You are on page 1of 25

Accepted Article

DR LUN FENG (Orcid ID : 0000-0002-6895-9950)


DR BILL FAHRENHOLTZ (Orcid ID : 0000-0002-8497-0092)

Article type : Article

Corresponding author mail-id: allen13521@gmail.com


Strength of single-phase high-entropy carbide ceramics up to 2300°C
Lun Feng*, Wei-Ting Chen, William G. Fahrenholtz, Gregory E. Hilmas
Department of Materials Science and Engineering, Missouri University of Science and Technology,
Rolla, MO 65409, USA

Abstract
The mechanical properties of single-phase (Hf,Zr,Ti,Ta,Nb)C high-entropy carbide ceramics were
investigated. Ceramics with relative density >99% and an average grain size of 0.9±0.3 µm were
produced by a two-step process that involved carbothermal reduction at 1600°C and hot pressing at
1900°C. At room temperature, Vickers hardness was 25.0±1.0 GPa at a load of 4.9 N, Young’s
modulus was 450 GPa, chevron notch fracture toughness was 3.5±0.3 MPa•m1/2, and four-point
flexural strength was 421±27 MPa. With increasing temperature, flexural strength stayed above ~400
MPa up to 1800°C, then decreased nearly linearly to 318±21 MPa at 2000°C and to 93±10 MPa at
2300°C. No significant changes in relative density or average grain size were noted after testing at
elevated temperatures. The degradation of flexural strength above 1800°C was attributed to a decrease
in dislocation density that was accompanied by an increase in dislocation motion. These are the first
reported flexural strengths of high-entropy carbide ceramics at elevated temperatures.
Key words: High-entropy carbides; microstructure; mechanical properties; high-temperature flexural
strength; dislocation;

1 Introduction

This article has been accepted for publication and undergone full peer review but has not been
through the copyediting, typesetting, pagination and proofreading process, which may lead to
differences between this version and the Version of Record. Please cite this article as doi:
10.1111/JACE.17443
This article is protected by copyright. All rights reserved
Transition metal carbides are being investigated for nuclear energy and hypersonic aerospace
Accepted Article
applications because of their high melting temperature (above 3000°C) and phase stability. [1-5]
Recently, single-phase solid solutions containing multiple transition metal carbides called high-
entropy carbide (HEC) ceramics have attracted interest due to their higher hardness and better
oxidation resistance compared to the individual carbide ceramics. [6, 7]
Several studies on the synthesis, densification, and microstructure of HEC ceramics have
recently been reported. [6-17] HEC powders have been synthesized from commercial carbide
powders by heating to form solid solutions (SS) [7], carbothermal reduction (CTR) of oxides followed
by SS formation [8, 9], liquid precursors [10], and reaction of elemental metals with carbon sources
[11]. Significant particle growth occurred during the synthesis due to the high temperatures required
for SS formation, which suppressed the densification of the resulting HEC ceramics. Previous studies
indicated that HEC ceramics with relative densities in the range of ~93% to ~99% could be produced
by spark plasma sintering (SPS) of commercial carbide powders at temperatures up to ~2300°C. [6,
16, 17] The resulting ceramics had grain sizes approaching 20 µm due to the high densification
temperatures, [16, 18] which decreased their flexural strength. Analysis indicated that (Hf,Ta,Zr,Nb)C
ceramics were composed of grains that all had the same chemical composition from the micron scale
down to the nano/atomic level without any detectable segregation. [18]
The mechanical properties of HEC ceramics have rarely been reported with most studies only
reporting hardness, Young’s modulus, and fracture toughness measured on small (i.e. 10 mm or 20
mm diameter) specimens. [6, 12-17] Castle et al. [6] reported that (Hf,Ta,Zr,Nb)C ceramics with a
relative density of ~99% produced by SPS had nanohardness values of 36.1±1.6 GPa and a Young’s
modulus of 598±15 GPa, which were measured using loads of 50 mN to 300 mN. Similarly, Ye et al.
[13] reported that (Hf,Zr,Ti,Ta,Nb)C ceramics with a relative density of 95.3% had hardness of
40.6±0.6 GPa and Young’s modulus values ranging from 514±10 GPa to 522±10 GPa measured by
nanoindentation with a load of 8 mN. The same study reported fracture toughness of 3.0±0.2 MPa.m1/2
that was measured by indentation using the direct crack method. These results indicated that HEC
ceramics exhibited higher hardness than HfC (31.5±1.3 GPa), TaC (20.6±1.2 GPa) and binary carbide
(Hf,Ta)C (32.9±1.8 GPa), which was presumably due to mass disorder and solid solution hardening in
the HEC. [6, 19-21] Recently, Wang et al. [22] reported that (Hf,Zr,Ti,Ta,Nb)C ceramics with relative

This article is protected by copyright. All rights reserved


densities of 92.7% to 94.9%, produced by SPS at 2000°C, had Vickers hardness values of 16.2 GPa to
Accepted Article
17.1 GPa at a load of 9.8 N. The flexural strengths ranged from 318 to 400 MPa, as measured in
three-point bending, while fracture toughness measured by the single-edge notched beam (SENB)
method was between 4.9 and 5.9 MPa•m1/2. The results indicated that HEC ceramics with finer grain
sizes had higher strength. Our previous study [12] reported that (Hf,Zr,Ti,Ta,Nb)C ceramics with a
relative density of 99.3% had a Vickers hardness of 24.8±0.8 GPa at a load of 9.8 N. The hardness
was higher than values ranging from 15 GPa to 18.8 GPa [13, 16, 22] reported in previous studies due
to higher relative density and finer grain size. Further, the Young’s modulus value of 452 GPa was
similar to values of 456 GPa [19] and 464 GPa [17] predicted as part of computational studies. To
date, mechanical properties of HEC ceramics at elevated temperatures have not been reported.
However, Demirskyi el al. [23] reported that a ternary (Ta,Zr,Nb)C ceramics had a room temperature
elastic modulus of 563±19 GPa, flexural strength of 460±24 MPa by three-point bending, and fracture
toughness of 2.9±0.3 MPa•m1/2 by SENB. With increasing the temperature, flexural strength
increased to 496±44 MPa at 1600°C, then decreased to 366±46 MPa at 1800°C and 139±32 MPa at
2000°C.
Our previous studies investigated the synthesis, densification, and microstructure of dense
(Hf,Zr,Ti,Ta,Nb)C ceramics with fine grain size. [8, 12] The purpose of the present study was to
investigate the mechanical properties of (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics with emphasis on the
flexural strength at temperatures up to 2300°C.

2 Experimental Procedures
2.1 Specimen preparation
Hafnium oxide (HfO2, 99%, -325 mesh; Alfa Aesar, Tewksbury, MA), zirconium oxide (ZrO2, 99%,
5 µm; Sigma-Aldrich, St. Louis, MO), titanium oxide (TiO2, 99.9%, 32 nm APS; Alfa Aesar),
tantalum pentoxide (Ta2O5, 99.8%, 1-5 µm; Atlantic Equipment Engineers, Upper Saddle River, NJ),
niobium oxide (Nb2O5, 99.5%, -100 mesh; Alfa Aesar) and carbon black (C, BP1100, Cabot,
Alpharetta, GA) were used as the starting materials. The carbothermal reduction procedure for the
synthesis of the HEC powder mixtures was previously reported in detail. [8] Briefly, oxides were
batched with a 1 wt% of excess carbon based on the carbothermal reduction reactions and then mixed

This article is protected by copyright. All rights reserved


by high-energy ball milling (HEBM). Synthesis was conducted at 1600°C under mild vacuum (~13
Accepted Article
Pa) in a furnace with a graphite heating element. Reacted powder mixtures were used to produce
(Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C billets (~30 mm × ~45 mm × ~5 mm) by hot pressing (HP) at 1900°C
under a uniaxial pressure of 32 MPa according to the results from our previous study. [12]

2.2 Characterization
Hot-pressed billets were ground on both sides to remove the graphite foil and reaction layers. Relative
densities were measured by Archimedes’ method using distilled water as the immersing medium.
Phase compositions of sintered billets were determined by x-diffraction analysis (XRD, PANalytical
X-Pert Pro, Malvern Panalytical Ltd., Royston, United Kingdom). Microstructures of specimens
before and after mechanical testing were examined in a dual-beam scanning electron microscopy-
focused ion beam system (SEM-FIB, Helios NanoLab 600, Thermo Fisher Scientific, Hillsboro, OR).
After mechanical testing, the both tensile surface and cross-section of the specimens were examined
after polishing to determine average grain sizes. The tensile surface was initially ground using a
diamond grinding wheel on a fully automated surface grinder (FSG-3A818, Chevalier, Santa Fe
Springs, CA) to remove about ~200 µm. Then the specimens were polished to 0.25 µm finish using
diamond suspensions. Polished tensile surfaces and cross-sections (from tensile regions of the test
specimens) were further thermally etched at 1600°C for 30 min in a flowing Ar atmosphere. [12] A
computer-based image analysis program (ImageJ, National Institutes of Health, Bethesda, MD) was
used to determine the average grain sizes by examining at least 250 grains per specimen. Specimens
were further analyzed using transmission electron microscopy (TEM, Tecnai F20, Thermo Fisher
Scientific). TEM foils (10 µm × 10 µm) were prepared by the lift-out method using a dual beam
SEM-FIB (Helios Nanolab 600) from the polished cross-sections in the tensile area.

2.3 Mechanical testing


At room temperature (RT), flexural strengths were measured following ASTM C1161 in four-point
bending using a fully-articulating test fixture and type-B bars (45 mm × 4 mm × 3 mm). Flexural
strengths were measured at elevated temperatures (1400°C, 1600°C, 1800°C, 2000°C, 2200°C, and
2300°C) in four point bending using the same sized bars and a semi-articulating test fixture following

This article is protected by copyright. All rights reserved


the procedures outlined in ASTM1211. Six bars were tested at room temperature and a minimum of
Accepted Article
five bars were tested at each elevated temperature. Bars were machined by diamond grinding on a
fully automated surface grinder. Tensile surfaces were ground using a #600 grit diamond grinding
wheel and then polished and chamfered using a series of successively finer diamond abrasives with a
final abrasive size of 0.25 µm. Room temperature testing was performed using a screw-driven
instrumented load frame (Model 5881, Instron, Norwood, MA). A deflectometer was used to measure
displacement at the center of the bar. Elevated temperature testing was performed in an environmental
chamber using an induction heated graphite hot zone and load train, under a flowing argon
atmosphere. [24] The heating profile was 60°C/min from room temperature to 900°C, then 100°C/min
to the desired testing temperatures, followed by a 5 min isothermal hold to allows the bars to reach
thermal equilibrium. Different crosshead rates were selected at elevated temperatures to ensure that
the load-displacement curves remained linear up to the failure point as outlined in ASTM C1211.
Room temperature fracture toughness was measured by the chevron notch beam method in four-
point bending using a fully-articulating test fixture using type-A bars (45 mm ×3 mm × 4 mm)
following ASTM Standard C1421. The chevron notch was machined using a dicing saw (Accu-cut
5200, Aremco Products, Ossining, NY) with a 0.15 mm thick diamond wafering blade. The crosshead
rate during toughness testing was 0.02 mm/min. The notch dimensions were measured after testing
using digital microscope (KH-3000, Hirox-USA, Hackensack, NJ). The reported fracture toughness
values were the average of six valid tests, and the variance was the standard deviation of the
individual calculated toughness values.

3 Results and discussion


3.1 Characterization
A single-phase in cubic structure was detected by XRD for (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics,
although trace amounts of oxides remained in the ceramics (Fig. 1). The lattice parameter was
0.4511(2) nm measured using Rietveld refinement, which was consistent with the value reported in
our previous study [12]. The value was higher than the value of 0.4488(2) nm reported in another
study [22], presumably due to the lower oxygen content of the HEC ceramics produced in the present
study. When oxygen dissolves into the carbide lattice, the lattice parameter decreases. Based on XRD,

This article is protected by copyright. All rights reserved


high purity HEC ceramics were produced.
Accepted ArticleRelative densities were above 99.0% for all of the materials produced in the present study. These
values are higher than the relative densities in the range of 92% to 95% that have been reported by
other groups as summarized in Table 1. [13, 22] The higher relative density achieved in the present
study was attributed to the submicron particle size and low oxygen content (less than 0.8 wt%) of the
powders. [8, 12] Dense microstructures were observed on the surface (perpendicular to the hot
pressing direction) (Fig. 2 (a)) and cross-section (parallel to the hot pressing direction) (Fig. 3 (a)) of
the bars prior to testing. Trace amounts of residual pores were present in the ceramics, which was
consistent with the measured relative densities. The average grain size measured on the surface
(0.9±0.3 µm) and cross-section (0.8±0.2 µm) were similar. (Table 1 and Table 2) In addition, no
microstructural texturing was obvious from the micrographs. This analysis supported the assertion
that the microstructures were homogeneous and did not exhibit texturing.

3.2 Room temperature mechanical properties


The mechanical properties measured at room temperature for (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics
are summarized in Table 1. Vickers hardness increased from 24.4±0.4 GPa for a load of 9.8 N to
25.0±1.0 GPa for a load of 4.9 N. The hardness value of 24.4±0.4 GPa at a load of 9.8 N in the
present study was much higher than the values of 18.8±0.4 GPa [13] and 16.2±1.0 GPa to 17.1±0.5
GPa [22] reported for same loads in previous studies, which can be attributed to the higher relative
density and finer grain size of the HEC ceramics in the present study. Young’s modulus values were
450±5 GPa by ultrasound and 454±5 GPa by bending, which were comparable to reported values of
443±5 GPa from experimental studies [17, 25], and values of 456 GPa [26] and 464 GPa [17] from
computational calculations. Both hardness and modulus values of HEC ceramics were higher than
those values (23±2 GPa [17] for hardness and 436±30 GPa [17] for modulus) predicted using a
volumetric rule of mixture (ROM) calculation based on the properties of the constituent binary
carbides, presumably due to the effects of mass disorder and solid solution hardening described in
other studies of high entropy ceramics. [6, 19-21]
Fracture toughness was 3.5±0.3 MPa•m1/2 at room temperature. The value was higher than the
value of 3.0±0.2 MPa•m1/2 measured by the direct crack method [13], but lower than values of 4.9

This article is protected by copyright. All rights reserved


MPa•m1/2 to 5.9 MPa•m1/2 measured by a single-edge notched beam (SENB) method. [22] However,
Accepted Article
direct comparison to the reported results is difficult due to the difference in relative densities, grain
sizes, and measurement methods. Nevertheless, toughness is in the range expected for carbide
ceramics.
The room temperature flexural strength was 421±27 MPa. The average strength of the HEC is
higher than values for other HEC ceramics, which are typically in the range of ~300 MPa to ~400
MPa [22]. The higher strengths in the present study are attributed to the higher relative density of the
ceramics in this study (~99%) compared to relative densities of 95% or lower for other studies.
Comparisons of room temperature strengths of carbide ceramics, in Fig. 4, indicated that
(Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics had a higher flexural strength than most binary [23] or
individual carbide [27, 28, 31] ceramics with the exception of TiC [30] ceramics. The enhancement of
flexural strength of HEC ceramics compared to the individual carbides may be due the refined
microstructure and higher (~99%) relative density of the ceramics in the present study in addition to
mass disorder and solid solution hardening effects mentioned in other studies. [6, 19-21]

3.3 Flexural strength at elevated temperatures


Flexural strengths of HEC ceramics were measured up to 2300°C. (Table 2) As described in the
standard, the crosshead displacement rate was increased as testing temperature increased with typical
load versus displacement curves shown in Fig. 5. The specimens exhibited linear-elastic behavior to
failure from room temperature to 2000°C. However, deviation from linear-elastic behavior was
observed at 2200°C and 2300°C even with the higher displacement rates.
Flexural strengths of the HEC ceramics were above ~400 MPa from room temperature up to
1800°C, then decreased nearly linearly from 318±21 MPa at 2000°C to 93±10 MPa at 2300°C. In
general, flexural strength of ceramics is affected by residual porosity and the grain size. [32, 33]
Based on the microstructures of both tensile surfaces and cross-sections after testing at elevated
temperatures, a small volume fraction (<1 vol%) of residual porosity was present in all specimens, but
the amount of porosity did not change significantly with testing temperature. (Fig. 2 and Fig. 3) In
addition, the average grain size remained about the same after testing at elevated temperatures. Fig. 6
shows the average grain sizes as a function of temperature for both the tensile surfaces and cross-

This article is protected by copyright. All rights reserved


sections of the HEC specimens. The results indicated that not only were the average grain sizes
Accepted Article
nominally the same for tensile surfaces and cross-sections, but that the average grain sizes were
around 0.9 µm for all of the testing temperatures. The only apparent deviation in average grain size
was an increase to about 1.2 µm at 1600°C; however, this can likely be attributed to not sampling
enough grains to get a representative average since the standard deviation was wider at 1600°C and
the grain sizes at higher testing temperatures were the same as the lower temperatures. Based on these
observations, the change in strength of HEC ceramics above 1800°C was not caused by changes in
microstructure such as the amount of porosity or the average grain size.
Comparisons in Fig. 7 showed that the HEC from the present study had higher strength than binary
and individual carbides at elevated temperatures. [29, 34] For example, one previous study [23]
showed that the strength of ternary TaZrNbC ceramics maintained their room temperature strength
until 1600°C. However, the present study showed that (HfZrTiTaNb)C ceramics their maintained
room temperature strength until a higher temperature of 1800°C, which indicates that HEC ceramics
can have higher strengths at elevated temperatures, which is likely due to mass disorder and solid
solution hardening effects.

3.4 Mechanism of strength retention


Previous studies have reported that HEC ceramics exhibited higher hardness than expected based on
the hardness values of the individual carbides, which was attributed to the effects of mass disorder
and solid solution hardening. [6, 35-37] Castle et al. [6] and Smith et al. [19, 37] pointed out that
hardness behavior was governed by the systems in the lattice for which the resistance to slip is lowest.
Solid solutions that incorporate multiple transition metal carbides into a single phase may contain
nonuniformities in the lattice structure that inhibit the motion of dislocations that result in a solid
solution hardening effect. For example, Csanadi et al. [36] found that HEC ceramics had higher
strengths in micropillar compression than single metal carbides. The dominant system was {111}
〈110〉, which possibly occurred by activation of partial dislocations. The present study revealed that
HEC ceramics maintained their room temperature strength up to at least 1800°C, which shows
strength retention to higher temperatures as compared to individual carbides, binary carbides, or
ternary carbides. (Fig. 7) Hence, dislocations were investigated further because they may play an

This article is protected by copyright. All rights reserved


important role in the evolution of flexural strength at elevated temperatures in HEC ceramics.
Accepted ArticleTo study possible dislocation effects on the strength of HEC ceramics at elevated temperatures,
TEM analysis was performed on specimens extracted from the tensile regions of strength bars that
had been tested at elevated temperatures. Because flexural strength decreased above ~1800°C, the
specimens tested at 1600°C and 2000°C were selected for TEM analysis to compare dislocations in
materials before and after the decrease in strength. Dislocations were observed in bright field images
as shown in Fig. 8. Based on the results reported in previous studies [36, 38, 39], the observed
dislocations appeared to be partial dislocations, which were presumably formed during deformation
by the {111}<110> slip system. Analysis under different tilt conditions indicated that the density of
dislocations was apparently higher for the specimen tested at 1600°C (Fig. 8 (a, c, d)) than in the
specimen tested at 2000°C (Fig. 8 (b, e)). Quantification of dislocation density and the mechanism of
dislocation motion (i.e., identification of Burgers’ vectors and active slip planes) in HEC ceramics
will be the subject of future TEM studies. Materials with higher dislocation densities tend to have
higher yield strengths due to the higher shear stress required to move the dislocations, according to
the Equation (1). [40]
𝜏 = 𝛼𝐺𝑏 𝜌 (1)
Where 𝜏 is the shear stress, 𝛼 is a proportionality, G is the shear modulus, b is the Burgers vector, and
𝜌 is the dislocation density. The dislocation density appeared to be higher in the ceramic tested at
1600°C compared to the one tested at 2000°C, which is consistent with the higher strength measured
at 1600°C. In addition, a higher lattice resistance to dislocation movement also increases yield
strength, while Equation (2) shows that lattice resistance decreases linearly with increasing
temperature. [40, 41]

𝜏 = 𝜏𝑝 ―
𝑘𝑇
𝑉 𝑙𝑛 ( )
𝜌𝑏2𝜇
𝑑𝛾 𝑑𝑡
(2)

Where 𝜏𝑝 is the Peierls stress, k is Boltzmann’s constant, T is temperature, V is the activation volume,
𝜇 is the attempt frequency by which the dislocation tries the move, and 𝛾 is the strain in the materials
caused by the solute. Lattice resistance presumably decreased as the temperature increased from
1600°C to 2000°C, which would increase the motion of dislocations and lower the yield strength. The
TEM observation, combined with trends in dislocation mobility, indicated that the decrease in flexural
strength of the HEC ceramics above 1800°C was consistent with a decrease in dislocation density

This article is protected by copyright. All rights reserved


with increasing temperature. However, other effects, particularly creep deformation or reduction in
Accepted Article
elastic modulus with temperature could also be responsible for the decrease in strength.

4 Conclusion
The strength of dense and single-phase (Hf,Zr,Ti,Ta,Nb)C high-entropy carbide ceramics was
measured up to 2300°C. Dense HEC ceramics had an average grain size of 0.9±0.3 µm, Vickers
hardness of 25.0±1.0 GPa, Young’s modulus of about 450 GPa, fracture toughness of 3.5±0.3
MPa•m1/2, and flexural strength of 421±27 MPa at room temperature. Flexural strength was relatively
constant at ~400 MPa up to 1800°C. Above 1800°C, strength decreased near linearly with
temperature from 318±21 MPa at 2000°C to 93±10 MPa at 2300°C. No significant changes in relative
densities or average grain size were observed after testing at elevated temperatures. The flexural
strength decreased above 1800°C, which may be due to a decrease in dislocation density caused by an
increase in dislocation motion with increasing temperature. This study is the first to analyze the
mechanical behaviors of high-entropy carbide ceramics at elevated temperatures.

Acknowledgements
This research was conducted as part of the Enabling Materials for Extreme Environments Signature
Area at Missouri S&T. The use of the Advanced Materials Characterization Laboratory is also
acknowledged.

References
1. Toth LE. Transition metal carbides and nitrides. New York, NY: Academic Press; 1971.
2. Storms EK. The refractory carbides. New York and London, UK: Academic Press, 1967; p. 18–
27.
3. Wuchina EJ, Opila E, Opeka MM, Fahrenholtz WG, Talmy IG, UHTCs: ultra-high temperature
ceramic materials for extreme environment applications. Interface. 2007;16:30–6.
4. Katoh Y, Vasudevamurthy G, Nozawa T, Snead L. Properties of zirconium carbide for nuclear
fuel application. J Nul Mater. 2013;441:718–42.
5. Florez R, Crespillo ML, He XQ, White TA, Hilmas GE, Fahrenholtz WG, Graham J. J Eur Ceram

This article is protected by copyright. All rights reserved


Soc. 2020;40:1791–800.
Accepted Article
6. Castle EG, Csanadi T, Grasso S, Dusza J, Reece M. Processing and properties of high-entropy
ultra-high temperature carbides. Sci Rep. 2018;8:8609–20.
7. Zhou JY, Zhang JY, Zhang F, Niu B, Lei LW, Wang WM. High-entropy carbide: A novel class of
multicomponent ceramics. Ceram Int. 2018;44:22014–8 .
8. Feng L, Fahrenholtz WG, Hilmas GE, Zhou Y. Synthesis of single-phase high-entropy carbide
powders. Scr Mater. 2019;162:90–3.
9. Ye BL, Ning SS, Liu D, Wen TQ, Chu YH. One-step synthesis of coral-like high-entropy carbide
powders. 2019;102:6372–78.
10. Li F, Lu Y, Wang XG, Bao WC, Liu JX, Xu FF, Zhang GJ. Liquid precursor-derived high-
entropy carbide nanopowders. Ceram Int. 2019;45:22437–41.
11. Lun HL, Zeng Y, Xiong X, Ye ZM, Qian TX, Sun W, Wang YL, Chen ZK. Synthesis of carbide
solid solution with multiple components using elemental powder. Adv Powder Technol.
2020;31:505–9.
12. Feng L, Fahrenholtz WG, Hilmas GE. Low-temperature sintering of single-phase, high-entropy
carbide ceramics. J Am Ceram Soc. 2019;102:7217–24.
13. Ye BL, Wen TQ, Huang KH, Wang CZ, Chu YH. First-principles study, fabrication, and
characterization of (Hf0.2Zr0.2Ta0.2Nb0.2Ti0.2)C high-entropy ceramic. J. Am Ceram Soc.
2019;102:4344–52.
14. Wei XF, Liu JX, Li F, Qin Y, Liang YC, Zhang GJ. High entropy carbide ceramics from different
starting materials. 2019;39:2989–94.
15. Zhang Z, Fu S, Aversano F, Bortolotti M, Zhang HW, Hu CF, Grasso S. Arc melting: a novel
method to prepare homogeneous solid solutions of transition metal carbides (Zr, Ta, Hf). Ceram
Int. 2019;45:9316–19.
16. Yan XL, Contantin L, Lu YF, Silvain JF, Nastasi M, Cui B. (Hf0.2Zr0.2Ta0.2Nb0.2 Ti0.2)C high-
entropy ceramics with low thermal conductivity. J Am Ceram Soc. 2018;101:4486–91.
17. Sarker P, Harrington T, Toher C, Oses C, Samiee M, Maria JP, Brenner DW, Vecchio KS,
Curtarolo S. High-entropy high-hardness metal carbides discovered by entropy descriptors. Nat.
Commun. 2018;9:4980–9.

This article is protected by copyright. All rights reserved


18. Dusza J, Svec P, Girman V, Sedlak R, Castle EG, Csanadi T, Kovalcikova A, Reece MJ.
Accepted Article
Microstructure of (Hf-Ta-Zr-Nb)C high-entropy carbide at micro and nano/atomic level. J Eur
Ceram Soc. 2018;38:4303–7.
19. Smith CJ, Yu XX, Guo QY, Weinberger CR, Thompson GB. Phase, hardness, and deformation
slip behavior in mixed HfxTa1-xC. Acta Mater. 2018;145:142–53.
20. Ye YF, Wang Q, Lu J, Liu CT, Yang Y. High-entropy alloy: challenges and prospects. Mater
Today. 2016;19:349–62.
21. Senkov ON, Senkova, SV, Woodward C, Miracle DB. Low-density, refractory multi-principal
element alloys of the Cr-Nb-Ti-V-Zr system: Microstructure and phase analysis. Acta Mater.
2013;61:1545–57.
22. Wang F, Zhang X, Yan XL, Lu YF, Nastasi M, Chen Y, Cui B. The effect of submicron grain size
on thermal stability and mechanical properties of high-entropy carbides ceramics. J Am Ceram
Soc. 2020;00:1–10.
23. Demirskyi D, Borodianska H, Suzuki TS, Sakka Y, Yoshimi K, Vasylkiv O. High-temperature
flexural strength performance of ternary high-entropy carbide consolidated via spark plasma
sintering of TaC, ZrC and NbC. Scr Mater. 2019;164:12–6.
24. Neuman EW, Brown-Shaklee HJ, Watts JL, Hilmas GE, Fahrenholtz WG. Case study: Building
an ultra-high temperature mechanical testing system. Am Ceram Soc Bull. 2013;92:36–8
25. Harrington TJ, Gild J, Sarker P, Toher C, Rost CM, Dippo OF, McElfresh C, Kaufmann K, Marin
E, Borowski L, Hopkins PE, Luo J, Curtarolo S. Brenner DW, Vecchio KS. Phase stability and
mechanical properties of novel high entropy transition metal carbides. Acta Mater. 2019;166:271–
80.
26. Yang Y, Wang W, Gan GY, Shi XF, Tang BY. Structural, mechanical and electronic properties of
(TaNbHfTiZr)C high entropy carbide under pressure: Ab initio investigation. Physica B.
2018;550:163–70.
27. Silvestroni L, Bellosi A, Melandri C, Sciti D, Liu JX, Zhang GJ. Microstructure and properties of
HfC and TaC-based ceramics obtained by ultrafine powder. J Eur Ceram Soc. 2011;31:619–27.
28. Feng L, Fahrenholtz WG, Hilmas GE, Watts J. Zhou Y. Densification, microstructure, and
mechanical properties of ZrC-SiC ceramics. J Am Ceram Soc. 2019;102:5786–95.

This article is protected by copyright. All rights reserved


29. Demirskyi D, Nishimura T, Sakka Y, Vasylkiv O. High-strength TiB2-TaC ceramic composites
Accepted Article
prepared using reactive spark plasma consolidation. Ceram Int. 2016;42: 1298–306.
30. Nguyen VH, Pazhouhanfar Y, Ali Delbari S, Shaddel S, Babapoor, Mohammadpourderakhshi Y,
Le QV, Shokouhimehr M, Shahedi Asl M, Namini AS. Beneficial role of carbon black on the
properties of TiC ceramics. Ceram Int. 2020 (In press)
31. Ordan’yan SS, Stepanenko EK, Sokolov IV. Strength of NbC-NbB2 sintered composites. Izv
Vyssh Uchebn Zaved Khim Khim T. 1984;27:1201–3.
32. Salem J, Hilmas GE, Fahrenholtz WG. Mechanical properties and processing ceramic binary,
ternary, and composite systems: ceramic engineering and science proceedings. Wiley Am Ceram
Soc. 2008.
33. Chantikul P, Bennison SJ, Lawn B. Role of grain size in the strength and R-Curve properties of
Alumina. J Am Ceram Soc. 1990;73:2419–27.
34. A Kelly, Rowcliffe DJ. Deformation of poly crystalline transition metal carbides. J Am Ceram
Soc. 1967;50:253–6.
35. Han XX, Girman V, Sedlak R, Dusza J, Castle EG, Wang YC, Reece M, Zhang CY. Improved
creep resistance of high entropy transition metal carbides. J Eur Ceram Soc, 2020;40:2709–15.
36. Csanadi T, Castle E, Reece MJ, Dusza J. Strength enhancement and slip behavior of high-entropy
grains during micro-compression. Sci Rep. 2019;9:10200.
37. Smith CJ, Yu XX, Guo Q, Weinberger CR, Thompson GB. Phase, hardness, and deformation slip
behavior in mixed HfxTa1-xC. Acta Mater. 2018;145:142–53.
38. Green DJ. An introduction to the mechanical properties of ceramics. Cambridge Solid State
Science Series, Eds. Ckarke, D.R., Suresh, S., Ward, I.M. 1998.
39. Fahrenholtz WG, Wuchina EJ, Lee WE, Zhou YC. Ultra-high temperature ceramic: materials for
extreme environment applications. Chapter 10, Wiley Am Ceram Soc. 2014.
40. Green DJ. An introduction to the mechanical properties of ceramics. Cambridge Solid State
Science Series, Eds. Ckarke, D.R., Suresh, S., Ward, I.M. 1998.
41. Fahrenholtz WG, Wuchina EJ, Lee WE, Zhou YC. Ultra-high temperature ceramic: materials for
extreme environment applications. Chapter 10, Wiley Am Ceram Soc. 2014.

This article is protected by copyright. All rights reserved


Tables
Accepted Article
Table 1 Sintering conditions, lattice parameters, theoretical densities, measured densities, average
grain sizes, and room temperature mechanical properties of (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics
compared to results from previous studies.
Table 2 Flexural strength and average grain sizes of HEC ceramics after testing at elevated
temperatures in Argon.

This article is protected by copyright. All rights reserved


Figures
Accepted Article
Fig. 1 XRD pattern of (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics produced at 1900°C by HP.
Fig. 2 Microstructures of polished and thermally etched tensile surface of
(Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics following mechanical testing at (a) RT, (b) 1400°C, (c)
1600°C, (d) 1800°C, (e) 2000°C, (f) 2200°C, and (g) 2300°C.
Fig. 3 Microstructures of polished and thermally etched cross-sections of
(Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics following mechanical testing at (a) RT, (b) 1400°C, (c)
1600°C, (d) 1800°C, (e) 2000°C, (f) 2200°C, and (g) 2300°C.
Fig. 4 Room temperature strength of (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics compared to individual
carbides, binary carbides, ternary carbides, and HEC ceramics reported in previous studies.
Fig. 5 Load versus displacement curves for HEC ceramics as a function of testing temperature.
Fig. 6 Average grain sizes as a function of mechanical testing temperature for tensile surfaces and
cross-sections of HEC ceramics.
Fig. 7 Flexural strength as a function of temperature for HEC ceramics compared to the results of
individual carbides, binary carbides, ternary carbides, and HEC ceramics reported in previous studies.
Fig. 8 TEM images (bright field) at different magnifications for HEC ceramics after testing at (a, c, an
d d) 1600°C and (b and e) 2000°C. White arrows show the dislocations.

This article is protected by copyright. All rights reserved


Accepted Article
Table 1 Sintering conditions, lattice parameters, theoretical densities, measured densities, average grain sizes, and room temperature mechanical
properties of (Hf0.2,Zr0.2,Ti0.2,Ta0.2,Nb0.2)C ceramics compared to results from previous studies.

Parameters Lattice Theoretical Measured Relative Average Vickers Young’s Flexural Fracture
(°C/MPa/min) parameters density density density grain size hardness modulus strength toughness Reference
(nm) (g/cm3) (g/cm3) (%) (μm) (GPa) (GPa) (MPa) (MPa•m1/2)

25.0±1.0a 450±5c Present


HP/1900/32/60 0.4511(2) 9.43 9.33 >99.0 0.9±0.3 421±27 3.5±0.3
24.4±0.4b 454±5d study

SPS/2000/30/5 8.45 94.9 16.5±4.2 16.2±1.0b — 318±25 4.9±0.5

SPS/2000/30/5 0.4488 (2) — 20


8.25 92.7 0.6±0.2 17.1±0.5b — 400±27 5.9±0.7
+1800/30/15
514±10~
HP/1800/30/30 0.4497 9.52 9.08 95.3 — 18.8±0.4b — 3.0±0.2 11
522±10e

a-HV0.5; b-HV1; c-measured by ultrasound; d-measured by bending; e-measured by nanoindentation.

This article is protected by copyright. All rights reserved


Accepted Article
Table 2 Flexural strength and average grain sizes of HEC ceramics after testing at elevated
temperatures in Argon.

Flexural Average grain size (µm)


Temperature Crosshead rate
strength
(°C) (mm/min) Tensile surface Cross-section
(MPa)

RT 0.5 421±27 0.9±0.3 0.8±0.2


1400 2.0 467±70 1.0±0.4 0.9±0.4
1600 2.0 387±19 1.2±0.6 1.3±0.8
1800 2.0 423±5 0.9±0.4 1.0±0.4
2000 2.5 318±21 0.8±0.3 0.9±0.3
2200 3.0 167±25 0.9±0.3 0.8±0.3
2300 3.5 93±10 1.0±0.5 1.0±0.5

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f1.tif

This article is protected by copyright. All rights reserved


Accepted Article

This article is protected by copyright. All rights reserved


Accepted Article

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f4.tif

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f5.tif

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f6.tif

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f7.tif

This article is protected by copyright. All rights reserved


Accepted Article

jace_17443_f8.tif
This article is protected by copyright. All rights reserved

You might also like