Next Article in Journal
Exergy Analysis of a Two-Pass Reverse Osmosis (RO) Desalination Unit with and without an Energy Recovery Turbine (ERT) and Pressure Exchanger (PX)
Next Article in Special Issue
Increasing Hydrogen Density with the Cation-Anion Pair BH4-NH4+ in Perovskite-Type NH4Ca(BH4)3
Previous Article in Journal
Coal and Coalbed Methane Co-Extraction Technology Based on the Ground Movement in the Yangquan Coalfield, China
Previous Article in Special Issue
Dehydriding Process and Hydrogen–Deuterium Exchange of LiBH4–Mg2FeD6 Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The improved Hydrogen Storage Performances of the Multi-Component Composite: 2Mg(NH2)2–3LiH–LiBH4

1
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
2
Graduate School of the Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Energies 2015, 8(7), 6898-6909; https://doi.org/10.3390/en8076898
Submission received: 25 May 2015 / Revised: 28 June 2015 / Accepted: 30 June 2015 / Published: 10 July 2015
(This article belongs to the Special Issue Hydrides: Fundamentals and Applications)

Abstract

:
2Mg(NH2)2–3LiH–LiBH4 composite exhibits an improved kinetic and thermodynamic properties in hydrogen storage in comparison with 2Mg(NH2)2–3LiH. The peak temperature of hydrogen desorption drops about 10 K and the peak width shrinks about 50 K compared with the neat 2Mg(NH2)2–3LiH. Its isothermal dehydrogenation and re-hydrogenation rates are respectively 2 times and 18 times as fast as those of 2Mg(NH2)2–3LiH. A slope desorption region with higher equilibrium pressure is observed. By means of X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR) and nuclear magnetic resonance (NMR) analyses, the existence of Li2BNH6 is identified and its roles in kinetic and thermodynamic enhancement are discussed.

1. Introduction

Owing to the worldwide demand for the renewable energy sources, hydrogen will be an ideal energy carrier if it can be stored safely, efficiently and conveniently [1]. Numerous solid-state hydrogen storage materials have been developed to store hydrogen in an energy or volume efficient way [2]. Complex hydrides, e.g., alanates [3,4,5], borohydrides [6,7,8], and amide-hydride systems [9,10,11,12,13,14,15,16,17,18,19,20] are promising to fulfill the on-board hydrogen storage requirements. In particular, a number of amide–hydride systems, such as Li–N–H [10], Li–Mg–N–H [11,13,17,18,19,20], Li–Ca–N–H [14], Li–Al–N–H [15], Mg–N–H [12], Ca–N–H [16], and so on [9,14,21,22] have been investigated since 2002. As for the Li–Mg–N–H system, Li2Mg(NH)2 can be obtained by releasing hydrogen from MgH2–2LiNH2 composite or Mg(NH2)2–2LiH composite, which has attracted considerable attention due to its favorable thermodynamics (ΔH ≈ 40 kJ/mol-H2), relatively high hydrogen capacity (5.6 wt%) and good reversibility.
Hydrogenation/dehydrogenation processes of Li2Mg(NH)2 were probed under the pressure composition isotherms conditions by ex-situ X-ray diffraction (XRD) [17] and in-situ neutron measurements [23]. During the hydrogenation process, Li2Mg(NH)2 is first converted to Li2Mg2(NH)3, LiNH2 and LiH at a low pressure slope, and then form Mg(NH2)2 and 2LiH at the high pressure plateau. In the dehydrogenation, Mg(NH2)2 and 2LiH take the reverse reaction route and go back to Li2Mg(NH)2.
The reactions can be described as follows:
2Mg(NH2)2 + 3LiH ⇌ Li2Mg2(NH)3 + LiNH2 + 2H2
Li2Mg2(NH)3 + LiNH2 + LiH ⇌ Li2Mg(NH)2 + 2H2
Considerable works have been carried out to lower down the operating temperature and improve the dehydrogenation kinetics of Mg(NH2)2–2LiH composite by doping additives, such as alkali metal compounds [24,25,26,27] and metal borohydrides [28,29]. The addition of KH significantly reduced the operating temperature from ca. 523 to 380 K, while the equilibrium pressure was ca. 0.2 MPa. Rb based compounds were found to be more effective for improving the kinetic properties of Mg(NH2)2–2LiH composite than the potassium based additives, and the active species was identified to be RbH [24,30].Yang et al. [31] reported that a multi-component composite of LiNH2, MgH2 and LiBH4 with a molar ratio of 2:1:1 exhibited enhanced low-temperature desorption kinetics and a significant reduction in ammonia liberations. As a kind of self-catalyzing hydrogen storage material, LiBH4 catalyzed the dehydrogenation of 2LiNH2–MgH2 composite to form Li2Mg(NH)2, furthermore Li2Mg(NH)2 could react with LiBH4 to release more hydrogen. Following this work, Hu et al. [29] doped 10 mol% LiBH4 in Mg(NH2)2–2LiH composite and found the doping of LiBH4 not only improved the kinetics but also reduced the heat of dehydrogenation from 40 kJ/mol-H2 of Mg(NH2)2–2LiH composite to 36.5 kJ/mol-H2 of Mg(NH2)2–2LiH–0.1LiBH4 composite. Other borohydrides, such as Mg(BH4)2 and Ca(BH4)2[32,33,34], have similar effects as LiBH4 in Mg(NH2)2–2LiH composite because they convert to LiBH4 after a metathesis reaction. Li et al. [35] introduced 5 mol% LiBr to 2LiNH2–MgH2 composite and found certain effects on its thermodynamics and kinetics. In Cao’s recentwork [36], thermodynamic properties of the dehydrogenation of 2Mg(NH2)2–3LiH composite were improved by stabilizing the dehydrogenated product, LiNH2, in Reaction (1). LiNH2 reacts with LiBH4, LiI and LiBr exothermically and forms more stable compounds, i.e., Li4(NH2)3BH4, Li3(NH2)2I, and Li2(NH2)Br; the reactions are as follows:
2Mg(NH2)2 + 3LiH + 1/2LiI ⇌ Li2Mg2(NH)3 + 1/2Li3(NH2)2I + 3H2
2Mg(NH2)2 + 3LiH + LiBr ⇌ Li2Mg2(NH)3 + Li2NH2Br + 3H2
2Mg(NH2)2 + 3LiH + 1/3LiBH4⇌ Li2Mg2(NH)3 + 1/3Li4BN3H10 + 3H2
The additions of LiBH4, LiI and LiBr noticeably reduced the heat of dehydrogenation from 40 kJ/mol-H2 of 2Mg(NH2)2–3LiH composite to 35.8, 33.3 and 31.9 kJ/mol-H2, respectively, which suggests that hydrogen release at 0.1 MPa equilibrium pressure can be thermodynamically allowed at 337, 333 and 320 K, respectively. Quaternary complex hydrides, such as Li2BNH6 [37] and Li4BN3H10 [38], were found by ball milling or heating the mixtures of LiNH2 and LiBH4 with corresponding molar ratios. If with more LiBH4 to facilitate the formation of Li2BNH6, Reaction (5) can be rewritten as Reaction (6):
2Mg(NH2)2 + 3LiH + LiBH4 ⇌ Li2Mg2(NH)3 + Li2BNH6 + 3H2
Herein we report the modification of 2Mg(NH2)2–3LiH system by adding a different ratio of LiBH4. The dehydrogenation and re-hydrogenation performances of Mg(NH2)2–2LiH–1LiBH4 composite were investigated by Thermogravimetry coupled differential thermal analysis (TG-DTA) measurement, temperature programmed desorption-mass spectroscopy (TPD-MS), isothermal volumetric release, and soak. We found that the increase in the amount of LiBH4 improved the kinetics of 2Mg(NH2)2–3LiH composite.
An interesting phenomenon was observed that there was a higher pressure slope in pressure composition isotherm (PCI) measurements of 2Mg(NH2)2–3LiH–LiBH4 composite than 2Mg(NH2)2–3LiH–1/3LiBH4 composite [36], which reveals the change of thermodynamic properties. Samples at this stage were analyzed by means of X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and magic angle spinning nuclear magnetic resonance (MAS NMR).

2. Results and Discussion

2.1. The hydrogen Desorption Properties of 2Mg(NH2)2–3LiH–LiBH4 Composite

TG curve of 2Mg(NH2)2–3LiH–LiBH4 composite (T2) is shown in Figure 1a. The thermal desorption of this sample exhibits two main steps, one is from 400 K to 450 K and the other is from 475 K to 575 K. The weight losses of the first step and the second step are 3.63 wt% and 6.24 wt%, respectively. Moreover, the first-step weight loss is near the theoretical loss of H2 (3.7 wt%) deduced from Reaction (6).
Figure 1. TG-DTA curves of 2Mg(NH2)2–3LiH–LiBH4 composite (T2) (a) TG and (b) DTA.
Figure 1. TG-DTA curves of 2Mg(NH2)2–3LiH–LiBH4 composite (T2) (a) TG and (b) DTA.
Energies 08 06898 g001
The DTA curve of T2 gives two main endothermic signals corresponding to the two desorption steps in TG curve. The small endothermic peak in Figure 1 at 363 K can be assigned to the phase transformation of LiBH4 in starting materials.
The kinetic desorption behaviors of 2Mg(NH2)2–3LiH (Tp) and 2Mg(NH2)2–3LiH–LiBH4 (T2) composites were investigated by TPD-MS too. H2 and NH3 signals are shown in Figure 2a,b, respectively. As shown in Figure 2b, the byproduct ammonia can be effectively inhibited in T2 sample during desorption process, and the two-step dehydrogenation of T2 sample is corresponding to the result of TG. In the first dehydrogenation step of T2 sample, 3.63 wt% hydrogen can be released according to the thermogravimetry analysis, which is near the weight loss of 3.7 wt% deduced from Reaction (6). The first dehydrogenation step of T2 sample occurred at the temperature of 400 K, and finished at 460 K. The first dehydrogenation peak temperature of T2 sample lowers ca. 10 K compared with that of Tp sample (462 K), and the peak width shrinks ca. 50 K from that of Tp samples. Such a sharp dehydrogenation peak reflects a fast rate of dehydrogenation near the peak temperature. Furthermore, the dehydrogenation activation energies (Ea) of T2 sample and Tp sample were determined by the Kissinger’s method [36]. Through the calculation, the activation energy (Ea) of T2 sample is 109 kJ/mol, which is lower than that of Tp (127 kJ/mol).
Figure 2. TPD-MS curves of 2Mg(NH2)2–3LiH composite (Tp), and 2Mg(NH2)2–3LiH–LiBH4 composite (T2): (a) H2 signals and (b) NH3 signals.
Figure 2. TPD-MS curves of 2Mg(NH2)2–3LiH composite (Tp), and 2Mg(NH2)2–3LiH–LiBH4 composite (T2): (a) H2 signals and (b) NH3 signals.
Energies 08 06898 g002
Isothermal dehydrogenation and re-hydrogenation of T2 and Tp samples at 416 K were performed. Figure 3 (more detailed information is showed in Figure S1) shows that desorption and absorption kinetics of T2 sample are accelerated remarkably in comparison with the Tp sample. In 100 min T2 sample releases ca. 80% hydrogen, while Tp sample only releases 40% hydrogen, where the total amount of hydrogen is calculated according to Reaction (6) and Reaction (1), respectively. The tangent slopes of the initial linear parts also indicate that the dehydrogenation rate of the T2 sample is ca. 2 times as fast as that of Tp sample. In the re-hydrogenation test, ca. 80% hydrogen can be charged back to T2 sample within 50 min, while under the same condition, only ca. 20% hydrogen can be soaked in Tp sample. And the tangent slopes of the initial linear parts also show that T2 is ca. 18 times as fast as that of the Tp sample.
Figure 3. Isothermal hydrogen desorption and sorption of the 2Mg(NH2)2–3LiH composite (Tp), 2Mg(NH2)2–3LiH–LiBH4 composite (T2) at 416 K under the pressure of 0.01 and 6 MPa, respectively.
Figure 3. Isothermal hydrogen desorption and sorption of the 2Mg(NH2)2–3LiH composite (Tp), 2Mg(NH2)2–3LiH–LiBH4 composite (T2) at 416 K under the pressure of 0.01 and 6 MPa, respectively.
Energies 08 06898 g003
XRD patterns of T2 samples after dehydrogenation and re-hydrogenation at 460 K were collected for structural analyses. As shown in Figure 4a, the dehydrogenation products of T2 sample are Li2Mg2(NH)3 and Li4BN3H10, which is somehow out of our initial expectation. Li2BNH6 has a hexagonal structure, and melts at ~365 K. Li4BN3H10 has a body-centered cubic structure, and melts at ~465 K. Li2BNH6 is less stable than Li4BN3H10 at temperatures above the melting temperature and decomposes to Li4BN3H10 and LiBH4 [39]. However, we did not observe crystalline LiBH4 phase in XRD patterns in Figure 4a, which may become amorpous during this process. After the following re-hydrogenation, Mg(NH2)2 and LiH are regenarated and Li4BN3H10 still remains (seen in Figure 4a). The melting of Li4BN3H10 and/or Li2BNH6 may create a unique reaction environment, allowing interface reaction and mass transport of Reaction (1) to proceed at a faster rate [40]. Figure 4b shows the FTIR spectra of dehydrogenation and re-hydrogenation samples. The symmetric and asymmetric N–H stretching vibrations of Li4BN3H10 are at 3243 and 3302 cm−1, while the symmetric and asymmetric N–H stretching vibrations of LiNH2 are at 3257 and 3310 cm−1. Corresponding to the results of XRD characterization, the vibrations of Li4BN3H10 can be found in the dehydrogenation and re-hydrogenation products of T2 sample. Isothermal dehydrogenation/re-hydrogenation processes of T2 sample are not fully reversible due to the kinetic and/or thermodynamic reasons. Li2Mg2(NH)3 or MgNH is found in the re-hydrogenated sample, which exhibits the vibration at 3197 cm−1.
Figure 4. (a) XRD patterns and (b) FT-IR spectrum of 2Mg(NH2)2–3LiH (Tp) and 2Mg(NH2)2–3LiH–LiBH4 (T2) samples release (R) and soak (S) at 460 K.
Figure 4. (a) XRD patterns and (b) FT-IR spectrum of 2Mg(NH2)2–3LiH (Tp) and 2Mg(NH2)2–3LiH–LiBH4 (T2) samples release (R) and soak (S) at 460 K.
Energies 08 06898 g004

2.2. Pressure–Composition–Isotherm (PCI) Dehydrogenation at High Pressure

Pressure–Composition–Isotherm (PCI) dehydrogenation curves at 473 K of 2Mg(NH2)2–3LiH (Tp), 2Mg(NH2)2–3LiH–LiBH4 (T2) and 2Mg(NH2)2–3LiH–1/3LiBH4 (T4) samples are shown in Figure 5. It can be seen that the equilibrium plateau pressures of T2 and T4 composites at 473 K are almost the same. However, in the region where the amount of H2 desorption is between 0 and 1.2 wt% (2 hydrogen atoms), the dehydrogenation curves of T2 and T4 at the higher pressure show certain differences (see the insert figure), i.e., there is a higher pressure slope of the T2 (2Mg(NH2)2–3LiH–LiBH4) composite than T4 (2Mg(NH2)2–3LiH–1/3LiBH4) composite [36]. The sloping curve indicates that the constitution of the material changes along with the hydrogen desorption. T2 samples with different amount of hydrogen desorption are characterized by means of X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and magic angle spinning nuclear magnetic resonance (MAS NMR).
Figure 5. Dehydrogenation Pressure–Composition–Isotherm (PCI) curves of 2Mg(NH2)2–3LiH (Tp), 2Mg(NH2)2–3LiH–LiBH4 (T2) and 2Mg(NH2)2–3LiH–1/3LiBH4 (T4) composite at 473 K. The inset is the PCI curves at H/(mol of T2 and T4): 0 to 2.
Figure 5. Dehydrogenation Pressure–Composition–Isotherm (PCI) curves of 2Mg(NH2)2–3LiH (Tp), 2Mg(NH2)2–3LiH–LiBH4 (T2) and 2Mg(NH2)2–3LiH–1/3LiBH4 (T4) composite at 473 K. The inset is the PCI curves at H/(mol of T2 and T4): 0 to 2.
Energies 08 06898 g005
XRD patterns of T2 after ball milling and desorbing different amount of hydrogen at 460 K are showed in Figure 6, the pattern of fresh-made sample confirms the presence of starting materials Mg(NH2)2, LiH and LiBH4. When T2 desorbs 0.25 wt% hydrogen, the peaks of LiBH4 disappear and new peaks assignable to Li4BN3H10 are visible. As the amount of desorption is up to 0.58 wt%, peaks belonging to Li2BNH6 and Li4BN3H10 appear at the same time. However, with the increasing amount of released hydrogen, the peaks of Li4BN3H10 become stronger, the peaks of Li2BNH6 are getting weaker and finally disappear upon 1.4 wt% hydrogen is desorbed. Because diffraction peaks of Li2BNH6 are weak and some of them are overlapped with those of Li4BN3H10, FTIR are preformed synchronously.
Figure 6. XRD patterns of T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: 0 wt%, ball-milling stage; 0.25 wt%; 0.38 wt%; 0.58 wt%; 0.71 wt%; 0.94 wt%; and 1.4 wt%.
Figure 6. XRD patterns of T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: 0 wt%, ball-milling stage; 0.25 wt%; 0.38 wt%; 0.58 wt%; 0.71 wt%; 0.94 wt%; and 1.4 wt%.
Energies 08 06898 g006
FTIR spectra of N–H stretching vibrations from the corresponding samples are shown in Figure 7. The B–H stretching frequencies (vibrations from 2000 to 2600 cm−1) were also investigated (Figure S2), but due to the sensitivity of the equipment of FT-IR, there is not much useful information that can be gained. In Figure 7, the post-milled composite gives typical N–H stretching vibrations at 3327 and 3273 cm−1 assigned to Mg(NH2)2, which means Mg(NH2)2 does not react with LiBH4 during the ball milling process. During the PCI dehydrogenation, LiNH2 produced by Reaction (1) is immediately combined with the nearby LiBH4, which causes that the typical N–H stretching vibrations at 3301 and 3258 cm−1 assigned to LiNH2 are not observed. Furthermore, the spectral signals at 3243 and 3301 cm−1, which are assigned to the characteristic vibrations of Li4BN3H10, suggest that LiNH2 reacts with LiBH4 to form Li4BN3H10 at the beginning of PCI dehydrogenation. It is likely that the LiNH2–LiBH4 interface reaction produces the Li2BNH6 layer and Li4BN3H10 layer. However, the possibility of melting and decomposition of Li2BNH6 to Li4BN3H10 and LiBH4 may also exist during the PCI measures. The vibrations at 3246 cm−1 and 3294 cm−1 assigned to Li2BNH6 appear after the dehydrogenation of 0.58 wt%, which corresponds well with the XRD measurement. The solid-state phase of Li2BNH6 may originate from the quenching of the melts. The vibrations of Li2BNH6 are generally weak and broad, and companied with those of Li4BN3H10. Li2BNH6 finally disappears upon desorption of 1.4 wt% hydrogen. We tentatively propose that the formation of Li2BNH6 and Li2Mg2(NH)3 might be the reason for the high pressure slope in the initial dehydrogenation step.
As shown in Figure 8, 11B MAS NMR spectra of T2 samples at the different dehydrogenation degree can be fitted into three species, i.e., LiBH4 (−41.62 ppm), Li4BN3H10 (−39.81 ppm) and Li2BNH6 (−37.88 ppm). Figure 8a shows that the peak of 11B signal gradually becomes asymmetric and shifts to the lower chemical shift in pace with the increasing degree of dehydrogenation of T2 sample. After the weight loss of hydrogen reached 0.58 wt%, the board peak starts to appear the signal of Li2BNH6 (−37.88 ppm), which exhibits strongest intensity when desorption weight of hydrogen comes up to 0.94 wt%, plotted in Figure 8b. To summarize the tendency of the peak shift, Figure 8c reveals that the signal area of Li4BN3H10 increases and the signal peak of LiBH4 decreases with the dehydrogenation. The signal of Li2BNH6 can be distinguished upon releasing 0.58 wt% hydrogen and it soon becomes invisible upon releasing 1.4 wt% hydrogen, which is consistent with the findings of XRD and FT-IR results.
Figure 7. FT-IR spectra of N–H stretching vibrations from T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: 0 wt%, ball-milling stage; 0.25 wt%; 0.38 wt%; 0.58 wt%; 0.71 wt%; 0.94 wt%; and 1.4 wt%.
Figure 7. FT-IR spectra of N–H stretching vibrations from T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: 0 wt%, ball-milling stage; 0.25 wt%; 0.38 wt%; 0.58 wt%; 0.71 wt%; 0.94 wt%; and 1.4 wt%.
Energies 08 06898 g007
Figure 8. (a) 11B magic angle spinning (MAS) NMR spectra of T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: fresh-made, 0.25 wt%, 0.38 wt%, 0.58 wt%, 0.71 wt%, 0.94 wt%, and 1.4 wt%; (b) hydrogen desorption fitted NMR spectra of Figure 8a (0.94 wt%) in comparison with LiBH4, Li2BNH6 and Li4BN3H10; and (c) the normalized 11B magic angle spinning (MAS) NMR peak intensities of LiBH4 (−41.62 ppm), Li2BNH6 (−37.88 ppm) and Li4BN3H10 (−39.81 ppm) in samples with different degree of hydrogen desorption.
Figure 8. (a) 11B magic angle spinning (MAS) NMR spectra of T2 (2Mg(NH2)2–3LiH–LiBH4) samples collected at 460 K at different degree of dehydrogenation: fresh-made, 0.25 wt%, 0.38 wt%, 0.58 wt%, 0.71 wt%, 0.94 wt%, and 1.4 wt%; (b) hydrogen desorption fitted NMR spectra of Figure 8a (0.94 wt%) in comparison with LiBH4, Li2BNH6 and Li4BN3H10; and (c) the normalized 11B magic angle spinning (MAS) NMR peak intensities of LiBH4 (−41.62 ppm), Li2BNH6 (−37.88 ppm) and Li4BN3H10 (−39.81 ppm) in samples with different degree of hydrogen desorption.
Energies 08 06898 g008
Based on above analyses, the reaction of equimolar LiNH2 and LiBH4 is very likely to happen during the high-pressure slope of T2 sample. The melted LiNH2–LiBH4 is cooled to room temperature, where Li4BN3H10, LiBH4, and a small amount of solid Li2BNH6 can be found [39]. The existence of melted LiNH2–LiBH4 may result in a higher desorption pressure. Future work is still needed to make the mechanism clear.

3. Experimental Section

Lithium hydride (LiH, 99%) (Alfa Aesar, Ward Hill, MA, USA) and Lithium borohydride (LiBH4, 95%) (Sigma Aldrich Fine Chemicals, St. Louis, MO, USA) were used without further purification. Magnesium amide (Mg(NH2)2) was synthesized by reacting metallic Mg powder (99%, Sigma) with NH3.
A Retsch PM400 planetary (Haan, Germany) mill was used to ball mill the mixtures of Mg(NH2)2, LiH and LiBH4 with different molar ratios, such as 2:3:1 and 2:3:1/3 (abbreviated as T2 and T4) at 200 rpm for 36 h. Pristine 2Mg(NH2)2–3LiH composite (short for Tp) without additives was also prepared under the same conditions for comparison. Li2Mg2(NH)3 was synthesized by heating post-milled Mg(NH2)2–LiH composite under vacuum for 24 h. Moreover, Li2BNH6 was synthesized by ball-milling the mixture of LiBH4 and LiNH2 with the same molar number for 24 h under the atmosphere of argon. To avoid oxygen and moisture contaminations, all the sample loadings were conducted inside a glove box that was filled with purified argon (O2 < 10 ppm, H2O < 0.1 ppm).
Thermal decomposition properties of samples were carried on a custom-built temperature programmed desorption (TPD)-mass spectrometer (MS, Hiden Analytical Limited, Warrington, UK) combined system. About 10 mg sample was tested each time at a ramping rate of 2 °C/min. H2 and NH3 signals were detected at the m/z ratios of 2 and 15, respectively. Thermogravimetry coupled differential thermal analysis (TG-DTA) measurements were conducted on a STA-449C (Netzsch, Wittelsbacherstraße, Germany), which was installed in the glove box as mentioned above. The heating rate was 2 °C/min and Ar was the carrier gas.
Pressure–Composition–Isotherm (PCI) measurements and dehydrogenation/hydrogenation experiments were carried out on an automatic Sieverts-type apparatus (Hy-Energy scientific instruments PCT pro-2000, Newark, CA, USA). A sample of ca. 200 mg was used each time. Initial pressures in the sample chamber for dehydrogenation and re-hydrogenation experiments were 0.01 and 6 Mpa, respectively.
Isothermal dehydrogenation and re-hydrogenation experiments were carried out on an automatic Sieverts-type apparatus (Advanced Materials Co., PCT, Pittsburgh, PA, USA). Initial pressure in sample chamber for hydrogen desorption was 0.01 MPa, and for absorption was 6 MPa.
Powder X-ray diffraction (XRD) patterns were recorded over a 2θ range of 5°–80° on an X’Pert Pro diffractometer (PANalytical, Almelo, The Netherlands) with Kα radiation at 40 kV and 40 mA. The powder was placed on a self-made cell sealed with an air-tight hood. Fourier transform IR (FTIR) measurements were conducted on a Varian 3100 unit in DRIFT mode to detect the N–H vibration for metal amides or imides. Solid-state 11B nuclear magnetic resonance (NMR) measurements were conducted on a Bruker Advance III 500 NMR spectrometer (Berlin, Germany) with a 4 mm MAS NMR probe working at a frequency of 128.28 MHz.

4. Conclusions

Increasing LiBH4 amount in the 2Mg(NH2)2–3LiH composite leads to a higher dehydrogenation pressure in the slope region. The reason for this change may be due to the melted reaction of LiNH2–LiBH4. Moreover, the kinetics of 2Mg(NH2)2–3LiH–LiBH4 composite are significantly improved in comparison with pristine 2Mg(NH2)2–3LiH sample. Especially, it desorbs ca. 80% hydrogen in 100 min and re-hydrogenates 80% hydrogen in 50 min isothermally at 416 K, which is ca. 2 times and 18 times as fast as those of 2Mg(NH2)2–3LiH composite.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1996-1073/8/7/6898/s1.

Acknowledgments

We acknowledge financial support from the project of National Natural Science Funds for Distinguished Young Scholar (51225206), projects of National Natural Science Foundation of China (Grant Nos 21273229, U1232120, 51301161, 21473181 and 51472237) and CAS-Helmholtz Association Collaborative Funding.

Author Contributions

Han Wang conceived and designed the experiments; Han Wang and Hujun Cao performed the experiments; Han Wang, Ping Chen and Guotao Wu analyzed the data; Teng He and Guotao Wu contributed analysis tools; Han Wang wrote the paper; Guotao Wu and Ping Chen revised it critically for intellectual content; Guotao Wu is the person for final approval of the version to be published.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schlapbach, L.; Zuttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, P.; Zhu, M. Recent progress in hydrogen storage. Mater. Today 2008, 11, 36–43. [Google Scholar] [CrossRef]
  3. Bogdanovic, B.; Schwickardi, M. Ti-doped alkali metal aluminium hydrides as potential novel reversible hydrogen storage materials. J. Alloy. Compd. 1997, 253, 1–9. [Google Scholar] [CrossRef]
  4. Chen, J.; Kuriyama, N.; Xu, Q.; Takeshita, H.T.; Sakai, T. Reversible hydrogen storage via titanium-catalyzed LiAlH4 and Li3AlH6. J. Phys. Chem. B 2001, 105, 11214–11220. [Google Scholar] [CrossRef]
  5. Fichtner, M.; Fuhr, O.; Kircher, O. Magnesium alanate—A material for reversible hydrogen storage. J. Alloy. Compd. 2003, 356, 418–422. [Google Scholar] [CrossRef]
  6. Chlopek, K.; Frommen, C.; Leon, A.; Zabara, O.; Fichtner, M. Synthesis and properties of magnesium tetrahydroborate, Mg(BH4)2. J. Mater. Chem. 2007, 17, 3496–3503. [Google Scholar] [CrossRef]
  7. Miwa, K.; Aoki, M.; Noritake, T.; Ohba, N.; Nakamori, Y.; Towata, S.; Zuttel, A.; Orimo, S. Thermodynamical stability of calcium borohydride Ca(BH4)2. Phys. Rev. B 2006, 74. [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. Chen, P.; Xiong, Z. Metal–N–H systems for the hydrogen storage. Scr. Mater. 2007, 56, 817–822. [Google Scholar] [CrossRef]
  10. Chen, P.; Xiong, Z.T.; Luo, J.Z.; Lin, J.Y.; Tan, K.L. Interaction of hydrogen with metal nitrides and imides. Nature 2002, 420, 302–304. [Google Scholar] [CrossRef] [PubMed]
  11. Luo, W. (LiNH2–MgH2): A viable hydrogen storage system. J. Alloy. Compd. 2004, 381, 284–287. [Google Scholar] [CrossRef]
  12. Nakamori, Y.; Kitahara, G.; Orimo, S. Synthesis and dehydriding studies of Mg–N–H systems. J. Power Sources 2004, 138, 309–312. [Google Scholar] [CrossRef]
  13. Xiong, Z.; Wu, G.; Hu, J.J.; Chen, P. Ternary imides for hydrogen storage. Adv. Mater. 2004, 16, 1522–1525. [Google Scholar] [CrossRef]
  14. Wu, H. Structure of ternary imide Li2Ca(NH)2 and hydrogen storage mechanisms in amide–hydride system. J. Am. Chem. Soc. 2008, 130, 6515–6522. [Google Scholar] [CrossRef] [PubMed]
  15. Xiong, Z.; Wu, G.; Hu, J.; Liu, Y.; Chen, P.; Luo, W.; Wang, J. Reversible hydrogen storage by a Li–Al–N–H complex. Adv. Funct. Mater. 2007, 17, 1137–1142. [Google Scholar] [CrossRef]
  16. Hino, S.; Ichikawa, T.; Leng, H.Y.; Fujii, H. Hydrogen desorption properties of the Ca–N–H system. J. Alloy. Compd. 2005, 398, 62–66. [Google Scholar] [CrossRef]
  17. Hu, J.; Liu, Y.; Wu, G.; Xiong, Z.; Chen, P. Structural and compositional changes during hydrogenation/dehydrogenation of the Li–Mg–N–H system. J. Phys. Chem. C 2007, 111, 18439–18443. [Google Scholar] [CrossRef]
  18. Leng, H.; Ichikawa, T.; Fujii, H. Hydrogen storage properties of Li–Mg–N–H systems with different ratios of LiH/Mg(NH2)2. J. Phys. Chem. B 2006, 110, 12964–12968. [Google Scholar] [CrossRef] [PubMed]
  19. Luo, W.; Wang, J.; Stewart, K.; Clift, M.; Gross, K. Li–Mg–N–H: Recent investigations and development. J. Alloy. Compd. 2007, 446, 336–341. [Google Scholar] [CrossRef]
  20. Xiong, Z.T.; Wu, G.T.; Hu, J.J.; Chen, P.; Luo, W.F.; Wang, J. Investigations on hydrogen storage over Li–Mg–N–H complex—The effect of compositional changes. J. Alloy. Compd. 2006, 417, 190–194. [Google Scholar] [CrossRef]
  21. Chen, X.Y.; Guo, Y.H.; Yu, X.B. Enhanced dehydrogenation properties of modified Mg(NH2)2–LiBH4 composites. J. Phys. Chem. C 2010, 114, 17947–17953. [Google Scholar] [CrossRef]
  22. Noritake, T.; Aoki, M.; Towata, S.; Ninomiya, A.; Nakamori, Y.; Orimo, S. Crystal structure analysis of novel complex hydrides formed by the combination of LiBH4 and LiNH2. Appl. Phys. A 2006, 83, 277–279. [Google Scholar] [CrossRef]
  23. Weidner, E.; Dolci, F.; Hu, J.J.; Lohstroh, W.; Hansen, T.; Bull, D.J.; Fichtner, M. Hydrogenation reaction pathway in Li2Mg(NH)2. J. Phys. Chem. C 2009, 113, 15772–15777. [Google Scholar] [CrossRef]
  24. Li, C.; Liu, Y.F.; Gu, Y.J.; Gao, M.X.; Pan, H.G. Improved hydrogen-storage thermodynamics and kinetics for an RbF-doped Mg(NH2)2–2LiH system. Chem.Asian J. 2013, 8, 2136–2143. [Google Scholar] [CrossRef] [PubMed]
  25. Liang, C.; Liu, Y.F.; Gao, M.X.; Pan, H.G. Understanding the role of K in the significantly improved hydrogen storage properties of a KOH-doped Li–Mg–N–H system. J. Mater. Chem. A 2013, 1, 5031–5036. [Google Scholar] [CrossRef]
  26. Liu, Y.F.; Li, C.; Li, B.; Gao, M.X.; Pan, H.G. Metathesis reaction-induced significant improvement in hydrogen storage properties of the KF-added Mg(NH2)2–2LiH system. J. Phys. Chem. C 2013, 117, 866–875. [Google Scholar] [CrossRef]
  27. Wang, J.H.; Liu, T.; Wu, G.T.; Li, W.; Liu, Y.F.; Araujo, C.M.; Scheicher, R.H.; Blomqvist, A.; Ahuja, R.; Xiong, Z.T.; et al. Potassium-modified Mg(NH2)2/2 LiH system for hydrogen storage. Angew. Chem. Int. Ed. 2009, 48, 5828–5832. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, J.J.; Fichtner, M.; Chen, P. Investigation on the properties of the mixture consisting of Mg(NH2)2, LiH, and LiBH4 as a hydrogen storage material. Chem. Mater. 2008, 20, 7089–7094. [Google Scholar] [CrossRef]
  29. Hu, J.J.; Liu, Y.F.; Wu, G.T.; Xiong, Z.T.; Chua, Y.S.; Chen, P. Improvement of hydrogen storage properties of the Li–Mg–N–H system by addition of LiBH4. Chem. Mater. 2008, 20, 4398–4402. [Google Scholar] [CrossRef]
  30. Durojaiye, T.; Hayes, J.; Goudy, A. Rubidium hydride: An exceptional dehydrogenation catalyst for the lithium amide/magnesium hydride system. J. Phys. Chem. C 2013, 117, 6554–6560. [Google Scholar] [CrossRef]
  31. Yang, J.; Sudik, A.; Siegel, D.J.; Halliday, D.; Drews, A.; Carter, R.O., 3rd; Wolverton, C.; Lewis, G.J.; Sachtler, J.W.; Low, J.J.; et al. A self-catalyzing hydrogen-storage material. Angew. Chem. Int. Ed. 2008, 47, 882–887. [Google Scholar] [CrossRef] [PubMed]
  32. Li, B.; Liu, Y.F.; Gu, J.; Gao, M.X.; Pan, H.G. Synergetic effects of in situ formed CaH2 and LiBH4 on hydrogen storage properties of the Li–Mg–N–H system. Chem. Asian J. 2013, 8, 374–384. [Google Scholar] [CrossRef] [PubMed]
  33. Li, B.; Liu, Y.F.; Gu, J.; Gu, Y.J.; Gao, M.X.; Pan, H.G. Mechanistic investigations on significantly improved hydrogen storage performance of the Ca(BH4)2-added 2LiNH2/MgH2 system. Int. J. Hydrog. Energy 2013, 38, 5030–5038. [Google Scholar] [CrossRef]
  34. Pan, H.G.; Shi, S.B.; Liu, Y.F.; Li, B.; Yang, Y.J.; Gao, M.X. Improved hydrogen storage kinetics of the Li–Mg–N–H system by addition of Mg(BH4)2. Dalton Trans. 2013, 42, 3802–3811. [Google Scholar] [CrossRef] [PubMed]
  35. Li, B.; Liu, Y.F.; Li, C.; Gao, M.X.; Pan, H.G. In situ formation of lithium fast-ion conductors and improved hydrogen desorption properties of the LiNH2–MgH2 system with the addition of lithium halides. J. Mater. Chem. A 2014, 2, 3155–3162. [Google Scholar] [CrossRef]
  36. Cao, H.J.; Wu, G.T.; Zhang, Y.; Xiong, Z.T.; Qiu, J.S.; Chen, P. Effective thermodynamic alteration to Mg(NH2)2–LiH system: Achieving near ambient-temperature hydrogen storage. J. Mater. Chem. A 2014, 2, 15816–15822. [Google Scholar] [CrossRef]
  37. Chater, P.A.; David, W.I.F.; Anderson, P.A. Synthesis and structure of the new complex hydride Li2BH4NH2. Chem. Commun. 2007, 45, 4770–4772. [Google Scholar] [CrossRef] [PubMed]
  38. Wu, H.; Zhou, W.; Udovic, T.J.; Rush, J.J.; Yildirim, T. Structures and crystal chemistry of Li2BNH6 and Li4BN3H10. Chem. Mater. 2008, 20, 1245–1247. [Google Scholar] [CrossRef]
  39. Borgschulte, A.; Jones, M.O.; Callini, E.; Probst, B.; Kato, S.; Zuttel, A.; David, W.I.F.; Orimo, S. Surface and bulk reactions in borohydrides and amides. Energy Environ. Sci. 2012, 5, 6823–6832. [Google Scholar] [CrossRef]
  40. Matsuo, M.; Remhof, A.; Martelli, P.; Caputo, R.; Ernst, M.; Miura, Y.; Sato, T.; Oguchi, H.; Maekawa, H.; Takamura, H.; et al. Complex hydrides with (BH4) and (NH2) anions as new lithium fast-ion conductors. J. Am. Chem. Soc. 2009, 131, 16389–16391. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Wang, H.; Cao, H.; Wu, G.; He, T.; Chen, P. The improved Hydrogen Storage Performances of the Multi-Component Composite: 2Mg(NH2)2–3LiH–LiBH4. Energies 2015, 8, 6898-6909. https://doi.org/10.3390/en8076898

AMA Style

Wang H, Cao H, Wu G, He T, Chen P. The improved Hydrogen Storage Performances of the Multi-Component Composite: 2Mg(NH2)2–3LiH–LiBH4. Energies. 2015; 8(7):6898-6909. https://doi.org/10.3390/en8076898

Chicago/Turabian Style

Wang, Han, Hujun Cao, Guotao Wu, Teng He, and Ping Chen. 2015. "The improved Hydrogen Storage Performances of the Multi-Component Composite: 2Mg(NH2)2–3LiH–LiBH4" Energies 8, no. 7: 6898-6909. https://doi.org/10.3390/en8076898

Article Metrics

Back to TopTop