Next Article in Journal
Interfacial Microstructure and Formation of Direct Laser Welded CFRP/Ti-6Al-4V Joint
Previous Article in Journal
Understanding the Role of β Recrystallization on β Microtexture Evolution in Hot Processing of a Near-β Titanium Alloy (Ti-10V-2Fe-3Al)
Previous Article in Special Issue
Theoretical Study on Thermoelectric Properties and Doping Regulation of Mg3X2 (X = As, Sb, Bi)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Local Lattice Distortion in High-Entropy Carbide Ceramics

1
Beijing Advanced Innovation Center for Materials Genome Engineering, State Key Laboratory for Advanced Metals and Materials, University of Science and Technology Beijing, Beijing 100083, China
2
Institute for Applied Physics, University of Science and Technology Beijing, Beijing 100083, China
3
College of Physics and Telecommunication Engineering, Zhoukou Normal University, Zhoukou 466001, China
*
Authors to whom correspondence should be addressed.
Metals 2021, 11(9), 1399; https://doi.org/10.3390/met11091399
Submission received: 10 August 2021 / Revised: 27 August 2021 / Accepted: 31 August 2021 / Published: 3 September 2021
(This article belongs to the Special Issue Advances in First-Principles Calculations on Metallic Materials)

Abstract

:
Using the ab initio calculations, we study the lattice distortion of HfNbTaTiVC5, HfNbTaTiZrC5 and MoNbTaTiVC5 high-entropy carbide (HEC) ceramics. Results indicate that the Bader atomic radius and charge transfer in HECs is very close to those from binary carbide. The degree of lattice distortion strongly depends on the alloying element. The Bader atomic radius can excellently describe the lattice distortion in HEC. Further, the corresponding atomic radius and formation enthalpy of binary carbides may be indicators to predict the single-phase HECs.

1. Introduction

High-entropy carbides (HECs) [1,2,3,4,5,6], as one of the high entropy materials, are being paid more attention due to their excellent properties, for instance, the high hardness, the remarkable oxidation resistance, and the adjustable range of thermal conductivity, etc. The high configuration entropy [7], as one of the key factors, is considered to induce excellent performance in high-entropy materials, including high-entropy alloys (HEAs) [8,9], high-entropy oxides (HEOs) [10], and high entropy diborides (HEBs) [11].
(Ti0.2Zr0.2Hf0.2Nb0.2Ta0.2)C powder adopts the two face-centered cubic (fcc) nested salt rock structured solid solution. It is thermodynamically more stable than its original components [1]. Using the descriptor of entropy forming ability and experimental method, Sarker et al. found the nine single phase HECs with the high hardness [12], etc. [13,14,15,16,17]. Herein, the cation elements are composed of refractory metals (Ti, Zr, Hf, V, Nb, Ta, Mo and W). Different from single phase equimolar quinary HEAs with Sconfig ~1.61 R, the salt-rock HEC structure has two sublattices, and the configuration entropy on the sublattice model is ~0.8 R. While in order to show the high configuration entropy of HECs, the quinary HECs have Sconfig ~1.61 R, according to the new configuration metric [18]. To study the effect of alloying elements on the properties of HECs, the (Ti0.33Zr0.33Nb0.33)C ternary carbide [19], (NbTaZr)C-based HECs [20] and (Hf0.25Ta0.25Zr0.25Ti0.25)C, (Hf0.25Ta0.25Zr0.25Nb0.25)C, (Zr0.25Ta0.25Nb0.25Ti0.25)C quaternary HECs were reported [21]. Li et al. found that in the quinary HECs, the W and/or Hf addition produces a large effect on the hardness and fracture toughness of (NbTaZr)C-based HECs [20].
As one of four core effects in HEAs, the lattice distortion is considered as a key role to enhance the performance. For HECs, these are used in the theoretical simulation to study the lattice distortion. Ye et al. found that the anion C atom severely deviates from its ideal lattice sites based on the ab initio calculations [16,22]. In this work, we propose the general formula to describe the local lattice distortion in HECs based on the Bader atomic radius derived from the local chemical environment. Interestingly, we find that the Bader atomic radius of MC (M = refractory metals) can be used to predict the lattice distortion and the formation of simple phases on HECs.

2. Methodology

High-entropy carbide ceramics adopt the NaCl structure, where there are two different Wyckoff position sites. Cl sites (0, 0, 0) are fully occupied by C atoms. Na positions (0.5, 0.5, 0.5) are occupied randomly by five refractory metal elements. To describe the disorder in HECs, we used the similar atomic environment (SAE) method [23,24] to construct a 3 × 3 × 5 supercell for quinary equimolar HECs based on the primary cell of the NaCl structure (a = b = c, α = β = γ = 60°) (see Figure 1). The size of the supercell was determined based on the alloying chemical composition and our available computer resources. Different from the effective medium method, the supercell method is often applied to the specific chemical composition of an alloy. The SAE method was applied to simulate the disorder and short-range order in high entropy alloys [25,26]. We took into account the number of nearest-neighboring (NN) atomic pairs up to the 7th NN distance, while the three-atom cluster is about the 3rd NN distance. The SAE supercell has 90 atoms, in which the number of C atoms is 45, and the number of atoms for each refractory metal is 9. In the support material, we list the initial and relaxed SAE structure parameter and the corresponding cartesian coordinates of atoms.
To obtain the optimized structures of HECs, we employed the Vienna ab initio simulation package (VASP) to perform the electronic density of state [27]. The exchange–correlation potentials were treated by using the Perdew-Burkey-Ernzerh functional in the framework of the generalized gradient approximation [28]. The ion–electron interactions were performed by using the projector augmented wave (PAW) method [29]. The valence electron configuration of element is 3s23p63d24s2 for Ti, 3s23p63d34s2 for V, 4s24p64d25s2 for Zr, 4s24p64d35s2 for Nb, 4s24p64d45s2 for Mo, 5p65d26s2 for Hf, and 5p65d36s2 for Ta. The Brillouin zone sampling was performed using the special k points generated by the Monkhorst–Pack scheme with a density parameter of 0.2 Å−1. The plane wave energy cutoff of 400 eV provided accurate results. The tolerances of energy and maximum force were set to 1.0 × 10−5 eV and 2.0 × 10−2 eV Å−1, respectively. Constant pressure and full relaxations with Methfessel–Paxton smearing (with a width of 0.2 eV) were used. With the Bader method [30], the atomic radius and charge transfer were extracted from the ab initio electronic density of state.

3. Results and Discussion

Considering that the MC binary carbides (M is the refractory metals, except for MoC and WC) have the same crystal structure as HECs, we first compare the atomic radius and charge transfer in MC and HECs. Herein, we employed the Wigner–Seitz (WS) radius to describe the complex bonding topology of each alloying element in HECs. Figure 2 shows the Bader WS radius of MC (M = Hf, Mo, Nb, Ta, Ti, V, Zr). We can see that the atomic radius of carbon strongly depends on the bonding in MC, with a change from 1.269 to 1.433 Å. Among the MC carbides, the carbon in VC has the smallest atomic radius, while the atomic radius of carbon is the largest for ZrC. With the increase in atomic number, the atomic radius difference between carbon and the refractory metal becomes large. Taking the MC carbide as the reference ground structure, we calculated the formation enthalpy ΔH1 of HECs listed in Table 1. ΔH2 was calculated based on the energy of graphite, HCP Ti, Zr, Hf and BCC V, Nb, Ta, and Mo. Experimentally, it has been found that the enthalpy of graphite is 0.019 eV lower than the enthalpy of diamond. To obtain E(C), we calculated the energy of diamond and shifted it by applying the experimental difference value. Compared to ΔH2, ΔH1 has a less negative value, which is close to the range of the single phase HEAs [31].
We show in Figure 3 the Bader WS radius and charge transfer of each element for HECs. From the right panel in Figure 3, we can find that the charge transfer between elements in HECs is consistent with that of per element in binary carbides. The charge transfer mainly occurs between carbons and refractory metal elements. The carbons obtain about ~1.2 average charge from the refractory metal elements, while the refractory metal elements have the discrete values of charge transfer. It suggests that the charge transfer on refractory metals is sensitive to the local chemical environment (different bonding lengths), but the average value of charge transfer on refractory metals in HECs is very close to that in MC. Similarly, we can see from the left panel in Figure 3 that the Bader atomic radius of carbon is a small deviation range of 0.02 Å, whereas the Bader atomic radius of refractory metals depends on the local chemical environment. The present HECs all contain Nb, Ta and Ti refractory elements, but the Bader atomic radius is about the range of 1.385–1.522 Å for Nb in HfNbTaTiVC5, while the change of range region is only 0.022 Å for Nb in MoNbTaTiVC5. The average Bader WS radius of each element in HECs is close to the corresponding radius in MC. From Figure 3, we can see that the change of the Bader atomic radius is consistent with the charge transfer. Further, one may use the atomic radius and charge transfer of MC binary alloys to represent the corresponding information of HECs. For example, the formula of the atomic radius difference is defined as δ = 100 i = 1 n c i ( 1 r i / r ¯ ) , where r ¯ = i = 1 n c i r i , with ci and ri being the atomic percentage and the Bader atomic radius of the ith alloying element, respectively. According to its key physical parameter, we obtain the degree of atomic misfitting of 4.514% for HfNbTaTiVC5 and 3.884% for HfNbTaTiZrC5. This tendency is consistent with the available experiments [12] (see Table 1).
We also define the local lattice distortion in terms of the change of atomic coordinate position with respect to the ideal lattice sites
Δ d = 1 N i ( x i x i ) 2 + ( y i y i ) 2 + ( z i z i ) 2
where N is the number of atoms, and (xi, yi, zi) and (xi, y′i, z′i) are the reduced coordinates of the ideal and distorted positions of each atom for individual alloy component. The local lattice distortion of some individual alloy components depends on the alloy system, for instance, the local lattice distortion of Ti is 0.050 Å in HfNbTaTiVC5, whereas 0.027 Å in MoNbTaTiVC5. From Table 1 and Figure 4, we find that Hf, Ta and Zr have small local lattice distortion. While C, Mo and V have severe lattice distortion. The large change of atomic coordinate induces the severe lattice distortion of 0.059 Å for HfNbTaTiVC5 and 0.060 Å for MoNbTaTiVC5. Although the local lattice distortion of carbon is up to 0.062 Å, and other alloy components have small lattice distortion, it results in small lattice distortion of HfNbTaTiZrC5. We can find that Δ d is consistent with experiments based on the peak broadening in XRD patterns [12]. The local lattice distortions of Nb, Ta, and Ti strongly depend on the nearest chemical environment (see Figure 4). For example, the Nb atoms in HfNbTaTiVC5 have large lattice distortion with respect to those in MoNbTaTiVC5 and HfNbTaTiZrC5. Note that the severely local lattice distortion of V and Mo does not only occur in the first nearest neighboring shell but also in other nearest neighboring shells.
Table 2 lists the lattice parameters of the present HECs with and without lattice distortion. Interestingly, the lattice parameter is not sensitive to the local lattice distortion. The calculated lattice parameters are consistent with the available experiments (the difference is within 0.8%). However, our calculation was performed at 0 K. Considering the temperature effect on the lattice thermal expansion, our results are larger than experiments. One possible reason is our exchange-correlation potential of PBE. It often overestimates the lattice parameter of a material. Another possible reason is the carbon vacancy that existed in the present HECs. The vacancy results in a decrease in the lattice parameter in the experiment, whereas our calculations are based on the ideal carbon composition.

4. Conclusions

In summary, we have employed the ab initio calculation based on density functional theory in combination with the similar atomic environment to study the local lattice distortion of the HfNbTaTiVC5, HfNbTaTiVZrC5, and MoNbTaTiVC5 high-entropy carbide ceramics. The binary carbide has a similar Bader radius and charge transfer between C and refractory metals to the present HECs. Based on the physical parameter of the binary carbides, the formation enthalpy and classical atomic radius difference can be used to predict the single-phase formation and lattice distortion. HfNbTaTiZrC5 has a larger local lattice distortion than HfNbTaTiVC5 and MoNbTaTiVC5. The alloying element V and Mo show the large local lattice distortion in their first nearest neighboring shell and other neighboring shells.

Author Contributions

H.G. and C.C.: conceptualization, data curation, formal analysis, and writing—original draft. H.S. and F.T.: formal analysis, funding acquisition, resources, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 51831003, 51527801, 51771015, U1804123), the Funds for Creative Research Groups of China (Grant No. 51921001), the National High-level Personnel of Special Support Program (Grant No. ZYZZ2021001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, J.; Zhang, J.; Zhang, F.; Niu, B.; Lei, L.; Wang, W. High-entropy carbide: A novel class of multicomponent ceramics. Ceram. Int. 2018, 44, 22014–22018. [Google Scholar] [CrossRef]
  2. Nisar, A.; Zhang, C.; Boesl, B.; Agarwal, A. A perspective on challenges and opportunities in developing high entropy-ultra high temperature ceramics. Ceram. Int. 2020, 46, 25845–25853. [Google Scholar] [CrossRef]
  3. Xiang, H.; Xing, Y.; Dai, F.; Wang, H.; Su, L.; Wang, H.; Zhao, B.; Li, J.; Zhou, Y. High-entropy ceramics: Present status, challenges, and a look forward. J. Adv. Ceram. 2021, 10, 385–441. [Google Scholar] [CrossRef]
  4. Oses, C.; Toher, C.; Curtarolo, S. High-entropy ceramics. Nat. Rev. Mater. 2020, 5, 295–309. [Google Scholar] [CrossRef]
  5. Alvi, S.A.; Zhang, H.; Akhtar, F. High-Entropy Ceramics. In Engineering Steels and High Entropy-Alloys; IntechOpen: London, UK, 2019. [Google Scholar]
  6. Wang, Z.; Li, Z.T.; Zhao, S.J.; Wu, Z.G. High-entropy carbide ceramics: A perspective review. Tungsten 2021, 3, 131–142. [Google Scholar] [CrossRef]
  7. Lei, Z.; Liu, X.; Wang, H.; Wu, Y.; Jiang, S.; Lu, Z. Development of advanced materials via entropy engineering. Scr. Mater. 2019, 165, 164–169. [Google Scholar] [CrossRef]
  8. Cantor, B.; Chang, I.T.H.; Knight, P.; Vincent, A.J.B. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  9. Yeh, J.W.; 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]
  10. Rost, C.M.; Sachet, E.; Borman, T.; Moballegh, A.; Dickey, E.C.; Hou, D.; Jones, J.L.; Curtarolo, S.; Maria, J. Entropy-stabilized oxides. Nat. Commun. 2015, 6, 1–8. [Google Scholar] [CrossRef] [Green Version]
  11. Gild, J.; Zhang, Y.; Harrington, T.; Jiang, S.; Hu, T.; Quinn, M.C.; Mellor, W.M.; Zhou, N.; Vecchio, K.; Luo, J. High-Entropy Metal Diborides: A New Class of High-Entropy Materials and a New Type of Ultrahigh Temperature Ceramics. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef]
  12. Sarker, P.; Harrington, T.; Toher, C.; Oses, C.; Samiee, M.; Maria, J.P.; Brenner, D.W.; Vecchio, K.S.; Curtarolo, S. High-entropy high-hardness metal carbides discovered by entropy descriptors. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wei, X.F.; Liu, J.X.; Li, F.; Qin, Y.; Liang, Y.C.; Zhang, G.J. High entropy carbide ceramics from different starting materials. J. Eur. Ceram. Soc. 2019, 39, 2989–2994. [Google Scholar] [CrossRef]
  14. Liu, D.; Zhang, A.; Jia, J.; Meng, J.; Su, B. Phase evolution and properties of (VNbTaMoW) C high entropy carbide prepared by reaction synthesis. J. Eur. Ceram. Soc. 2020, 40, 2746–2751. [Google Scholar] [CrossRef]
  15. Dai, F.Z.; Wen, B.; Sun, Y.; Xiang, H.; Zhou, Y. Theoretical prediction on thermal and mechanical properties of high entropy (Zr0.2Hf0.2Ti0.2Nb0.2Ta0.2)C by deep learning potential. J. Mater. Sci. Technol. 2020, 43, 168–174. [Google Scholar] [CrossRef]
  16. Ye, B.; Wen, T.; Huang, K.; Wang, C.Z.; Chu, Y. First-principles study, fabrication, and characterization of (Hf0.2Zr0.2Ta0.2Nb0.2Ti0.2 )C high-entropy ceramic. J. Am. Ceram. Soc. 2019, 102, 4344–4352. [Google Scholar] [CrossRef] [Green Version]
  17. Harrington, T.J.; Gild, J.; Sarker, P.; Toher, C.; Rost, C.M.; Dippo, O.F.; McElfresh, C.; Kaufmann, K.; Marin, E.; Borowski, L.; et al. Phase stability and mechanical properties of novel high entropy transition metal carbides. Acta Mater. 2019, 166, 271–280. [Google Scholar] [CrossRef] [Green Version]
  18. Dippo, O.F.; Vecchio, K.S. A universal configurational entropy metric for high-entropy materials. Scr. Mater. 2021, 201, 113974. [Google Scholar] [CrossRef]
  19. Ye, B.; Chu, Y.; Huang, K.; Liu, D. Synthesis and characterization of (Zr1/3Nb1/3Ti1/3)C metal carbide solid-solution ceramic. J. Am. Ceram. Soc. 2019, 102, 919–923. [Google Scholar] [CrossRef]
  20. Li, Z.; Wang, Z.; Wu, Z.; Xu, B.; Zhao, S.; Zhang, W.; Lin, N. Phase, microstructure and related mechanical properties of a series of (NbTaZr)C-Based high entropy ceramics. Ceram. Int. 2021, 47, 14341–14347. [Google Scholar] [CrossRef]
  21. Castle, E.; Csanádi, T.; Grasso, S.; Dusza, J.; Reece, M. Processing and Properties of High-Entropy Ultra-High Temperature Carbides. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zhao, S. Lattice distortion in high-entropy carbide ceramics from first-principles calculations. J. Am. Ceram. Soc. 2021, 104, 1874–1886. [Google Scholar] [CrossRef]
  23. Song, H.; Tian, F.; Hu, Q.; Vitos, L.; Wang, Y.; Shen, J.; Chen, N. Local lattice distortion in high-entropy alloys. Phys. Rev. Mater. 2017, 1, 023404. [Google Scholar] [CrossRef] [Green Version]
  24. Tian, F.; Lin, D.Y.; Gao, X.; Zhao, Y.F.; Song, H.F. A structural modeling approach to solid solutions based on the similar atomic environment. J. Chem. Phys. 2020, 153, 034101. [Google Scholar] [CrossRef]
  25. Guan, H.; Huang, S.; Ding, J.; Tian, F.; Xu, Q.; Zhao, J. Chemical environment and magnetic moment effects on point defect formations in CoCrNi-based concentrated solid-solution alloys. Acta Mater. 2020, 187, 122–134. [Google Scholar] [CrossRef]
  26. Tian, F.; Yang, Z.; Lin, D.Y.; Zhao, Y.F. Lattice distortion inducing local antiferromagnetic behaviors in FeAl alloys. J. Phys. Condens. Matter 2020, 32. [Google Scholar] [CrossRef]
  27. Kresse, G.; Marsman, M.; Furthmüller, J. Vienna Ab-initio Package Vienna Simulation: VASP the GUIDE; VASP Software GmbH: Vienna, Austria, 2016. [Google Scholar]
  28. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 2009, 21, 084204. [Google Scholar] [CrossRef]
  31. Tian, F.; Varga, L.K.; Chen, N.; Shen, J.; Vitos, L. Empirical design of single phase high-entropy alloys with high hardness. Intermetallics 2015, 58, 1–6. [Google Scholar] [CrossRef]
Figure 1. The similar atomic environment supercell containing 90 atoms for HEC, enlarged by the 3 × 3 × 5 primitive rock-salt structure.
Figure 1. The similar atomic environment supercell containing 90 atoms for HEC, enlarged by the 3 × 3 × 5 primitive rock-salt structure.
Metals 11 01399 g001
Figure 2. The Bader WS radius (Å) for MC (M = Hf, Mo, Nb, Ta, Ti, V, Zr).
Figure 2. The Bader WS radius (Å) for MC (M = Hf, Mo, Nb, Ta, Ti, V, Zr).
Metals 11 01399 g002
Figure 3. Shown are the Bader WS radius and the Bader charge transfer of individual alloy components.
Figure 3. Shown are the Bader WS radius and the Bader charge transfer of individual alloy components.
Metals 11 01399 g003
Figure 4. The number of atomic pairs as a function of the radial distribution distance for the present HEC with lattice distortion. The black vertical dot lines stand for the ideal lattice position of different nearest neighboring atomic pairs.
Figure 4. The number of atomic pairs as a function of the radial distribution distance for the present HEC with lattice distortion. The black vertical dot lines stand for the ideal lattice position of different nearest neighboring atomic pairs.
Metals 11 01399 g004
Table 1. Listed are the formation enthalpy ΔH (in KJ/mol), average radial displacement Δd (in Å) for individual alloy components, total radial displacement tot., and atomic misfitting difference δ in the considered HECs. The available experiments are from Reference [12]. ΔH1  = E H E C i c i E M C , with EMC for the energy of binary carbides, ΔH2 = E H E C i c i E i , with ci for the composition and Ei for the energy of ground state of elements (graphite for HCP, Hf, Ti, Zr, BCC for Mo, Nb, Ta, V).
Table 1. Listed are the formation enthalpy ΔH (in KJ/mol), average radial displacement Δd (in Å) for individual alloy components, total radial displacement tot., and atomic misfitting difference δ in the considered HECs. The available experiments are from Reference [12]. ΔH1  = E H E C i c i E M C , with EMC for the energy of binary carbides, ΔH2 = E H E C i c i E i , with ci for the composition and Ei for the energy of ground state of elements (graphite for HCP, Hf, Ti, Zr, BCC for Mo, Nb, Ta, V).
HECsΔH1ΔH2ΔdδExp.
CHfMoNbTaTiVZrtot.
HfNbTaTiVC51.330−64.850.0620.045--0.0650.0390.0500.084--0.0594.5140.107
HfNbTaTiZrC5−0.902−74.820.0620.039--0.0330.0380.027--0.0370.0483.8840.094
MoNbTaTiVC5−1.276−47.410.049--0.1020.0460.0440.0430.105--0.0604.223--
Table 2. The converted lattice parameter L of the equilibrium state (Å), the available experiments, lattice parameters (a, b, c and α, β, γ) of the SAE supercell for the high entropy carbides (HECs) with/without distortion.
Table 2. The converted lattice parameter L of the equilibrium state (Å), the available experiments, lattice parameters (a, b, c and α, β, γ) of the SAE supercell for the high entropy carbides (HECs) with/without distortion.
HECsLExp.LLDIdeal
abcαβγabc
HfNbTaTiVC54.4324.4159.4039.40215.67759.92359.96660.0129.4209.42015.700
HfNbTaTiZrC54.5194.5009.5899.59115.97359.97159.98259.9859.5859.58515.975
MoNbTaTiVC54.375--9.2829.27815.46159.99060.07159.9929.2859.28515.475
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ge, H.; Cui, C.; Song, H.; Tian, F. Local Lattice Distortion in High-Entropy Carbide Ceramics. Metals 2021, 11, 1399. https://doi.org/10.3390/met11091399

AMA Style

Ge H, Cui C, Song H, Tian F. Local Lattice Distortion in High-Entropy Carbide Ceramics. Metals. 2021; 11(9):1399. https://doi.org/10.3390/met11091399

Chicago/Turabian Style

Ge, Huijuan, Chengfeng Cui, Hongquan Song, and Fuyang Tian. 2021. "Local Lattice Distortion in High-Entropy Carbide Ceramics" Metals 11, no. 9: 1399. https://doi.org/10.3390/met11091399

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