Next Article in Journal
Influence of the Composition on the Compressive Behaviour of a Semi-Metallic Brake-Pad Material
Next Article in Special Issue
The Influence of Cryogenic Treatment on the Microstructure and Mechanical Characteristics of Aluminum Silicon Carbide Matrix Composites
Previous Article in Journal
Effect of Cellulose Nanofibrils on the Physical Properties and Frost Resistance of Pervious Concrete
Previous Article in Special Issue
Synergistic Effects of La and Y on the Microstructure and Mechanical Properties of Cast Al-Si-Cu Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Mechanical Properties of W-Al2O3 Alloy Plates Prepared by a Wet Chemical Method and Rolling Process

1
School of Materials Science and Engineering, Henan University of Science and Technology, Luoyang 471000, China
2
Henan Key Laboratory of High-Temperature Structural and Functional Materials, Henan University of Science and Technology, Luoyang 471003, China
3
Longmen Laboratory, Luoyang 471000, China
4
Scientific Research Platform Service Center of Henan Province, Zhengzhou 450003, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(22), 7910; https://doi.org/10.3390/ma15227910
Submission received: 14 October 2022 / Revised: 5 November 2022 / Accepted: 6 November 2022 / Published: 9 November 2022
(This article belongs to the Special Issue Advances in High-Performance Non-ferrous Materials)

Abstract

:
The uneven distribution and large size of the second phase weakens the effect of dispersion strengthening in ODS-W alloys. In this article, the W-Al2O3 composite powders were fabricated using a wet chemical method, resulting in a finer powder and uniformly dispersed Al2O3 particles in the tungsten-based alloy. The particle size of the pure tungsten powder is 1.05 μm and the particle size of W-0.2 wt.%Al2O3 is 727 nm. Subsequently, the W-Al2O3 alloy plates were successfully obtained by induction sintering and rolling processes. Al2O3 effectively refined grain size from powder-making to sintering. The micro-hardness of the tungsten alloy plates reached 512 HV0.2, which is 43.7% higher than that of pure tungsten plates. The nano-hardness reached 14.2 GPa, which is 24.1% higher than that of the pure tungsten plate; the compressive strength reached 2224 MPa, which is 37.2% higher than that of the pure tungsten.

1. Introduction

Tungsten or tungsten-base alloys are used in the aerospace industry to make nozzles for rocket propulsion and armor-piercing warheads in the defense military industry owing to a high melting point (3420 °C), high density, excellent thermal shock resistance, and high-temperature strength [1]. Because of the excellent sputtering resistance and low tritium retention, tungsten is also considered to be the most ideal material for use in the first wall to resist fusion plasma irradiation in the nuclear industry [2,3,4,5,6]. However, the room temperature brittleness, irradiation embrittlement, and recrystallization embrittlement of tungsten limit its wider applications [7].
The introduction of oxides and carbides into the tungsten matrix is considered to be an effective method to solve the above problems of tungsten [8,9,10,11]. Oxides such as La2O3 [12], Y2O3 [13], ZrO2 [14], and Al2O3 [15] have been widely studied. Al2O3 has been added to metal matrix composites (such as Al, Mo, and Mg) as a strengthening phase because of the excellent high-temperature properties, good chemical stability, and low cost [16,17,18,19]. The introduction of a second phase using conventional mechanical alloying methods not only results in a non-uniform distribution of the second phase but also tends to introduce impurities during the alloying process; the use of wet chemical methods was able to circumvent these problems and prepare powders to the nanoscale [20,21,22]. In our previous work, the addition of finer Al2O3 to tungsten alloys using a wet chemical method was found to significantly improve hardness and toughness [15] and increase the recrystallization temperature of the alloys [23].
This study used a wet chemical method to achieve a homogeneous distribution of the second phase Al2O3 in the alloy. The alloy powder prepared using this method also allows for effective grain refinement, resulting in significant improvements in the properties and machinability of the alloy. The tungsten alloys were unidirectionally rolled into plates, and the mechanical properties of the tungsten alloy plates were investigated by conducting nano-hardness, micro-hardness, and room-temperature compression tests. The effects of the microstructure and weaving on the mechanical properties of the alloy plates were studied by performing scanning electron microscopy (SEM), electron backscatter diffraction (EBSD), and transmission electron microscopy (TEM).

2. Materials and Methods

The W-Al2O3 composite powders were prepared from aluminum nitrate (Al(NO3)3) and ammonium para tungstate ((NH₄)10W12O41·5H₂O). Aluminum nitrate (Al(NO3)3) was dissolved in deionized water by stirring using a magnetic mixer. The obtained solution was added to a concentrated nitric acid solution to adjust the pH (pH = 0.5). The treated solution was poured into a hydrothermal reactor and placed in a drying oven for hydrothermal heating (170 °C, 15 h). Ammonium para tungstate was added to deionized water and dissolved completely. After the hydrothermal reaction, the precursor solution of Al2O3 was poured into the aqueous solution of ammonium para tungstate by applying magnetic stirring + ultrasonic shaking (20–24 h). The evenly stirred mixture was poured into a vessel and placed in a drying oven (80 °C). The dried composite powder was crushed using a pulverizer. The crushed powder was placed in a muffle furnace and calcined (500 °C, 2 h) to obtain the composite powder of WO3 and Al2O3. The composite powder was placed in a hydrogen reduction furnace for two-stage reduction (620 °C, 6 h, 920 °C, 8 h). Finally, the W-Al2O3 composite powders were obtained. The W-Al2O3 composite powders were loaded into a rubber mold and rolled into a square billet by employing cold isostatic pressing (CIP). The billets were placed in an induction sintering furnace at a temperature of 2350 °C for a sintering time of 8 h. The hydrogen atmosphere was maintained throughout the sintering process to prevent oxidation of the alloy. The sintered body with a thickness of 24 mm was heated to 1500–1600 °C in the heating furnace and the tungsten alloy plates with a thickness of 3 mm were finally obtained after several unidirectional rolling. The total thickness reduction after rolling was 87.5%. Five compositions of tungsten alloy plates with Al2O3 contents of 0, 0.2 wt.%, 0.4 wt.%, 0.6 wt.%, and 0.8 wt.% were prepared to investigate the effect of different contents of Al2O3 on the properties and microstructure of the plates.
The physical phase of the prepared alloy powder was examined by performing X-ray diffraction (Bruker-D8) equipped with a CuKα target. The voltage and current used in XRD experiments were 40 kV and 40 mA, respectively. The microscopic morphology and the distribution of Al2O3 in the W-Al2O3 composite powder were analyzed by using JEOL JSM-IT 800 SHL scanning electron microscope, and the particle size of the W-Al2O3 composite powder was tested by using an LS-909 laser particle sizer (Zhuhai OMEC Instruments Co., Ltd., Zhuhai, China). After rolling the alloy, samples (6 mm × 6 mm) were cut from the rolled surface. After mounting and polishing, the micro-hardness of the alloy plates was tested using an HV-1000 Vickers micro-hardness tester (Laizhou Huayin Test Instrument Co. Ltd., Laizhou, China) with a load of 200 g and a holding time of 15 s. Each sample was tested 20 times, then the maximum and minimum values were removed, and the remaining values were averaged, which was conducted to minimize error. The nano-indentation hardness was characterized using a G200 Nano-Indenter instrument (Keysight Technologies) equipped with a Berkovich indenter. In order to ensure the accuracy of the test data, the tungsten samples were polished using the diamond suspensions and 36 points were tested for each sample under the constant strain rate of 0.05 s−1. The pressing depth and dwell time were 1500 nm and 15 s, respectively. The maximum load in the nanoindentation tests was 870 mN. The hardness of tungsten alloys was measured by continuous stiffness method (CSM) and calculated by Oliver–Pharr method. Room-temperature compression samples (3 mm × 3 mm × 6 mm) were cut along the rolling direction of the plates, and room-temperature compression experiments were performed in a Instron 5582 Double Column Electronic Universal Testing Machine at a rate of 0.5 mm/min. Samples (3 mm × 8 mm) were cut in the tangential direction of the plates and polished in an argon-ion polishing machine (IB-19530CP, JEOL, Akishima-shi, Tokyo, Japan). After the sample was set, smoothed, and polished by using an argon-ion polisher, the weaving and microstructure of the deformed plates were observed using a JEOL JSM-IT 800 SHL scanning electron microscope (Akishima-shi, Tokyo, Japan) and an Oxford c-nano EBSD system with an EBSD scan step of 160 nm. Analysis of the interface and microstructure of the rolled sheet using FEI Talos F200X Transmission electron microscope. Figure 1 shows the schematic diagram of the examined sample.

3. Results and Discussion

3.1. Powder Characterization

Figure 2 shows the XRD patterns of the composite powders. The XRD peaks of pure tungsten and W-Al2O3 alloy powders with different compositions are the same, with no diffraction peaks of other phases except those of tungsten; this may be because the content of Al2O3 is extremely low to be detected in XRD. In addition, the intensity of the diffraction peaks is slightly different; this may be because the crystallinity of tungsten is affected by the addition of Al2O3.
Figure 3 shows the SEM images of the composite powders with different compositions. The average particle size of the pure tungsten powder is 1.05 μm, as shown in Figure 3a. The particles appear to be bonded and the particles are significantly larger than those of the composite powders with other compositions. When the content of Al2O3 is 0.2 wt.% (Figure 3b), the particles are better than those of the powders with other compositions in terms of both morphology and size; the average particle size is only 727 nm, which is 300 nm smaller than that of pure tungsten powder. The particle sizes in the other compositions do not differ considerably are shown in Figure 3c–e. However, compared with those presented in Figure 3b, these compositions contain many large Al2O3 particles; this is one of the reasons for the deterioration of the performance with the increase in Al2O3 content. Figure 3f shows the average particle size of the powders with different compositions. The standard deviations of composite powders with Al2O3 content from 0 to 0.8 wt.% are 0.2, 0.09, 0.11, 0.12, and 0.12, respectively. The D50 of composite powders with Al2O3 content from 0 to 0.8 wt.% are 1.05, 0.73, 0.83, 0.94, and 0.93 μm, respectively. The particle size of the powder prepared using the wet chemical method reaches the nanometer level. A distinct difference between the particle size of pure tungsten and the other powders with Al2O3 can be observed, indicating that adding Al2O3 effectively reduces the particle size during the hydrothermal synthesis and the reduction process. This has a crucial effect on the subsequent sintering and rolling of the tungsten alloys.

3.2. Microstructure of Sintered and Rolled Alloys

After the blanks are pressed and sintered and the sample is cut for SEM observation, pores appear in varying degrees in each composition. The reason is that tungsten has a high melting point, and it is difficult to achieve full density even with high temperature and sintering over a long time period. Therefore, the rolling process was adopted to eliminate the pores. The SEM images of the tungsten alloys are shown in Figure 4 and the grain sizes were determined by the linear intercept method. At least 200 intercepts of grains were measured and at least 8 SEM micrographs from random and spread-out locations were analyzed for each condition. As shown in Figure 4a, the grain size of sintered pure tungsten is 24 μm. Figure 4b shows the grain size of sintered W-0.2 wt.%Al2O3 alloy is only 9 μm. When the content of Al2O3 is 0.4 wt.%, 0.6 wt.%, and 0.8 wt.% (Figure 4c–e), the grain sizes are 10 μm, 14 μm, and 12 μm, respectively, corresponding to the grain sizes of the alloy powder measured previously. Al2O3 distributed at the grain boundaries during the sintering process can hinder the diffusion of grain boundaries, thus inhibiting grain growth and achieving grain refinement. Excessive addition of Al2O3 grains leads to agglomeration, and the ability to inhibit W grain growth is weakened, but the grain size is substantially reduced compared with that of pure tungsten.
As shown in Figure 5a–e, the grains of the sintered alloy are elongated after rolling. The degrees of elongation and the proportions of large-angle grain boundaries differ for plates with different compositions. Among these, the pure tungsten plate has the largest grains, maximum Feret diameter of 10.2 μm, and smallest percentage (only 39.6%) of large-angle grain boundaries. The W-0.2 wt.% Al2O3 alloy plate has a maximum Ferret diameter of 6.3 μm and a percentage of large-angle grain boundaries of 48.6%. The remaining plates of the other compositions have no significant advantage in these two aspects. Because of the large plate deformation under extreme stress, many dislocations in the form of dislocation walls and networks are widely distributed in the grain. With the movement of dislocations, a series of interactions between various dislocations cause the dislocation entanglement phenomenon. With the development of this entanglement phenomenon, the grains break into sub-grains; this is why the grains become smaller after rolling. The increase in the dislocation density and sub-grain boundaries gradually increases the deformation resistance of tungsten. The addition of Al2O3 enhances this phenomenon. Pure tungsten without Al2O3 impedes dislocations, with long strips of grains appearing after rolling. For several other compositions, aggregation, and uneven distribution due to excessive Al2O3 addition also cause such phenomena. Thus, the maximum Ferret diameter is the smallest for W-0.2 wt.% Al2O3. Figure 5f shows the phase distribution of the W-0.2 wt.% Al2O3 alloy plate. Al2O3 is distributed more uniformly in this plate compared with those in the plates with other compositions, and the Al2O3 particles are aggregated to the lowest degree. The uniform distribution of Al2O3 in the matrix can hinder the migration of grain boundaries during the sintering process, thus inhibiting grain growth. The finer the grains, the smaller the maximum Ferret diameter of the deformed grains. This is the reason for the highest percentage of large-angle grain boundaries, which can influence grain boundary strengthening, and one of the reasons for the higher hardness and compressive strength of the plates with this composition. The plates with other compositions have higher Al2O3 contents; thus, Al2O3 tends to aggregate and form large particles. This is why the mechanical properties of the W-0.2 wt.% Al2O3 alloy plates are better than the properties of the plates with other compositions.
As shown in Figure 6a–f, the grain orientation distribution function (ODF) diagram of the alloy plates with different compositions after rolling shows a distinct texture. Tungsten is known to have a body-centered cubic (BCC) structure, the texture of which is usually more obvious at the Euler angle of φ2 = 45°. Figure 6a shows the standard ODF diagram of the BCC structure at the Euler angle of φ2 = 45°. The typical {0 0 1} <1 1 0> BCC plate texture appears in the ODF diagram of pure tungsten plates, and the addition of Al2O3 transforms the original texture into a {1 1 1} <1 1 0> texture (Figure 6c–f). The strength of the texture structure varies with the change in Al2O3 content. This knowledge helps us understand the evolution of texture during the rolling process. and allows us to adjust the process parameters and the second phase addition ratio to adapt the texture depending on the application needs. The texture structure mainly depends on the motion of slip and twin systems, and the Schmidt factor diagram theory can explain the evolution of the texture structure. Figure 7 shows the Schmidt factor diagrams for pure tungsten and alloy plates, combined with the orientation distribution function diagrams. The higher Schmidt factor means that the slip system is easier to slip. This is why the {1 1 1} <1 1 0> slip system is easier to open, changing the structure of the plate after adding Al2O3.

3.3. Microstructure and Interface Feature

The samples of the rolled plates were subjected to TEM imaging and most of the Al2O3 grains present in the plates and their fineness are visible in Figure 8a. Figure 8b shows the grain size distribution of the Al2O3 particles with an average grain size of approximately 73.7 nm. Point 1 in Figure 8c represents many sub-grain boundaries around Al2O3 particles, as indicated in energy-dispersive X-ray spectroscopy (Figure 8d). This is because Al2O3, as a high-hardness ceramic material, accumulates around Al2O3 during the rolling process. The dislocations cannot pass through Al2O3 during slip. Hence, many dislocations accumulate to form sub-grain boundaries and these sub-grain boundaries strongly impede the movement of dislocations. This is an important reason why Al2O3 can enhance the strength of the plate.
The selected area electron diffraction (SAED) pattern (Figure 9c) of the tungsten matrix part was analyzed. The crystal plane spacing was 0.159 nm, 0.222 nm, and 0.129 nm, corresponding to the crystal planes of ( 2 ¯ 0 0), (0 1 ¯ 1), and ( 2 ¯   1 ¯ 1), respectively. The angular relationship of each crystal plane was verified and determined to be consistent with the (0 1 1) crystal band axis parameter of W. The SAED pattern (Figure 9d) of the Al2O3 part was analyzed. The crystalline band axis was calculated based on the crystalline plane spacing and angle, and it was consistent with the ( 2 ¯ 2 1 ¯ ) crystalline band axis corresponding to α-Al2O3. Figure 9a shows a clear interface between Al2O3 and the tungsten matrix. The crystalline plane and the lattice stripe spacing are determined using the Fourier transform, as shown in Figure 9b. The (1 1 0) crystalline plane spacing in the tungsten matrix is 2.23 Å, and the (1 0 4) crystalline plane spacing in α-Al2O3 is 2.55 Å. According to the crystalline plane mismatch degree formula, the lattice mismatch degree of the two-phase interface is 33%, and the interface is a non-co-grained relationship. As Al2O3 is a high-hardness inorganic non-metallic material with plasticity very different from that of tungsten, a metallic material, the different deformation ability of the two phases leads to the non-coordinated deformation in the rolling process. A transition layer of approximately 3-5 nm appears at the interface of the two phases after rolling, but no similar transition layer is observed in the sintered state. The Al2O3 particles are strongly bonded to the W matrix and do not separate during the deformation process (compressive stress state). Therefore, the tungsten grains on the surface layer of the Al2O3 particles undergo a large slip along the surface of the Al2O3 particles, causing the tungsten atoms to deviate from their equilibrium position and show amorphous properties. This amorphous transition layer can effectively prevent dislocation movement because amorphous grains do not have many slip systems as in crystals or dense row surfaces in which dislocations can barely move, thus strengthening the tungsten matrix.

3.4. Mechanical Properties

The micro-hardness of tungsten alloy plates with different compositions is shown in Figure 10. The micro-hardness of pure tungsten plate is 356 HV0.2, and the micro-hardness of the W-0.2 wt.% Al2O3 alloy plate is 512 HV0.2, which is an increase of 43.7%. However, the micro-hardness of the W-0.4 wt.% Al2O3 alloy plate decreases to 495 HV0.2, and with an increase in the Al2O3 content, the micro-hardness decreases further. The micro-hardness of the W-0.6 wt.% Al2O3 alloy plate is 491 HV0.2. Among all the tested alloy plates, the microhardness of the W-0.8 wt.% Al2O3 alloy plate is the lowest (472 HV0.2). The SEM images in Figure 4 of the composite powders show the low content of Al2O3 in the W-0.2 wt.% Al2O3 and Al2O3 dispersed around the tungsten grains. After sintering is performed, Al2O3 is uniformly distributed in the alloy. When subjected to external forces, dislocations are generated inside the alloy, and the fine Al2O3 particles dispersed inside the alloy play a role in pinning. Al2O3 particles dispersed in the alloy contribute to nailing dislocations and impede the movement of dislocations. However, when the content of Al2O3 increases, the fine Al2O3 particles agglomerate, thereby ineffectively inhibiting the grain growth during the sintering process and reducing the effect of nailing dislocations. The agglomeration of Al2O3 also results in the inhomogeneous microstructure of the tungsten alloys (see Figure 5), which eventually leads to the decrease in micro-hardness. Therefore, when the Al2O3 content exceeds a certain limit, the micro-hardness does not continue to increase but rather decreases to values even lower than that of pure tungsten.
The nano-indentation characterization technique allows the hardness of microscopic regions within individual grains to be measured and compared with micro-hardness to determine the variation of nano-hardness with indentation depth continuously. Hence, nano-hardness can eliminate the influence of grain boundaries on hardness and thus can directly reflect the effect of microstructure on hardness. Based on previously published work, the minimum indent spacing can be 10 times the indentation depth for a Berkovich tip [24]. In this work, the ratio of indent spacing to indent depth is 10, which can ensure that the data from the nano-indentation tests is valid. Figure 11a shows that the distance between each indentation is moderate and does not affect the test results. Figure 11b shows the nano-hardness of the plates with different compositions. The nano-hardness of pure tungsten is 11.44 GPa. After 0.2 wt.% Al2O3 is added, and the nano-hardness of the plates reaches 14.2 GPa, which is 24.1% higher than the nano-hardness of pure tungsten. After more Al2O3 is added, the nano-hardness gradually decreases, but even the lowest is 13.2% higher than that of pure tungsten. The increase in nano-hardness may be due to the interaction mechanism between the dispersed Al2O3 particles and dislocation motion, mainly the Orowan mechanism. During the downward pressure of the indenter, the dislocations caused by the load are blocked by the dispersed Al2O3 particles at the grain boundaries, causing an increase in nano-hardness [23]. However, when the Al2O3 content increases, the fine Al2O3 particles agglomerate to form large particles, and the uneven distribution reduces the hindering effect of Al2O3, resulting in a decrease in hardness.
Figure 12 presents the room-temperature compressive stress–strain curves of tungsten alloy plates with different compositions and the corresponding maximum compressive strength and failure strain are shown in Figure 13. The curves reveal that the compressive strength of pure tungsten is 1620 MPa. The highest compressive strength of 2224 MPa is achieved when 0.2 wt.% Al2O3 is added, an increase of approximately 603 MPa or 37.2% compared with that of pure tungsten. When the Al2O3 content reaches 0.8 wt.%, the compressive strength is lower than that of pure tungsten, but the deformation of the W-0.8 wt.% Al2O3 alloy plates is greater than that of pure tungsten. Pure tungsten fractures at 13.9% deformation, but the fracture of the W-0.8 wt.% Al2O3 alloy plate occurs when the deformation reaches 20%. The deformation of the W-0.2 wt.% Al2O3 alloy plates is 29%. In general, the compressive strength is influenced by the Al2O3 particles. Under stress, the Al2O3 particles in the tungsten matrix can hinder the movement of dislocations and the expansion of cracks, and at the same time, because the grains are inhibited from growing during sintering and broken into many fine sub-crystals during rolling, the resistance to crack sprouting increases, causing the superior toughness and higher compressive strength of tungsten alloy plates. The compressive strength is about 604 MPa higher than that of pure tungsten plates, and the maximum strain at fracture increases from 13.9% to 29%, improving tungsten’s ductility.
The mechanical properties of some typical tungsten alloys from the literature are summarized in Table 1 and a visual comparison of these data graphically is shown in Figure 14, indicating that W-2 wt% HfC and W-10 wt.% Ta in the green shaded area has the same or even higher failure strains at room temperature than W-0.2 wt.% Al2O3 plates, but the strength is more than 240 MPa lower than W-0.2 wt.% Al2O3 plates. The compressive strength of WMoNbTaV-1ZrO2 and Wf/Zr41.2Ti13.8Cu12.5Ni10Be22.5 is slightly lower than that of the W-0.2 wt.% Al2O3 plate, but the failure strain is much lower than that of the W-0.2 wt.% Al2O3 plates. The rest of the tungsten alloys shaded in blue are far inferior to the W-0.2 wt.% Al2O3 plates in terms of failure strain and compressive strength. However, compared to ZrO2, Al2O3 is cheaper and can be used to save costs in large batches for industrial production. Comparing the difference in compressive strength and failure strain between spin-forged and sintered W-1.5 ZrO2, as well as the significant difference in compressive strength and failure strain between W-0.2 wt.% Al2O3 sheet and sintered W-0.25 wt.% Al2O3 alloy of similar composition, it can be seen that forging and extrusion are effective methods to increase the strength and ductility of the material.

4. Conclusions

  • Among the alloy powders of different compositions, the average grain size of the W-0.2 wt.% Al2O3 powder was only 727 nm, while the grain size of the rest of the powders was about 1 μm. The XRD pattern showed no diffraction peaks of Al2O3 due to the low Al2O3 content.
  • In the sintered specimens, the average grain size of pure tungsten was 24 μm, and that of W-0.2 wt.% Al2O3 was only 9 m. The addition of Al2O3 was effective in refining the grains.
  • The maximum Ferret diameter is 10 μm for pure tungsten and 6 μm for W-0.2 wt.% Al2O3 plates, with the same size trend as in the sintered state, due to dislocation cleavage of the grains during the rolling process. The texture of the plates changes from {001} <110> to {111} <110> after the addition of Al2O3.
  • As can be observed in the TEM image, the average grain size of the Al2O3 in the plate is approximately 100 nm, with a transition layer at the interface between the two phases and a non-co-grained interface relationship.
  • Comparing the mechanical properties of pure tungsten plates with those of tungsten alloy plates of different compositions, the microhardness has increased by up to 156 HV0.2, an increase of 43%, while the nano-hardness increased by 2.75 GPa, an increase of 24.1%. The compressive strength of pure tungsten plates was 1620 MPa, but the addition of Al2O3 increased the compressive strength by up to 600 MPa, an increase of 37%. Pure tungsten plates fractured at a compression deflection of 13.9%, and at an Al2O3 content of 0.2 wt.%, after which the plates fractured only at a deflection of 29%. This proves that the doping of Al2O3 has enhanced the strength and toughness of the alloy plates.

Author Contributions

Conceptualization, C.W. and C.C.; Data curation, X.D.; Formal analysis, X.D. and M.X.; Investigation, F.M.; Methodology, C.W., Y.L. and X.W.; Project administration, Y.L., S.W., K.P. and C.C.; Resources, C.W.; Supervision, S.W., K.P., C.Z. and H.Y.; Validation, K.P. and C.Z.; Visualization, T.J.; Writing—original draft, C.W.; Writing—review & editing, K.P., C.Z., M.X., F.M., T.J., H.Y. and C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 52201118), Natural Science Foundation of Henan Province (Grant No. 222300420155), Program for Science & Technology Innovation Talents in the University of Henan Province (Grant No. 22HASTIT1006), Program for Central Plains Talents (Grant No. ZYYCYU202012172) and Ministry of Education, Singapore (Grant No. RG70/20 (2020-T1-001-023)).

Informed Consent Statement

Not applicable.

Acknowledgments

The authors greatly acknowledge the financial support from Provincial and Ministerial Co-construction of Collaborative Innovation Center for Non-ferrous Metal New Materials and Advanced Processing Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Talignani, A.; Seede, R.; Whitt, A.; Zheng, S.; Ye, J.; Karaman, I.; Kirka, M.M.; Katoh, Y.; Wang, Y.M. A review on additive manufacturing of refractory tungsten and tungsten alloys. Addit. Manuf. 2022, 58, 103009. [Google Scholar] [CrossRef]
  2. Zhou, H.-B.; Liu, Y.-L.; Jin, S.; Zhang, Y.; Luo, G.N.; Lu, G.-H. Towards suppressing H blistering by investigating the physical origin of the H–He interaction in W. Nuclear Fusion 2010, 50, 115010. [Google Scholar] [CrossRef]
  3. Xu, Q.; Sato, K.; Cao, X.Z.; Zhang, P.; Wang, B.Y.; Yoshiie, T.; Watanabe, H.; Yoshida, N. Interaction of deuterium with vacancies induced by ion irradiation in W. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2013, 315, 146–148. [Google Scholar] [CrossRef]
  4. Wang, H.; Gao, Y.; Fu, E.; Yang, T.; Xue, J.; Yan, S.; Chu, P.K.; Wang, Y. Irradiation effects on multilayered W/ZrO2 film under 4MeV Au ions. J. Nucl. Mater. 2014, 455, 86–90. [Google Scholar] [CrossRef]
  5. Miao, S.; Xie, Z.M.; Zhang, T.; Wang, X.P.; Fang, Q.F.; Liu, C.S.; Luo, G.N.; Liu, X.; Lian, Y.Y. Mechanical properties and thermal stability of rolled W-0.5wt% TiC alloys. Mater. Sci. Eng. A 2016, 671, 87–95. [Google Scholar] [CrossRef]
  6. Battabyal, M.; Schäublin, R.; Spätig, P.; Baluc, N. W–2wt.%Y2O3 composite: Microstructure and mechanical properties. Mater. Sci. Eng. A 2012, 538, 53–57. [Google Scholar] [CrossRef]
  7. Chakraborty, S.P.; Banerjee, S.; Sanyal, G.; Bhave, V.S.; Paul, B.; Sharma, I.G.; Suri, A.K. Studies on the synthesis of a Mo–30wt% W alloy by non-conventional approaches. J. Alloys Compd. 2010, 501, 211–217. [Google Scholar] [CrossRef]
  8. Veleva, L.; Oksiuta, Z.; Vogt, U.; Baluc, N. Sintering and characterization of W–Y and W–Y2O3 materials. Fusion Eng. Des. 2009, 84, 1920–1924. [Google Scholar] [CrossRef]
  9. Ryu, H.J.; Hong, S.H. Fabrication and properties of mechanically alloyed oxide-dispersed tungsten heavy alloys. Mater. Sci. Eng. A 2003, 363, 179–184. [Google Scholar] [CrossRef]
  10. Kim, Y.; Lee, K.H.; Kim, E.-P.; Cheong, D.-I.; Hong, S.H. Fabrication of high temperature oxides dispersion strengthened tungsten composites by spark plasma sintering process. Int. J. Refract. Met. Hard Mater. 2009, 27, 842–846. [Google Scholar] [CrossRef]
  11. Kim, Y.; Hong, M.-H.; Lee, S.H.; Kim, E.-P.; Lee, S.; Noh, J.-W. The effect of yttrium oxide on the sintering behavior and hardness of tungsten. Met. Mater. Int. 2006, 12, 245–248. [Google Scholar] [CrossRef]
  12. Mabuchi, M.; Okamoto, K.; Saito, N.; Asahina, T.; Igarashi, T. Deformation behavior and strengthening mechanisms at intermediate temperatures in W-La2O3. Mater. Sci. Eng. A 1997, 237, 241–249. [Google Scholar] [CrossRef]
  13. Zhao, M.; Zhou, Z.; Zhong, M.; Tan, J.; Lian, Y.; Liu, X. Thermal shock behavior of fine grained W–Y2O3 materials fabricated via two different manufacturing technologies. J. Nucl. Mater. 2016, 470, 236–243. [Google Scholar] [CrossRef]
  14. Li, Z.; Xu, L.; Wei, S.; Chen, C.; Xiao, F. Fabrication and mechanical properties of tungsten alloys reinforced with c-ZrO2 particles. J. Alloys Compd. 2018, 769, 694–705. [Google Scholar] [CrossRef]
  15. Wang, C.; Zhang, L.; Wei, S.; Pan, K.; Wu, X.; Li, Q. Preparation, microstructure, and constitutive equation of W-0.25 wt% Al2O3 alloy. Mater. Sci. Eng. A 2019, 744, 79–85. [Google Scholar] [CrossRef]
  16. Zhou, Y.; Gao, Y.; Wei, S.; Pan, K.; Hu, Y. Preparation and characterization of Mo/Al2O3 composites. Int. J. Refract. Met. Hard Mater. 2016, 54, 186–195. [Google Scholar] [CrossRef]
  17. Zabihi, M.; Toroghinejad, M.R.; Shafyei, A. Application of powder metallurgy and hot rolling processes for manufacturing aluminum/alumina composite strips. Mater. Sci. Eng. A 2013, 560, 567–574. [Google Scholar] [CrossRef]
  18. Tian, B.; Liu, P.; Song, K.; Li, Y.; Liu, Y.; Ren, F.; Su, J. Microstructure and properties at elevated temperature of a nano-Al2O3 particles dispersion-strengthened copper base composite. Mater. Sci. Eng. A 2006, 435–436, 705–710. [Google Scholar] [CrossRef]
  19. Rajkovic, V.; Bozic, D.; Jovanovic, M.T. Effects of copper and Al2O3 particles on characteristics of Cu–Al2O3 composites. Mater. Des. 2010, 31, 1962–1970. [Google Scholar] [CrossRef]
  20. Wang, J.; Gan, X.; Li, Z.; Zhou, K. Microstructure and gas sensing property of porous spherical In2O3 particles prepared by hydrothermal method. Powder Technol. 2016, 303, 138–146. [Google Scholar] [CrossRef]
  21. Visa, M. Synthesis and characterization of new zeolite materials obtained from fly ash for heavy metals removal in advanced wastewater treatment. Powder Technol. 2016, 294, 338–347. [Google Scholar] [CrossRef]
  22. Selvarajan, S.; Suganthi, A.; Rajarajan, M.; Arunprasath, K. Highly efficient BiVO4/WO3 nanocomposite towards superior photocatalytic performance. Powder Technol. 2017, 307, 203–212. [Google Scholar] [CrossRef]
  23. Wang, C.; Zhang, L.; Pan, K.; Wei, S.; Wu, X.; Li, Q. Effect of Al2O3 content and swaging on microstructure and mechanical properties of Al2O3/W alloys. Int. J. Refract. Met. Hard Mater. 2020, 86, 105082. [Google Scholar] [CrossRef]
  24. Sudharshan Phani, P.; Oliver, W.C. A critical assessment of the effect of indentation spacing on the measurement of hardness and modulus using instrumented indentation testing. Mater. Des. 2019, 164, 107563. [Google Scholar] [CrossRef]
  25. Debata, M.; Acharya, T.S.; Sengupta, P.; Acharya, P.P.; Bajpai, S.; Jayasankar, K. Effect of high energy ball milling on structure and properties of 95W-3.5Ni-1.5Fe heavy alloys. Int. J. Refract. Met. Hard Mater. 2017, 69, 170–179. [Google Scholar] [CrossRef]
  26. Zong, L.; Xu, L.; Luo, C.; Li, Z.; Zhao, Y.; Xu, Z.; Zhu, C.; Wei, S. Fabrication of nano-ZrO2 strengthened WMoNbTaV refractory high-entropy alloy by spark plasma sintering. Mater. Sci. Eng. A 2022, 843, 143113. [Google Scholar] [CrossRef]
  27. Senkov, O.N.; Wilks, G.B.; Scott, J.M.; Miracle, D.B. Mechanical properties of Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20 refractory high entropy alloys. Intermetallics 2011, 19, 698–706. [Google Scholar] [CrossRef]
  28. Han, Z.D.; Luan, H.W.; Liu, X.; Chen, N.; Li, X.Y.; Shao, Y.; Yao, K.F. Microstructures and mechanical properties of TixNbMoTaW refractory high-entropy alloys. Mater. Sci. Eng. A 2018, 712, 380–385. [Google Scholar] [CrossRef]
  29. Han, Z.D.; Chen, N.; Zhao, S.F.; Fan, L.W.; Yang, G.N.; Shao, Y.; Yao, K.F. Effect of Ti additions on mechanical properties of NbMoTaW and VNbMoTaW refractory high entropy alloys. Intermetallics 2017, 84, 153–157. [Google Scholar] [CrossRef]
  30. Yao, H.W.; Qiao, J.W.; Gao, M.C.; Hawk, J.A.; Ma, S.G.; Zhou, H.F.; Zhang, Y. NbTaV-(Ti,W) refractory high-entropy alloys: Experiments and modeling. Mater. Sci. Eng. A 2016, 674, 203–211. [Google Scholar] [CrossRef]
  31. Li, Z.; Chen, Y.; Wei, S.; Pan, K.; Shen, H.; Xu, L. Effect of rotary swaging and subsequent annealing on microstructure and mechanical properties of W-1.5ZrO2 alloys. J. Alloy. Compd. 2021, 875, 160041. [Google Scholar] [CrossRef]
  32. Wang, C.; Zhang, L.; Wei, S.; Pan, K.; Wu, X.; Li, Q. Effect of ZrO2 content on microstructure and mechanical properties of W alloys fabricated by spark plasma sintering. Int. J. Refract. Met. Hard Mater. 2019, 79, 79–89. [Google Scholar] [CrossRef]
  33. Xiao, F.; Xu, L.; Zhou, Y.; Pan, K.; Li, J.; Liu, W.; Wei, S. A hybrid microstructure design strategy achieving W-ZrO2(Y) alloy with high compressive strength and critical failure strain. J. Alloys Compd. 2017, 708, 202–212. [Google Scholar] [CrossRef]
  34. Zong, H.T.; Ma, M.Z.; Liu, L.; Zhang, X.Y.; Bai, B.W.; Yu, P.F.; Qi, L.; Jing, Q.; Li, G.; Liu, R.P. Wf/Zr41.2Ti13.8Cu12.5Ni10Be22.5 bulk metallic glass composites prepared by a new melt infiltrating method. J. Alloys Compd. 2010, 504, S106–S109. [Google Scholar] [CrossRef]
  35. Duan, X.; Huang, Y.; Liu, W.; Cai, Q.; Liu, W.; Ma, Y. Effect of Ta on the microstructure and mechanical properties of WTa alloys prepared by arc melting. Mater. Charact. 2022, 188, 111823. [Google Scholar] [CrossRef]
  36. Zhang, J.; Ma, S.; Zhu, J.; Kang, K.; Luo, G.; Wu, C.; Shen, Q.; Zhang, L. Microstructure and Compression Strength of W/HfC Composites Synthesized by Plasma Activated Sintering. Met. Mater. Int. 2018, 25, 416–424. [Google Scholar] [CrossRef]
  37. Liu, W.; Huang, Y.; Wang, Y.; Zhang, Y.; Duan, X.; Liu, W.; Ma, Y. Microstructure and mechanical properties of W-10 wt-%Ta alloys prepared by spark plasma sintering. Mater. Sci. Technol. 2022, 38, 159–168. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of test specimens cut from the tungsten alloy plates.
Figure 1. Schematic diagram of test specimens cut from the tungsten alloy plates.
Materials 15 07910 g001
Figure 2. XRD patterns of the composite powders.
Figure 2. XRD patterns of the composite powders.
Materials 15 07910 g002
Figure 3. SEM images of the composite powders: (a) pure tungsten, (b) W-0.2 wt.%Al2O3, (c) W-0.4 wt.%Al2O3, (d) W-0.6 wt.%Al2O3, (e) W-0.8 wt.%Al2O3, (f) grain size of the composite powders.
Figure 3. SEM images of the composite powders: (a) pure tungsten, (b) W-0.2 wt.%Al2O3, (c) W-0.4 wt.%Al2O3, (d) W-0.6 wt.%Al2O3, (e) W-0.8 wt.%Al2O3, (f) grain size of the composite powders.
Materials 15 07910 g003
Figure 4. SEM images of the tungsten alloys: (a) pure tungsten, (b) W-0.2 wt.% Al2O3, (c) W-0.4 wt.% Al2O3, (d) W-0.6 wt.% Al2O3, (e) W-0.8 wt.%Al2O3, (f) grain size of the tungsten alloys.
Figure 4. SEM images of the tungsten alloys: (a) pure tungsten, (b) W-0.2 wt.% Al2O3, (c) W-0.4 wt.% Al2O3, (d) W-0.6 wt.% Al2O3, (e) W-0.8 wt.%Al2O3, (f) grain size of the tungsten alloys.
Materials 15 07910 g004
Figure 5. Grain boundaries of the plates: (a) pure tungsten, (b) W-0.2 wt.% Al2O3, (c) W-0.4 wt.% Al2O3, (d) W-0.6 wt.% Al2O3, (e) W-0.8 wt.% Al2O3, (f) phase distribution of W-0.2 wt.% Al2O3.
Figure 5. Grain boundaries of the plates: (a) pure tungsten, (b) W-0.2 wt.% Al2O3, (c) W-0.4 wt.% Al2O3, (d) W-0.6 wt.% Al2O3, (e) W-0.8 wt.% Al2O3, (f) phase distribution of W-0.2 wt.% Al2O3.
Materials 15 07910 g005
Figure 6. (a) standard ODF at φ2 = 45°.ODF diagram for different composition plates, (b) pure tungsten, (c) W-0.2 wt.% Al2O3, (d) W-0.4 wt.% Al2O3, (e) W-0.6 wt.% Al2O3, (f) W-0.8 wt.% Al2O3.
Figure 6. (a) standard ODF at φ2 = 45°.ODF diagram for different composition plates, (b) pure tungsten, (c) W-0.2 wt.% Al2O3, (d) W-0.4 wt.% Al2O3, (e) W-0.6 wt.% Al2O3, (f) W-0.8 wt.% Al2O3.
Materials 15 07910 g006
Figure 7. Schmidt coefficient plots and Schmidt coefficient frequency distributions for (a,c) pure tungsten, (b,d) W-0.2 wt.% Al2O3.
Figure 7. Schmidt coefficient plots and Schmidt coefficient frequency distributions for (a,c) pure tungsten, (b,d) W-0.2 wt.% Al2O3.
Materials 15 07910 g007
Figure 8. TEM images of the rolled W-0.2 wt.% Al2O3 plates: (a) Al2O3 present in the tungsten matrix, (b) Al2O3 average grain size, (c) sub-boundary around point 1, (d) EDS of point 1.
Figure 8. TEM images of the rolled W-0.2 wt.% Al2O3 plates: (a) Al2O3 present in the tungsten matrix, (b) Al2O3 average grain size, (c) sub-boundary around point 1, (d) EDS of point 1.
Materials 15 07910 g008
Figure 9. TEM images and selected area electron diffraction patterns analysis of W-0.2 wt.% Al2O3 plate: (a) bright field image, (b) HRTEM image of the interface, (c) SEAD of tungsten, (d) SEAD of Al2O3.
Figure 9. TEM images and selected area electron diffraction patterns analysis of W-0.2 wt.% Al2O3 plate: (a) bright field image, (b) HRTEM image of the interface, (c) SEAD of tungsten, (d) SEAD of Al2O3.
Materials 15 07910 g009
Figure 10. Microhardness of rolled samples with different Al2O3 contents.
Figure 10. Microhardness of rolled samples with different Al2O3 contents.
Materials 15 07910 g010
Figure 11. (a) Nanoindentation of W-0.2 wt.% Al2O3 alloy, (b) nano-hardness of rolled samples with different Al2O3 contents.
Figure 11. (a) Nanoindentation of W-0.2 wt.% Al2O3 alloy, (b) nano-hardness of rolled samples with different Al2O3 contents.
Materials 15 07910 g011
Figure 12. Compressive stress–strain curves of rolled samples with different Al2O3 contents.
Figure 12. Compressive stress–strain curves of rolled samples with different Al2O3 contents.
Materials 15 07910 g012
Figure 13. Compressive strength and failure strain of the samples with different Al2O3 contents.
Figure 13. Compressive strength and failure strain of the samples with different Al2O3 contents.
Materials 15 07910 g013
Figure 14. Compressive strength and failure strain diagrams of some tungsten alloys at room temperature.
Figure 14. Compressive strength and failure strain diagrams of some tungsten alloys at room temperature.
Materials 15 07910 g014
Table 1. Compressive mechanical properties of typical tungsten alloys at room temperature.
Table 1. Compressive mechanical properties of typical tungsten alloys at room temperature.
AlloyStateCompressive StrengthElongationReferences
95W-3.5Ni-1.5FeSwaged1434 MPa4.6%[25]
WMoNbTaVSintered1284.6 MPa9.6%[26]
WMoNbTaV-
1ZrO2
Sintered2171.1 MPa12.7%[26]
WMoNbTaVSintered1246 MPa1.7%[27]
WMoNbTaSintered1058 MPa2.1%[27]
WMoNbTaTi0.25Sintered1109 MPa2.5%[28]
WMoNbTaTi0.5Sintered1211 MPa5.9%[28]
WMoNbTaTi0.75Sintered1304 MPa8.4%[28]
WMoNbTaTiSintered1343 MPa14.1%[29]
WMoNbTaVTiSintered1515 MPa10.6%[29]
WNbTaVSintered1530 MPa12%[30]
WNbTaVTiSintered1420 MPa20%[30]
W-1.5 ZrO2Swaged2235 MPa38.2%[31]
W-1.5 ZrO2Sintered1628 MPa21%[32]
W-1.5% ZrO2(Y)Sintered1680 MPa23.9[33]
Wf/Zr41.2Ti13.8Cu12.5Ni10Be22.5Sintered2146 MPa21.4[34]
W-40 wt.% TaSintered1630 MPa9.54%[35]
W/2 wt.% HfCSintered1980 MPa34.7%[36]
W-10 wt.% TaSintered1500 MPa29%[37]
W-0.25 wt.% Al2O3Sintered1318 MPa23%[15]
W-0.2 wt.% Al2O3Rolled2224 MPa29%This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, C.; Dong, X.; Liu, Y.; Wei, S.; Pan, K.; Zhang, C.; Xiong, M.; Mao, F.; Jiang, T.; Yu, H.; et al. Microstructure and Mechanical Properties of W-Al2O3 Alloy Plates Prepared by a Wet Chemical Method and Rolling Process. Materials 2022, 15, 7910. https://doi.org/10.3390/ma15227910

AMA Style

Wang C, Dong X, Liu Y, Wei S, Pan K, Zhang C, Xiong M, Mao F, Jiang T, Yu H, et al. Microstructure and Mechanical Properties of W-Al2O3 Alloy Plates Prepared by a Wet Chemical Method and Rolling Process. Materials. 2022; 15(22):7910. https://doi.org/10.3390/ma15227910

Chicago/Turabian Style

Wang, Changji, Xiaonan Dong, Yao Liu, Shizhong Wei, Kunming Pan, Cheng Zhang, Mei Xiong, Feng Mao, Tao Jiang, Hua Yu, and et al. 2022. "Microstructure and Mechanical Properties of W-Al2O3 Alloy Plates Prepared by a Wet Chemical Method and Rolling Process" Materials 15, no. 22: 7910. https://doi.org/10.3390/ma15227910

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