Next Article in Journal
Effect of Ho Addition on the Glass-Forming Ability and Crystallization Behaviors of Zr54Cu29Al10Ni7 Bulk Metallic Glass
Next Article in Special Issue
High Humidity Response of Sol–Gel-Synthesized BiFeO3 Ferroelectric Film
Previous Article in Journal
Effects of a ZnCuO-Nanocoated Ti-6Al-4V Surface on Bacterial and Host Cells
Previous Article in Special Issue
Studies on Electron Escape Condition in Semiconductor Nanomaterials via Photodeposition Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Mechanical Properties of Composites Obtained by Spark Plasma Sintering of Ti3SiC2-15 vol.%Cu Mixtures

1
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
2
State Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou 730000, China
3
Sichuan Province Engineering Technology Research Center of Powder Metallurgy, 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.
Submission received: 13 February 2022 / Revised: 25 March 2022 / Accepted: 28 March 2022 / Published: 29 March 2022
(This article belongs to the Special Issue Microstructural Design and Processing Control of Advanced Ceramics)

Abstract

:
Method of soft metal (Cu) strengthening of Ti3SiC2 was conducted to increase the hardness and improve the wear resistance of Ti3SiC2. Ti3SiC2/Cu composites containing 15 vol.% Cu were fabricated by Spark Plasma Sintering (SPS) in a vacuum. The effect of the sintering temperature on the phase composition, microstructure and mechanical properties of the composites was investigated in detail. The as-synthesized composites were thoroughly characterized by scanning electron micrography (SEM), optical micrography (OM) and X-ray diffractometry (XRD), respectively. The results indicated that the constituent of the Ti3SiC2/Cu composites sintered at different temperatures included Ti3SiC2, Cu3Si and TiC. The formation of Cu3Si and TiC originated from the reaction between Ti3SiC2 and Cu, which was induced by the presence of Cu and the de-intercalation of Si atoms Ti3SiC2. OM analysis showed that with the increase in the sintering temperature, the reaction between Ti3SiC2 and Cu was severe, leading to the Ti3SiC2 getting smaller and smaller. SEM measurements illustrated that the uniformity of the microstructure distribution of the composites was restricted by the agglomeration of Cu, controlling the mechanical behaviors of the composites. At 1000 °C, the distribution of Cu in the composites was relatively even; thus, the composites exhibited the highest density, relatively high hardness and compressive strength. The relationships of the temperature, the current and the axial dimension with the time during the sintering process were further discussed. Additionally, a schematic illustration was proposed to explain the related sintering characteristic of the composites sintered by SPS. The as-synthesized Ti3SiC2/Cu composites were expected to improve the wear resistance of polycrystalline Ti3SiC2.

1. Introduction

As a typical ternary layered MAX phase, Ti3SiC2 exhibited combined characteristics of both metals and ceramics. It possessed good machinability, good electrical and thermal conductivity, good ductility, thermal stability, oxidation resistance, etc. [1,2,3,4]. It has potential for applications in electro friction materials, novel structural/functional ceramic materials and high-temperature lubricating materials.
The crystalline structure of Ti3SiC2 can be described as a sandwich structure: a layer of Si and the twin boundary of TiC. The chemical bonding between the Si atom and the other atoms, such as Ti and C, was to a certain extent relatively weak compared to the strong Ti-C bonding [5,6]. Based on its unique structure, Ti3SiC2 was chemically reactive when it contacted with metals at high temperatures. Li et al. [7] fabricated Ti3SiC2/Ni and Ti3SiC2/Co through vacuum sintering. It was found that the metals (Ni and Co) were prone to aggregating towards their surfaces due to the poor wettability between Ti3SiC2 and Ni or Co. Gu et al. [8] investigated the possibility of fabricating Ti3SiC2-Ti composites. They discovered that the as-obtained composites were mainly composed of Ti3SiC2, TiCx, Ti5Si3, and TiSi2. It was inferred that the decomposition of Ti3SiC2 was owing to the de-intercalation of Si and the separation of carbon from Ti3SiC2. Gu et al. [9] studied the reactions between Ti3SiC2 and Al in the temperature range of 600–650 °C, and they found that besides Ti3SiC2 and Al, new phases Al3Ti, Al4SiC4 and Al4C3 were generated, the formation of which relied on the time, temperature and the relative amount of Al and Ti3SiC2. Kothalkar et al. [10] synthesized NiTi-Ti3SiC2 composite, and they discovered that the composites showed higher damping up to applied stress of 200 MPa. For example, the energy dissipation of the composite was thirteen times larger than pure Ti3SiC2 and two times larger than pure NiTi.
Except for the Ti3SiC2-metals composites described above, researchers concentrated on the investigation of Ti3SiC2/Cu composites. Dang et al. [11] fabricated Ti3SiC2/Cu composites with different contents of Cu by mechanical alloying and spark plasma sintering. They inferred that the presence of Cu leads to the decomposition of Ti3SiC2 to form TiCx, Ti5Si3Cy, Cu3Si, and TiSi2Cz. In another paper, Dang et al. [12] synthesized Ti3SiC2/Cu/Al/SiC composites by powder metallurgy/spark plasma sintering and found that the addition of Al could inhibit the decomposition of Ti3SiC2. Lu et al. [5] found chemical reaction between Cu and Ti3SiC2 contributes to the wettability. Zhou and his colleagues [13] investigated the chemical reactions and stability of Ti3SiC2 in Cu for the Cu/Ti3SiC2 composites. They found that at low content of Ti3SiC2 or below 1000 °C, Cu (Si) solid solution and TiCx were generated, whereas at high temperature or high content of Ti3SiC2, Cu-Si intermetallic compounds, such as Cu5Si, Cu15Si4, and TiCx, were generated. In our recent publication [14], Ti3SiC2/Cu composites were synthesized by spark plasma sintering technique at various temperatures. The microstructure, composition and mechanical properties of the as-obtained composites were investigated. The results indicated that the Ti3SiC2/Cu composite sintered at 1100 °C exhibited superior mechanical properties. While these results give partial information on the reactions occurring between Ti3SiC2 and Cu, they did not permit elucidation of the entire sintering behaviors occurring during the sintering process.
In this paper, the Ti3SiC2/Cu composites were fabricated by Spark Plasma Sintering (SPS) at different sintering temperatures. The sintering behaviors of the composites were explored by the characterization of the phase composition, microstructure and mechanical properties of the composites based on the relationship of the temperature, the current and the pressure with the time during the sintering process.

2. Experiment

2.1. Samples Preparation

Ti3SiC2/Cu composites were fabricated using powder mixture of Ti3SiC2 (average particle size: 38 μm, ≥98% purity, 11 technology Co., Ltd., Jilin, China) and Cu (average particle size: 74 μm, ≥99.9% purity, Macklin Biochemical Co., Ltd., Shanghai, China). The content of Cu in Ti3SiC2/Cu composites was 15 vol.%. The mixture with designed composition was first mixed by a ball-milling machine (PMQD2LB, Nanjing Chishun Technology Development Co., Ltd., Nanjing, China) with a rotational speed of 150 rpm and a ratio of ball to powder of 3 for 6 h, then filled into a graphite die (inner diameter: Φ25 mm). Finally, it was sintered by Spark Plasma Sintering (SPS, Model Labox-350, Xinxie, Japan) at different sintering temperatures (950, 1000 and 1050 °C, respectively) under a pressure of 35 MPa in vacuum for 20 min and cooled with the furnace. The heating rate was set as follows. From the room temperature to 600 °C, the heating rate was 100 °C/min. Additionally, from 600 °C to the desired sintering temperature, the heating rate was 50 °C/min.

2.2. Mechanical Property

The density of the as-prepared Ti3SiC2/Cu composites was measured by the Archimedes method. Cylinder specimens with a size of φ5 mm × 12 mm were machined for compressive strength testing, which was performed on WDW-100 universal materials testing machine (Jinan Hansen Precision Instrument Co., Ltd., Jinan, China). The Vickers hardness of the composites was determined in an MHVD-50AP microhardness tester (Shanghai Jujing Precision Instrument Manufacturing Co., Ltd., Shanghai, China) at a load of 1 kg with a dwell time of 10 s.

2.3. Analysis

The as-synthesized Ti3SiC2/Cu composites were polished using 0.5 um polishing paste by an automatic polishing machine (AutoMetTM250, Yigong Testing and Measuring Instrument Co., Ltd. Shanghai, China) for microscopic evaluation. In order to expose the grains, the polished samples were etched using a 1:1:1 by volume HF:HNO3:H2O solution and observed under optical microscopy (MDJ-DM, Chongqing Auto Optical Instrument Co., Ltd., Chongqing, China). The compression fracture morphology of Ti3SiC2/Cu composites was observed by scanning electron microscopy (SEM, JSM-6510LA, JEOL Japan Electronics Co., Ltd., Zhaodao, Japan) equipped with energy dispersive spectroscopy (EDS). X-ray diffraction (XRD) analysis was carried out on a DX-2700B diffractometer (Dandong Haoyuan Instrument Co., Ltd., Dandong, China) with Cu Kα radiation at a scanning rate of 7.2 °/min to identify the phase composition of Ti3SiC2/Cu composites.

3. Results and Discussion

3.1. Phase Composition and Microstructure

Figure 1 shows XRD patterns of the as-synthesized Ti3SiC2/Cu composites at different sintered temperatures. It was clearly observed that the as-synthesized Ti3SiC2/Cu composites were all composed of Ti3SiC2, Cu3Si and TiC, which was independent of the sintered temperature.
Based on the unique sandwich structure of Ti3SiC2, it exhibited high reactivity when contacting with metal phases [5,11,15,16,17,18,19]. On the one hand, the Si atoms were easily de-intercalated from Ti3SiC2. On the other hand, Si-containing solid solution or intermetallic compounds was prone to form when the metal phases contacted with Ti3SiC2. The presence of contacted metal phases accelerated the decomposition of Ti3SiC2. As for Ti3SiC2/Cu composites, Cu can form Cu (Si) solid solution or react with Si to form CuxSiy intermetallic compounds [18]. Therefore, the composition of Ti3SiC2/Cu composites consisted of Cu(Si) solid solution, CuxSiy intermetallic compounds, and TiCz, which were commonly examined by researchers [5,11,13].
In this study, with the change in sintered temperatures, Cu3Si was the only Cu-Si intermetallic compound, and TiC was the only decomposed product of Ti3SiC2. According to Guo et al. [20], when the content of Cu was less than that of Si, Cu3Si was the preferential product of the Cu-Si system. Because based on the binary phase diagram of Cu-Si, the content of Si in Cu3Si (22.2–25.2%) was the maximum among six copper silicides (Cu (Si), Cu7Si, Cu5Si, Cu4Si, Cu15Si4 and Cu3Si) [20]. This explained the reason why Cu3Si was the single Cu-Si intermetallic compound in our present study.
The effect of sintered temperature on the microstructure is shown in Figure 2. As seen in Figure 2a, at 950 °C, the typical plate-like morphology of Ti3SiC2 grains was evidently inhibited by the addition of Cu. A considerable amount of Ti3SiC2 equiaxed grains appeared due to the reaction between Ti3SiC2 and Cu. Additionally, with the increase in the sintered temperature from 950 °C to 1050 °C, the Ti3SiC2 granules decreased, which was accompanied by fewer pores or holes, indicating higher reactivity of Ti3SiC2 and Cu.
The back scattering electron images of polished and etched Ti3SiC2/Cu composites sintered at 950 °C, 1000 °C and 1050 °C are shown in Figure 3. As mentioned above, the main composition of Ti3SiC2/Cu composites were Ti3SiC2, Cu3Si and TiC. Ti3SiC2 was located at the dark grey area in Figure 3; both Cu3Si and TiC were located at the light grey area in Figure 3. It was clearly seen from Figure 3 that the as-formed Cu3Si distributed along the grain boundary of Ti3SiC2, and it was accompanied by the formation of hard TiC particles. Moreover, with the increase in the sintered temperature, the reaction between Ti3SiC2 and Cu became more severe, resulting in the formation of Cu3Si with a non-negligible amount, especially at 1000 °C. The reaction mechanism between Ti3SiC2 and Cu was proposed as follows. After milling, the Cu powder is relatively evenly distributed in Ti3SiC2 powder. At elevated temperatures (for example, 900 °C), the Si atoms de-intercalated from Ti3SiC2 grains, diffused rapidly around Cu and reacted with Cu to form Cu3Si. Meanwhile, the original Ti3SiC2 skeleton structure transformed to TiC structure, which was concomitant with the formation of pores or holes due to the mismatch of skeletal density between Ti3SiC2 and TiC. When the sintered temperature raised (for example, 1000 °C), the de-intercalation of Si atoms from the Ti3SiC2 skeleton was accelerated by the defects (pores or holes) and high temperature. On the other hand, Cu had a tendency to melt and flow around the grain boundaries of Ti3SiC2 grains and the as-formed defects mentioned above. All these led to more violent reactions between Ti3SiC2 and Cu, causing the Ti3SiC2 grains to become smaller and smaller. This reaction process well coincided with Zhou et al. [13]. Theoretically, with the increase in temperature, the reactivity of Ti3SiC2 and Cu increased, producing more Cu3Si. However, when the sintered temperature reached 1050 °C, a large amount of Cu melted and released from the graphite die, leading to the loss of Cu. As seen in Figure 3, visually, the content of Cu3Si was largest at 1000 °C, not at 1050 °C.
Figure 4 illustrates the mapping of elemental distribution for Ti3SiC2/Cu composites sintered at different temperatures. It was irrefutable evidence for the explanation of the sintering effect of the composites. At 950 °C, the Cu considerably aggregated in the Ti3SiC2/Cu composites (see Figure 4a). At this temperature, a solid–solid sintering process occurred. The agglomeration of Cu relied on the uniformity of its distribution in Ti3SiC2 during ball milling. As we know, ball milling cannot avoid material agglomeration. At 1000 °C, it was obviously observed that Cu was relatively evenly distributed in the Ti3SiC2/Cu composites (see Figure 4b). It was speculated that solid–liquid sintering process took place at 1000 °C. The appropriate flowability of the quasi-liquid Cu contributed to not only its reaction with Si atoms but also its uniform distribution in the composites. At 1050 °C, the Cu melted and rapidly flowed around the Ti3SiC2 grains, leading to the relatively obvious aggregation of Cu (see Figure 4c). Moreover, the release of liquid Cu during sintering caused the loss of Cu to a certain extent. Therefore, it was concluded that the optimal sintering temperature was 1000 °C for the Ti3SiC2/Cu composites in our study.

3.2. Mechanical Property

The variation in relative density, hardness and compressive strength of the Ti3SiC2/Cu composites with the sintered temperature are shown in Figure 5. As seen in Figure 5, the relative density of the Ti3SiC2/Cu composites sintered at 950 °C was 95.35%, which was slightly lower than that sintered at 1000 °C (96.43%). However, the relative density of the composite sintered at 1050 °C was 93.85%, which was the lowest among the three samples. The higher relative density of the Ti3SiC2/Cu composites sintered at 1000 °C was attributed to the synergetic effect of high temperature, high pressure and pulse current during the sintering process. The most important thing was that the quasi-liquid Cu possessed proper mobility, which was beneficial for the densification of the composites as well. At 1050 °C, its lowest relative density was also related to the flowability of Cu. In such circumstances, Cu flowed easily and released rapidly, causing the aggregation of Cu (see Figure 4c) and the loss of Cu (see Figure 3c).
As shown in Figure 5, both the hardness and the compressive strength of the Ti3SiC2/Cu composites increased with the increase in the sintered temperature. It was clearly seen that the deviation of the hardness of the composites sintered at 950 °C was the highest among the three samples, which originated from the agglomeration of Cu in the composites (see Figure 3a and Figure 4a). Compared with the polycrystalline Ti3SiC2 (5.5 GPa) [21], the higher hardness of the composites at different temperatures came from the formation of hard TiC product, which was detected by XRD analysis (see Figure 1).
Additionally, in comparison with the polycrystalline Ti3SiC2, the compressive strength of the composites sintered at different temperatures exhibited an equivalent or higher value [22], which was due to the appropriate reaction between Ti3SiC2 and Cu. As mentioned above, during the sintering process, the de-intercalation of Si from Ti3SiC2 and thereafter the dissolution of it in the liquid Cu phase led to the formation of TiC and Cu3Si. Moreover, the Cu3Si is uniformly distributed along the grain boundary of Ti3SiC2. The as-obtained fine TiC and Cu3Si grains were uniformly distributed along the boundary of Ti3SiC2 grains, which was a benefit for the higher compressive strength of the composite. As seen in Figure 6a, the Ti3SiC2/Cu composites showed a brittle fracture character, which was identical with polycrystalline Ti3SiC2. From the compression fracture morphology of the composites (see Figure 6b–d), it was apparently seen that both the intergranular fracture and transgranular fracture were present on the compression fracture surface of the composites sintered at different temperatures. It indicated that the fracture mode of the composites was independent of the sintered temperature.

3.3. Sintering Behaviors of the Ti3SiC2/Cu Composites

It is instructive to explain the sintering process of the Ti3SiC2/Cu composites at different temperatures. The relationships of temperature, current and axial dimension with time is illustrated in Figure 7. In comparison, although the temperature difference for the three sintering temperatures was the same (950–1000 °C and 1000–1050 °C), the change in the axial dimension was different. The change in the axial dimension for 950–1000 °C was obviously lower than that for 1000–1050 °C.
The profile of temperature, current and axial dimension with time was insensitive with the sintering temperature (see Figure 7a–c). We took 1000 °C as an example to elaborate the related connections among the temperature, the current, the axial dimension and time. As seen in Figure 7b, from room temperature to 600 °C (Note: without monitoring temperature by the infrared thermometer), the heating rate was fast (approximate 100 °C/min) in order to remove moisture to dry powder completely, thus the current increased rapidly to 2300 A. During this period, the decrease in the axial dimension was due to the action of pressure and the softening due to the heating, and the first maximum of the axial dimension at about 5 min resulted from the thermal expansion and contraction. Hereafter, from 600 to the sintering temperature (for example, 1000 °C), the heating rate was set as 50 °C/min; therefore, the current rapidly adjusted to a low value (about 750 A) and then increased at a proper rate to make sure the sintering temperature intelligently controlled. At the same time, the axial dimension was controlled by a comprehensive impact of pressure, current and the thermal expansion and contraction, and it kept relatively stable before 10 min, then it remarkably decreased from 800 °C to the sintering temperature (for example, 1000 °C). The rapid decrease in the axial dimension was mainly attributed to the densification effect due to the continuous pressure at the sintering temperature and possibly due to the reaction between Ti3SiC2 and Cu. Finally, in the holding period, both the temperature and the current were constant, and the axial dimension slowly decreased as a result of the further densification. Additionally, an extensive reaction between Ti3SiC2 and Cu was partly beneficial for the decrease in the axial dimension.
The proposed sintering process of the Ti3SiC2/Cu composites is shown in Figure 8. In the initial state (see Figure 8a), Ti3SiC2 and Cu were relatively uniformly distributed in the graphite die, which corresponded to the state of the first 10 min in Figure 7. During the initial stage, there was no reaction between Ti3SiC2 and Cu, and the softening of the powder and the thermal expansion and extraction were due to the action of the current and the pressure. In the transition state (see Figure 8b), under the synergetic effect of the current, temperature and pressure, Cu locally melted and was extruded into the grain boundary of Ti3SiC2.
On the other hand, Ti3SiC2 grains underwent plastic deformation under the same condition. All these factors increased the contact area of Ti3SiC2 and Cu. Additionally, the de-intercalation of Si atoms from Ti3SiC2 and its diffusion contributed to the reaction between Ti3SiC2 and Cu, leading to the formation of Cu3Si and TiC, which was distributed along the grain boundary of Ti3SiC2 grains. Moreover, the quasi-liquid Cu diffused into the inner part of Ti3SiC2 grains through the holes or pores produced by the mismatch of skeleton density of the original Ti3SiC2 and the as-formed TiC and reacted further with the Si atoms de-intercalated from the inner Ti3SiC2 grains. For the Ti3SiC2 grain surrounded by Cu, its surface was continuously transformed into TiC, which was accompanied by the formation of Cu3Si. Accordingly, the grain consisting of Cu3Si and TiC, which was embedded by Ti3SiC2, was expected (see the inset of Figure 3b and Figure 8b). Therefore, the grain size of Ti3SiC2 became smaller and smaller, which was inconsistence with the result in Figure 2. This state (the transition state in Figure 8b) corresponded to the rapid decrease in the axial dimension in Figure 7. In the final state (see Figure 8c), the reaction of Ti3SiC2 and Cu continued, and the pores were filled by the product, densifying the composites. This final state corresponds to the holding period in Figure 7. Consequently, the Ti3SiC2/Cu composites with superior mechanical properties were obtained at 1000 °C for 20 min by SPS. The proposed sintering process undoubtedly inspired us to further explore the sintering of Ti3SiC2/Cu composites in detail.

4. Conclusions

Ti3SiC2/Cu composites were sintered by Spark Plasma Sintering (SPS) in a vacuum under a pressure of 35 MPa at 950–1050 °C. The as-synthesized composites were systematically characterized by scanning electron micrography (SEM), optical micrography (OM) and X-ray diffractometry (XRD), respectively. The results indicated that the constituent of the Ti3SiC2/Cu composites included Ti3SiC2, Cu3Si and TiC, which was insensitive to the sintering temperature. The reaction between Ti3SiC2 and Cu produced Cu3Si and TiC, the formation of which was attributed to the presence of Cu and the de-intercalation of Si atoms Ti3SiC2. OM analysis confirmed that with the increase in the sintering temperature, the Ti3SiC2 became smaller and smaller, which resulted from the more violent reaction between Ti3SiC2 and Cu. Moreover, SEM analysis demonstrated that the agglomeration of Cu in the composites limited the uniformity of the microstructure distribution of the composites, governing the mechanical behaviors of the composites. At 1000 °C, the distribution of Cu in the composites was relatively uniform; thus, the composites exhibited the highest density, relatively high hardness and compressive strength. The relationships of the temperature, the current and the axial dimension with the time during the sintering process were further discussed, and a corresponding schematic illustration was proposed to explain the related sintering behaviors of the composites sintered by SPS.

Author Contributions

Conceptualization, R.Z.; Data curation, B.C. and H.Z.; Formal analysis, R.Z., B.C., M.S. and C.W.; Investigation, R.Z., F.L., M.S. and C.W.; Methodology, R.Z. and B.C.; Writing—original draft, R.Z., B.C. 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 was funded by the Open Project of State Key Laboratory of Solid Lubrication, Chinese Academy of Sciences (LSL-2115), and Sichuan Province Key Research and Development Program (22SYX0017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

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. Zhou, Y.; Sun, Z.; Chen, S.; Zhang, Y. In-situ hot pressing/solid-liquid reaction synthesis of dense titanium silicon carbide bulk ceramics. Mater. Res. Innov. 1998, 2, 142–146. [Google Scholar] [CrossRef]
  3. Barsoum, M.W.; El-Raghy, T. Synthesis and characterization of a remarkable ceramic: Ti3SiC2. J. Am. Ceram. Soc. 1996, 79, 1953–1956. [Google Scholar] [CrossRef]
  4. El-Raghy, T.; Zavaliangos, A.; Barsoum, M.W.; Kalidindi, S.R. Damage Mechanisms around Hardness Indentations in Ti3SiC2. J. Am. Ceram. Soc. 1997, 80, 513–516. [Google Scholar] [CrossRef]
  5. 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. 2015, 114, 39–44. [Google Scholar] [CrossRef]
  6. Low, I.M. Depth profiling of phase composition in a novel Ti3SiC2–TiC system with graded interfaces. Mater. Lett. 2004, 58, 927–932. [Google Scholar] [CrossRef]
  7. Li, H.; Peng, L.M.; Gong, M.; He, L.H.; Zhao, J.H.; Zhang, Y.F. Processing and microstructure of Ti3SiC2/M (M=Ni or Co) composites. Mater. Lett. 2005, 59, 2647–2649. [Google Scholar] [CrossRef]
  8. Gu, W.-L.; Zhou, Y.-C. Reactions between Ti and Ti3SiC2 in temperature range of 1273–1573 K. Trans. Nonferrous Met. Soc. China 2006, 16, 1281–1288. [Google Scholar] [CrossRef]
  9. Gu, W.L.; Yan, C.K.; Zhou, Y.C. Reactions between Al and Ti3SiC2 in temperature range of 600–650 °C. Scr. Mater. 2003, 49, 1075–1080. [Google Scholar] [CrossRef]
  10. Kothalkar, A.D.; Benitez, R.; Hu, L.; Radovic, M.; Karaman, I. Thermo-mechanical Response and Damping Behavior of Shape Memory Alloy–MAX Phase Composites. Metall. Mater. Trans. A 2014, 45, 2646–2658. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Zhou, Y.C.; Gu, W.L. 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]
  14. Zhang, R.; Liu, F.; Tulugan, K. Self-lubricating behavior caused by tribo-oxidation of Ti3SiC2/Cu composites in a wide temperature range. Ceram. Int. 2022, in press. [Google Scholar]
  15. Zhou, Y.; Chen, B.; Wang, X.; Yan, C. Mechanical properties of Ti3SiC2 particulate reinforced copper prepared by hot pressing of copper coated Ti3SiC2 and copper powder. Mater. Sci. Tech. Lond. 2004, 20, 661–665. [Google Scholar] [CrossRef]
  16. Gao, N.F.; Miyamoto, Y. Joining of Ti3SiC2 with Ti–6Al–4V Alloy. J. Mater. Res. 2011, 17, 52–59. [Google Scholar] [CrossRef]
  17. El-Raghy, T.; Barsoum, M.W.; Sika, M. Reaction of Al with Ti3SiC2 in the 800–1000 °C temperature range. Mat. Sci. Eng. A 2001, 298, 174–178. [Google Scholar] [CrossRef]
  18. Tzenov, N.; Barsoum, M.W.; 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]
  19. Yin, X.H.; Li, M.S.; Zhou, Y.C. Microstructure and mechanical strength of diffusion-bonded Ti3SiC2/Ni joints. J. Mater. Res. 2011, 21, 2415–2421. [Google Scholar] [CrossRef]
  20. Guo, H.; Zhang, J.; Li, F.; Liu, Y.; Yin, J.; Zhou, Y. Surface strengthening of Ti3SiC2 through magnetron sputtering Cu and subsequent annealing. J. Eur. Ceram. Soc. 2008, 28, 2099–2107. [Google Scholar] [CrossRef]
  21. 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]
  22. 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 Pt A, 10380–10386. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ti3SiC2/Cu composites at different sintered temperatures.
Figure 1. XRD patterns of Ti3SiC2/Cu composites at different sintered temperatures.
Materials 15 02515 g001
Figure 2. Optical micrographs of polished and etched Ti3SiC2/Cu composites sintered at 950 °C (a), 1000 °C (b) and 1050 °C (c).
Figure 2. Optical micrographs of polished and etched Ti3SiC2/Cu composites sintered at 950 °C (a), 1000 °C (b) and 1050 °C (c).
Materials 15 02515 g002
Figure 3. BSE micrographs of the Ti3SiC2/Cu composites sintered at (a) 950 °C, (b) 1000 °C (inset showed the structure at a higher magnification) and (c) 1050 °C (inset showed the structure at a higher magnification).
Figure 3. BSE micrographs of the Ti3SiC2/Cu composites sintered at (a) 950 °C, (b) 1000 °C (inset showed the structure at a higher magnification) and (c) 1050 °C (inset showed the structure at a higher magnification).
Materials 15 02515 g003
Figure 4. Mapping of elemental distribution for Ti3SiC2/Cu composites sintered at 950 °C (a), 1000 °C (b) and 1050 °C (c).
Figure 4. Mapping of elemental distribution for Ti3SiC2/Cu composites sintered at 950 °C (a), 1000 °C (b) and 1050 °C (c).
Materials 15 02515 g004
Figure 5. The variation in relative density, hardness and compressive strength of the Ti3SiC2/Cu composites versus the sintered temperature.
Figure 5. The variation in relative density, hardness and compressive strength of the Ti3SiC2/Cu composites versus the sintered temperature.
Materials 15 02515 g005
Figure 6. (a) Compressive stress–strain curve of Ti3SiC2/Cu composites, and the fracture morphology of Ti3SiC2/Cu composites sintered at (b) 950 °C (inset showed the structure at a higher magnification), (c) 1000 °C and (d) 1050 °C.
Figure 6. (a) Compressive stress–strain curve of Ti3SiC2/Cu composites, and the fracture morphology of Ti3SiC2/Cu composites sintered at (b) 950 °C (inset showed the structure at a higher magnification), (c) 1000 °C and (d) 1050 °C.
Materials 15 02515 g006
Figure 7. The relationship of the temperature, the current and the dimensional change with sintering time for the Ti3SiC2/Cu composites sintered at (a) 950 °C, (b) 1000 °C and (c) 1050 °C.
Figure 7. The relationship of the temperature, the current and the dimensional change with sintering time for the Ti3SiC2/Cu composites sintered at (a) 950 °C, (b) 1000 °C and (c) 1050 °C.
Materials 15 02515 g007
Figure 8. Proposed schematic illustration showing the sintering process of the Ti3SiC2/Cu composites: (a) initial state, (b) transition state and (c) final state.
Figure 8. Proposed schematic illustration showing the sintering process of the Ti3SiC2/Cu composites: (a) initial state, (b) transition state and (c) final state.
Materials 15 02515 g008
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.; 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. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15072515

AMA Style

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(7):2515. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15072515

Chicago/Turabian Style

Zhang, Rui, Biao Chen, Fuyan Liu, Miao Sun, Huiming Zhang, and Chenlong Wu. 2022. "Microstructure and Mechanical Properties of Composites Obtained by Spark Plasma Sintering of Ti3SiC2-15 vol.%Cu Mixtures" Materials 15, no. 7: 2515. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15072515

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