 and substantially decreasing the wear rate (7.63 × 10−7 mm3·N−1·m−1). Combined with the wear scar morphology in Figure 7, the transfer film formed during room-temperature friction primarily consisted of W and S elements, along with trace amounts of O. XPS analysis of the B/WS2 composite coating at 450 °C revealed the disappearance of metallic W (W4f7/2 31.6 eV, W4f5/2 33.7 eV), with the coating mainly composed of WS2, WB4, and WO3. The formation of WO3 occurred via two pathways: (1) the partial oxidation of metallic W during high-temperature friction reacting with atmospheric oxygen and impurities to form minor WO3 (accounting for the depletion of metallic W) and (2) the extensive oxidation of WS2 in the coating under elevated temperatures, generating abundant WO3, the distribution of which matched the oxygen enrichment pattern observed in Figure 7. The high-temperature transfer film was predominantly composed of WO3.
To further investigate the friction and wear mechanisms of the B/WS2 composite coating, wear debris samples were prepared on microgrid copper meshes for transmission electron microscopy (TEM) analysis (Figure 9). The TEM results demonstrate that the wear debris of the B/WS2 composite coating at room temperature primarily consists of WS2, the excellent lubricating properties of which contribute to the coating’s favorable friction coefficient (0.11). Notably, the WB4 lattice can be clearly observed in the wear debris. Combined with XRD and XPS analyses, this finding confirms the formation of WB4 during the friction process. The presence of WB4 significantly enhances the load-bearing capacity of the coating through both solid solution strengthening and its inherent high hardness characteristics (with the coating’s microhardness increasing from 0.83 GPa to 3.604 GPa). After exposure to 450 °C, the wear debris exhibits spherical particle accumulation, with a morphology resembling WO3. The measured interplanar spacing of approximately 0.39 nm matches the standard value for WO3, which is further corroborated via the EDS analysis of the wear scar (Figure 7) and the XPS spectra of the steel ball’s wear scar (Figure 8), confirming the formation of WO3 compounds during high-temperature friction. Our analysis suggests that WO3 functions as a high-temperature solid lubricant with a moderate lubricating capability. At room temperature, the coating contains abundant inherent WS2, the superior lubricity of which surpasses that of WO3, resulting in favorable friction coefficients. Additionally, B doping enhances the load-bearing capacity of the coating, while improved hardness and wear resistance further optimize its tribological performance. Under elevated temperatures, the combined effects of thermal conditions and grinding heat promote the reaction between W and O to form WO3. The high-temperature lubricating properties of WO3, coupled with the rolling effect of its spherical particles, collectively reduce the friction coefficient, endowing the coating with excellent friction and wear resistance, even at 450 °C.

4. Conclusions

A comparative analysis of pure WS2 coatings, B/WS2 composite coatings, and B/WS2 composite coatings under high-temperature friction and wear test conditions yielded the following conclusions: The pure WS2 coating exhibited poor wear resistance, with its wear mechanism dominated by typical adhesive wear. With boron doping, the average friction coefficient decreased from 0.11 to 0.071, while the wear rate reduced from 3.63 × 10−6 mm3·N−1·m−1 to 7.63 × 10−7 mm3·N−1·m−1. The wear mechanism of the composite coating manifested as a combination of abrasive wear and spalling. The B/WS2 composite coatings tested at elevated temperatures (150 °C and 300 °C) demonstrated superior friction coefficients and wear rates compared to those tested under room-temperature conditions, exhibiting wear characteristics of both abrasive and adhesive wear. Notably, even at 450 °C, the B/WS2 composite coating maintained excellent tribological performance at a sputtering power of 70 W, with its wear mechanism consisting of abrasive wear accompanied by adhesive wear.

Author Contributions

Conceptualization, S.Z. and H.C.; data curation, X.Z.; funding acquisition, S.Z.; investigation, H.C. and X.Z.; methodology, H.C. and Z.H.; project administration, Y.X.; supervision, L.P.; visualization, X.Z. and B.K.; writing—original draft, S.Z.; writing—review and editing, H.C., X.Z. and Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Technologies Research and Development Program, Grant No. 2021YFB3400401 and the Key Research Projects of Colleges and Universities in Henan Province, Grant No. 25CY007.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

Authors Songmin Zhang, Haichao Cai, Zixuan Huang and Yujun Xue were employed by the company High-End Bearing Henan Collaborative Innovation Center. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Guo, Q.; Liu, H.; Sun, C.; Cao, Y.; Lu, X.; Du, Y.; Ru, X.; Xu, H.; Song, K.; Fu, P.; et al. Dual-gradient structure enhances wear resistance of aero-engine bearing steel by suppressing strain localization. Acta Mater. 2025, 289, 120919. [Google Scholar] [CrossRef]
  2. Song, F.; Pu, J. Research Status of Low-friction Coatings in Wide Temperature Range. China Surf. Eng. 2022, 35, 1–15. [Google Scholar]
  3. Fan, X.; Shi, Y.; Cui, M.; Ren, S.; Wang, H.; Pu, J. MoS2/WS2 nanosheet-based composite films irradiated by atomic oxygen: Implications for lubrication in space. ACS Appl. Nano Mater. 2021, 4, 10307–10320. [Google Scholar] [CrossRef]
  4. Li, X.; Deng, J.; Yue, H.; Ge, D.; Zou, X. Wear performance of electrohydrodynamically atomized WS2 coatings deposited on biomimetic shark-skin textured surfaces. Tribol. Int. 2019, 134, 240–251. [Google Scholar] [CrossRef]
  5. Li, Z.; Wang, D.; Ye, Q.; He, J.; Sun, J.; Weng, L.; Hu, M.; Gao, X. The effect of oxygen introduced by WO3 on the structure and tribological performance of WSO sputtered films. Tribol. Int. 2024, 193, 109408. [Google Scholar] [CrossRef]
  6. Babuska, T.F.; Curry, J.F.; Thorpe, R.; Cjowdhury, M.; Strandwitz, N.C.; Krick, B.A. High-sensitivity low-energy ion spectroscopy with sub-nanometer depth resolution reveals oxidation resistance of MoS2 increases with film density and shear-induced nanostructural modifications of the surface. ACS Appl. Nano Mater. 2023, 6, 1153–1160. [Google Scholar] [CrossRef]
  7. Feng, X.; Zhou, H.; Zheng, Y.; Zhang, K.; Zhang, Y. Tribological behavior and wear mechanism of Ti/MoS2 films deposited on plasma nitrided CF170 steel sliding against different mating materials. Vacuum 2021, 194, 110623. [Google Scholar] [CrossRef]
  8. Yu, C.; Xie, L.; Wang, Q.; Niu, Y.; Chen, M.; Wang, F. Investigation of the Synergistic Effect of Cr and CeO2 on the Mechanical, Oxidation, and Tribological Properties of Ni-Based Self-lubricating Composites. Metall. Mater. Trans. A 2024, 55, 2773–2789. [Google Scholar] [CrossRef]
  9. Liu, J.; Yan, Z.; Hao, J.; Liu, W. The influence of proton radiation on the friction and wear of Ti: WS2/P201 composite lubricating materials for space application. Wear 2024, 538, 205190. [Google Scholar] [CrossRef]
  10. Kaushik, A.; Bisht, P.; Senapati, S.; Singh, J.P. Ag nanoparticle functionalized vertical WS2 nanoflakes as SERS substrate for chemical sensing. Surf. Interfaces 2024, 52, 104951. [Google Scholar] [CrossRef]
  11. Romanov, R.I.; Fominskii, D.V.; Kas’yanenko, V.A.; Gritskevich, M.D.; Fominskii, V.Y. Effect of Interface Ni and Ti Films on Antifriction Properties of Nanolayer Thin-Film WS2/g-C Coatings. Phys. At. Nucl. 2023, 85, 1816–1820. [Google Scholar] [CrossRef]
  12. Yang, J.; Wang, D.; Fu, Y.; Wang, Q.; Hu, M.; Jiang, D.; Gao, X.; Sun, J.; Weng, L. Improving the tribological and anti-corrosion property of the WS2 film through Ta doping. Vacuum 2021, 192, 110485. [Google Scholar] [CrossRef]
  13. Zhou, X.; Yuan, L.; Yue, Q.; Dan, Z.; Li, H. Structural regulation of WS2/XSe (X= Al, Ga, In) heterostructure from first-principles calculations for bandstructure engineering. Mater. Sci. Semicond. Process. 2025, 189, 109305. [Google Scholar] [CrossRef]
  14. Cao, H.; De Hosson, J.T.M.; Pei, Y. Effect of carbon concentration and argon flow rate on the microstructure and triboperformance of magnetron sputtered WS2/aC coatings. Surf. Coat. Technol. 2017, 332, 142–152. [Google Scholar] [CrossRef]
  15. Baglio, J.; Kamieniecki, E.; DeCola, N.; Struck, C.; Marzik, J.; Dwight, K.; Wold, A. Growth and characterization of n-WS2 and niobium-doped p-WS2 single crystals. J. Solid State Chem. 1983, 49, 166–179. [Google Scholar] [CrossRef]
  16. Cai, H.; Xue, Y.; Ye, J.; Wang, J.; Pang, B.; Li, H. Effect of Sputtering Power on High Temperature Tribological Behavior of La-Ti/WS2 Composite Films. RARE Met. Mater. Eng. 2023, 52, 1201–1209. [Google Scholar]
  17. Tian, C.; Cai, H.; Xue, Y. Effect of Working Pressure on Tribological Properties of Ce-Ti/MoS2 Coatings Using Magnetron Sputter. Coatings 2022, 12, 1576. [Google Scholar] [CrossRef]
  18. Xu, S.; Hu, M.; Sun, J.; Weng, L.; Liu, W.; Gao, X. A simple strategy to tailor the microstructure and wear-resistance of sputtered WS2 films. Mater. Lett. 2018, 216, 179–181. [Google Scholar] [CrossRef]
  19. Xu, S.; Gao, X.; Hu, M.; Wang, D.; Jiang, D.; Sun, J.; Zhou, F.; Weng, L.; Liu, W. Microstructure evolution and enhanced tribological properties of Cu-doped WS2 films. Tribol. Lett. 2014, 55, 1–13. [Google Scholar] [CrossRef]
  20. Xu, S.; Gao, X.; Hu, M.; Sun, J.; Jiang, D.; Wang, D.; Zhou, F.; Weng, L.; Liu, W. Dependence of atomic oxygen resistance and the tribological properties on microstructures of WS2 films. Appl. Surf. Sci. 2014, 298, 36–43. [Google Scholar] [CrossRef]
  21. Xu, S.; Gao, X.; Hu, M.; Sun, J.; Jiang, D.; Zhou, F.; Liu, W.; Weng, L. Nanostructured WS2–Ni composite films for improved oxidation, resistance and tribological performance. Appl. Surf. Sci. 2014, 288, 15–25. [Google Scholar] [CrossRef]
  22. Dai, M.; Wei, C.; Zhou, K.; Zhu, M.; Hou, H.; Lin, S.; Tong, X. Properties of W/DLC/W–S–C composite films fabricated by magnetron sputtering. Trans. Nonferrous Met. Soc. China 2015, 25, 3002–3011. [Google Scholar] [CrossRef]
  23. Hou, G.; Chen, J.; Chen, W.; Chen, J.; Cheng, J.; Zhu, S.; Yang, J. Wear compatibility of super-hard WB4-B tribo-pair materials in moist environments. Tribol. Int. 2025, 204, 110485. [Google Scholar] [CrossRef]
  24. Hou, G.; Zhu, S.; Tan, H.; Chen, W.; Chen, J.; Sun, Q.; Chen, J.; Cheng, J.; Li, P.; Yi, W.; et al. Near-zero-wear with super-hard WB4 and a self-repairing tribo-chemical layer. Commun. Mater. 2024, 5, 222. [Google Scholar] [CrossRef]
  25. Yue, X.; Chen, J.; Han, J.; Shan, Y.; Shen, S.; Wu, W.; Liu, B.; Li, L.; Chen, Y.; Zhang, R.; et al. Monitoring and engineering interface coupling between monolayer WS2 and substrate through controllably introducing interfacial strain. Sci. China Mater. 2024, 67, 3012–3020. [Google Scholar] [CrossRef]
  26. Zhen, Y.; Liu, J.; Shi, X.; Wang, C.; Lu, X.; Hao, J. Tribological properties and transfer behaviors of WS2-Ag nanocomposite films with structure evolution for aerospace application. Appl. Surf. Sci. 2024, 674, 160928. [Google Scholar] [CrossRef]
  27. Sara, R.; Zhang, C.; Harimkar, S.; Boesl, B.; Agarwal, A. Effect of WS2 addition on tribological behavior of aluminum at room and elevated temperatures. Tribol. Lett. 2017, 65, 76. [Google Scholar]
  28. Lulu, P.; Kang, B.; Cai, H.; Xue, Y. Excellent Tribological Properties of WS2 Films in Air by Doping Copper. Coatings 2025, 15, 173. [Google Scholar] [CrossRef]
  29. Cai, H.; Xue, Y.; Pang, B.; Wang, J.; Ye, J. Effect of rare earth La on friction and wear resistance of WS2-based composite coating at high temperature. J. Mater. Sci. 2022, 57, 16875–16891. [Google Scholar] [CrossRef]
  30. Ning, Z.; Wang, Y.; Chen, T.; Tan, Z.; Wen, M. Research progress of silver films/coatings deposited by magnetron sputtering. Rare Met. Mater. Eng. 2022, 51, 4773–4782. [Google Scholar]
  31. Gao, X.; Fu, Y.; Jiang, D.; Wang, D.; Yang, J.; Weng, L.; Hu, M.; Sun, J. Structural, mechanical, and tribological properties of WS2-Al nanocomposite film for space application. Tribol. Lett. 2018, 66, 137. [Google Scholar] [CrossRef]
  32. Hu, M.; Dong, Y.; Wang, Y.; Wang, D.; Jiang, D.; Fu, Y.; Weng, L.; Sun, J.; Gao, X. Low deposition temperature-induced changes of the microstructure and tribological property of WS2 film. Coatings 2019, 9, 227. [Google Scholar] [CrossRef]
  33. Zhang, K.; Deng, J.; Lei, S.; Yu, X. Effect of micro/nano-textures and burnished MoS2 addition on the tribological properties of PVD TiAlN coatings against AISI 316 stainless steel. Surf. Coat. Technol. 2016, 291, 382–395. [Google Scholar] [CrossRef]
  34. Long, Y.; Wu, Z.; Zheng, X.; Lin, H.; Zhang, F. Mechanochemical synthesis and annealing of tungsten di-and tetra-boride. J. Am. Ceram. Soc. 2020, 103, 831–838. [Google Scholar] [CrossRef]
  35. Qin, Z.; Gong, W.; Song, X.; Wang, M.; Wang, H.; Li, Q. Effect of pressure on the structural, electronic and mechanical properties of ultraincompressible W2B. RSC Adv. 2018, 8, 35664–35671. [Google Scholar] [CrossRef]
Figure 1. FESEM surface morphology images of B/WS2 composite coatings.
Figure 1. FESEM surface morphology images of B/WS2 composite coatings.
Lubricants 13 00332 g001
Figure 2. XRD diffraction patterns of B/WS2 composite coatings.
Figure 2. XRD diffraction patterns of B/WS2 composite coatings.
Lubricants 13 00332 g002
Figure 3. Microhardness and elastic modulus of B/WS2 composite coatings.
Figure 3. Microhardness and elastic modulus of B/WS2 composite coatings.
Lubricants 13 00332 g003
Figure 4. H/E and H3/E2 values of B/WS2 composite coatings.
Figure 4. H/E and H3/E2 values of B/WS2 composite coatings.
Lubricants 13 00332 g004
Figure 5. Friction coefficient curve of B/WS2 composite coatings.
Figure 5. Friction coefficient curve of B/WS2 composite coatings.
Lubricants 13 00332 g005
Figure 6. Wear scar morphology of B/WS2 composite coatings.
Figure 6. Wear scar morphology of B/WS2 composite coatings.
Lubricants 13 00332 g006
Figure 7. Steel ball wear scar morphology and EDS energy spectrum of B/WS2 composite coatings.
Figure 7. Steel ball wear scar morphology and EDS energy spectrum of B/WS2 composite coatings.
Lubricants 13 00332 g007aLubricants 13 00332 g007b
Figure 8. Steel ball XPS energy spectrum of B/WS2 composite coatings.
Figure 8. Steel ball XPS energy spectrum of B/WS2 composite coatings.
Lubricants 13 00332 g008
Figure 9. Microgrid TEM images of B/WS2 composite coatings.
Figure 9. Microgrid TEM images of B/WS2 composite coatings.
Lubricants 13 00332 g009
Table 1. B/WS2 composite coating preparation and experimental parameters.
Table 1. B/WS2 composite coating preparation and experimental parameters.
Vacuum Degree/Pa5 × 10−4
Sputtering power of target B/W70
Argon flow rate/Sccm50
Sputtering temperature/°C300
WS2 target power/W200
Deposition time/min60
Friction and wear test temperature/°C25–450
Table 2. Wear rate of B/WS2 composite coatings.
Table 2. Wear rate of B/WS2 composite coatings.
Sample NumberCoating NameTarget B PowerTest
Temperature
Wear rate/mm3·N−1·m−1Wear Rate SD RateAverage Friction
Coefficient Curve
1WS20 W25 °C3.63 × 10−61.56266 × 1070.11
2B/WS270 W25 °C7.63 × 10−73.81655 × 1080.071
3B/WS270 W150 °C1.89 × 10−79.46452 × 1090.015
4B/WS270 W300 °C2.26 × 10−71.13475 × 1080.024
5B/WS270 W450 °C8.96 × 10−74.4836 × 1080.053
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

Zhang, S.; Zhang, X.; Cai, H.; Huang, Z.; Xue, Y.; Pei, L.; Kang, B. Effect of B Element Doping on High-Temperature Tribological Properties of WS2-Based Composite Coatings. Lubricants 2025, 13, 332. https://doi.org/10.3390/lubricants13080332

AMA Style

Zhang S, Zhang X, Cai H, Huang Z, Xue Y, Pei L, Kang B. Effect of B Element Doping on High-Temperature Tribological Properties of WS2-Based Composite Coatings. Lubricants. 2025; 13(8):332. https://doi.org/10.3390/lubricants13080332

Chicago/Turabian Style

Zhang, Songmin, Xiaopeng Zhang, Haichao Cai, Zixuan Huang, Yujun Xue, Lulu Pei, and Bowei Kang. 2025. "Effect of B Element Doping on High-Temperature Tribological Properties of WS2-Based Composite Coatings" Lubricants 13, no. 8: 332. https://doi.org/10.3390/lubricants13080332

APA Style

Zhang, S., Zhang, X., Cai, H., Huang, Z., Xue, Y., Pei, L., & Kang, B. (2025). Effect of B Element Doping on High-Temperature Tribological Properties of WS2-Based Composite Coatings. Lubricants, 13(8), 332. https://doi.org/10.3390/lubricants13080332

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