Next Article in Journal
Machinability and Tribological Performance of Advanced Alloys
Previous Article in Journal
Spectroscopical Characterization of Copper–Iron (Cu-Fe) Alloy Plasma Using LIBS, ICP-AES, and EDX
Previous Article in Special Issue
Manufacturing of Novel Nanostructured TiCrC Carbides Using Mechanical Alloying and Spark Plasma Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Production of a Reinforced Refractory Multielement Alloy via High-Energy Ball Milling and Spark Plasma Sintering

Department of Industrial Engineering, University of Trento, Via Sommarive 9, 38123 Trento, Italy
*
Author to whom correspondence should be addressed.
Metals 2023, 13(7), 1189; https://doi.org/10.3390/met13071189
Submission received: 24 May 2023 / Revised: 15 June 2023 / Accepted: 22 June 2023 / Published: 27 June 2023
(This article belongs to the Special Issue High Energy Ball Milling and Consolidation of Nanocomposite Powders)

Abstract

:
Refractory high entropy alloys have shown potential to be developed as structural materials for elevated temperature applications. In the present research, the multielement alloy Fe2TiVZrW0.5 was produced by high-energy ball milling of elemental powders in the air to promote the formation of reinforcing oxide and nitride particles followed by spark plasma sintering consolidation. The sintering temperature was optimized to achieve a full-density material that was characterized from the microstructural and mechanical points of view. Hardness and KIC were measured in the as-sintered condition as well as after thermal treatment at 1100 °C. TEM observations showed the presence of a fine distribution of ZrO2 and Ti(V)-N in the microstructure mainly constituted by the bcc Fe-V and Fe-V-W phases. The fine distribution of ceramic particles in a metallic multielement matrix is responsible for the consistent hardness and thermal stability of this alloy.

1. Introduction

Among the different families of HEAs, refractory ones show very interesting properties. Refractory high-entropy alloys (RHEAs) contain at least four of the nine refractory elements: Cr, Hf, Mo, Nb, Ta, Ti, V, W, and Zr, and have outstanding mechanical strength at extremely high temperatures. They generally have bcc microstructures constituted by elements that have this crystal structure [1,2]. High interest and a great number of studies have recently focused on this class of alloys [1,3,4,5,6,7,8]. They exhibit remarkable mechanical properties up to 1000 °C and above, and can therefore be considered a valid substitution for Ni-based superalloys as new high-temperature structural materials [9,10,11,12,13]. Refractory HEAs can possess either a single-phase bcc/B2 structure or multi-phase structures in which some intermetallic compounds are dispersed in the bcc/B2 matrix [14,15,16]. For the production of HEAs, mechanical alloying through a high-energy ball milling process can be used [11,17,18,19,20,21] since the high-energy ball–powder collisions enable the alloying effect among different powders, together with reducing the grain size. Moreover, using the powder metallurgy route, all the problems related to casting (segregations, dendritic microstructure, porosity) can be avoided. To keep the grain size as fine as possible, which ensures high mechanical strength, Spark Plasma Sintering (SPS) can be used to consolidate the powders. SPS is a field-assisted sintering technique similar to conventional hot pressing. However, instead of external heating, a pulsed direct current flows through a graphite die and two graphite punches, which contain the powder, whilst mechanical uniaxial pressure is applied. On increasing temperature under a small pressure (the minimum required to establish the electrical contact between punches and powder), densification is progressively promoted by rearrangement of the powder particles, localized deformation at the contact points, and bulk deformation of the particles [22]. SPS is a suitable method to sinter milled powders since it does not require long times at high temperatures, which could delete the refining effect obtained through the milling process. Therefore, SPS is a consolidation technique that is used for sintering milled HEA powders since it works with different classes of materials, leading to homogeneous microstructures and high densification in short times [23,24,25,26]. The application of the pressure is important to obtain a suitable densification of the compact during the sintering cycle and it was seen that it is more effective if it is applied after powder degassing when the powder particles start to plastically deform [27,28].
One of the main issues concerning materials prepared via ball milling is that they are very sensitive to process conditions (milling type, energy input, type and amount of process control agent, atmosphere, etc.). Specifically, the role of the milling atmosphere is considered of main importance. Several authors have investigated the influence of the milling atmosphere on the formation of secondary phases, on the powder refinement, and on the microstructural evolution of the milled powders [11,17,29,30,31,32,33,34]. In particular, they have studied the effect of oxygen and nitrogen uptake in the case of milling in air, considering the oxygen adsorption not as a contamination, but as a source of oxides, together with the process control agent and the oxide layers of the initial powders. The intentional oxygen uptake, leading to the formation of oxides, was first developed by Benjamin [35] in the preparation of oxide-dispersion-strengthened superalloys, using a sealed air atmosphere in an attritor mill. Other authors reported that mechanical milling in air promoted the formation of oxides, nitrides, and oxynitrides uniformly dispersed into a metal matrix [11,17,30,31,32,33,34,36]. These compounds are not necessarily harmful to the alloy system since they can contribute to the material dispersion strengthening leading to higher hardness and strength. Moreover, it should be considered that milling in the air is easier and cheaper than using a protective atmosphere; therefore, it is promising for industrial applications.
In the present investigation, powder milling in the air was used to produce a multielement alloy composed of Ti, V, Zr, W, and Fe to take advantage of this milling atmosphere to promote the formation of dispersed reinforcing oxide and nitride particles. Milled powder was spark plasma sintered and characterized in terms of microstructure, hardness, and fracture toughness.

2. Materials and Methods

Five initial high-purity powders (Fisher Scientific, Darmstadt, Germany) were used to produce the alloy Fe2TiVZrW0.5. The particle size, amount % in weight, and atomic fraction of the initial powders are summarized in Table 1. W powders used have a very small particle size to promote a fine W dispersion into the alloy since in preliminary laboratory experiments it was seen that bigger particles do not allow a complete dissolution of W, even after a prolonged milling time. For the same reason, the alloy composition was not designed in equiatomic proportion (0.2 atomic fraction of each element) but W was reduced to a lower amount, correspondingly increasing the amount of Fe.
According to the definition proposed by Yeh [37], the ideal configurational entropy of mixing of an alloy (SSS,ideal) is calculated with the Formula (1):
SSS,ideal = −RΣci × ln(ci)
with ci = atomic fraction of element i.
The SSS,ideal must be higher than 1.5 in order to have a high-entropy alloy. Calculation of S with the atomic fractions reported in Table 1 gives S = 1.511.
Another important parameter used to identify high-entropy alloys is the Valence Electron Concentration (VEC). It is used to predict the formation of solid solution phases. The empirical VEC rule was first proposed by Guo et al. [38,39] based on their experimental observation of around 20 multi-component systems such as AlxCrCuyFezMnNi, AlxCoyCrzCu0.5FevNiw, MoNbTaVW, and CoCrFeMnNi, etc. It was found that FCC is stable at VEC ≥ 8, BCC is stable at VEC ≤ 6.87, and a mixture of FCC and BCC phases exists at 6.87 ≤ VEC ≤ 8. The calculation of the alloy VEC is based on the VEC values reported in Table 1, and it is carried out through Equation (2) [1]:
V E C ( H E A ) = i 1 n C i ( V E C ) i
where Ci and (VEC)i are the atomic fraction and the VEC of element I, respectively. The present alloy has the following VEC:
VEC = ∑ci × VECi = 5.6
indicating the formation of a bcc solid solution, which is the desired phase for refractory alloys [7,33].
The production of the alloy was carried out through high-energy ball milling and SPS. Ball milling was carried out in a Fritsch Pulverisette six-ball mill (Fritsch, Idar-Oberstein, Germany) with steel vials and balls. A rotation speed of 400 rpm and a ball-to-powder ratio (BPR) of 10 were used. Milling was stopped every 5 h in order to check the powder alloying and refinement. The milling process was carried out in the air to promote oxygen uptake and subsequent oxide formation. To avoid any powder overheating, each milling cycle comprised 20 min of milling followed by a 10 min pause.
Sintering was carried out through Spark Plasma Sintering (DR. SINTER 1050® SPS (Sumitomo Coal and Mining Ltd., Tokyo, Japan) at three different temperatures (1050 °C, 1250 °C, and 1350 °C) with a holding time of 5 min at the maximum temperature and applying a pressure of 30 MPa from 650 °C up to the end of the sintering cycle. The heating rate was 50 °C/min. During sintering, the temperature was monitored through a pyrometer recording the temperature from 570 °C.
Thermal treatment was carried out in a tubular laboratory furnace under Ar atmosphere for 1, 4, and 8 h at 900 °C and 1100 °C since the majority of refractory alloys exhibit interesting mechanical properties up to the range 800–1000 °C, showing marked decreases in properties above this temperature [40,41,42]. Moreover, 1100 °C is considered the highest application temperature of state-of-the-art Ni-based superalloys [43,44], which is a reference alloy category for the RHEAs.
A JEOL IT300 scanning electron microscope (JEOL Ltd., Akishima, Japan) operating at 20 keV accelerating voltage and equipped with an energy-dispersive X-ray spectroscope (EDXS) was used to examine the powders and the sintered alloys.
X-ray diffraction (XRD) analyses were carried out to evaluate the phase composition after milling and after sintering. The measurements were carried out with an ItalStructures IPD3000 diffractometer (GNR Analytical Instruments Group, Novara, Italy) using Co Kα radiation, and an Intel CPS120 detector (Inel, Artenay, France) that measured the signal concurrently throughout an angular range of 5°–120°. The acquired diffraction spectra were elaborated by Rietveld’s method using MAUD software (Materials Analysis Using Diffraction [45].
In order to investigate the microstructure, the alloy was also observed with a high-resolution Scanning Transmission Electron Microscope (S/TEM), ThermoFisher Talos F200S (Thermo Fisher Scientific, Waltham, MA, USA) at the maximum accelerating voltage of 200 kV. The instrument is equipped with an EDXS system using two windowless Silicon Drift Detectors (SDD) with a total active area of 60 mm2. Thin foils were prepared by starting from 1 mm thick alloy slices, followed by mechanical polishing on both sides down to a thickness of about 70 μm and surface roughness of 1 μm. Three mm discs were punched from the thin foil, and dimple-grinded in the central area down to 20 μm thickness.
The electron transparence was achieved through ion milling in a Leica RES102 apparatus (Leica Microsystems, Wetzlar, Germany) applying an accelerating voltage of 7 KV and a current of 2,6 mA, at a beam incident angle of 5 degrees.
Phase transformations occurring during the heating of the milled powders were studied by means of Differential Scanning Calorimetry (DSC) using a Netzsch STA 409 Luxx apparatus (Netzsch, Selb, Germany) under a protective flux of Argon (100 mL/min) in the chamber. The DSC heating was carried out at 10 °C/min up to 1350 °C, which is the maximum sintering temperature used.
The density of the sintered discs was measured by the water displacement method, and the oxygen content was measured by a LECO TC400 machine (LECO Corporation, St. Joseph, MI, USA). Image Analysis was carried out using the software Image J.

3. Results

3.1. Milled Powders

Powders milled for 5, 10, 15, and 20 h are reported in Figure 1a–d. Powders milled up to 15 h contained large particles formed by cold welding and plastic deformation of different powder particles. The big particles disappeared after 20 h of milling, when particle fragmentation prevails on cold welding [29]. Powder milling of 20 h is shown at a high magnification in Figure 2 with its EDXS. The amount of oxygen and nitrogen adsorbed during the milling process, measured through LECO analysis, are 5.5 %wt. and 6.1 %wt., respectively. As can be seen from Figure 2, powder particles are very fine (<5 µm). X-ray diffraction analysis on the same powder (Figure 3) indicates the presence of two cubic bcc structures, namely 1 and 2 on the XRD spectrum, with a large prevalence of phase bcc 1 (95%) with respect to phase bcc 2 (5%). Phase bcc 1 has a cell size parameter a = 3.178 Å and represents the most part of the alloy, phase bcc 2 has a lower cell size a = 2.905 Å and constitutes only 5% of the alloy. No oxides were detected by XRD thus suggesting that the oxygen detected by LECO is present as a supersaturated solid solution, created by the high-energy milling process, or, as reported by other authors, they are present as amorphous phases [30,46]. There were two sources of oxygen: the initial content of the elemental powders (less than 0.1 %wt. declared by each powder producer) and the atmosphere (air) under which milling was carried out.

3.2. Sintered Samples

3.2.1. Microstructural Analysis

The microstructures of samples sintered at 1250 °C and 1350 °C are shown in Figure 4. The specimen sintered at 1050 °C is not reported because of its excessively high amount of macro-porosity. The sample sintered at 1250 °C also shows the presence of some residual porosity, indicated by arrows in Figure 4a. The amount of porosity measured through Archimedes’ method is 3.6 %vol. Instead, the alloy sintered at 1350 °C is almost full density (residual porosity evaluated through Image Analysis < 0.5 %vol.). The microstructure of both specimens, observed under a scanning electron microscope at high magnification (8000×) is very fine and constituted by a homogeneous mix of different light gray, dark gray, and white areas.
LECO analysis on sintered samples revealed an oxygen content of 5.34 %wt. and 5.22 %wt. after sintering at 1250 and 1350 °C, respectively, which are almost the same quantity measured on the milled powder.
Since the material of interest is the one sintered at 1350 °C, because it is full density, on this sample XRD analysis was carried out. The spectrum obtained is shown in Figure 5. As can be seen, there is a complex mix of phases in which only three are clearly detected: Fe-bcc, ZrO2, and TiN indicated by the numbers 1, 2 and 3 on the XRD spectrum. Other peaks were not identified using the database of known phases.
TEM analysis carried out in different points of the sample sintered at 1350 °C confirmed the very fine grain size of this alloy ranging around 1 μm, as shown in the TEM micrograph of Figure 6. Moreover, through the distribution map of elements (Figure 7) and the corresponding diffraction analysis, four different constituents were identified as documented in Figure 8, Figure 9, Figure 10, Figure 11 and Figure 12.
The four phases identified and indicated with the letters A-B-C-D in Figure 8, Figure 9, Figure 10, Figure 11 and Figure 12 are as follows:
Fe-V bcc, ZrO2, and Ti(V)-N phases are ascribed to the three phases also detected by XRD analysis. Differently, Fe-V-W results were undetected on the XRD spectrum due to the limitation of the MAUD database. The presence of ZrO2 and Ti(V)-N particles in the Fe-V phase creates a sort of composite material in which hard particles are homogeneously distributed into a soft “matrix” as can be observed in Figure 9 and Figure 10. The Fe-V system containing an amount of V up to about 20 %wt. is a soft bcc solid solution having a hardness of 200 HV, while for higher V contents a thermal treatment above 700 °C induces the formation of σ phase with a consequent hardening [47]. In the present investigation, the F-V phase contained about 15 %wt. of V, as indicated by EDXS analysis carried out on different F-V areas. One of them is reported in Figure 9.
As can be seen in Figure 10, which shows some ZrO2 particles, there was an evident twinning in these oxides. Monoclinic ZrO2 was mostly twinned on (1 1 0) and (1 0 0) planes [48]. The ZrO2 particles measured on TEM images had a mean size of about 300 nm, they were homogeneously distributed in the matrix, and their total amount, measured through Image Analysis on EDXS maps, was 18% vol. These features, together with the fine grain size of the other phases, are responsible for the high hardness of this alloy and the hardness stability at high temperatures. Zirconia is stable up to very high temperatures, about 1250 °C in working conditions [49]. Regarding the Fe-V-W phase, EDXS analysis indicates the presence of about 57 %wt. W, 33 %wt. Fe and 5 %wt. V, as shown in Figure 12b.
Moreover, in addition to the four phases (A–D), a fifth phase was identified, corresponding to spongy areas homogeneously distributed in the material, as shown in Figure 13. These areas contain the five elements constituting the alloy. Due to the complexity of the system, it was not possible to index the diffraction pattern with the structure present in the database containing these elements.

3.2.2. Hardness and Fracture Toughness

The hardness of the samples sintered at 1250 °C and 1350 °C are 1316 and 1348 HV5, respectively. Hardness maintained remarkable values even after 8 h at 1100 °C (852 HV5). Similar behavior was also observed by Sathiyamoorthi et al. [50] who found a very stable hardness in temperature in a high entropy alloy with different composition but similar microstructural features, i.e., constituted by a very fine fcc matrix reinforced by the precipitation of carbides and oxides.
Hardness values were used to calculate KIC through Equation (3) proposed by Shubert et al. which refers to the Palmqvist method [51]:
K I C = 0.0028 · H · P l i
where H = Vickers hardness in N/mm2, P = load in N, li = crack lengths in mm.
An example of hardness indentation with cracks starting at the corners of the imprint, used for the calculation of KIC, is shown in Figure 14. The KIC values found were 3.55 and 3.97 MPa m1/2 for the samples sintered at 1250 and 1350 °C, respectively. The very low values of fracture toughness are more similar to those of ceramic materials than metals due to the presence of ZrO2 and TiN and the intrinsic brittleness of the two phases that can be considered at high entropy (Fe-W-V and the spongy Fe-W-V-Ti-Zr). While, for example, Al2O3 in Al alloys contributes to significant improvements in strength, fracture toughness, stress corrosion resistance, and fatigue crack growth [29,52], in the present investigation ZrO2 formed in a matrix constituted by highly disordered phases (Fe-V-W and Fe-V-W-Ti-Zr), which were not ductile. For this reason, its contribution to the improvement of fracture toughness is counterbalanced by a growing brittleness.
Hardness and KIC values of the samples sintered at 1350 °C and heat-treated at 900 °C and 1100 °C for different times are reported in Figure 15 and Table 2, respectively.
Hardness values showed a consistent decrease only at 1100 °C, remaining anyway above 850 HV5. KIC after 8 h at 1100 °C increased to 5.90 MPa m1/2, but it was still a quite low value. KIC data presented in the review by Li [52] varied between less than 1 to more than 200 MPa m1/2 depending on the alloy composition, phase constituents, processing methods, and temperature. KIC represents a key issue to be improved for the application of these alloys in different fields, particularly in structural engineering parts.

3.3. Sintering Behavior

The sintering behavior of the milled powder was studied by analyzing different parameters recorded during the whole SPS process: voltage applied, current flowing through the sample, displacement of the SPS lower punch, and its displacement rate.
The current flowing vs. voltage applied is shown in Figure 16. This curve was smooth and continuous, indicating the absence of any oxide layer breakdown. The breakdown of the oxide layer covering the powder particles was revealed by a sudden change in slope and an evident curve discontinuity and irregularity [53]. This confirms the absence of a surface oxide layer on the milled powder particles (in line with XRD data of Figure 3) and the presence of oxygen as a supersaturated solid solution that forms oxides particles on heating (i.e., during the sintering cycle) as confirmed by XRD data of Figure 5 and by the TEM investigation.
In Figure 17 the displacement of the SPS lower punch is shown as a function of the temperature. The displacement rate is also reported. It is interesting to observe that the displacement increased up to 1280 °C where the slope started to decrease since the densification process was slowing down and going to finish. At the sintering temperature of 1250 °C the densification was not completed; therefore, residual porosity was present in the alloy sintered at this temperature, as indicated in Figure 4a. At 1350 °C the sintered sample was instead almost full density. Regarding the displacement rate reported in Figure 17, it can be observed that it showed three peaks. These points correspond to maximum densification rate points, i.e., temperatures at which the particles' accommodation, plastic deformation, phase transformations, and local sintering of the powders promote densification processes. The highest peak of the displacement rate (i.e., densification rate) was observed at 1222 °C. This peak indicates the occurrence of high-temperature plastic deformation, which is the main mechanism of consolidation and densification of this material. The temperature required to deform the powder compact is quite high because of the hardness of the milled powder related to its fine grain size, the alloying effect, and the dispersion of fine oxides formed during the heating step of the sintering process (up to this temperature).
DSC analysis carried out on the milled powder is reported in Figure 18. It shows four peaks, indicated by the numbers 1–4 on the curve. The first peak was exothermic and centered at 400 °C, the second was also exothermic and started at 740 °C. Peak 3 was the only endothermic peak as shown in the square graph at higher magnification, and it started at about 1120 °C and had its minimum at 1170 °C. The last peak (4) was exothermic and showed its maximum at 1274 °C.
Peak 1 is ascribed to the formation of TiN [32,54,55,56]. Peak 2 can be due to the precipitation of ZrO2 from the oxygen-rich supersaturated solid solution [57], which increases the hardness of the alloy slowing down the phenomena of plastic deformation of the powder particles responsible for sintering, indicated by a reduction of the displacement rate, as shown in Figure 17. A decrease in the displacement rate was observed starting from 754 °C. The endothermic peak 3 was due to the transformation of the crystal structure of ZrO2 from monoclinic to tetragonal [58] which led to a corresponding increase of ductility and reduction in volume that resulted in the displacement rate increase observed in Figure 17, which had a maximum at 1222 °C [59]. Above this temperature, the densification rate started to decrease but sintering was still ongoing (the sample sintered at 1250 °C still had some residual porosity). Lastly, the exothermic peak 4 is attributed to the sintering process of the powders at the end of the DSC test which, even without any pressure applied, were no more dispersed but consolidated into the crucible. The heat released on sintering was very small and showed its maximum at a temperature (1274 °C) which is in line with the temperature of the deviation of displacement from the linearity observed in Figure 17 (1280 °C), which indicates the consolidation and densification of the powder compact.

4. Conclusions

A multielement refractory alloy was produced through high-energy ball milling of elemental powders in an air atmosphere, followed by spark plasma sintering consolidation. After 20 h of milling, an almost homogeneous bcc structure was obtained that evolved into a multiphase system during the sintering cycle, in which reinforcing particles such as ZrO2 and TiN formed due to the oxygen and nitrogen uptake that occurred during the milling process. The sintering temperature was optimized to obtain a full-density material. The very fine multiphase microstructure of this alloy, analyzed through a TEM microscope, comprised particles of ZrO2 and TiN dispersed in a matrix made of three different phases and showed a very high hardness stable with temperature.

Author Contributions

Conceptualization, C.M., L.E. and K.N.R.S.; methodology, C.M. and G.I.; formal analysis, K.N.R.S. and L.E.; investigation, C.M. and K.N.R.S.; resources, C.M.; writing—original draft preparation, C.M., K.N.R.S. and L.E.; writing—review and editing, C.M., L.E. and G.I.; supervision, C.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gao, M.C.; Liaw, P.K.; Yeh, J.W.; Zhang, Y. High-Entropy Alloys: Fundamentals and Applications; Springer: Berlin, Germany, 2016. [Google Scholar]
  2. Liu, F.; Liaw, P.K. Recent Progress with BCC-Structured High-Entropy Alloys. Metals 2022, 12, 501. [Google Scholar] [CrossRef]
  3. Gao, X.; Wang, L.; Guo, N.; Luo, L.; Zhu, G.; Shi, C.; Su, Y.; Guo, J. Microstructure and Mechanical Properties of Multi-Phase Reinforced Hf-Mo-Nb-Ti-Zr Refractory High-Entropy Alloys. Int. J. Refract. Met. Hard Mater. 2022, 102, 105723. [Google Scholar] [CrossRef]
  4. Real, C.; Alcalá, M.D.; Trigo, I.; Fombella, I.; Córdoba, J.M. Fabrication and Characterization of FeCoNiCrMn,(Al) High Entropy Alloy Based (Ti,Ta,Nb)(C,N) Cermet. Int. J. Refract. Met. Hard Mater. 2021, 101, 105694. [Google Scholar] [CrossRef]
  5. Liu, Z.; Shi, X.; Zhang, M.; Qiao, J. Development of Refractory High Entropy Alloys with Tensile Ductility at Room Temperature. Metals 2023, 13, 329. [Google Scholar] [CrossRef]
  6. Senkov, O.; Miracle, D.; Chaput, K.; Couzinie, J. Development and exploration of refractory high entropy alloys—A review. J. Mat. Res. 2018, 33, 3092–3128. [Google Scholar] [CrossRef] [Green Version]
  7. Pickering, E.J.; Carruthers, A.W.; Barron, P.J.; Middleburgh, S.C.; Armstrong, D.E.J.; Gandy, A.S. High-Entropy Alloys for Advanced Nuclear Applications. Entropy 2021, 23, 98. [Google Scholar] [CrossRef]
  8. Tian, Y.-S.; Zhou, W.-Z.; Tan, Q.-B.; Wu, M.-X.; Qiao, S.; Zhu, G.-L.; Dong, A.-P.; Shu, D.; Sun, B.-D. A Review of Refractory High-Entropy Alloys. Trans. Nonferrous Met. Soc. China 2022, 32, 3487–3515. [Google Scholar] [CrossRef]
  9. Chen, S.; Qi, C.; Liu, J.; Zhang, J.; Wu, Y. Recent Advances in W-Containing Refractory High-Entropy Alloys—An Overview. Entropy 2022, 24, 1553. [Google Scholar] [CrossRef]
  10. Senkov, O.N.; Wilks, G.B.; Miracle, D.B.; Chuang, C.P.; Liaw, P.K. Refractory High-Entropy Alloys. Intermetallics 2010, 18, 1758–1765. [Google Scholar] [CrossRef]
  11. Li, Q.; Zhang, H.; Li, D.; Chen, Z.; Qi, Z. The Effect of Configurational Entropy on Mechanical Properties of Single BCC Structural Refractory High-Entropy Alloys Systems. Int. J. Refract. Met. Hard Mater. 2020, 93, 105370. [Google Scholar] [CrossRef]
  12. Vlasák, T.; Čížek, J.; Melikhova, O.; Lukáč, F.; Preisler, D.; Janeček, M.; Harcuba, P.; Zimina, M.; Srba, O. Thermal Stability of Microstructure of High-Entropy Alloys Based on Refractory Metals Hf, Nb, Ta, Ti, V, and Zr. Metals 2022, 12, 394. [Google Scholar] [CrossRef]
  13. Lu, S.; Li, X.; Liang, X.; Yang, W.; Chen, J. Effect of V and Ti on the Oxidation Resistance of Wmotanb Refractory High-Entropy Alloy at High Temperatures. Metals 2022, 12, 41. [Google Scholar] [CrossRef]
  14. Yurchenko, N.Y.; Stepanov, N.D.; Gridneva, A.O.; Mishunin, M.V.; Salishchev, G.A.; Zherebtsov, S.V. Effect of Cr and Zr on Phase Stability of Refractory Al-Cr-Nb-Ti-V-Zr High-Entropy Alloys. J. Alloys Compd. 2018, 757, 403–414. [Google Scholar] [CrossRef]
  15. Senkov, O.N.; Miracle, D.B.; Chaput, K.J.; Couzinie, J.P.; Yurchenko, N.Y.; Stepanov, N.D.; Gridneva, A.O.; Mishunin, M.V.; Salishchev, G.A.; Zherebtsov, S.V.; et al. Phase Stability in High Entropy Alloys: Formation of Solid-Solution Phase or Amorphous Phase. Powder Metall. 2018, 61, 115–122. [Google Scholar]
  16. Chen, H.; Kauffmann, A.; Seils, S.; Boll, T.; Liebscher, C.H.; Harding, I.; Kumar, K.S.; Szabó, D.V.; Schlabach, S.; Kauffmann-Weiss, S.; et al. Crystallographic Ordering in a Series of Al-Containing Refractory High Entropy Alloys Ta–Nb–Mo–Cr–Ti–Al. Acta Mater. 2019, 176, 123–133. [Google Scholar] [CrossRef]
  17. Moravcik, I.; Kubicek, A.; Moravcikova-Gouvea, L.; Adam, O.; Kana, V.; Pouchly, V.; Zadera, A.; Dlouhy, I. The Origins of High-Entropy Alloy Contamination Induced by Mechanical Alloying and Sintering. Metals 2020, 10, 1186. [Google Scholar] [CrossRef]
  18. Xiong, W.; Guo, A.X.Y.; Zhan, S.; Liu, C.T.; Cao, S.C. Refractory High-Entropy Alloys: A Focused Review of Preparation Methods and Properties. J. Mater. Sci. Technol. 2023, 142, 196–215. [Google Scholar] [CrossRef]
  19. Yang, Q.; Tang, Y.; Wen, Y.; Zhang, Q.; Deng, D.; Nai, X. Microstructures and Properties of CoCrCuFeNiMox High-Entropy Alloys Fabricated by Mechanical Alloying and Spark Plasma Sintering. Powder Metall. 2018, 61, 115–122. [Google Scholar] [CrossRef]
  20. Zhang, M.; Zhang, W.; Liu, Y.; Liu, B.; Wang, J. FeCoCrNiMo High-Entropy Alloys Prepared by Powder Metallurgy Processing for Diamond Tool Applications. Powder Metall. 2018, 61, 123–130. [Google Scholar] [CrossRef]
  21. Shkodich, N.; Sedegov, A.; Kuskov, K.; Busurin, S.; Scheck, Y.; Vadchenko, S.; Moskovskikh, D. Refractory High-Entropy Hftatinbzr-Based Alloys by Combined Use of Ball Milling and Spark Plasma Sintering: Effect of Milling Intensity. Metals 2020, 10, 1268. [Google Scholar] [CrossRef]
  22. Diouf, S.; Menapace, C.; Molinari, A. Study of Effect of Particle Size on Densification of Copper during Spark Plasma Sintering. Powder Metall. 2012, 55, 228–234. [Google Scholar] [CrossRef]
  23. Fu, Z.; Chen, W.; Xiao, H.; Zhou, L.; Zhu, D.; Yang, S. Fabrication and Properties of Nanocrystalline Co0.5FeNiCrTi0.5 High Entropy Alloy by MA–SPS Technique. Mater. Des. 2013, 44, 535–539. [Google Scholar] [CrossRef]
  24. Wang, C.; Ji, W.; Fu, Z. Mechanical Alloying and Spark Plasma Sintering of CoCrFeNiMnAl High-Entropy Alloy. Adv. Powder Technol. 2014, 25, 1334–1338. [Google Scholar] [CrossRef]
  25. Fang, S.; Chen, W.; Fu, Z. Microstructure and Mechanical Properties of Twinned Al0.5CrFeNiCo0.3C0.2 High Entropy Alloy Processed by Mechanical Alloying and Spark Plasma Sintering. Mater. Des. 2014, 54, 973–979. [Google Scholar] [CrossRef]
  26. Joo, S.H.; Kato, H.; Jang, M.J.; Moon, J.; Kim, E.B.; Hong, S.J.; Kim, H.S. Structure and Properties of Ultrafine-Grained CoCrFeMnNi High-Entropy Alloys Produced by Mechanical Alloying and Spark Plasma Sintering. J. Alloys Compd. 2017, 698, 591–604. [Google Scholar] [CrossRef]
  27. Diouf, S.; Menapace, C.; D’Incau, M.; Molinari, A.; Ischia, G. Spark Plasma Sintering of Cryomilled Copper Powder. Powder Metall. 2013, 56, 420–426. [Google Scholar] [CrossRef]
  28. Straffelini, G.; Nogueira, A.P.; Muterlle, P.; Menapace, C. Spark Plasma Sintering and Hot Compression Behaviour of AZ91 Mg Alloy. Mater. Sci. Technol. 2011, 27, 1582–1587. [Google Scholar] [CrossRef]
  29. Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  30. Cuevas, F.G.; Cintas, J.; Montes, J.M.; Gallardo, J.M. Powder Metallurgy On Selection of Milling Atmosphere for Production of Al-10%Ti Powders On Selection of Milling Atmosphere for Production of Al-10%Ti Powders. Powder Metall. 2009, 52, 116–123. [Google Scholar] [CrossRef]
  31. Rajendrachari, S. An Overview of High-Entropy Alloys Prepared by Mechanical Alloying Followed by the Characterization of Their Microstructure and Various Properties. Alloys 2022, 1, 116–132. [Google Scholar] [CrossRef]
  32. Lu, C.J.; Zhang, J.; Li, Z.Q. Structural Evolution of Titanium Powder during Ball Milling in Different Atmospheres. J. Alloys Compd. 2004, 381, 278–283. [Google Scholar] [CrossRef]
  33. Olier, P.; Couvrat, M.; Cayron, C.; Lochet, N.; Chaffron, L. Incidence of Mechanical Alloying Contamination on Oxides and Carbides Formation in ODS Ferritic Steels. J. Nucl. Mater. 2013, 442, S106–S111. [Google Scholar] [CrossRef]
  34. Noh, S.; Choi, B.K.; Kang, S.H.; Kim, T.K. Influence of Mechanical Alloying Atmospheres on the Microstructures and Mechanical Properties of 15Cr Ods Steels. Nucl. Eng. Technol. 2014, 46, 857–862. [Google Scholar] [CrossRef] [Green Version]
  35. Benjamin, J.S. Dispersion Strengthened Superalloys by Mechanical Alloying. Metall. Trans. 1970, 1, 2943–2951. [Google Scholar] [CrossRef]
  36. Liu, W.H.; Lu, Z.P.; He, J.Y.; Luan, J.H.; Wang, Z.J.; Liu, B.; Liu, Y.; Chen, M.W.; Liu, C.T. Ductile CoCrFeNiMox High Entropy Alloys Strengthened by Hard Intermetallic Phases. Acta Mater. 2016, 116, 332–342. [Google Scholar] [CrossRef]
  37. Yeh, J.W. Recent Progress in High-Entropy Alloys. Ann. Chim. Sci. Mater. 2006, 31, 633–648. [Google Scholar] [CrossRef]
  38. Guo, S.; Liu, C.T. Phase Stability in High Entropy Alloys: Formation of Solid-Solution Phase or Amorphous Phase. Prog. Nat. Sci. Mater. Int. 2011, 21, 433–446. [Google Scholar] [CrossRef] [Green Version]
  39. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of Valence Electron Concentration on Stability of Fcc or Bcc Phase in High Entropy Alloys Related Articles Effect of Valence Electron Concentration on Stability of Fcc or Bcc Phase in High Entropy Alloys. Addit. Inf. J. Appl. Phys. J. Homepage 2011, 109, 144503. [Google Scholar]
  40. Senkov, O.N.; Gorsse, S.; Miracle, D.B. High Temperature Strength of Refractory Complex Concentrated Alloys. Acta Mater. 2019, 175, 394–405. [Google Scholar] [CrossRef]
  41. Senkov, O.N.; Jensen, J.K.; Pilchak, A.L.; Miracle, D.B.; Fraser, H.L. Compositional Variation Effects on the Microstructure and Properties of a Refractory High-Entropy Superalloy AlMo0.5NbTa0.5TiZr. Mater. Des. 2018, 139, 498–511. [Google Scholar] [CrossRef]
  42. Srikanth, M.; Raja Annamalai, A.; Muthuchamy, A.; Jen, C.P. A Review of the Latest Developments in the Field of Refractory High-Entropy Alloys. Crystals 2021, 11, 612. [Google Scholar] [CrossRef]
  43. Atli, K.C.; Karaman, I. A Short Review on the Ultra-High Temperature Mechanical Properties of Refractory High Entropy Alloys. Front. Met. Alloy. 2023, 2, 1135826. [Google Scholar] [CrossRef]
  44. Rame, J.; Utada, S.; Bortoluci Ormastroni, L.M.; Mataveli-Suave, L.; Menou, E.; Després, L.; Kontis, P.; Cormier, J. Platinum-Containing New Generation Nickel-Based Superalloy for Single Crystalline Applications. In Proceedings of the Minerals, Metals and Materials Series, Seven Springs, PA, USA, 13–17 September 2020. [Google Scholar]
  45. Rietveld, H.M. A Profile Refinement Method for Nuclear and Magnetic Structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  46. Aoki, K.; Memezawa, A.; Masumoto, T. Amorphization of Ti-Zr Powders by Mechanical Alloying in H2, N 2 and 02 Atmospheres. Mater. Sci. Eng. 1994, 181, 1263–1267. [Google Scholar] [CrossRef]
  47. Ustinovshikov, Y.; Pushkarev, B.; Sapegina, I. Phase Transformations in Alloys of the Fe–V System. J. Alloys Compd. 2005, 398, 133–138. [Google Scholar] [CrossRef]
  48. Hayakawa, M.; Kuntani, N.; Oka, M. Structural Study on the Tetragonal to Monoclinic Transformation in Arc-Melted ZrO2-2mol.%Y2O3—I. Experimental Observations. Acta Metall. 1989, 37, 2223–2228. [Google Scholar] [CrossRef]
  49. Leib, E.W.; Pasquarelli, R.M.; Blankenburg, M.; Müller, M.; Schreyer, A.; Janssen, R.; Weller, H.; Vossmeyer, T. High-Temperature Stable Zirconia Particles Doped with Yttrium, Lanthanum, and Gadolinium. Part. Part. Syst. Charact. 2016, 33, 645–655. [Google Scholar] [CrossRef] [Green Version]
  50. Sathiyamoorthi, P.; Basu, J.; Kashyap, S.; Pradeep, K.G.; Kottada, R.S. Thermal Stability and Grain Boundary Strengthening in Ultrafine-Grained CoCrFeNi High Entropy Alloy Composite. Mater. Des. 2017, 134, 426–433. [Google Scholar] [CrossRef]
  51. Sheikh, S.; M’Saoubi, R.; Flasar, P.; Schwind, M.; Persson, T.; Yang, J.; Llanes, L. Fracture Toughness of Cemented Carbides: Testing Method and Microstructural Effects. Int. J. Refract. Met. Hard Mater. 2015, 49, 153–160. [Google Scholar] [CrossRef] [Green Version]
  52. Li, W.; Liaw, P.K.; Gao, Y. Fracture Resistance of High Entropy Alloys: A Review. Intermetallics 2018, 99, 69–83. [Google Scholar] [CrossRef]
  53. Menapace, C.; Cipolloni, G.; Hebda, M.; Ischia, G. Spark Plasma Sintering Behaviour of Copper Powders Having Different Particle Sizes and Oxygen Contents. Powder Technol. 2016, 291, 170–177. [Google Scholar] [CrossRef]
  54. Sundgren, J.-E. Formation and characterization of titanium nitride and titanium carbide films prepared by reactive sputtering. In Linkoping Stud. Sci. Technol. Diss. No. 79; Dep. Phys. Meas. Technol. Linkoping Inst. Technol.: Linkoping, Sweden, 1982; Available online: https://inis.iaea.org/collection/NCLCollectionStore/_Public/13/693/13693374.pdf (accessed on 23 June 2023).
  55. Mary, C.; Le Mogne, T.; Beaugiraud, B.; Vacher, B.; Martin, J.M.; Fouvry, S. Tribochemistry of a Ti Alloy under Fretting in Air: Evidence of Titanium Nitride Formation. Tribol. Lett. 2009, 34, 211–222. [Google Scholar] [CrossRef]
  56. Ettmayer, P.; Kolaska, H. Ti (C, N) Cermets—Metallurgy and Properties. Int. J. 1995, 13, 343–351. [Google Scholar] [CrossRef]
  57. Taguchi, M.; Sumitomo, H.; Ishibashi, R.; Aono, Y. Effect of Zirconium Oxide Addition on Mechanical Properties in Ultrafine Grained Ferritic Stainless Steels. Mater. Trans. 2008, 49, 1303–1310. [Google Scholar] [CrossRef] [Green Version]
  58. Gauna, M.R.; Conconi, M.S.; Gomez, S.; Suárez, G.; Aglietti, E.F.; Rendtorff, N.M. Monoclinic-Tetragonal Zirconia Quantification of Commercial Nanopowder Mixtures by XRD and DTA. Ceram. Silikaty 2015, 50, 318–325. [Google Scholar]
  59. Zhang, Y.; Chen, H.X.; Duan, L.; Fan, J.B.; Ni, L.; Ji, V. A Comparison Study of the Structural and Mechanical Properties of Cubic, Tetragonal, Monoclinic, and Three Orthorhombic Phases of ZrO2. J. Alloys Compd. 2018, 749, 283–292. [Google Scholar] [CrossRef]
Figure 1. Fe2TiVZrW0.5 powder high-energy milled for 5 h (a), 10 h (b), 15 h (c), and 20 h (d).
Figure 1. Fe2TiVZrW0.5 powder high-energy milled for 5 h (a), 10 h (b), 15 h (c), and 20 h (d).
Metals 13 01189 g001
Figure 2. Powders milled for 20 h with EDXS analysis result taken in the area indicated on the micrograph.
Figure 2. Powders milled for 20 h with EDXS analysis result taken in the area indicated on the micrograph.
Metals 13 01189 g002
Figure 3. X-Ray diffraction pattern of 20 h milled powder.
Figure 3. X-Ray diffraction pattern of 20 h milled powder.
Metals 13 01189 g003
Figure 4. Microstructure of Fe2TiVZrW0.5 sintered at 1250 °C (a) and 1350 °C (b).
Figure 4. Microstructure of Fe2TiVZrW0.5 sintered at 1250 °C (a) and 1350 °C (b).
Metals 13 01189 g004
Figure 5. XRD spectrum of Fe2TiVZrW0.5 sintered at 1350 °C.
Figure 5. XRD spectrum of Fe2TiVZrW0.5 sintered at 1350 °C.
Metals 13 01189 g005
Figure 6. TEM micrograph of sample sintered at 1350 °C.
Figure 6. TEM micrograph of sample sintered at 1350 °C.
Metals 13 01189 g006
Figure 7. Distribution map of elements.
Figure 7. Distribution map of elements.
Metals 13 01189 g007
Figure 8. Identification of area A containing Fe and V and having a bcc structure (Fe bcc ZA 1 0 0).
Figure 8. Identification of area A containing Fe and V and having a bcc structure (Fe bcc ZA 1 0 0).
Metals 13 01189 g008
Figure 9. EDXS analysis carried out on an F-V bcc area.
Figure 9. EDXS analysis carried out on an F-V bcc area.
Metals 13 01189 g009
Figure 10. Identification of B particles, indicated by red dots, as monoclinic ZrO2. * is a B, referring to the phase B.
Figure 10. Identification of B particles, indicated by red dots, as monoclinic ZrO2. * is a B, referring to the phase B.
Metals 13 01189 g010
Figure 11. Identification of C particles, indicated by red dots, as cubic Ti(V)-N.
Figure 11. Identification of C particles, indicated by red dots, as cubic Ti(V)-N.
Metals 13 01189 g011
Figure 12. (a) D areas containing Fe-W-V (b) EDXS analysis on one of these areas.
Figure 12. (a) D areas containing Fe-W-V (b) EDXS analysis on one of these areas.
Metals 13 01189 g012
Figure 13. EDXS analysis of a spongy area containing all the alloy constituent elements.
Figure 13. EDXS analysis of a spongy area containing all the alloy constituent elements.
Metals 13 01189 g013
Figure 14. Micrographs of a Vickers indentation and induced cracks at the corners.
Figure 14. Micrographs of a Vickers indentation and induced cracks at the corners.
Metals 13 01189 g014
Figure 15. Hardness values after heat treatments at 900 °C and 1100 °C.
Figure 15. Hardness values after heat treatments at 900 °C and 1100 °C.
Metals 13 01189 g015
Figure 16. Current vs. voltage during the sintering cycle.
Figure 16. Current vs. voltage during the sintering cycle.
Metals 13 01189 g016
Figure 17. Displacement and displacement rate vs. temperature during the sintering cycle.
Figure 17. Displacement and displacement rate vs. temperature during the sintering cycle.
Metals 13 01189 g017
Figure 18. DSC curve of the milled powder.
Figure 18. DSC curve of the milled powder.
Metals 13 01189 g018
Table 1. Initial powders.
Table 1. Initial powders.
ElementParticle Size (μm)Purity Level
(%)
Weight
(wt.%)
Atomic Fraction (-)VEC
(-)
Fe<7499.0028.40.358
Ti<14999.9012.20.184
V<4499.5013.00.185
Zr<4499.5023.10.184
W0.6–0.999.9523.40.16
Table 2. KIC values of samples heat treated at 900 °C and 1100 °C.
Table 2. KIC values of samples heat treated at 900 °C and 1100 °C.
KIC Measured through Palmqvist Method
MPa m1/2
treatment timeT = 900 °CT = 1100 °C
1 h4.105.05
4 h4.685.71
8 h4.875.90
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Menapace, C.; Shaik, K.N.R.; Emanuelli, L.; Ischia, G. Production of a Reinforced Refractory Multielement Alloy via High-Energy Ball Milling and Spark Plasma Sintering. Metals 2023, 13, 1189. https://doi.org/10.3390/met13071189

AMA Style

Menapace C, Shaik KNR, Emanuelli L, Ischia G. Production of a Reinforced Refractory Multielement Alloy via High-Energy Ball Milling and Spark Plasma Sintering. Metals. 2023; 13(7):1189. https://doi.org/10.3390/met13071189

Chicago/Turabian Style

Menapace, Cinzia, Khaja Naib Rasool Shaik, Lorena Emanuelli, and Gloria Ischia. 2023. "Production of a Reinforced Refractory Multielement Alloy via High-Energy Ball Milling and Spark Plasma Sintering" Metals 13, no. 7: 1189. https://doi.org/10.3390/met13071189

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