Previous Article in Journal
First-Principles Linear Combination of Atomic Orbitals Calculations of K2SiF6 Crystal: Structural, Electronic, Elastic, Vibrational and Dielectric Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sintering and Tribological Properties of Ti3SiC2-TiSix Composite Sintered by High-Pressure High-Temperature Technology

1
Intelligent Manufacturing and Electrical Engineering, Nanyang Normal University, Nanyang 473061, China
2
Henan Academy of Science, Zhengzhou 450002, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(19), 4866; https://doi.org/10.3390/ma17194866 (registering DOI)
Submission received: 27 August 2024 / Revised: 21 September 2024 / Accepted: 1 October 2024 / Published: 3 October 2024
(This article belongs to the Section Materials Physics)

Abstract

:
The Ti3SiC2TiSix ceramic composite was synthesized in situ from a mixture of 3Ti:1.5Si:1.2C powders under pressures ranging from 2 to 5 GPa and temperatures of 1150 °C to 1400 °C. At medium and high temperatures (4–5 GPa and 1400 °C), Ti3SiC2 dissolves into the cubic TiC phase. SEM analysis revealed that the high-pressure-produced multilayer structure of Ti3SiC2 remained intact. The friction properties of Ti3SiC2-TiSix composites combined with copper and aluminum were studied under both dry and lubricated conditions. After the break-in period, the Ti3SiC2-TiSix/Al combination exhibited the lowest friction coefficient: approximately 0.2. In dry-sliding conditions, the friction coefficient varies between 0.5 and 0.8. The wear mechanisms for Ti3SiC2-TiSix composites paired with aluminum primarily involve pear groove wear and adhesive wear during dry friction. Irregularly shaped aluminum balls accumulate in the pear grooves and adhere to each other. With increasing sintering pressure, the average friction coefficient of Ti3SiC2-TiSix composites against Cu ball pairs first increases and then decreases. The wear rate of the samples did not vary significantly as the sintering pressure increased, whereas the wear rate of Cu balls decreased with increasing sintering pressure. The adhesive wear of the Ti3SiC2-TiSix composite with its Cu counterpart is stronger than that of the Al counterpart. Abrasive chips of Cu balls appeared in flake form and adhered to the contact interface.

1. Introduction

The ternary lamellar ceramic Ti3SiC2 offers a remarkable combination of the superior properties found in both ceramics and metals [1]. As a result, it has wide-ranging applications, including use as a binder for super-hard materials [2,3,4,5], in copper matrix composites [6,7,8], and in ceramic composites [9,10,11,12,13,14,15]. The unique properties of Ti3SiC2 have spurred significant research into its preparation and application, with a focus on understanding the fundamental principles of the microstructure–property relationship [16,17]. The properties are highly dependent on the microstructures, which are influenced by the synthesis method. The most commonly used method is hot pressing, applying approximately 35 MPa to produce dense Ti3SiC2 solids [9].
Spark plasma sintering (SPS) at pressures ranging from 50 MPa to 80 MPa [11] and hot isostatic pressing (HIP) at approximately 200 MPa have been shown to enhance the microstructure and performance of Ti3SiC2. High sintering pressure, on the order of several GPa, facilitates the synthesis of dense Ti3SiC2, particularly in super-hard composites. Some studies have successfully synthesized diamond or cBN composites with Ti3SiC2 and related MAX-phase compounds at pressures exceeding 3 GPa [2,3,4,5,18]. The ternary Ti–Si–C alloy phase diagram has been investigated for the synthesis of Ti3SiC2 at temperatures ranging from 1250 to 2877 °C. However, there are conflicting conclusions regarding the synthesis and stable phase region of Ti3SiC2 under high pressure. Qin et al. [19] reported that Ti3SiC2 powders become unstable between 3 and 5 GPa, with Ti3SiC2 breaking down into TiC at temperatures above 1000 °C under 3 GPa. The critical disintegration pressure and temperature change linearly. Meng et al. [20] investigated the formation mechanism of Ti3SiC2 with different raw material types (Si/SiC or TiC/C) and proportions. Excess silicon is beneficial for the synthesis of high-purity Ti3SiC2 at 1300 °C. Depending on the ratio of starting materials, reactions in the Ti-Si-C system may yield TiC, SiC, TiS2, Ti5Si3C, or Ti3SiC2 [21]. The influence of sintering pressure on the formation mechanism is essential for further study. TiSi2, as an intermediate phase in the synthesis of Ti3SiC2, indirectly reflects how different temperatures and pressures affect the stability of Ti3SiC2.
To prevent the decomposition of Ti3SiC2 caused by the diffusion of silicon atomic layers under high pressure, Zhu et al. [3] employed a high-pressure, high-temperature sintering process to synthesize polycrystalline diamond, using Ti3SiC2 and Si as the binder. The Ti3SiC2-Si binder composite began to decompose into TiC and SiC at temperatures above 1350 °C under 5.5 GPa. Among the Ti3SiC2-based composite materials, most studies have focused on Ti3SiC2-TiC [22] or Ti3SiC2-SiC [9,10,12] composites, with only limited research on Ti3SiC2-TiSix composites. TiSi2 is an excellent reinforcement material due to its high melting point (1540 °C), oxidation resistance, and mechanical stability [23]. Additionally, it is widely used in the semiconductor industry for its outstanding electrical properties [24,25].
The study of TiSi2 matrix composites prepared under high pressure provides indirect insights into the stability of Ti3SiC2. The mechanism of Ti3SiC2 synthesis under high pressure, in the presence of abundant titanium and silicon, was investigated. The effect of raw material composition (3Ti:1.5Si:1.2C) was explored in comparison to previous studies. Additionally, the friction properties of Ti3SiC2-TiSix composites synthesized under varying pressures were examined.

2. Materials and Methods

To synthesize Ti3SiC2-TiSix composites, titanium (Ti, 325 mesh, 99.3 wt.% purity), silicon (Si, 325 mesh, >99.7 wt.% purity), and graphite (325 mesh, 99% purity) powders were measured in a molar ratio of 3:1.5:1.2. The powders were then mixed in a mixer at 300 rpm/min for 12 hrs. The mixture was compacted into tablets with a diameter of 14 mm and a height of 5 mm using a cemented carbide die.
Ti3SiC2-TiSix samples were prepared using a high-pressure sintering (HPS) apparatus (SPD 6× 1200, Xianyang Superhard Materials Equipment (Group) Co., Ltd., Xian, China) at temperatures ranging from 1150 °C to 1400 °C and pressures between 1 and 5 GPa. The temperature and pressure were increased over a period of approximately 3 min. Both were maintained for a predetermined holding time. Upon completion of the holding time, the power was immediately cut off, and the pressure was gradually released over a period of about 15 min. The sintering process followed the procedure described in previous studies [4,5]. A schematic representation of the high-pressure and high-temperature sintering process is provided in Figure 1.
The composition of the sintered compacts was analyzed using X-ray diffraction (XRD, Brukeraxs Co., Karlsruhe, Germany). The X-ray utilized a copper (Cu) target, with a loading voltage of 40 kV and a current of 40 mA. The cross-section morphology and microstructure of sintered Ti3SiC2-TiSix composites was examined using scanning electron microscopy (SEM, JSM-6390LV, JEOL, Tokyo, Japan).
The friction and wear tests of sintered Ti3SiC2-TiSix bulks were conducted using the pin-on-disk-type CFT-I material surface performance tester (Zhongke Kaihua Instrument Equipment Co., Ltd., Beijing, China). The surface of the Ti3SiC2-TiSix sample was treated with sandpaper prior to the test, achieving a surface roughness of approximately 0.1 mm. Small differences in surface roughness or variations in material composition can lead to discrepancies in friction and wear results. Typically, the surface roughness and compositional homogeneity of samples synthesized at high temperatures and pressures are consistent. The Ti3SiC2-TiSix sample was secured using hollow disk bolts under compression. The counter-abrasives were aluminum or copper balls with a diameter of Φ 4 mm. The test balls and samples were cleaned ultrasonically with alcohol before testing. The dry friction and wet grinding reciprocating sliding tests were conducted at room temperature with a sliding distance of 5 mm, a drive motor speed of 300 rpm/min, a load of 12 N, and a test duration of 30 min. The dynamic real-time friction coefficient was automatically recorded by the computer during the test. Fluctuations in the load force were due to vibrations in the drive mechanism. The variation in the error between the actual value and the set value of the load with time is illustrated in Supplementary Figure S1.

3. Results

3.1. XRD Results of Ti3SiC2-TiSix Composite

Figure 2 presents the XRD patterns of samples synthesized from mixtures containing a molar ratio of 3Ti/1.5Si/1.2C under pressures ranging from 1 to 5 GPa for 30 min at 1150 °C. At 1 GPa, distinct Ti3SiC2 peaks are visible, as shown in Figure 2. With an increase in synthetic pressure to 2–3 GPa, the prominent peaks of Ti3SiC2 and the intermediate phase Ti5Si3 diminished. Concurrently, the TiSi2 phase and residual carbon formed. At 3.5 GPa, the intensity of the distinctive Ti5Si3 distinctive peaks was at its lowest, and nearly vanished. Interestingly, Ti5Si3 peaks increased when the synthesizing pressure rose to 4 GPa, then decreased again as the pressure increased to 5 GPa. Between 2 GPa and 5 GPa, TiSi2 became the predominant phase, with its characteristic peaks shifting towards smaller angles. TiSi2 crystallizes in an orthorhombic structure with Cmcm (C49) and Fddd (C54) [26]. The TiSi2 phase was stable, exhibiting a contraction of the lattice constant of about 0.01 during pressurization from 0 to 5 GPa [27]. There is the possibility of the formation of solid solutions of orthorhombic TiSi2; however, hexagonal TiSi2 was not observed. This result is analogous to the reaction where TiC starts to react with Si, forming the TiSi2 phase at 1150 °C under 2 GPa [28].
Figure 3 presents the XRD patterns of Ti3SiC2-TiSix composites synthesized from a molar ratio of 3Ti/1.5Si/1.2C at various pressures for 30 min at 1250 °C. Ti3SiC2 is the dominant phase; however, Ti5Si3 and TiSi2 phases coexisted under 3 GPa. This contrasts with the pulse discharge sintering conducted at a pressure of 50 MPa. The synthesis of Ti3SiC2 was achieved during sintering at temperatures of 1250 °C and higher using pulse discharge sintering [29]. At pressures exceeding 3 GPa, the sintering pressure had no impact on the chemical composition, as shown in Figure 3. However, upon increasing the pressure to 4 GPa, the characteristic peak of Ti3SiC2 (lattice plane 008) diminished. Interestingly, while the TiSi2 phase decreased, the Ti3SiC2 peaks increased at 4.5 GPa; the situation differed significantly at 5 GPa. The primary component synthesized at pressures ranging from 3 to 5 GPa and temperatures of 1150 °C or 1250 °C is TiSi2, indirectly indicating that this temperature–pressure interval is more suitable for the high-pressure preparation of TiSi2.
Figure 4 presents the XRD patterns of Ti3SiC2-TiSix composites fabricated from a mixture with a molar ratio of 3Ti/1.5Si/1.2C at 1400 °C under pressures of 4 to 5 GPa for 30 min. The predominant phase is Ti3SiC2, which coexists with trace amounts of TiS2 and Ti5Si3. The excessively high sintering pressure and temperature led to the emergence of ZrO2 peaks. The sintered results are comparable to those synthesized at 1150 °C under 1 GPa and 1250 °C under 3 GPa. Ti3SiC2 completely decomposes at 5 GPa and 1300 °C [19]. The enhanced stability of Ti3SiC2 can be explained in two ways: 1) the Si content of the starting material is greater than the stoichiometric ratio of Ti3SiC2; 2) the secondary-phase TiSx exhibits high hardness (8.7 GPa for TiSi2 and 9.8 GPa for Ti5Si3) [30] and high Young’s modulus (256 GPa for TiSi2 and 156 GPa for Ti5Si3) [31]. The excess Si enhances the stability of the Ti3SiC2 phase under high pressure, similar to the preparation of the Ti3SiC2 phase through chemical vapor deposition (CVD) [32], arc melting [33], self-propagating high-temperature synthesis (SHS) [34], pressure-less synthesis [35], reactive melt infiltration (RMI) [36], and pulse discharge sintering (PDS) processes [37]. The excess silicon likely compensates for losses due to evaporation. Currently, there is no thermodynamic phase diagram available for the stability of the high-pressure Ti3SiC2 phase. Phase diagrams of Ti-Si-C at 1200 °C under atmospheric pressure indicate that all samples with excess silicon have comparable and relatively low amounts of TiSi2 [38]. When the synthesis temperature exceeds 1300 °C, silicon can form a low melting point eutectic with the TiSi2 alloy, as depicted in the calculated Ti-Si-C ternary phase diagram at 1400 °C and 1800 °C, respectively [39]. The hard-phase TiSi2 and Ti5Si3 compartmentalize the Ti3SiC2 into hermetically sealed units at high pressure, thereby inhibiting Si diffusion escape. Under high-pressure conditions, these hard phases (such as diamond or cBN) act to compartmentalize the structure, creating barriers to Si atom diffusion. Referring to the study of Ti3SiC2 as a superhard material binder, the graphical abstract in Ref. [2], Figure 4b in Ref. [40], and Figure 15 in Ref. [18] illustrate the mechanism of sealing the bonded phase with the hard phase under high pressure.

3.2. Microstructure of Ti3SiC2-TiSix Composites

Figure 5 presents fracture surface micrographs of sintering products fabricated from 3Ti/1.5Si/1.2C under pressure at temperatures ranging from 1150 to 1400 °C for 30 min. The sample consists of Ti3SiC2, TiSi2, Ti5Si3, and graphite, as shown in Figure 5a. The high-pressure sintered sample in Figure 5b is nearly fully dense and exhibits the characteristic layered structure of Ti3SiC2. As deduced from Figure 2, Ti3SiC2 can be synthesized at this temperature and pressure under 1 GPa. The Ti3SiC2-TiSix composites demonstrate good interfacial bonding, akin to cBN-Ti3AlC2 composites [4] and cBN-Ti3SiC2 composites [5]. The presence of Ti3SiC2 is further confirmed by the typical layered structure observed in Figure 5b.
Figure 5c displays a low-magnification image of the sample sintered at 1150 °C under 3 GPa for 30 min. The primary phases observed include orthorhombic TiSi2 and layered Ti3SiC2, highlighted in the red rectangle, which aligns with the XRD result in Figure 2. TiSi2 adopts an orthorhombic structure, and a potential phase change under pressure during high-temperature high-pressure sintering could result in the formation of microcracks observed during the friction test. High-magnification images of the layered Ti3SiC2 structure and orthorhombic TiSi2 are provided in Figure 5d,e, respectively. The interaction between TiSi2 particles and the stacked Ti3SiC2 layers is depicted in Figure 5f.

3.3. Friction Behavior of Ti3SiC2-TiSix Composites

Figure 6 presents the friction coefficient (COF) of Ti3SiC2-TiSix composites versus sliding time. The frictional behavior is highly sensitive to the test conditions. The average COF of Ti3SiC2-TiSix composites sliding against an Al ball was measured with a normal load of 12 N.
Following the break-in period (150 s), the Ti3SiC2-TiSix/Al pair exhibited the lowest friction coefficient (approximately 0.2), with minimal fluctuations during the stable period. These slight fluctuations in COF during the steady period can be attributed to the plastic deformation of stressed surfaces, and the reduction in stiffness during the wet-sliding test [41]. Notably, the COF remained unaffected by variations in sintering pressure.
The COF increases significantly under dry-sliding conditions compared to wet-sliding conditions. As the sintering pressure increases, the COF initially decreases and then rises. The COFs of Ti3SiC2-TiSix composites sintered at 2 GPa and 4.5 GPa are 0.7316 and 0.6437, respectively. These COF values for Ti3SiC2-TiSix composites are similar to those of Ti3SiC2-PbO-Ag composites tested against the Inconel 78 alloy [42]. Relatively significant fluctuations were observed during the sliding phase of both test pairings. The Ti3SiC2-TiSix/Al pairing exhibited a COF of approximately 0.5073. The COF ranges from 0.5 to 0.73, depending on the chemical composition of the Ti3SiC2-TiSix composites. This suggests that, under similar testing conditions, the COF is highly sensitive to the material composition.
The wear rates of Ti3SiC2-TiSix composites sliding against an Al ball are shown in Figure 7. Ti3SiC2-TiSix composites exhibit a higher wear rate under dry-sliding conditions. The Ti3SiC2-TiSix composites sintered at 4 GPa exhibit the highest wear rate. The wear rate of Ti3SiC2-TiSix composites initially increases and then decreases as the sintering pressure increases.
Figure 8 displays SEM images of wear tracks on virgin Ti3SiC2-TiSix composites sliding against an Al ball, as examined under an optical microscope. The variation in the scar area correlates with the changes in COF, as supported by optical microscope images. Under dry-sliding situations, the scar diameter of the Al ball increased from 1.966 mm to 2.536 mm before decreasing to 1.606 mm. Optical microscope images of Ti3SiC2-TiSix composites with Al balls are shown in Figure S2. The primary wear mechanisms of Ti3SiC2-TiSix composites against their Al counterparts involve pear groove wear and adhesive wear during dry sliding. The irregularly shaped Al chips accumulate in the pear grooves and adhere to one another. During wet sliding, pear groove wear predominates, while adhesive wear is significantly reduced. The abrasive chips are encapsulated by the lubricating fluid, leading to agglomeration within the medium.
Figure 9 presents the typical measurement curves of the dynamic COF of Ti3SiC2-TiSix composites against Cu ball pairs under a 12 N load and a sliding speed of 0.05 m/s over a 30 min test duration. Following a 5 min running-in period, the COF of Ti3SiC2-TiSix composites sintered at 2 GPa and 4.5 GPa against the Cu ball pairs ranged from 0.5 to 0.7. In contrast, the Ti3SiC2-TiSix composites sintered at 4 GPa displayed a higher COF of 0.77, along with random fluctuation behavior. The COFs of Ti3SiC2-TiSix composites against Cu balls are lower than those of Ti3SiC2-ZnO composites tested against Inconel 78 alloy [15].
In the context of dry-sliding friction, researchers have identified two primary behaviors; transition behavior, where the COF shifts from an initial value to a steady state during the early phase of continuous friction, and random fluctuation behavior, characterized by turbulent variations in the entire friction process [43]. Different counterfaces exhibit varying frictional behavior under the same conditions, as compared to the Ti3SiC2-TiSix/Al ball tribocouple. This suggests that the counterparts play a crucial role in the tribological performance of Ti3SiC2-TiSix composites.
Figure 10 illustrates the wear rates of Ti3SiC2-TiSix composites against Cu ball pairs at a load of 12 N and a sliding speed of 0.05 m/s. The wear rate of Ti3SiC2-TiSix composites decreases with increasing sintering pressure. The wear rates of the Cu balls exhibited a similar trend.
Figure 11 presents SEM images of wear tracks from virgin Ti3SiC2-TiSix composites sliding against Cu balls. As the sintering pressure rises, the scar length diminishes. The wear mechanism of the Ti3SiC2-TiSix composite in contact with its Cu counterpart includes pear groove wear and adhesive wear, with adhesive wear being more pronounced compared to the Al counterpart. The abrasive chips from the Cu balls are flakes that adhere to the contact interface. Optical microscope images of Ti3SiC2-TiSix composites paired with Cu balls are shown in Figure S3.

4. Discussion

Ti3SiC2 can be synthesized from Ti, Si/SiC, and graphite at 4 GPa and 1100 °C for varying soaking durations, as indicated by previous research [44]. At 2.0 GPa, the synthesis and formation mechanisms of Ti3SiC2 were investigated using the reactant species Ti/Si/C, Ti/SiC/TiC, Ti/SiC/C, Ti/SiC/C, and Ti/TiC/Si [20]. Ti3SiC2 forms at 1050 °C and 4.5 GPa, similar to the constituent combinations of Ti/Si/C/cBN, Ti/SiC/TiC/cBN, Ti/SiC/C/cBN, and Ti/TiC/Si/cBN [5]. If the initial elements are weighted according to the stoichiometric ratio, the presence of the impurity TiC becomes inevitable, as indicated by these findings [5,20,44]. A deficiency of Si promotes the formation of TiC, while an excess of Si favors the formation of TiSi2.
In Figure 12, when Ti3SiC2 is employed as the raw material, the dotted line on the left divides the stable zone at high pressure. Ti3SiC2 decomposes into TiC as temperature and pressure rise. Ti3SiC2 composites can be synthesized at high pressures of 4.5 GPa or 5.5 GPa by combining titanium powder, silicon powder, carbon powder, or aluminum powder with a hard phase (such as diamond or cubic boron nitride). TiSi2 is used as the secondary phase enhancement in this study. TiSi2 serves as an intermediate in the synthesis of Ti3SiC2, helping to inhibit its decomposition. Additionally, Ti3SiC2-TiSix can be synthesized over a broader range of temperatures (1150–1250 °C) and pressures (3–5 GPa) compared to silicon enrichment. The presence of the second phase can enhance the stability of Ti3SiC2 by expanding its stable region at high pressure. The recommended parameters for the high-pressure synthesis of Ti3SiC2-TiSix are 4–5 GPa and 1400 °C. Considering the application purposes of previous studies on the superhard material binder, the excessive addition of too much Si to the Ti3SiC2 binder or Al to Ti3AlC2 may result in the formation of secondary phase TiSix alloys or TiAlx alloys. It is meaningful to synthesize TiSix-Ti3SiC2 as a superhard material bonding agent by varying the material ratios.

5. Conclusions

Ti3SiC2-TiSix composites were synthesized by a high-pressure and high-temperature method. The phase evolution of cermet Ti3SiC2 powder was studied under high pressure (1–5 GPa) and high temperature (1150–1400 °C). At 1 GPa and 1150 °C, high-content Ti3SiC2 is synthesized. Above 2 GPa, it transforms to TiSi2. In the middle-temperature zone (1250 °C), TiSix content increases with sintering pressure. The friction and wear properties of high-pressure synthetic Ti3SiC2 with aluminum and copper were investigated under normal temperature and dry/wet conditions. Under wet friction, the friction coefficient of Ti3SiC2-TiSix and Al balls is about 0.2. Under dry friction, it is comparable to that of Cu balls. Future research will focus on the high-temperature tribological properties and mechanisms of Ti3SiC2-TiSix composites with different ceramic counterparts for high-temperature composite ceramic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17194866/s1, Figure S1: Variation in the error between the actual load value and the set value over time; Figure S2: Optical microscope images of Ti3SiC2-TiSix composites sliding against Al balls; Figure S3: Optical microscope images of Ti3SiC2-TiSix composites sliding against Cu balls.

Author Contributions

Conceptualization, Y.C. and L.L.; methodology, J.L. and L.L.; data curation, J.L. and L.L.; writing—original draft preparation, Y.C., J.L. and L.L.; writing—review and editing, Y.C., C.S. and L.L. project administration, Y.C., M.H. and L.L.; funding acquisition, Y.C., M.H. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Open Cooperation Project of Henan Academy of Science, grant number 220909003; the Youth Fund Project of the Natural Science Foundation of Henan Province, grant number 242300421464; the Doctoral Special Fund Project of Nanyang Normal University, grant numbers 2024ZX025 and 2019ZX018; and the Cultivation Project of the National Natural Science Foundation of Nanyang Normal University, grant numbers 2024PY023 and 2023PY011.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available on reasonable request from the corresponding authors, Yuqi Chen and Liang Li.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, H.B.; Bao, Y.W.; Zhou, Y.C. Current status in layered ternary carbide Ti3SiC2, a review. J. Mater. Sci. Technol 2009, 25, 1–38. [Google Scholar]
  2. Ji, H.; Liang, Y.; Jiang, Z.; Li, Z.; Zhu, Y. Controllable HTHP sintering and property of cBN/diamond composites containing Ti3SiC2. Ceram. Int. 2020, 46, 13807–13812. [Google Scholar] [CrossRef]
  3. Lv, X.; Jian, Q.; Li, Z.; Sun, K.; Ji, H.; Zhu, Y. Effect of controllable decomposition of MAX phase (Ti3SiC2) on mechanical properties of rapidly sintered polycrystalline diamond by HPHT. Ceram. Int. 2019, 45, 16564–16568. [Google Scholar] [CrossRef]
  4. Li, L.; Zhou, A.; Wang, L.; Li, S.; Wu, D.; Yan, C. In situ synthesis of cBN–Ti3AlC2 composites by high-pressure and high-temperature technology. Diam. Relat. Mater. 2012, 29, 8–12. [Google Scholar] [CrossRef]
  5. Li, Z.; Zhou, A.; Li, L.; Wang, L.; Hu, M.; Li, S.; Gupta, S. Synthesis and characterization of novel Ti3SiC2–cBN composites. Diam. Relat. Mater. 2014, 43, 29–33. [Google Scholar] [CrossRef]
  6. Wu, Z.; Jiang, X.; Li, Y.; Christian, P.; Sun, H.; Zhang, Y.; Fang, Y.; Shu, R. Microstructures and properties of graphene nanoplatelets reinforced Cu/Ti3SiC2/C nanocomposites with efficient dispersion and strengthening achieved by high-pressure torsion. Mater. Charact. 2022, 193, 112308. [Google Scholar] [CrossRef]
  7. Zhang, R.; Chen, B.; Liu, F.; Sun, M.; Zhang, H.; Wu, C. Microstructure and mechanical properties of composites obtained by spark plasma sintering of Ti3SiC2-15 vol.% Cu mixtures. Materials 2022, 15, 2515. [Google Scholar] [CrossRef]
  8. Zhang, R.; Du, C.; Liu, F.; Wu, C. Electrochemical Corrosion Behavior and the Related Mechanism of Ti3SiC2/Cu Composites in a Strong Acid Environment. Materials 2024, 17, 4035. [Google Scholar] [CrossRef]
  9. Amiri, S.H.; Kakroudi, M.G.; Vafa, N.P.; Asl, M.S. Synthesis and sintering of Ti3SiC2–SiC composites through reactive hot-pressing of TiC and Si precursors. Silicon 2021, 14, 4227–4235. [Google Scholar] [CrossRef]
  10. He, G.; Xu, J.; Zhang, Z.; Qian, Y.; Zuo, J.; Li, M.; Liu, C. Interfacial reactions and mechanical properties of SiC fiber reinforced Ti3SiC2 and Ti3(SiAl)C2 composites. Mater. Sci. Eng. A 2021, 827, 142069. [Google Scholar] [CrossRef]
  11. Singh, J.; Wani, M. Fretting wear of spark plasma sintered Ti3SiC2/GNP ceramic composite against Si3N4. Ceram. Int. 2021, 47, 5648–5655. [Google Scholar] [CrossRef]
  12. Yang, J.; Ye, F.; Cheng, L. In-situ formation of Ti3SiC2 interphase in SiCf/SiC composites by molten salt synthesis. J. Eur. Ceram. Soc. 2021, 42, 1197–1207. [Google Scholar] [CrossRef]
  13. Giuranno, D.; Gambaro, S.; Bruzda, G.; Nowak, R.; Polkowski, W.; Sobczak, N.; Delsante, S.; Novakovic, R. Interface design in lightweight SiC/TiSi2 composites fabricated by reactive infiltration process: Interaction phenomena between liquid Si-rich Si-Ti alloys and glassy carbon. Materials 2021, 14, 3746. [Google Scholar] [CrossRef] [PubMed]
  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. Zhang, R.; Feng, W.; Wei, Q.; Ma, S. Synthesis and tribological characterization of Ti3SiC2/ZnO composites. Materials 2021, 14, 6088. [Google Scholar] [CrossRef]
  16. 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]
  17. Magnus, C.; Sharp, J.; Rainforth, W.M. The lubricating properties of spark plasma sintered (SPS) Ti3SiC2 MAX phase compound and composite. Tribol. Trans. 2020, 63, 38–51. [Google Scholar] [CrossRef]
  18. Yuqi, C.; Liang, L.; Shibang, M.; Chao, L.; Songhao, Z.; Wucheng, L.; Libo, W.; Aiguo, Z.; Xing, W. Preparation of Ti3Si0.8Al0.2C2 bonded diamond composites and their friction properties coupled with different counterfaces. Adv. Mater. Sci. Eng. 2023, 2023, 1740345. [Google Scholar] [CrossRef]
  19. Qin, J.; He, D. Phase stability of Ti3SiC2 at high pressure and high temperature. Ceram. Int. 2013, 39, 9361–9367. [Google Scholar] [CrossRef]
  20. Li, X.; Xu, L.; Chen, Q.; Cao, X.; Liu, L.; Wang, Y.; Zhang, H.; Meng, C.; Wu, Q. Investigation of formation mechanism of Ti3SiC2 by high pressure and high-temperature synthesis. High Press. Res. 2018, 38, 440–447. [Google Scholar] [CrossRef]
  21. Jaworska, L.; Stobierski, L.; Twardowska, A.; Królicka, D. Preparation of materials based on Ti–Si–C system using high temperature–high pressure method. J. Mater. Process. Technol. 2005, 162, 184–189. [Google Scholar] [CrossRef]
  22. Li, Y.; Zhang, X.; Zhang, S.; Song, X.; Wang, Y.; Chen, Z. First principles study of stability, electronic structure and fracture toughness of Ti3SiC2/TiC interface. Vacuum 2021, 196, 110745. [Google Scholar] [CrossRef]
  23. Jung, Y.I.; Park, D.J.; Park, J.H.; Park, J.Y.; Kim, H.G.; Koo, Y.H. Effect of TiSi2/Ti3SiC2 matrix phases in a reaction-bonded SiC on mechanical and high-temperature oxidation properties. J. Eur. Ceram. Soc. 2016, 36, 1343–1348. [Google Scholar] [CrossRef]
  24. Wan, C.; Wang, Y.; Wang, N.; Koumoto, K. Low-thermal-conductivity (MS)1+ x(TiS2)2 (M = Pb, Bi, Sn) misfit layer compounds for bulk thermoelectric materials. Materials 2010, 3, 2606–2617. [Google Scholar] [CrossRef]
  25. Ursi, F.; Virga, S.; Garcìa-Espejo, G.; Masciocchi, N.; Martorana, A.; Giannici, F. Long-term stability of TiS2–Alkylamine hybrid materials. Materials 2022, 15, 8297. [Google Scholar] [CrossRef]
  26. Niranjan, M.K. Anisotropy in elastic properties of TiSi2 (C49, C40 and C54), TiSi and Ti5Si3: An ab-initio density functional study. Mater. Res. Express 2015, 2, 096302. [Google Scholar] [CrossRef]
  27. Li, C.; Yu, Z.; Liu, H.; Lü, T. The crystallographic stability and anisotropic compressibility of C54-type TiSi2 under high pressure. J. Phys. Chem. Solids 2013, 74, 1291–1294. [Google Scholar] [CrossRef]
  28. Chen, Q.; Liu, L.; Xu, B.; Meng, C.; Li, X. TiSi2-SiC agglomerates toughened TiC composites prepared by in-situ reaction under high pressure. High Press. Res. 2019, 39, 598–607. [Google Scholar] [CrossRef]
  29. Tian, W.; Sun, Z.; Hashimoto, H.; Du, Y. Synthesis, microstructure and mechanical properties of Ti3SiC2–TiC composites pulse discharge sintered from Ti/Si/TiC powder mixture. Mater. Sci. Eng. A 2009, 526, 16–21. [Google Scholar] [CrossRef]
  30. Raju, G.; Basu, B.; Tak, N.; Cho, S. Temperature dependent hardness and strength properties of TiB2 with TiSi2 sinter-aid. J. Eur. Ceram. Soc. 2009, 29, 2119–2128. [Google Scholar] [CrossRef]
  31. Frommeyer, G.; Rosenkranz, R. Structures and properties of the refractory silicides Ti5Si3 and TiSi2 and Ti-Si-(Al) eutectic alloys. In Metallic Materials with High Structural Efficiency; Springer: Berlin/Heidelberg, Germany, 2004; pp. 287–308. [Google Scholar]
  32. Racault, C.; Langlais, F.; Bernard, C. On the chemical vapour deposition of Ti3SiC2 from TiCl4-SiCl4-CH4-H2 gas mixtures: Part IA thermodynamic approach. J. Mater. Sci. 1994, 29, 5023–5040. [Google Scholar] [CrossRef]
  33. Abu, M.J.; Mohamed, J.J.; Ahmad, Z.A. Effect of excess silicon on the formation of Ti3SiC2 using free Ti/Si/C powders synthesized via arc melting. Int. Sch. Res. Not. 2012, 2012, 341285. [Google Scholar]
  34. El Saeed, M.; Deorsola, F.A.; Rashad, R. Optimization of the Ti3SiC2 MAX phase synthesis. Int. J. Refract. Met. Hard Mater. 2012, 35, 127–131. [Google Scholar] [CrossRef]
  35. Foratirad, H.; Baharvandi, H.; Maraghe, M.G. Effect of excess silicon content on the formation of nano-layered Ti3SiC2 ceramic via infiltration of TiC preforms. J. Eur. Ceram. Soc. 2017, 37, 451–457. [Google Scholar] [CrossRef]
  36. Hosseinizadeh, S.A.; Pourebrahim, A.; Baharvandi, H.; Ehsani, N. Synthesis of nano-layered Ti3SiC2 MAX phase through reactive melt infiltration (RMI): Metallurgical and thermodynamical parameters. Ceram. Int. 2020, 46, 22208–22220. [Google Scholar] [CrossRef]
  37. Zhang, Z.F.; Sun, Z.M.; Hashimoto, H.; Abe, T. Effects of sintering temperature and Si content on the purity of Ti3SiC2 synthesized from Ti/Si/TiC powders. J. Alloys Compd. 2003, 352, 283–289. [Google Scholar] [CrossRef]
  38. Kero, I.; Antti, M.-L.; Odén, M. Synthesis of Ti3SiC2 by reaction of TiC and Si powders. In Proceedings of the International Conference on Advanced Ceramics and Composites: 27/01/2008–01/02/2008, Daytona Beach, FL, USA, 18–23 January 2009; pp. 21–30. [Google Scholar]
  39. Du, Y.; Schuster, J.C.; Seifert, H.J.; Aldinger, F. Experimental investigation and thermodynamic calculation of the titanium–silicon–carbon system. J. Am. Ceram. Soc. 2000, 83, 197–203. [Google Scholar] [CrossRef]
  40. Li, L.; Chen, Y. The Influence of sintering pressure on the preparation, friction properties, and magnetic properties of Ti2AlC-TiC and Ti3AlC2-TiC composites under high-pressure and high-temperature. Adv. Mater. Sci. Eng. 2022, 2022, 9108736. [Google Scholar] [CrossRef]
  41. Zhu, Y.; Zhou, A.; Ji, Y.; Jia, J.; Wang, L.; Wu, B.; Zan, Q. Tribological properties of Ti3SiC2 coupled with different counterfaces. Ceram. Int. 2015, 41, 6950–6955. [Google Scholar] [CrossRef]
  42. 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. [Google Scholar] [CrossRef]
  43. Zhimei, S.; Yanchun, Z.; Shu, L. Tribological behavior of Ti3SiC2-based material. J. Mater. Sci. Technol. 2002, 18, 142. [Google Scholar]
  44. Zhou, A.G.; Li, L.; Su, T.C.; Li, S.S. Synthesize Ti3SiC2 and Ti3SiC2-diamond composites at high pressure and high temperature. In Proceedings of the Key Engineering Materials, Singapore, 26–28 February 2012; pp. 671–675. [Google Scholar]
Figure 1. The experimental procedure of high-pressure high-temperature sintering technology.
Figure 1. The experimental procedure of high-pressure high-temperature sintering technology.
Materials 17 04866 g001
Figure 2. XRD patterns of Ti3SiC2-TiSix composites sintered under 1~5 GPa at 1150 °C for 30 min.
Figure 2. XRD patterns of Ti3SiC2-TiSix composites sintered under 1~5 GPa at 1150 °C for 30 min.
Materials 17 04866 g002
Figure 3. XRD patterns of Ti3SiC2-TiSix composites sintered under 3~5 GPa at 1250 °C for 30 min.
Figure 3. XRD patterns of Ti3SiC2-TiSix composites sintered under 3~5 GPa at 1250 °C for 30 min.
Materials 17 04866 g003
Figure 4. XRD patterns of Ti3SiC2-TiSix composites sintered under 4~5 GPa at 1400 °C for 30 min.
Figure 4. XRD patterns of Ti3SiC2-TiSix composites sintered under 4~5 GPa at 1400 °C for 30 min.
Materials 17 04866 g004
Figure 5. SEM images of the products from 3Ti/1.5Si/1.2C sintered at different conditions: (a) 1150 °C and 1 GPa; (b) layered structure of Ti3SiC2; (c) 1150 °C and 3 GPa low-magnification image; (d) high-magnification image of layered structure of Ti3SiC2; (e) high-magnification image of orthorhombic TiSi2; (f) interface between TiSi2 and layered Ti3SiC2 structure.
Figure 5. SEM images of the products from 3Ti/1.5Si/1.2C sintered at different conditions: (a) 1150 °C and 1 GPa; (b) layered structure of Ti3SiC2; (c) 1150 °C and 3 GPa low-magnification image; (d) high-magnification image of layered structure of Ti3SiC2; (e) high-magnification image of orthorhombic TiSi2; (f) interface between TiSi2 and layered Ti3SiC2 structure.
Materials 17 04866 g005
Figure 6. COF of Ti3SiC2-TiSix composites versus sliding time.
Figure 6. COF of Ti3SiC2-TiSix composites versus sliding time.
Materials 17 04866 g006
Figure 7. Wear rates of Ti3SiC2-TiSix composites sliding against Al ball.
Figure 7. Wear rates of Ti3SiC2-TiSix composites sliding against Al ball.
Materials 17 04866 g007
Figure 8. SEM images of Ti3SiC2-TiSix composites sliding against Al ball; (a) 2 GPa and dry sliding; (b) 4 GPa and dry sliding; (c) 4.5 GPa and dry sliding; (d) 2 GPa and wet sliding; (e) 4 GPa and wet sliding; (f) 4.5 GPa and wet sliding.
Figure 8. SEM images of Ti3SiC2-TiSix composites sliding against Al ball; (a) 2 GPa and dry sliding; (b) 4 GPa and dry sliding; (c) 4.5 GPa and dry sliding; (d) 2 GPa and wet sliding; (e) 4 GPa and wet sliding; (f) 4.5 GPa and wet sliding.
Materials 17 04866 g008
Figure 9. Typical measuring curves of the friction coefficients of Ti3SiC2-TiSix composites against the Cu ball pair at a load of 12 N and sliding speed of 0.05 m/s.
Figure 9. Typical measuring curves of the friction coefficients of Ti3SiC2-TiSix composites against the Cu ball pair at a load of 12 N and sliding speed of 0.05 m/s.
Materials 17 04866 g009
Figure 10. Variations in the wear rate of Ti3SiC2-TiSix composites against a Cu ball pair at a sliding speed of 0.05 m/s with a load of 12 N.
Figure 10. Variations in the wear rate of Ti3SiC2-TiSix composites against a Cu ball pair at a sliding speed of 0.05 m/s with a load of 12 N.
Materials 17 04866 g010
Figure 11. SEM images of Ti3SiC2-TiSix composites sliding against a Cu ball; (a) 2 GPa and dry sliding; (b) 4 GPa and dry sliding; (c) 4.5 GPa and dry sliding.
Figure 11. SEM images of Ti3SiC2-TiSix composites sliding against a Cu ball; (a) 2 GPa and dry sliding; (b) 4 GPa and dry sliding; (c) 4.5 GPa and dry sliding.
Materials 17 04866 g011aMaterials 17 04866 g011b
Figure 12. Comparison of stable areas of Ti3SiC2 [5,10,21] and Ti3SiC2-based composites.
Figure 12. Comparison of stable areas of Ti3SiC2 [5,10,21] and Ti3SiC2-based composites.
Materials 17 04866 g012
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

Chen, Y.; Li, L.; Han, M.; Sun, C.; Li, J. Sintering and Tribological Properties of Ti3SiC2-TiSix Composite Sintered by High-Pressure High-Temperature Technology. Materials 2024, 17, 4866. https://doi.org/10.3390/ma17194866

AMA Style

Chen Y, Li L, Han M, Sun C, Li J. Sintering and Tribological Properties of Ti3SiC2-TiSix Composite Sintered by High-Pressure High-Temperature Technology. Materials. 2024; 17(19):4866. https://doi.org/10.3390/ma17194866

Chicago/Turabian Style

Chen, Yuqi, Liang Li, Ming Han, Chaofan Sun, and Jin Li. 2024. "Sintering and Tribological Properties of Ti3SiC2-TiSix Composite Sintered by High-Pressure High-Temperature Technology" Materials 17, no. 19: 4866. https://doi.org/10.3390/ma17194866

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop