You are on page 1of 6

Journal of the European Ceramic Society 41 (2021) 2855–2860

Contents lists available at ScienceDirect

Journal of the European Ceramic Society


journal homepage: www.elsevier.com/locate/jeurceramsoc

Ultrafast densification of high-entropy oxide


(La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7 by reactive flash sintering
Hai-Rong Mao a, Rui-Fen Guo a, Yue Cao a, Shen-Bao Jin b, Xiao-Ming Qiu a, Ping Shen a, *
a
Key Laboratory of Automobile Materials (Ministry of Education), School of Materials Science and Engineering, Jilin University, No. 5988 Renmin Street, Changchun,
130022, PR China
b
Herbert Gleiter Institute of Nanoscience, School of Materials Science and Engineering, Nanjing University of Science and Technology, Nanjing, 210094, China

A R T I C L E I N F O A B S T R A C T

Keywords: Pyrochlore-type high-entropy oxides (HEOs) are usually sintered at high temperatures for a long time to achieve
High-entropy pyrochlore oxide (HEPO) full density. Herein, we synthesized pyrochlore-structured (La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7 HEOs with den­
Reactive flash sintering (RFS) sities up to 99 % at a furnace temperature of 1200 ◦ C in seconds via reactive flash sintering (RFS). The resultant
Mechanical property
HEOs achieved compositional uniformity at the atomic level and exhibited superior modulus, hardness and
fracture toughness compared to the counterparts prepared by conventional solid-state sintering (at 1600 ◦ C for
6 h). The underlying mechanisms for the ultrafast densification of the RFSed-HEOs were addressed in view of the
roles of electric field, rapid heating, external pressure and internal reactions.

1. Introduction (Ti0.2Y0.2Zr0.2Sn0.2Hf0.2)O3-x and (Bi0.2Na0.2K0.2Ba0.2Ca0.2)TiO3 perov­


skite oxide ceramics using RFS. To our knowledge, there are no reports
Since Rost et al. first prepared a high-entropy oxide (HEO), on the synthesis of dense HEPOs by RFS. In addition, despite that RFS
Mg0.2Ni0.2Co0.2Cu0.2Zn0.2O, via solid-state sintering in 2015 [1], a series has been demonstrated to be an effective method for the synthesis of
of HEOs with unique thermophysical, mechanical and optical/elec­ high-entropy oxides, the mechanisms for the rapid densification remain
trical/magnetic properties have been developed [2–5]. Among them, largely unclear so far.
high-entropy pyrochlore oxides (HEPOs) have attracted great attention In this work, we prepared a (La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7
due to their extremely low thermal conductivities, high melting points (hereafter abbreviated as 5RE2Zr2O7) HEPO with a 99 % relative density
and sluggish grain growth rates [6–9]. However, these unique properties via RFS at 1200 ◦ C in seconds. The mechanical properties of the RFSed-
also make the sintering of the HEPOs to full density via conventional 5RE2Zr2O7 HEPOs were evaluated by nano-indentation tests and
techniques rather difficult. For instance, Li et al. synthesized a compared with those of the counterparts that were conventionally sin­
(La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7 HEPO at 1500 ◦ C for 3 h and ob­ tered at 1600 ◦ C for 6 h. Besides, the mechanisms for the ultrafast
tained only 71.4 % relative density [8]. In contrast, Teng et al. [9] densification during the RFS process were addressed.
prepared a nearly fully dense (La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7 HEPO
after sintering at 1500 ◦ C for 40 h. 2. Experimental
An alternative method to prepare high-density HEOs is reactive flash
sintering (RFS). In recent years, RFS has been applied to synthesize Commercially available La2O3, Nd2O3, Sm2O3, Eu2O3, Gd2O3 and
single-phase compounds by dissolving and densifying oxide mixtures or ZrO2 powders (all in 99.9 wt% purity, D50 = 30 nm) were used as raw
amorphous phase precursors at lower temperatures [10–12]. Recently, materials. Firstly, stoichiometric amounts of the respective oxides for
Yoon et al. [13], Liu et al. [14] and Kumar et al. [15], respectively, re­ the synthesis of 5RE2Zr2O7 were ball-milled for 60 min in a high-speed
ported the RFS of a rocksalt Mg0.2Ni0.2Co0.2Cu0.2Zn0.2O HEO, which can three-dimensional pendulum ball-mill machine. The powder mixtures
readily be sintered to a high density by conventional solid-state sintering with 2 wt% polyvinyl alcohol (PVA) as binder were cold pressed into
at low temperatures (≤ 1050 ◦ C) [16]. More recently, Wang et al. [17] cylindrical pellets with sizes of φ 6 mm × 3.5 mm under a uniaxial
and Liu et al. [18] have prepared high entropy Sr pressure of 400 MPa. The pellets were then pre-sintered at 600 ◦ C for 1 h

* Corresponding author.
E-mail address: shenping@jlu.edu.cn (P. Shen).

https://doi.org/10.1016/j.jeurceramsoc.2020.11.052
Received 13 August 2020; Received in revised form 24 November 2020; Accepted 29 November 2020
Available online 10 December 2020
0955-2219/© 2020 Elsevier Ltd. All rights reserved.
H.-R. Mao et al. Journal of the European Ceramic Society 41 (2021) 2855–2860

in air to remove the PVA binder. The relative density of the pre-sintered indenter at room temperature. The average of five measurements was
samples was ~50 %. For the RFS, the pellet was placed between two presented.
graphite electrodes in a boron nitride tube inside a graphite mold, as
schematically shown in Fig. 1(a). The furnace was evacuated to approx. 3. Results and discussion
5 Pa and then heated to 1200 ◦ C at 10 ◦ C/min. After isothermal dwelling
for 15 min, an AC power (100 Hz) with a preset electric field strength (E, Fig. 1(b) shows the typical variations in electric field, current and
9 V/mm) and a current density (I, within 60–200 mA/mm2) was applied power with time during the RFS process. After a short incubation period
to the pellet. Once the flash was initiated, the AC power automatically (stage I), the current rapidly increased and the AC power changed from
switched from voltage control to current control. A pressure of 10 MPa voltage control to current control. Accordingly, the power showed a
was applied during the RFS process to ensure that the electrodes closely sharp peak (stage II) and then rapidly decreased to a stable value (stage
contacted the sample. Finally, the sample was cooled at 20 ◦ C/min to III). At a current density of 120 mA/mm2, the peak power in stage II
600 ◦ C and subsequently furnace-cooled to room temperature. reached 1115 mW/mm3. Such an instantaneous high peak power caused
The densities of the sintered samples were determined by the rapid heating and cooling in the sample, which was beneficial to the
Archimedes method. The crystal structures of the polished RFSed sam­ formation and stabilization of a high-entropy phase [13]. Note that the
ples (to remove the pollution from graphite) were identified by X-ray electric resistance (E/I) of the sample was fairly stable in stage III,
diffraction (XRD, D/Max 2500PC, Rigaku, Japan). The microstructures implying that the primary densification might have been finished in
were observed by using a field emission electron microscope (FE-SEM, stage II, since the electrical resistance was closely related to sintering
JSM-6700 F, JEOL, Japan) and a transmission electron microscope density and temperature [11,19]. In addition, according to the feedback
(TEM, JEM-2100 F, JEOL, Japan) equipped with an energy dispersive from pressure sensor during the RFS process (Fig. S1), the sample un­
spectrometer (EDS). The specimens for TEM were firstly mechanically derwent rapid shrinkage during the peak power stage (stage II), con­
ground to a thickness of approx. 50 μm, and then ion-beam milled firming that the densification mainly occurred during this period (within
(EMRES101, Leica, Germany) with a beam energy of 5 kV and an angle 5 s).
of 10◦ . Furthermore, an atom probe tomography (APT) measurement XRD examinations (Fig. 1(c)) show that there was no significant
was performed with a local electrode atom probe instrument (LEAP difference in peak patterns between the RFSed samples obtained under
4000X Si, Cameca, USA) at a target specimen temperature of 40 K under different current densities and the conventionally solid-state sintered
a pulsing UV laser with a pulsing energy of 60 pJ, a pulse rate of (hereafter abbreviated as CSSed) samples, indicating that the crystal­
200 kHz, and a detector efficiency of ~55 %. Needle-shaped APT sam­ lographic phases formed in both cases were similar. According to liter­
ples were prepared by using a focused ion beam (FIB, Zeiss Auriga SEM/ ature [8], the 5RE2Zr2O7 HEPO product has a pyrochlore structure with
FIB, Germany). Nano-indentation tests were carried out on a nano a lattice parameter of 10.625 Å. Here, the lattice parameters of the
indenter system (G200, KLA-Tencor, USA) with a Berkovich diamond as-prepared 5RE2Zr2O7 HEPOs were calculated to be in the range of

Fig. 1. (a) Schematic diagram of RFS device; (b) typical variations in electric field strength (E), current density (I) and power density (P) with time during RFS
process; (c) XRD patterns and (d) relative densities of the RFSed samples sintered under different current densities and the CSSed samples sintered at 1600 ◦ C for 3 h
and 6 h (The insert in (d) shows an image of the RFSed sample).

2856
H.-R. Mao et al. Journal of the European Ceramic Society 41 (2021) 2855–2860

10.615–10.628 Å, consisting with the reported result. 5RE2Zr2O7 HEPO had no elemental segregation at the grain boundaries.
Fig. 1(d) shows the sintered densities of the RFSed-5RE2Zr2O7 HEPOs Furthermore, atom probe tomography was also used to characterize the
under different current densities together with those of the CSSed- compositional homogeneity of the RFSed-5RE2Zr2O7 HEPO. The three-
5RE2Zr2O7 HEPOs prepared at 1600 ◦ C for 3 h and 6 h (the theoretical dimensional (3D) reconstructions of the atomic maps (Fig. 4(d)) for
density of 5RE2Zr2O7 was reported to be 6.554 g⋅ cm− 3 [8]). The relative the RFSed-5RE2Zr2O7 HEPO (under 120 mA/mm2) strongly confirmed
density of the RFSed samples increased noticeably with increasing cur­ the homogenous distribution of the compositions at the atomic scale.
rent density, reaching 99 % when the current density was 200 mA/mm2. As we have mentioned before, the mechanisms for ultrafast densifi­
For comparison, the relative density of the samples sintered in the same cation in reactive flash sintering have not yet been clarified so far. Here,
furnace at 1200 ◦ C for 15 min but without the application of the AC we suppose that the ultrafast densification of the RFSed-5RE2Zr2O7
power was 62 % (the porous structure is shown in Fig. S2(a)). Never­ HEPOs is presumably attributed to the following aspects: (1) The com­
theless, a high-entropy pyrochlore phase had also formed (Fig. S2(b)), bination of transient high temperature (caused by Joule heat effect)
suggesting that the formation of the HEPO phase is much easier than the during the peak power stage (stage II) and external pressure (10 MPa)
achievement of the high density. In contrast, the relative densities of the should be responsible for the rapid densification, as demonstrated by the
CSSed samples reached 94 % and 97 % after sintering at 1600 ◦ C for 3 h fact that the diameter of the RFSed samples was equivalent to the inner
and 6 h, respectively. In comparison, the RFSed samples show only diameter of the boron nitride protective tube in the graphite mold. (2)
uniformly dispersed micropores (Fig. 2(a–f)), while the CSSed samples The presence of lattice defects in the RFSed sample promoted the rapid
have scattered larger pores (Fig. 2(g, h)), which was related to the diffusion of ions. It is believed that a large number of Frenkel-pair de­
long-period high-temperature diffusion and merging of micropores. In fects such as positively charged oxygen vacancies and interstitial oxygen
addition, it is worth mentioning that the final diameters of the RFSed ions can be generated in the oxide powders under the application of an
samples were equivalent to the inner diameter (6.1 mm) of the boron electric field [14,20,21]. These high-energy charged lattice defects can
nitride protective tube in the graphite mold, suggesting that the HEPO greatly reduce the diffusion activation energy of the ions, thereby pro­
samples underwent plastic rheology under 10 MPa axial pressure during moting the rapid homogenization and densification of the high-entropy
the flash sintering process. Without the application of pressure, the oxides; (3) Rapid heating provided high sintering activity in the nano­
HEPO samples sintered at a current density of 150 mA/mm2 reached a particles. During the peak power stage (~650–1550 mW/mm3) in the
relative density of ~94 % and their diameters shrank to ~4.95 mm. RFS process, the samples underwent rapid heating (~103–104 ◦ C/min)
The fracture morphologies of all RFSed-5RE2Zr2O7 samples and and dramatic shrinkage. The rapid heating and short duration at the
CSSed-5RE2Zr2O7 sample (1600 ◦ C–6 h) show a flat transgranular frac­ maximum temperature avoided grain coarsening and neck growth
ture mode (Fig. 3(a–f)), indicating that they have relatively high encountered in traditional solid-state sintering process [22]; (4) Internal
strength at grain boundaries. As the current density increased from reactions provided additional chemical driving force. The RFS process
60 mA/mm2 to 200 mA/mm2, the porosity in the RFSed samples grad­ was accompanied with phase transformations and solid solution re­
ually decreased, and the average grain size increased from 0.6 μm to actions, which might provide additional impetus for the sintering.
2.4 μm. Note that the average grain size of the CSSed sample was much The mechanical properties of the RFSed- and CSSed-5RE2Zr2O7
larger, reaching 5.2 μm. The smaller grain sizes in the RFSed samples (1600 ◦ C-6 h) HEPOs were evaluated by using nano-indentation. Fig. 5
were mainly attributed to the ultrafast densification under rapid heating (b) shows the loading-unloading curves of the RFSed- and CSSed-
and external pressure. Fig. 3(g) shows the SEM image of the polished 5RE2Zr2O7 HEPOs. The nano-hardness and Young’s modulus of the
surface of the RFSed-5RE2Zr2O7 sample (under 200 mA/mm2) together RFSed HEPOs increased with increasing current density during sintering
with the corresponding EDS mapping graphs. All the components were (Fig. 5(c)). The RFSed sample sintered at 200 mA/mm2 showed a slight
homogeneously distributed over a wide range without noticeable decrease in hardness, presumably owing to the grain coarsening in its
segregation or clustering. microstructure (Fig. 3(e)). Similarly, the CSSed sample showed an even
Fig. 4(a) shows a TEM image of the RFSed-5RE2Zr2O7 sample (under lower hardness than the RFSed sample (under 200 mA/mm2) due to its
120 mA/mm2) at a triangular boundary junction. Each grain was ho­ larger grain size (Fig. 3(f)). It is noteworthy that the modulus of the
mogeneous without containing a secondary phase. Selected area elec­ CSSed sample was much lower than that of the RFSed samples. Jia et al.
tron diffraction (SAED) pattern (Fig. 4(b)) and HRTEM image (Fig. 4(c)) observed a similar phenomenon but did not explain the mechanism
confirmed that the RFSed-5RE2Zr2O7 sample had a single-phase pyro­ [23]. We speculate that this phenomenon may be related to the different
chlore structure. EDS analysis (Fig. S3) demonstrated that the pore structures in the RFSed samples and the CSSed samples (Fig. 2), or

Fig. 2. Optical micrographs of (a–f) the RFSed-5RE2Zr2O7 samples sintered under different current densities and (g–h) the CSSed-5RE2Zr2O7 samples sintered at
1600 ◦ C for 3 h and 6 h.

2857
H.-R. Mao et al. Journal of the European Ceramic Society 41 (2021) 2855–2860

Fig. 3. Fracture morphologies of (a–e) the RFSed-5RE2Zr2O7 samples under different current densities and (f) the CSSed-5RE2Zr2O7 sample (the insets show the
statistic diagrams of average grain size (AGS) obtained by using Nano Measurer software); (g) SEM image of the polished surface of the RFSed-5RE2Zr2O7 sample
(under 200 mA/mm2) together with the corresponding EDS mappings of Zr, La, Nd, O, Sm, Eu and Gd elements.

to the influence of internal crystal defects or stress. However, the toughness of the nearly fully dense RFSed-5RE2Zr2O7 HEPO sintered at
detailed mechanism for this result needs to be further explored. The 200 mA/mm2 was as high as 2.8 ± 0.2 MPa⋅m1/2, more than twice that
indentation toughness Kc of the 5RE2Zr2O7 HEPOs was calculated ac­ of the single-component rare-earth zirconate RE2Zr2O7 (RE = La, Nd,
cording to the following equation [24]: Sm, Eu, or Gd) [2,25]. In comparison, the fracture toughness of the
( / ) CCSed samples was much lower (1.1 ± 0.1 MPa⋅m1/2), which was
KC = 0.015(a/l)1/2 (E/Hv )2/3 P c3/2 (1) ascribed to the presence of relatively large pores (Fig. 2(h)). We suggest
that the superior mechanical properties of the RFSed-5RE2Zr2O7 HEPOs
where E is the Young’s modulus (GPa), Hv is the hardness (GPa), P is the are related to the current effect, rapid heating/cooling and uniform
load (N), l is the crack length (m), and a is the half-diagonal of the distribution of micropores, but the specific mechanism needs to be
indentation impression (c = l + a, as shown in Fig. 5(a)). The calculation further explored.
results (Fig. 5(d)) indicate that the fracture toughness of the RFSed-
5RE2Zr2O7 samples noticeably decreased with increasing current den­ 4. Conclusions
sity and kept a plateau after the current density exceeded 120 mA/mm2.
However, it is worth mentioning that the calculated toughness values for Nearly fully dense (La0.2Nd0.2Sm0.2Eu0.2Gd0.2)2Zr2O7 HEPOs were
the low-density RFSed samples (sintered at 60 and 80 mA/mm2, marked synthesized at a low furnace temperature in a short time by reactive
by brown dots in Fig. 5(d)) could have a large error since the cracks at flash sintering. The application of electric field, rapid heating, external
the indentation corners were inconspicuous. Surprisingly, the fracture pressure and internal reactions should play important roles in the rapid

2858
H.-R. Mao et al. Journal of the European Ceramic Society 41 (2021) 2855–2860

Fig. 4. (a) TEM image, (b) SAED pattern, (c) HRTEM image, and (d) 3D atomic distribution maps of the RFSed-5RE2Zr2O7 sample (under 120 mA/mm2).

Fig. 5. (a) Nano-indentation morphology of the RFSed-5RE2Zr2O7 HEPO (under 200 mA/mm2); (b) loading-unloading curves, (c) Young’s modulus, hardness and (d)
fracture toughness of the RFSed- and CSSed-5RE2Zr2O7 HEPOs (error bar represents standard deviation).

densification of the 5RE2Zr2O7 HEPOs. The resultant HEPOs had a Declaration of Competing Interest
uniform composition at the atomic scale and superior modulus, hardness
and fracture toughness to the CCSed counterparts. This work demon­ The authors declare that they have no known competing financial
strates that the RFS is a potentially advanced technique for the prepa­ interests or personal relationships that could have appeared to influence
ration of HEOs, especially for those with difficulty in achieving high the work reported in this paper.
densification.
Acknowledgments

We are grateful to Drs. Yu-Hua Liu, Liang Zhao and Gui-Xun Sun at

2859
H.-R. Mao et al. Journal of the European Ceramic Society 41 (2021) 2855–2860

Key Laboratory of Automobile Materials, Jilin University, for their helps [11] V. Avila, R. Raj, Reactive flash sintering of powders of four constituents into a
single phase of a complex oxide in a few seconds below 700 ◦ C, J. Am. Ceram. Soc.
in characterization and testing. This work is supported by the National
102 (2019) 6443–6448.
Natural Science Foundation of China (Nos. 51871110 and 51604156) [12] V. Avila, B. Yoon, R.R.I. Neto, R.S. Silva, S. Ghose, R. Raj, L.M. Jesus, Reactive flash
and the Changbai Mountain Scholars Program of Jilin Province (No. sintering of the complex oxide Li0.5La0.5TiO3 starting from an amorphous precursor
2015011). powder, Scripta Mater. 176 (2020) 78–82.
[13] B. Yoon, V. Avila, R. Raj, L.M. Jesus, Reactive flash sintering of the entropy-
stabilized oxide Mg0.2Ni0.2Co0.2Cu0.2Zn0.2O, Scripta Mater. 181 (2020) 48–52.
References [14] D. Liu, X. Peng, J. Liu, L. Chen, Y. Yang, L. An, Ultrafast synthesis of entropy-
stabilized oxide at room temperature, J. Eur. Ceram. Soc. 40 (2020) 2504–2508.
[1] C.M. Rost, E. Sachet, T. Borman, A. Moballegh, E.C. Dickey, D. Hou, J.L. Jones, [15] A. Kumar, G. Sharma, A. Aftab, M.I. Ahmad, Flash assisted synthesis and
S. Curtarolo, J.P. Maria, Entropy-stabilized oxides, Nat. Commun. 6 (2015) 8485. densification of five component high entropy oxide (Mg, Co, Cu, Ni, Zn)O at 350 ◦ C
[2] K. Ren, Q. Wang, G. Shao, X. Zhao, Y. Wang, Multicomponent high-entropy in 3 min, J. Eur. Ceram. Soc. 40 (2020) 3358–3362.
zirconates with comprehensive properties for advanced thermal barrier coating, [16] M. Biesuz, L. Spiridigliozzi, G. Dell’Agli, M. Bortolotti, V.M. Sglavo, Synthesis and
Scripta Mater. 178 (2020) 382–386. sintering of (Mg, Co, Ni, Cu, Zn)O entropy-stabilized oxides obtained by wet
[3] S. Jiang, T. Hu, J. Gild, N. Zhou, J. Nie, M. Qin, T. Harrington, K. Vecchio, J. Luo, chemical methods, J. Mater. Sci. 53 (2018) 8074–8085.
A new class of high-entropy perovskite oxides, Scripta Mater. 142 (2018) 116–120. [17] K.W. Wang, B.S. Ma, T. Lia, C.X. Xie, Z.Z. Sun, D.G. Liu, J.L. Liu, L.N. An,
[4] Z. Grzesik, G. Smoła, M. Miszczak, M. Stygar, J. Dąbrowa, M. Zajusz, K. Świerczek, Fabrication of high-entropy perovskite oxide by reactive flash sintering, Ceram. Int.
M. Danielewski, Defect structure and transport properties of (Co, Cr, Fe, Mn, 46 (2020) 18358–18361.
Ni)3O4 spinel structured high entropy oxide, J. Eur. Ceram. Soc. 40 (2020) [18] J. Liu, K. Ren, C.Y. Ma, H.L. Dua, Y.G. Wang, Dielectric and energy storage
835–839. properties of flash-sintered high-entropy (Bi0.2Na0.2K0.2Ba0.2Ca0.2)TiO3 ceramic,
[5] J. Gild, M. Samiee, J.L. Braun, T. Harrington, H. Vega, P.E. Hopkins, K. Vecchio, Ceram. Int. 46 (2020) 20576–20581.
J. Luo, High-entropy fluorite oxides, J. Eur. Ceram. Soc. 38 (2018) 3578–3584. [19] J.S.C. Francis, R. Raj, Influence of the field and the current limit on flash sintering
[6] K.B. Zhang, W.W. Li, J.J. Zeng, T. Deng, B.Z. Luo, H.B. Zhang, X.G. Huang, at isothermal furnace temperatures, J. Am. Ceram. Soc. 96 (2013) 2754–2758.
Preparation of (La0.2Nd0.2Sm0.2Gd0.2Yb0.2)2Zr2O7 high-entropy transparent [20] K. Ren, S.S. Huang, Y.J. Cao, G. Shao, Y.G. Wang, The densification behavior of
ceramic using combustion synthesized nanopowder, J. Alloys. Compd. 817 (2020), flash sintered BaTiO3, Scripta Mater. 186 (2020) 362–365.
153328. [21] J.S.C. Francis, M. Cologna, R. Raj, Particle size effects in flash sintering, J. Eur.
[7] Z. Zhao, H. Xiang, F.Z. Dai, Z. Peng, Y. Zhou, (La0.2Ce0.2Nd0.2Sm0.2Eu0.2)2Zr2O7: a Ceram. Soc. 32 (2012) 3129–3136.
novel high-entropy ceramic with low thermal conductivity and sluggish grain [22] W. Ji, B. Parker, S. Falco, J.Y. Zhang, Z.Y. Fu, R.I. Todd, Ultra-fast firing: effect of
growth rate, J. Mater. Sci. Technol. 35 (2019) 2647–2651. heating rate on sintering of 3YSZ, with and without an electric field, J. Eur. Ceram.
[8] F. Li, L. Zhou, J.X. Liu, Y. Liang, G.J. Zhang, High-entropy pyrochlores with low Soc. 37 (2017) 2547–2551.
thermal conductivity for thermal barrier coating materials, J. Adv. Ceram. 8 (2019) [23] Y.J. Jia, X.H. Su, Y.J. Wu, Z.J. Wang, L.C. Meng, X.Q. Xu, A.N. Zhang, Flash
576–582. sintering of 3YSZ/Al2O3-platelet composites, J. Am. Ceram. Soc. 103 (2020)
[9] Z. Teng, L.N. Zhu, Y.Q. Tan, S.F. Zeng, Y.H. Xia, Y.G. Wang, H.B. Zhang, Synthesis 2351–2361.
and structures of high-entropy pyrochlore oxides, J. Eur. Ceram. Soc. 40 (2020) [24] R.D. Dukino, M.V. Swain, Comparative measurement of indentation fracture
1639–1643. toughness with berkovich and vickers indenters, J. Am. Ceram. Soc. 75 (1992)
[10] B. Yoon, D.Y.S. Ghose, R. Raj, Reactive flash sintering: MgO and α-Al2O3 transform 3299–3304.
and sinter into single-phase polycrystals of MgAl2O4, J. Am. Ceram. Soc. 102 [25] J. Feng, B. Xiao, C.L. Wan, Z.X. Qu, Z.C. Huang, J.C. Chen, R. Zhou, W. Pan,
(2019) 2294–2303. Electronic structure, mechanical properties and thermal conductivity of Ln2Zr2O7
(Ln = La, Pr, Nd, Sm, Eu and Gd) pyrochlore, Acta Mater. 59 (2011) 1742–1760.

2860

You might also like