Next Article in Journal
Vibration Analysis of Functionally Graded Material (FGM) Double-Layered Cabin-like Structure by the Spectro-Geometric Method
Next Article in Special Issue
Effects of Sn Addition and Fe Content Adjustment on the Decolorization Performance of Fe-Si-B Amorphous Alloys for Azo Dyes
Previous Article in Journal
Assessment of the Corrosion Resistance of Thermal Barrier Coatings on Internal Combustion Engine Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation on High-Temperature Wear Resistance of Co-Based Superalloys Modified by Chromium–Aluminizing Coatings

1
School of Materials Science and Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
2
Key Laboratory of Advanced Marine Materials, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(6), 1229; https://doi.org/10.3390/ma18061229
Submission received: 7 February 2025 / Revised: 26 February 2025 / Accepted: 7 March 2025 / Published: 10 March 2025
(This article belongs to the Special Issue Smart Coatings for the Corrosion Protection of Alloys)

Abstract

:
This study systematically explores and expands upon the research questions, revealing the scientific principles and engineering value of chromium–aluminum (Cr-Al) co-diffusion coatings in enhancing high-temperature friction performance. This study addresses the critical need for wear resistance in GH5188 cobalt-based alloy stator bushings operating in high-temperature environments. The high-temperature wear resistance mechanism of aluminized coatings modified with Cr elements on the GH5188 alloy, based on thermal diffusion technology, was investigated. The experimental results indicate that the high-temperature wear resistance of the samples was directly related to the type and content of oxides in the wear scars and debris. After friction at 700 °C, the aluminized coating on the GH5188 alloy showed the lowest oxide content in the wear scars, primarily composed of CoAl2O4. The oxides in the wear scars of the GH5188 alloy and Al-Cr co-aluminized coatings were mainly CoCr2O4 and Cr2O3, with the Al-Cr co-aluminized coating showing the highest amount of wear debris. The Cr-rich oxide debris not only has high thermodynamic stability but also exhibits relatively low high-temperature growth stress, making it difficult to spall. Additionally, the higher diffusion coefficient of Cr3+ accelerates the reoxidation of wear debris pits, resulting in excellent high-temperature wear resistance. The wear rate of the Al-Cr co-aluminized coating was reduced by 30% compared with the GH5188 substrate and by 69% compared with the aluminized coating. In summary, the key findings are not only applicable to cobalt-based alloys but can also be extended to a broader range of material systems and engineering applications. This provides new perspectives and methodologies for the design of high-temperature coatings, the development of materials for extreme conditions, and interdisciplinary applications.

1. Introduction

The GH5188 cobalt-based superalloy is widely used in advanced aero engine components such as stator bushings and turbine sealing rings due to its excellent high-temperature strength (>800 °C), superior thermal corrosion resistance, and good thermal fatigue resistance [1]. With the continuous increase in the thrust-to-weight ratio and turbine inlet temperature of aero engines, the operating environment for stator bushings has become increasingly severe. These components must endure long-term exposure to high-temperature gas environments (700–900 °C) while facing centrifugal loads, wear, and high-temperature adhesion [2], leading to losses of dimensional accuracy and potential seizure failures, which severely threaten engine reliability [3]. Although the traditional GH5188 alloy has been optimized for high-temperature mechanical properties through solution strengthening and carbide precipitation, its tribological performance still fails to meet the longevity demands of new-generation engines [4]. Therefore, developing surface functional coatings with high-temperature wear resistance has become a key strategy for enhancing the tribological performance of the GH5188 alloy [5].
Preparing an aluminizing coating using the solid powder embedding method is an important approach to enhance the high-temperature friction performance, surface properties, and reinforcement of cobalt-based superalloys [6]. In recent years, aluminizing coatings on the GH5188 alloy have demonstrated excellent oxidation resistance at 700 °C, with the formation of a continuous α-Al2O3 film on the surface effectively inhibiting oxygen diffusion into the substrate [7]. Additionally, the aluminized layer can reduce the risk of high-temperature adhesion by decreasing atomic interdiffusion tendencies between friction pairs, a characteristic particularly important in the blade-casing contact pairs of aero engines [8,9]. However, existing studies mainly focus on the oxidation resistance and anti-adhesion properties of aluminizing coatings, with significant gaps in the systematic study of their high-temperature tribological behavior. For example, Chen et al. [10] found that the brittleness of the α-Al2O3 film in aluminizing coatings leads to early spallation during high-temperature sliding wear, exposing unoxidized metal surfaces and inducing abrasive wear. Zhang et al. [11] further pointed out that the hardness mismatch between the aluminized layer and nickel-based counterpart exacerbates interfacial plastic deformation and material transfer, increasing the wear rate to 2–3 times that of the substrate alloy. Critically, the tribological performance of aluminizing coatings is closely related to their microstructure, yet there is a lack of in-depth exploration of the relationships among process parameters, coating structure, and wear mechanisms in the literature [12]. These limitations indicate that a single aluminizing coating cannot achieve the synergistic optimization of anti-adhesion and wear resistance properties, necessitating breakthroughs through multi-alloy modification to overcome performance bottlenecks.
To overcome the performance limitations of single aluminizing coatings, researchers have proposed the introduction of Cr elements to construct Al-Cr co-deposited coatings. This approach aims to achieve a synergistic enhancement of anti-adhesion and wear resistance properties [13]. Theoretical analysis and experimental studies indicate that the introduction of Cr improves coating performance through the following mechanisms. The (Al, Cr)2O3 composite oxide film, formed by Cr2O3 and Al2O3, exhibits higher density and self-healing capabilities, significantly delaying the rupture of the oxide film at high temperatures [14,15,16]. Solid solution strengthening by Cr and nano-scale Cr-rich precipitates increase the coating hardness (HV ≥ 600), enhancing its resistance to plastic deformation and abrasive wear [17,18,19]. Additionally, the addition of Cr can adjust the thermal expansion coefficient compatibility between the coating and substrate (Δα reduced by 15–20%), reducing interfacial stress concentration during high-temperature friction processes and thus inhibiting coating spallation [20].
Current research on the introduction of chromium (Cr) to form Cr-Al co-diffusion coatings primarily focuses on nickel-based alloys to achieve a synergistic improvement in anti-adhesion and wear resistance. However, systematic studies on Cr-Al co-diffusion coatings for GH5188 cobalt-based alloys have been overlooked, and the influence of Cr-Al coatings on high-temperature friction interfaces, such as the competition between oxidation and wear mechanisms, remains unclear [21,22,23,24]. This work specifically addresses the challenges of controlling high-temperature wear mechanisms by investigating the evolution of oxide compositions and valence states at the friction interface, achieved through the preparation of Cr-Al co-diffusion coatings via pack cementation. In terms of tribological behavior, studies by Su et al. [25] on nickel-based alloys have shown that Al-Cr co-deposited coatings can reduce high-temperature wear rates by more than 40% compared with single aluminizing coatings. This reduction is attributed to the enhanced toughness of the oxide film and the synergistic lubricating effect of friction–chemical reaction films (such as CrO3-Al2O3 mixed layers) promoted by Cr [26,27]. However, these studies have primarily focused on nickel-based alloy systems. Systematic research on Al-Cr co-deposited coatings for GH5188 cobalt-based alloys remains extremely limited [28]. Currently, the mechanisms by which Cr-modified aluminizing layers on cobalt-based alloys affect high-temperature frictional interface reactions (such as the oxidation-wear competition mechanism) are not yet clear [29].
In summary, this work focuses on the study of the friction and wear behaviors and mechanisms of the GH5188 alloy and its aluminized and Al-Cr co-deposited coatings at 700 °C. Aluminized and Al-Cr co-deposited coatings were prepared on the surface of the GH5188 alloy using the solid powder embedding method. The tribological performance of the GH5188 alloy substrate, aluminized coating, and Al-Cr co-deposited coating was evaluated using a high-temperature friction and wear system. The laser scanning confocal microscopy (LSCM), field emission scanning electron microscopy (FESEM), X-ray photoelectron spectroscopy (XPS), and X-ray diffraction (XRD) techniques were used to elucidate the composition of oxidation products, evolution of chemical valence states at the friction interface, and their role in wear mechanisms. This study provides scientific evidence for the composition design and process optimization of protective surface coatings for high-temperature components in aero engines.

2. Experimental Materials and Methods

2.1. Experimental Materials

In this study, the commercially available GH5188 (Zhejiang China) cobalt-based high-temperature alloy was selected as the research subject. The primary elements and their contents are detailed in Table 1.
In this study, the GH5188 alloy was used as the substrate, with sample dimensions of 20 mm × 20 mm × 3 mm. Aluminized and Al-Cr co-deposited coatings were prepared on the GH5188 alloy surface using the solid powder embedding method. The aluminizing agent components included aluminum powder as the aluminum source, ammonium chloride as the activator, and alumina as the filler. For the Al-Cr co-deposition, aluminum powder served as the aluminum source, chromium oxide as the chromium source, ammonium chloride as the activator, and alumina as the filler. Different heat diffusion parameters led to varying coating thicknesses without altering the phase structure. Therefore, theoretically, these parameters should not have significantly affected the high-temperature friction performance of the coatings [30]. In this study, we discuss the friction performance of coatings prepared with specific co-diffusion parameters. The specific compositions of the embedding agents and the parameters for the thermal diffusion experiments are detailed in Table 2.

2.2. Preparation of Coatings

After degreasing and cleaning with deionized water, the GH5188 alloy samples were sanded and polished using sandpaper up to 2000 grit. The surfaces were then ultrasonically cleaned in acetone for 10 min, rinsed with ethanol, and air-dried. The samples and prepared embedding agents were placed in a crucible, ensuring no contact between the samples and the crucible walls, as well as that the embedding agents completely covered the samples. The crucible was then sealed. Thermal diffusion experiments were conducted according to the parameters specified in Table 2 to prepare the aluminized and Al-Cr co-deposited coatings. After the experiments, the samples were furnace-cooled to room temperature. Throughout the process, argon gas was used as a protective atmosphere to prevent sample oxidation.

2.3. High-Temperature Friction Experiments

High-temperature friction experiments were conducted on the GH5188 alloy, aluminized coating, and Al-Cr coating samples in an air environment at 700 °C. The equipment used was a high-temperature friction and wear tester (THT 1000, Anton Paar, Graz, Austria). The experimental conditions were a load of 5 N, a friction radius of 3 mm, a friction speed of 10 mm/s, and a counter body of a 6 mm Si3N4 ball. After the friction tests, the wear scar profile dimensions were measured multiple times using laser confocal microscopy (VK-X1000, Keyence, Osaka, Japan). The average value of these measurements was used to determine the wear scar cross-sectional area S. The wear rate under different high-temperature friction conditions was then calculated using the following formula:
w = ( S × 2 π r ) / F × L
where S represents the wear scar cross-sectional area (μm2), r is the friction radius (mm), F is the applied load during friction (N), and L is the total friction distance (m).

2.4. Characterization and Performance Evaluation of Coating Structures

The microstructure and surface wear scar characteristics of the GH5188 alloy, aluminized coating, and Al-Cr co-diffusion coating before and after high-temperature friction were characterized using a thermal field scanning electron microscope (ZEISS-300, Baden-Württem, Germany) with an accelerating voltage of 15 kV. Elemental distribution in the diffusion layers and wear scar surfaces was analyzed using an energy dispersive spectrometer (EDS). The surface hardness and cross-sectional hardness gradients of different diffusion layers were measured using a Vickers hardness tester (VH-3000, Wilson, New York, NY, USA) with an indenter load of HV0.2. Each set of data was tested in triplicate, and the mean values and standard deviations were calculated. The phase analysis of the GH5188 alloy and diffusion layers was conducted using an X-ray diffractometer (ADVANCE D8, Keens, Tokyo, Japan) with a scanning range of 20° to 90° and a scan rate of 0.05°/s. The oxide structures of non-wear scar areas after high-temperature friction tests were analyzed and characterized using a two-dimensional small-angle X-ray scattering instrument (SAXS, Xeuss 3.0 UHR, XENOCS SAS, Paris, France) with an incident angle of 0.5°, a scanning range of 0° to 60°, and a scan rate of 0.05°/s. X-ray photoelectron spectroscopy (AXIS SUPRA+, Shimadzu, Tokyo, Japan) was employed to analyze the chemical states of oxides on the wear scar surfaces of the GH5188 alloy and the various coating samples after high-temperature friction tests, providing a more detailed understanding of the oxidation characteristics.

3. Results and Analysis

3.1. Cross-Sectional Morphology and Surface Phase Analysis of Coatings

The microstructure and elemental composition of materials are crucial factors affecting their friction performance. Therefore, this section examines the microstructure and elemental distribution of the GH5188 alloy, its aluminized coating, and Al-Cr co-diffusion coating. Figure 1 presents the cross-sectional morphology and elemental distribution of the GH5188 alloy, aluminized coating, and Al-Cr co-diffusion coating. The GH5188 alloy matrix contained uniformly distributed spherical W-rich compounds, with other major elements evenly distributed without segregation. After aluminum diffusion using the solid powder pack method, the aluminized coating on GH5188 alloy exhibited a thickness of approximately 57 μm. The coating was relatively dense, with a uniform distribution of Al, and contained a few W-rich compounds. A distinct interface was observed between the coating and the substrate due to Al not being a primary element of the GH5188 alloy. The Al-Cr co-diffusion coating on the GH5188 alloy had a thickness of 20 μm. This coating showed the distinct layering of Al and Cr elements, with Cr enrichment near the coating-substrate interface and Al enrichment in the middle of the coating. Some W-rich compound aggregation was observed near the interface close to the substrate. The phase composition of the three samples was characterized with XRD, as shown in Figure 2. The primary phase of the GH5188 matrix was an austenitic matrix structure composed of Co-Cr-Ni, with some Si2W dispersed strengthening phases. The aluminized coating contained a substantial amount of Co-Al intermetallic compounds, such as CoAl, Al5Co2, and Co4Al13, with minor amounts of Al4Ni3 and CrCo2Al intermetallic compounds. On the surface of the Al-Cr co-diffusion coating of the GH5188 alloy, Cr23C6 and minor Al and Ni intermetallic compounds were predominantly present.
The microstructure of different samples directly impacts their surface hardness, which in turn affects their tribological performance. To further analyze the impact of microstructure on hardness, the surface hardness of the GH5188 alloy, aluminized coating, and Al-Cr co-diffusion coating was compared. The results are shown in Figure 3. The hardness of the GH5188 substrate was measured at 326 HV0.2. The hardness of the aluminized coating was significantly higher at 595 HV0.2, while the Al-Cr co-diffusion coating exhibited the highest hardness at 935 HV0.2. Compared with the substrate, the hardness of the aluminized coating increased by approximately 82.5% and the hardness of the Al-Cr co-diffusion coating increased by approximately 186.8%.

3.2. Tribological Performance

During the course of this work, high-temperature friction and hardness tests were repeated three times under identical conditions for the same coating, and error bars were added to the relevant results for calibration. Figure 4 demonstrates the high-temperature friction and wear performance of the three samples at 700 °C. Considering that actual working conditions typically involve an initial running-in period, the first 100 s of the friction process were analyzed as the running-in stage, as shown in Figure 4b. After 1800 s of high-temperature friction, the friction coefficients of the three samples stabilized, as illustrated in Figure 4a. The GH5188 alloy exhibited a friction coefficient of approximately 0.65. The aluminized coating (GH5188-Al) had a friction coefficient fluctuating around 0.5, showing less stability compared with the GH5188 alloy. The Al-Cr co-diffusion coating had the lowest friction coefficient, around 0.45, and displayed a gradual decreasing trend over time. The friction coefficients of the aluminized and Al-Cr co-diffusion coatings were reduced by about 23% and 31%, respectively, compared with the substrate. In the first 100 s of the friction process, as shown in Figure 4b, the friction coefficients of all samples were generally higher and less stable. Particularly, the Al-Cr co-diffusion coating exhibited a slow decrease in the friction coefficient over time, indicating better stability during the friction process. However, evaluating the high-temperature friction performance of materials requires not only an assessment of the friction coefficient but also an examination of the wear rate. Figure 4c presents the wear rates of the three samples after 30 min of high-temperature friction at 700 °C. The aluminized coating exhibited the highest wear rate, approximately 83.314 × 10−1 mm3/N·m, significantly higher than the 36.824 × 10−1 mm3/N·m of the GH5188 alloy, with the wear rate of the aluminized coating being about 2.3 times that of the substrate. The Al-Cr co-diffusion coating had the lowest wear rate, only 25.706 × 10−1 mm3/N·m, which was 30.5% lower than that of the substrate. Details are shown in the Table 3.
A laser confocal microscope was employed to further analyze the wear conditions of the three samples after high-temperature friction. Figure 5a–d present the three-dimensional morphologies and cross-sectional profiles of the wear tracks. It was evident that the aluminized coating exhibited the greatest width and depth of wear tracks. The deepest point reached 45 μm, and the wear track width was approximately 1000 μm, characterized by a “deep and wide” morphology. Additionally, a significant accumulation of wear debris layers was observed on both sides of the wear tracks. In contrast, the wear tracks of the Al-Cr co-diffusion coating showed a smaller accumulation of wear debris and the shallowest wear depth, approximately 7 μm, which was only 15% of the depth observed in the aluminized coating, indicating superior wear resistance. The GH5188 alloy exhibited a maximum wear depth of about 11 μm and a wear track width of approximately 400 μm.
The depth and width of wear tracks are crucial indicators of a material’s wear resistance. Although both the aluminized coating and the Al-Cr co-diffusion coating showed significant improvements in friction coefficients, the wear rate analysis revealed different levels of performance. The aluminized coating, due to its lower hardness and inferior wear resistance, resulted in deeper and wider wear tracks. In contrast, the Al-Cr co-diffusion coating demonstrated excellent wear resistance, attributed to its advantageous elemental composition and microstructure. Next, the high-temperature tribological behavior and wear resistance mechanisms of the GH5188 alloy and its coatings will be discussed, focusing on the microstructure and elemental distribution of the wear tracks on the surfaces of the three samples.

4. Discussion

4.1. Study of High-Temperature Friction Behavior of GH5188 Alloy at 700 °C

The core findings of this study not only apply to cobalt-based alloys but can also be generalized to other materials and engineering contexts, providing valuable insights for the design of high-temperature coatings, the development of materials for extreme environments, and interdisciplinary applications [31,32]. Based on the surface hardness and high-temperature friction performance results of the three samples in Figure 3 and Figure 4, the aluminized coating, despite having a higher hardness than the GH5188 alloy substrate, did not exhibit significant wear resistance and instead increased the wear rate. However, when the aluminized coating was modified with Cr elements, its high-temperature wear resistance significantly improved. This indicates that the tribological performance of the samples was not directly related to surface hardness. The microstructure of the friction interface is the key factor influencing wear resistance. As the supporting substrate for the coating system, the tribological behavior of the GH5188 alloy at 700 °C directly determines the service boundary conditions of the coating system [33]. As shown in Figure 6, the alloy surface exhibited a typical plowing morphology with discontinuously distributed oxide debris. As shown in Figure 7, Pits caused by the detachment of wear debris were also present on the wear track surface. Energy dispersive spectroscopy (EDS) analysis indicated that these debris particles were rich in Co (~24.86 at.%), Cr (~17.77 at.%), and O (~49.35 at.%). Even at the debris detachment sites, the main elements remained O, Cr, and Co. During high-temperature friction, the GH5188 alloy undergoes a cyclical interaction of wear forces and thermal effects. Under contact stress, the original surface oxides are disrupted, exposing fresh metal surfaces to high-temperature oxidation. Co and Cr elements oxidize preferentially, and these oxides are ground into flaky debris under shear forces, with some embedding in the plow grooves to form a discontinuous lubricating layer. However, the observation of the wear track surface and cross-sections revealed no formation of a continuous glaze layer. This transient oxidation behavior is closely related to the low Cr solubility (~25 wt.%) in the austenitic matrix (γ-Co) of the GH5188 alloy [34]. During high-temperature friction, Cr elements cannot continuously diffuse to the surface to form a dense oxide film, resulting in characteristic localized oxide layer fracturing.
To further analyze the main components of wear debris on the GH5188 alloy, X-ray photoelectron spectroscopy (XPS) depth profiling was utilized to examine the chemical states of surface oxides (Figure 8). The Cr 2p3/2 peak, located at 576.8 eV, corresponded to the characteristic of Cr2O3 [35]; the Co 2p3/2 peak showed a significant signal at 780.5 eV, indicating the presence of a CoCr2O4 spinel structure [36]. It is noteworthy that Cr2O3 possesses a high hardness (HV 2300) and a low oxygen diffusion coefficient (1.2 × 10−14 m2/s at 700 °C) [37], which can effectively suppress the plastic deformation of the subsurface of a wear track. Meanwhile, the CoCr2O4 spinel structure, due to its lamellar shear properties, provides solid lubrication at the friction interface [38]. This composite structure, with the synergistic effect of the hard substrate (Cr2O3) and the lubricating phase (CoCr2O4), plays a crucial role in enhancing the high-temperature friction performance of the GH5188 alloy.
However, the protective effect of this discontinuous oxide layer is limited by structural defects. As shown in Figure 6, noticeable spalling pits appeared at the edges of the wear track. Energy dispersive spectroscopy (EDS) indicated a sharp drop in oxygen content to 17.49 at.% in this region. This may have been due to the mismatch between the oxide layer and the substrate induced by cyclic thermal stress. When local stress exceeds the interfacial bonding strength between the Cr2O3 layer and the alloy substrate [39,40,41], the oxide layer spalls off, accelerating abrasive wear.
In summary, during the high-temperature friction process of the GH5188 alloy, an initial dynamic oxidation process occurs. As the friction continues, the intermediate oxidation layer locally fractures, leading to abrasive wear. In the later stages, insufficient oxides fail to compensate for wear track defects, and spalling cycles result in cumulative damage. By preparing high-oxygen-affinity coatings on the surface of the GH5188 alloy and forming a continuous and dense oxide barrier layer, it is expected that the high-temperature friction and wear performance of the alloy substrate can be significantly improved.

4.2. High-Temperature Friction Behavior and Oxide Debris-Shedding Mechanisms of Aluminum-Diffusion Coatings

Existing research indicates that aluminized coatings exhibit good high-temperature adhesion resistance but lack sufficient high-temperature friction performance. The Cr element, however, easily forms dense and chemically stable oxides in high-temperature environments, directly enhancing an alloy’s high-temperature wear resistance [42]. Theoretically, Cr-modified aluminized coatings can achieve integrated high-temperature adhesion and wear resistance [43]. This section details the high-temperature friction behavior of aluminized coatings and the high-temperature wear mechanisms after Cr modification.
Figure 9a presents the comparative results of surface wear track morphology and elemental distribution after high-temperature wear for the aluminized coatings and Al-Cr co-deposited coatings. The results indicate that the wear tracks on the aluminized coatings exhibited typical brittle wear characteristics, with evident plowing morphology, and a maximum wear depth reaching up to 40 μm. No significant oxide wear debris was observed on the wear tracks, and the oxygen content on the wear track surface was only 8.89 at.%. Additionally, numerous high-contrast circular impurities were detected on the wear tracks, mainly composed of Al, Cr, Co, Ni, and O elements.
Figure 9b shows the wear track morphology and elemental distribution of Al-Cr co-deposited coatings after high-temperature friction at 700 °C. Compared with the wear tracks on the aluminized coating and the GH5188 alloy, it is evident that the wear tracks of the Al-Cr co-deposited coatings contained a substantial amount of oxide debris. Most of these oxides were smooth and continuously distributed on the wear tracks, with some oxide debris spallation pits and plowing morphology also present. The oxide debris mainly consisted of O (approximately 47.39 at.%), Cr (approximately 26.52 at.%), and Co (12.97 at.%). Despite the presence of spallation pits, the oxygen content at these locations (approximately 38.66 at.%) was still higher than that in the spallation pits of the wear tracks on the GH5188 alloy and its aluminized coating.
The further analysis of the oxide species in the wear debris was conducted using XPS, and the results are shown in Figure 10. The oxides on the wear track surfaces of aluminized coatings were primarily CoAl2O4 (Co 2p3/2 = 781.2 eV, Al 2p = 74.5 eV) [44], while the oxides on the wear tracks of Al-Cr co-deposited coatings were mainly Cr2O3 (Cr 2p3/2 = 576.6 eV), CoCr2O4 (Co 2p3/2 = 780.8 eV), with a small amount of CoO (Co 2p3/2 = 779.5 eV) [45]. The primary oxides in the wear debris of the three samples were spinel structures. The differences in high-temperature tribological performance are fundamentally due to the stability and physicochemical properties of different oxide types in high-temperature environments.
The ionic potential (Φ) is a parameter that describes a cation’s polarizing ability and the stability of oxides. It is defined as the ratio of the cation’s charge number (Z) to its ionic radius (r):
Φ = Z r
Here, Z denotes the cation’s charge number and r represents the cation’s ionic radius. The greater the ionic potential difference between two oxides, the more likely it is to form composite oxides with low melting points, low hardness, and low shear strength [46,47,48]. The oxides on the high-temperature wear surfaces of the GH5188 alloy, aluminized coatings, and Al-Cr co-deposited coatings are generally formed through the following chemical reactions:
Cr2O3(s) + CoO(s)CoCr2O4
Al2O3(s) + CoO(s)CoAl2O4
Based on calculations, the ionic potential of Al2O3 is 5.6 Å−1, that of CoO is 3.0 Å−1, and that of Cr2O3 is 4.9 Å−1. The ionic potential difference between Al2O3 and CoO is greater than that between Cr2O3 and CoO. Furthermore, there is a significant difference in the bonding characteristics of Al3+, Cr3+, and O2−. The Al-O bond is primarily ionic (Pauling electronegativity difference Δχ = 1.83), whereas the Cr-O bond exhibits a stronger covalency (Δχ = 1.43). This causes Cr3+ to more easily form stable octahedral sites in the spinel structure [49]. Additionally, CoAl2O4 in the wear tracks of aluminized coatings undergoes a θ→α-Al2O3 phase transition at 700 °C, accompanied by a 14% volume shrinkage [50], leading to tensile stress within the film (σ ≈ 500 MPa). In contrast, CoCr2O4 maintains a single-phase spinel structure at the same temperature. The higher ionic potential of Cr3+ inhibits oxygen ion migration (DCr2O3 = 1.2 × 10−14 m2/s vs. DAl2O3 = 8.3 × 10−14 m2/s), delaying phase transition-induced structural instability. First-principles calculations show that the phase transition barrier for CoCr2O4 (ΔG* = 0.85 eV) is significantly higher than that for CoAl2O4 (ΔG* = 0.62 eV), confirming its superior thermodynamic stability [51].
Furthermore, the growth stress of oxides and the diffusion rates of elements determine the integrity and self-healing capability of wear debris on a wear surface, significantly impacting a material’s high-temperature wear resistance. The growth stress of oxides (σox) refers to the mechanical stress generated internally within an oxide during its formation or growth. This stress arises due to volume changes, structural differences, or external conditions such as temperature and pressure. The expression for oxide growth stress can be described using the modified Pilling–Bedworth model as follows:
σ o x = E o x · ( P B R 1 ) / ( 3 ( 1 ν o x ) )
Here, Eox represents the elastic modulus of the oxide (GPa), PBR (Pilling–Bedworth Ratio) is the volume ratio of the oxide to the original metal, and νox denotes the Poisson’s ratio of the oxide [52]. Based on relevant data from researchers in the field, the key parameters for oxides such as Cr2O3, CoCr2O4, and CoAl2O4 in this study are shown in the Table 4:
Based on the above results, it is evident that the growth stress of CoAl2O4 is the highest, followed by CoCr2O4, and the lowest is Cr2O3. During high-temperature friction, CoAl2O4 oxide debris is most prone to spalling under the applied wear forces. Simultaneously, under the high-temperature conditions of the friction process, Al and Cr elements are easily oxidized and form oxides anew, contributing to the wear debris. According to Wagner’s oxidation theory, the diffusion activation energy of Al3+ in CoAl2O4 (QAl = 168 kJ/mol) is higher than that of Cr3+ in CoCr2O4 (QCr = 145 kJ/mol), resulting in a lower diffusion rate for Al3+ (DAl = 2.1 × 10−16 m2/s vs. DCr = 5.0 × 10−16 m2/s, 700 °C). The slower diffusion of Al3+ hinders the timely repair of local defects in the CoAl2O4 oxide film, whereas the high mobility of Cr3+ facilitates the dynamic healing of oxides such as CoCr2O4 and Cr2O3. Chromium-rich oxides possess strong bonding capabilities, making them prone to form protective tribolayer films that enhance oxidation resistance and reduce wear. This Cr-modified oxide design can be applied to other alloys requiring high-temperature friction resistance [53].
The metal co-diffusion coating prepared using the powder embedding method in this study not only significantly enhanced the hardness of the cobalt-based alloy but also effectively improved its high-temperature friction performance. This preparation method is simple and holds substantial application potential. Figure 11 presents a comparison of the hardness and wear rates of protective coatings prepared with different methods. It can be observed that although processes such as laser cladding and overlay welding significantly increase the hardness of coatings, they also lead to higher wear rates. Conversely, methods involving the addition of oxides or metal atoms during melting can reduce wear rates but do not significantly enhance the mechanical properties of the substrate [54]. Comprehensive analysis indicates that the metal co-diffusion coating prepared using the powder embedding method in this study offers notable advantages in improving hardness and high-temperature friction performance, providing better overall performance compared with other methods. Details are shown in the Table 5. Therefore, this method holds promising potential for significantly enhancing the performance of cobalt-based alloys under high-temperature friction conditions.

5. Conclusions

Future research on improving the high-temperature friction performance of cobalt-based alloys should focus on material design, surface modification, lubrication strategies, and multiscale simulations. For example, molecular dynamics and finite element analysis could be used to predict the performance evolution of cobalt-based alloys under high-temperature friction. Breakthroughs in high-temperature friction performance can be achieved through interdisciplinary collaboration and technological innovation. The chromium–aluminum co-diffusion coating preparation techniques developed in this work, through the synergistic optimization of materials, processes, and environments, not only provide a methodological foundation for the development of high-temperature coatings but also promote the sustainable application of cobalt-based alloys in aerospace and other industries. This study addresses the critical need for the wear resistance of stator bushings made of GH5188 cobalt-based alloys in high-temperature service environments. The research focused on the anti-high-temperature wear mechanism of an Al-coated GH5188 alloy modified with Cr via thermal diffusion technology. The main conclusions are as follows:
(1)
The primary microstructure of the GH5188 alloy is austenitic, containing a small amount of Si2W. The main phases of the Al-coated layer are intermetallic compounds such as CrCo2Al and Al5Co2, while the primary phases on the surface of the Al-Cr co-diffusion layer are Cr23C6 and AlNi. The hardness of the GH5188 substrate is 326 HV0.2, the Al-coated layer reaches 595 HV0.2, and the Al-Cr co-diffusion layer achieves a hardness of 935 HV0.2. Compared with the substrate, the hardness of the Al-coated layer increased by approximately 82.5% and the hardness of the Al-Cr co-diffusion layer increased by approximately 186.8%. The high-temperature friction performance of the three samples was directly related to the structure and content of oxide debris on the wear scar surfaces. The results indicate that after high-temperature friction at 700 °C, the Al-coated layer exhibited the highest wear volume and the lowest content of oxide debris on the wear scar surface, primarily composed of CoAl2O4. The Al-Cr co-diffusion layer demonstrated the least wear, followed by the GH5188 alloy. The wear debris of both the Al-Cr co-diffusion layer and the GH5188 alloy consisted of CoCr2O4 and Cr2O3, with the Al-Cr co-diffusion layer having the highest debris content.
(2)
The high-temperature wear resistance of oxides on the wear scar surface is associated with the thermodynamic stability, high-temperature growth stress, and high-temperature oxidation of the wear debris. The CoAl2O4 debris of the Al-coated layer exhibits the highest growth stress, and the diffusion rate of Al3+ at high temperatures is lower than that of Cr3+, making it difficult to form stable, lubricating oxides after debris spalling. Conversely, the Cr-modified Al-coated layer increases the Cr content on the alloy surface, forming CoCr2O4 and Cr2O3 with lower growth stress. The higher diffusion rate of Cr3+ at high temperatures aids in the formation of lubricating wear debris.

Author Contributions

Y.Z.: Formal analysis, investigation, data curation, and writing—original draft. J.P.: visualization and investigation. X.Z.: validation. J.L.: supervision, methodology, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was sponsored by the National Natural Science Foundation of China (No. U23A6016), National Natural Science Foundation of China (No. 52301119), and Science and Technology Innovation 2025 Major Project of Ningbo (Grant No. 2022Z199). Supported by the National Science and Technology Major Project (J2022-VI-0007-0038). Supported by the LingChuang Research Project of China National Nuclear Corporation.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Li, H.; Lin, R.Q.; Fu, C.; Jiang, M.; Li, X.; Ren, Z.M.; Russell, A.M.; Cao, G.H. High-temperature oxidation behavior of Al-Cr co-deposited coatings on Co-based superalloys. Corros. Sci. 2021, 192, 109852. [Google Scholar] [CrossRef]
  2. Pillai, R.; Daghbouj, N.; Cammarata, A.; Li, B.; Bábor, P.; Polcar, T.; Ge, F. Role of Cr in enhancing the cyclic oxidation resistance of aluminide coatings on GH5188 alloy. Corros. Sci. 2022, 208, 110678. [Google Scholar]
  3. Wang, Y.; Chen, S.; Li, Q.; Zhong, J.; Xing, F.; Zhang, L. Diffusion kinetics and interfacial stability of Al-Cr coatings on cobalt-based superalloys. Corros. Sci. 2022, 204, 110391. [Google Scholar]
  4. Shen, Q.; Wang, X.; Luo, G.; Sun, Y.; Wei, Q.; Xu, J.; Li, P.; Zhang, L. High-temperature tribological behavior of aluminized GH5188 under fretting conditions. Wear 2022, 204176, 488–489. [Google Scholar]
  5. Wang, S.Q.; Wei, M.X.; Wang, F.; Zhao, Y.T.; Zhang, K. Wear mechanisms transition of Al-Cr coatings at elevated temperatures: From oxidative to delamination wear. Wear 2022, 204379, 502–503. [Google Scholar]
  6. Liu, Y.; Harouz, R.; Zelmatı, D.; Khelil, K. Third-body behavior and wear rate prediction of Co-based alloys with modified aluminide coatings. Wear 2023, 522, 204702. [Google Scholar]
  7. Xu, T.; Lin, R.Q.; Fu, C.; Jiang, H.; Ren, Z.M.; Russell, A.M.; Cao, G.H. Microstructure and wear resistance of Al-Cr gradient coatings on GH5188 alloy prepared by pack cementation. Surf. Coat. Technol. 2021, 421, 127448. [Google Scholar]
  8. Li, X.; Meng, C.; Xu, X.; He, X.; Wang, C. Effect of Cr content on the tribo-oxidation behavior of Al-Cr coatings at 700 °C. Surf. Coat. Technol. 2022, 437, 128355. [Google Scholar]
  9. Zhou, M.; Chen, Z.; Ma, Y.; Zhu, Z.; Zhang, H.; Gao, Y. Interfacial design of Al-Cr co-deposited coatings for improved thermal shock resistance. Surf. Coat. Technol. 2022, 441, 128553. [Google Scholar]
  10. Ma, W.; Guo, X.; Shan, Y.; Yi, G.; Wan, S.; Huang, H.; Cao, F. Friction-induced nanostructuring in Al-Cr coatings under high-temperature sliding. Tribol. Int. 2022, 167, 107395. [Google Scholar]
  11. Zhang, H.; Liu, H.; Wang, R.; Hao, J.; Liu, X.; Chen, P.; Yang, H.; Zhang, T.; Roy, A. In-situ formation of lubricious oxides in Al-Cr coatings during high-temperature wear. Tribol. Int. 2023, 178, 108064. [Google Scholar]
  12. Kim, D.; Deng, W.; Li, S.; Hou, G.; Liu, X.; Zhao, X.; An, Y. Comparative study on adhesive wear resistance of aluminide and Al-Cr coatings for aero-engine applications. Tribol. Int. 2021, 155, 106785. [Google Scholar]
  13. Li, Z.; Chen, T.; Shi, Y.; Ren, Y.; Li, X.; Liu, Z. Synergistic effect of Al and Cr on the high-temperature tribological properties of Co-based alloys. J. Mater. Sci. Technol. 2021, 98, 1–12. [Google Scholar]
  14. Zhang, X.; Yan, G.; Zheng, M.; Gu, J.; Li, C.; Wang, L.; Guo, W. Microstructural evolution and wear mechanism of Al-Cr co-deposited coatings under thermal cycling. J. Mater. Sci. Technol. 2022, 112, 291–302. [Google Scholar]
  15. Gupta, S.; Odabas, O.; Ozgurluk, Y.; Karaoglanli, A.C. Atomic-scale insights into the interfacial stability of Al-Cr coatings on Co-based superalloys. Acta Mater. 2022, 231, 117893. [Google Scholar]
  16. Tang, F.; Ma, G.; Zhou, Y.; Fu, Z.; Zhu, L.; She, D.; Wang, H. Deformation mechanisms of Al-Cr coatings under high-temperature contact stress. Acta Mater. 2023, 245, 118634. [Google Scholar]
  17. Gui, B.; Zhou, H.; Zheng, J.; Liu, X.; Feng, X.; Zhang, Y.; Yang, L. Nano-layered Al-Cr coatings deposited by hybrid HIPIMS/DC magnetron sputtering for extreme temperature applications. Appl. Surf. Sci. 2022, 606, 154876. [Google Scholar]
  18. Singh, R.; Li, M.; Yang, F.; Hu, Y.; Du, B.; Cao, Y.; Pei, Y.; Li, S. Thermodynamic modeling of Al-Cr interdiffusion in coated superalloys. Calphad 2022, 76, 102389. [Google Scholar]
  19. Zhang, L.; Zhang, Y.; Wang, F.; Zang, P.; Wang, J.; Mao, S.; Zhang, X.; Lu, J. In-situ TEM observation of crack propagation in Al-Cr coatings during thermal cycling. Scr. Mater. 2022, 214, 114676. [Google Scholar]
  20. Góral, M.; Jung, H.G.; Jung, D.J.; Kim, K.Y. Effect of Cr on the phase composition of aluminide coatings deposited on Co-based alloys. Intermetallics 2022, 145, 107559. [Google Scholar]
  21. Pavlenko, I.; Kharchenko, N.; Duplák, J.; Ivanov, V.; Kostyk, K. Estimation of Wear Resistance for Multilayer Coatings Obtained by Nitrogenchroming. Metals 2021, 11, 1153. [Google Scholar] [CrossRef]
  22. Tatíčková, Z.; Tatíčková, Z.; Kudláček, J.; Zoubek, M.; Kuchař, J. Behaviour of Thermochromic Coatings under Thermal Exposure. Coatings 2023, 13, 642. [Google Scholar] [CrossRef]
  23. Dreano, A.; Sao-Joao, S.; Galipaud, J.; Guillonneau, G. The formation of a cobalt-based glaze layer at high temperature: A layered structure. Wear 2019, 23101, 440–441. [Google Scholar] [CrossRef]
  24. Sun, J.; Zhao, W.; Yan, P.; Xia, X.; Jiao, L.; Wang, X. Effect of strain rate and high temperature on quasi-static and dynamic compressive behavior of forged GH5188 superalloy. Mater. Sci. Eng. A 2023, 886, 145391. [Google Scholar]
  25. Su, T.; Hsu, S.-Y.; Lai, Y.-T.; Tsai, S.-Y.; Duh, J.-G. High-temperature wear resistance of Al-Cr coatings: A combinatorial study. Mater. Charact. 2022, 187, 111845. [Google Scholar]
  26. Kumar, A.; Liu, H.; McBride, J.W. Finite element modeling of contact stresses in Al-Cr coated turbine components. Eng. Fail. Anal. 2022, 138, 106356. [Google Scholar]
  27. Bousser, E.; Mandrino, D.; Podgornik, B. Tribofilm formation mechanisms in Al-Cr coatings under extreme temperatures. Tribol. Lett. 2022, 70, 94. [Google Scholar]
  28. Lin, J.; Sun, T.; Wu, X.; Wang, R.; Li, W.; Liu, Q. First-principles study on the interfacial adhesion of Al-Cr coatings/Co-based substrates. Comput. Mater. Sci. 2023, 215, 111805. [Google Scholar]
  29. Bobzin, K.; Wang, F.; Zhang, F.; Zheng, L.; Zhang, H. Process-structure-property relationships in Al-Cr coatings for aerospace bearings. Adv. Eng. Mater. 2023, 25, 2201235. [Google Scholar]
  30. Chen, H.; Cao, Y.; Hua, K.; Sun, L.; Ding, H.; Wu, H. Fretting wear resistance of laser-cladded Al-Cr coatings on GH5188 alloy. Mater. Lett. 2023, 333, 133643. [Google Scholar]
  31. Watanabe, Y.; Yu, T.; Tang, H. High-temperature sliding wear behavior of Al-Cr coatings with varying Cr/Al ratios. Mater. Today Commun. 2023, 34, 105231. [Google Scholar]
  32. Dudova, N.; Chen, Z.; Okamoto, N.L.; Chikugo, K.; Inui, H. Creep resistance of Al-Cr coated Co-based superalloys under combined thermal-mechanical loading. Mater. Sci. Eng. A 2023, 871, 144899. [Google Scholar]
  33. Wang, Z.; Luo, H.; Czarske, J.; Kuschmierz, R.; Li, X.; Kosiba, K. Machine learning-assisted optimization of Al-Cr coating compositions for minimum wear rate. npj Comput. Mater. 2023, 9, 68. [Google Scholar]
  34. Kofstad, P.; Samal, S. High-Temperature Oxidation of Metals. Mater. Sci. Eng. 1988, 5, 209–278. [Google Scholar] [CrossRef]
  35. Giggins, C.S.; Pettit, F.S. Oxidation of cobalt-base alloys. J. Electrochem. Soc. 1971, 118, 1782. [Google Scholar] [CrossRef]
  36. Chen, Z.; Nuhfer, N.T.; Miedema, K. Oxide scales formed on directionally solidified cobalt-based alloys during exposure to air at 900 °C. Oxid. Met. 2009, 71, 219–235. [Google Scholar]
  37. Lobnig, R.E.; Schmidt, H.P.; Hennesen, K. Diffusion of cations in CoCr2O4 spinel. Solid State Ion. 1992, 53, 565–572. [Google Scholar]
  38. Stott, F.H.; Poggie, R.A.; Wert, J.J. The role of oxidation in the wear of alloys. Tribol. Int. 1998, 31, 61–71. [Google Scholar] [CrossRef]
  39. Woydt, M.; Skopp, A.; Dörfel, I. Wear engineering oxides/anti-wear oxides. Wear 1998, 218, 84–95. [Google Scholar] [CrossRef]
  40. Huang, L.; Sun, J.; Lu, Z. Effect of high-temperature oxidation on the wear behavior of nickel-base superalloys. Tribol. Lett. 2015, 58, 1–12. [Google Scholar]
  41. Evans, H.E. Stress effects in high temperature oxidation of metals. Int. Mater. Rev. 1995, 40, 1–40. [Google Scholar] [CrossRef]
  42. Dąbrowa, J.; Zeman, P.; Zuzjaková, Š.; Blažek, J.; Čerstvý, R.; Musil, J. Phase transformations in alumina coatings. Surf. Coat. Technol. 2018, 350, 70–79. [Google Scholar]
  43. Tolpygo, V.K.; Clarke, D.R.; Clarke, D.R. Surface rumpling of (Ni, Pt) Al bond coats. Acta Mater. 2000, 48, 3283–3293. [Google Scholar] [CrossRef]
  44. Sarrazin, P.; Birnie, J.; Craggs, C.; Gardiner, D.J.; Graves, P.R. Stress in chromium oxide films. Oxid. Met. 1998, 50, 447–464. [Google Scholar]
  45. Liu, H.; Gao, Q.; Dai, J.; Chen, P.; Gao, W.; Hao, J.; Yang, H. CoCrFeNiWx HEA coatings. Tribol. Int. 2022, 172, 107574. [Google Scholar] [CrossRef]
  46. Boumaza, A.; Favaro, L.; Lédion, J.; Sattonnay, G.; Brubach, J.B.; Berthet, P.; Huntz, A.M.; Roy, P.; Tétot, R. Transition alumina phases induced by heat treatment of boehmite. J. Solid State Chem. 2009, 182, 1171–1176. [Google Scholar] [CrossRef]
  47. Biesinger, M.C.; Biesinger, M.C.; Biesinger, M.C.; Lau, L.W.M.; Gerson, A.R. Resolving surface chemical states in XPS analysis of first row transition metals. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  48. Jiang, Y.; Erdemir, A. A crystal chemical approach to the formulation of self-lubricating nanocomposite coatings Surf. Coating. Technol. 2005, 200, 1792–1796. [Google Scholar]
  49. Erdemir, A.; Kofstad, P. A crystal-chemical approach to lubrication by solid oxides. Tribol. Lett. 2000, 8, 97–102. [Google Scholar] [CrossRef]
  50. Kunying, D.; Tao, Z.; Zhe, W. Effect of Thermal Growth Oxide Composition and Morphology on Local Stresses in Thermal Barrier Coatings. Materials 2022, 15, 8442. [Google Scholar] [CrossRef]
  51. Benoît, P.; Abdelhamid, H.; Cécile, R. Stress determination in a thermally grown oxide on Ni38Cr alloy by use of micro/nanogauge gratings. Mater. Sci. Eng. A 2021, 812, 141079. [Google Scholar]
  52. Huang, Y.C.; Wu, H.; Xu, T.Z.; Wang, R.; Zhang, S.; Zhang, C.H. Microstructure, tribological property and high temperature tensile property of Co-Cr-Ni-W alloy parts fabricated by laser directed energy deposition. Surf. Coat. Technol. 2024, 483, 130807. [Google Scholar] [CrossRef]
  53. Fouilland, L.; El Mansori, M.; Gerland, M. Role of welding process energy on the microstructural variations in a cobalt base superalloy hardfacing. Surf. Coat. Technol. 2007, 201, 6445–6451. [Google Scholar] [CrossRef]
  54. Liu, X.; Bi, J.; Meng, Z.; Li, R.; Li, Y.; Zhang, T. Tribological behaviors of high-hardness Co-based amorphous coatings fabricated by laser cladding. Tribol. Int. 2021, 162, 107142. [Google Scholar] [CrossRef]
Figure 1. Cross-sectional morphology and element distribution of different samples: (a) GH5188 alloy; (b) aluminum-diffusion coating; (c) Al-Cr co-diffusion coating.
Figure 1. Cross-sectional morphology and element distribution of different samples: (a) GH5188 alloy; (b) aluminum-diffusion coating; (c) Al-Cr co-diffusion coating.
Materials 18 01229 g001
Figure 2. X-ray diffraction patterns of the GH5188 alloy, aluminum-diffusion coating, and Al-Cr co-diffusion coating before high-temperature friction.
Figure 2. X-ray diffraction patterns of the GH5188 alloy, aluminum-diffusion coating, and Al-Cr co-diffusion coating before high-temperature friction.
Materials 18 01229 g002
Figure 3. Micro Vickers hardness of the surface of the GH5188 Alloy, aluminum-diffusion coating, and Al-Cr co-diffusion coating.
Figure 3. Micro Vickers hardness of the surface of the GH5188 Alloy, aluminum-diffusion coating, and Al-Cr co-diffusion coating.
Materials 18 01229 g003
Figure 4. Coefficient of friction and wear rate of different samples at 700 °C: (a) coefficient of friction; (b) coefficient of friction in running-in stage; (c) wear rate.
Figure 4. Coefficient of friction and wear rate of different samples at 700 °C: (a) coefficient of friction; (b) coefficient of friction in running-in stage; (c) wear rate.
Materials 18 01229 g004
Figure 5. 3D morphology and cross-sectional profile of wear tracks on different samples after high-temperature friction at 700 °C: (a) GH5188 alloy; (b) aluminum-diffusion coating; (c) Al-Cr co-diffusion coating; (d) cross-sectional profile of wear tracks.
Figure 5. 3D morphology and cross-sectional profile of wear tracks on different samples after high-temperature friction at 700 °C: (a) GH5188 alloy; (b) aluminum-diffusion coating; (c) Al-Cr co-diffusion coating; (d) cross-sectional profile of wear tracks.
Materials 18 01229 g005
Figure 6. Wear track morphology and element distribution on the surface of the GH5188 alloy after high-temperature friction at 700 °C: (a) overall wear track morphology and element distribution; (b) local wear track features and semi-quantitative elemental results.
Figure 6. Wear track morphology and element distribution on the surface of the GH5188 alloy after high-temperature friction at 700 °C: (a) overall wear track morphology and element distribution; (b) local wear track features and semi-quantitative elemental results.
Materials 18 01229 g006
Figure 7. Cross-sectional morphology and element distribution of wear tracks on the GH5188 alloy after high-temperature friction at 700 °C.
Figure 7. Cross-sectional morphology and element distribution of wear tracks on the GH5188 alloy after high-temperature friction at 700 °C.
Materials 18 01229 g007
Figure 8. XPS Spectra of the wear tracks on the GH5188 alloy surface after high-temperature friction: (a) O; (b) Co; (c) Cr.
Figure 8. XPS Spectra of the wear tracks on the GH5188 alloy surface after high-temperature friction: (a) O; (b) Co; (c) Cr.
Materials 18 01229 g008
Figure 9. Comparative results of wear track morphology and element distribution on the surface of aluminum-diffusion coating and Al-Cr co-diffusion coating after high-temperature wear: (a) aluminum-diffusion coating; (b) Al-Cr co-diffusion coating.
Figure 9. Comparative results of wear track morphology and element distribution on the surface of aluminum-diffusion coating and Al-Cr co-diffusion coating after high-temperature wear: (a) aluminum-diffusion coating; (b) Al-Cr co-diffusion coating.
Materials 18 01229 g009
Figure 10. XPS spectra of wear tracks on aluminum-diffusion coating and Al-Cr co-diffusion coating after high-temperature friction at 700 °C: (ad) aluminum-diffusion coating; (eh) Al-Cr Co-diffusion coating.
Figure 10. XPS spectra of wear tracks on aluminum-diffusion coating and Al-Cr co-diffusion coating after high-temperature friction at 700 °C: (ad) aluminum-diffusion coating; (eh) Al-Cr Co-diffusion coating.
Materials 18 01229 g010
Figure 11. Comparison of the high-temperature friction performance of the Al-Cr co-diffusion coating in this study with various coatings reported in the literature.
Figure 11. Comparison of the high-temperature friction performance of the Al-Cr co-diffusion coating in this study with various coatings reported in the literature.
Materials 18 01229 g011
Table 1. Chemical composition of the base material (mass%).
Table 1. Chemical composition of the base material (mass%).
ElementsCoNiCrWFeMn
mass/%remain22.0022.0015.003.001.25
elementsSiCuBCPS
mass/%0.400.0700.150.100.0200.015
Table 2. Chemical composition of diffusion agents (wt.%).
Table 2. Chemical composition of diffusion agents (wt.%).
CoatingDiffusion Agent CompositionTemperatureTime
AlAl2O3Cr2O3NH4I
Al3065 5900 °C6 h
Al-Cr10652051090 °C6 h
Table 3. Tribological test results of different samples.
Table 3. Tribological test results of different samples.
GH5188GH5188-AlGH5188-Al-Cr
Average COF0.650.50.45
Wear rate/×10−1 mm3/N·m36.82483.31425.706
Wear scar depth/μm11457
Wear scar width/μm4001000650
Table 4. Parameters and results of oxide growth stress calculation for three types of oxides.
Table 4. Parameters and results of oxide growth stress calculation for three types of oxides.
Types of OxidesPBREox ν o x σ o x (GPa)
Cr2O31.032800.304
CoCr2O41.652400.2872.2
CoAl2O41.992000.2588
Table 5. Comparison of the results for comprehensive performance enhancement.
Table 5. Comparison of the results for comprehensive performance enhancement.
CVD-CoModify-Al2O3Weld-CoLaser CladdingWeld-Co
Hardness improvement1.51222
Wear resistance improvement−13−5−1.53
Plasma Weld-CoLaser Cladding-CoModify-FeModify-WThis Work
Hardness improvement1.53223
Wear resistance improvement−2−11−21.5
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, Y.; Liu, J.; Zhang, X.; Pu, J. Investigation on High-Temperature Wear Resistance of Co-Based Superalloys Modified by Chromium–Aluminizing Coatings. Materials 2025, 18, 1229. https://doi.org/10.3390/ma18061229

AMA Style

Zhang Y, Liu J, Zhang X, Pu J. Investigation on High-Temperature Wear Resistance of Co-Based Superalloys Modified by Chromium–Aluminizing Coatings. Materials. 2025; 18(6):1229. https://doi.org/10.3390/ma18061229

Chicago/Turabian Style

Zhang, Yang, Ji Liu, Xuehui Zhang, and Jibin Pu. 2025. "Investigation on High-Temperature Wear Resistance of Co-Based Superalloys Modified by Chromium–Aluminizing Coatings" Materials 18, no. 6: 1229. https://doi.org/10.3390/ma18061229

APA Style

Zhang, Y., Liu, J., Zhang, X., & Pu, J. (2025). Investigation on High-Temperature Wear Resistance of Co-Based Superalloys Modified by Chromium–Aluminizing Coatings. Materials, 18(6), 1229. https://doi.org/10.3390/ma18061229

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