Next Article in Journal
Effect of Heat Treatment on Microstructures and Mechanical Properties of a Ti-Al-V-Cr-Fe-Based Alloy
Previous Article in Journal
Tuning Electromagnetically Induced Transparency in a Double GaAs/AlGaAs Quantum Well with Modulated Doping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Crystallography and Morphology of (Gd,Y)H2 Hydride in a Mg-Gd-Y-Al Alloy

1
National Engineering Research Center of Light Alloy Net Forming and State Key Laboratory of Metal Matrix Composite, Shanghai Jiao Tong University, Shanghai 200240, China
2
Centre for Additive Manufacturing, School of Engineering, RMIT University, Melbourne, VIC 3053, Australia
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(3), 249; https://doi.org/10.3390/cryst15030249
Submission received: 20 February 2025 / Revised: 4 March 2025 / Accepted: 4 March 2025 / Published: 6 March 2025
(This article belongs to the Special Issue Preparation and Characterization of Precipitates in Magnesium Alloys)

Abstract

:
Hydrogen can be easily captured by the rare-earth (RE) elements in hydrogen-rich environments, which significantly affect the phase compositions and mechanical performance of Mg-RE based alloys. However, the morphology of RE hydrides and their orientation relationships (ORs) with the Mg matrix have not been well explained. Here, a stable face-centered cubic (FCC) G d , Y H 2 hydride was introduced and uniformly distributed in a Mg-15Gd-2.5Y-1Al alloy after hydrogenation treatment at 500 °C and 2 MPa for 40 h. The plate-like G d , Y H 2 hydride with six variants was identified to exhibit an OR with the magnesium (Mg) matrix, which is [ 0001 ] M g // [ 001 ] G d , Y H 2 , ( 10 1 ¯ 0 ) M g 10.5 from ( 002 ) G d , Y H 2 , ( 1 2 ¯ 10 ) M g 10.5 from ( 020 ) G d , Y H 2 . Further crystallographic matching calculations based on the edge-to-edge matching model suggest that such an OR is energetically favorable and provides the actual interface between the RE hydrides and the Mg matrix during precipitation. Our findings offer new insights into the microstructural regulation of Mg alloys in hydrogenation environments.

1. Introduction

Hydrogen has garnered significant attention as a promising alternative to fossil fuels in the emerging energy sector [1]. Metals play a crucial role in the transportation, storage, and utilization of hydrogen [2]. However, hydrogen can significantly reduce the fracture toughness and ductility of metals and thus lead to catastrophic failure, which is known as hydrogen embrittlement [3]. This poses a considerable risk to the widespread and safe application of hydrogen. Although hydrogen embrittlement has been intensively investigated in steel [4,5], aluminum alloys [6,7], and titanium alloys [8,9], the relevant research on the lightest structural metal—magnesium (Mg)—alloys [10,11] is far from sufficient. Although it has been reported that hydrogen segregation at grain boundaries can lead to premature intergranular cracking in Mg alloys [12,13], some recent studies have suggested that the ductility of hydrogen-treated rare-earth (RE) Mg alloys can be significantly improved [14,15,16]. This is mainly attributed to the high affinity of RE elements for hydrogen and the formation of RE hydrides [17,18]. In other words, hydrogenation treatment of Mg-RE alloys will dramatically change the phase compositions and microstructures. As a secondary phase, the morphology of RE hydrides and their orientation relationships (ORs) with the Mg matrix can significantly affect the mechanical properties of the alloys [19]. However, while some research has focused on the morphology of RE hydrides [20,21,22], there is still a significant gap in understanding their ORs with the Mg matrix.
In this work, a Mg-Gd-Y-Al alloy was selected to investigate the precipitation behavior of Mg alloys following hydrogen treatment. Gadolinium (Gd) and yttrium (Y) are the most commonly used RE elements in Mg alloys due to their high solubility, which results in remarkable solid solution strengthening and precipitation strengthening effects [23,24]. They were also reported to readily combine with hydrogen to form dihydride [17,20]. To gain higher strength and corrosion resistance, Al was added to the alloy to refine the grain size and form a passive film [25,26,27]. The microstructure of this Mg-Gd-Y-Al alloy before and after hydrogenation treatment was studied. The morphological crystallography of RE hydrides was comprehensively characterized by X-ray diffraction (XRD), electron backscattered diffraction (EBSD), and transmission electron microscopy (TEM) and rationalized by contemporary interphase crystallographic models.

2. Materials and Methods

2.1. Materials

The Mg-Gd-Y-Al alloy was prepared with pure Mg (99.9%), Mg-30Gd master alloy (99.9%), Mg-30Y master alloy (99.9%), and pure Al (99.9%) using a gravity casting method under the protection of C O 2 and S F 6 . The chemical composition of the alloy was measured to be Mg-15Gd-2.5Y-1Al (wt.%) using an inductively coupled plasma optical emission spectrometer (Avio 500, Perkin Elmer, Waltham, MA, USA). Hydrogenation experiments were conducted using an Ultra High Pressure Gas Sorption Analyzer (BSD-PHU, Beishide Company, Beijing, China) with high-purity H₂ (99.999%) as the adsorption gas. Cubic samples with dimensions of 10 mm * 10 mm * 10 mm were cut from the casting alloy and then placed in the hydrogen adsorption chamber. The samples were treated with hydrogen permeation conditions at 500 °C and 2 MPa for 40 h.

2.2. Methods

The phase composition of the alloy before and after hydrogenation was analyzed by XRD and EBSD. The EBSD sample was sequentially polished using 320#, 800#, and 2000# grit sandpaper, followed by polishing with 3 μm, 1 μm, and 0.05 μm polishing suspensions. The ORs between the hydrides and the Mg matrix were determined by EBSD and TEM (JEOL ARM200F, JEOL, Tokyo Metropolis, Japan). The sample was thinned to ~60 μm using 1200#, 3000#, and 7000# grit sandpaper. Double jet polishing was carried out using an electrolytic double jet polisher with a 4 vol.% perchloric acid in alcohol solution at a working voltage of 40 V. Subsequently, the sample was further thinned for 20 min using an ion thinning instrument under conditions of ± 2.5 ° tilt and 1 keV energy. The crystallographic matching between the Mg matrix and the RE hydrides was analyzed using the edge-to-edge matching (E2EM) model, which has been utilized to explain and predict the ORs between precipitates and matrices in numerous alloy systems [28]. The fundamental principles of this model have been previously described [29]. By inputting the crystal structures and lattice parameters of both the Mg matrix and the RE hydrides, the energetically favorable ORs can be identified.

3. Results and Discussion

Figure 1 presents the XRD results of the Mg-Gd-Y-Al alloy before and after hydrogenation. The as-cast alloy contains two types of eutectic phases, namely M g 24 ( G d , Y ) 5 and A l 2 ( G d , Y ) . After hydrogenation, both of these second phases disappeared, while two types of hydrides, i.e., M g H 2 and G d , Y H 2 , were observed. Figure 2 summarizes the phase maps, inverse pole figure (IPF), and pole figure (PF) of the alloy before and after hydrogenation. As shown in Figure 2a, M g 24 ( G d , Y ) 5 , indicated in cyan color, mainly distributes along the grain boundaries or at the grain boundary junctions. The A l 2 ( G d , Y ) phase is randomly dispersed in the intergranular or intragranular regions. Figure 2b,c show the phase composition after hydrogenation. Consistent with the XRD results, the original second phases in the as-cast alloy disappeared, while M g H 2 and G d , Y H 2 formed. It can be seen that most of the hydrides are G d , Y H 2 and very few are M g H 2 (as indicated by the white circle in Figure 2c). This suggests that the hydrogenation of the Mg matrix is greatly inhibited under the present hydrogenation conditions and most of the hydrogen is bonded to the RE elements. Figure 2c,d show the magnified phase map and IPF of the region highlighted by the white box in Figure 2b, respectively. As highlighted in the black box, a large number of plate-like G d , Y H 2 phases precipitated from the Mg matrix. The {0001}-Mg PF and {001}- G d , Y H 2 PF in Figure 2e suggest that the G d , Y H 2 phase shows a specific OR with the Mg matrix with six different variants, and these six variants are sequentially numbered from 1 to 6. All these variants share a common (001) pole point, which coincides with the projection of the (0001) pole point of the Mg matrix. Details of the OR will be discussed later. In addition to intragranular precipitated G d , Y H 2 , continuous distributions of the G d , Y H 2 phase along the grain boundaries and hydride clusters were also observed. It is likely that the original M g 24 ( G d , Y ) 5 and A l 2 G d , Y served as precursors for the formation of these G d , Y H 2 precipitates.
The OR between the intragranular plate-like G d , Y H 2 and the Mg matrix was further investigated by TEM. Figure 3a shows the bright-field (BF) TEM image of the sample observed along the [ 0001 ] M g direction. The average thickness of the plate-like G d , Y H 2 is approximately 100 nm. As indicated by the blue lines, six different variants can be observed within a single grain. Herein, three of these variants are selected, as boxed in red, and the corresponding Selective Area Electron Diffraction (SAED) patterns are given in Figure 3b–d. As can be seen from Figure 3b, the four diffraction spots closest to the transmitted beam connected by a green square confirm that the FCC-structured G d , Y H 2 phase is oriented along the [001] direction. As indicated by the yellow dashed line, the ( 002 ) plane of G d , Y H 2 and the ( 0 1 ¯ 10 ) plane of the Mg matrix are misaligned by an angle of 10.1°. Thus, the OR can be expressed as follows: [ 0001 ] M g // [ 001 ] G d , Y H 2 , ( 0 1 ¯ 10 ) M g 10.1 from ( 002 ) G d , Y H 2 , ( 2 11 ¯ 0 ) M g 10.1 from ( 020 ) G d , Y H 2 . Similar to that in Figure 3b, the OR in Figure 3c can be represented as [ 0001 ] M g // [ 001 ] G d , Y H 2 , ( 10 1 ¯ 0 ) M g 10.5 from ( 002 ) G d , Y H 2 , ( 1 2 ¯ 10 ) M g 10.5 from ( 020 ) G d , Y H 2 . The OR in Figure 3d can be expressed as [ 0001 ] M g // [ 001 ] G d , Y H 2 , ( 1 1 ¯ 00 ) M g 10.7 from ( 002 ) G d , Y H 2 , ( 11 ¯ 20 ) M g 10.7 from ( 020 ) G d , Y H 2 . All these three ORs are essentially equivalent within the experimental uncertainty. Because there are three equivalent { 10 1 ¯ 0 } M g planes and the diffraction spot of the { 002 } G d , Y H 2 plane can stay on either side of { 10 1 ¯ 0 } M g with the same misalignment angle, this OR has six equivalent variants, which explains the six traces in the BF image (Figure 3a) as well as the six pairs of {010} pole points in Figure 2e.
The occurrence of this unusual OR between hexagonal close-packed (HCP, the Mg matrix) and face-centered cubic (FCC, the plate-like G d , Y H 2 precipitates) can be well explained from a crystallographic matching perspective. In the HCP/FCC system, the atomic row mismatch along the close-packed directions [ 0001 ] M g and [ 001 ] G d , Y H 2 is only 0.85%. Thus, this is considered to be a pair of matching rows owing to the low interatomic mismatch along the row. As required by the E2EM model, this pair of matching rows must coincide or match at the interface to minimize the interface energy [29]. Figure 4a presents the simulated superimposed diffraction pattern of the Mg matrix and G d , Y H 2 along the [ 0001 ] M g // [ 001 ] G d , Y H 2 direction based on the E2EM model. In this case, the pair of matching planes that carry this matching pair of rows to the interface is ( 0 1 ¯ 10 ) M g and ( 002 ) G d , Y H 2 with an interplanar spacing misfit of 5.98%. To maintain atomic row alignment at the interface, these two matching planes must rotate by a certain degree relative to each other about the matching direction [30]. The calculated misalignment angle between ( 0 1 ¯ 10 ) M g and ( 002 ) G d , Y H 2 is 9.5° according to the second Δg parallelism rule, i.e., g 1 // g 2 , where g 1 = g 1 2 ¯ 10 g 020 and g 2 = g 10 1 ¯ 0 g 002 . It should be noted that there exists a ~1° discrepancy between the experimentally observed and simulated ORs in terms of the angle between the ( 0 1 ¯ 10 ) M g and ( 002 ) G d , Y H 2 planes. This may be attributed to the difference between the nominal lattice constants of G d , Y H 2 used for crystallographic matching calculation and the actual lattice constants, due to the complex structure of the RE hydrides [17,18,31,32]. In practice, the atomic ratio of (Gd, Y) and H in the RE hydrides can vary between 1.9 and 2.3 [17], potentially affecting the lattice constants and, consequently, the OR.
The energetically favorable interface orientation can also be calculated, defined by this pair of parallel Δg vectors, which can be expressed as ( 0.93   0.36   1.30 ¯   0.00 ) M g // ( 0.00   0.40   0.92 ¯ ) G d , Y H 2 . The green dashed line in Figure 4a represents the interface trace line perpendicular to the parallel g vectors, which is also consistent with the actual interface (see the trace of precipitate No. 1, or (c) in Figure 3a) between the RE hydrides and the Mg matrix during precipitation. Figure 4b shows a simulated atomic structure of this high-index interface observed along the [ 0001 ] M g // [ 001 ] G d , Y H 2 direction. The projections of atomic rows in the Mg and G d , Y H 2 lattices are represented by open circles and solid circles, respectively. The atomic species within a single lattice are not distinguished. For brevity, only the terraces and steps in the Mg lattice are illustrated by solid lines. The terrace plane is the ( 0 1 ¯ 10 ) M g plane on the Mg side and the ( 01 1 ¯ ) G d , Y H 2 plane on the G d , Y H 2 side. There is an angular discrepancy of 5.5° between these two planes, which allows the matching of terrace planes along the edges across the interface. The primary matching planes, [ 100 ] M g and [ 002 ] G d , Y H 2 , are highlighted by thick, gray solid lines, and they match each other in an “edge-to-edge” manner along the trace of the predicted interface. This stepped interface structure effectively eliminates all lattice mismatches between adjacent rows along the matching direction at the interface, thereby minimizing the structural component of the interfacial energy [33].

4. Conclusions

In summary, the microstructure evolution of a Mg-Gd-Y-Al alloy after hydrogenation treatment was investigated in detail in this work. The initial eutectic phases, including M g 24 ( G d , Y ) 5 a n d A l 2 ( G d , Y ) , present in the as-cast alloy were found to have completely disappeared after hydrogenation, confirming their full conversion through hydrogenation. Massive plate-like G d , Y H 2 precipitated from the matrix, while the formation of M g H 2 was almost suppressed. The plate-like G d , Y H 2 exhibited a specific OR with the Mg matrix with six equivalent variants, which can be written as [ 0001 ] M g // [ 001 ] G d , Y H 2 , ( 10 1 ¯ 0 ) M g ~ 10 from ( 002 ) G d , Y H 2 , ( 1 2 ¯ 10 ) M g ~ 10 from ( 020 ) G d , Y H 2 . Further crystallographic matching calculations suggest that such an OR is energetically favorable. These findings contribute to understanding the precipitation behavior of RE hydrides and improving the mechanical properties of the material by controlling the microstructure of the precipitated phases.

Author Contributions

Conceptualization, Y.L., D.Q. and X.Z.; methodology, K.C. and D.Q.; validation, K.C., S.C., D.Q. and Z.X.; formal analysis, D.Q.; investigation, K.C., Y.S., S.C. and Z.X.; resources, Y.L., D.Q. and X.Z.; data curation, K.C., Y.L, Y.S. and S.C.; writing—original draft preparation, K.C., Y.L., Y.S., S.C. and Z.X.; writing—review and editing, K.C., Y.L., S.C. and D.Q.; visualization, Y.L., S.C. and D.Q.; supervision, Y.L., D.Q. and X.Z.; project administration, Y.L. and X.Z.; funding acquisition, Y.L. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (No. 2022YFB3709300), the Space Utilization System of China Manned Space Engineering (No. KJZ-YY-WCL04, KJZ-YY-WCL0402), the Natural Science Foundation of Shanghai (No. 23ZR1431100), and the National Natural Science Foundation of China (No. 52305158).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, R.; Li, Y.H.; Zhang, N.; Li, Z.; Lin, X.; Zhu, W.; Lu, C.; Ding, W.J.; Zou, J.X. Nanostructuring of Mg-Based Hydrogen Storage Materials: Recent Advances for Promoting Key Applications. Nano-Micro Lett. 2023, 15, 36–62. [Google Scholar]
  2. Chen, Y.-S.; Huang, C.; Liu, P.-Y.; Yen, H.-W.; Niu, R.; Burr, P.; Moore, K.L.; Martínez-Pañeda, E.; Atrens, A.; Cairney, J.M. Hydrogen trapping and embrittlement in metals—A review. Int. J. Hydrog. Energy 2024, in press. [Google Scholar] [CrossRef]
  3. Behvar, A.; Haghshenas, M.; Djukic, M.B. Hydrogen embrittlement and hydrogen-induced crack initiation in additively manufactured metals: A critical review on mechanical and cyclic loading. Int. J. Hydrog. Energy 2024, 58, 1214–1239. [Google Scholar] [CrossRef]
  4. Lindblom, D.; Halilović, A.E.; Woracek, R.; Tengattini, A.; Helfen, L.; Dahlberg, C.F.O. In-situ neutron imaging of delayed crack propagation of high strength martensitic steel under hydrogen embrittlement conditions. Mater. Sci. Eng. A 2024, 895, 146215. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Ye, Q.Z.; Yan, Y. Processing, microstructure, mechanical properties, and hydrogen embrittlement of medium-Mn steels: A review. J. Mater. Sci. Technol. 2024, 201, 44–57. [Google Scholar] [CrossRef]
  6. Wang, Y.H.; Liu, J.Y.; Wang, Z.R.; Du, F.S. Effect of pre-strain on hydrogen embrittlement of 7075 aluminum alloy and molecular dynamics simulation. Int. J. Hydrog. Energy 2024, 88, 626–637. [Google Scholar] [CrossRef]
  7. Safyari, M.; Schnall, M.; Haunreiter, F.; Moshtaghi, M. Design of hydrogen embrittlement resistant 7xxx-T6 aluminum alloys based on wire arc additive manufacturing: Changing nanochemistry of strengthening precipitates. Mater. Des. 2024, 243, 113030. [Google Scholar] [CrossRef]
  8. Cheng, H.X.; Xu, J.J.; Luo, H.; Duan, G.P.; Zhao, Q.C.; Guo, Y.L. Enhanced hydrogen embrittlement resistance of additive manufacturing Ti-6Al-4V alloy with basket weave structure. Corros. Sci. 2024, 236, 112231. [Google Scholar] [CrossRef]
  9. Feng, X.D.; Xu, Y.W.; Shi, Y.; Gu, Y.F.; Korzhyk, V. Effects of microstructure and morphological distribution on hydrogen-embrittlement sensitivity of Ti–6Al–4V alloy welded joint. Int. J. Hydrog. Energy 2024, 50, 361–371. [Google Scholar] [CrossRef]
  10. Sun, B.Z.; Ren, J.X.; Wang, J.; Lv, X.L.; Zeng, Y.Z.; Chen, J.S. On Shapes and Orientations of Precipitation Strengthening Phases in an Aged Mg–La Alloy. Cryst. Growth Des. 2023, 23, 997–1013. [Google Scholar] [CrossRef]
  11. Ni, R.; Boehlert, C.J.; Zeng, Y.; Chen, B.; Huang, S.J.; Zheng, J.; Zhou, H.; Wang, Q.D.; Yin, D.D. Automated analysis framework of strain partitioning and deformation mechanisms via multimodal fusion and computer vision. Int. J. Plast. 2024, 182, 104119. [Google Scholar] [CrossRef]
  12. Kamilyan, M.; Silverstein, R.; Eliezer, D. Hydrogen trapping and hydrogen embrittlement of Mg alloys. J. Mater. Sci. 2017, 52, 11091–11100. [Google Scholar] [CrossRef]
  13. Wang, S.; Xu, D.K.; Wang, B.J.; Zhang, Z.Q.; Xu, X.B.; Wang, D.L.; Lv, X. Effect of I-phase formation on hydrogen embrittlement behaviour of as-cast Mg-8 wt%Li based alloys. Corros. Sci. 2024, 240, 112467. [Google Scholar] [CrossRef]
  14. He, B.; Li, Y.X.; Xiong, Z.H.; Chu, S.F.; Chen, K.; Li, M.W.; Zhou, L.P.; Zeng, X.Q. Effect of high-temperature hydrogenation on mechanical properties of a Mg-B4C composite. J. Alloys Compd. 2024, 975, 172949. [Google Scholar] [CrossRef]
  15. Chiu, C.; Su, C.-J.; Yu, W.-H.; Rabkin, E. Microstructure and mechanical properties of Mg–GdH2 composite prepared by internal hydrogenation. J. Mater. Sci. 2022, 57, 11649–11662. [Google Scholar] [CrossRef]
  16. Lapovok, R.; Zolotoyabko, E.; Berner, A.; Skripnyuk, V.; Lakin, E.; Larianovsky, N.; Xu, C.J.; Rabkin, E. Hydrogenation effect on microstructure and mechanical properties of Mg-Gd-Y-Zn-Zr alloys. Mater. Sci. Eng. A 2018, 719, 171–177. [Google Scholar] [CrossRef]
  17. Vajda, P. Hydrogen in rare-earth metals, including RH2+x phases. Handb. Phys. Chem. Rare Earths 1995, 20, 207–291. [Google Scholar]
  18. Vajda, P.; Andre, G. Commensurate and incommensurate magnetic structures in rare-earth. J. Alloys Compd. 2001, 326, 151–156. [Google Scholar] [CrossRef]
  19. Zhang, M.-X.; Kelly, P.M. Crystallographic features of phase transformations in solids. Prog. Mater. Sci. 2009, 54, 1101–1170. [Google Scholar] [CrossRef]
  20. Huang, Y.D.; Yang, L.; You, S.H.; Gan, W.M.; Kainer, K.U.; Hort, N. Unexpected formation of hydrides in heavy rare earth containing magnesium alloys. J. Magnesium Alloys 2016, 4, 173–180. [Google Scholar] [CrossRef]
  21. Peng, Q.M.; Huang, Y.D.; Meng, J.; Li, Y.D.; Kainer, K.U. Strain induced GdH2 precipitate in Mg–Gd based alloys. Intermetallics 2011, 19, 382–389. [Google Scholar] [CrossRef]
  22. Wu, L.Y.; Li, Y.a.; Cheng, Y.L.; Linghu, F.; Jiang, F.L.; Chen, G.; Teng, J.; Fu, D.F.; Zhang, H. Microstructure evolution and corrosion resistance improvement of Mg–Gd–Y–Zn–Zr alloys via surface hydrogen treatment. Corros. Sci. 2021, 191, 109746. [Google Scholar] [CrossRef]
  23. Wei, Q.H.; Yuan, L.; Ma, X.; Zheng, M.Y.; Shan, D.B.; Guo, B. Strengthening of low-cost rare earth magnesium alloy Mg-7Gd-2Y–1Zn-0.5Zr through multi-directional forging. Mater. Sci. Eng. A 2022, 831, 142144. [Google Scholar] [CrossRef]
  24. Xu, W.L.; Su, C.; Chen, X.H.; Tan, J.; Feng, L.; Wen, C.; Bai, J.Y.; Pan, F.S. Achieving superior elevated-temperature strength of Mg-12Gd-3Y alloys by Nd addition. Mater. Sci. Eng. A 2023, 867, 144730. [Google Scholar] [CrossRef]
  25. Zhu, Q.C.; Li, Y.X.; Cao, F.Y.; Qiu, D.; Yang, Y.; Wang, J.Y.; Zhang, H.; Ying, T.; Ding, W.J.; Zeng, X.Q. Towards development of a high-strength stainless Mg alloy with Al-assisted growth of passive film. Nat. Commun. 2022, 13, 5838. [Google Scholar] [CrossRef]
  26. Zhu, Q.C.; Shang, X.Q.; Zhang, H.; Qi, X.X.; Li, Y.X.; Zeng, X.Q. Influence of Al2Y particles on mechanical properties of Mg-11Y-1Al alloy with different grain sizes. Mater. Sci. Eng. A 2022, 831, 142166. [Google Scholar] [CrossRef]
  27. Wang, Z.P.; Shen, Z.; Liu, Y.; Zhao, Y.H.; Zhu, Q.C.; Chen, Y.W.; Wang, J.Y.; Li, Y.X.; Lozano-Perez, S.; Zeng, X.Q. The effect of LPSO phase on the high-temperature oxidation of a stainless Mg-Y-Al alloy. J. Magnesium Alloys 2024, 12, 4045–4052. [Google Scholar] [CrossRef]
  28. Easton, M.A.; Gibson, M.A.; Qiu, D.; Zhu, S.M.; Gröbner, J.; Schmid-Fetzer, R.; Nie, J.F.; Zhang, M.X. The role of crystallography and thermodynamics on phase selection in binary magnesium–rare earth (Ce or Nd) alloys. Acta Mater. 2012, 60, 4420–4430. [Google Scholar] [CrossRef]
  29. Kelly, P.M.; Zhang, M.X. Edge-to-Edge Matching—The Fundamentals. Metall. Mater. Trans. A 2006, 37, 833–839. [Google Scholar] [CrossRef]
  30. Zhang, M.X.; Kelly, P.M.; Easton, M.A.; Taylor, J.A. Crystallographic study of grain refinement in aluminum alloys using the edge-to-edge matching model. Acta Mater. 2005, 53, 1427–1438. [Google Scholar] [CrossRef]
  31. Fu, K.; Li, G.L.; Li, J.G.; Liu, Y.; Tian, W.H.; Zheng, J.; Li, X.G. Study on the thermodynamics of the gadolinium-hydrogen binary system (H/Gd = 0.0–2.0) and implications to metallic gadolinium purification. J. Alloys Compd. 2016, 673, 131–137. [Google Scholar] [CrossRef]
  32. Zhang, W.Z.; Weatherly, G.C. On the crystallography of precipitation. Prog. Mater. Sci. 2005, 50, 181–292. [Google Scholar] [CrossRef]
  33. Hwangsun, K.; Howook, C.; Juhyun, O.; Sangmin, L.; Ho, K.; Eun, S.P.; Lee, S.w.; Lee, G.D.; Miyoung, K.; Heung, N.H. Elucidating the role of a unique step-like interfacialstructure of η4 precipitates in Al-Zn-Mg alloy. Sci. Adv. 2023, 9, 7426. [Google Scholar]
Figure 1. The XRD results of the alloy before and after hydrogenation.
Figure 1. The XRD results of the alloy before and after hydrogenation.
Crystals 15 00249 g001
Figure 2. EBSD analysis of samples before and after hydrogenation. (a) Phase map before hydrogenation. (b,c) Phase maps after hydrogenation. (d) IPF of the region highlighted by the white box in (b). (e) PF of the area within the black box in (d).
Figure 2. EBSD analysis of samples before and after hydrogenation. (a) Phase map before hydrogenation. (b,c) Phase maps after hydrogenation. (d) IPF of the region highlighted by the white box in (b). (e) PF of the area within the black box in (d).
Crystals 15 00249 g002
Figure 3. Intragranular plate-like hydride precipitates analyzed by TEM. (a) A mosaic of bright-field images showing the six variants of hydrides sharing the same OR. Number these six variants from NO. 1 to 6 (blue Arabic numerals). (bd) Superimposed SAED patterns of the Mg matrix and RE hydrides in the vicinity of three variants (No. 1, 2, and 4, respectively) as indicated in (a). The specific positions are shown in the red box in (a), with the corresponding red letters at the top corresponding to (bd). Note the Δg vectors are defined as the difference of g{ 10 1 ¯ 0 } and g(002), as shown in the inset of (bd).
Figure 3. Intragranular plate-like hydride precipitates analyzed by TEM. (a) A mosaic of bright-field images showing the six variants of hydrides sharing the same OR. Number these six variants from NO. 1 to 6 (blue Arabic numerals). (bd) Superimposed SAED patterns of the Mg matrix and RE hydrides in the vicinity of three variants (No. 1, 2, and 4, respectively) as indicated in (a). The specific positions are shown in the red box in (a), with the corresponding red letters at the top corresponding to (bd). Note the Δg vectors are defined as the difference of g{ 10 1 ¯ 0 } and g(002), as shown in the inset of (bd).
Crystals 15 00249 g003
Figure 4. (a) Simulated diffraction patterns of intragranular plate-like hydrides and their Mg matrix at the OR suggested by the second Δg parallelism rule. The green dashed line indicates the trace of the habit plane. The large gray circle denotes the Mg matrix, while the small black circles represent G d , Y H 2 . (b) Local interfacial structure showing the “edge-to-edge” matching manner between ( 10 1 ¯ 0 ) M g and ( 002 ) G d , Y H 2 across the interface at the predicted OR. The regularly distributed red circles are well-matched atoms along the stepped interface with the terrace plane of ( 0 1 ¯ 10 ) M g | ( 01 1 ¯ ) G d , Y H 2 .
Figure 4. (a) Simulated diffraction patterns of intragranular plate-like hydrides and their Mg matrix at the OR suggested by the second Δg parallelism rule. The green dashed line indicates the trace of the habit plane. The large gray circle denotes the Mg matrix, while the small black circles represent G d , Y H 2 . (b) Local interfacial structure showing the “edge-to-edge” matching manner between ( 10 1 ¯ 0 ) M g and ( 002 ) G d , Y H 2 across the interface at the predicted OR. The regularly distributed red circles are well-matched atoms along the stepped interface with the terrace plane of ( 0 1 ¯ 10 ) M g | ( 01 1 ¯ ) G d , Y H 2 .
Crystals 15 00249 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, K.; Li, Y.; Su, Y.; Chu, S.; Xiong, Z.; Qiu, D.; Zeng, X. Crystallography and Morphology of (Gd,Y)H2 Hydride in a Mg-Gd-Y-Al Alloy. Crystals 2025, 15, 249. https://doi.org/10.3390/cryst15030249

AMA Style

Chen K, Li Y, Su Y, Chu S, Xiong Z, Qiu D, Zeng X. Crystallography and Morphology of (Gd,Y)H2 Hydride in a Mg-Gd-Y-Al Alloy. Crystals. 2025; 15(3):249. https://doi.org/10.3390/cryst15030249

Chicago/Turabian Style

Chen, Kun, Yangxin Li, Yang Su, Shufen Chu, Zhihao Xiong, Dong Qiu, and Xiaoqin Zeng. 2025. "Crystallography and Morphology of (Gd,Y)H2 Hydride in a Mg-Gd-Y-Al Alloy" Crystals 15, no. 3: 249. https://doi.org/10.3390/cryst15030249

APA Style

Chen, K., Li, Y., Su, Y., Chu, S., Xiong, Z., Qiu, D., & Zeng, X. (2025). Crystallography and Morphology of (Gd,Y)H2 Hydride in a Mg-Gd-Y-Al Alloy. Crystals, 15(3), 249. https://doi.org/10.3390/cryst15030249

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