Next Article in Journal
Experimental Electron Density Distribution in Two Cocrystals of Betaines with p-Hydroxybenzoic Acid
Previous Article in Journal
Mechanochemical Synthesis and Crystal Structure of the Lidocaine-Phloroglucinol Hydrate 1:1:1 Complex
Previous Article in Special Issue
The Effects of Nanostructure on the Hydrogen Sorption Properties of Magnesium-Based Metallic Compounds: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Rehydrogenation of LiBH4 from H-Deficient Li2B12H12−x

1
Center for Materials Crystallography (CMC), the Department of Chemistry and the Interdisciplinary Nanoscience Center (iNANO), Aarhus University, Langelandsgade 140, 8000 Aarhus, Denmark
2
School of Materials Science and Engineering, Guangdong Provincial Key Laboratory of Advanced Energy Storage Materials, South China University of Technology, Guangzhou 510640, China
3
School of Materials and Energy, Guangdong University of Technology, Guangzhou 510006, China
4
Empa-Swiss Federal Laboratories for Materials Science and Technology, 8600 Dübendorf, Switzerland
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(3), 131; https://doi.org/10.3390/cryst8030131
Submission received: 22 January 2018 / Revised: 23 February 2018 / Accepted: 6 March 2018 / Published: 9 March 2018
(This article belongs to the Special Issue Properties and Applications of Novel Light Metal Hydrides)

Abstract

:
Li2B12H12 is commonly considered as a boron sink hindering the reversible hydrogen sorption of LiBH4. Recently, in the dehydrogenation process of LiBH4 an amorphous H-deficient Li2B12H12−x phase was observed. In the present study, we investigate the rehydrogenation properties of Li2B12H12−x to form LiBH4. With addition of nanostructured cobalt boride in a 1:1 mass ratio, the rehydrogenation properties of Li2B12H12−x are improved, where LiBH4 forms under milder conditions (e.g., 400 °C, 100 bar H2) with a yield of 68%. The active catalytic species in the reversible sorption reaction is suggested to be nonmetallic CoxB (x = 1) based on 11B MAS NMR experiments and its role has been discussed.

1. Introduction

Hydrogen is considered to be an ideal synthetic energy carrier to replace the limited quantity of fossil fuels available. Wide utilization of hydrogen as a fuel source for mobile applications requires the storage material to be safe, efficiently store hydrogen, and transportable. Owing to high gravimetric and volumetric densities of hydrogen, metal borohydrides have been intensively investigated for solid-state hydrogen storage over the last decade [1,2,3,4,5,6]. Among them, lithium borohydride (LiBH4), exhibiting a hydrogen density of 18.5 wt %, is one of the currently most discussed lightweight complex hydrides [7,8,9,10,11,12,13,14,15,16,17,18]. It crystalizes in two polymorphs, with structural transition from an orthorhombic low-temperature phase to a hexagonal high-temperature (HT) phase above 110 °C [7].
LiBH4 melts at Tm = 280 °C and releases considerable amounts of hydrogen from the liquid state. The decomposition pathway of LiBH4 depends on temperature and H2 pressure with Li2B12H12 formed as the main intermediate compound following a two-step route [10,13]:
LiBH4 → 5/6 LiH + 1/12 Li2B12H12 + 13/12 H2→ LiH + B + 3/2 H2
Experimentally, a H-deficient Li2B12H12−x phase has been identified in the solid residue of the thermal decomposition of LiBH4 [13,14]. Li2B12H12−x further decomposes into elemental boron above 650 °C [14]. Owing to the high thermal stability and low chemical reactivity, Li2B12H12 is generally considered as a boron sink in the hydrogen sorption process of LiBH4-based compounds hindering the efficient rehydrogenation reaction. Many efforts have been taken to circumvent the formation of Li2B12H12 in dehydrogenation process of LiBH4 and to improve the reversibility [18,19,20,21,22,23,24,25,26,27,28,29]. However, much less work has been done on the hydrogenation properties of Li2B12H12 itself, especially of the H-deficient Li2B12H12−x formed from the decomposition of LiBH4, which is of great importance for improving the hydrogen storage function of LiBH4.
The reformation of LiBH4 from its decomposition products was observed at 600 °C under a H2 pressure of 150 to 350 bar [9,11]. However, due to the amorphous state of the boron-containing compound in the decomposition product, the reaction pathway of the reformation of LiBH4 is not well documented. Recently, the reactivity of crystalline Li2B12H12 and LiH with a molar ratio of 1 to 10 has been examined, which convert to LiBH4 at 500 °C within 72 h. However, a high H2 pressure of 1000 bar is required to overcome the high kinetic barrier in the hydrogenation reaction [17].
In the present study, we systematically investigated the rehydrogenation properties of Li2B12H12−x to form LiBH4. First, pure LiBH4 was decomposed to Li2B12H12−x at 600 °C. The rehydrogenation of the Li2B12H12−x was carried out under the conditions of 350 bar H2, 500 to 600 °C and 24 h. Second, nanostructured cobalt boride was added to LiBH4 in a weight ratio of 1:1, enabling the decomposition of LiBH4 to Li2B12H12−x already at 350 °C. In presence of cobalt boride, the rehydrogenation of Li2B12H12−x is facilitated, where the reformation of LiBH4 is achieved under relatively mild conditions (e.g., 400 °C and 100 bar H2). We investigated the active catalytic species of cobalt boride and discuss the catalytic mechanism.

2. Experimental

The starting material, LiBH4 (purity, 95%) was purchased from Sigma-Aldrich Corp (St. Louis, MO, USA). Waxberry-like nanostructured cobalt boride was synthesized based on a wet-chemistry method described in literature [15,30,31]. The nanostructured cobalt boride shows a specific surface area of 39.7 m2/g and approximate average composition of Co1.34B [15]. The as-synthesized Co1.34B and LiBH4 were mechanically milled in a weight ratio of 1:1 using vibration milling (QM-3C, Nanjing Nanda Instrument Plant, Nanjing, China) for 1 h with a ball to powder ratio of 120:1 under Ar atmosphere. In the as-prepared sample 80.5 mol % of the boron originate from LiBH4, the remaining 19.5 mol % from Co1.34B. The H2 desorption of the as-prepared LiBH4-Co1.34B composite was performed using a custom-made pressure-composition-temperature apparatus under dynamic vacuum (lower than 10−4 mbar). The hydrogen amount was determined by the gas flow by means of a flow meter.
Solid state 11B magic angle spinning (MAS) NMR experiments were performed on a Bruker Avance-400 NMR spectrometer (Bruker BioSpin AG, Fällanden, Switzerland) using a 4 mm CP-MAS probe. The 11B MAS NMR spectra were recorded at 128.4 MHz at 12 kHz sample rotation applying a Hahn echo pulse sequence to suppress the broad background resonance of boron nitride in the probe. Pulse lengths of 1.5 µs (π/12 pulse) and 3.0 µs were applied for the excitation and echo pulses, respectively. For selected samples, 1H-11B-cross polarization magic angle spinning (CP-MAS) NMR experiments were performed using weak radio-frequency powers for spin locking of the 11B nucleus on resonance with mixing times of 50 µs. The setup was performed using LiBH4, for a sample of B(OH)3 only a very weak CP transfer efficiency was observed (<2% of signal intensity compared to a single pulse experiment).11B NMR chemical shifts are reported in parts per million (ppm) externally referenced to a 1 M B(OH)3 aqueous solution at 19.6 ppm as external standard sample. Quadrupolar parameters of B(III) sites and relative amounts of three- and four-fold coordinated boron atoms were determined by non-linear least-square fits of the regions of interest using the software DMFIT [32].

3. Results

To improve the reversibility the de- and rehydrogenation reaction of LiBH4, 50 wt % waxberry-like nanostructured CoxB (x = 1.34) was introduced into LiBH4 by ball milling. In the past, different CoxB:LiBH4 ratios were investigated. The 1:1 ratio showed the optimal hydrogen sorption performance [15]. The present study is to further investigate the hydrogen sorption mechanism of the CoxB:LiBH4 composite. Therefore, only the 1:1 ratio sample is investigated here. Figure 1a depicts the hydrogen desorption profile up to 500 °C. The LiBH4-Co1.34B composite shows two hydrogen desorption events at 200 and 375 °C, respectively, which are in agreement with previously reported results [15]. The major hydrogen desorption occurs around 375 °C. Figure 1b shows the isothermal dehydrogenation at 350 °C in the first two cycles. The rehydrogenation was carried out at 400 °C and 100 bar H2 for 24 h after the first dehydrogenation. The composite releases 5.1 wt % and 3.6 wt % hydrogen in the first two cycles, respectively, indicating a reversibility of 68%.
Figure 2a shows the full-range 11B MAS NMR spectra of the LiBH4-Co1.34B composite at different reaction states, compared to LiBH4. The center band in the 11B NMR spectrum in the as-prepared composite is observed at −41.9 ppm corresponding to the resonance of LiBH4. The signal maximum shifts to −10.3 ppm after dehydrogenation indicating the formation of intermediate products and returns to −41.9 ppm after rehydrogenation confirming that a large portion of LiBH4 has been reformed. Note that the as-prepared LiBH4-Co1.34B composite shows very strong sidebands in the 11B MAS NMR spectrum, while the side bands are much weaker in the spectra of the other samples shown in Figure 2a. We attribute the intense spinning side bands over a large chemical shift range in the as-prepared sample to the presence of ferromagnetism in the initial Co1.34B, as discussed in more detail below.
Figure 2b,c compare the central parts of the 11B MAS and the 1H-11B CP-MAS NMR spectra of the products after dehydrogenation at 350 °C and rehydrogenation at 400 °C. In Figure 2b the observed center band resonance can be deconvoluted into a main resonance at −10.3 ppm, and minor resonances at −41.0, 5.3 and 17.0 ppm. The resonances at −10.3 ppm and at −41.0 ppm are assignable to Li2B12H12−x and LiBH4, respectively. In the 1H-11B CP-MAS NMR spectra both signals still are present, whereas all other resonances belong to boron containing chemical species not attached to protons. The resonances at 5.3 and 17.0 ppm showing a typical second-order quadrupole pattern represent about 19 mol % of the boron atoms in the dehydrogenated state (Figure 2b), corresponding to the initial LiBH4:CoxB ratio. Therefore, these resonances are tentatively attributed to CoxB. On the other hand, the shape of these resonances resembles the one of the line shape of B(OH)3 [33]. However, the evaluated quadrupolar coupling constant and the chemical shift are slightly different and a strong contamination with oxygen is unlikely. In Figure 2c the main signal at −41 ppm is assigned to LiBH4 and a minor resonance at −15.5 ppm originates from stoichiometric Li2B12H12. An unambiguous quantification of CoxB by NMR is hampered by the presence of magnetic and chemical impurities.
Previous XRD studies indicate the formation of a new compound i.e., CoxB (x = 1) [15]. In the present experiment we see a strong decrease of the ferromagnetic signal by the reduced intensity of the spinning side bands in the 11B MAS NMR spectra upon the first hydrogen cycling. This is an indirect evidence of the formation of CoxB (x = 1). The magnetism of CoxB is known to decreases with increasing boron concentration, i.e., Co3B and Co2B are ferromagnetic while CoB is non-magnetic [34,35].

4. Discussion

The direct rehydrogenation of LiBH4 from its decomposition products only occurs at harsh conductions, e.g., 600 °C under 155 bar H2 [11]. We significantly improve the rehydrogenation properties of LiBH4 from its decomposition products Li2B12H12−x and LiH by addition of nanocrystalline cobalt boride (Co1.34B), the rehydrogenation of LiBH4 from its decomposition product Li2B12H12−x and LiH occurs already at much lower temperature (400 °C) and pressure (100 bar H2) with a yield of 68%. In addition, under catalysis of cobalt boride, the dehydrogenation reaction of LiBH4 to Li2B12H12−x is feasible at a lower temperature, i.e., 350 °C. The chemical state of cobalt changed during the hydrogen sorption process from the Co-rich, ferromagnetic Co1.34B to the non-magnetic CoxB (x = 1).
The catalytic effect of cobalt borides has been reported in other hydrogen-related reactions [36,37,38]. For instance, metallic Co and metallic-like Co2B were reported to decrease remarkably the dehydrogenation temperature in the LiBH4/LiNH2 system [36]. Nanocrystalline Co2B was also found to act as an efficient catalyst for hydrogen production from the hydrolysis of NaBH4 and in the field of electrochemical water splitting [36,38]. This is different from the result in the present study, where 11B MAS NMR results suggest that the active catalytic species must be nonmetallic and nonmagnetic with a composition close to CoB (1:1 in molar ratio). The catalytic effect of cobalt borides could be attributed to their non-compensated electronic structure, where electron transfers from B to a vacant d-orbital of metallic Co. making B electron-deficient and Co. electron-enriched [39,40]. Thereby cobalt borides may be able to promote the formation of B-H bonds of [BH4] during dehydrogenation and enable the break of B-H bonds in Li2B12H12−x in the rehydrogenation process. In the present case, the formation of CoxB (x = 1) from Co1.34B requires the addition of boron, which may originate from Li2B12H12−x, leading to a partial decomposition of the stable B12 units. In this scenario, cobalt boride would act as an additive rather than a catalyst. Further investigations are under progress to unveil how cobalt boride catalyzes the hydrogenation reaction of Li2B12H12−x and to improve the completeness of the reformation of LiBH4.

5. Conclusions

We demonstrate the improved rehydrogenation of H-deficient Li2B12H12−x to reform LiBH4. In presence of nanocrystalline cobalt boride, reformation of LiBH4 from Li2B12H12−x is achieved under relatively mild conditions (e.g., 400 °C, 100 bar H2) with a yield of 68%. The active species in the reversible sorption reaction step is suggested to be CoxB (x = 1) based on 11B MAS NMR results. It provides important insights on catalyzing LiBH4 as a potential reversible hydrogen storage material toward practical applications.

Acknowledgments

We are grateful to the Danish research council (HyNanoBorN). We also like to thank the Swiss National Science Foundation for financial support within the Sinergia project ‘Novel ionic conductors’ under contract number CRSII2_160749/1. The NMR hardware was partially granted by the Swiss National Science Foundation (SNFS, under contract number 206021_150638/1).

Author Contributions

Y.Y., M.Z. and A.R. conceived and designed the experiments; H.W. and W.C. prepared the CoxB samples, Y.Y. and D.R. performed the experiments and analyzed the data; the paper was written by Y.Y., A.R. and D.R. All authors agreed on the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yu, X.; Tang, Z.; Sun, D.; Ouyang, L.; Zhu, M. Recent advances and remaining challenges of nanostructured materials for hydrogen storage applications. Prog. Mater. Sci. 2017, 88, 1–48. [Google Scholar] [CrossRef]
  2. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Cerny, R.; Ravnsbaek, D.B.; Filinchuk, Y.; Dornheim, M.; Besenbacherf, F.; Jensen, T.R. Metal borohydrides and derivatives - synthesis, structure and properties. Chem. Soc. Rev. 2017, 46, 1565–1634. [Google Scholar] [CrossRef] [PubMed]
  3. Moller, K.T.; Jensen, T.R.; Akiba, E.; Li, H.W. Hydrogen—A sustainable energy carrier. Prog. Nat. Sci.-Mater. 2017, 27, 34–40. [Google Scholar] [CrossRef]
  4. He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen carriers. Nat. Rev. Mater. 2016, 1, 16059. [Google Scholar] [CrossRef]
  5. Lai, Q.; Paskevicius, M.; Sheppard, D.A.; Buckley, C.E.; Thornton, A.W.; Hill, M.R.; Gu, Q.; Mao, J.; Huang, Z.; Liu, H.K.; et al. Hydrogen storage materials for mobile and stationary applications: Current state of the art. Chemsuschem 2015, 8, 2789–2825. [Google Scholar] [CrossRef] [PubMed]
  6. Orimo, S.I.; Nakamori, Y.; Eliseo, J.R.; Zuttel, A.; Jensen, C.M. Complex hydrides for hydrogen storage. Chem. Rev. 2007, 107, 4111–4132. [Google Scholar] [CrossRef] [PubMed]
  7. Soulie, J.P.; Renaudin, G.; Cerny, R.; Yvon, K. Lithium boro-hydride LiBH4. J. Alloy Compd. 2002, 346, 200–205. [Google Scholar] [CrossRef]
  8. Zuttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage properties of LiBH4. J. Alloy Compd. 2003, 356, 515–520. [Google Scholar] [CrossRef]
  9. Orimo, S.; Nakamori, Y.; Kitahara, G.; Miwa, K.; Ohba, N.; Towata, S.; Züttel, A. Dehydriding and rehydriding reactions of LiBH4. J. Alloy Compd. 2005, 404, 427–430. [Google Scholar] [CrossRef]
  10. Orimo, S.I.; Nakamori, Y.; Ohba, N.; Miwa, K.; Aoki, M.; Towata, S.; Zuttel, A. Experimental studies on intermediate compound of LiBH4. Appl. Phys. Lett. 2006, 89, 021920. [Google Scholar] [CrossRef]
  11. Mauron, P.; Buchter, F.; Friedrichs, O.; Remhof, A.; Bielmann, M.; Zwicky, C.N.; Zuttel, A. Stability and reversibility of LiBH4. J. Phys. Chem. B 2008, 112, 906–910. [Google Scholar] [CrossRef] [PubMed]
  12. Her, J.H.; Yousufuddin, M.; Zhou, W.; Jalisatgi, S.S.; Kulleck, J.G.; Zan, J.A.; Hwang, S.J.; Bowman, R.C.; Udovic, T.J. Crystal structure of Li2B12H12: A possible intermediate species in the decomposition of LiBH4. Inorg. Chem. 2008, 47, 9757–9759. [Google Scholar] [CrossRef] [PubMed]
  13. Yan, Y.G.; Remhof, A.; Hwang, S.J.; Li, H.W.; Mauron, P.; Orimo, S.; Zuttel, A. Pressure and temperature dependence of the decomposition pathway of LiBH4. Phys. Chem. Chem. Phys. 2012, 14, 6514–6519. [Google Scholar] [CrossRef] [PubMed]
  14. Pitt, M.P.; Paskevicius, M.; Brown, D.H.; Sheppard, D.A.; Buckley, C.E. Thermal stability of Li2B12H12 and its role in the decomposition of LiBH4. J. Am. Chem. Soc. 2013, 135, 6930–6941. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, W.; Wang, H.; Liu, J.; Jiao, L.; Wang, Y.; Ouyang, L.; Sun, T.; Sun, D.; Wang, H.; Yao, X.; et al. Towards easy reversible dehydrogenation of LiBH4 by catalyzing hierarchic nanostructured cob. Nano Energy 2014, 10, 235–244. [Google Scholar] [CrossRef]
  16. Shao, J.; Xiao, X.Z.; Fan, X.L.; Huang, X.; Zhai, B.; Li, S.Q.; Ge, H.W.; Wang, Q.D.; Chen, L.X. Enhanced hydrogen storage capacity and reversibility of LiBH4 nanoconfined in the densified zeolite-templated carbon with high mechanical stability. Nano Energy 2015, 15, 244–255. [Google Scholar] [CrossRef]
  17. White, J.L.; Newhouse, R.J.; Zhang, J.Z.; Udovic, T.J.; Stavila, V. Understanding and mitigating the effects of stable dodecahydro-closo-dodecaborate intermediates on hydrogen-storage reactions. J. Phys. Chem. C 2016, 120, 25725–25731. [Google Scholar] [CrossRef]
  18. Ngene, P.; Verkuijlen, M.H.W.; Barre, C.; Kentgens, A.P.M.; de Jongh, P.E. Reversible li-insertion in nanoscaffolds: A promising strategy to alter the hydrogen sorption properties of li-based complex hydrides. Nano Energy 2016, 22, 169–178. [Google Scholar] [CrossRef]
  19. Vajo, J.J.; Skeith, S.L.; Mertens, F. Reversible storage of hydrogen in destabilized LiBH4. J. Phys. Chem. B 2005, 109, 3719–3722. [Google Scholar] [CrossRef] [PubMed]
  20. Pinkerton, F.E.; Meyer, M.S.; Meisner, G.P.; Balogh, M.P.; Vajo, J.J. Phase boundaries and reversibility of LiBH4/MgH2 hydrogen storage material. J. Phys. Chem. C 2007, 111, 12881–12885. [Google Scholar] [CrossRef]
  21. Au, M.; Jurgensen, A.R.; Spencer, W.A.; Anton, D.L.; Pinkerton, F.E.; Hwang, S.J.; Kim, C.; Bowman, R.C. Stability and reversibility of lithium borohydrides doped by metal halides and hydrides. J. Phys. Chem. C 2008, 112, 18661–18671. [Google Scholar] [CrossRef]
  22. Blanchard, D.; Shi, Q.; Boothroyd, C.B.; Vegge, T. Reversibility of al/ti modified LiBH4. J. Phys. Chem. C 2009, 113, 14059–14066. [Google Scholar] [CrossRef]
  23. Nielsen, T.K.; Bosenberg, U.; Gosalawit, R.; Dornheim, M.; Cerenius, Y.; Besenbacher, F.; Jensen, T.R. A reversible nanoconfined chemical reaction. ACS Nano 2010, 4, 3903–3908. [Google Scholar] [CrossRef] [PubMed]
  24. Shim, J.H.; Lim, J.H.; Rather, S.U.; Lee, Y.S.; Reed, D.; Kim, Y.; Book, D.; Cho, Y.W. Effect of hydrogen back pressure on dehydrogenation behavior of LiBH4-based reactive hydride composites. J. Phys. Chem. Lett. 2010, 1, 59–63. [Google Scholar] [CrossRef]
  25. Choi, Y.J.; Lu, J.; Sohn, H.Y.; Fang, Z.Z.; Kim, C.; Bowman, R.C.; Hwang, S.J. Reaction mechanisms in the Li3AlH6/LiBH4 and Al/LiBH4 systems for reversible hydrogen storage. Part 2: Solid-state NMR studies. J. Phys. Chem. C 2011, 115, 6048–6056. [Google Scholar] [CrossRef]
  26. Cai, W.T.; Wang, H.; Sun, D.L.; Zhu, M. Nanosize-controlled reversibility for a destabilizing reaction in the LiBH4-NdH2+x system. J. Phys. Chem. C 2013, 117, 9566–9572. [Google Scholar] [CrossRef]
  27. Javadian, P.; Sheppard, D.A.; Buckley, C.E.; Jensen, T.R. Hydrogen storage properties of nanoconfined LiBH4-Ca(BH4)2. Nano Energy 2015, 11, 96–103. [Google Scholar] [CrossRef]
  28. Puszkiel, J.A.; Riglos, M.V.C.; Karimi, F.; Santoru, A.; Pistidda, C.; Klassen, T.; von Colbe, J.M.B.; Dornheim, M. Changing the dehydrogenation pathway of LiBH4-MgH2 via nanosized lithiated TiO2. Phys. Chem. Chem. Phys. 2017, 19, 7455–7460. [Google Scholar] [CrossRef] [PubMed]
  29. Kang, X.-D.; Wang, P.; Ma, L.-P.; Cheng, H.-M. Reversible hydrogen storage in LiBH4 destabilized by milling with al. Appl. Phys. A 2007, 89, 963–966. [Google Scholar] [CrossRef]
  30. Song, D.; Wang, Y.; Wang, Y.; Jiao, L.; Yuan, H. Preparation and characterization of novel structure Co–B hydrogen storage alloy. Electrochem. Commun. 2008, 10, 1486–1489. [Google Scholar] [CrossRef]
  31. Wang, Q.; Jiao, L.; Du, H.; Song, D.; Peng, W.; Si, Y.; Wang, Y.; Yuan, H. Facile preparation and good electrochemical hydrogen storage properties of chain-like and rod-like Co-B nanomaterials. Electrochim. Acta 2010, 55, 7199–7203. [Google Scholar] [CrossRef]
  32. Massiot, D.; Fayon, F.; Capron, M.; King, I.; LeCalve, S.; Alonso, B.; Durand, J.O.; Bujoli, B.; Gan, Z.H.; Hoatson, G. Modelling one- and two-dimensional solid-state nmr spectra. Magn. Reson. Chem. 2002, 40, 70–76. [Google Scholar] [CrossRef]
  33. Yan, Y.G.; Remhof, A.; Mauron, P.; Rentsch, D.; Lodziana, Z.; Lee, Y.S.; Lee, H.S.; Cho, Y.W.; Zuttel, A. Controlling the dehydrogenation reaction toward reversibility of the LiBH4-Ca(BH4)2 eutectic system. J. Phys. Chem. C 2013, 117, 8878–8886. [Google Scholar] [CrossRef]
  34. Bykova, E.; Tsirlin, A.A.; Gou, H.; Dubrovinsky, L.; Dubrovinskaia, N. Novel non-magnetic hard boride Co5B16 synthesized under high pressure. J. Alloy Compd. 2014, 608, 69–72. [Google Scholar] [CrossRef]
  35. Chikazumi, S. Physics of Ferromagnetism, 2nd ed.; Oxford University Press: Oxford, UK, 1997. [Google Scholar]
  36. Wu, C.; Wu, F.; Bai, Y.; Yi, B.; Zhang, H. Cobalt boride catalysts for hydrogen generation from alkaline nabh4 solution. Mater. Lett. 2005, 59, 1748–1751. [Google Scholar] [CrossRef]
  37. Tang, W.S.; Wu, G.; Liu, T.; Wee, A.T.S.; Yong, C.K.; Xiong, Z.; Hor, A.T.S.; Chen, P. Cobalt-catalyzed hydrogen desorption from the LiNH2-LiBH4 system. Dalton Trans. 2008, 0, 2395–2399. [Google Scholar] [CrossRef] [PubMed]
  38. Masa, J.; Weide, P.; Peeters, D.; Sinev, I.; Xia, W.; Sun, Z.; Somsen, C.; Muhler, M.; Schuhmann, W. Amorphous cobalt boride (Co2B) as a highly efficient nonprecious catalyst for electrochemical water splitting: Oxygen and hydrogen evolution. Adv. Energy Mater. 2016, 6, 15023131. [Google Scholar] [CrossRef]
  39. Fernandes, R.; Patel, N.; Miotello, A.; Filippi, M. Studies on catalytic behavior of Co–Ni–B in hydrogen production by hydrolysis of NaBH4. J. Mol. Catal. A 2009, 298, 1–6. [Google Scholar] [CrossRef]
  40. Carenco, S.; Portehault, D.; Boissière, C.; Mézailles, N.; Sanchez, C. Nanoscaled metal borides and phosphides: Recent developments and perspectives. Chem. Rev. 2013, 113, 7981–8065. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Hydrogen desorption profile up to 500 °C; (b) isothermal hydrogen release of LiBH4-Co1.34B at 350 °C in the first two cycles.
Figure 1. (a) Hydrogen desorption profile up to 500 °C; (b) isothermal hydrogen release of LiBH4-Co1.34B at 350 °C in the first two cycles.
Crystals 08 00131 g001
Figure 2. (a) Full range of 11B MAS NMR spectra of LiBH4-Co1.34B composite at different reaction stages: as-prepared, dehydrogenated (DeH) at 350 °C, rehydrogenated (ReH) at 400 °C, and of pure LiBH4 as reference; (b,c) 11B MAS NMR and 1H-11B CP-MAS NMR spectra of the LiBH4-Co1.34B composite dehydrogenated at 350 °C and rehydrogenated at 400 °C, respectively. The experimental data (exp.) are shown as solid, individual components (ind.) and fitting results as different dotted (···) and broken (---) lines, respectively. The stars (*) indicate spinning side bands.
Figure 2. (a) Full range of 11B MAS NMR spectra of LiBH4-Co1.34B composite at different reaction stages: as-prepared, dehydrogenated (DeH) at 350 °C, rehydrogenated (ReH) at 400 °C, and of pure LiBH4 as reference; (b,c) 11B MAS NMR and 1H-11B CP-MAS NMR spectra of the LiBH4-Co1.34B composite dehydrogenated at 350 °C and rehydrogenated at 400 °C, respectively. The experimental data (exp.) are shown as solid, individual components (ind.) and fitting results as different dotted (···) and broken (---) lines, respectively. The stars (*) indicate spinning side bands.
Crystals 08 00131 g002

Share and Cite

MDPI and ACS Style

Yan, Y.; Wang, H.; Zhu, M.; Cai, W.; Rentsch, D.; Remhof, A. Direct Rehydrogenation of LiBH4 from H-Deficient Li2B12H12−x. Crystals 2018, 8, 131. https://doi.org/10.3390/cryst8030131

AMA Style

Yan Y, Wang H, Zhu M, Cai W, Rentsch D, Remhof A. Direct Rehydrogenation of LiBH4 from H-Deficient Li2B12H12−x. Crystals. 2018; 8(3):131. https://doi.org/10.3390/cryst8030131

Chicago/Turabian Style

Yan, Yigang, Hui Wang, Min Zhu, Weitong Cai, Daniel Rentsch, and Arndt Remhof. 2018. "Direct Rehydrogenation of LiBH4 from H-Deficient Li2B12H12−x" Crystals 8, no. 3: 131. https://doi.org/10.3390/cryst8030131

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