Study of the hydrogen-induced amorphization in the LaNi2.28 alloy

Ping Lia, Fuqiang Zhaib, Qi Wana, Kuifei Zhaoa, Ziliang Lia, Alex A. Volinskyc and Xuanhui Qu*a
aInstitute for Advanced Materials and Technology, University of Science and Technology Beijing, Beijing 100083, China. E-mail: quxh@ustb.edu.cn; Fax: +86-10-62334311; Tel: 86-10-8277286
bDepartament Física Aplicada, EETAC, Universitat Politècnica de Catalunya-BarcelonaTech, 08860 Castelldefels, Spain
cDepartment of Mechanical Engineering, University of South Florida, Tampa, FL 33620, USA

Received 16th March 2014 , Accepted 28th May 2014

First published on 28th May 2014


Abstract

To understand the fundamental reason leading to the poor stability of LaNi2.28 hydrogen storage alloy, the cycling behavior of LaNi2.28 alloy under hydrogen has been investigated. The present study describes the hydrogen-induced amorphization (HIA) of LaNi2.28 alloy. The phase composition, microstructure and morphology of the phases are observed and analyzed using X-ray diffraction and scanning electron microscopy. The crystal structure and chemical composition of the amorphized alloy show that HIA is found in the sample at different cycles of the hydrogenation process. During the first hydrogenation cycle, the degree of HIA is indeclinable with the increased cycle. In addition, the formation processes of amorphous alloys, the occurrence of HIA and the relationship between the structure of the hydrogen-induced amorphous alloy and the stability of LaNi2.28 are given. The conclusions can be summarized as follows: 2LaNi2.28 + H2 → LaNi2.28Hx (amorphous) + LaNi5Hy + LaH2, and amorphous LaNi2.28Hx and LaNi5Hy can desorb 50% hydrogen of the absorption capacity of LaNi2.28, and the absorption–desorption capacity is stable after the first cycle.


1. Introduction

Because of the energy crisis and environmental issues, extensive efforts have been devoted to the development of high performance hydrogen storage alloys that can be used as the negative electrode of the nickel–metal hydride (Ni–MH) batteries in order to significantly reduce the adverse effects of consumption and dependence on fossil fuels.1–3 To date, the investigated hydrogen storage alloys for the negative electrode are mainly classified into the following categories: AB5 (CaCu5 type), AB3 (PuNi3 or CeNi3 type), A2B7 (Ce2Ni7 type), AB2 (MgCu2 or MgZn2 Laves phase), AB (CsCl type), Mg-based alloys and Zr-based Laves phase alloy.4–7 Among them, rare earth-based AB5-type and Zr-based Laves phase hydrogen storage alloys have been widely used in the commercial Ni–MH rechargeable battery due to their long-term cycling stability and environmental friendliness.8–10 However, with the rapid development of electronic devices requiring a Ni–MH rechargeable battery with higher overall performance, the limited discharge capacity of rare earth-based AB5-type (300 mA h g) and the difficult activation of Zr-based Laves phase alloy electrodes became the main obstacles to the extension of their applications in the field of Ni–MH rechargeable batteries.11,12

It has come to light that the deterioration of the hydrogen storage alloy capacity used as the negative electrode is mainly associated with pulverization, oxidation or corrosion of the alloy electrode during electrochemical charge–discharge cycling.13–15 Furthermore, it should be noted that the hydrogen-induced amorphization (HIA) also plays a crucial role in destroying the hydrogen storage capacity and cycle stability of AB3-type negative electrodes.16–19 Zhang et al.19 found that the AB3-type structure consists of one-third AB5 and two-third AB2 subunits, forming an alternating stacking structure. In the LaNi3 crystal structure, its LaNi2 subunits absorb hydrogen, resulting in the formation of an amorphous phase of LaNi2Hx, while its LaNi5 subunits transfer to the crystalline phase of LaNi5Hx, which leads to crack down of the LaNi3 crystal structure, and decreases the cycle stability of the corresponding alloy. The detailed reaction process may be described as: 3LaNi3 → LaNi2Hx (amorphous) + LaNi5Hy → 3/5LaH2 + 2/5LaNi5Hz. Similar to the structure of AB3-type alloy, A2B7-type alloy consists of both half AB5 and AB2 subunits, while A5B19-type alloy consists of three-fifth AB5 and two-fifth AB2 subunits. It was determined that the structural stability of the AB2 subunits determines the hydrogen storage properties of AB3-, A2B7- and A5B19-type alloys. With increase in the atomic proportion of the component element B among AB3, A2B7 and A5B19-type alloy, the content of AB2 subunits would decline resulting in improving in the cycle stability, but decrease in the theoretical hydrogen storage and electrochemical capacity. To date, HIA has been discovered in a large numbers of intermetallics, including A3B, A2B, AB, AB2, AB3, and A2B7-type alloy.16–23 LaNi2.28 (La7Ni16) alloy contains less Ni as compared with LaNi3 alloy, which indicates that LaNi2.28 alloy could be a promising candidate as the negative electrode of Ni–MH rechargeable batteries due to its high theoretical hydrogen storage capacity. Therefore, it is necessary to investigate the mechanism generating HIA in LaNi2.28 alloy during the absorption–desorption cycle process and the effects of HIA on the hydrogen storage capacity of LaNi2.28 alloy. Unfortunately, to date, the systematic investigation of HIA in LaNi2.28 alloy has not been conducted.24–27

In order to obtain a better understanding of the underlying mechanism of why and how to avoid or lessen the HIA in LaNi2.28 alloy to improve the cycling stability of hydrogen storage alloys as the negative electrode material, in this work the HIA process in LaNi2.28 alloy evolving during cycling under hydrogen, as well as the relationship between HIA and the microstructure of LaNi2.28 alloy, were investigated by X-ray diffraction (XRD), scanning electron microscopy (SEM) coupled with energy-dispersive spectroscopy (EDS), pressure-composition-temperature (PCT) measurement and transmission electron microscopy (TEM).

2. Experimental

Alloy ingots with nominal composition LaNi2.28 were prepared by arc-melting mixtures of pure La (99.9 mass%) and Ni (99.9 mass%) in high purity argon atmosphere. To ensure the homogeneity of the prepared alloy, the alloy ingots were turned over and re-melted at least three times. According to the phase diagram of La–Ni binary alloy, the annealing treatment was set at 700 °C in a bid to obtain the maximum quantity of LaNi2.28 phases. Consequently, the alloy samples sealed inside the quartz tubes were annealed at 700 °C for 24 h in a high-purity argon atmosphere. After completing the holding time at 700 °C, the annealed quartz tubes were immediately quenched in ice water to achieve relatively larger cooling speed resulting in more the tetragonal LaNi2.28 phases. To achieve fine powders from the annealed alloy samples, the annealed alloys were used to conduct the mechanical ball milling using a high-energy Spex mill (QM-3B) at the rate of 1200 rpm for 30 min. About 1.5 g of the annealed alloy sample was loaded into the stainless milling vial with a ball to powder weight ratio of 20[thin space (1/6-em)]:[thin space (1/6-em)]1 in a high-purity argon-filled glovebox (H2O: <10 ppm; O2: <10 ppm). Then, the alloy samples were mechanically ground into the powders with an average 35 μm diameter. Herein, it should be noted that the steel vial should rest for 5 min after every milling 10 min during the milling process in order to prevent increase in the temperature inside the steel vial.

X-ray diffraction (XRD) measurements were performed to characterize the amorphous nature and phase composition of the annealed LaNi2.28 alloys using a Rigaku D/Max diffractometer with Cu Kα radiation (λ = 1.5406 × 10−10 m). The morphologies of as annealed samples were observed using a Cambridge S360 model scanning electron microscope (SEM) coupled with energy-dispersive spectrometer (EDS). The hydriding/dehydriding properties of the LaNi2.28 alloy samples were measured using a pressure-composition-temperature (PCT) apparatus (Beijing Nonferrous Metal Research Institute, China). The details of the apparatus are given in previous reports.28,29 Typically, 0.5 g powder sample was loaded into the vessel, then the measuring system was pumped to a hard vacuum, and then the sample was heated up to 120 °C at a 6 °C min−1 heating rate under a pressure of 0.1 atm. After the completion of the first absorption process, the sample was subjected to re-hydrogenation at 120 °C under 3.5 MPa hydrogen pressure. Subsequently, the rehydrogenated sample was dehydrogenated at 120 °C under 0.1 MPa pressure to complete one cycle. The variation of phase composition with cycling process was examined by XRD and transmission electron microscopy (TEM).

3. Results and discussion

3.1 Structural characteristics

Fig. 1 shows the XRD pattern of LaNi2.28 alloy annealed at 700 °C for 24 h. As can be seen in Fig. 1, the annealed sample mainly consists of LaNi2.28 phase, but still contains traces LaNi3 second phase. In order to further prove the phase composition of the annealed LaNi2.28 alloy, Fig. 2a exhibits the back-scattered SEM image of the annealed LaNi2.28 alloys annealed at 700 °C for 24 h. It is obvious that the overall image is divided into two regions corresponding to grey and black region. From EDS measurements for these two regions, it can be concluded that the grey region corresponds to the LaNi2.28 phase and the black region should be LaNi3, as shown in Fig. 2b and c. By analyzing the phase diagram of La–Ni alloy,30 it seems that the LaNi2 phase could be obtained by the peritectic reaction under equilibrium solidification, while in practice it is difficult to obtain the single LaNi2 phase. Moreover, Klimyenko et al.31 found that La and Ni atoms are easier to be formed into a LaNi2.28 phase when the atomic ratio between La and Ni is 1[thin space (1/6-em)]:[thin space (1/6-em)]2, which is mainly because the atomic radius ratio between La and Ni (1.506) is relatively higher than the ideal atomic radius ratio of the Laves phase (1.225). Therefore, it is easier to form a LaNi2.28 phase when the atomic ratio between La and Ni is 1[thin space (1/6-em)]:[thin space (1/6-em)]2, which explains why more LaNi2.28 phases appear in Fig. 1.
image file: c4ra02302a-f1.tif
Fig. 1 XRD pattern of LaNi2.28 alloy annealed at 700 °C for 24 h.

image file: c4ra02302a-f2.tif
Fig. 2 (a) Back-scattered SEM image of the annealed LaNi2.28 alloys annealed at 700 °C for 24 h; energy dispersive spectroscopy (EDS) results of the (b) grey region A and (c) black region B.

3.2 Hydrogen absorption–desorption properties

A previous report on LaNi2 found that the LaNi2 alloy could achieve the equilibrium compound, corresponding to LaNi2–H2 when the LaNi2 alloy occurs the absorption–desorption cycle.32 However, so far there is no detailed theory that has been established between absorption–desorption cycle and LaNi2.28 alloy. Herein, the hydrogenation properties of LaNi2.28 alloy were investigated to understand the effects of HIA on the hydrogenation properties of LaNi2.28 alloy. Fig. 3a shows the hydrogen absorption curves of the as annealed LaNi2.28 alloy at 120 °C under a hydrogen pressure of 3.5 MPa without any activation process. As shown in Fig. 3, the as annealed LaNi2.28 alloy can absorb 1.6 wt% hydrogen within 0.5 h in the above-mentioned absorption condition. Compared with the traditional LaNi3 alloy,16–18 the LaNi2.28 alloy exhibits higher hydrogen absorption behavior at relatively low temperatures (≤150 °C). However, the cyclic stability of LaNi2.28 alloy is very poor as compared with other hydrogen storage alloys. It is necessary to explore the reason for the poor cyclic stability. The desorption kinetics of LaNi2.28 was measured at 120 °C at an initial hydrogen pressure of 3.5 MPa without any activation process. As can be seen in Fig. 3b, the LaNi2.28 alloy shows very sluggish hydrogen desorption kinetics. The LaNi2.28 only desorbed about 1.0 wt% of hydrogen at 120 °C within 1 h, indicating that the desorption kinetics of LaNi2.28 are worse than its absorption kinetics.
image file: c4ra02302a-f3.tif
Fig. 3 First hydrogen absorption–desorption curves of LaNi2.28 at 120 °C under a hydrogen pressure of 3.5 MPa: (a) hydrogen absorption and (b) hydrogen desorption.

Fig. 4 shows the isothermal absorption–desorption curves of LaNi2.28 alloy from the first to fourth cycle at 120 °C. Before measuring the second cycle, the sample was dehydrogenated at 120 °C for 0.5 h. It should be noted that 1.6 wt% hydrogen remains inside the alloy sample after the first absorption process. The sloping plateau appears and the hydrogen absorption clearly decreases in the second absorption–desorption process, suggesting that the isothermal absorption–desorption curves of the second cycle is remarkably different from the first one. There is a distinct decrease in hydrogen absorption during the absorption–desorption cycle of LaNi2.28 alloy, indicating the formation of amorphous hydrides during the cycling process.


image file: c4ra02302a-f4.tif
Fig. 4 The hydrogen capacity of pre-annealed LaNi2.28 alloy during four hydriding/dehydriding cycles at 120 °C.

3.3 Structural characteristics after hydriding/dehydriding

In order to understand the change of the structure and morphology of LaNi2.28 alloy after absorption–desorption cycles, the as-annealed alloys after hydriding/dehydriding were observed by SEM, and the results are shown in Fig. 5. By comparing SEM images of Fig. 5a–d, it can be observed that the change of granule sizes of the alloy samples is not obvious. This result shows that the pulverization basically occur in the process of hydriding/dehydriding, indicating that the HIA of the LaNi2.28 alloy is the main reason, leading to be more difficult on the alloy hydrogenation. The formation of amorphous hydrides inside the LaNi2.28 alloy impedes the diffusion of hydrogen atoms and decreases the dynamic property in the process of hydrogen absorption and desorption.33 Therefore, it can be concluded that the most important approach to enhance the lifecycle of the La–Ni system (LaNi2.28-type) hydrogen storage electrode alloys is to improve their anti-amorphization.
image file: c4ra02302a-f5.tif
Fig. 5 SEM images of as-cast and hydriding/dehydriding of the microstructure: (a) as-cast; (b) 1 hydriding/dehydriding cycle; (c) four hydrogen hydriding/dehydriding cycles and (d) ten hydriding/dehydriding cycles.

Fig. 6a shows the XRD pattern of as-annealed alloy after the first absorption. Compared with Fig. 1, the LaNi2.28 and LaNi3 phases disappear, and the main phases are LaNi5Hy and LaH2. Moreover, a small amount of LaNi5 can also be observed, and the diffraction peaks are broadened accordingly. Combined with the TEM analysis (Fig. 7), it can be observed that a few amorphous phases are formed. Moreover, the SAED diffraction rings can be indexed with crystal planes of (111) for LaNi5Hy, (111) for LaNi5 and (101) for LaH2, which is in good agreement with XRD analysis. Fig. 6b–d show the XRD patterns of the alloy after the first, fourth and tenth cycles, respectively. There are no diffraction peaks observed in Fig. 6b–d, indicating that the detected alloy exists in the form of amorphous state. Combined with the TEM analysis (Fig. 7b–d), it can be concluded that the major phases are amorphous states and that minor crystal phases, corresponding to LaNi5 and LaH2 can be seen, which can be verified by SAED analysis.


image file: c4ra02302a-f6.tif
Fig. 6 XRD patterns of the pre-annealed LaNi2.28 alloy cycled under 3.5 MPa H2 after (a) 1 hydrogen absorption cycle; (b) 1 hydrogen desorption cycle; (c) 4 hydrogen desorption cycles and (d) 10 hydrogen desorption cycles.

image file: c4ra02302a-f7.tif
Fig. 7 TEM images and selected area diffraction patterns of pre-annealed LaNi2.28 alloy after (a) 1 hydrogen absorption cycle; (b) 1 hydrogen desorption cycle; (c) 4 hydrogen desorption cycles and (d) 10 hydrogen desorption cycles. The circle and rectangle region in (a)–(d) show the crystalline and amorphous region, respectively; The insets in (a)–(d) show the selected area electron diffraction (SAED) patterns.

It can be seen from Fig. 6a and 7a that only partial amorphous phases can be found after absorption. A disproportionation reaction occurred between LaNi2.28 and LaNi3 after LaH2 and LaNi5 formation. The major phases exist in the form of amorphous states after the first absorption, while the minor phases are LaH2 and LaNi5 crystalline phases. To our best knowledge, the above mentioned results are not reported in the literature; moreover, this phenomenon is not consistent with the process of HIA, but it could be related to the process of sample preparation. The first absorption was conducted at 120 °C under 3.5 MPa hydrogen pressure, whereas the XRD and TEM measurements were performed after the sample was cooled in an electric furnace from 120 °C to room temperature; i.e., the sample was in amorphous state after the first absorption, but the crystallization process for the sample results from cooling in the electric furnace. For the samples after the first cycle, the XRD measurement was performed after the sample was cooled in air from 120 °C to room temperature. It can be seen from Fig. 4 and 5 that the absorption capacity is about 1.6 wt% for the first absorption, but the desorption capacity is only 0.8 wt% for the first desorption, which indicates that a disproportionation reaction occurs during the absorption process after LaH2 and LaNi5H6 formation, according to the reaction formula (1). As a support for above, the precipitation of LaH2 and LaNi5Hz from LaNi2Hx was also observed by Chung and Lee.34 According to Fig. 7, after the first, fourth and tenth cycles at 120 °C, the SADP of LaNi5 is pronounced. LaH2 is very stable and can not desorb hydrogen at 120 °C, resulting in a decreased desorption capacity. According to the reaction formula (1), if all La7Ni16 (LaNi2.28) participated in the disproportionation reaction, the desorption capacity of the formed LaNi5H6 would be 71.6% of the total absorption capacity; however, the actual desorption capacity is 50% of the total absorption capacity, which indicates that the disproportionation reaction is incomplete. After the absorption, amorphous LaNi2.28, LaNi5H6 and LaH2 are formed as shown in the reaction formula (2). The absorption and desorption capacities are 0.8 wt% from the second to fourth cycle, indicating that the amount of LaH2 is stable after the first cycle and the disproportionation reaction is completed during the first cycle.

 
image file: c4ra02302a-t1.tif(1)
 
image file: c4ra02302a-t2.tif(2)

4. Conclusions

In the present work, the hydrogen storage behavior of intermetallic LaNi2.28 is systematically studied after absorption–desorption cycling under hydrogen gas, during which they undergo a degree of amorphization. As a result we can be conclude the following.

The hydrogen-induced amorphization (HIA) is mainly responsible for the poor reversibility of the LaNi2.28 alloy, and it mainly occurs after the first cycle. With an increase in the number of cycles, the degree of HIA does not increase. For the LaNi2.28 alloy, the hydrogen absorption–desorption of its LaNi2.28 phase causes the local lattice to become a disordered one, leading to the amorphous phase formation. At the same time, the LaNi5Hx crystallites formed by the LaNi5 subunit start to precipitate. Combining the TEM measurements in Fig. 7 with the XRD results in Fig. 6 describes what occurred in the LaNi2.28 alloy: after the first absorption, amorphous LaNi2.28Hx formed, and it formed LaNi5Hy and LaH2 by the disproportionation reaction; amorphous LaNi2.28Hx and LaNi5Hy can desorb 50% hydrogen of the absorption capacity of LaNi2.28; and the absorption–desorption capacity is stable after the first cycle.

Acknowledgements

This work was supported by the National High-Tech R&D Program (863 Program) of China (2011AA03A408) for financial support, and the National Key Laboratory of New Metal (Z2011-11). Fuqiang Zhai thanks China Scholarship Council (CSC) for providing the scholarship.

References

  1. H. Reardon, J. M. Hanlon, R. W. Hughes, A. Godula-Jopek, T. K. Mandal and D. H. Gregory, Energy Environ. Sci., 2012, 5, 5951–5979 CAS.
  2. Y. B. Tan and X. B. Yu, RSC Adv., 2013, 3, 23879–23894 RSC.
  3. X. D. Kang, K. K. Wang, Y. J. Zhong, B. Yang and P. Wang, Phys. Chem. Chem. Phys., 2013, 15, 2153–2158 RSC.
  4. Z. Rogulski, J. Dłubak, M. Karwowska, M. Krebs, E. Pytlik, M. Schmalz, A. Gumkowska and A. Czerwiński, J. Power Sources, 2010, 195, 7517–7523 CrossRef CAS PubMed.
  5. X. Y. Zhao, L. Q. Ma, Y. Yao, Y. Ding and X. D. Shen, Energy Environ. Sci., 2010, 3, 1316–1321 CAS.
  6. Y. F. Liu, H. G. Pan, M. X. Gao and Q. D. Wang, J. Mater. Chem., 2011, 21, 4743–4755 RSC.
  7. K. Young, J. Nei, B. Huang and M. A. Fetcenko, Int. J. Hydrogen Energy, 2011, 36, 11146–11154 CrossRef CAS PubMed.
  8. F. Feng, M. Geng and D. O. Northwood, Int. J. Hydrogen Energy, 2001, 26, 725–734 CrossRef CAS.
  9. T. Kohno, H. Yoshida, F. Kawashima, T. Inaba, I. Sakai, M. Yamamoto and M. Kanda, J. Alloys Compd., 2000, 311, L5–7 CrossRef CAS.
  10. C. Khaldi, S. Boussami, R. B. Ben, H. Mathlouthi and J. Lamloumi, Mater. Sci. Eng., B, 2010, 175, 22–28 CrossRef CAS PubMed.
  11. S. K. Pandey, A. Srivastava and O. N. Srivastava, Int. J. Hydrogen Energy, 2007, 32, 2461–2465 CrossRef CAS PubMed.
  12. K. Young, T. Ouchi, B. Reichman, J. Koch and M. A. Fetcenko, J. Alloys Compd., 2011, 509, 7611–7617 CrossRef CAS PubMed.
  13. X. Y. Zhao, J. J. Li, Y. Yao, L. Q. Ma and X. D. Shen, RSC Adv., 2012, 2, 2149–2153 RSC.
  14. M. F. Wen, C. L. Song, L. Chen and Y. C. Zhai, J. Mater. Chem., 2002, 12, 3761–2545 RSC.
  15. X. Y. Zhao, L. Q. Ma and X. D. Shen, J. Mater. Chem., 2012, 22, 277–285 RSC.
  16. W. Wang, Y. G. Chen, M. D. Tao and C. L. Wu, J. Rare Earths, 2010, 28, 443–446 CrossRef CAS.
  17. J. Chen, H. T. Takeshita, H. Tanaka, N. Kuriyama, T. Sakai, I. Uehara and M. Haruta, J. Alloys Compd., 2000, 302, 304–313 CrossRef CAS.
  18. J. Zhang, F. Fang, S. Y. Zheng, J. Zhu, G. R. Chen, D. L. Sun, M. Latroche and A. Percheron-Guegan, J. Power Sources, 2007, 172, 446–450 CrossRef CAS PubMed.
  19. J. Zhang, G. Y. Zhou, G. R. Chen, M. Latroche, A. Percheron-Guegan and D. L. Sun, Acta Mater., 2008, 56, 5388–5394 CrossRef CAS PubMed.
  20. T. Kohno, H. Yoshida, F. Kawashima, T. Inaba, I. Sakai, M. Yamamoto and M. Kanda, J. Alloys Compd., 2000, 311, L5–L7 CrossRef CAS.
  21. C. Zhong, D. Chao, Y. Chen, W. Wang, D. Zhu and C. Wu, Electrochim. Acta, 2011, 58, 668–673 CrossRef CAS PubMed.
  22. K. Funaki, S. Orimo and H. Fujii, J. Alloys Compd., 1997, 253–254, 110–113 CrossRef CAS.
  23. D. Y. Zhang, J. K. Tang and K. A. Jr Gschneidner, J. Less-Common Met., 1991, 1, 45–53 Search PubMed.
  24. F. L. Zhang, Y. C. Luo, Y. C. Zhang, A. Q. Deng, L. Kang and J. H. Chen, J. Rare Earths, 2006, 24, 592–598 CAS.
  25. Z. L. Zhou, Y. Q. Song, S. Cui, C. G. Lin and X. H. Qu, Chin. J. of Nonferrous Met., 2007, 17, 45–52 CAS.
  26. L. Kang, X. G. Tian, Y. C. Luo and R. X. Yan, J. Lanzhou Univ. Technol., 2008, 34, 1–5 CAS.
  27. I. Park, N. Terashita and E. Abe, J. Alloys Compd., 2013, 580, S81–S84 CrossRef CAS PubMed.
  28. X. P. Zheng, P. Li, I. S. Humail, F. Q. An, G. Q. Wang and X. H. Qu, Int. J. Hydrogen Energy, 2007, 32, 1141–4960 CrossRef CAS PubMed.
  29. F. Q. Zhai, P. Li, A. Z. Sun, S. Wu, Q. Wan, W. N. Zhang, Y. L. Li, L. Q. Cui and X. H. Qu, J. Phys. Chem. C, 2012, 116, 11939–11945 CAS.
  30. H. J. Okamoto, J. Phase Equilib., 1991, 12, 615–616 CrossRef.
  31. A. V. Klimyenko, J. Seuntjens, L. L. Miller, B. J. Beaudry, R. A. Jacobson and K. A. Jr. Gschneidner, J. Less Common Met., 1988, 144, 133–141 CrossRef.
  32. W. K. Hu, R. V. Denys, C. C. Nwakwuo, T. Holm, J. P. Maehlena, J. K. Solberg and V. A. Yartys, Electrochim. Acta, 2013, 96, 27–33 CrossRef CAS PubMed.
  33. V. A. Yartys, O. Gutfleisch, V. V. Panasyuk and I. R. Harris, J. Alloys Compd., 1997, 253–254, 128–133 CrossRef CAS.
  34. U. I. Chung and J. Y. Lee, J. Non-Cryst. Solids, 1989, 110, 203–210 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2014