Next Article in Journal
A Comprehensive Review of Hot In-Place Recycling Technology: Classification, Factors Affecting Performance of Asphalt Mixtures, and Benefits Analysis
Previous Article in Journal
Preparation and Properties of a Composite Glass Protective Lubricating Coating for the Forging of Ti-6Al-4V Alloy
Previous Article in Special Issue
Preparation and Hydrogen Resistance Property of CrCN Coating by Magnetron Sputtering Method Under Different C-Target Powers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Si Content on the Microstructure and Properties of Hydrogenated Amorphous Carbon Films Deposited by Magnetron Sputtering Technique

1
AECC Zhongchuan Transmission Machiney Co., Ltd., 248 No. Middle of Guoliang Road, Wangcheng, Changsha 410200, China
2
School of Materials Science and Engineering, Anhui University of Technology, Maanshan 243002, China
3
School of Metallurgical Engineering, Anhui University of Technology, Maanshan 243002, China
4
Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University, Huizhou 516007, China
5
China International Science and Technology Cooperation Base on Intelligent Equipment Manufacturing in Special Service Environment, Anhui University of Technology, Maanshan 243002, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(7), 793; https://doi.org/10.3390/coatings15070793
Submission received: 17 June 2025 / Revised: 3 July 2025 / Accepted: 4 July 2025 / Published: 6 July 2025
(This article belongs to the Special Issue Sputtering Deposition for Advanced Materials and Interfaces)

Abstract

Hydrogenated amorphous carbon (a-C:H) films are widely valued for their excellent mechanical strength and low friction, but their performance significantly degrades at elevated temperatures, limiting practical applications in aerospace environments. In this work, we aimed to enhance the high-temperature tribological behavior of a-C:H films through controlled silicon (Si) doping. A series of a-C:H:Si films with varying Si contents were fabricated via direct current magnetron sputtering, and their microstructure, mechanical properties, and friction behavior were systematically evaluated from room temperature up to 400 °C. Results show that moderate Si doping (8.3 at.%) substantially enhances hardness and wear resistance, while enabling ultralow friction (as low as 0.0034) at 400 °C. This superior performance is attributed to the synergistic effects of transfer layer formation, preferential Si oxidation, and tribo-induced graphitization. This study provides new insights into the high-temperature lubrication mechanisms of Si-doped a-C:H films and demonstrates the critical role of Si content optimization, highlighting a viable strategy for extending the thermal stability and lifespan of solid-lubricating films.

1. Introduction

With the rapid advancement of aviation technology, increasing demands have been placed on the high-temperature lubrication of critical aero-engine components. Conventional liquid lubricants undergo rapid thermal degradation above approximately 350 °C, rendering them unsuitable for such extreme environments [1,2]. Hydrogenated amorphous carbon films (a-C:H), depending on their excellent mechanical strength, chemical inertness, and intrinsically low friction coefficient, have emerged as promising solid-lubrication candidates for aerospace and other high-performance applications [3].
Amorphous carbon lacks a long-range crystalline structure and comprises a metastable mixture of sp2- and sp3-bonded carbon. Under elevated temperatures, this metastability promotes conversion of sp3-C into sp2-C and concurrent hydrogen desorption, leading to structural graphitization and diminished wear resistance. Indeed, non-equilibrium magnetron-sputtered diamond-like carbon (DLC) films prepared by Ni et al. exhibited pronounced frictional degradation beginning at 240 °C and complete failure at 300 °C [4], while microwave sheath-voltage plasma deposition by Deng et al. showed onset of property decay at 200 °C and rapid loss of lubricity upon short-term sliding at 300 °C [5]. Erdemiret and coworkers further reported that DLC’s wear rate at 250 °C (1.39 × 10−6 mm3 N−1 m−1) is orders of magnitude higher than at room temperature (0.000186 × 10−6 mm3 N−1 m−1) [6]. Although Liu et al. demonstrated relatively stable wear behavior up to 200 °C in humid air against Al2O3, the films suffered catastrophic failure by graphitization above 300 °C despite a friction coefficient reduction from 0.15 to as low as 0.02, indicating a severely limited high-temperature lifetime [7].
Elemental doping (e.g., F [8], Si [9], Zr [10], Cr [11,12], W [13], Mo [14]) has proven effective at enhancing both mechanical properties and high-temperature tribological performance of amorphous carbon. Silicon, in particular, offers unique benefits: the longer Si-C bond relative to C-C relieves intrinsic compressive stress [15], preferentially increases the sp3-bonded fraction, and reduces sp2 cluster size, thereby fortifying the film matrix [16]. HiPIMS-deposited Si-doped a-C films reported by Zhang et al. showed elevated sp3 content and reduced residual stress due to Si substitution in sp2 sites [17]. Previous studies reported that Si doping levels of approximately 2–4 at.% could significantly reduce the wear rate of a-C:H films in the 10−7 mm3/N·m order of magnitude and lower the friction coefficient to values as low as 0.04 at elevated temperatures [18,19], due to suppressed oxidation and the formation of Si-containing transfer layers [18,19,20,21].
Despite extensive efforts, a comprehensive understanding of how varying Si content influences the microstructure and high-temperature tribological behavior of hydrogenated amorphous carbon films remains limited. In this study, we systematically investigate a-C:H:Si films prepared by DC magnetron sputtering with Si contents ranging from 0 to 14.7 at.%, focusing on their structural evolution, mechanical performance, and friction behavior at room temperature, 300 °C, and 400 °C. The objective of this work is to elucidate the role of Si doping in enhancing the high-temperature durability of a-C:H films and to establish correlations between microstructure and tribological response, which have not been comprehensively addressed in prior literature.

2. Experimental

2.1. Film Deposition

In this work, a-C:H and Si-doped a-C:H (a-C:H:Si) films were deposited onto 9Cr18 stainless-steel substrates (for friction tests) and Silicon wafers (for SEM, Raman, TEM, and hardness tests) using a dual-source DC magnetron sputtering system (Figure 1). Prior to insertion into the vacuum chamber, all substrates underwent a rigorous cleaning protocol to ensure surface cleanliness and promote strong film adhesion. First, the 9Cr18 discs were mechanically polished with successive grades of SiC abrasive paper, then mirror-finished. They were subsequently ultrasonically degreased in metal-detergent solution, rinsed in anhydrous ethanol for 20 min, and finally in deionized water for another 20 min. After each bath, the substrates were blow-dried with hot air and loaded into the sputtering chamber. Silicon wafers were cleaned in an identical fashion (40 min in ethanol, 20 min in deionized water) to monitor film thickness and composition. Once the base pressure in the chamber fell below 5 × 10−3 Pa, the sputtering targets were pre-cleaned by Ar plasma. The substrates were then etched in situ by Ar+ ion bombardment at a flow rate of 500 sccm (99.999% Ar) under −1000 V bias for 20 min to remove any residual oxides and adsorbates. During deposition, a rectangular graphite target (300 mm × 75 mm × 6 mm) and a silicon target of identical size were powered independently by two DC magnetrons. The substrates were aligned parallel to the targets. The reactive gas mixture of Ar and C2H2 was maintained at a flow ratio of 4:1 (92 sccm Ar, 23 sccm C2H2). The graphite target power was fixed at 1000 W, while the substrate bias was fixed at −100 V. Si incorporation was controlled by varying the Si-target power between 50, 100, and 200 W. A thin Si interlayer was deposited to improve the adhesion between the substrate and films. The distance between the target and substrates was approximately 80 mm. All films were grown for 120 min. The deposition parameters of films are provided in Table 1.

2.2. Film Characterization

The as-deposited films’ surface morphology, cross-sectional thickness, and elemental distributions were examined using a field-emission scanning electron microscope (FE-SEM; TESCAN MIRA4, Brno, Czechia) equipped with an energy-dispersive X-ray spectrometer (EDS). Film bonding configurations, both before and after high-temperature tribological testing, were probed by Raman spectroscopy (Renishaw InVia, λ = 532 nm, Wotton-under-Edge, UK). Detailed microstructural information was obtained by high-resolution transmission electron microscopy (HRTEM; Talos F200X, Thermo Fisher Scientific, Waltham, MA, USA). Mechanical properties were characterized via Knoop microhardness testing (MH-5LD, Shanghai Optical Instrument Factory, Shanghai, China) under a load of 25 gf and a dwell time of 15 s. To minimize measurement scatter, at least five indentations were made at different locations on each sample; reported values thus represent the composite hardness of the coating and substrate.
Room-temperature tribological performance was evaluated on an A & P TRB3 ball-on-disk tribometer. A 9Cr18 stainless-steel ball (6.35 mm in diameter) served as the counterpart material. The wear test was conducted under a normal load of 5 N at 400 rpm and a track radius of 6 mm for 60 min in an ambient environment with a relative humidity of 60%. High-temperature friction tests were carried out on a UMT-3 tribometer (BrukerCETR, Billerica, MA, USA) equipped with a furnace module. Tests were performed at 300 °C and 400 °C under a 5 N load, 200 rpm (track radius 5 mm), using a 6.35 mm Al2O3 ceramic ball as the counterpart. Wear-track morphology and worn volume were measured by optical metallography and white-light interferometry. Wear rates were calculated according to the Archard equation:
k = V/F·S,
where k is the wear rate (mm3·N−1·m−1), V the worn volume (mm3), F the normal load (N), and S the sliding distance (m) [22].

3. Results and Discussion

Figure 2a presents the elemental composition of a-C:H:Si films deposited at various target powers of 50 W, 100 W, and 200 W, as determined by EDS. Carbon originates from both the C2H2 precursor gas and the graphite sputtering target, silicon from the Si target, and oxygen from residual O2 in the deposition chamber as well as post-deposition atmospheric adsorption. Since hydrogen cannot be quantified by EDS, the reported atom fractions are normalized to the sum of C + Si + O. As the Si target power increases from 50 W to 200 W, the Si content rises from 4.6 at.% to 14.7 at.%, while C correspondingly decreases from 89.8 at.% to 83.4 at.%. The increase in Si content is attributed to the deeper sputtering of Si atoms at higher power, yielding a greater silicon flux to the growing film. Figure 2b shows the corresponding thicknesses and deposition rates of films. Compared with the undoped a-C:H film, all a-C:H:Si coatings exhibit lower thicknesses and deposition rates under identical sputtering conditions. However, as the Si target power increases, both thickness and deposition rate increase gradually. This behavior arises from a balance between two competing effects: on one hand, energetic Si ions under negative substrate bias can sputter loosely bound hydrocarbon species from the film’s surface, reducing film growth; on the other hand, higher Si target power enhances plasma density and precursor ionization, which promotes film growth. At low Si power, the sputtering (etching) effect dominates, whereas at higher power, the increased ionization and film-forming flux prevail, resulting in thicker films and faster deposition.
Surface and cross-sectional SEM micrographs of the as-deposited films are presented in Figure 3. All films exhibit a compact columnar structure without apparent surface cracks, delamination, or large voids. The film surfaces are characterized by a typical nodular (spherical-cluster) morphology. Compared to the undoped a-C:H coating, the a-C:H:Si films display noticeably larger, more three-dimensional clusters, whose size increases with Si content. Cross-section images further reveal that the columnar structure of the a-C:H:Si films is significantly more compact than that of a-C:H film. This densification is attributed to the increased discharge power applied to the Si target, which enhances the overall sputtering flux and promotes higher energy ion bombardment at the substrate surface, thereby improving surface mobility of adatoms and leading to a denser microstructure [23].
For further investigation of the microstructure of the film, the TEM technique was adopted. As shown in Figure 4a, the low-magnification cross-sectional TEM image of the a-C:H:Si (8.3 at.% Si) film reveals two distinct layers: a thin interfacial transition zone and the overlying a-C:H:Si film. The interface is continuous and defect-free, with no evidence of delamination or cracking. Figure 4b presents elemental-mapping EDS images, demonstrating a uniform spatial distribution of C and Si throughout the film; the modest O signal arises from residual O2 in the deposition chamber. The high-resolution TEM micrograph in Figure 4c displays a featureless, glassy morphology characteristic of amorphous materials. Consistently, the selected-area electron diffraction (SAED) pattern in Figure 4d exhibits only diffuse halos and no discrete diffraction rings, confirming the fully amorphous nature of the a-C:H:Si film, as previously reported [17,24]. Additionally, the appearance of short-range, curved chain-like motifs within the amorphous matrix indicates a substantial fraction of sp2-C bonds [25].
Figure 5 presents the Raman spectra of the a-C:H:Si films. Each spectrum exhibits a broad envelope around 1500 cm−1 that can be deconvoluted into two Gaussian components: The D peak near 1350 cm−1, arising from the stretching vibration of sp2-C bond sites in rings and chains, and the G peak near 1580 cm−1, corresponding to the in-plane breathing mode of sp2-C bond in graphitic rings [26,27] (Figure 5a). Three key parameters are used to quantify structural disorder: the G-peak full width at half maximum (FWHM), which reflects bond-angle and bond-length distortions; the G-peak position, which correlates with the size and topology of sp2 clusters; and the intensity ratio ID/IG, which indicates the relative abundance of ring-like versus chain-like sp2-C configurations, since a low ID/IG suggests predominantly chain-bound sp2 carbon with limited π-electron delocalization in rings [28].
As Si content increases, the G-peak FWHM first decreases from 158.94 cm−1 to 148.79 cm−1 and then rises to 156.33 cm−1, indicating an initial reduction in bond disorder followed by a slight reintroduction of structural distortion at the highest Si level. The G-peak position shifts upward from 1545.75 cm−1 to 1552.52 cm−1 before decreasing to 1548.19 cm−1, suggesting a nonmonotonic variation in sp2 cluster size and shape distribution. Likewise, ID/IG increases from 1.117 to 1.367, then decreases to 1.250, implying that moderate Si incorporation promotes the formation of larger graphitic clusters, whereas excessive Si begins to disrupt ring-like order. Collectively, these trends demonstrate that Si doping exerts a complex influence on the short- and medium-range ordering of the amorphous carbon network.
Figure 6 illustrates the Knoop microhardness of a-C:H:Si films with various Si contents. The undoped a-C:H coating exhibits a hardness of 557.0 ± 23.1 HK0.025. With 4.6 at.% Si, the hardness increases to 673.3 ± 23.0 HK0.025, reaching a maximum of 864.0 ± 42.1 HK0.025 at 8.3 at.% Si. Further increasing the Si content beyond this level results in a gradual decline in hardness. The initial hardness enhancement is attributed to Si’s preferential bonding with sp3-hybridized carbon, which raises the overall sp3 fraction and thus reinforces the film’s resistance to indentation. However, excessive Si doping leads to increased free volume and progressive stress relaxation within the amorphous matrix, both of which contribute to a reduction in film hardness at higher Si concentrations [16,17].
Figure 7a shows the friction coefficient of a-C:H:Si films with varying Si contents, measured at room temperature. All films exhibit an initial running-in period of approximately 400 s, during which the friction coefficient stabilizes near ~0.12, followed by a gradual, slight increase over the remainder of the 60 min test. Figure 7b summarizes the average friction coefficients and wear rates extracted from these tests. Incorporation of Si causes a modest increase in the mean friction coefficient, which rises from 0.116 for the undoped a-C:H film to a maximum of 0.126 at 8.3 at.% Si, then decreases slightly at higher Si levels. In contrast, the wear rate of film decreases initially with Si addition, reaching a minimum value of 0.78 × 10−7 mm3·N−1·m−1 at 8.3 at.% Si. Then increasing to 0.99 × 10−7 mm3·N−1·m−1 at 14.7 at.% Si. The combination of the highest hardness and the lowest wear rate was observed at 8.3 at.% Si confirms that hardness enhancement effectively mitigates material removal. Overall, Si doping exerts only a marginal influence on the tribological performance under ambient conditions.
Optical-microscope images of the wear tracks produced on a-C:H:Si films with varying Si contents under ambient conditions are shown in Figure 8. All films display well-defined wear scars, together with clearly visible transfer films and accumulated wear debris. The 8.3 at.% Si film exhibits the narrowest wear track, approximately at 132 µm in width, whereas the 4.6 at.% and 14.7 at.% Si films show somewhat wider tracks of 143 µm and 145 µm, respectively. The undoped a-C:H film endures the greatest damage, with a maximum track width of 153 µm. At low or zero Si content, insufficient formation of a lubricious C-Si-O transfer layer leads to more severe material removal; conversely, at high Si contents, in-situ generation of hard SiC particles during sliding accelerates abrasive wear [17].
High-temperature friction tests on a-C:H:Si films with different Si contents are shown in Figure 9. The tribological properties of films tested at high temperatures show more obvious differences compared with those tested at room temperature. The friction coefficient at 300 °C is shown in Figure 9a. Each film reached a relatively stable value after a period of the running-in stage. The friction coefficient of the film with 8.3 at.% Si began to stabilize at about 680 s and maintained an ultra-low friction coefficient of 0.01 all the time. In contrast, the friction coefficients of the undoped a-C:H film and other films fluctuated around 0.1 after stabilizing. Moreover, with the increase of Si content, the time required for the running-in stage also increases, from about 300 s at 4.6 at.% to about 900 s at 14.8 at.%. The running-in stage time of a-C:H film is comparable to the doping content of 8.3 at.% Si, which indicates that low-doping Si content is conducive to rapid running-in. The wear rates of each film at 300 °C are shown in Figure 9b. The a-C:H:Si film with 4.6 at.% Si content has the highest wear rate, which is 5.3 × 10−6 mm3 N−1·m−1, and the wear rate with 14.8 at.% Si content is 2.03 × 10−6 mm3 N−1·m−1. The a-C:H film has the lowest wear rate of 0.96 × 10−6 mm3 N−1·m−1. Notably, the a-C:H:Si film with a Si content of 8.3 at.% exhibits the lowest coefficient of friction while having a relatively low wear rate of 1.2 × 10−6 mm3 N−1·m−1.
The friction coefficient of films at 400 °C is shown in Figure 9c. The a-C:H films have the highest initial friction coefficient and the longest running-in period. After 230 s of running-in, the friction coefficients of a-C:H:Si films with Si contents of 4.6 at. % and 14.8 at.% remained at 0.01, which is the same as the a-C:H films during the stable stage. Meanwhile, the friction coefficient of a-C:H:Si films with 8.3 at.% Si gradually increased at the beginning. It reached a maximum of 0.068 at about 220 s, then gradually decreased, dropped below 0.01 at 870 s, and continued to decrease slowly, reducing to about 0.0034 at the end of the friction test. During the high-temperature friction process, the friction coefficient may occasionally show a sharp peak shape, and this phenomenon can be regarded as caused by the instability of the transfer film [29]. The wear rate at 400 °C is shown in Figure 9d. The wear rate of the a-C:H film is 4.6 × 10−6 mm3 N−1·m−1, and the wear rate of 4.6 at.% Si film is 3.32 × 10−6 mm3 N−1·m−1. The film wear rates of the films containing 8.3 at.% and 14.8 at.% Si are comparable, which are 2.52 × 10−6 mm3 N−1·m−1 and 2.35 × 10−6 mm3 N−1·m−1, respectively. It can thus be concluded that at 400 °C, the incorporation of silicon effectively reduces the wear rate of the amorphous carbon films, with the anti-wear performance becoming increasingly pronounced as the Si doping concentration rises.
Compared with the room temperature environment, the friction coefficient has been significantly reduced in the high-temperature environments of 300 °C and 400 °C, but the wear rate has increased. The reduction in friction coefficient is primarily driven by thermally induced graphitization and the formation of highly lubricious transfer films, which dramatically lower shear strength at the sliding interface. Meanwhile, the structural transformation of the a-C:H:Si thin film caused by high temperature leads to a decrease in mechanical properties, thereby resulting in an increase in the wear rate.
The wear track morphologies of the films after friction testing at 300 °C and 400 °C are shown in Figure 10. The depths of all wear tracks remained below the thickness of the deposited films, indicating that the films were not worn through during the tests. All films exhibited relatively narrow wear tracks. At 300 °C, the edges of the wear scars on both the undoped a-C:H film and the a-C:H:Si film with 4.6 at.% Si was smooth, whereas other samples displayed pronounced plastic deformation and raised edges. As the temperature increased to 400 °C, noticeable changes in wear morphology were observed. The wear track edges became flatter, but distinct ploughing grooves appeared within the wear tracks. This suggests a transition in the dominant wear mechanism from adhesive wear at 300 °C to abrasive wear at 400 °C [30].
At a test temperature of 400 °C, the a-C:H:Si film with 8.3 at.% Si exhibited outstanding tribological performance, characterized by both a low coefficient of friction and a low wear rate. As illustrated in Figure 11, the wear scar of the friction ball and wear track of the film were analyzed by SEM, EDS, and Raman spectroscopy to investigate the underlying mechanisms. As shown in Figure 11a, debris accumulation was observed near the wear scar, where localized enrichment of both carbon and silicon occurred. Similarly, Figure 11c reveals the presence of wear debris around the wear track, with comparable elemental distributions. The wear track itself appears smooth, suggesting shallow wear depth, consistent with results obtained from optical microscopy and white light interferometry. Elemental mapping further confirmed that the worn regions (both scar and track) were rich in Si, whereas the surrounding wear debris was predominantly composed of carbon. The superior performance of this film can be attributed to the protective role of silicon during high-temperature sliding. Silicon has a lower work function (4.85 eV) compared to carbon (5.0 eV), indicating a higher tendency to lose electrons and undergo oxidation. Consequently, Si atoms preferentially react with environmental oxygen or moisture, forming a passivating layer that protects the underlying carbon network from degradation. This contributes to the film’s enhanced lubricity and stability in oxidative atmospheres [25].
Raman spectra of the unworn surface, wear track, and wear scar (Figure 11b) indicate the formation of a tribolayer during sliding. Specifically, the wear scar region shows a pronounced amorphous carbon signal, suggesting transfer film formation in the contact zone. Compared with the unworn area, both the D peak near 1350 cm−1 becomes sharper and more intense, and the G peak near 1560 cm−1 increases in intensity and shifts to a higher wavenumber. These spectral changes suggest a structural transformation from sp3-C to sp2-C during the friction process, indicative of thermally induced graphitization. The growth of sp2 clusters under high temperature and stress may further promote the formation of graphene-like structures, enhancing the lubricating behavior of the tribolayer [31].
The above results demonstrate that silicon doping can effectively reduce the high-temperature friction coefficient of amorphous carbon (a-C:H) films and even exhibit unexpectedly outstanding performance when optimally doped. As illustrated in Figure 12, the Si-doped a-C:H:Si film possesses a coupled structure in which silicon atoms are incorporated into the sp2/sp3 amorphous carbon matrix. Notably, Si atoms preferentially bond with sp3-hybridized carbon atoms to form Si-C bonds, while the sp2-C network remains largely unaffected.
During sliding contact between the a-C:H:Si film and the counterface ball, the interfacial region undergoes complex processes including carbon dissociation, surface reactions, and interfacial interactions. These processes result in the adsorption and attachment of oxygen species, water molecules, and free amorphous carbon fragments to the counterface, ultimately forming a carbon-rich transfer layer. In contrast to conventional hard carbide-forming dopants that tend to disrupt the carbon matrix and elevate internal stresses—promoting delamination or acting as abrasive particles [32]—Si plays a more beneficial role.
Due to its lower work function, Si is more readily oxidized than carbon. During friction, this preferential oxidation of Si helps suppress the oxidation of the graphitic transfer layer, preserving its lubricating function. Furthermore, the generated oxides are often encapsulated within the graphitic matrix, preventing their direct participation in tribological interactions, thereby enhancing the stability and durability of the lubricating film. At the tribological interface, which is subject to both high temperature and mechanical stress, Si atoms can also react with water molecules in air, forming bonds with -OH or -H groups [33,34]. These terminations serve to passivate dangling bonds in a manner similar to hydrogen atoms already present in the film. The resulting H-H interactions across the sliding interface are associated with weak van der Waals forces, which contribute significantly to the low-friction behavior.
Additionally, under tribo-thermal activation, a structural reorganization occurs involving the transformation from sp3-C to sp2-C configurations. This graphitization process may lead to the formation of graphene-like domains [35,36,37], which are characterized by interlayer weak π*-π* interactions, another critical factor enabling superlubricity. Therefore, the low-friction behavior of Si-doped a-C:H:Si films at elevated temperatures can be attributed to the synergistic effect of interfacial H-H interactions and π*-π* stacking between graphene-like lamellae. Together, these mechanisms contribute to the formation of a stable, low-shear tribological interface that sustains excellent lubricating performance under harsh thermal–mechanical conditions [38].

4. Conclusions

This study addresses the temperature sensitivity limitation of hydrogenated amorphous carbon (a-C:H) films by exploring the effects of silicon doping on their high-temperature tribological performance. A series of a-C:H:Si films with varying Si content were deposited via direct current magnetron sputtering, and their microstructure, mechanical properties, and tribological behavior were systematically investigated at elevated temperatures (300 °C and 400 °C). The findings demonstrate that moderate Si doping significantly enhances the wear resistance and tribological performance of a-C:H films. Notably, the film with 8.3 at.% Si exhibited an ultralow friction coefficient of 0.0034 at 400 °C, which is attributed to the synergistic effects of a carbon-rich transfer layer formation, preferential Si oxidation, and tribo-induced phase transformation to graphitic structures at the sliding interface. The results also reveal that Si content influences film hardness, with the maximum hardness observed at 8.3 at.% Si, followed by a decrease at higher doping levels. These findings emphasize the critical role of controlled Si doping in tailoring the structural and tribological properties of a-C:H:Si films, providing new insights for the development of high-performance films in high-temperature environments.

Author Contributions

Writing—original draft preparation, Z.Y.; methodology, J.S. and H.Z.; software, Q.W.; formal analysis, H.M.; investigation and data curation, D.Z. and X.L.; writing—review and editing and supervision, J.D.; project administration, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Anhui Postdoctoral Scientific Research Program Foundation (No.2024C988) and the Open Project of China International Science and Technology Cooperation Base on Intelligent Equipment Manufacturing in Special Service Environment (No. ISTC2023KF02).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Zhen Yu was employed by the company AECC Zhongchuan Transmission Machiney Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relation-ships that could be construed as a potential conflict of interest.

References

  1. Zhu, S.; Cheng, J.; Qiao, Z.; Yang, J. High temperature solid-lubricating materials: A review. Tribol. Int. 2019, 133, 206–223. [Google Scholar] [CrossRef]
  2. Sliney, H.E. Solid lubricant materials for high temperatures—A review. Tribol. Int. 1982, 15, 303–315. [Google Scholar] [CrossRef]
  3. Sarkar, M.; Mandal, N. Solid lubricant materials for high temperature application: A review. Mater. Today Proc. 2022, 66, 3762–3768. [Google Scholar] [CrossRef]
  4. Ni, W.; Cheng, Y.-T.; Weiner, A.M.; Perry, T.A. Tribological behavior of diamond-like-carbon (DLC) coatings against aluminum alloys at elevated temperatures. Surf. Coat. Technol. 2006, 201, 3229–3234. [Google Scholar] [CrossRef]
  5. Deng, X.; Kousaka, H.; Tokoroyama, T.; Umehara, N. Thermal Stability and High-Temperature Tribological Properties of a-C:H and Si-DLC Deposited by Microwave Sheath Voltage Combination Plasma. Tribol. Online 2013, 8, 257–264. [Google Scholar] [CrossRef]
  6. Erdemir, A.; Donnet, C. Tribology of Diamond and Diamond-like Carbon Films. In Wear–Materials, Mechanisms and Practice; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2005; pp. 191–222. [Google Scholar]
  7. Liu, H.; Tanaka, A.; Umeda, K. The tribological characteristics of diamond-like carbon films at elevated temperatures. Thin Solid Films 1999, 346, 162–168. [Google Scholar] [CrossRef]
  8. Bhowmick, S.; Banerji, A.; Khan, M.Z.U.; Lukitsch, M.J.; Alpas, A.T. High temperature tribological behavior of tetrahedral amorphous carbon (ta-C) and fluorinated ta-C coatings against aluminum alloys. Tribol. Int. 2015, 284, 14–25. [Google Scholar] [CrossRef]
  9. Yu, W.; Huang, W.; Wang, J.; Su, Y.; Long, Q.; Wang, L.; Zhu, L. High-temperature tribological performance of the Si-gradually doped diamond-like carbon film. Vacuum 2021, 191, 110387. [Google Scholar] [CrossRef]
  10. Vitu, T.; Escudeiro, A.; Polcar, T.; Cavaleiro, A. Sliding properties of Zr-DLC coatings: The effect of tribolayer formation. Surf. Coat. Technol. 2014, 258, 734–745. [Google Scholar] [CrossRef]
  11. Santiago, J.A.; Fernández-Martínez, I.; Sánchez-López, J.C.; Rojas, T.C.; Wennberg, A.; Bellido-González, V.; Molina-Aldareguia, J.M.; Monclús, M.A.; González-Arrabal, R. Tribomechanical properties of hard Cr-doped DLC coatings deposited by low-frequency HiPIMS. Tribol. Int. 2020, 382, 124899. [Google Scholar] [CrossRef]
  12. Zheng, J.; Shang, J.; Zhuang, W.; Ding, J.C.; Mei, H.; Yang, Y.; Ran, S. Structural and tribomechanical properties of Cr-DLC films deposited by reactive high power impulse magnetron sputtering. Vacuum 2024, 230, 113611. [Google Scholar] [CrossRef]
  13. Yu, W.; Huang, W.; Wang, J.; Su, Y.; Gong, X. Further improving the mechanical and high-temperature tribological properties of silicon-doped diamond-like carbon film by tungsten incorporation. Surf. Rev. Lett. 2024, 31, 2450035. [Google Scholar] [CrossRef]
  14. Zhao, D.; Mei, H.; Ding, J.C.; Cheng, Y.; Zhang, L.; Zhang, T.F.; Kwang, H.K.; Zheng, J. Microstructure and properties of Mo doped DLC nanocomposite films deposited by a hybrid sputtering system. Vacuum 2023, 208, 111732. [Google Scholar] [CrossRef]
  15. Li, X.-W.; Joe, M.-W.; Wang, A.-Y.; Lee, K.-R. Stress reduction of diamond-like carbon by Si incorporation: A molecular dynamics study. Tribol. Int. 2013, 228, S190–S193. [Google Scholar] [CrossRef]
  16. Sharifahmadian, O.; Pakseresht, A.; Mosas, K.K.A.; Galusek, D.; Wang, L.; Liu, Y.; Chen, H.; Wang, M.; Zhang, S.; Yan, M.; et al. Doping effects on the tribological performance of diamond-like carbon coatings: A review. J. Mater. Res. Technol. 2023, 27, 7748–7765. [Google Scholar] [CrossRef]
  17. Zhang, T.F.; Wan, Z.X.; Ding, J.C.; Zhang, S.; Wang, Q.M.; Kim, K.H. Microstructure and high-temperature tribological properties of Si-doped hydrogenated diamond-like carbon films. Appl. Surf. Sci. 2018, 435, 963–973. [Google Scholar] [CrossRef]
  18. Chen, Y.; Su, F.; Li, Q.; Sun, J.; Lin, S.; Ma, G. Extremely enhanced friction and wear performance of hydrogen-free DLC film at elevated temperatures via Si doping. Tribol. Int. 2024, 199, 109981. [Google Scholar] [CrossRef]
  19. Wang, L.; Li, H.; Pu, C.; Wang, L.; Wang, J. Elevated temperature tribological properties of a-C:Si films sputtered from SiC target in the presence of C2H2 gas. Mater. Today Commun. 2024, 38, 108226. [Google Scholar] [CrossRef]
  20. Lu, S.; Duan, F. High Temperature Induced Low Friction and Wear in a-C:Si via Formation of a “H Passivation Layer/a-SiO2/H Passivation Layer” Structure in a Humid Environment. Langmuir 2024, 40, 18078–18088. [Google Scholar] [CrossRef]
  21. Li, Z.; Xu, M.; Zhang, H.; He, W.; Zhang, G.; Lu, Z. Effect of silicon-doping on the wide-temperature tribological behavior and lubrication mechanism of WC/a-C film. Wear 2023, 516–517, 204614. [Google Scholar] [CrossRef]
  22. Merkle, A.P.; Erdemir, A.; Eryilmaz, O.L.; Johnson, J.A.; Marks, L.D.; Archard, J.F. Contact and Rubbing of Flat Surfaces. J. Appl. Phys. 1953, 24, 981–988. [Google Scholar]
  23. Ding, J.C.; Dai, W.; Zhang, T.F.; Zhao, P.; Yun, J.M.; Kim, K.H.; Wang, Q.M.; Kajita, S.; Righi, M.C. Microstructure and properties of Nb-doped diamond-like carbon films deposited by high power impulse magnetron sputtering. Thin Solid Films 2018, 663, 159–167. [Google Scholar] [CrossRef]
  24. Kim, J.-I.; Jang, J.-W.; Umehara, N.; Jang, Y.-J. Counterpart-dependent friction and wear of hydrogenated tetrahedral amorphous carbon under ultra-high vacuum. Carbon 2025, 238, 120223. [Google Scholar] [CrossRef]
  25. Ma, Q.; Fu, Y.; Xu, Y.; Ma, T. Achieving superlubricity and high adhesion strength of hydrogenated amorphous carbon film with Al/Cr/Si-doping. Carbon 2023, 215, 118424. [Google Scholar] [CrossRef]
  26. Ferrari, A.C.; Robertson, J. Resonant Raman spectroscopy of disordered, amorphous, and diamondlike carbon. Phys. Rev. B 2001, 64, 075414. [Google Scholar] [CrossRef]
  27. Liskiewicz, T.; Yang, J.; Chen, Z.; Yuan, Y.; Chen, Q.; Feng, X.; Qin, L.; Zhang, Y. Tuning C-C sp2/sp3 ratio of DLC films in FCVA system for biomedical application. Bioact. Mater. 2020, 5, 192–200. [Google Scholar]
  28. Ferrari, A.C.; Robertson, J. Interpretation of Raman spectra of disordered and amorphous carbon. Phys. Rev. B 2000, 61, 14095–14107. [Google Scholar] [CrossRef]
  29. Jantschner, O.; Field, S.K.; Holec, D.; Fian, A.; Music, D.; Schneider, J.M.; Zorn, K.; Mitterer, C. Origin of temperature-induced low friction of sputtered Si-containing amorphous carbon coatings. Acta Mater. 2015, 82, 437–446. [Google Scholar] [CrossRef]
  30. Dong, Y.; Zhao, Y.; Li, J.; Chen, J.; Zheng, J.; Sun, D.; Zhang, S.; Wang, S. Super-hard and well-tough (TiZrVCrCoNi)Nx high entropy nitride coatings with biphasic nanocomposite structure. Tribol. Int. 2024, 224, 113110. [Google Scholar] [CrossRef]
  31. Chen, C.; Xue, P.; Fan, X.; Wang, C.; Diao, D. Friction-induced rapid restructuring of graphene nanocrystallite cap layer at sliding surfaces: Short run-in period. Carbon 2018, 130, 215–221. [Google Scholar] [CrossRef]
  32. He, F.; Aebersold, R.; Baker, M.S.; Bian, X.; Bo, X.; Chan, D.W.; Chang, C.; Chen, L.; Chen, X.; Chen, Y.-J.; et al. Structure, Mechanical and Tribological Properties of Me-Doped Diamond-like Carbon (DLC) (Me = Al, Ti, or Nb) Hydrogenated Amorphous Carbon Coatings. Coatings 2018, 8, 370. [Google Scholar]
  33. Tran, N.V.; Righi, M.C. Ab initio insights into the interaction mechanisms between H2, H2O, and O2 molecules with diamond surfaces. Carbon 2022, 487, 497–507. [Google Scholar] [CrossRef]
  34. Kajita, S.; Righi, M.C. Insigths into the Tribochemistry of Silicon-doped Carbon-Based Films by Ab Initio Analysis of Water–Surface Interactions. Tribol. Lett. 2016, 61, 17. [Google Scholar] [CrossRef]
  35. Venturi, F. Carbon-based films: A review on mechanical and tribology properties. Mater. Chem. Phys. 2024, 325, 129716. [Google Scholar] [CrossRef]
  36. Wang, Y.; Gao, K.; Zhang, B.; Wang, Q.; Zhang, J. Structure effects of sp2-rich carbon films under super-low friction contact. Mater. Today Proc. 2018, 137, 49–56. [Google Scholar] [CrossRef]
  37. Ma, T.-B.; Wang, L.-F.; Hu, Y.-Z.; Li, X.; Wang, H. A shear localization mechanism for lubricity of amorphous carbon materials. Tribol. Int. 2015, 4, 3662. [Google Scholar] [CrossRef]
  38. Sun, K.; Sun, M.; Chen, C.; Chen, S.; Fan, J.; Diao, D. Structural and tribological behaviours of silicon doped amorphous carbon films. Carbon 2023, 215, 118434. [Google Scholar] [CrossRef]
Figure 1. Schematic of a magnetron sputtering deposition system.
Figure 1. Schematic of a magnetron sputtering deposition system.
Coatings 15 00793 g001
Figure 2. (a) The element composition, and (b) deposition rate of a-C:H:Si films with various Si target powers.
Figure 2. (a) The element composition, and (b) deposition rate of a-C:H:Si films with various Si target powers.
Coatings 15 00793 g002
Figure 3. Surface and cross-sectional morphology of a-C:H:Si films with various Si contents; (a) 0 at.%; (b) 4.6 at.%; (c) 8.3 at.%; (d) 14.7 at.%.
Figure 3. Surface and cross-sectional morphology of a-C:H:Si films with various Si contents; (a) 0 at.%; (b) 4.6 at.%; (c) 8.3 at.%; (d) 14.7 at.%.
Coatings 15 00793 g003
Figure 4. TEM images of a-C:H:Si films with Si content of 8.3 at.%: (a) bright-field cross-section image; (b) cross-section elemental distribution mapping; (c) HRTEM image; (d) selected area electron diffraction (SAED) image.
Figure 4. TEM images of a-C:H:Si films with Si content of 8.3 at.%: (a) bright-field cross-section image; (b) cross-section elemental distribution mapping; (c) HRTEM image; (d) selected area electron diffraction (SAED) image.
Coatings 15 00793 g004
Figure 5. Raman spectroscopy of a-C:H:Si films with various Si contents:. (a) Raman spectrum; (b) FWHM of G peak, position G peak, and ID/IG ratio.
Figure 5. Raman spectroscopy of a-C:H:Si films with various Si contents:. (a) Raman spectrum; (b) FWHM of G peak, position G peak, and ID/IG ratio.
Coatings 15 00793 g005
Figure 6. Microhardness of films with different Si contents.
Figure 6. Microhardness of films with different Si contents.
Coatings 15 00793 g006
Figure 7. (a) The friction coefficient curves, and (b) wear rate of a-C:H:Si films with different Si contents.
Figure 7. (a) The friction coefficient curves, and (b) wear rate of a-C:H:Si films with different Si contents.
Coatings 15 00793 g007
Figure 8. Morphology of wear tracks on a-C:H:Si films with different Si contents: (a) 0 at.%; (b) 4.6 at.%; (c) 8.3 at.%; (d) 14.7 at.%.
Figure 8. Morphology of wear tracks on a-C:H:Si films with different Si contents: (a) 0 at.%; (b) 4.6 at.%; (c) 8.3 at.%; (d) 14.7 at.%.
Coatings 15 00793 g008
Figure 9. Friction coefficient and wear rate of a-C:H:Si films with different Si contents in high temperature conditions: (a) coefficient of friction and (b) wear rate under 300 °C; (c) coefficient of friction and (d) wear rate under 400 °C.
Figure 9. Friction coefficient and wear rate of a-C:H:Si films with different Si contents in high temperature conditions: (a) coefficient of friction and (b) wear rate under 300 °C; (c) coefficient of friction and (d) wear rate under 400 °C.
Coatings 15 00793 g009
Figure 10. The wear tracks, 3D image of the wear tracks, and wear depths for films with various Si contents after high temperature friction test: (a) 0 at.% Si at 300 °C; (b) 4.6 at.% Si at 300 °C; (c) 8.3 at.% Si at 300 °C; (d) 14.7 at.% Si at 300 °C; (e) 0 at.% Si at 400 °C; (f) 4.6 at.% Si at 400 °C; (g) 8.3 at.% Si at 400 °C; (h) 14.7 at.% Si at 400 °C.
Figure 10. The wear tracks, 3D image of the wear tracks, and wear depths for films with various Si contents after high temperature friction test: (a) 0 at.% Si at 300 °C; (b) 4.6 at.% Si at 300 °C; (c) 8.3 at.% Si at 300 °C; (d) 14.7 at.% Si at 300 °C; (e) 0 at.% Si at 400 °C; (f) 4.6 at.% Si at 400 °C; (g) 8.3 at.% Si at 400 °C; (h) 14.7 at.% Si at 400 °C.
Coatings 15 00793 g010
Figure 11. Characterization of wear scar on friction ball and wear track of a-C:H:Si film with Si content of 8.3 at.% at 400 °C: (a) Morphology of wear scar and mapping of C, Si and O elements; (b) Raman spectra of unworn area, wear scar and wear track of the film; (c) Morphology of wear track and C, Si element mapping.
Figure 11. Characterization of wear scar on friction ball and wear track of a-C:H:Si film with Si content of 8.3 at.% at 400 °C: (a) Morphology of wear scar and mapping of C, Si and O elements; (b) Raman spectra of unworn area, wear scar and wear track of the film; (c) Morphology of wear track and C, Si element mapping.
Coatings 15 00793 g011
Figure 12. The high-temperature friction mechanism of a-C:H:Si film.
Figure 12. The high-temperature friction mechanism of a-C:H:Si film.
Coatings 15 00793 g012
Table 1. Parameters of film deposition.
Table 1. Parameters of film deposition.
Parametersa-C:H:Si
Base pressure (Pa)5 × 10−3
Working pressure (Pa)~0.6
Ar/C2H2 ratio4:1
Working temperature (°C)120
Bias voltage (V)
Substrate rotation (r/min)
−100
3
Deposition time (min)120
Carbon target power (W)1000
Silicon target power (W)0/50/100/200
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

Yu, Z.; Shang, J.; Wang, Q.; Zheng, H.; Mei, H.; Zhao, D.; Liu, X.; Ding, J.; Zheng, J. Influence of Si Content on the Microstructure and Properties of Hydrogenated Amorphous Carbon Films Deposited by Magnetron Sputtering Technique. Coatings 2025, 15, 793. https://doi.org/10.3390/coatings15070793

AMA Style

Yu Z, Shang J, Wang Q, Zheng H, Mei H, Zhao D, Liu X, Ding J, Zheng J. Influence of Si Content on the Microstructure and Properties of Hydrogenated Amorphous Carbon Films Deposited by Magnetron Sputtering Technique. Coatings. 2025; 15(7):793. https://doi.org/10.3390/coatings15070793

Chicago/Turabian Style

Yu, Zhen, Jiale Shang, Qingye Wang, Haoxiang Zheng, Haijuan Mei, Dongcai Zhao, Xingguang Liu, Jicheng Ding, and Jun Zheng. 2025. "Influence of Si Content on the Microstructure and Properties of Hydrogenated Amorphous Carbon Films Deposited by Magnetron Sputtering Technique" Coatings 15, no. 7: 793. https://doi.org/10.3390/coatings15070793

APA Style

Yu, Z., Shang, J., Wang, Q., Zheng, H., Mei, H., Zhao, D., Liu, X., Ding, J., & Zheng, J. (2025). Influence of Si Content on the Microstructure and Properties of Hydrogenated Amorphous Carbon Films Deposited by Magnetron Sputtering Technique. Coatings, 15(7), 793. https://doi.org/10.3390/coatings15070793

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