首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 93 毫秒
1.
MoSi2-particulate-reinforced α-SiAlON ceramic composites containing 10, 20, 25, and 30 vol% were prepared by hot pressing at 1750°-1800°C. The α-SiAlON matrix was of the composition (Y0.48Si10.00A12.30O1.17N15.29). The hardness for the fully dense samples changed from HV10 = 22.5 to 15.3 GPa and the toughness from 3.2 to around 5.2 MPa.m1/2 when up to 30 vol% MoSi2 was present. Two interesting microstructural features have been found. First, with an increasing amount of MoSi2 a pronounced coalescence of MoSi2 particles formed a "dual phase" material. The second effect was the growth of elongated α-SiAlON grains in the matrix with 10 vol% MoSi2 added. The oxidation resistance has been determined to be unaffected by the addition of 2hd vol % MoSi2 at 1250°C in oxygen gas of l atm pressure.  相似文献   

2.
Hafnium diboride (HfB2)- and hafnium carbide (HfC)-based materials containing MoSi2 as sintering aid in the volumetric range 1%–9% were densified by spark plasma sintering at temperatures between 1750° and 1950°C. Fully dense samples were obtained with an initial MoSi2 content of 3 and 9 vol% at 1750°–1800°C. When the doping level was reduced, it was necessary to raise the sintering temperature in order to obtain samples with densities higher than 97%. Undoped powders had to be sintered at 2100°–2200°C. For doped materials, fine microstructures were obtained when the thermal treatment was lower than 1850°C. Silicon carbide formation was observed in both carbide- and boride-based materials. Nanoindentation hardness values were in the range of 25–28 GPa and were independent of the starting composition. The nanoindentation Young's modulus and the fracture toughness of the HfB2-based materials were higher than those of the HfC-based materials. The flexural strength of the HfB2-based material with 9 vol% of MoSi2 was higher at 1500°C than at room temperature.  相似文献   

3.
The physical and mechanical properties of hot-pressed Si3N4–MoSi2 particulate composites containing 15 and 30 vol% MoSi2 were studied. The average room-temperature four-point bend strength, fracture toughness, and electrical resistivity are 522 MPa, 3.6 MPa·√m, and 6.3 × 105Χ·cm for the 15 vol% MoSi2 composite, and 487 MPa, 5.3 MPa·√m, and 0.31 Ω·cm for the 30 vol% MoSi2 composite. The mechanical properties of the composites are very close to those of hot-pressed Si3N4 ceramics. The high electrical conductivity of the 30 vol% MoSi2 composite was attributed to the percolation effect of MoSi2 particles. Parabolic oxidation behaviors were observed for the 30 vol% MoSi2 composite during the 1200°C long-term oxidation experiments.  相似文献   

4.
Abrasive Wear Behavior of a Si3N4-MoSi2 Composite   总被引:5,自引:0,他引:5  
MoSi2 particles (20 vol%) have been added to Si3N4 to form ceramic matrix-intermetallic composites. Benefits associated with the addition of the MoSi2 to Si3N4 include higher strength, higher fracture toughness, no loss in oxidation resistance, and lower electrical resistivity. However, because the hardness of MoSi2 is approximately half that of Si3N4, a decrease in the specific wear rate of the Si3N4-20 vol% MoSi2 composite is expected to result from the incorporation of the MoSi2 into the Si3N4. In this U.S. Bureau of Mines and Los Alamos National Laboratory study, it is found, however, that the specific wear rate of the composite during two-body abrasion by SiC particles is equivalent in magnitude to the specific wear rate of monolithic Si3N4. The specific wear rates of both the Si3N4-20 vol% MoSi2composites and monolithic Si3N4 are four to five times less than that of monolithic MoSi2.  相似文献   

5.
The thermal and electrical properties of MoSi2 and/or SiC-containing ZrB2-based composites and the effects of MoSi2 and SiC contents were examined in hot-pressed ZrB2–MoSi2–SiC composites. The thermal conductivity and electrical conductivity of the ZrB2–MoSi2–SiC composites were measured at room temperature by a nanoflash technique and a current–voltage method, respectively. The results indicate that the thermal and electrical conductivities of ZrB2–MoSi2–SiC composites are dependent on the amount of MoSi2 and SiC. The thermal conductivities observed for all of the compositions were more than 75 W·(m·K)−1. A maximum conductivity of 97.55 W·(m·K)−1 was measured for the 20 vol% MoSi2-30 vol% SiC-containing ZrB2 composite. On the other hand, the electrical conductivities observed for all of the compositions were in the range from 4.07 × 10–8.11 × 10 Ω−1·cm−1.  相似文献   

6.
The microstructure of two pressureless-sintered ultra-high-temperature ceramics, namely ZrC+20 vol% MoSi2 and HfC+20 vol% MoSi2, was characterized by scanning and transmission electron microscopy. With regard to the ZrC–MoSi2 system, Zr x Si y compounds and SiC were detected. In the HfC–MoSi2 system, a mixed phase was detected at the triple points and identified as (Mo,Hf)5Si3. For both the systems investigated, the high wettability of the silicide-based phases on the matrix grains suggests that sintering is assisted by a liquid phase. This contribution reports for the first time on the sintering mechanisms of early transition metal carbides doped with MoSi2 as a sinter additive, on the basis of the microstructural evolution observed upon sintering and in the light of phase diagrams and thermodynamical calculations.  相似文献   

7.
The densification of non-oxide ceramics like titanium boride (TiB2) has always been a major challenge. The use of metallic binders to obtain a high density in liquid phase-sintered borides is investigated and reported. However, a non-metallic sintering additive needs to be used to obtain dense borides for high-temperature applications. This contribution, for the first time, reports the sintering, microstructure, and properties of TiB2 materials densified using a MoSi2 sinter-additive. The densification experiments were carried out using a hot-pressing and pressureless sintering route. The binderless densification of monolithic TiB2 to 98% theoretical density with 2–5 μm grain size was achieved by hot pressing at 1800°C for 1 h in vacuum. The addition of 10–20 wt% MoSi2 enables us to achieve 97%–99%ρth in the composites at 1700°C under similar hot-pressing conditions. The densification mechanism is dominated by liquid-phase sintering in the presence of TiSi2. In the pressureless sintering route, a maximum of 90%ρth is achieved after sintering at 1900°C for 2 h in an (Ar+H2) atmosphere. The hot-pressed TiB2–10 wt% MoSi2 composites exhibit high Vickers hardness (∼26–27 GPa) and modest indentation toughness (∼4–5 MPa·m1/2).  相似文献   

8.
Particulate ceramic composites that were composed of a combustion-synthesized β';-SiAlON matrix and dispersed MoSi2 particles were hot pressed at 1600°C in a nitrogen atmosphere. The physical and mechanical properties of the composites that contained 15, 30, and 45 vol% MoSi2 were evaluated. The average four-point bend strength, fracture toughness, and Vickers hardness of the composites were in the ranges of 500-600 MPa, 3-4 MP·am1/2, and 11-13 GPa, respectively. The measured mechanical strength and hardness were very similar to the values that were predicted from the rule of mixtures. The fracture toughness of the combustion-synthesized β';-SiAlON (2.5 MPa·m1/2) was apparently enhanced by the MoSi2 particles that were added. The increase in the fracture toughness was predominately attributed to the residual thermal stress that was induced by the thermal expansion mismatch between the MoSi2 particles and the β';-SiAlON matrix. The composites showed improved electrical conductivity and oxidation resistance over monolithic β';-SiAlON. High-resolution transmission electron microscopy examination of the composites indicated that the MoSi2 was chemically well compatible with the β';-SiAlON.  相似文献   

9.
Continuously graded MoSi2-ZrO2(2Y) materials with high density (97.5% of theoretical) have been fabricated by uniaxial wet-molding, followed by hot pressing (1000°C/1 h/30 MPa) and hot isostatic pressing (1400°C/2 h/196 MPa). Their composition profiles are greatly influenced by the viscosity of mixed solutions of glycerin and ethanol used as a dispersion medium; a linear compositional gradient from MoSi2/ZrO2(2Y) 70/30 to 20/80 mol% is obtained from the solution (50/50 vol%) with a viscosity of 20 mPa s. Vickers hardness (Hv) and fracture toughness (KIC) increase from 9.7 to 12.4 GPa and from 5.1 to 12.5 MPa m1/2, respectively, with increasing ZrO2(2Y) composition.  相似文献   

10.
Dense SiC/MoSi2 nanocomposites were fabricated by reactive hot pressing the mixed powders of Mo, Si, and nano-SiC particles coated homogeneously on the surface of Si powder by polymer processing. Phase composition and microstructure were determined by methods of X-ray diffraction, scanning electron microscopy, transmission electron microscopy, and energy-dispersive spectrometry. The nanocomposites obtained consisted of MoSi2, β-SiC, less Mo5Si3, and SiO2. A uniform dispersion of nano-SiC particles was obtained in the MoSi2 matrix. The relative densities of the monolithic material and nanocomposite were above 98%. The room-temperature flexural strength of 15 vol% SiC/MoSi2 nanocomposite was 610 MPa, which increased 141% compared with that of the monolithic MoSi2. The fracture toughness of the nanocomposite exceeded that of pure MoSi2, and the 1200°C yield strength measured for the nanocomposite reached 720 MPa.  相似文献   

11.
The presence of Mo5Si3 in MoSi2 preforms hinders the reactive infiltration of aluminum. To understand the role of Mo5Si3, the kinetics of aluminum infiltration into pure Mo5Si3 is studied. Irrespective of the initial composition (MoSi2 or Mo5Si3) of the preform, the final product always contains Mo(Al,Si)2. However, the aluminum content in the two cases is different: when the preform is MoSi2, the aluminum content is 14–18 at.%, and, when the preform is Mo5Si3, the aluminum content is 25–27 at.%. The activation energy for the reactive infiltration of aluminum into the Mo5Si3 preform is ∼26 kJ/mol.  相似文献   

12.
In an earlier work, it was observed that the use of MoSi2 (up to 10 wt%) enhanced the densification and mechanical properties of TiB2. Therefore, the motivation of this study is three-fold: (a) to assess whether a small amount of MoSi2 addition can enhance wear resistance property, (b) to study whether the MoSi2 addition will influence the formation of a tribochemical layer, and (c) to correlate the wear resistance with material properties in TiB2–MoSi2 materials. In order to address these issues, a series of fretting experiments were conducted systematically by varying load (2, 5, and 10 N) at an oscillating frequency of 4 Hz and a 100 μm linear stroke, for a duration of 100 000 cycles with a cemented carbide (WC—6 wt% Co cermet) ball as a counterbody. The average coefficient of friction of the TiB2 samples varied within a narrow range (0.50–0.54), without being much affected by either the sintering additive or the load. The wear volume increased with increasing load, while the specific wear rate of all the TiB2 compositions falls within a mild wear regime (1.1–3.4 × 10−6 mm3/Nm). Based on the experimental results, it can be said that the addition of MoSi2 degrades neither the wear resistance properties nor the frictional properties of TiB2, within the investigated load regime. The microcracking-induced spalling has been found to be the dominant mechanism and, consequently, the wear volume is observed to have a linear dependency on the abrasion parameter. It is noteworthy that the tribo-oxidation as well as the formation of finer wear debris particles occurs to a limited extent.  相似文献   

13.
Crack propagation in SiC-whisker-reinforced MoSi2 was studied. In particular, the deflection angles of the cracks were examined to determine the degree to which they are affected by the whisker reinforcement. The composite studied was hot-pressed MoSi2 with 20 vol% vapor-liquid-solid β-SiC whiskers. A substantial difference was found between the deflection angles of cracks formed in the reinforced MoSi2 and those in a control sample with no whiskers, showing the process of crack deflection as an important, but not the only, toughening mechanism.  相似文献   

14.
Processing Temperature Effects on Molybdenum Disilicide   总被引:1,自引:0,他引:1  
A series of MoSi2 compacts were fabricated at increasing hot-pressing temperatures to achieve different grain sizes. The materials were evaluated by Vickers indentation fracture to determine room-temperature fracture toughness, hardness, and fracture mode. From 1500° to 1800°C, MoSi2 had a constant 67% transgranular fracture and linearly increasing grain size from 14 to 21 μm. Above 1800°C, the fracture percentage increased rapidly to 97% transgranular at 1920°C (32-μm grain size). Fracture toughness and hardness decreased slightly with increasing temperature. MoSi2 processed at 1600°C had the highest fracture toughness and hardness values of 3.6 MPa.m1/2 and 9.9 GPa, respectively. The effects of SiO2 formation from oxygen impurities in the MoSi2 starting powders and MoSi2–Mo5Si3 eutectic liquid formation were studied.  相似文献   

15.
Al2O3-MoSi2 composites were prepared by reactive hot pressing using molybdenum, aluminum, and mullite powders as precursors. The Gibbs free energy was highly negative for the composite-forming reaction, which indicated that the products were stable relative to the reactants. After the reaction, the composites had high relative density, ∼96%. Based on the composite-forming reaction, the composites should have contained 18 vol% MoSi2 in an Al2O3 matrix. Scanning electron microscopy revealed that the MoSi2 inclusions were elongated, with an average thickness of ∼5 μm and inclusion lengths that ranged from 5 to 50 μm. Average composite strength was 467 MPa, and toughness was 3.7 MPa·m1/2.  相似文献   

16.
Details of the fabrication and microstructures of hot-pressed MoSi2 reinforced–Si3N4 matrix composites were investigated as a function of MoSi2 phase size and volume fraction, and amount of MgO densification aid. No reactions were observed between MoSi2 and Si3N4 at the fabrication temperature of 1750°C. Composite microstructures varied from particle–matrix to cermet morphologies with increasing MoSi2 phase content. The MgO densification aid was present only in the Si3N4 phase. An amorphous glassy phase was observed at the MoSi2–Si3N4 phase boundaries, the extent of which decreased with decreased MgO level. No general microcracking was observed in the MoSi2–Si3N4 composites, despite the presence of a substantial thermal expansion mismatch between the MoSi2 and Si3N4 phases. The critical MoSi2 particle diameter for microcracking was calculated to be 3 μm. MoSi2 particles as large as 20 μm resulted in no composite microcracking; this indicated that significant stress relief occurred in these composites, probably because of plastic deformation of the MoSi2 phase.  相似文献   

17.
A composite consisting of 20 vol% Sic whiskers in a hot-pressed MoSi2 matrix was investigated. The composite displayed an ∼100% increase inflexural strength and a 54% improvement in fracture toughness, compared to the values measured for the matrix material. The improvements are attributed to the change in the micro-structure of the MoSi2 afforded by the presence of the whiskers.  相似文献   

18.
A ZrB2-based composite was fully densified by pressureless sintering at 1850°C with addition of 20 vol% MoSi2. The microstructure was very fine, with mean dimensions of ZrB2 grains around 2.5 μm. The four-point flexural strength in air was in excess of 500 MPa up to 1500°C.  相似文献   

19.
The influence of additions of molybdenum disilicide (MoSi2) on the microstructure and the mechanical properties of a silicon nitride (Si3N4) material, with neodymium oxide (Nd2O3) and aluminum nitride (AIN) as sintering aids, was studied. The composites, containing 5, 10, and 17.6 wt% MoSi2, were fabricated by hot pressing. All materials exhibited a similar phase composition, detected by X-ray diffractometry. Up to MoSi2 additions of 10 wt%, mechanical properties such as strength, fracture toughness, or creep at 1400°C were not affected significantly, in comparison to that of monolithic Si3N4. The oxidation resistance of the composites, in terms of weight gain, degraded. After 1000 h of oxidation at 1400° and 1450°C in air, a greater weight gain (by a factor of approximately three) was obtained, in comparison to that of the material without MoSi2. Nevertheless, after 1000 h of oxidation, the degradation in strength of the composites was considerably less severe than that of the material without MoSi2. An additional layer was formed, caused by processes at the surface of the Si3N4 material, preventing the formation of pores, cracks, or glassy-phase-rich areas, which are common features of oxidation damage in Si3N4 materials. This surface layer, containing Mo5Si3 and silicon oxynitride (Si2ON2), was the result of reactions between MoSi2, Si3N4, and the oxygen penetrating by diffusion into the material during the hightemperature treatment.  相似文献   

20.
High-pressure sintering behavior in the B6O– c -BN system was investigated using in-laboratory-synthesized B6O and commercially available c -BN powders (with an average grain size of 0.5, 3, or 6 μm). No reaction occurred between the two components under the high-pressure (4–6 GPa) and high-temperature (1500°–1800°C) conditions that have been investigated. Well-dispersed, sintered B6O– x ( c -BN) composites (where x = 0–60 vol%) of almost-full density were prepared by sintering at a pressure of 6 GPa and temperature of 1800°C for 20 min. The maximum Vickers microhardness (46 GPa) of these composites was attained by adding 40 vol% c -BN with an average grain size of 0.5 μm. The fracture toughness of these composites increased as the c -BN content increased; the maximum fracture toughness (1.5–1.8 MPa.m1/2) was observed for x = 40–60 vol%. Crack deflection along the B6O– c -BN grain boundary contributed to increasing the fracture toughness.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号