Next Article in Journal
Prospects of Using Pseudobrookite as an Iron-Bearing Mineral for the Alkaline Electrolytic Production of Iron
Next Article in Special Issue
Studies on Electron Escape Condition in Semiconductor Nanomaterials via Photodeposition Reaction
Previous Article in Journal
Syntheses and Characterizations of CuIn1-xZnxSe2 Chalcopyrite Nanoparticles
Previous Article in Special Issue
Effect of Al2TiO5 Content and Sintering Temperature on the Microstructure and Residual Stress of Al2O3–Al2TiO5 Ceramic Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Tribological Properties of Spark-Plasma-Sintered Ti3SiC2-Pb-Ag Composites at Elevated Temperatures

1
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
2
Sichuan Province Engineering Technology Research Center of Powder Metallurgy, Chengdu University, Chengdu 610106, China
3
Institute for Advanced Materials Deformation and Damage from Multi-Scale, Chengdu University, Chengdu 610106, China
4
School of Mechanical Engineering, Xinjiang University, Urumqi 830000, China
5
School of Chemical Engineering and Materials, Changzhou Institute of Technology, Changzhou 213032, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(4), 1437; https://doi.org/10.3390/ma15041437
Submission received: 29 December 2021 / Revised: 25 January 2022 / Accepted: 10 February 2022 / Published: 15 February 2022
(This article belongs to the Special Issue Microstructural Design and Processing Control of Advanced Ceramics)

Abstract

:
Ti3SiC2-PbO-Ag composites (TSC-PA) were successfully prepared using the spark plasma sintering (SPS) technique. The ingredient and morphology of the as-synthesized composites were elaborately investigated. The tribological properties of the TSC-PA pin sliding against Inconel 718 alloys disk at room temperature (RT) to 800 °C were examined in air. The wear mechanisms were argued elaborately. The results showed that the TSC-PA was mainly composed of Ti3SiC2, Pb, and Ag. The average friction coefficient of TSC-PA gradually decreased from 0.72 (RT) to 0.3 (800 °C), with the temperature increasing from RT to 800 °C. The wear rate of TSC-PA showed a decreasing trend, with the temperature rising from RT to 800 °C. The wear rate of Inconel 718 exhibited positive wear at RT and negative wear at elevated temperatures. The tribological property of TSC-PA was related to the tribo-chemistry, and the abrasive and adhesive wear.

1. Introduction

Ti3SiC2 belongs to one of the MAX phases (polycrystalline nanolaminates of ternary carbides and nitrides, over 60 + phases) [1]. It possesses the combined characteristics of metal and ceramics, such as being easily machinable, being electrically and thermally conductive, being oxidation resistant, and having a high melting point. Ti3SiC2 can be a potential candidate for structural functional components. Similar to graphite and MoS2, it has a hexagonal structure, endowing it with an exceptional lubricating property. Research on the tribological properties of Ti3SiC2 demonstrated that it is not a special solid lubricant. Contrarily, Ti3SiC2 generally has a high friction coefficient (μ > 0.6) and a high wear rate (WR: in the order of 10−2–10−3 mm3/N m), especially at medium and low temperatures (RT) [2,3,4,5,6,7,8]. It has been reported that, in some circumstances (such as with temperature > 600 °C, at a speed of >5 m/s, or sliding against some particular tribopair material), polycrystalline Ti3SiC2 displays the relatively low μ and wear rate, which mainly results from the tribo-oxidation transfer films generated on the contact surface [2,6,8,9,10]. In other situations, the fracture and pulling-out of Ti3SiC2 grains causes three-body abrasive wear, which is the main wear mechanisms for Ti3SiC2.
Subsequently, most of scientific enthusiasts concentrate on the intensifying research of polycrystalline Ti3SiC2 by uniting metal and/or ceramic to enhance its friction and wear behaviors. Ren et al. [11] prepared Ti3SiC2/Cu/Al/SiC composites, and they found that the tribological behaviors of the composite were better than those of polycrystalline Ti3SiC2 at RT and 200 °C, while the wear properties of the composite were worse than that of polycrystalline Ti3SiC2 at elevated temperatures (≥400 °C). The incorporation of Al, Cu, SiC and Al2O3 played an important role in reinforcing the bonding of Ti3SiC2 grains and in fixing the soft Ti3SiC2 matrix around them during reciprocate sliding at RT–200 °C, where the wear mechanism was abrasive wear. At elevated temperatures (≥400 °C), the plastic flow of tribo-oxides layers led to the higher wear rate of the composite than that of Ti3SiC2 where the wear mechanism was adhesive wear. Yang et al. [12] prepared Cu-Ti3SiC2 co-continuous composites. They showed excellent electrical conductivity and wear resistance due to the addition of Cu. Yang et al. [13] explored the tribological behaviors of (TiB2 + TiC)/Ti3SiC2 composites. It was discovered that μ of the composites was bigger compared with that of Ti3SiC2. The addition of TiB2 and TiC could fix the soft matrix around them and scatter the shear stresses. In our earlier work [14], we discussed the tribological behavior of Ti3SiC2/CaF2 composites at elevated temperatures. The prepared Ti3SiC2/CaF2 exhibited better friction and wear property than Ti3SiC2 in a wide temperature range due to the tribo-oxide competition. Islak et al. [15] investigated the effect of reinforcing TiB2 particles (5, 10, and 15 wt.%) on the properties of as-synthesized Ti3SiC2. The addition of TiB2 improved the mechanical properties and thermal diffusivity of the composites. Moreover, the addition of TiB2 significantly increased the wear properties of Ti3SiC2 matrix. Magnus et al. [16] prepared the Ti3SiC2-TiSi2-TiC composites by spark plasma sintering (SPS). They discovered that the tribological property of the composites was attributed to the intrinsic lubricity and the addition of second phase TiC particles. During the sliding, the tribo-oxidative wear was the main wear mechanism; then, it changed to the deformation-induced wear. The transition resulted from the breakdown of the formed tribofilm.
Silver shows some outstanding characteristics (such as good thermal conductivity, favorable ductility, and special mechanical behaviors) and can be widely applied in many fields, such as air-foil bearings, and biomedical and thermal interface materials. It is commonly utilized as a solid lubricant at temperatures below 500 °C. Additionally, it has a high diffusion coefficient and easily generates a lower shear stress intersection at the sliding interface [17,18]. Thus, it was significant and practical to incorporate good properties of Ag with Ti3SiC2. The goal of this work was to synthesize Ti3SiC2/Ag composites and to uncover the mechanical and tribological behaviors of the composite at raising temperatures. F. AlAnazi et al. [19] prepared metals-based (Ag and Bi) composites incorporation with 5 vol%, 10 vol%, 20 vol%, and 30 vol% Ti3SiC2 and investigated their tribological properties. They found that the incorporation of Ti3SiC2 improved the hardness and compressive yield strength of the two composites. Additionally, the addition of Ti3SiC2 enhanced the tribological performance of the two composites. Zeng et al. [20] inspected the tribological behaviors of Ti3SiC2/Ag composites with Inconel 718 as the tribopair at RT. At different sliding speeds, the composites showed lower μ and wear rate than that of Ti3SiC2.
Pb is lubricous due to its low shear strength. In our previous work [21,22], we found that the incorporation of PbO (melting point: 888 °C) instead of Pb (melting point: 327 °C) in Ti3SiC2 can successfully prepare Ti3SiC2/Pb composites and avoid the loss of Pb during the sintering process. The as-prepared composite showed better friction and wear behaviors than Ti3SiC2 at elevated temperatures.
As a result of the above-mentioned facts, it was necessary to suggest some novel composites that can move forward a single step in strengthening the tribological performance of Ti3SiC2 in a wide temperature range. Therefore, in this work, the metal Pb was added to the Ti3SiC2/Ag composite to prepare Ti3SiC2-Pb-Ag composites (TSC-PA). To prevent the loss of Pb during the process of preparing the TSC-PA, PbO was chosen as the source of Pb. The target of this job was to synthesize TSC-PA and to discuss the mechanical and tribological behaviors of the as-synthesized compound at elevated temperatures.

2. Experimental Procedure

2.1. Materials

The TSC-PA composites were sintered using the powder mixture of Ti3SiC2, PbO and Ag. The average particle size of Ti3SiC2 was 3 μm and its purity was ≥98%. It was purchased from Jinhezhi Materials Ltd., Beijing, China. The average particle size of PbO and Ag was 5 μm, and their purity was ≥99%. They were bought from Xilong Chemical Ltd., Shantou, China. The powder composition was 70 vol% Ti3SiC2-15 vol%PbO-15 vol%Ag. The mixture with a designated composition was blended by ball mill for 6 h. Next, it was loaded in a graphite die. In the end, it was sintered by SPS furnace (Shanghai Chenhua Electric Furnace Co., Ltd., Shanghai, China) in vacuum at 1130 ℃.

2.2. Mechanical Properties

The densities of the composites were determined using Archimedes’s principle. Micro-hardness was measured on a Micro-hardness Tester (MH-5-VM, Shanghai Hengyi Technology Co., Ltd., Shanghai, China). with a load of 500 g and a dwell time of 10 s. Flexural strength and compression strength were measured using a universal material tester (SANS-CMT5205, Shenzhen New Sansi Material Testing Co., Ltd., Shenzhen, China). The three-point bending test was conducted to obtain the flexural strength, and the samples were cut to 3 mm high × 4 mm wide × 20 mm long. The cross-head speed and span were 0.05 mm/min and 16 mm, respectively. The sizes of the samples for the compression strength test were Φ 5 mm × 12.5 mm, and the cross-head speed was 0.2 mm/min.

2.3. Friction and Wear Test

The friction and wear experiments were carried out in air on a high-temperature tribometer with a pin-on-disk setup (THT01-04015, CSM Instruments SA, Peseux, Switzerland). The TSC-PA was used as a pin, the size of which was Φ 6 mm × 12 mm. The Inconel 718 was used as a disk, and its size was Φ 32 mm × 8 mm. The surfaces of the disk and pin were ground and polished into a surface roughness (Ra) of 0.06 μm. The tribopairs slid against each other for 200 m at a sliding speed of 0.1 m/s under a normal load of 5 N. The friction coefficient was automatically documented by computer during the experiment. The wear volume was obtained by quantifying the volume loss of TSC-PA pin by optical microscopy and by examining the cross-sectional area of the worn area of Inconel 718 disk by 3D surface profilometry (NanoMap-D, Columbus, OH, USA). The wear rates were acquired by Equation (1).
Wear   rate   ( mm 3 N · m ) = W e a r   v o l u m e   ( mm 3 ) S l i d i n g   d i s t a n c e   ( m ) × N o r m a l   l o a d   ( N )

2.4. Analysis

The morphology of TSC-PA was investigated by scanning electron microscopy (SEM, JSM-5600LV, JEOL, Tokyo, Japan) equipped with energy dispersive X-ray spectroscopy. The EDS (IE250, Oxford Instrument, Abingdon, Oxfordshire, UK) was equipped with X-Max Silicon Drift Detector (SDD; the spectral resolution of 124 eV at Mn Kα) and an ultra-thin window; the active area of the SDD was 50 mm2. The phase composition of TSC-PA was analyzed by X-ray diffraction (XRD, Philips X’Pert Pro, PANalytical Netherlands, Almelo, Netherland) using Cu Kα radiation (λ = 0.15418 nm) within the 2θ-angle range from 5° to 90°. X-ray photoelectron spectroscope (XPS, PHI-5702, Physical Electronics Corporation, Chanhassen, MN, USA) was used to analyze the elemental chemical states of TSC-PA on the worn surfaces, and the binding energy of adventitious carbon (C1s: 284.8 eV) was used as the reference.

3. Result and Discussion

3.1. Phase Composition and Microstructure of TSC-PA

The XRD pattern of TSC-PA is displayed in Figure 1. As noticed in Figure 1, the main phases of TSC-PA were Ti3SiC2, Pb, and Ag. The Ti3SiC2 was temporarily stable and did not break down during the sintering process of TSC-PA. To the best of the authors’ knowledge, the Ti3SiC2 revealed high reaction activity when in contact with the metal phases, such as Cu [23,24,25], Al [26,27], and Fe [28]. In this study, the Ti3SiC2 did not decompose with the co-addition of Ag and PbO at the sintering temperature of 1130 °C. In a previous paper [21], we reported that PbO in the Ti3SiC2-PbO system was deoxidized to Pb by C or Si during the preparation process and Ti3SiC2 did not decompose at 1200 °C. The relevant chemical reactions were described in Reference [21]. In comparison, the addition of Ag did not change the composition of the Ti3SiC2-PbO system.
As observed in Figure 2, the morphology and the distributions of elements of the cross section of the TSC-PA was analyzed using SEM. It can be observed that the Pb particles were homogeneously scattered in the Ti3SiC2 matrix, while the Ag particles were agglomerated to a certain extent. As the melting point of silver was merely 961 °C, some amount of molten silver nearby flowed together during sintering. Therefore, the agglomeration of Ag occurred in the TSC-PA.

3.2. Mechanical Performance of TSC-PA

The relative density, microhardness, flexural strength, and compressive strength of TSC-PA and TSC are listed in Table 1. It can be seen that the relative densities of TSC-PA and TSC were 95.35% and 98.24%, respectively. The microhardness of TSC-PA was less than half that of TSC, and their microhardness values were 2.24 and 5.5 GPa, respectively. Pb and Ag in the TSC-PA belong to soft metals and caused the low hardness of TSC-PA. The flexural strength of TSC-PA was less than that of TSC, and their values were 311 and 428 MPa, respectively. The compressive strength of TSC-PA was only half that of TSC, and their values were 654 MPa and 1230 MPa, respectively.
As shown in Figure 3a, the TSC-PA only suffered from elastic deformation before fracture, which was a typical brittle fracture. The fracture surface of TSC-PA (see Figure 3b) indicated that the fracture surface of TSC-PA was featured by intergranular and transgranular fracture. This fracture mode of TSC-PA was similar to TSC. As well known, the diverse energy-absorbing strategies including the diffuse of the microcracks, lamination, crack deflection, and grain pull-out were discovered to be in charge of the high flexural strength of TSC [29,30]. In addition to intergranular and transgranular fracture, some pores and cracks resulting from the unsatisfactory compatibility (e.g., wettability) of the TSC matrix and Pb and Ag were observed (see Figure 3b). The thermal expansion coefficients (for instance: TSC: 9.1 × 10−6, Pb: 29.3 × 10−6, and Ag: 19.5 × 10−6 °C−1) of different phases were not matched with each other, leading to the weak bonding between granules and the formation of some defects (such as micro-pores and micro-cracks) during the preparation procedure. The as-formed defects were detrimental to the flexural strength. An additional major justification for the poor fracture strength of the TSC-PA was that the distribution of Pb and Ag around the TSC grains to a certain extent hindered the diverse energy-assimilating strategies, which let fracture take place between the grains.

3.3. Tribological Behavior of TSC-PA

Figure 4a showed the plot of μ versus distance curve of the TSC-PA/Inconel 718 tribopair at elevated temperatures. It can be easily construed that the temperature had a direct effect on the μ of the tribopair. At RT and 200 °C, the μ had a big fluctuation. At 400, 600, and 800 °C, the incipient μ had a big fluctuation and, then, it soon achieved a steady state with low μ values. Specifically, at RT, the μ showed a larger fluctuation than that at other temperatures. Particularly, at 600 °C, the μ exhibited a turbulent behavior in contrasted with that at other temperatures, which indicated an incipient short break-in period with relatively low μ values; then, it reached high μ values and fluctuated largely. Later, it approached a stable state with relatively low μ values. In comparison, the fluctuation of the μ decreased with the temperature increasing from RT to 800 °C. Figure 4b compared the μmean of TSC-PA/Inconel 718 tribopair at elevated temperatures. It can be seen that the μmean decreased with the temperature increasing from RT to 800 °C.
Figure 5a plotted the wear rates of the TSC-PA pin at elevated temperatures. The wear rates of the pin decreased with the temperature increasing from RT to 400 °C. Moreover, the wear rates of the pin were the same when the temperature was at 400, 600, and 800 °C. No other results were found in the literature for a comparison with the results in this work because Ti3SiC2 composites with the co-addition of PbO and Ag have not been reported before. Moreover, it was not easy to make a comparison between tribology results as they were carried out under distinct conditions on the basis of the detailed demands of an application. For interpretative aspirations, Table 2 compared tribological behavior of Ti3SiC2 composites with the addition of the different amount of Ag or the addition of the different amount of PbO. Comparatively, the incorporation of relative low content (5–20 vol%) of Ti3SiC2 can diminish the wear rate of the Ag-based composite during sliding against Al2O3 at 5N and 0.5 m/s [19]. In this work, the situation was different because the sample was Ti3SiC2 based-composite with 70 vol% of Ti3SiC2. At RT, comparatively, Ti3SiC2 and composites of Ti3SiC2 with 15 vol% Ag, with 15 vol% PbO, and with 15 vol% PbO-15 vol% Ag showed wear rates of 2 × 10−3 mm3/(N·m), 3 × 10−5 mm3/(N·m), 2 × 10−4 mm3/(N·m), and 2 × 10−4 mm3/(N·m), respectively, during sliding against Inconel 718 at 5N and 0.1 m/s [20,22]. These results showed that the incorporation of 15 vol% Ag or 15 vol% PbO or 15 vol% PbO-15 vol% Ag can decrease the wear rate of Ti3SiC2 by at least an order of magnitude at RT. At elevated temperatures (>200 °C), the composite of Ti3SiC2 with 15 vol% PbO showed better wear resistance than both Ti3SiC2 and the composite of Ti3SiC2 with 15 vol% PbO-15 vol% Ag (TSC-PA). The tribological behavior of the composite of Ti3SiC2 with 15 vol% Ag at elevated temperature was unavailable and incomparable. It can be seen from Table 2 that TSC-PA showed a lower friction than TSC at > 200 °C and a lower friction than Ti3SiC2-15 vol% PbO at <600 °C. It was concluded that the addition of PbO and Ag apparently improved the tribological behavior of TSC-PA, in comparison with TSC and Ti3SiC2-15 vol% PbO.
Figure 5b compared the wear rates of the Inconel 718 disk sliding against TSC-PA at elevated temperatures. As the temperature raised from RT to 800 °C, the wear rates of the disk decreased. At RT, a positive wear rate was found for the disk. Interestingly, negative wear rates were detected for the disk and increased with the temperature rising from 200 to 800 °C. Moreover, the wear rate of the disk at 600 °C was identical to that at 800 °C. The detection of Ti, Si, Pb, and Ag on the worn surface of Inconel 718 was in line with the negative wear of the disk (see Table 3).

3.4. Tribological Mechanisms

Figure 6 showed the SEM assessment of tribosurfaces of TSC-PA/Inconel 718 tribocouple after tribological testing at elevated temperature. TSC-PA was full of scars resulting from the abrasive wear, while the Inconel 718 surface was covered with an even tribolayer delivered from the TSC-PA surface. The surface was surrounded by the tribolayer made up of incompletely oxidized TSC-PA (Table 3 and Figure 6). With the rise in temperature, the tribofilm formed on the surface of TSC-PA was smoother and denser. This film was formed undoubtably by a coalescence of diverse mechanical blending, fracturing, resintering, and mechanically vitalized oxidation and other chemical processes, perhaps, continuously refined during the heat cycle. The EDS data showed that the elements of Ni, Cr, and Fe from Inconel 718 alloy appeared on the surface of TSC-PA (Figure 6a,c,e,g,i in Table 3) and the elements of Ti, Si, Pb, and Ag from TSC-PA appeared on the surface of Inconel718 alloy (Figure 6b,d,f,h,j in Table 3). Moreover, tribo-oxidation was a predominant factor that influenced, and even controlled, the wear mechanism and finally brought about the change of friction and wear features [31]. The chemical states of Ti, Si, Pb, and Ag elements on the worn surface of TSC-PA at RT–800 °C were investigated in detail by XPS (see Figure 7).
At RT, some wear debris dispersed on the worn surface of the TSC-PA with several dark continent regions (see Figure 6a). It was worth mentioning that some loose cracks and pores scattered on the worn surface, which was due to the peeling-off and decentralization of the granules. Such an episode was in accordance with Ti3SiC2 [6]. Therefore, it can be deduced that the addition of Pb and Ag did not affect the wear behavior of Ti3SiC2 at RT. Though the weak bonding between Ti3SiC2 grains in the TSC-PA led to the fracture and pulling out of Ti3SiC2 granules, soft metals (including Pb and Ag) distributed around the Ti3SiC2 grains, inhibiting the peeling-off and decentralization of Ti3SiC2 matrix; therefore, the wear rate of TSC-PA was only one tenth of Ti3SiC2 (see Table 2). A third body was formed and entrapped between the pin and disk, leading to the severe wear of TSC-PA and Inconel 718 at RT. The obvious evidence was that wear grooves were found on the worn surface of Inconel 718 after tribology testing at RT (see Figure 6b) and mutual material transfer between the pin and disk was detected by EDS analysis (see Table 3). The abrasive wear was the primary wear mechanism for TSC-PA. A small amount of SiO2 and PbO was detected on the worn surface of TSC-PA (see Figure 7). However, the temperature was too low to form an enough thick tribo-oxidation film between the contact surface of TSC-PA and Inconel 718, so high friction coefficients and high wear rates were present at RT (see Figure 4 and Figure 5a).
At 200 °C, some compacted wear debris was found on the worn surface of TSC-PA (see Figure 6c). Compared with RT, besides SiO2 and PbO, other oxides, TiO2 and Ag2O, were formed on the surface of TSC-PA (evidently owing to the XPS at 200 °C in Figure 7), generating a thicker tribo-oxidation film. The formation of the tribo-oxidation film did not inhibit the fracture and pulling-out of Ti3SiC2 grains. The formation and entrapment of the third body (the formed wear debris) between the pin and disk contributed to the wear of TSC-PA and Inconel 718. The mutual material transfer between the pin and disk at 200 °C was detected by EDS analysis (see Table 3). A tribo-oxidation film with some wear grooves was transferred to the Inconel 718, so the counterpart showed relatively flat region and slight negative wear (see Figure 5b and Figure 6d). At 200 °C, the abrasive wear mechanism was the main wear mechanism with very slight adhesive wear.
At 400 °C, obviously, the TSC-PA was covered with some large wear debris after tribology testing (see Figure 6e). It was expected that, initially, a tribo-oxidation film containing TiO2, SiO2, PbO and Ag2O was formed on its surface (as evident owing to the XPS at 400 °C in Figure 7), which was similar to the oxidation composition at 200 °C. Then, the peel-off and removal of the tribo-oxidation film occurred for TSC-PA. As the sliding finished, the mutual material transfer between the pin and disk was detected by EDS analysis (see Table 3). It was seen from Figure 6f that a relatively smooth film was built up on the surface of the Inconel 718. The plastic flow, in some way, retarded the abrasive wear of TSC-PA; therefore, the wear rate of TSC-PA at 400 °C exhibited a distinct decline in contrast to those at RT and 200 °C. Moreover, the addition of Ag and Pb could strengthen the ductility of TSC-PA. On other hand, when the asperities of the tribo-pair were in contact with each other, the instant flash temperature may reach 3000 °C [32]. Therefore, a high temperature was beneficial for the formation of oxides, such as TiO2, SiO2, PbO, and Ag2O. Therefore, the adhesive wear and plastic flow was the main wear mechanism with slight abrasive wear (as distinct on account of the slight wear trenches on the Inconel 718 disk).
At 600 °C, the worn surface of TSC-PA was surrounded by a relatively smooth and continuous tribolayer after tribology testing (see Figure 6g). This surface was covered with a sufficiently thick tribo-oxidation film containing TiO2, SiO2, PbO, and AgO (see Figure 7). In comparison with 400 °C, AgO was found on the worn surface of TSC-PA at 600 °C instead of Ag2O. Interestingly, the existence of AgO played a significant role in the decrease of friction and wear of the tribopair. The worn surface of Inconel 718 was embraced by a tribo-oxidation film, which was transferred from TSC-PA (see Figure 6h). The serious plastic flowing resulted in the adhesive wear. Mutual transfer of matter between the pin and the disk was verified by the EDS results in Table 3. Moreover, with the increase in temperature, especially at above 400 °C, oxygen contents of the pin and the disk were higher, which indicated the thicker thickness of the tribo-oxidation film (see Table 3). Therefore, the friction coefficient of TSC-PA was lower and more stable and its wear rate was low due to the moderately thick oxides. The wear mechanism was adhesive wear at this temperature.
At 800 °C, the worn surface of TSC-PA was surrounded by a tribo-oxidation film, which was smoothest and densest among all temperatures (see Figure 6). With the thickness of the tribofilm increasing, the covering of the Inconel 718 surface increased. Matter accumulation on the Inconel 718 led to seizure of the contact. The tribofilm reached a critical thickness so that it was easily peeled off from TSC-PA (see Figure 6i). The Inconel 718 showed a smooth tribofilm with large negative wear (see Figure 5b and Figure 6j). Of course, mutual transfer of matter between the pin and the disk was also confirmed by the EDS data in Table 3. The wear mechanism was adhesive wear and tribochemistry of elements of Pb, Ag, Ti, and Si to form TiO2, SiO2, PbO, and AgO (see Figure 7).
Conclusively at RT and 200 °C, the abrasive wear ascribed to the peeling-off and pulled-out of Ti3SiC2 grains dominated the wear mechanism of TSC-PA, leading to the indispensable friction and wear. As the temperature increased, the formation of tribo-oxidation films was favorable for the reduction in friction and wear of the tribopair. As for the reason why the Inconel 718 disk showed negative wear at elevated temperatures. We proposed that with the temperature increasing, especially at 800 °C, the formed oxidation film was so thick that it was easily peeled-off from the TSC-PA pin and transferred to the Inconel 718 disk, leading to the negative wear of the disk. Additionally, the transition of wear mechanism from abrasive wear to adhesive wear occurred with the increasing temperature. The co-addition of PbO and Ag worked together in the improvement of the tribological behavior of TSC-PA at elevated temperatures, which shed light on the effectiveness of co-addition of solid lubricants in MAX phases at elevated temperatures.

4. Conclusions

The Ti3SiC2-Pb-Ag self-lubricating composite (TSC-PA) was successfully sintered by spark plasma sintering technique in vacuum, and its friction and wear behavior were investigated by sliding it against Inconel 718 using a pin-on disk setup at RT and elevated temperatures in air. The results found that the TSC-PA was mainly composed of Ti3SiC2, Pb, and Ag. The mechanical properties of the TSC-PA were relatively weaker than that of TSC monolithic. As the temperature increased, the average friction coefficient of TSC-PA/Inconel 718 decreased. The wear rate of TSC-PA exhibited a downward trend and that of Inconel 718 decreased, too, with the rise in the temperature. Interestingly, the wear rate of Inconel 718 was positive wear only at RT and changed to negative wear at other temperatures. The wear mechanism was abrasive wear at low and medium temperatures and turned to adhesive wear at higher and elevated temperatures. The tribochemistry of the sliding contact plays an important part in the transition of the wear theories. It was an effective method of the co-addition of solid lubricants in MAX phases.

Author Contributions

Conceptualization, R.Z.; Data curation, H.Z.; Formal analysis, R.Z. and F.L.; Investigation, R.Z.; Methodology, R.Z. and H.Z.; Writing—original draft, R.Z. and F.L.; Writing—review and editing, R.Z. and F.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article.

Acknowledgments

The author acknowledges the financial support from the “Xinjiang Tianchi hundred Talents Program (Zhang Rui)”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barsoum, M.W. The MN+1AXN phases: A new class of solids: Thermodynamically stable nanolaminates. Prog. Solid State Chem. 2000, 28, 201–281. [Google Scholar] [CrossRef]
  2. Gupta, S.; Filimonov, D.; Zaitsev, V.; Palanisamy, T.; Barsoum, M. Ambient and 550°C tribological behavior of select MAX phases against Ni-based superalloys. Wear 2008, 264, 270–278. [Google Scholar] [CrossRef]
  3. Gupta, S.; Barsoum, M. On the tribology of the MAX phases and their composites during dry sliding: A review. Wear 2011, 271, 1878–1894. [Google Scholar] [CrossRef]
  4. El-Raghy, T.; Blau, P.; Barsoum, M.W. Effect of grain size on friction and wear behavior of Ti3SiC2. Wear 2000, 238, 125–130. [Google Scholar] [CrossRef]
  5. Hu, C.; Zhou, Y.; Bao, Y.; Wan, D. Tribological Properties of Polycrystalline Ti3SiC2 and Al2O3-Reinforced Ti3SiC2 Composites. J. Am. Ceram. Soc. 2006, 89, 3456–3461. [Google Scholar] [CrossRef]
  6. Ren, S.; Meng, J.; Lu, J.; Yang, S. Tribological Behavior of Ti3SiC2 Sliding Against Ni-based Alloys at Elevated Temperatures. Tribol. Lett. 2008, 31, 129–137. [Google Scholar] [CrossRef]
  7. Souchet, A.; Fontaine, J.; Belin, M.; Le Mogne, T.; Loubet, J.-L.; Barsoum, M.W. Tribological duality of Ti3SiC2. Tribol. Lett. 2005, 18, 341–352. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Ding, G.P.; Zhou, Y.C.; Cai, B.C. Ti3SiC2-a self-lubricating ceramic. Mater. Lett. 2002, 55, 285–289. [Google Scholar] [CrossRef]
  9. Zhai, H.; Huang, Z.; Zhou, Y.; Zhang, Z.; Wang, Y.; Ai, M. Oxidation layer in sliding friction surface of high-purity Ti3SiC2. J. Mater. Sci. 2004, 39, 6635–6637. [Google Scholar] [CrossRef]
  10. Ren, S.; Meng, J.; Lu, J.; Yang, S.; Wang, J. Tribo-physical and tribo-chemical aspects of WC-based cermet/Ti3SiC2 tribo-pair at elevated temperatures. Tribol. Int. 2010, 43, 512–517. [Google Scholar] [CrossRef]
  11. Dang, W.; Ren, S.; Zhou, J.; Yu, Y.; Wang, L. The tribological properties of Ti3SiC2/Cu/Al/SiC composite at elevated temperatures. Tribol. Int. 2016, 104, 294–302. [Google Scholar] [CrossRef]
  12. Yang, D.; Zhou, Y.; Yan, X.; Wang, H.; Zhou, X. Highly conductive wear resistant Cu/Ti3SiC2(TiC/SiC) co-continuous composites via vacuum infiltration process. J. Adv. Ceram. 2020, 9, 83–93. [Google Scholar] [CrossRef] [Green Version]
  13. Yang, J.; Gu, W.; Pan, L.M.; Song, K.; Chen, X.; Qiu, T. Friction and wear properties of in situ (TiB2+TiC)/Ti3SiC2 composites. Wear 2011, 271, 2940–2946. [Google Scholar] [CrossRef]
  14. Zhang, R.; Feng, W.; Liu, F. Tribo-oxide Competition and Oxide Layer Formation of Ti3SiC2/CaF2 Self-Lubricating Composites during the Friction Process in a Wide Temperature Range. Materials 2021, 14, 7466. [Google Scholar] [CrossRef] [PubMed]
  15. Islak, B.Y.; Candar, D. Synthesis and properties of TiB2/Ti3SiC2 composites. Ceram. Int. 2021, 47, 1439–1446. [Google Scholar] [CrossRef]
  16. Magnus, C.; Cooper, D.; Ma, L.; Rainforth, W.M. Microstructures and intrinsic lubricity of in situ Ti3SiC2–TiSi2–TiC MAX phase composite fabricated by reactive spark plasma sintering (SPS). Wear 2019, 448–449, 203169. [Google Scholar] [CrossRef]
  17. Yang, S.H.; Kong, H.; Yoon, E.-S.; Kim, D.E. A wear map of bearing steel lubricated by silver films. Wear 2003, 255, 883–892. [Google Scholar] [CrossRef]
  18. Akbulut, M.; Alig, A.R.G.; Israelachvili, J. Friction and tribochemical reactions occurring at shearing interfaces of nanothin silver films on various substrates. J. Chem. Phys. 2006, 124, 174703. [Google Scholar] [CrossRef]
  19. AlAnazi, F.; Ghosh, S.; Dunnigan, R.; Gupta, S. Synthesis and tribological behavior of novel Ag- and Bi-based composites reinforced with Ti3SiC2. Wear 2017, 376–377, 1074–1083. [Google Scholar] [CrossRef]
  20. Zeng, J.L.; Hai, W.X.; Meng, J.H.; Lu, J.J. Friction and wear of Ti3SiC2-Ag/Inconel 718 tribo-pair under a hemisphere-on-disk contact. Key Eng. Mater. 2014, 602–603, 507–510. [Google Scholar] [CrossRef]
  21. Zhang, R.; Feng, K.; Meng, J.; Su, B.; Ren, S.; Hai, W. Synthesis and characterization of spark plasma sintered Ti3SiC2/Pb composites. Ceram. Int. 2015, 41, 10380–10386. [Google Scholar] [CrossRef]
  22. Zhang, R.; Feng, K.; Meng, J.; Liu, F.; Ren, S.; Hai, W.; Zhang, A. Tribological behavior of Ti3SiC2 and Ti3SiC2/Pb composites sliding against Ni-based alloys at elevated temperatures. Ceram. Int. 2016, 42, 7107–7117. [Google Scholar] [CrossRef]
  23. Lu, J.R.; Zhou, Y.; Zheng, Y.; Li, H.Y.; Li, S.B. Interface structure and wetting behaviour of Cu/Ti3SiC2 system. Adv. Appl. Ceram. 2014, 114, 39–44. [Google Scholar] [CrossRef]
  24. Zhou, Y.; Gu, W. Chemical reaction and stability of Ti3SiC2 in Cu during high-temperature processing of Cu/Ti3SiC2 composites. Z. Metallkd. 2004, 95, 50–56. [Google Scholar] [CrossRef]
  25. Dang, W.; Ren, S.; Zhou, J.; Yu, Y.; Li, Z.; Wang, L. Influence of Cu on the mechanical and tribological properties of Ti3SiC2. Ceram. Int. 2016, 42, 9972–9980. [Google Scholar] [CrossRef]
  26. Yang, S.; Sun, Z.; Yang, Q.; Hashimoto, H. Effect of Al addition on the synthesis of Ti3SiC2 bulk material by pulse discharge sintering process. J. Eur. Ceram. Soc. 2007, 27, 4807–4812. [Google Scholar] [CrossRef]
  27. El-Raghy, T.; Barsoum, M.W.; Sika, M. Reaction of Al with Ti3SiC2 in the 800–1000 °C temperature range. Mater. Sci. Eng. A 2001, 298, 174–178. [Google Scholar] [CrossRef]
  28. Tzenov, N.; Barsoum, M.; El-Raghy, T. Influence of small amounts of Fe and V on the synthesis and stability of Ti3SiC2. J. Eur. Ceram. Soc. 2000, 20, 801–806. [Google Scholar] [CrossRef]
  29. Zhang, Z.F.; Sun, Z.M.; Zhang, H.; Hashimoto, H. Micron-scale deformation and damage mechanisms of Ti3SiC2 crystals induced by indentation. Adv. Eng. Mater. 2004, 6, 980–983. [Google Scholar] [CrossRef]
  30. El-Raghy, T.; Zavaliangos, A.; Barsoum, M.W.; Kalidindi, S.R. Damage Mechanisms around Hardness Indentations in Ti3SiC2. J. Am. Ceram. Soc. 2005, 80, 513–516. [Google Scholar] [CrossRef]
  31. Woydt, M.; Skopp, A.; Dörfel, I.; Witke, K. Wear engineering oxides/anti-wear oxides. Wear 1998, 218, 84–95. [Google Scholar] [CrossRef]
  32. Kalin, M. Influence of flash temperatures on the tribological behaviour in low-speed sliding: A review. Mater. Sci. Eng. A 2004, 374, 390–397. [Google Scholar] [CrossRef]
Figure 1. The XRD pattern of TSC-PA.
Figure 1. The XRD pattern of TSC-PA.
Materials 15 01437 g001
Figure 2. SEM micrograph (a) and distribution of Ti (b), Si (c), Pb (d), Ag (e) elements in the cross section of TSC-PA.
Figure 2. SEM micrograph (a) and distribution of Ti (b), Si (c), Pb (d), Ag (e) elements in the cross section of TSC-PA.
Materials 15 01437 g002
Figure 3. (a) Compressive stress–strain curve of TSC-PA and (b) SEM image of the fractured surface of TSC-PA after the three-point bending test.
Figure 3. (a) Compressive stress–strain curve of TSC-PA and (b) SEM image of the fractured surface of TSC-PA after the three-point bending test.
Materials 15 01437 g003
Figure 4. Plot of (a) friction coefficient (μ) versus distance and (b) the average friction coefficients of TSC-PA sliding against Inconel 718 alloys with different temperature.
Figure 4. Plot of (a) friction coefficient (μ) versus distance and (b) the average friction coefficients of TSC-PA sliding against Inconel 718 alloys with different temperature.
Materials 15 01437 g004
Figure 5. The wear rates of (a) pin and (b) disk for TSC—PA/Inconel 718 alloy tribo-pair as a function of temperature.
Figure 5. The wear rates of (a) pin and (b) disk for TSC—PA/Inconel 718 alloy tribo-pair as a function of temperature.
Materials 15 01437 g005
Figure 6. SEM micrographs of (a) TSC-PA at 25 °C, (b) Inconel 718 alloy at 25 °C, (c) TSC-PA at 200 °C, (d) Inconel 718 alloy at 200 °C, (e) TSC-PA at 400 °C, (f) Inconel 718 alloy at 400 °C, (g) TSC-PA at 600 °C, (h) Inconel 718 alloy at 600 °C, (i) TSC-PA at 800 °C, and (j) Inconel 718 alloy at 800 °C after tribological testing.
Figure 6. SEM micrographs of (a) TSC-PA at 25 °C, (b) Inconel 718 alloy at 25 °C, (c) TSC-PA at 200 °C, (d) Inconel 718 alloy at 200 °C, (e) TSC-PA at 400 °C, (f) Inconel 718 alloy at 400 °C, (g) TSC-PA at 600 °C, (h) Inconel 718 alloy at 600 °C, (i) TSC-PA at 800 °C, and (j) Inconel 718 alloy at 800 °C after tribological testing.
Materials 15 01437 g006
Figure 7. X-ray photoelectron spectroscopy for (a) Ti2p, (b) Si2pc, (c) Pb4f, (d) Ag3d, and (e) C1s on the worn surface of TSC-PA after sliding against Inconel 718 alloy with different temperature.
Figure 7. X-ray photoelectron spectroscopy for (a) Ti2p, (b) Si2pc, (c) Pb4f, (d) Ag3d, and (e) C1s on the worn surface of TSC-PA after sliding against Inconel 718 alloy with different temperature.
Materials 15 01437 g007
Table 1. Properties of TSC and TSC-PA.
Table 1. Properties of TSC and TSC-PA.
SampleNo.Relative Density
/%
Microhardness
/GPa
Flexural Strength
/MPa
Compressive Strength
/MPa
Ti3SiC2 (1250 °C)TSC98.245.5 ± 0.2428 ± 101230 ± 13
Ti3SiC2-PbO-AgTSC-PA95.352.24 ± 0.14311 ± 22654
Note: The data in the first line were from Reference [21].
Table 2. Comparison of tribological behavior of different Ti3SiC2-based composites.
Table 2. Comparison of tribological behavior of different Ti3SiC2-based composites.
CompositionCounter
-Surface
ConditionsWear rate (mm3/Nm)μReference
Ag-5 vol % Ti3SiC2Al2O3Block (Tab)-on-Disc, 5N, 0.5 m/s, air, RT3.2 × 10−50.38[19]
Ag-10 vol % Ti3SiC2Block (Tab)-on-Disc, 5N, 0.5 m/s, air, RT2.9 × 10−50.35
Ag-20 vol % Ti3SiC2Block (Tab)-on-Disc, 5N, 0.5 m/s, air, RT4.1 × 10−60.29
Ag-30 vol % Ti3SiC2Block (Tab)-on-Disc, 5N, 0.5 m/s, air, RT2.5 × 10−50.3
AgBlock (Tab)-on-Disc, 5N, 0.5 m/s, air, RT4.9 × 10−50.38
Ti3SiC2Inconel 718Pin-on-disk, 5N, 0.01 m/s, air, RT4.0 × 10−30.62[20]
Pin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−30.61
Pin-on-disk, 5N, 1 m/s, air, RT6.0 × 10−30.57
Ti3SiC2-5 vol% AgPin-on-disk, 5N, 0.01 m/s, air, RT7.0 × 10−50.42
Pin-on-disk, 5N, 0.1 m/s, air, RT8.0 × 10−50.51
Pin-on-disk, 5N, 1 m/s, air, RT5.0 × 10−30.56
Ti3SiC2-10 vol% AgPin-on-disk, 5N, 0.01 m/s, air, RT4.0 × 10−50.4
Pin-on-disk, 5N, 0.1 m/s, air, RT7.0 × 10−60.51
Pin-on-disk, 5N, 1 m/s, air, RT3.0 × 10−30.5
Ti3SiC2-15 vol% AgPin-on-disk, 5N, 0.01 m/s, air, RT5.0 × 10−50.4
Pin-on-disk, 5N, 0.1 m/s, air, RT3.0 × 10−50.5
Pin-on-disk, 5N, 1 m/s, air, RT2.0 × 10−50.45
Ti3SiC2-20 vol% AgPin-on-disk, 5N, 0.01 m/s, air, RT2.0 × 10−50.31
Pin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−40.45
Pin-on-disk, 5N, 1 m/s, air, RT1.0 × 10−50.4
Ti3SiC2Inconel 718 diskPin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−30.65[22]
Pin-on-disk, 5N, 0.1 m/s, air, 200 °C3.0 × 10−40.65
Pin-on-disk, 5N, 0.1 m/s, air, 400 °C5.0 × 10−50.62
Pin-on-disk, 5N, 0.1 m/s, air, 600 °C2.5 × 10−50.65
Pin-on-disk, 5N, 0.1 m/s, air, 800 °C5.0 × 10−60.4
Ti3SiC2-5 vol% PbOInconel 718 diskPin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−30.72
Pin-on-disk, 5N, 0.1 m/s, air, 200 °C1.5 × 10−30.66
Pin-on-disk, 5N, 0.1 m/s, air, 400 °C9.0 × 10−50.65
Pin-on-disk, 5N, 0.1 m/s, air, 600 °C4.0 × 10−50.58
Pin-on-disk, 5N, 0.1 m/s, air, 800 °C7.0 × 10−60.45
Ti3SiC2-10 vol% PbOInconel 718 diskPin-on-disk, 5N, 0.1 m/s, air, RT6.0 × 10−40.67
Pin-on-disk, 5N, 0.1 m/s, air, 200 °C6.0 × 10−40.65
Pin-on-disk, 5N, 0.1 m/s, air, 400 °C3.0 × 10−50.66
Pin-on-disk, 5N, 0.1 m/s, air, 600 °C1.0 × 10−50.46
Pin-on-disk, 5N, 0.1 m/s, air, 800 °C5.0 × 10−60.3
Ti3SiC2-15 vol% PbOInconel 718 diskPin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−40.71
Pin-on-disk, 5N, 0.1 m/s, air, 200 °C2.0 × 10−40.75
Pin-on-disk, 5N, 0.1 m/s, air, 400 °C1.0 × 10−50.65
Pin-on-disk, 5N, 0.1 m/s, air, 600 °C5.0 × 10−60.35
Pin-on-disk, 5N, 0.1 m/s, air, 800 °C2.0 × 10−60.2
Ti3SiC2-15 vol%PbO-15 vol % AgInconel 718 diskPin-on-disk, 5N, 0.1 m/s, air, RT2.0 × 10−40.72This work
Pin-on-disk, 5N, 0.1 m/s, air, 200 °C1.5 × 10−40.67
Pin-on-disk, 5N, 0.1 m/s, air, 400 °C6.0 × 10−50.56
Pin-on-disk, 5N, 0.1 m/s, air, 600 °C6.0 × 10−50.51
Pin-on-disk, 5N, 0.1 m/s, air, 800 °C6.0 × 10−50.3
Table 3. The chemical components of samples corresponding to Figure 6 as determined by EDS.
Table 3. The chemical components of samples corresponding to Figure 6 as determined by EDS.
PositionsSamplesTemperature/°CAtomic Percentage/at. %
Figure 5aTSC-PA pin2514.9%Ti, 5.2%Si, 29.9%C, 1.4%Al, 1.9%Ag, 42.8%O, 1.0%Pb, 1.3%Ni, 0.7%Cr, 0.9%Fe
Figure 5bInconel 718 disk252.4%Ti, 1.3%Si, 28.8%C, 0.9%Al, 0.4%Ag, 25.4%O, 21.0%Ni, 9.4%Cr, 8.4%Fe, 1.3%Nb, 0.7%S
Figure 6cTSC-PA pin20017.3%Ti, 6.3%Si, 31.2%C, 1.3%Al, 1.6%Ag, 39.5%O, 1.0%Pb, 0.8%Ni, 0.5%Cr, 0.5%Fe
Figure 6dInconel 718 disk2006.5%Ti, 3.3%Si, 31.7%C, 1.4%Al, 0.9%Ag, 33.6%O, 0.5%Pb, 11.1%Ni, 5.3%Cr, 4.6%Fe, 0.7%Nb, 0.4%S
Figure 6eTSC-PA pin40020.5%Ti, 6.1%Si, 23.4%C, 2.0%Al, 2.0%Ag, 42.0%O, 1.2%Pb, 1.5%Ni, 0.6%Cr, 0.7%Fe
Figure 6fInconel 718 disk40013.0%Ti, 5.7%Si, 18.2%C, 1.9%Al, 2.0%Ag, 46.4%O, 1.0%Pb, 6.1%Ni, 3.0%Cr, 2.7%Fe
Figure 6gTSC-PA pin6009.5%Ti, 3.0%Si, 11.0%C, 0.8%Al, 0.5%Ag, 61.0%O, 0.4%Pb, 7.4%Ni, 3.2%Cr, 2.7%Fe, 0.5%Nb
Figure 6hInconel 718 disk6002.3%Ti, 0.7%Si, 15.4%C, 0.5%Al, 45.3%O, 1.1%Pb, 17.9%Ni, 7.8%Cr, 6.9%Fe, 1.3%Nb, 0.8%S
Figure 6iTSC-PA pin8009.7%Ti, 2.9%Si, 10.7%C, 0.9%Al, 0.7%Ag, 59.7%O, 0.6%Pb, 7.8%Ni, 3.5%Cr, 2.9%Fe, 0.6%Nb
Figure 6jInconel 718 disk80011.9%Ti, 5.9%Si, 18.7%C, 1.7%Al, 2.3%Ag, 46.9%O, 0.9%Pb, 5.9%Ni, 3.2%Cr, 2.6%Fe
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, R.; Zhang, H.; Liu, F. Microstructure and Tribological Properties of Spark-Plasma-Sintered Ti3SiC2-Pb-Ag Composites at Elevated Temperatures. Materials 2022, 15, 1437. https://doi.org/10.3390/ma15041437

AMA Style

Zhang R, Zhang H, Liu F. Microstructure and Tribological Properties of Spark-Plasma-Sintered Ti3SiC2-Pb-Ag Composites at Elevated Temperatures. Materials. 2022; 15(4):1437. https://doi.org/10.3390/ma15041437

Chicago/Turabian Style

Zhang, Rui, Huiming Zhang, and Fuyan Liu. 2022. "Microstructure and Tribological Properties of Spark-Plasma-Sintered Ti3SiC2-Pb-Ag Composites at Elevated Temperatures" Materials 15, no. 4: 1437. https://doi.org/10.3390/ma15041437

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