Next Article in Journal
Study on the Modification of Silty Soil Sites Using Nanosilica and Methylsilicate
Previous Article in Journal
Quantifying Decay Due to Wet Atmospheric Deposition on Basalt
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Mo2C Addition on the Tribological Behavior of Ti(C,N)-Based Cermets

National Engineering Research Centre of Near-Net-Shape Forming Technology for Metallic Materials, South China University of Technology, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(16), 5645; https://doi.org/10.3390/ma16165645
Submission received: 17 July 2023 / Revised: 8 August 2023 / Accepted: 11 August 2023 / Published: 16 August 2023

Abstract

:
Due to the excellent properties of Ti (C,N)-based ceramics, such as high hardness, excellent wear resistance, exceptional thermal deformation resistance, and sound chemical stability, they have been widely used in cutting tools or molds. Thus, revealing their tribological behavior against hard materials is of great significance. Some studies have reported the tribological behavior of Ti(C,N)-based cermets and hard cermets, but so far, the effects of Mo2C additions on the frictional properties of Ti(C,N)-based cermets are still unclear. In this study, Ti(C,N)-10WC-1Cr3C2-5Co-10Ni-x Mo2C cermets (x = 4, 6, 8, 10 and 12 wt.%) were sintered using a vacuum hot-pressing furnace. Furthermore, the core–rim morphologies of the sintered samples were observed in SEM images. Then, the wear resistance of the cermets was studied against a Si3N4 ball at a 50 N load using the fretting wear test. Finally, the wear mechanism was characterized using a combination of SEM, EDS and XPS. The experimental results indicated that the wear mechanisms of the cermets were mainly abrasive wear, adhesive wear, and the formation of an oxide film. As the content of Mo2C increased from 4 wt.% to 12 wt.%, the friction coefficient and wear volume had a variation law of first decreasing and then decreasing, and reached minimum values at 6 wt.% and 12 wt.%, and the lowest friction coefficient and wear rate were 0.49 and 0.9 × 10−6 mm3/Nm, respectively. The 6 wt.% Mo2C greatly improved the hardness and fracture toughness of the cermet, while the 12 wt.% Mo2C promoted the formation of an oxide film and protected the friction surface. The cermet with 6 wt.% Mo2C is recommended because it has comprehensive advantages in terms of its mechanical properties, tribological properties, and cost.

1. Introduction

Ti(C,N)-based cermets exhibit significant potential for diverse applications, including cutting tools, sea ling rings, and wear parts. This is owing to their remarkable attributes, such as high hardness, exceptional wear resistance, extraordinary resistance to thermal deformation, and sound chemical stability [1,2,3,4,5]. Ti(C,N)-based cermets have demonstrated superior oxidation resistance and high temperature toughness when compared to traditional WC–Co cemented carbide. As a result, they have become an increasingly competitive rival [6,7]. Consequently, Ti(C,N)-based cermets have found widespread and successful implementation within the metal cutting industry, especially in high-speed cutting applications [8,9].
Ti(C,N)-based cermets comprise two phases: the cermet phase, such as Ti(C,N), which has high hardness, and a metal binder phase, such as Co and Ni, which is used as a binder phase to bond with the cermet phase to make Ti(C,N)-based composites with a high relative density, high hardness, and high toughness [10]. However, the poor wettability between the binder phase and the Ti(C,N) hard phase, as well as the low fracture toughness of the cermets, hinder the application of Ti(C,N) cermets.
So far, many technologies have been developed to fabricate Ti(C,N)-based cermets, such as hot-press sintering (HPS) [11], microwave sintering [12], and hot isostatic pressing sintering (HIPS) [13]. Microwave sintering can improve the density of the cermets, but causes a reduction in the hardness to a certain extent. HIPS does not need a mold, but the preparation efficiency is low. Compared with microwave sintering and HIPS, HPS can economically obtain highly dense materials, a uniform fine-grained microstructure, and excellent physical and mechanical properties with high reliability [14]. Thus, HPS has been widely used for sintering and developing various materials, including composite materials such as hard metals and cermets [15,16].
In the Ti(C,N)-based cermets, Mo2C is generally used as a sintering aid to enhance the wettability between the Ti(C,N) grain and the binder phase, and to improve the fracture toughness of cermets [17]. Earlier studies have indicated that the addition of Mo2C to Ti(C,N)-based cermets is beneficial to the wettability between the Ti(C,N) cermet phase and the (Ni, Co) metallic phase, and a finer microstructure and better mechanical properties have been obtained [2,18,19,20,21].
TiCN-based cermets have been thoroughly studied for their wear behavior under a wide range of dry ambient testing conditions. Extensive investigations have been carried out to understand the friction and wear characteristics of TiCN-based cermets when interacting with steel. Previous researchers have reported a considerable variation in wear rates, varying from 10−7 mm3/Nm to 10−6 mm3/Nm. The complicated wear behavior of TiCN-based cermets is defined by adhesion, abrasion, tribo-chemical wear, oxidation, plastic deformation, and fracture according to the experimental conditions of wear testing [22,23,24,25].
There are significant differences in the tribological performance of different friction pairs; however, there is little information on hard cermet counter bodies, and the effects of Mo2C additions on the frictional properties of Ti(C,N)-based cermets against hard cermet counter bodies are still unclear. The relationship between the addition of Mo2C, the “core–rim” structure and the strengthening and toughening mechanism remains unclear among the studies. As the application of TiCN cermet is extending to hard metals and ceramics, in order to evaluate their full potential, a thorough grasp of tribological wear behavior against hard counter bodies is very necessary. The aim of this study is to clarify the effects of Mo2C on the microstructure and tribological properties of Ti(C,N)-based cermets.

2. Materials and Methods

2.1. Preparation of Cermets

The particle sizes and purity of the initial powders are listed in Table 1. The powders were precisely weighed and mixed based on the specified composition provided in Table 2. The content range of Mo2C addition was determined based on the previous studies of our team and other researchers [18,26,27]. Subsequently, they were wet-milled in an ethanol bath using YG8 cemented carbide balls in a planetary ball mill (QM-3SP4, NanDa Instrument Plant, Nanjing, China) for 24 h at a rotational speed of 200 r/min. The powder/ball/ethanol mass ratio was maintained at 1:5:0.8. Finally, the mixture was dried at 70 °C in an electrically heated drying oven.
Before the sintering, the mixed powder was filled into a cylindrical graphite mold that was 30 mm in diameter and 60 mm in height. A layer of graphite paper was placed between the graphite mold and the powder. Under vacuum conditions (10−3 Pa), the powders were first pre-pressed at room temperature and then hot-pressing sintered at 1500 °C and 20 MPa for 1 h, following furnace cooling.
Figure 1 shows the sintering curve of the Ti(C,N)-based cermets. In order to achieve a more uniform furnace temperature and facilitate the release of gases (CO2, CO, N2, etc.) before the liquid-phase sintering stage (in which the binder phases Ni and Co melted), specific dwell times of 30 min at 800 °C and 60 min at 1220 °C were implemented. These conditions were chosen to promote the densification of the sample [28,29].

2.2. Characterization

After sintering, the specimens were cut into 8 × 8 × 5 mm3 pieces using an EDM wire-cutting machine. Subsequently, the pieces were ground and polished for metallographic analysis. The surface morphology of the sample was examined using a metallographic microscope (Leica DMi8 C) and a high-resolution scanning electron microscope (FEI Nova Nano SEM 430). The densities of the samples were determined using the Archimedes immersion method.
The Vickers hardness was measured using a THVP-30 hardness tester (Beijing Era United Technology Co., Ltd., Beijing, China) with a 10 kgf load, and the obtained result was determined according to the average of seven measurements. To determine the fracture toughness (KIC) of the specimens, the crack lengths departing from the Vickers indentation corners were measured, and the expression derived by Shetty et al. was employed [30]:
K I C = A H V · P i = 1 4 l i
where HV is the hardness (N·m−1); P is the load (N); l is the sum of crack lengths (mm); and A is a constant, which is determined by the following equation: A = 1/[3(1 − ν2)(21/2π5/2 tan θ)1/3], where ν is Poisson’s ratio of the material and θ is the angle of the opposite faces of Vickers indenter (2θ = 136° in our test). For ν = 0.22 (typical value of cermet), A is calculated as 0.0889 [31]. In our study, the hardness of specimens was tested under a load of 98 N. Thus, the Equation can be simplified as follows:
K I C = 0.88 H V 10 i = 1 4 l i
where KIC is the fracture toughness (MPa·m1/2); HV10 is the Vickers hardness (GPa); and li is the length of four cracks at the corner of the indentation (mm).

2.3. Fretting Wear Test

The fretting wear test was conducted utilizing an oscillating friction and wear tester (SRV IV, Optimol, Munich, Germany), with a ball-on-block contact configuration. The detailed parameters of the fretting wear tests are shown in Table 3. The specimens were cut into blocks with dimensions of 8 × 8 × 8 mm for the test, and their testing surfaces were polished. Si3N4 balls with a diameter of 10 mm, a hardness of 1700 HV, and a surface roughness of 0.025 µm were chosen as the counterbody due to their high hardness and chemical inertness under wear conditions. The experiment was conducted at room temperature (approximately 25 °C) with an air humidity of 40–45%, and the coefficient of friction (CoF) was continually recorded during the test. A load of 5 N was preloaded for 5 min before the test for running-in. In order to ensure that enough friction products could be observed, previous work was studied [32,33,34], and a high load of 50 N was set. The mean CoF and average wear volume of each material state were determined using the results of at least three conducted tests.
Both before and after the fretting wear test, the specimen was cleaned in ethanol and then dried. The profiles of the wear scars were measured using a universal 3D profilemeter (UP Dual Model, Rtec Instruments, San Jose, CA, USA), and the wear volume of the specimens was calculated using the Gwyddion software (Version 2.61, Czech Metrology Institute, Brno, Czechia).
In addition, the wear surfaces of the samples were characterized using X-ray photoelectron spectroscopy (XPS, AXIS SUPRA+, Shimadzu, Kyoto, Japan). The C1s peak was used as a fixed reference point at 284.8 eV to calibrate all binding energies. Thermo Avantage software was employed for the analysis of the obtained spectra (version 5.9922).

3. Results and Discussion

3.1. Phase Analysis of TiCN-Based Cermets

The XRD analysis was carried out to confirm the crystallographic structures of the TiCN-based cermets, as shown in Figure 2. The hard phase (Ti,M)(C,N) can be observed in all specimens, where M represents W or/and Mo in the molecular formula (Ti,M)(C,N). The binder-phase Ni–Co–M peaks, which are slightly adrift from the pattern of Ni and Co, indicate the existence of Ti, Mo or W dissolved in the binder phase. No diffraction peaks corresponding to WC or Mo2C were detected in any of the specimens. This suggests that most and even all of the WC and Mo2C underwent dissolution and formed hard phases of (Ti,W,Mo)(C,N) solid solution.
The lattice constants calculated based on the XRD patterns of various phases are shown in Table 4. When the Mo2C content exceeds 8 wt.%, the lattice constants of the white core and grey rim exhibit minimal variation. Consequently, the content of W and Mo remains stable in the white core–grey rim, aligning with the findings of Lindahl et al.’s research [35].

3.2. Core–Rim Morphology of Sintered TiCN-Based Cermet

Figure 3 shows the microstructure of the Ti(C,N)-based cermets with Mo2C contents of 4 wt.%, 6 wt.%, 8 wt.%, 10 wt.% and 12 wt.%. Due to the addition of Mo2C, the Ti(C,N)-based cermets form a unique core–rim structure, namely black core–white inner rim–gray outer rim structure. As shown in Figure 3a–e, elements such as C, N, Ti, Cr, Co, Ni, Mo, and W were detected in the black core phase and rim phase of the five types of cermets. The composition difference between the black core phase and the rim phase is manifested in the higher content of C, N, and Ti in the black core phase, while it is manifested in the higher content of W and Mo in the rim phase, which is related to the formation of the black core phase and the rim phase. As the sintering temperature increases, Mo, TiC, TiN, and Ni diffuse with each other. Before the liquid phase appears, the undissolved TiC and TiN hard-phase particles become black core phases. Then, during the solid-state sintering stage, due to the occurrence of dissolution and precipitation, a white inner rim phase rich in W and Mo is formed around the black core phase. And in the later stage of liquid-phase sintering, excess TiC, TiN, WC, and Mo2C precipitate from the bonding phase, forming a gray outer rim phase around the white inner rim phase. As seen in Figure 3, when the content of Mo2C increases from 4 wt.% to 6 wt.%, the microstructures are significantly uniform, and the hard-phase grains are significantly refined, and the average grain diameter decreases from 1 μm to 0.5 μm. This is because sufficient Mo2C can form a complete ring phase via the dissolution precipitation mechanism, thereby improving the wettability of the bonding relative to the hard phase and reducing the direct contact between hard grains. However, as the Mo2C content continues to increase to 12 wt.%, the volume fraction of the rim phase increases and the proportion of core phase decreases. In addition, as the Mo2C content increases, the Mo content in the black core phase, white inner rim phase, and gray outer rim phase increases, while the Ti content in the gray outer rim phase decreases, which is also related to the increase in the volume fraction of the rim phase.

3.3. Mechanical Properties of Sintered TiCN-Based Cermet

The correlation between the Mo2C content and the relative density of the Ti(C,N)-based cermets is presented in Figure 4. As can be observed, the cermet achieves a constant relative density of more than 96% when the content of Mo2C is greater than 6 wt.%. This is due to the fact that a higher Mo2C concentration improves the wettability between the hard phase and the metal phase, and enhances the sinterability of cermet, leading to a denser microstructure [27]. As can be observed from Figure 4b,c, Mo2C also has a significant effect on the mechanical characteristics of cermets. The addition of 6 weight percent of Mo2C results in optimal mechanical properties, with a Vickers hardness of 17.86 GPa and a fracture toughness of 7.27 MPa.m1/2, which are significantly higher than the Ti(C,N)-based cermets without Mo2C addition (16.34 GPa [32] and 7.0 MPa·m1/2 [36]). The strengthening effects were mainly due to the core–rim structure induced by the addition of Mo2C.
As the Mo2C content is increased from 4 wt.% to 6 wt.%, the fracture toughness shows improvement, attributed to the refinement of hard-phase grains and the presence of thin rims. Thin rims are advantageous for fracture toughness, because thick rims can be more easily passed through by cracks or induce them under loading [37,38]. When the Mo2C content exceeds 6 wt.%, excessive Mo2C leads to the agglomerations of fine Ti(C,N) particles, which reduce the wettability of the binder phase on the cermet phase [39]. The excessive Mo2C addition inhibits the development of rims surrounding large Ti(C,N) cores as well. These cause negative effects on the fracture toughness of cermet. Additionally, the high fracture toughness of cermets also depends on the mean free path of the large binder [40], together with the increasing W content in the rims [41]. Therefore, the fracture toughness of cermets decreases with the increase in the Mo2C content, which results in a lower W content in the rims and a smaller binder mean free path.

3.4. Frictional Behavior and Wear Results

The tribological properties of Ti(C,N)-based cermets were assessed. Figure 5 shows the dynamic coefficients of friction (CoF), the mean coefficients of friction, and the wear volume of Ti(C,N)-based cermets with different Mo2C contents. MC14 is excluded due to its poor mechanical properties. As shown in Figure 6, three sorts of areas—unworn areas, debris accumulation areas, and worn areas—were formed on the attack surface of the cermets after the wear experiment along the sliding direction. Following the fretting wear test, the morphology of the worn surfaces of the cermets was investigated.
As demonstrated in Figure 5a, CoFs of all specimens show clear fluctuations throughout the running-in phase (10–20 min after the beginning), which is mostly brought on by the direct contact between rough peaks and the friction surfaces. After that, in the stable stage of the CoF curves, the mean CoFs are determined and the wear volume is estimated. As seen in Figure 5 and Figure 6, when the content of Mo2C increases from 4 wt.% to 12 wt.%, the friction coefficient and wear volume have a variation law of first decreasing, then increasing, and then decreasing. Specifically, as the Mo2C content increases from 4 wt.% to 6 wt.%, the wear volume (Figure 5c) and the mean CoF (Figure 5b) decrease due to an increase in the relative density hardness of the specimens (Figure 4b) When the Mo2C content goes from 6 wt.% to 8 wt.%, the hardness of the specimen diminishes, and the results of the friction test show that the CoF climbs to 0.61 and the wear volume also increases. Interestingly, as the Mo2C content rises from 8 wt.% to 12 wt.%, the CoF again declines from 0.61 to 0.54, and the wear loss drops as well. When the Mo2C contents are 6 wt.% and 12 wt.%, respectively, the specimens show low CoFs and wear volumes. Even so, the wear loss of the cermets with 12 wt.% Mo2C is slightly lower.

3.5. SEM and EDS Analysis of Worn Surfaces

After the fretting wear test, the detailed morphology of the worn surfaces of the cermets was studied in order to establish the relationship between the Mo2C content and the characteristics of the tribological surface, as shown in Figure 7. As seen in Figure 7a,b, when the Mo2C content is 4 wt.%, some scratches can be observed on the surface of the cermet, and Si is not detected, which means that the wear between the friction pairs is mainly abrasive wear. When the Mo2C content is low, the formed rim phase is incomplete, and the rim phase cannot completely wrap the hard phase, which results in the hard phase having a larger grain size and a poor interfacial bonding strength between the rim phase and the hard phase. Therefore, during friction and wear, the rim phase in the cermet is broken and the hard phase becomes abrasive particles.
As seen in Figure 7c,d, when the Mo2C content is 6 wt.%, the surface of the cermet has fewer scratches and debris, and the surface is smooth, which is mainly because the cermet with 6 wt.% Mo2C has the best mechanical properties (hardness and fracture toughness, as shown in Figure 4); in addition, the proper core–rim structure can improve the wettability between the liquid bonding phase and the solid hard phase, as well as the uniformity of the organizational structure, so as to increase the interfacial bonding strength between the rim phase and the hard phase. In the process of friction and wear, the rim phase is not easily broken, so it has good friction performance.
When the Mo2C content increases to 8–10 wt.%, as can be seen in Figure 7e–h, more scratches and debris appear on the surface of the cermet. From Table 5, there is a small amount of Si on the surface of the cermet after friction, which indicates that there exists abrasive wear and adhesive wear. With the increase in the Mo2C content, the thickness of the rim phase further increases, and the uniformity of the structure deteriorates. At this time, the brittleness of the rim phase plays a major role, the solid solution strengthening effect of the rim phase is not obvious, and the interfacial bonding strength between the rim phase and the hard phase decreases. During the friction and wear process, the rim phase is prone to breakage, resulting in abrasive wear and adhesive wear. However, as can be seen in Figure 7i,j, when the Mo2C content increases to 12 wt.%, there are fewer scratches and debris on the surface of the cermet, and the surface is smoother. According to the experimental results shown in Figure 7, the wear resistance of metal cermets with a Mo2C content of 6 wt.% and 12 wt.% is better, which is consistent with their wear volume in Figure 5c.
From Table 5, as the Mo2C content increases, the content of oxygen on the surface of the cermet after friction and wear increases, while Si mainly appears on the surface with a Mo2C content of 10 wt.%. This indicates that oxide forms during the friction and wear process, and Figure 7f shows that when the Mo2C content is 12 wt.%, the cermet surface wear is reduced. The formation of an oxide film improves the wear performance of the cermet and avoids further wear on the surface of the friction pair.

3.6. XPS Analysis of Worn Surfaces

XPS was used to analyze the surface characteristics of the sample. Figure 8a shows the wide-scan survey spectra of the cermets, corresponding to W4f, Mo3d, C1s, N1s, Ti2p and O1s at 231.9 eV, 284.8 eV, 396.4 eV, 455.44 eV and 530.37 eV, respectively. As can be seen from the spectra, the presence of O, Ti, Mo, C, and N was confirmed [42,43]. C1s at 284.8 eV was used as a standard for calibration [44], while C1s at 281.9 eV was assigned to TiCN in this study [45].
As can be seen in Figure 9, the Ti2p spectra show that there are two spin splitting peaks, namely 3/2 and 1/2, and that the distance between the two spin splitting peaks is about 6.0 eV. The satellite peak of TiN is also detected [46]. The W4f spectra exhibit two distinct double peaks: one at 32.1 and 34.0 eV for W4f7/2, and another at 35.3 and 37.5 eV for the W4f5/2 spectra. The low-energy side of the W4f double peaks is thought to be connected to the W–N bonding [47]. The XPS spectra of the O1s energy region for the Ti(C,N)-based cermets include two peaks. That at 529.76 eV is deemed to be due to Mo or W oxides, and that at 531.32 eV is consistent with the Ti oxides [46,48].
The XPS narrow-scan spectrum of Mo 3d is shown in Figure 9d. Two peaks at 231.92 eV and 235.41 eV can be attributed to the presence of MoO3 formation in the tribological films [42], while that at 228.6 eV corresponds to Mo in the hard phase (Ti,W,Mo)CN. This indicates that friction triggers the transformation of Mo2C into Mo oxides. It is noteworthy that the presence of MoO2 and MoO3 on the friction interface enhances the wear resistance properties [49]. This accounts for the observed decrease in the CoF and wear volume with the increasing Mo2C content.
Table 6 shows the relative atomic content of each compound calculated from the XPS spectrum fitting results. According to Table 6, it can be found that the products on the friction surface include TiN, WC, WO2, Mo2C, MoO2, Mo2O5, MoO3 and other compounds. When the Mo2C content increases from 4 wt.% to 12 wt.%, the relative contents of TiN, WC and WO2 show a variation law of first increasing, then decreasing, and then increasing. When the content of Mo2C is 6 wt.% and 12 wt.%, the content of WC has the maximum value (17.7 wt.% and 30.5 wt.%, respectively). The increase in WC helps to improve the toughness of the cermets, and thus enhances their friction and wear properties. In addition, as can be seen in Table 6, when the content of Mo2C increases from 4 wt.% to 12 wt.%, the content of MoO3 presents a variation law of first decreasing and then increasing, and the generation of MoO3 plays a certain protective role on the friction surface. When the content of Mo2C is 4 wt.%, the content of MoO3 is 26.9 wt.%. However, the friction coefficient and wear performance are poor (Figure 5 and Figure 7), which indicates that the friction and wear behavior of the ceramics at this time is mainly determined by the mechanical properties (Figure 4). The adhesive phase and hard phase are prone to shape deformation, resulting in particle stripping along the wear track, and the resulting abrasive wear increases the wear amount and friction coefficient. When the content of Mo2C is 6 wt.% and 8 wt.%, and the content of MoO3 is 24.1 wt.% and 19.3 wt.%, respectively, the mechanical properties of the cermet are close to each other (Figure 4). However, Mo2C with a content of 8 wt.% has the largest friction coefficient and wear amount (Figure 5 and Figure 7), which indicates that a sufficiently thick oxide friction film can avoid the further wear of the friction pair. When the content of Mo2C is 12 wt.%, the content of MoO3 is 20.5 wt.%, and the mechanical properties of the cermet are better (Figure 5); in addition, the tribological properties of the cermet are better under the dual action of mechanical properties and friction film. According to the results shown in Figure 6, Figure 7, Figure 8 and Figure 9 and Table 6, it can be inferred that the friction and wear behavior of the TiCN cermet is determined by the mechanical properties and oxide friction film.
The effects of Mo2C addition on the wear resistance of cermets could be divided into several aspects. It was confirmed that the addition of Mo2C could increase the hardness. Moreover, compared with Ti(C,N)-based cermets without Mo2C [50], the addition of Mo2C also enhanced the ability to form oxide film during the wear process, which greatly helped in the protection of the cermets’ surface from tear. However, the excessive addition of Mo2C could cause a decrease in the toughness, which was also illustrated by other researchers [27]. Considering the comprehensive influence of Mo2C, cermets with 6 wt.% Mo2C hold the best tribological properties.

4. Conclusions

Ti(C,N)-based cermets have been widely used in cutting tools or molds, and thus revealing their tribological behavior against hard materials is of great significance. This article conducted fretting wear tests and studied the effect of Mo2C on the tribological behavior of Ti(C,N)-based cermets. The main conclusions are as follows:
  • The addition of Mo2C could induce the formation of a “core–rim” structure, which exhibited significant strengthening effects on the mechanical properties, including hardness and fracture toughness. The core phase is mainly composed of Ti(C,N), while the rim phase is mainly composed of (W,Mo,Ti)(C,N). When the content of Mo2C is 6 wt.%, the cermets have optimal mechanical properties.
  • Due to the Mo2C addition and “core–rim” structure, the wear mechanisms of cermets are mainly abrasive wear, adhesive wear, and oxidation wear. As the content of Mo2C increases from 4 wt.% to 12 wt.%, the friction coefficient and wear volume have a variation law of first decreasing, then increasing, and then decreasing, and reach minimum values at 6 wt.% and 12 wt.%, which is the result of the comprehensive effect of the Mo2C strengthening and the oxide film.
  • The tribological behavior of Ti(C,N)-based cermets was determined according to the hardness, toughness and formation of an oxide film. On the one hand, the addition of Mo2C helped to increase the hardness and to form an oxide film, which plays a role in protecting the surface. On the other hand, the excessive carbide reduced the toughness and increased the risk of crushing. Combining these findings with practical engineering applications, and considering the mechanical properties, tribological properties, and cost factors, cermets with a Mo2C content of 6 wt.% are feasible.

Author Contributions

Conceptualization, H.Q.; methodology, H.Q. and C.P.; investigation, H.Q.; writing—original draft preparation, H.Q.; writing—review and editing, C.P. and J.F.; supervision, X.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China, grant number 51474108; the Natural Science Foundation of Guangdong Province, grant number 208081725043; the GF Basic Scientific Research Project, grant number JCKY2020205C002; and Innovation-driven Development Science and Technology Project of Economic Development District of Yichun, grant number 1203.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data utilized in the present work can be obtained from this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lengauer, W.; Scagnetto, F. Ti(C,N)-Based Cermets: Critical Review of Achievements and Recent Developments. Solid State Phenom. 2018, 274, 53–100. [Google Scholar] [CrossRef]
  2. Peng, Y.; Miao, H.; Peng, Z. Development of TiCN-Based Cermets: Mechanical Properties and Wear Mechanism. Int. J. Refract. Met. Hard Mater. 2013, 39, 78–89. [Google Scholar] [CrossRef]
  3. Ettmayer, P.; Kolaska, H.; Lengauer, W.; Dreyer, K. Ti(C,N) Cermets—Metallurgy and Properties. Int. J. Refract. Met. Hard Mater. 1995, 13, 343–351. [Google Scholar] [CrossRef]
  4. Rajabi, A.; Ghazali, M.J.; Syarif, J.; Daud, A.R. Development and Application of Tool Wear: A Review of the Characterization of TiC-Based Cermets with Different Binders. Chem. Eng. J. 2014, 255, 445–452. [Google Scholar] [CrossRef]
  5. Heydari, L.; Lietor, P.F.; Corpas-Iglesias, F.A.; Laguna, O.H. Ti(C,N) and WC-Based Cermets: A Review of Synthesis, Properties and Applications in Additive Manufacturing. Materials 2021, 14, 6786. [Google Scholar] [CrossRef]
  6. Zhuang, Q.; Lin, N.; He, Y.; Kang, X. Influence of Temperature on Sintering Behavior and Properties of TiC-Fe-Co-Ni-Cr-Mo Cermets. Ceram. Int. 2017, 43, 15992–15998. [Google Scholar] [CrossRef]
  7. Hadi, M.; Rafiaei, S.M.; Fernandes, F. The Effect of Mo2C Additions on the Oxidation Resistance of (Ti,W)CN Cermets as Base Material for the Production of Cutting Tools. Ceram. Int. 2023, 49, 21538–21545. [Google Scholar] [CrossRef]
  8. Verma, V.; Kumar, B.V.M. Sliding Wear Behavior of SPS Processed TaC-Containing Ti(CN)-WC-Ni/Co Cermets against Silicon Carbide. Wear 2017, 376–377, 1570–1579. [Google Scholar] [CrossRef]
  9. Kwon, W.T.; Park, J.S.; Kim, S.W.; Kang, S.; Tae, W.; Seuk, J.; Kim, S.W.; Kang, S.; Kwon, W.T.; Park, J.S.; et al. Effect of WC and Group IV Carbides on the Cutting Performance of Ti(C,N) Cermet Tools. Int. J. Mach. Tools Manuf. 2004, 44, 341–346. [Google Scholar] [CrossRef]
  10. Ahn, S.Y.; Kang, S. Formation of Core/Rim Structures in Ti(C,N)-WC-Ni Cermets via a Dissolution and Precipitation Process. J. Am. Ceram. Soc. 2000, 83, 1489–1494. [Google Scholar] [CrossRef]
  11. Binder, S.; Lengauer, W.; Ettmayer, P.; Bauer, J.; Debuigne, J.; Bohn, M. Phase Equilibria in the Systems TiCN, ZrCN and HfCN. J. Alloys Compd. 1995, 217, 128–136. [Google Scholar] [CrossRef]
  12. Hu, H.; Cheng, Y.; Yin, Z.; Zhang, Y.; Lu, T. Mechanical Properties and Microstructure of Ti(C, N) Based Cermet Cutting Tool Materials Fabricated by Microwave Sintering. Ceram. Int. 2015, 41, 15017–15023. [Google Scholar] [CrossRef]
  13. Rodriguez, N.; Sanchez, J.M.; Aristizabal, M. Consolidation of (Ti,Mo)(C,N)–Ni Cermets by Glass Encapsulated Hot Isostatic Pressing. Mater. Sci. Eng. A 2011, 528, 4453–4461. [Google Scholar] [CrossRef]
  14. Lemboub, S.; Boudebane, S.; Gotor, F.J.J.; Haouli, S.; Mezrag, S.; Bouhedja, S.; Hesser, G.; Chadli, H.; Chouchane, T. Core-Rim Structure Formation in TiC-Ni Based Cermets Fabricated by a Combined Thermal Explosion/Hot-Pressing Process. Int. J. Refract. Met. Hard Mater. 2018, 70, 84–92. [Google Scholar] [CrossRef]
  15. Chicardi, E.; Gotor, F.J.; Medri, V.; Guicciardi, S.; Lascano, S.; Córdoba, J.M. Hot-Pressing of (Ti,Mt)(C,N)-Co-Mo2C (Mt = Ta,Nb) Powdered Cermets Synthesized by a Mechanically Induced Self-Sustaining Reaction. Chem. Eng. J. 2016, 292, 51–61. [Google Scholar] [CrossRef]
  16. Monteverde, F.; Medri, V.; Bellosi, A. Microstructure of Hot-Pressed Ti(C,N)-Based Cermets. J. Eur. Ceram. Soc. 2002, 22, 2587–2593. [Google Scholar] [CrossRef]
  17. Sparling, W.A.; Plucknett, K.P. The Effects of Mo2C Additions on the Sintering Response of TiC0.3N0.7–Ni3Al Cermets. Int. J. Refract. Met. Hard Mater. 2016, 61, 98–106. [Google Scholar] [CrossRef]
  18. Xu, Q.; Zhao, J.; Ai, X.; Qin, W.; Wang, D.; Huang, W. Effect of Mo2C/(Mo2C+WC) Weight Ratio on the Microstructure and Mechanical Properties of Ti(C,N)-Based Cermet Tool Materials. J. Alloys Compd. 2015, 649, 885–890. [Google Scholar] [CrossRef]
  19. Lin, Z.; Xiong, J.; Guo, Z.; Zhou, W.; Wan, W.; Yang, L. Effect of Mo2C Addition on the Microstructure and Fracture Behavior of (W,Ti)C-Based Cemented Carbides. Ceram. Int. 2014, 40, 16421–16428. [Google Scholar] [CrossRef]
  20. Lindahl, P.; Rolander, U.; Andrén, H.O. Atom-Probe Analysis of the Binder Phase in TiC-TiN-Mo2C-(Ni, Co) Cermets. Int. J. Refract. Met. Hard Mater. 1993, 12, 115–119. [Google Scholar] [CrossRef]
  21. Rafiaei, S.M.; Bahrami, A.; Shokouhimehr, M. Influence of Ni/Co Binders and Mo2C on the Microstructure Evolution and Mechanical Properties of (Ti0.93W0.07)C–Based Cermets. Ceram. Int. 2018, 44, 17655–17659. [Google Scholar] [CrossRef]
  22. Engqvist, H.; Högberg, H.; Botton, G.; Ederyd, S.; Axén, N. Tribofilm Formation on Cemented Carbides in Dry Sliding Conformal Contact. Wear 2000, 239, 219–228. [Google Scholar] [CrossRef]
  23. Espinosa-Fernandez, L.; Borrell, A.; Salvador, M.D.; Gutierrez-Gonzalez, C.F. Sliding Wear Behavior of WC-Co-Cr3C2-VC Composites Fabricated by Conventional and Non-Conventional Techniques. Wear 2013, 307, 60–67. [Google Scholar] [CrossRef]
  24. Manoj Kumar, B.V.; Basu, B.; Kalin, M.; Vizintin, J. Load-Dependent Transition in Sliding Wear Properties of TiCN–WC–Ni Cermets. J. Am. Ceram. Soc. 2007, 90, 1534–1540. [Google Scholar] [CrossRef]
  25. Verma, V.; Manoj Kumar, B. V Tribological Characteristics of Conventionally Sintered TiCN-WC-Ni/Co Cermets against Cemented Carbide. Ceram. Int. 2017, 43, 368–375. [Google Scholar] [CrossRef]
  26. Qiu, H.; Li, X.; Pan, C.; Fan, J.; Qu, S. Effect of Mo2C Addition on the Formation of Core-Rim Structure and Mechanical Properties of Ti(C,N)-WC-Mo2C-(Ni,Co) Cermet. J. Mater. Res. Technol. 2023, 25, 750–762. [Google Scholar] [CrossRef]
  27. Liu, C.; Lin, N.; He, Y.H. Influence of Mo2C and TaC Additions on the Microstructure and Mechanical Properties of Ti(C,N)-Based Cermets. Ceram. Int. 2016, 42, 3569–3574. [Google Scholar] [CrossRef]
  28. Lóh, N.J.; Simão, L.; Faller, C.A.; De Noni, A.; Montedo, O.R.K. A Review of Two-Step Sintering for Ceramics. Ceram. Int. 2016, 42, 12556–12572. [Google Scholar] [CrossRef]
  29. Yang, Q.; Xiong, W.; Zhang, M.; Huang, B.; Chen, S. Microstructure and Mechanical Properties of Mo-Free Ti(C,N)-Based Cermets with Ni–XCr Binders. J. Alloys Compd. 2015, 636, 270–274. [Google Scholar] [CrossRef]
  30. Shetty, D.K.; Wright, I.G.; Mincer, P.N.; Clauer, A.H. Indentation Fracture of WC-Co Cermets. J. Mater. Sci. 1985, 20, 1873–1882. [Google Scholar] [CrossRef]
  31. Cardinal, S.; Malchère, A.; Garnier, V.; Fantozzi, G. Microstructure and Mechanical Properties of TiC-TiN Based Cermets for Tools Application. Int. J. Refract. Met. Hard Mater. 2009, 27, 521–527. [Google Scholar] [CrossRef]
  32. Xu, L.W.; Lin, N.; Zhao, L.B.; Ma, C.; Wang, Z.Y.; He, Y.H. Improvement in Wear and Corrosion Resistance of Ti(C,N)-Based Cermets via Aluminum Nitride Additions. Vacuum 2019, 168, 108849. [Google Scholar] [CrossRef]
  33. Ye, J.; Xiong, J.; Guo, Z.; Yang, T.; Zhou, L.; Liu, J.; Li, S.; Li, T. Preparation of Self-Lubricating Ti(C,N)-Based Cermets by Solid Carburization and Wear Behavior. Int. J. Refract. Met. Hard Mater. 2017, 67, 68–73. [Google Scholar] [CrossRef]
  34. Akhtar, F.; Guo, S.J. Microstructure, Mechanical and Fretting Wear Properties of TiC-Stainless Steel Composites. Mater. Charact. 2008, 59, 84–90. [Google Scholar] [CrossRef]
  35. Lindahl, P.; Gustafson, P.; Rolander, U.; Stals, L.; Andrén, H.O. Microstructure of Model Cermets with High Mo or W Content. Int. J. Refract. Met. Hard Mater. 1999, 17, 411–421. [Google Scholar] [CrossRef]
  36. Demoly, A.; Lengauer, W.; Veitsch, C.; Rabitsch, K. Effect of Submicron Ti(C,N) on the Microstructure and the Mechanical Properties of Ti(C,N)-Based Cermets. Int. J. Refract. Met. Hard Mater. 2011, 29, 716–723. [Google Scholar] [CrossRef]
  37. Xiong, H.; Wen, Y.; Gan, X.; Li, Z.; Chai, L. Influence of Coarse TiCN Content on the Morphology and Mechanical Properties of Ultrafine TiCN-Based Cermets. Mater. Sci. Eng. A 2017, 682, 648–655. [Google Scholar] [CrossRef]
  38. García, J.; Collado Ciprés, V.; Blomqvist, A.; Kaplan, B. Cemented Carbide Microstructures: A Review. Int. J. Refract. Met. Hard Mater. 2019, 80, 40–68. [Google Scholar] [CrossRef]
  39. Aramian, A.; Sadeghian, Z.; Narimani, M.; Razavi, N.; Berto, F. A Review on the Microstructure and Properties of TiC and Ti(C,N) Based Cermets. Int. J. Refract. Met. Hard Mater. 2023, 115, 106320. [Google Scholar] [CrossRef]
  40. Walbrühl, M.; Linder, D.; Ågren, J.; Borgenstam, A. Diffusion Modeling in Cemented Carbides: Solubility Assessment for Co, Fe and Ni Binder Systems. Int. J. Refract. Met. Hard Mater. 2017, 68, 41–48. [Google Scholar] [CrossRef]
  41. Zheng, Y.; Wang, S.; You, M.; Tan, H.; Xiong, W. Fabrication of Nanocomposite Ti(C,N)-Based Cermet by Spark Plasma Sintering. Mater. Chem. Phys. 2005, 92, 64–70. [Google Scholar] [CrossRef]
  42. Zhou, Y.L.; Luo, D.M. Corrosion Behavior of Ti-Mo Alloys Cold Rolled and Heat Treated. J. Alloys Compd. 2011, 509, 6267–6272. [Google Scholar] [CrossRef]
  43. Oshikawa, K.; Nagai, M.; Omi, S. Characterization of Molybdenum Carbides for Methane Reforming by TPR, XRD, and XPS. J. Phys. Chem. B 2001, 105, 9124–9131. [Google Scholar] [CrossRef]
  44. Gengenbach, T.R.; Major, G.H.; Linford, M.R.; Easton, C.D. Practical Guides for X-ray Photoelectron Spectroscopy (XPS): Interpreting the Carbon 1s Spectrum. J. Vac. Sci. Technol. A 2021, 39, 13204. [Google Scholar] [CrossRef]
  45. Meng, J.; Lu, J.; Wang, J.; Yang, S. Tribological Behavior of TiCN-Based Cermets at Elevated Temperatures. Mater. Sci. Eng. A 2006, 418, 68–76. [Google Scholar] [CrossRef]
  46. Zhang, G.; Zhao, Z.; Zheng, H.; Zhang, S.; Chen, Z.; Wang, S.; Zhao, X. Effects of the Addition of Multi-Walled Carbon Nanotubes on the Microstructure and Properties of Ti(C,N)-Based Cermets Prepared by Spark Plasma Sintering. Mater. Charact. 2022, 193, 112289. [Google Scholar] [CrossRef]
  47. Ren, B.; Li, D.; Jin, Q.; Cui, H.; Wang, C. A Self-Supported Porous WN Nanowire Array: An Efficient 3D Electrocatalyst for the Hydrogen Evolution Reaction. J. Mater. Chem. A 2017, 5, 19072–19078. [Google Scholar] [CrossRef]
  48. Qiao, X.-Q.; Wang, Z.; Li, C.; Zhang, H.; Hou, D.; Lan, Y.-Q.; Li, D.-S. A New, Efficient and Durable MoO2/Mo2C-C Cocatalyst with the Optimized Composition and Electronic Structure Via In-Situ Carburization for Photocatalytic H2 Evolution. SSRN Electron. J. 2022, 455, 140791. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Feng, K.; Yao, C.; Li, Z. Effect of MoO3 on the Microstructure and Tribological Properties of Laser-Clad Ni60/NanoCu/h-BN/MoO3 Composite Coatings over Wide Temperature Range. Surf. Coat. Technol. 2020, 387, 125477. [Google Scholar] [CrossRef]
  50. Yi, M.; Wang, J.; Li, C.; Bai, X.; Wei, G.; Zhang, J.; Xiao, G.; Chen, Z.; Zhou, T.; Wang, L.; et al. Friction and Wear Behavior of Ti(C,N) Self-Lubricating Cermet Materials with Multilayer Core-Shell Microstructure. Int. J. Refract. Met. Hard Mater. 2021, 100, 105629. [Google Scholar] [CrossRef]
Figure 1. The temperature curve of the vacuum hot press sintering of TiCN-based cermets.
Figure 1. The temperature curve of the vacuum hot press sintering of TiCN-based cermets.
Materials 16 05645 g001
Figure 2. XRD patterns of sintered cermets with various Mo2C contents: (a) the complete pattern; (b) enlarged peaks of hard phases; (c) enlarged peaks of binder and pure Co/Ni.
Figure 2. XRD patterns of sintered cermets with various Mo2C contents: (a) the complete pattern; (b) enlarged peaks of hard phases; (c) enlarged peaks of binder and pure Co/Ni.
Materials 16 05645 g002
Figure 3. SEM–BSE micrographs and EDS point analysis of the specimens with various Mo2C contents: (a) 4 wt.%, (b) 6 wt.%, (c) 8 wt.%, (d) 10 wt.%, (e) 12 wt.%.
Figure 3. SEM–BSE micrographs and EDS point analysis of the specimens with various Mo2C contents: (a) 4 wt.%, (b) 6 wt.%, (c) 8 wt.%, (d) 10 wt.%, (e) 12 wt.%.
Materials 16 05645 g003
Figure 4. Mechanical properties of cermets: (a) relative density, (b) hardness and (c) fracture toughness.
Figure 4. Mechanical properties of cermets: (a) relative density, (b) hardness and (c) fracture toughness.
Materials 16 05645 g004
Figure 5. Wear results of Ti(C,N)-based cermets: (a) dynamic CoF; (b) mean CoF; (c) wear volume.
Figure 5. Wear results of Ti(C,N)-based cermets: (a) dynamic CoF; (b) mean CoF; (c) wear volume.
Materials 16 05645 g005
Figure 6. Images of the wear tracks on the Ti(C,N)-based cermets with various Mo2C contents: (a,b) 4 wt.%, (c,d) 6 wt.%, (e,f) 8 wt.%, (g,h) 10 wt.%, (i,j) 12 wt.%.
Figure 6. Images of the wear tracks on the Ti(C,N)-based cermets with various Mo2C contents: (a,b) 4 wt.%, (c,d) 6 wt.%, (e,f) 8 wt.%, (g,h) 10 wt.%, (i,j) 12 wt.%.
Materials 16 05645 g006
Figure 7. SEM–BSE micrographs of the cermets’ worn surfaces after fretting at a 50 N load with various Mo2C contents: (a,b) 4 wt.%, (c,d) 6 wt.%, (e,f) 8 wt.%, (g,h) 10 wt.%, (i,j) 12 wt.%.
Figure 7. SEM–BSE micrographs of the cermets’ worn surfaces after fretting at a 50 N load with various Mo2C contents: (a,b) 4 wt.%, (c,d) 6 wt.%, (e,f) 8 wt.%, (g,h) 10 wt.%, (i,j) 12 wt.%.
Materials 16 05645 g007
Figure 8. XPS survey spectra of specimens with various Mo2C contents: (a) wide scan; (b) C 1s.
Figure 8. XPS survey spectra of specimens with various Mo2C contents: (a) wide scan; (b) C 1s.
Materials 16 05645 g008
Figure 9. XPS spectra of specimens with various Mo2C contents: (a) Ti 2p; (b) W 4f; (c) O 1s; (d) Mo 3d.
Figure 9. XPS spectra of specimens with various Mo2C contents: (a) Ti 2p; (b) W 4f; (c) O 1s; (d) Mo 3d.
Materials 16 05645 g009
Table 1. The particle sizes and the purity of the initial powders.
Table 1. The particle sizes and the purity of the initial powders.
PowderTiCNWCMo2CCr3C2NiCo
Particle size (μm)<1<0.82–5<0.8<5<5
purity99.9%99.9%99.9%99.9%99.8%99.8%
Table 2. The nominal composition of Ti(C,N)-based cermets (wt.%).
Table 2. The nominal composition of Ti(C,N)-based cermets (wt.%).
SpecimenComposition (wt.%)
TiCNCoNiMo2CWCCr3C2
MC4Bal.5104101
MC6Bal.5106101
MC8Bal.5108101
MC10Bal.51010101
MC12Bal.51012101
MC14Bal.51014101
Table 3. Parameters of fretting wear test.
Table 3. Parameters of fretting wear test.
ParameterSet Value
Stroke (μm)200
Frequency (Hz)20
Preloading Time (min)5
Preloading Force (N)5
Normal Force (N)50
Loading Time (min)50
Table 4. Lattice parameters (Å) of various phases based on the XRD patterns in Figure 2.
Table 4. Lattice parameters (Å) of various phases based on the XRD patterns in Figure 2.
SamplesMo2C
(wt.%)
Black Core/Grey RimWhite Core/Grey RimBinder
MC444.287-3.557
MC664.2964.2863.568
MC884.3024.2903.572
MC10104.3024.2893.573
MC12124.3084.2903.577
Table 5. EDS analysis of point A in Figure 7 on the worn surfaces of cermets after fretting (wt.%).
Table 5. EDS analysis of point A in Figure 7 on the worn surfaces of cermets after fretting (wt.%).
SamplesCNOTiSiCrCoNiMoW
MC410.039.871.3666.73- 0.532.688.79
MC69.059.663.8361.4- 1.122.153.799.01
MC88.02 16.6350.24-0.450.71.199.7513.01
MC104.23015.9249.71.190.520.380.712.1715.2
MC125.241.2614.7652.82-0.52.194.459.019.79
Table 6. Analysis of XPS spectra on the worn surface of cermets.
Table 6. Analysis of XPS spectra on the worn surface of cermets.
Elements and
Compounds
Binding
Energy (eV)
at. %Elements and
Compounds
Binding
Energy (eV)
at. %
MC4Ti 2p3/2455.167.51Mo2C 3d5/2228.5611.69
Ti 2p1/2461.168.31Mo2C 3d3/2231.8611.66
Ti 2p3/2 sat.458.3536.57MoO2 3d5/2228.6117.93
Ti 2p1/2 sat.462.9313.37MoO2 3d3/2231.6117.74
Ti others455.8434.25Mo2O5 3d5/2230.177.01
WC 4f7/232.404.33Mo2O5 3d3/2233.177.02
WC 4f5/234.474.33MoO3 3d5/2232.4713.47
WO2 4f7/233.461.14MoO3 3d3/2235.4713.48
WO2 4f5/235.611.15
W 4f others34.5061.98
W 4f others36.9927.06
MC6Ti 2p3/2455.165.31Mo2C 3d5/2228.647.91
Ti 2p1/2461.165.91Mo2C 3d3/2231.647.92
Ti 2p3/2 sat.458.2635.15MoO2 3d5/2228.9322.16
Ti 2p1/2 sat.462.7718.50MoO2 3d3/2231.9321.90
Ti others455.9235.13Mo2O5 3d5/2230.178.03
WC 4f7/232.406.70Mo2O5 3d3/2233.178.04
WC 4f5/234.4710.99MoO3 3d5/2235.5812.03
WO2 4f7/233.462.96MoO3 3d3/2232.5812.02
WO2 4f5/235.537.89
W 4f others35.0945.03
W 4f others37.3526.44
MC8Ti 2p3/2455.166.70Mo2C 3d5/2228.6411.44
Ti 2p1/2461.167.45Mo2C 3d3/2231.8511.45
Ti 2p3/2 sat.458.363.58MoO2 3d5/2228.935.91
Ti 2p1/2 sat.462.357.93MoO2 3d3/2234.857.92
Ti others456.0474.34Mo2O5 3d5/2230.1721.97
WC 4f7/232.403.97Mo2O5 3d3/2233.1722.00
WC 4f5/234.473.97MoO3 3d5/2232.479.65
WO2 4f7/233.461.15MoO3 3d3/2235.779.66
WO2 4f5/235.531.16
W 4f others34.6547.66
W 4f others37.0142.09
MC10Ti 2p3/2455.169.30Mo2C 3d5/2228.6421.04
Ti 2p1/2460.9610.15Mo2C 3d3/2231.8021.06
Ti 2p3/2 sat.457.8112.06MoO2 3d5/2228.9313.24
Ti 2p1/2 sat.462.3413.39MoO2 3d3/2231.9313.25
Ti others455.5855.09Mo2O5 3d5/2230.177.73
WC 4f7/232.234.76Mo2O5 3d3/2233.177.69
WC 4f5/234.364.76MoO3 3d5/2232.477.99
WO2 4f7/233.462.81MoO3 3d3/2235.478.00
WO2 4f5/235.462.81
W 4f others34.9643.09
W 4f others37.2941.77
MC12Ti 2p3/2455.168.98Mo2C 3d5/2228.6417.07
Ti 2p1/2460.969.85Mo2C 3d3/2231.7811.29
Ti 2p3/2 sat.458.1021.60MoO2 3d5/2228.9311.69
Ti 2p1/2 sat.462.4717.93MoO2 3d3/2232.227.73
Ti others456.0432.52Mo2O5 3d5/2230.1719.10
WC 4f7/232.4017.18Mo2O5 3d3/2233.2512.63
WC 4f5/234.5013.33MoO3 3d5/2232.4712.34
WO2 4f7/233.463.57MoO3 3d3/2235.558.16
WO2 4f5/235.132.82
W 4f others35.2141.74
W 4f others37.6121.36
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

Qiu, H.; Li, X.; Pan, C.; Fan, J. Effect of Mo2C Addition on the Tribological Behavior of Ti(C,N)-Based Cermets. Materials 2023, 16, 5645. https://doi.org/10.3390/ma16165645

AMA Style

Qiu H, Li X, Pan C, Fan J. Effect of Mo2C Addition on the Tribological Behavior of Ti(C,N)-Based Cermets. Materials. 2023; 16(16):5645. https://doi.org/10.3390/ma16165645

Chicago/Turabian Style

Qiu, Hao, Xiaoqiang Li, Cunliang Pan, and Jiafeng Fan. 2023. "Effect of Mo2C Addition on the Tribological Behavior of Ti(C,N)-Based Cermets" Materials 16, no. 16: 5645. https://doi.org/10.3390/ma16165645

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