Next Article in Journal
High Manganese TWIP Steel with Increased Corrosion Resistance
Previous Article in Journal
Corrosion Behavior of Mg-xGd-1Zn-0.4Zr Alloys with Different Gd Additions for Biomedical Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication, Microstructure, and Physico-Mechanical Properties of Fe–Cr–Ni–Mo–W High-Entropy Alloys from Elemental Powders

by
Alexander Yurievich Ivannikov
*,
Ivan Konstantinovich Grebennikov
,
Yulia Alexandrovna Klychevskikh
,
Anna Vladimirovna Mikhailova
,
Konstantin Victorovich Sergienko
,
Mikhail Alexandrovich Kaplan
,
Anton Sergeevich Lysenkov
and
Mikhail Anatolievich Sevostyanov
Baikov Institute of Metallurgy and Material Science, Russian Academy of Sciences, 119334 Moscow, Russia
*
Author to whom correspondence should be addressed.
Metals 2022, 12(10), 1764; https://doi.org/10.3390/met12101764
Submission received: 26 September 2022 / Revised: 10 October 2022 / Accepted: 17 October 2022 / Published: 20 October 2022
(This article belongs to the Section Powder Metallurgy)

Abstract

:
In this work, 35Fe30Cr20Ni10Mo5W (wt.%) and 30Fe30Cr20Ni10Mo10W (wt.%) high-entropy alloys were fabricated using a powder metallurgy route. Powder mixtures for a hot-pressure process were obtained by the mixing and mechanical alloying of elemental powders. Mechanical alloying was carried out for 1, 2.5, 5, and 10 h. X-ray phase analysis of the powder mixtures showed that with increasing time of mechanical alloying, Face-Centered Cubic (FCC), Body-Centered Cubic (BCC), and nickel–iron intermetallic phases were formed in the structure, and the volume content of molybdenum and tungsten decreased. The hot-pressing was carried out at a pressure of 30 MPa and a temperature of 1200 °C for 30 min. The maximum densities of 8.14 ± 0.02 and 8.40 ± 0.01 g/cm3 and compressive strengths of 2430 ± 30 MPa and 2460 ± 35 MPa for consolidated materials were achieved using powder mixtures after 10 h of mechanical milling, for compositions with 5 wt.% W and 10 wt.% W, respectively. The workpieces fabricated with a pressure-assisted sintering process from milled powders were found to consist of FCC, BCC, and sigma phases.

1. Introduction

In recent years, material scientists and engineers worldwide have paid increased attention to a new class of high-entropy alloys (HEAs) due to their unique compositions, microstructure, and properties, such as excellent strength, fracture toughness, high thermal stability, superior corrosion, and wear resistance [1,2,3,4]. Unlike traditional alloys, HEAs contain five or more elements with concentrations between 5 and 35 at%. The high configurational entropy of these first-generation HEAs results in stable solid solutions, such as face-centered cubic (FCC), body-centered cubic (BCC), and hexagonal close-packed (HCP) structures instead of ordered and intermetallic compounds [4,5].
Historically, the vacuum remelting process was used to produce the Cantor alloy [2]. Nowadays, HEAs are produced through traditional metallurgical approaches, which are also used in the fabrication of commercial alloys [4,5,6,7]. These processing routes can be classified on the basis of the type of mixing procedure into three categories: from gas phase (spraying), from liquid phase (induction and arc melting), and from solid state (powder metallurgy). Routes of these types have some advantages and disadvantages.
Gas phase methods are widely used in producing coatings [8,9,10]. Such coatings are used in several areas, such as information storage, superconductivity, power transmission, high-speed switching, high-speed signal transmission for computing, and high-speed trains with magnetic levitation. Many of these applications require superior mechanical properties at various operating temperatures, and high-entropy alloys show great potential in these areas.
Most HEAs are produced through melting methods, which have certain flaws, including coarse grain size with elemental segregation and dendritic structures during cooling. For this reason, the obtained multicomponent alloys have poor properties. Therefore, these alloys need further processing. Traditionally, long-time thermomechanical treatments were used to achieve homogeneous composition and recrystallized microstructures [11,12,13,14,15].
Powder metallurgy (PM) processes implement solid-state procedures of HEA manufacturing [16,17]. These typical bottom-up processes help to avoid or reduce metal removal processes and thereby significantly reduce manufacturing costs. The following methods are widely used to produce powders: mechanical alloying (MA) [18,19,20,21,22], self-propagating high-temperature synthesis [23,24,25], and gas and plasma atomization [26,27].
It is worth noting that the method of mixing elemental powders, which is widely used in the fabrication of medium-entropy alloys, is not used for high-entropy alloys [28].
To consolidate HEA powders, conventional sintering methods, which include hot-pressing vacuum sintering [4,29,30], are used. Other modern methods for consolidating HEA powders are, for example, microwave sintering, hot isostatic pressing, and spark plasma sintering (SPS) [31,32,33,34]. The powders can also be used for thermal spraying on coatings of specimens and parts [35,36,37]. Spherical powders from HEAs are used in the additive manufacturing process [38,39,40].
The use of powder metallurgy has extended the concept of high-entropy alloys. Materials with two- and three-phase structures are considered to be high-entropy alloys of the second generation [4,14]. Such alloys demonstrate unique mechanical properties and are promising for use in tribological applications since they have high wear resistance.
Powder metallurgy has found particular application in the hardening of the first-generation HEAs. FCC-type HEAs, such as CoCrFeNiMn, exhibit outstanding ductility and fracture toughness, even down to the liquid nitrogen temperature. However, they are relatively weak in strength for structural applications. Very recently, J.Y. He et al. [41] found that nanosized coherent reinforcing phase could be obtained and extraordinary balanced tensile properties at room temperature achieved mainly through precipitation strengthening. For this reason, W.H. Liu et al. [42] added Mo into the CoCrFeNi matrix. The precipitation of hard σ and μ intermetallic compounds tremendously strengthened the CoCrFeNiMo0.3 HEA, but without causing serious embrittlement. The addition of refractory elements into the composition is being considered to increase the strength of high-entropy alloys. Molybdenum is used to harden the solid solution, and tungsten and niobium are used similarly [4,14,41,42,43,44].
A separate direction in the field of HEAs is the development of cobalt-free alloys due to their high price [45]. Therefore, the search for variants of compositions with refractory components has become relevant [45,46]. These types of compositions are characterized by increased strength and wear resistance at different temperatures.
In this study, the alloying additions of Mo and W were added into the FeCrNi matrix. Mo has a moderately large atomic size for both solid solution and precipitation hardening [42]. Furthermore, it is known that Mo and W are capable of forming hard intermetallic phases with Fe, Cr, and Ni elements [41,42]. Moreover, despite their poor plasticity in the bulk state at ambient temperatures, Mo- and W-based intermetallic compounds usually offer an attractive combination of physical and mechanical properties, such as good thermal stability, oxidation, and wear resistance. Thus, Mo and W were selected as promising elements to add to the FeCrNi HEA for precipitation hardening.
The aim of this work is to evaluate the effect of the powder mixture preparation method and concentration of the W powders on the structure and physical and mechanical properties of Fe–Cr–Ni–Mo–W high-entropy alloys obtained by hot-pressing.

2. Materials and Methods

Elemental powders (Figure 1) with a purity of more than 99.9% were used to obtain powder mixtures. The iron and nickel powder had a particle size of less than 160 microns. Chromium, molybdenum, and tungsten powders had a particle size of less than 15 microns.
Two compositions were selected for preparation of powder mixtures: 35Fe30Cr20Ni10Mo5W (wt.%) and 30Fe30Cr20Ni10Mo10W (wt.%).
The mechanical mixture of powders for hot-pressing was obtained by intensive mixing in a turbulent mixer. The mixing time was 5 h.
The second method of preparing powder mixtures for hot-pressing was based on the use of a Retsch PM400 planetary mill (Retsch GmbH, Haan, Germany) with 250 mL steel cups. Hardened chromium steel balls with a diameter of 5.5 mm were used for milling with a ball-to-powder weight ratio of 5:1. The milling was carried out at rotational speed of 300 rpm and in an argon atmosphere to prevent oxidation. Mechanical alloying was performed for 1, 2.5, 5, and 10 h.
The bulk density was determined according to ISO 3923-1:2018, and the tap density was determined on a Logan TAP-2S/2SP in accordance with ISO 3953:2011.
The prepared powder mixtures were pressed at room temperature on a hydraulic press IP-1000-1 (Tochmashpribor Ltd., Ivanovo, Russia) in a cylindrical mold with a diameter of 25 mm at a pressure of 100 MPa. The obtained green bodies were wrapped in graphite foil, placed in a graphite mold, and pressed on a HP20-3560-20 machine (Thermal Technology Inc., Santa Rosa, CA, USA) at a pressure of 30 MPa and a temperature of 1200 °C for 30 min. The obtained consolidated powder samples were freed from graphite lining.
Cylindrical samples with a diameter of 4 mm and a height of 10 mm for mechanical tests were made on an electric erosion machine. Density, hardness, and structure were studied on separately manufactured samples with a height of 10 mm and a diameter of 10 mm.
Phase formation of mechanically alloyed powders and sintered alloys after hot-pressing was analyzed using a DRON 4 diffractometer with Co-Kα radiation (XRD, Saint Petersburg, Russia). Powder morphology, microstructure evolution, and chemical composition analysis were characterized using a Hitachi-4700 scanning electron microscope (SEM) (Scanning Electron Microscope Hitachi, Ltd., Tokyo, Japan) with energy-dispersive X-ray spectroscopy (EDS).
For all samples, the densities of sintered bulk alloys were measured with the Archimedes method.
Determination of compressive strength at room temperature was performed on an Instron 3382 (ITW Limited, High Wycombe, UK) with a loading rate of 1 mm/min. A hardness tester TR 5006 (Tochmashpribor Ltd., Ivanovo, Russia) was used to evaluate hardness according to the Rockwell method (C scale). At least 10 measurements were performed to calculate the average value.

3. Result and Discussion

3.1. Analysis of Powder Mixtures

The prepared powders were analyzed to determine bulk density and tap density. Figure 2 shows the effect of the powder preparation method on the bulk and tap density.
The minimum values of bulk and tap density were measured for mechanically mixed powders. The low values of the bulk density of mechanically mixed powders are related to the presence of large (between 60 and 160 microns) particles of iron and nickel powders. Increasing the milling time of the initial powder mixture contributes to increasing the bulk and tap density. This occurs due to the grinding of the initial powders and alignment of the average particle size of the resulting mixture. High values of bulk and tap density are important technological parameters for powder mixtures in powder metallurgy.
The bulk density and tap density of the powders after 1 h of mechanical alloying are low due to the lack of grinding of large particles of iron and nickel powders. Therefore, the powders after 1 h of mechanical mixing were not used for hot-pressing.
Figure 3 shows the morphology of powders obtained after mechanical alloying for 5 h. More than 90% of the particles had particle sizes of less than 10 microns. Agglomerates were also detected in the powder mixture.
These agglomerates (Figure 4) were obtained during the milling and cold welding of elemental powders. The morphologies of the agglomerates’ surfaces are shown in Figure 4, and the elemental composition data are presented in Figure 5.
The analysis of diffractograms (Figure 6) showed the presence of molybdenum and tungsten in the powders after mechanical alloying, as well as the BCC phase. The formation of the Ni3Fe intermetallic compound was also detected.
With the increasing milling time, the intensity of reflexes responsible for the identification of molybdenum and tungsten in the powder mixtures decreased. This indicates the grinding of molybdenum and tungsten powders and their introduction into the solid solution.
Ten hours of mechanical alloying led to the formation of the FCC phase in powders.
The obtained powder mixtures were suitable for green bodies and subsequent hot-pressing. Analysis of the density of green bodies showed that the highest density was achieved for workpieces from powders after 10 h of milling (Figure 2).

3.2. Structure and Phase Composition

The crystal structure of the hot-pressed samples was analyzed on a diffractometer. As the XRD results show in Figure 7, the diffractograms of the specimens after pressure-assisted sintering correspond to the FCC, BCC, and sigma phases. The hot-pressed specimens from the mixture of elemental powders detected the peaks showing the presence of molybdenum and tungsten in the structure of powders. The Ni3Fe intermetallic compound was revealed in powders after mechanical alloying, but no presence of peaks identifying Ni3Fe was detected in the specimens after pressure-assisted sintering.
The samples of the 35Fe30Cr20Ni10Mo5W alloy from the milled powders showed that with the increasing milling time, the intensity of the peak responsible for the content of the FCC phase and sigma phase increased while the content of the BCC phase decreased (Figure 7a).
A similar trend was observed for the 30Fe30Cr20Ni10Mo10W alloy samples, but for the powder sample after 2.5 h of milling, the intensity of the peak characterizing the BCC phase was greater than that of the peak characterizing the FCC phase (Figure 7b). This, in turn, affected the mechanical strength of the sample.
The microstructure of the treated materials is shown in Figure 8 and Figure 9. A finer grain size can be observed in the case of specimens from milled powders compared to the specimens from a mixture of elemental powders (Figure 8 and Figure 9a).
As shown in Figure 8 and Figure 9, the distribution of the elements Fe, Cr, and Ni could be clearly determined in the treated materials. The element distribution of Mo and W could be clearly determined only for the specimens sintered under hot pressure from a mixture of powders. The agglomerates of Mo and W powders were obtained during mixing (Figure 8 and Figure 9a). After pressure-assisted sintering, these agglomerates were fixed in the structure of the sintered materials. The specimens sintered from milled powders had uniform distribution of Mo and W.

3.3. Physical and Mechanical Properties of Consolidated Samples

Data on density, compressive strength, and hardness of consolidated samples are presented in Table 1.
High density is observed in the specimens from a mixture of elemental powders after hot-pressing, as well as in the specimens from powders after 10 h of milling.
Specimens from powders after 10 h of milling had the maximum value of physical density due to the high value of density of the green bodies and the phase composition of the powders (Figure 3).
Specimens after mechanical alloying for 2.5 and 5 h have low physical density due to residual porosity.
Figure 10a shows typical compressive stress–strain curves of the bulks produced by sintering under hot pressure from a mixture of elemental powders and milled powders at room temperature. With increasing milling time, the compressive strength increased substantially. Specimens made from powders after 10 h of milling had the maximum strain value due to the high value of the physical density and, consequently, low residual porosity.
Figure 10b shows a similar trend, but specimens that had been sintered under hot pressure from 2.5-h-milled powders had the minimum strength value. The low strength value is defined by the phase composition of the workpieces produced by pressure-assisted sintering (Figure 7b). This sample had low peaks of the FCC phase, high peaks of the BCC phase, and low density, hence the high value of the residual porosity.
The results of the analysis of the structure and phase composition indicated the formation of a three-phase structure in the bulks made with the pressure-assisted sintering process from milled powders. The FCC, BCC, and sigma phases were fixed in the structure. Moreover, the Cr-rich sigma phase was found to be very common in Cr-containing HEAs in sintered powder bulks [47,48], as well as in heat-treated cast alloys [49]. The presence of the sigma phase in the structure of the specimens determines the high values of strength, hardness, and low strain before fracture. Compressive strength values of more than 2400 MPa for specimens with 5 and 10wt.%W indicate the achievement of high mechanical strength, which corresponds to the best strength values attained for high-entropy alloys [14,44,46].
It is worth noting that the article [44] explored the preparation time of powders for 60 h to achieve a single-phase structure, and in this work, the maximum milling time was 10 h. Therefore, it is logical to analyze the effect of a long milling time on the structure and mechanical strength of the specimens. To increase the plasticity of the obtained HEAs, it is rational to carry out heat treatment and to investigate the effect of the type of heat treatment on the structure and mechanical properties of the HEAs.
The production of a multicomponent alloy from mechanically mixed powders resulted in the formation of a medium-entropy alloy that was disperse-hardened with molybdenum and tungsten powders. This means that this type of process route cannot be used to make HEAs with refractory elements.

4. Conclusions

Mechanical alloying helps to increase the bulk and tap density, as well as the density of the green bodies from powder mixtures.
Increasing the time of mechanical alloying led to a decrease in the molybdenum and tungsten peaks and also contributed to the formation of the BCC and FCC phases in the powders; the Ni3Fe intermetallic compound was also present in the powders.
In the process of mechanical alloying, agglomerates are formed, which are particles of initial powders welded together in the solid phase. The concentration of chemical elements on the surface is not homogeneous.
Sintering under hot pressure of mechanically mixed elemental powders resulted in the formation of consolidated high-density specimens with a physical density of 8.12 ± 0.02 and 8.35 ± 0.01 g/cm3 for compositions with 5 wt.% W and 10 wt.% W, respectively. The obtained bulks were medium-entropy Fe–Cr–Ni alloys, which were dispersion-strengthened by molybdenum and tungsten powders.
The physical density of the bulks produced through pressure-assisted sintering from milled powders increased with the increasing milling time. In this work, the maximum physical density values of 8.14 ± 0.02 and 8.40 ± 0.01 g/cm3 were present for the samples fabricated from the powders after 10 h of milling for compositions with 5 wt.% W and 10 wt.% W, respectively.
The low physical density values of bulks made by pressure-assisted sintering from powders after 2.5 and 5 h of milling are consistent with low compressive strength and hardness. Bulks with the maximum density from powders after 10 h of milling showed maximum compressive strength and hardness values at room temperature: 2430 ± 30 MPa and 60 ± 1 HRC for the bulks with 5 wt.% W and 2460 ± 35 MPa and 61 ± 1 HRC for the bulks with 10 wt.% W, respectively.
High hardness values of all specimens were associated with the presence of a brittle and high-strength sigma phase in the structure of the consolidated powder bulks.

Author Contributions

Supervision, conceptualization, funding acquisition, writing—original draft preparation, and powder preparation, A.Y.I.; X-ray analysis and methodology, M.A.S.; investigation of the structure and hardness, I.K.G. and Y.A.K.; SEM analysis, A.V.M.; X-ray analysis and pressure-assisted sintering, A.S.L.; preparation of the specimens, K.V.S.; mechanical testing, M.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation (project no. 22-29-00851).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yeh, B.J.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chang, S.Y. Nanostructured High-Entropy Alloys with Multiple Principal Elements: Novel Alloy Design Concepts and Outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  2. Cantor, B.; Chang, I.T.H.; Knight, P.; Vincent, A.J.N. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375–377, 213–218. [Google Scholar] [CrossRef]
  3. Miracle, D.B.; Senkov, O.N. Acta Materialia A critical review of high entropy alloys and related concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef] [Green Version]
  4. Gao, M.C.; Yeh, J.-W.; Liaw, P.K.; Zhang, Y. High-Entropy Alloys; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar]
  5. Yeh, J.-W. Recent progress in high-entropy alloys. Ann. Chim. Sci. Mat. 2006, 31, 633–648. [Google Scholar] [CrossRef]
  6. Ye, Y.F.; Wang, Q.; Lu, J.; Liu, C.T.; Yang, Y. High-entropy alloy: Challenges and prospects. Materialstoday 2016, 19, 349–362. [Google Scholar] [CrossRef]
  7. Tsai, M.-H.; Yeh, J.-W. High-entropy alloys: A critical review. Mater. Res. Lett. 2014, 2, 107–123. [Google Scholar] [CrossRef]
  8. Patel, P.; Roy, A.; Sharifi, N.; Stoyanov, P.; Chromik, R.R.; Moreau, C. Tribological Performance of High-Entropy Coatings (HECs): A Review. Materials 2022, 15, 3699. [Google Scholar] [CrossRef]
  9. Li, W.; Liu, P.; Liaw, P.K. Microstructures and Properties of High-Entropy Alloy Films and Coatings: A Review. Mater. Res. Lett. 2018, 6, 199–229. [Google Scholar] [CrossRef] [Green Version]
  10. Li, J.; Huang, Y.; Meng, X.; Xie, Y. A Review on High Entropy Alloys Coatings: Fabrication Processes and Property Assessment. Adv. Eng. Mater. 2019, 21, 1900343. [Google Scholar] [CrossRef]
  11. Kenedy, G.R.; Chemeli, K.R.; Cheng, W.-C. The Observation of Cellular Precipitation in an Ni36Co18Cr20Fe19Al7 High-Entropy Alloy after Quenching and Annealing. Materials 2022, 15, 6613. [Google Scholar] [CrossRef]
  12. Lohmuller, P.; Peltier, L.; Hazotte, A.; Zollinger, J.; Laheurte, P.; Fleury, E. Variations of the Elastic Properties of the CoCrFeMnNi High Entropy Alloy Deformed by Groove Cold Rolling. Materials 2018, 11, 1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Liu, F.; Liaw, P.K.; Zhang, Y. Recent Progress with BCC-Structured High-Entropy Alloys. Metals 2022, 12, 501. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Li, R. New Advances in High-Entropy Alloys. Entropy 2020, 22, 1158. [Google Scholar] [CrossRef] [PubMed]
  15. Choi, N.; Park, N.; Kim, J.-k.; Karasev, A.V.; Jönsson, P.G.; Park, J.H. Influence of Manufacturing Conditions on Inclusion Characteristics and Mechanical Properties of FeCrNiMnCo Alloy. Metals 2020, 10, 1286. [Google Scholar] [CrossRef]
  16. Lee, W.-H.; Park, K.B.; Yi, K.-W.; Lee, S.Y.; Park, K.; Lee, T.W.; Na, T.-W.; Park, H.-K. Synthesis of Spherical V-Nb-Mo-Ta-W High-Entropy Alloy Powder Using Hydrogen Embrittlement and Spheroidization by Thermal Plasma. Metals 2019, 9, 1296. [Google Scholar] [CrossRef] [Green Version]
  17. 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]
  18. Ji, W.; Fu, Z.; Wang, W.; Wang, H.; Zhang, J.; Wang, Y.; Zhang, F. Mechanical alloying synthesis and spark plasma sintering consolidation of CoCrFeNiAl high-entropy alloy. J. Alloys Compd. 2014, 589, 61–66. [Google Scholar] [CrossRef]
  19. Kumar, A.; Arora, A.; Chandrakar, R.; Rao, K.R.; Chopkar, M. Nano-crystalline high entropy alloys prepared by mechanical alloying. Mater. Today Proc. 2020, 27, 1310–1314. [Google Scholar] [CrossRef]
  20. Shivam, V.; Shadangi, Y.; Basu, J.; Mukhopadhyay, N. Evolution of phases, hardness and magnetic properties of AlCoCrFeNi high entropy alloy processed by mechanical alloying. J. Alloys Compd. 2020, 832, 154826. [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. Normand, J.; Moriche, R.; García-Garrido, C.; Sepúlveda Ferrer, R.E.; Chicardi, E. Development of a TiNbTaMoZr-Based High Entropy Alloy with Low Young’s Modulus by Mechanical Alloying Route. Metals 2020, 10, 1463. [Google Scholar] [CrossRef]
  23. Faruk, K.; Melih, Y.; İpek Selimoğlu, G.; Bora, D. Influence of Co content on microstructure and hardness of AlCoxCrFeNi (0 ≤ x ≤ 1) high-entropy alloys produced by self-propagating high-temperature synthesis. Eng. Sci. Technol. Int. J. 2022, 27, 101003. [Google Scholar]
  24. Nikolai, K.; Volker, V.; Nikita, S.; Dmitry, S.; Vladimir, S.; Zherebtsov, S. Laser beam welding of a CoCrFeNiMn-type high entropy alloy produced by self-propagating high-temperature synthesis. Intermetallics 2018, 96, 63–71. [Google Scholar]
  25. Evseev, N.S.; Matveev, A.E.; Nikitin, P.Y.; Abzaev, Y.A.; Zhukov, I.A. A theoretical and experimental investigation on the SHS synthesis of (HfTiCN)-TiB2 high-entropy composite. Ceram. Intern. 2022, 48, 16010–16014. [Google Scholar] [CrossRef]
  26. Xiong, K.; Huang, L.; Wang, X.; Yu, L.; Feng, W. Cooling-Rate Effect on Microstructure and Mechanical Properties of Al0.5CoCrFeNi High-Entropy Alloy. Metals 2022, 12, 1254. [Google Scholar] [CrossRef]
  27. Kolmakov, А.G.; Ivannikov, А.Y.; Kaplan, М.А.; Kirsankin, А.А.; Sevost’yanov, М.A. Коррозионностойкие стали в аддитивном производстве. Izvestiya. Ferr. Metall. 2021, 64, 619–650. (In Russian) [Google Scholar]
  28. Ustyukhin, A.S.; Ankudinov, A.B.; Zelensky, V.A.; Alymov, M.I.; Milyaev, I.M.; Vompe, T.A. Synthesis, thermal treatment, and characterization of sintered hard magnetic Fe–30Cr–16Co alloy. J. Alloys Compd. 2022, 902, 163754. [Google Scholar] [CrossRef]
  29. 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]
  30. Chlup, Z.; Fintová, S.; Hadraba, H.; Kuběna, I.; Vilémová, M.; Matějíček, J. Fatigue Behaviour and Crack Initiation in CoCrFeNiMn High-Entropy Alloy Processed by Powder Metallurgy. Metals 2019, 9, 1110. [Google Scholar] [CrossRef] [Green Version]
  31. Prabhu, G.; Kumar, N.A.; Sankaranarayana, M.; Nandy, T.K. Tensile and impact properties of microwave sintered tungsten heavy alloys. Mater. Sci. Eng. A 2014, 607, 63–70. [Google Scholar] [CrossRef]
  32. Rogachev, A.S.; Vadchenko, S.G.; Kochetov, N.A.; Rouvimov, S.; Kovalev, D.Y.; Shchukin, A.S.; Moskovskikh, D.O.; Nepa-pushev, A.A.; Mukasyan, A.S. Structure and properties of equiatomic CoCrFeNiMn alloy fabricated by high-energy ball milling and spark plasma sintering. J. Alloys Compd. 2019, 805, 1237–1245. [Google Scholar] [CrossRef]
  33. Murty, B.S.; Yeh, J.-W.; Ranganathan, S.; Bhattacharjee, P.P. High-Entropy Alloys; Elsevier: London, UK, 2019; ISBN 9780128160671. [Google Scholar]
  34. Liu, Q.; Wang, G.; Sui, X.; Liu, Y.; Li, X.; Yang, J. Microstructure and mechanical properties of ultra-fine grained MoNbTaTiV refractory high-entropy alloy fabricated by spark plasma sintering. J. Mater. Sci. Technol. 2019, 35, 2600–2607. [Google Scholar] [CrossRef]
  35. Lu, K.; Zhu, J.; Guo, D.; Yang, M.; Sun, H.; Wang, Z.; Hui, X.; Wu, Y. Microstructures, Corrosion Resistance and Wear Resistance of High-Entropy Alloys Coatings with Various Compositions Prepared by Laser Cladding: A Review. Coatings 2022, 12, 1023. [Google Scholar] [CrossRef]
  36. Arshad, M.; Amer, M.; Hayat, Q.; Janik, V.; Zhang, X.; Moradi, M.; Bai, M. High-Entropy Coatings (HEC) for High-Temperature Applications: Materials, Processing, and Properties. Coatings 2022, 12, 691. [Google Scholar] [CrossRef]
  37. Ghadami, F.; Davoudabadi, M.A.; Ghadami, S. Cyclic Oxidation Properties of the Nanocrystalline AlCrFeCoNi High-Entropy Alloy Coatings Applied by the Atmospheric Plasma Spraying Technique. Coatings 2022, 12, 372. [Google Scholar] [CrossRef]
  38. Chang, J.; He, J.; Mao, M.; Zhou, W.; Lei, Q.; Li, X.; Li, D.; Chua, C.-K.; Zhao, X. Advanced Material Strategies for Next-Generation Additive Manufacturing. Materials 2018, 11, 166. [Google Scholar] [CrossRef] [Green Version]
  39. Cobbinah, P.V.; Nzeukou, R.A.; Onawale, O.T.; Matizamhuka, W.R. Laser Powder Bed Fusion of Potential Superalloys: A Review. Metals 2021, 11, 58. [Google Scholar] [CrossRef]
  40. Luo, S.; Gao, P.; Yu, H.; Yang, J.; Wang, Z.; Zeng, X. Selective laser melting of an equiatomic AlCrCuFeNi high-entropy alloy: Processability, non-equilibrium microstructure and mechanical behavior. J. Alloys Compd. 2019, 771, 387–397. [Google Scholar] [CrossRef]
  41. He, J.Y.; Wang, H.; Huang, H.L.; Xu, X.D.; Chen, M.W.; Wu, Y.; Liu, X.J.; Nieh, T.G.; An, K.; Lu, Z.P. A precipitation-hardened high-entropy alloy with outstanding tensile properties. Acta Mater. 2016, 102, 187–196. [Google Scholar] [CrossRef] [Green Version]
  42. 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]
  43. Senkov, O.N.; Miracle, D.B.; Chaput, K.J.; Couzinie, J.-P. Development and exploration of refractory high entropy alloys—A review. J. Mater. Res. 2018, 33, 3092–3128. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, M.; Peng, Y.; Zhang, W.; Liu, Y.; Wang, L.; Hu, S.; Hu, Y. Gradient Distribution of Microstructures and Mechanical Properties in a FeCoCrNiMo High-Entropy Alloy during Spark Plasma Sintering. Metals 2019, 9, 351. [Google Scholar] [CrossRef] [Green Version]
  45. Xia, S.; Yang, X.; Chen, M.; Yang, T.; Zhang, Y. The Al Effects of Co-Free and V-Containing High-Entropy Alloys. Metals 2017, 7, 18. [Google Scholar] [CrossRef]
  46. Málek, J.; Zýka, J.; Lukáč, F.; Čížek, J.; Kunčická, L.; Kocich, R. Microstructure and Mechanical Properties of Sintered and Heat-Treated HfNbTaTiZr High Entropy Alloy. Metals 2019, 9, 1324. [Google Scholar] [CrossRef] [Green Version]
  47. Satyanarayana, P.V.; Sokkalingam, R.; Jena, P.K.; Sivaprasad, K.; Prashanth, K.G. Tungsten Matrix Composite Reinforced with CoCrFeMnNi High-Entropy Alloy: Impact of Processing Routes on Microstructure and Mechanical Properties. Metals 2019, 9, 992. [Google Scholar] [CrossRef] [Green Version]
  48. Maity, T.; Prashanth, K.G.; Janda, A.; Kim, J.T.; Spieckermann, F.; Eckert, J. Mechanism of high-pressure torsion-induced shear banding and lamellar thickness saturation in Co-Cr-Fe-Ni-Nb high-entropy composites. J. Mater. Res. 2019, 34, 2672–2682. [Google Scholar] [CrossRef]
  49. Tsai, M.H.; Chang, K.C.; Li, J.H.; Tsai, R.C.; Cheng, A.H. A second criterion for sigma phase formation in high-entropy alloys. Mater. Res. Lett. 2016, 4, 90–95. [Google Scholar] [CrossRef]
Figure 1. Morphology of the elemental powders: (a,b)—Fe; (c,d)—Cr; (e,f)—Ni; (g,h)—Mo; (i,j)—W.
Figure 1. Morphology of the elemental powders: (a,b)—Fe; (c,d)—Cr; (e,f)—Ni; (g,h)—Mo; (i,j)—W.
Metals 12 01764 g001
Figure 2. The effect of the powder preparation method on the bulk, tap density, and density of green bodies: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Figure 2. The effect of the powder preparation method on the bulk, tap density, and density of green bodies: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Metals 12 01764 g002
Figure 3. Morphology of the powders after 5 h of mechanical alloying: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Figure 3. Morphology of the powders after 5 h of mechanical alloying: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Metals 12 01764 g003
Figure 4. The surface morphology of agglomerates obtained during mechanical alloying after 5 h: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W; see Figure 5 for the specific points.
Figure 4. The surface morphology of agglomerates obtained during mechanical alloying after 5 h: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W; see Figure 5 for the specific points.
Metals 12 01764 g004
Figure 5. Elemental composition on the agglomerates’ surface according to EDS analysis: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W; see Figure 4 for the specific points.
Figure 5. Elemental composition on the agglomerates’ surface according to EDS analysis: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W; see Figure 4 for the specific points.
Metals 12 01764 g005
Figure 6. XRD patterns of the 35Fe30Cr20Ni10Mo5W powder produced by mechanical alloying.
Figure 6. XRD patterns of the 35Fe30Cr20Ni10Mo5W powder produced by mechanical alloying.
Metals 12 01764 g006
Figure 7. XRD patterns of the bulks produced during pressure-assisted sintering process: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Figure 7. XRD patterns of the bulks produced during pressure-assisted sintering process: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Metals 12 01764 g007
Figure 8. SEM image and EDX maps of Fe, Cr, Ni, Mo, and W for specimens produced by pressure-assisted sintering 35Fe30Cr20Ni10Mo5W: (a)—from 5 h of mixing; (b)—from milling for 2.5 h; (c)—from milling for 5 h.
Figure 8. SEM image and EDX maps of Fe, Cr, Ni, Mo, and W for specimens produced by pressure-assisted sintering 35Fe30Cr20Ni10Mo5W: (a)—from 5 h of mixing; (b)—from milling for 2.5 h; (c)—from milling for 5 h.
Metals 12 01764 g008
Figure 9. SEM image and EDX maps of Fe, Cr, Ni, Mo, and W for 30Fe30Cr20Ni10Mo10W specimens produced by pressure-assisted sintering: (a)—from 5 h of mixing; (b)—from milling for 2.5 h; (c)—from milling for 5 h.
Figure 9. SEM image and EDX maps of Fe, Cr, Ni, Mo, and W for 30Fe30Cr20Ni10Mo10W specimens produced by pressure-assisted sintering: (a)—from 5 h of mixing; (b)—from milling for 2.5 h; (c)—from milling for 5 h.
Metals 12 01764 g009
Figure 10. Compressive stress–strain curves of bulks produced by the pressure-assisted sintering process: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Figure 10. Compressive stress–strain curves of bulks produced by the pressure-assisted sintering process: (a)—35Fe30Cr20Ni10Mo5W; (b)—30Fe30Cr20Ni10Mo10W.
Metals 12 01764 g010
Table 1. Physical and mechanical properties of bulks produced during pressure-assisted sintering process.
Table 1. Physical and mechanical properties of bulks produced during pressure-assisted sintering process.
Chemical Composition35Fe30Cr20Ni10Mo5W30Fe30Cr20Ni10Mo10W
ModeDensity,
g/cm3
Strength,
MPa
HRCDensity,
g/cm3
Strength,
MPa
HRC
5 h of mixing8.12 ± 0.02860 ± 3053 ± 18.35 ± 0.011350 ± 4052 ± 1
2.5 h of milling8.06 ± 0.01970 ± 3552 ± 18.26 ± 0.01950 ± 2553 ± 1
5 h of milling8.08 ± 0.021350 ± 4056 ± 18.29 ± 0.021500 ± 4059 ± 1
10 h of milling8.17 ± 0.022430 ± 3060 ± 18.42 ± 0.012460 ± 3561 ± 1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ivannikov, A.Y.; Grebennikov, I.K.; Klychevskikh, Y.A.; Mikhailova, A.V.; Sergienko, K.V.; Kaplan, M.A.; Lysenkov, A.S.; Sevostyanov, M.A. Fabrication, Microstructure, and Physico-Mechanical Properties of Fe–Cr–Ni–Mo–W High-Entropy Alloys from Elemental Powders. Metals 2022, 12, 1764. https://doi.org/10.3390/met12101764

AMA Style

Ivannikov AY, Grebennikov IK, Klychevskikh YA, Mikhailova AV, Sergienko KV, Kaplan MA, Lysenkov AS, Sevostyanov MA. Fabrication, Microstructure, and Physico-Mechanical Properties of Fe–Cr–Ni–Mo–W High-Entropy Alloys from Elemental Powders. Metals. 2022; 12(10):1764. https://doi.org/10.3390/met12101764

Chicago/Turabian Style

Ivannikov, Alexander Yurievich, Ivan Konstantinovich Grebennikov, Yulia Alexandrovna Klychevskikh, Anna Vladimirovna Mikhailova, Konstantin Victorovich Sergienko, Mikhail Alexandrovich Kaplan, Anton Sergeevich Lysenkov, and Mikhail Anatolievich Sevostyanov. 2022. "Fabrication, Microstructure, and Physico-Mechanical Properties of Fe–Cr–Ni–Mo–W High-Entropy Alloys from Elemental Powders" Metals 12, no. 10: 1764. https://doi.org/10.3390/met12101764

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