3.1. Microstructure and Mechanical Properties of Carbon-Deficient WC-10Co Alloy
Figure 3 presents the microstructure photographs of the WC-10Co alloy following sintering at 1410 °C, 1450 °C, and 1500 °C under carbon-deficient conditions. In these images, the light-colored massive grains correspond to the WC hard phase, the dark black regions represent the bonded phase, and the grayish areas in between are identified as the η phase. To analyze the composition of these three phases, we conducted EDS tests at multiple points as illustrated in
Figure 4a, with the detection results shown in
Figure 4b. EDS analysis indicates a decreasing trend in C and W element contents across the light, dark, and gray phases, respectively, while Co element content exhibits an increasing trend. This distribution aligns with the phase composition characteristics of the hard phase, bonded phase, and η phase.
Under carbon-deficient conditions, a pronounced η brittle phase was observed in the alloy across all sintering temperatures. The formation mechanism of this phase is primarily attributed to carbon deficiency. The carbon content within the WC + γ two-phase region of WC-10Co cemented carbide is restricted to a narrow range, and any deviation from this equilibrium state, particularly a deficiency in carbon, alters the concentrations of W and C in the γ phase. Furthermore, the ultrafine WC powder inherently contains a low total carbon content, coupled with its fine particle size and large specific surface area, which predisposes it to oxygen adsorption during powder mixing and forming. This adsorbed oxygen subsequently reacts with carbon during sintering, exacerbating carbon depletion and promoting the formation of the η phase under carbon-deficient conditions.
In addition, it can be observed from
Figure 3 that as the sintering temperature increases, the grain size of WC undergoes notable growth, accompanied by the aggregation and growth of the η phase.
Table 4 presents the grain size and mechanical properties of the WC-10Co alloy sintered at different temperatures under carbon-deficient conditions. The average grain size of the carbon-deficient samples ranges from 0.4 to 0.5 μm, exhibiting a positive correlation with the sintering temperature. The mechanical properties of these samples, however, are suboptimal, with the hardness varying from 1630 HV to 1700 HV and decreasing as the sintering temperature rises. Similarly, the flexural strength ranges from 940 MPa to 1300 MPa, also showing a decreasing trend with increasing temperature.
The presence of the η phase leads to inhomogeneous grain growth within the alloy, resulting in the formation of a multi-scale grain organization and laminated structure. This alteration in microstructure has a detrimental effect on the mechanical properties of the alloy, including hardness, wear resistance, and compressive strength. The η phase, being brittle, increases the brittleness of the alloy and reduces its fracture toughness.
3.2. Effect of Carbon Distribution and Sintering Temperature on the Microstructure of WC-10CoNiFe and WC-10Co Alloys
Figure 5 and
Figure 6 illustrate the microstructures of WC-10Co samples from Group A and WC-10CoNiFe samples from Group B, sintered at three distinct temperatures with four varying carbon contents of 5.44%, 5.46%, 5.48%, and 5.50%, respectively. Metallographic and SEM observations revealed that, following carbon allocation, neither graphite phase nor brittle η phase emerged in the WC-10Co and WC-10CoNiFe alloys across all four carbon contents. This indicates that the added carbon content was maintained within a reasonable interval.
Cemented carbide primarily comprises two components: the hard phase and the bonded phase. As depicted in
Figure 4a, the light-colored, lumpy regions correspond to the hard phase, while the interspersed dark-colored portions represent the bonded phase.
Figure 5 and
Figure 6 show that the WC grains in both WC-10Co and WC-10CoNiFe alloys predominantly exhibit triangular and polygonal morphologies, with a significant size disparity between the coexisting large and small grains.
The size and morphology of the grains in these alloys demonstrate pronounced changes in response to variations in sintering temperature and carbon content. An increase in sintering temperature leads to a higher proportion of relatively coarse grains in the alloy, thereby elevating the average grain size. Conversely, as carbon content increases, the proportion of fine grains in the alloy initially rises and subsequently decreases.
The pristine WC powder employed in the experiments is nearly spherical, and the morphological transformation of WC grains during sintering can be attributed to the system’s tendency towards minimizing energy and alterations in interfacial tension. Taking the Co-based bonding-phase alloys of Group A as an example, the WC/Co contact interface tends to flatten in the presence of Co. When the sintering temperature reaches the eutectic temperature, the WC grains grow selectively and become more regular in shape under the dissolution–precipitation mechanism, ultimately resulting in the formation of triangular or polygonal grains with a flattened WC/Co interface. This shape does not represent the equilibrium state of WC grains but rather their growth state under the specific preparation conditions of 5.44–5.50% carbon content and sintering temperatures of 1410 °C, 1450 °C, and 1500 °C.
The grain shape of ultrafine-grained cemented carbide is influenced by carbon content, which alters the phase composition within the alloy. When the carbon content is moderate, the phase composition remains stable, enabling the grains to maintain a more regular shape and ensuring a more uniform growth process of the WC grains. However, excessively low carbon content may lead to the appearance of a decarburization phase, while excessively high carbon content may result in the emergence of a carburization phase.
Figure 7 presents the microstructures of WC-10Co alloy sintered at different temperatures with carbon contents of 5.46% and 5.48%. It is evident that as the sintering temperature increases, the grain size of the alloy undergoes a certain degree of growth. Notably, a Coarse–Fine grain interwoven structure, comprising grain sizes of 0.14–0.40 μm and 0.51–0.90 μm, emerged when the alloy was sintered at 1450 °C with a carbon content of 5.48%. This Coarse–Fine grain interwoven structure in cemented carbide exhibits a synergistic effect between the coarse and fine grains, where the coarse grains contribute to enhanced plasticity and toughness, while the fine grains provide increased hardness and wear resistance. Consequently, this structure enables the alloy to achieve improved wear resistance while maintaining toughness, resulting in an overall enhancement of its performance. Furthermore, the Coarse–Fine grain interwoven structure effectively restricts dislocation movement, which is beneficial for improving the tensile strength of the material.
Figure 8 illustrate the microstructures of WC-10CoNiFe alloy following sintering at various temperatures with carbon contents of 5.46% and 5.48%. It is apparent that the grain size of the WC-10CoNiFe alloy increases with rising sintering temperature. Additionally, a Coarse–Fine grain interwoven structure, characterized by large grains interspersed with small grains, is observed in the metallographic images.
X-ray diffraction (XRD) analysis was conducted on WC-10Co and WC-10CoNiFe samples, with phase identification results shown in
Figure 9. Comparison with ICDD PDF cards confirmed that both alloys are dominated by the WC hard phase (card no. 73-0471), whose diffraction peaks exhibit significantly higher intensity than the binder phase. The binder phase in WC-10Co is identified as face-centered cubic cobalt (Fcc-Co, card no. 15-0806), while the WC-10CoNiFe alloy forms a (Co-Ni-Fe) solid solution (card no. 12-0736). The dominant peaks of the (Co-Ni-Fe) solid solution correspond to the (111), (200), and (220) crystal planes, consistent with Fcc-Co structure. This suggests that the (Co-Ni-Fe) solid solution retains an Fcc structure. Calculations yield lattice constants of 2.876 Å for Co and 2.880 Å for the (Co-Ni-Fe) solid solution, indicating minimal structural impact from Ni/Fe addition. Notably, the WC diffraction peaks in WC-10CoNiFe are significantly stronger than those in WC-10Co, which correlates with the binder phase’s capacity to solubilize tungsten. Under identical preparation conditions, the synergistic solubilization effect of Ni-Fe in the (Co-Ni-Fe) solid solution effectively suppresses the dissolution–precipitation of WC phase, retaining more tungsten atoms within the hard-phase lattice and thereby enhancing WC peak intensity. No secondary phases such as η-phase or graphite were detected, demonstrating that the experimental design and processing parameters avoided phase imperfections in the alloys.
3.3. The Effect of Carbon Distribution and Sintering Temperature on the Grain Size of WC-10CoNiFe and WC-10Co Alloys
The grain size of WC grains exerts a significant influence on the properties of cemented carbide. Generally, as the grain size decreases, the hardness of the cemented carbide increases. However, an excessively small grain size may also lead to a reduction in the material’s toughness. Consequently, during the preparation of ultrafine-grained cemented carbide, it is imperative to comprehensively consider the effects of grain size on both hardness and toughness to achieve cemented carbide with excellent overall performance.
Table 5 presents the grain size data for WC-10Co alloy and WC-10CoNiFe alloy prepared with varying carbon contents and sintering temperatures. Among these, the Group A WC-10Co alloy exhibits the smallest grain size of 0.294 μm at a carbon content of 5.48%, while the Group B WC-10CoNiFe alloy achieves its smallest grain size at a carbon content of 5.46%.
Figure 10 and
Figure 11 illustrate the effect curves of carbon content and sintering temperature on the grain size of the two alloys. As depicted in
Figure 12, the grain size of both alloys demonstrates a trend of decreasing and then increasing with the rise in carbon content. Additionally, the higher the sintering temperature, the larger the grain size of the alloys. Furthermore, upon comparison, it is evident that at the same sintering temperature, WC-10CoNiFe alloy possesses a smaller grain size in comparison to WC-10Co alloy.
The growth of WC grains can be primarily categorized into normal growth and abnormal growth. During the sintering process, the growth of WC grains generally adheres to the Ostwald ripening mechanism, where the size distribution of normally grown grains is uniform, and the rate of grain growth is proportional to the driving force. However, in certain scenarios, the abnormal growth of WC grains may transpire due to factors such as irregular grain morphology and uneven grain boundary migration rates.
The increase in sintering temperature accelerates atomic diffusion during the sintering process. This thermodynamic relationship is mathematically described by the Gibbs-Boltzmann distribution:
where
p represents the diffusion rate;
V0 denotes the vibrational frequency of atoms;
Es signifies the energy barrier required for diffusion;
KB is the Boltzmann constant; and
T corresponds to the absolute temperature.
In other words, the atomic diffusion rate during the sintering stage is proportional to the sintering temperature. As the sintering temperature increases, the atomic migration rate accelerates, thereby expediting grain growth. Consequently, the grain size is also proportional to the sintering temperature.
Generally, as the carbon content increases, the volume of the liquid phase in the sintered body during cemented carbide production augments, and the retention time of the liquid phase is prolonged. This provides enhanced opportunities for the dissolution and precipitation of WC grains in the liquid phase, promoting grain growth. Nevertheless, the experimental results indicate that increasing the carbon content from 5.44% to 5.48% led to a decrease in the average grain size.
In the preceding ball-milling process, the raw material grains were refined, and the ultrafine WC crystals were inherently fine. The dissolved–precipitated WC grains during sintering were even coarser than those prior to sintering. However, the increase in carbon content reduces the solubility of WC in the binder phase and impedes the dissolved–precipitated behavior of the grains within the same sintering duration. Therefore, the increase in carbon content within a certain range results in a decrease in the grain size of the final alloy.
EDS analyses were conducted on Group A WC-10Co and Group B WC-10CoNiFe alloys, as presented in
Table 6. The results reveal that elements C and W are concentrated in the hard phase, while elements Co, Ni, Fe, Cr, and V are concentrated in the binder phase. Comparing Groups A and B, with the same total proportion of the hard phase in the alloy, element W in Group B is more concentrated in the hard phase. This suggests that upon completion of sintering, WC is less dissolved in the CoNiFe multi-major-element bonding phase compared to the pure Co bonding phase. Both Groups A and B incorporated VC and Cr
3C
2 as inhibitors to suppress grain growth. With the same total proportion of inhibitors in the alloy, Group A exhibits a significantly higher concentration of Cr and V elements in the bonding phase compared to Group B. In contrast, only a relatively small amount of Cr and V elements are present in the hard phase of Group B. This indicates that the inhibitors in Group B are more concentrated at the junction of the hard phase and the bonding phase. The inhibitors hinder the growth of WC grains by adsorbing onto the grain boundaries, resulting in more refined grains in Group B when the preparation process is identical.
3.4. The Effect of Carbon Distribution and Sintering Temperature on the Mechanical Properties of WC-10CoNiFe and WC-10Co Alloys
Figure 13 illustrate the relationship between the microhardness and carbon content of WC-10CoNiFe (Group B) and WC-10Co (Group A) alloys sintered at different temperatures. For Group A alloys, as the sintering temperature increases, the hardness of the alloy exhibits a downward trend. With the increase in carbon content, the hardness of the alloy shows a declining trend. For Group B alloys, the hardness of the samples sintered at 1450 °C is generally higher than that of the samples sintered at the other two temperatures. As the carbon content increases, the hardness of the alloys sintered at 1410 °C and 1450 °C shows a tendency to rise and fall, while the hardness of the alloys sintered at 1500 °C shows a tendency to fall and rise. The hardness of cemented carbide is profoundly influenced by the hard phase. The smaller the size of the WC grain, the higher the hardness of the alloy. This is because grain refinement leads to an increase in the number of grain boundaries and dislocations, which impedes dislocation movement and thereby enhances the hardness of the alloy. As shown in
Figure 13a, the lower the sintering temperature, the smaller the grain size, and the higher the hardness. Additionally, when the carbon content in the WC grain appears to be relatively reduced, the hardness of the alloy also increases. In Group B alloys, due to the strong affinity of Fe and Ni for C, these elements are widely present in the bonding phase. During the sintering process, carbon segregation occurs in the WC, and compounds are formed in the bonding phase with Fe and Ni. Higher sintering temperatures undoubtedly exacerbate this behavior. Therefore, the hardness of the alloy sintered at 1450 °C is higher than that of the alloy sintered at 1410 °C in
Figure 13b. In contrast, the alloy sintered at 1500 °C has the lowest hardness due to the large particle size. In summary, the hardness of the alloy is determined by the combined effect of grain size and carbon content in the WC grains.
As illustrated in
Figure 14, the flexural strength of Group A WC-10Co alloy decreases with the increase in sintering temperature. As the carbon content increases, the flexural strength of the alloy sintered at 1410 °C and 1450 °C shows a decreasing trend, while the alloy sintered at 1500 °C exhibits an increasing trend. For Group B WC-10CoNiFe alloy, the flexural strength of the sample sintered at 1450 °C is higher than that of the samples sintered at the other two temperatures. With the increase in carbon content, the flexural strength of the alloy shows a decreasing trend.
The transverse rupture strength of cemented carbide is influenced by several factors, including WC grain size, carbon content, and microphase defects. As the WC grain size decreases, the number of grain boundaries increases, which hinders crack propagation and enhances the flexural strength of the cemented carbide. Therefore, the alloy’s flexural strength increases with the decrease in WC grain size. However, excessive carbon content can lead to the formation of free carbon, while insufficient carbon can result in the η-phase, both of which can reduce the alloy’s strength. The bonding strength between WC grains and the Co bonding phase is stronger than that with Ni or Fe. Thus, substituting Co with Ni or Fe can relatively reduce the alloy’s transverse rupture strength. Internal defects in the alloy can cause stress concentration, accelerating crack formation and reducing the alloy’s strength. Microphase defects are also important factors affecting the transverse rupture strength.
A comprehensive analysis of
Figure 13 and
Figure 14 reveals that WC-10Co cemented carbide achieves the highest flexural strength of 3598 MPa and hardness of 1853 HV when the carbon content is 5.48% and the sintering temperature is 1410 °C. For WC-10CoNiFe, the optimal overall performance, with a flexural strength of 2999 MPa and hardness of 1765 HV, is obtained when the carbon content is 5.46% and the sintering temperature is 1450 °C.
As shown in
Table 7, in terms of transverse rupture strength, the WC-10CoNiFe alloy exhibits a strength of 2999 MPa. Although this value is lower than that of the WC-10Co alloy (3598 MPa), which serves as an important reference in this study, it still demonstrates a significant advantage compared to the 2280 MPa of the WC-8Co NiCu alloy. This indicates that while the WC-10CoNiFe alloy maintains a certain level of transverse rupture strength, it allows room for improvement in other properties, and its strength performance is still significantly superior to some other alloy systems.
In terms of hardness performance, the WC-10CoNiFe alloy reaches 1765 HV. This value is not only significantly higher than the 1508 HV of the WC-9Co alloy but also close to the hardness value of 1783 HV reported for the WC-10Co alloy from another literature source, showcasing its competitiveness in hardness. More remarkably, the WC-10CoNiFe alloy achieves significant grain refinement, with a grain size of only 0.261 μm. This size is significantly smaller than that of the WC-10Co alloy (0.294 μm), the WC-Cr2(C,N)-Co alloy (0.445 μm), and the WC-10Co alloy (0.64 μm) reported in another literature source. Grain refinement generally contributes to improving the hardness and strength of alloys. Therefore, the WC-10CoNiFe alloy performs exceptionally well in terms of hardness, and its fine grain structure offers potential for further performance enhancement.
Furthermore, the WC-10Co alloy, serving as a reference object in this study, exhibits excellent performance in both transverse rupture strength and hardness, with values of 3598 MPa and 1853 HV, respectively. These outstanding performances not only highlight the inherent advantages of the WC-10Co alloy but also provide an important reference direction for the performance improvement of the WC-10CoNiFe alloy. By comparing the performances of the WC-10Co alloy and the WC-10CoNiFe alloy, we can more clearly see the advantages of the WC-10CoNiFe alloy in grain refinement, as well as its gap with the WC-10Co alloy in terms of transverse rupture strength and hardness. This comparison not only helps us to deeply understand the performance characteristics of the WC-10CoNiFe alloy but also provides a clear direction for future performance optimization. Specifically, by further optimizing the alloy composition and preparation process, it is expected that the WC-10CoNiFe alloy can further enhance its mechanical properties while maintaining the advantage of grain refinement.
Not only on the surface but also in the internal grain structure of the alloy, a similar trend is observed. As shown in
Figure 15, the polygonal block shape of the grains in the cross-section is essentially the same as that observed under the electron microscope. This indicates that the inherent properties of the grains themselves are not significantly altered during the fracture process. A closer examination of the fracture surfaces reveals that cracks predominantly propagate along the bonding phase, as evidenced by the smooth appearance of the fracture paths. This indicates a strong preference for intergranular fracture, where cracks follow the grain boundaries rather than penetrating through the grains themselves, suggesting that the bonding phase acts as a weak link in the alloy’s fracture behavior.
However, it is also notable that a limited number of relatively larger WC grains exhibit transgranular fracture, where cracks do penetrate through the grains. This phenomenon is particularly evident in regions where the grain size is larger, possibly due to the increased difficulty in deflecting cracks around these larger grains. The rarity of such transgranular fractures further supports the overall brittle nature of these alloys but highlights the importance of grain size distribution in influencing fracture behavior.
To gain a deeper understanding of the fracture mechanism, we examined the influence of carbon content and sintering temperature on the alloy’s properties, as revealed by our experimental data. These parameters, while not substantially altering the inherent properties of the grains, such as their crystallographic structure, significantly affect the morphology and distribution of the bonding phase, thereby influencing the fracture behavior and the final mechanical properties of the alloys.
For instance, in the case of WC-10Co cemented carbides, the alloy reaches its peak performance at a carbon content of 5.48 wt% and a sintering temperature of 1410 °C, exhibiting a maximum transverse rupture strength of 3598 MPa and a hardness of 1853 HV, with a grain size of 0.294 µm. This excellent mechanical performance is closely related to the fine grain size and uniform distribution within the alloy. As the sintering temperature increases or the carbon content deviates from this optimum, both the transverse rupture strength and hardness of the alloy decrease. This suggests that the increase in grain size and the alteration in bonding-phase properties may facilitate crack propagation along the bonding phase, thereby reducing the alloy’s overall performance.
Similarly, for WC-10CoNiFe cemented carbides, our experimental data also demonstrates the significant impact of carbon content and sintering temperature on the alloy’s properties. The alloy exhibits its optimal mechanical performance at a carbon content of 5.46 wt% and a sintering temperature of 1450 °C, with a transverse rupture strength of 2999 MPa, a hardness of 1765 HV, and a grain size of 0.261 µm. Compared to WC-10Co alloys, WC-10CoNiFe alloys demonstrate a finer grain size under optimal conditions, primarily due to the introduction of the Co-Ni-Fe multi-principal-element binder phase, which effectively inhibits grain growth.
The difference in binder phases between Group A and Group B alloys, while not leading to significant structural differences in the overall grain morphology, results in variations in their mechanical properties and microstructures, particularly in terms of grain size and hardness. This further underscore the crucial role of the binder phase’s characteristics, including its composition, distribution, and interaction with WC grains, in determining the final properties of the alloys.
In conclusion, by carefully controlling the carbon content and sintering temperature, we can significantly optimize the microstructure and mechanical properties of ultrafine-grained WC-10CoNiFe and WC-10Co cemented carbides. The experimental data provides valuable insights for a deeper understanding of the fracture behavior of these alloys and offers guidance for the development of new cemented carbide materials with superior performance.