Next Article in Journal
Acceleration of Chemical Kinetics Computation with the Learned Intelligent Tabulation (LIT) Method
Next Article in Special Issue
A Novel Emergency Gas-to-Power System Based on an Efficient and Long-Lasting Solid-State Hydride Storage System: Modeling and Experimental Validation
Previous Article in Journal
Signal Analysis in Power Systems
Previous Article in Special Issue
Effects of the Chromium Content in (TiVNb)100−xCrx Body-Centered Cubic High Entropy Alloys Designed for Hydrogen Storage Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Hydrogen Storage Properties of Li-RHC System with In-House Synthesized AlTi3 Nanoparticles

by
Thi-Thu Le
1,
Claudio Pistidda
1,*,
Julián Puszkiel
1,2,
María Victoria Castro Riglos
3,
David Michael Dreistadt
1,
Thomas Klassen
1,2 and
Martin Dornheim
1,*
1
Institute of Hydrogen Technology, Helmholtz-Zentrum Hereon GmbH, D-21502 Geesthacht, Germany
2
Faculty of Mechanical Engineering, Helmut Schmidt University/University of the Federal Armed Forces, 22043 Hamburg, Germany
3
Department of Metalphysics, Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET) and Centro Atómico Bariloche, San Carlos de Bariloche R8402AGP, Argentina
*
Authors to whom correspondence should be addressed.
Energies 2021, 14(23), 7853; https://doi.org/10.3390/en14237853
Submission received: 11 October 2021 / Revised: 10 November 2021 / Accepted: 17 November 2021 / Published: 23 November 2021

Abstract

:
In recent years, the use of selected additives for improving the kinetic behavior of the system 2LiH + MgB2 (Li-RHC) has been investigated. As a result, it has been reported that some additives (e.g., 3TiCl3·AlCl3), by reacting with the Li-RHC components, form nanostructured phases (e.g., AlTi3) possessing peculiar microstructural properties capable of enhancing the system’s kinetic behavior. The effect of in-house-produced AlTi3 nanoparticles on the hydrogenation/dehydrogenation kinetics of the 2LiH + MgB2 (Li-RHC) system is explored in this work, with the aim of reaching high hydrogen storage performance. Experimental results show that the AlTi3 nanoparticles significantly improve the reaction rate of the Li-RHC system, mainly for the dehydrogenation process. The observed improvement is most likely due to the similar structural properties between AlTi3 and MgB2 phases which provide an energetically favored path for the nucleation of MgB2. In comparison with the pristine material, the Li-RHC doped with AlTi3 nanoparticles has about a nine times faster dehydrogenation rate. The results obtained from the kinetic modeling indicate a change in the Li-RHC hydrogenation reaction mechanism in the presence of AlTi3 nanoparticles.

Graphical Abstract

1. Introduction

Hydrogen plays a crucial role in the transition from a fossil-fuel-based to a renewable-energy-based society. In comparison with other energy carriers such as natural gas (13.9 kWh/kg) and liquid hydrocarbons (12.4 kWh/kg), hydrogen has the highest gravimetric energy density per mass (33.3 kWh/kg) [1,2]. As a result, in the context of a net-zero CO2 strategy, using hydrogen as an energy vector to replace environmentally harmful fossil fuels will enable, on the one hand, a drastic reduction in CO2 emissions and, on the other hand, a more efficient and synergic utilization of the intermittent and unevenly distributed renewable energy sources [3,4]. However, at standard conditions (0 °C and 1 bar), hydrogen has an extremely low volumetric density (0.089 kg/m3 at room temperature and under 1 bar) [5,6], which results in a low energy per unit volume (0.003 kWh/dm3) when compared with natural gas (0.008 kWh/dm3) and liquid hydrocarbons (10.5 kWh/dm3). In this scenario, the development of advanced storage methods is essential. Hydrogen is conventionally stored in liquid and gaseous form. However, these storage methods face cost and safety issues. For example, compressed gas storage systems need compressors to reach storage pressures of 350–700 bar. In fact, about 12% of the low heating volume (LHV) is necessary to reach a storage pressure of 350 bar (and 15% of the LHV for 700 bar) in the actual compression operation [7]. In comparison to hydrogen compression, the energy spent for hydrogen liquefaction is substantially higher (around 30% of the LHV) due to complex process steps that aim to achieve low temperatures (liquefaction temperature of H2: −252.9 °C) and to avoid the boiling off effect. Such demanding constraints make these physical storage methods unsuitable, especially for long-term storage applications [8,9]. A potentially plausible storage solution for large quantities of hydrogen is geological storage in suitable geological formations (e.g., underground salt caverns, sub-sea reservoirs) [4]. However, a disadvantage is the localized availability of suitable geological formations.
The adsorption of hydrogen on high-surface-area materials has been studied in recent decades as one of the chemical approaches for storing hydrogen. Several materials, including zeolites, carbon-based compounds, and metal–organic frameworks (MOFs), have been studied. The fundamental disadvantage of this method is that, to absorb substantial amounts of hydrogen, cryogenic temperatures must be applied [10,11]. In fact, at room temperature (RT), none of the above-mentioned materials can reach the gravimetric storage capacity given in the targets of the Department of Energy for onboard hydrogen storage in light-duty fuel cell (FC) vehicles [12]. Hydrogen clathrate hydrate structures offer another option to store hydrogen in molecular form. However, this method requires extreme pressures (~3000 bar at 77 °C) or the employment of foreign molecules such as tetrahydrofuran (THF) [6,13,14]. Another chemical method to store hydrogen is in the atomic form in hydrides-based solid-state systems. The advantages of using such a storage approach are the possibility to achieve high volumetric density, low operating pressures, and high safety standards [9,15,16]. Metal hydride-based hydrogen storage solutions have been shown at a scale that is appropriate for effective deployment in the context of the energy transition [17,18,19]. These applications, however, are dependent on the usage of so-called “low-temperature metal hydrides” (i.e., TiFe, Ti1.1CrMn, etc.) that have a poor gravimetric hydrogen capacity despite having excellent volumetric hydrogen density. Aiming at designing storage systems with high gravimetric capacity, complex metal hydrides and complex metal-hydride-based systems have attracted considerable attention. Among the complex metal hydrides, LiBH4 has been considered a potential candidate for solid-state hydrogen storage due to a theoretical gravimetric capacity of 18.5 wt. % H2 and a volumetric hydrogen storage capacity of 121 kg H2/m3 [20,21,22,23,24,25,26,27]. This hydride, however, was initially not considered suitable for reversible hydrogen storage purposes due to the lack of reversibility under moderate temperature and hydrogen pressure [9,10,11,12,13,14]. At the beginning of the 2000s, by applying an approach similar to the one proposed in 1967 for tuning the thermodynamic properties of the MgH2 [28], it was demonstrated that complex metal hydrides such as LiBH4 could be used in combination with other hydrides to store hydrogen under moderate temperature and hydrogen pressure conditions reversibly. These systems are the so-called reactive hydride composites (RHCs) [29,30,31]. These composites consist of the combination of two or more hydrides to form new compounds that might reduce the overall reaction enthalpy of the system and encourage its reversibility while still maintaining the hydrogen storage capacity of the original materials. Examples of such hydride mixtures are 2LiBH4 + MgH2 [30,31], 2NaBH4 + MgH2 [32,33], 2Mg(NH2)2 + 4LiH [34,35], or 6Mg(NH2)2 + 9LiH + LiBH4 [36]. However, RHCs often exhibit poor kinetic behaviors; thus, the dehydrogenation/hydrogen processes occur at temperatures higher than thermodynamically expected. The kinetic constraints are likely to be caused by the involvement of the system components in complex reaction steps taking place at their interfaces. In addition, the hydrogen desorption from a complex hydride such as a borohydride and amides requires high energy to break the strong bonds of anionic complexes (i.e., (BH4) and (NH2)), to recombine the H atoms to form H2 molecules and to promote the long-range diffusion of the reacting species.
Microstructure modifications via mechanical processes (ball milling [37,38,39,40,41], cold rolling [42], high-pressure torsion (HPT) [43], or equal channel angular pressing [44]) can improve the sorption properties of hydride materials. Nevertheless, this enhancement is only temporary due to the unavoidable growth and agglomeration phenomena occurring during dehydrogenation–rehydrogenation cycling. Long-lasting improvements of the kinetic properties of complex hydrides and RHCs can be achieved by using transition metals (TM)-based additives [36,45,46,47,48,49,50,51,52,53,54]. Several investigations on the effect of the addition of TM-based additives on the kinetic properties of the RHCs have been conducted [45,46,50,55,56,57,58,59,60,61]. In most cases, the added components reacted with the hydride system during milling, forming more stable TM compounds. These new compounds have a typical particle size in the nanometric range between 5–20 nm [45,46,50,57,60,61,62].
From metallurgical studies, it is well known that structural similarities between a newly forming phase and its parental phase, lower the interface energy due to the reduced interatomic and interface lattice misfit. In particular, the nucleation of a new phase on a parental one is promoted when their structural properties meet the so-called “Edge-to-Edge model’s criteria” [63], i.e., interatomic spacing of their most close-packed atom rows and their interplanar mismatch are lower than 10% and 6%, respectively.
For the 2LiH + MgB2 (or 2LiBH4 + MgH2) system, it is known that the addition of selected additives leads to the improvement of hydrogenation–dehydrogenation properties via the formation of nano-dispersed TM compounds that act as energetically favored nucleation centers for the formation of MgB2. In our previous work [50], we found that, in the 2LiH + 2MgB2 system (Li-RHC), the nano-sized hexagonal-AlTi3 particles formed in situ during high energy ball milling of 3TiCl3·AlCl3. The in situ formation of nano-sized hexagonal-AlTi3 particles notably enhanced the kinetic dehydrogenation behavior of the Li-RHC composite by promoting the formation of a hexagonal-MgB2 nucleus onto nano-sized hexagonal-AlTi3 particles. Motivated by these outstanding results and aimed at minimizing the reduction of hydrogen storage capacity due to the additive-RHC side reactions, which usually lead to the formation of stable species, e.g., containing Cl, in-house-synthesized AlTi3 nanoparticles were used as a dopant in the Li-RHC system. In particular, the effect of adding the hexagonal AlTi3 nanoparticles (hexa-AlTi3) on the hydrogenation-dehydrogenation kinetic properties of 2LiH + MgB2 was investigated. The obtained results are presented and discussed below.

2. Materials and Methods

Material preparation: lithium hydride (LiH, 95% purity, Sigma Aldrich, St. Louis, MO, USA), magnesium boride (MgB2, 99% purity, Alfa Aesar, Haverhill, MA, USA), aluminum (III) chloride-titanium (III) chloride (3TiCl3∙AlCl3, ~76–78% TiCl3 purity, Fisher Scientific, Waltham, MA, USA) and sodium hydride (NaH, 90% purity, Sigma-Aldrich) were purchased from commercial suppliers without further treatments. One of the most common methods for synthesizing of the Al-Ti system is ball milling [64], thus in this study, AlTi3 nanoparticles were also synthesized via the mechanochemical driven reaction between NaH and 3TiCl3·AlCl3 in the molar ratio of 12:1. The reaction is expected to proceed as follows:
12 NaH ( s ) + ( 3 TiCl 3 · AlCl 3 ) ( s ) 12 NaCl ( s ) + AlTi 3 ( s ) + H 2 ( g )
Two grams of a NaH—3TiCl3·AlCl3 mixture were milled for 180 min in an argon atmosphere using a hardened stainless steel vial and milling balls (ball to powder (BTP) ratio of 20:1) in an 8000M Mixer/Mill High-Energy Ball Mill. After milling, the obtained powder was first washed with distilled water, and then the solid residues were separated from the liquid fraction via filtration. Washing and filtering were carried out under atmospheric conditions. The washing process might have led to the formation of some oxides. Applying Gibbs minimization equilibrium calculations, the most favorable reaction for the AlTi3(s) and H2O(l) interaction at 25 °C and 1 atm was identified and the Gibbs free energy was calculated as a function of the temperature using the HSC Chemistry software [65]. The species considered for the calculation were: AlTi3(s), H2O(l), H2(g), Al2O3(s), TiH2(s), Al(s), TiO(s), TiO2(s), Ti2O3(s), O2(g), and AlTi(s). The thermodynamic data of the AlTi3 added to the database of the software were referred to [66]. The solid residues were dried in a glass oven under a dynamic vacuum (~10−2 bar) at 100 °C for 12 h and then characterized by X-ray diffraction (XRD), high-resolution transmission electron microscopy (HR-TEM).
The steps followed for preparing the system Li-RHC and the Li-RHC doped with the prepared AlTi3 nanoparticles (denoted as Li-RHC + AlTi3) are identical to those followed for the preparation of the system Li-RHC doped with 3TiCl3∙AlCl3 described in [50]. The amount of added AlTi3 was calculated to be equal to the amount of Al-Ti contained in the Li-RHC doped with 7.5(3TiCl3∙AlCl3), that is, about 1.7 wt. % of AlTi3 was added to the Li-RHC system. All materials were handled and prepared in MBraun Unilab glove boxes under a continuously purified air flow (O2 and H2O levels <2 ppm).
X-ray diffraction (XRD): the XRD characterization of the investigated specimens was carried out using a Bruker D8 Discover device. This diffractometer is equipped with a copper X-ray source (Cu Kα radiation, λ= 1.54184 Å) and a VANTEC general area detector from Bruker. The diffraction patterns were acquired in nine steps in the 2θ range from 10° to 90° with an exposure time of 300 s per step. The preparation of the specimens to be analyzed via XRD was carried out in an MBraun Unilab glove box under a continuously purified argon flow (O2 and H2O levels <2 ppm). For this purpose, a small amount of powder (≈5 mg) was placed onto a sample holder and sealed with an airtight transparent dome made of poly-methyl-methacrylate (PMMA) to avoid oxidation or hydrolysis phenomena during the pattern acquisition. For all the XRD patterns, the bump between 10° and ~25° comes from the PMMA dome.
Volumetric analysis: the dehydrogenation–hydrogenation behavior of the studied material was investigated using a Sieverts type apparatus designed by Hydro-Quebec/HERA Hydrogen Storage Systems Inc. (Longueuil, QC, Canada) or in a similar home-built volumetric device. Both machines operate in the same manner, in which a differential pressure gauge connects the sample volume and a reference volume, and thus all temperature-induced pressure changes are compensated and only differences due to sorption processes are recorded. By considering the exact sample mass, typically around 160 mg, the pressure changes are converted into changes of the gravimetric hydrogen content of the investigated samples, i.e., the results are given in the unit “wt. % H2”. If not otherwise described in the text, the dehydrogenation–hydrogenation cycles of each sample were measured at a hydrogen pressure of 3 bar and 400 °C for the dehydrogenation and a hydrogen pressure of 50 bar and 350 °C for the hydrogenation. The chosen pressure and temperature conditions are based on the published thermodynamic data and previous studies for the Li-RHC system [30,31,67]. The dehydrogenation–hydrogenation measurements were carried out until the measured values reached a plateau region that met the criteria: ∆wt. % = 0.0002 and ∆t (min) = 1. To determine the activation energy, several experiments performed on the Li-RHC + AlTi3 material were carried out in the range of temperature between 370 °C and 400 °C under 50 bar for hydrogenation and under 3 bar for dehydrogenation. These kinetic curves were fitted with a selected reaction model [68,69,70,71], and a kinetic constant rate (k) is acquired for each kinetic curve fitted. Then, the activation energy of the Li-RHC + AlTi3 sample can be obtained using the Arrhenius equation:
k = A e E a R T   or   ln   k = E a R T +   ln   A
Plotting ln k versus the inverse of temperature 1/T (T = temperature) yields a straight line with a slope of E a R that allows calculating the activation energy Ea.
High-resolution transmission electron microscopy (HR-TEM): HR-TEM images were obtained using high-resolution transmission electron microscopy (HR-TEM). Observations were completed with a Tecnai F30 microscope (FEI Company, Hillsboro, OR, USA) with an information limit of 0.12 nm and Schottky field emission gun operating at 300 kV. The samples were exposed to air for a very short time. HR-TEM image processing was conducted with the following programs: Digital Micrograph (License no. 90294175), i-TEM (License no. A2382500), and JEMs (License no. IEb59yBDflUMh).

3. Results and Discussion

Figure 1 shows the XRD patterns of the in-house-prepared AlTi3 nanoparticles and Li-RHC-doped AlTi3. The XRD pattern of the mixture of NaH and 3TiCl3·AlCl3 after milling reported in Figure 1a mainly exhibits the presence of NaCl (η). Paying particular attention to the pattern background, in the 2θ range between 30° and 40°, it is possible to notice the presence of low-intensity reflections ascribed to AlTi3 (π). The XRD pattern acquired for the material after washing and drying does not show the presence of the NaCl (Figure 1b). However, the broad diffraction peaks of hexagonal AlTi3 are visible at 2θ = 35.7°, 38.9°, and 40.6°. This result indicates that NaCl was completely removed during the washing process. Moreover, the interaction with water might result in the formation of some oxides. In this regard, composition phase equilibrium calculations were performed to consider the possible formation of oxide species (Supplementary Materials ESI-Figure S1). Taking into account excess water, the most thermodynamically favorable reaction would lead to the formation of Al2O3(s), TiO2(s), and hydrogen release as indicated by Equation (3):
2AlTi3(s) + 15H2O(l) ⇆ Al2O3(s) + 6TiO2(s) + 15H2(g), ΔG0 = −3488.7 kJ/mol
None of the oxide species are seen in the XRD (Figure 1b) and this fact could be ascribed to their low amount or/and their nanometric nature. For the Li-RHC + AlTi3 sample after milling (Figure 1c), the XRD diffraction patterns show the presence of the starting phases (e.g., LiH, MgB2, and AlTi3), indicating the AlTi3 nanoparticles do not react with the other reactants during milling.
The existence of hexagonal AlTi3 nanoparticles as well as possible oxide species in the washed material was verified by employing the HR-TEM technique. Supplementary Materials ESI-Figure S2 shows the TEM diffraction patterns and possible phases in the sample NaH + 3TiCl3·AlCl3 after washing and drying. As shown there, besides the hexagonal AlTi3 phase (observed via XRD technique), other phases such as Al3Ti (tetragonal/hexagonal), AlTi2 (cubic), AlTi3 (cubic), AlTi (tetragonal), and Al2O3 might be present as well. Further details of the TEM characterization are presented in Figure 2 and Figure 3 HR-TEM micrographs and their respective fast Fourier transform (FFT). Simulated diffraction patterns (DP) in the zone axis condition (ZA) for each possible phase were calculated and compared until the best match was found, thus determining to which phase the particles belonged. This type of analysis was performed on several particles. HR-TEM observations combined with the corresponding FFT + simulated DP analysis were performed on several particles. In zone A (Figure 2c) and zone B (Figure 2d), it is possible to observe the presence of hexagonal AlTi3 particles of approximately 10 nm. In addition to the AlTi3 hexagonal nanoparticles, Al2O3 particles were also found in the washed/dried mixture, as observed by HR-TEM (Figure 3). The formation of Al2O3 can be attributed to the partial oxidation of the Al-Ti particles during the washing in distilled water, though the presence of titanium oxides was not verified by HR-TEM.
The effect of the addition of the AlTi3 nanoparticles on the dynamic hydrogen uptake and release properties of the Li-RHC system were examined via the volumetric technique. For comparison, the hydrogenation of the pristine Li-RHC sample is also shown. Figure 4 shows the first hydrogenation both of the pristine Li-RHC and the doped Li-RHC samples. As seen in Figure 4, the hydrogenation kinetic behavior of the Li-RHC material is improved with the presence of additive (i.e., AlTi3 nanoparticles), compared with the pristine Li-RHC material. More specifically, for the pristine material seen in Figure 4B, the hydrogenation starts at approximately 230 °C and reaches an amount of hydrogen absorbed of roughly 2.4 wt. % at 350 °C. After 25 h of isothermal treatment at 350 °C, the hydrogen uptake reaches almost 10 wt. % (Figure 4A). This value is slightly lower than the theoretical hydrogen capacity found for this system (theoretical hydrogen storage capacity of Li-RHC: 11.45 wt. % H2) [31]. The hydrogenation reaction of the doped materials, i.e., Li-RHC + AlTi3 starts at about 260 °C (Figure 4B). The amount of hydrogen stored in the doped materials during the heating to 350 °C is higher than that of the pristine sample. For Li-RHC + AlTi3, 3.7 wt. % H2 was stored. Eventually, a total amount of hydrogen uptake of approx. 9.6 wt. % after 20 h of isothermal treatment at 350 °C is achieved for Li-RHC + AlTi3, respectively (Figure 4A). As observed, the hydrogen storage capacities obtained experimentally are lower than the theoretical value. This is most likely due to the addition of AlTi3 weight. Another possible reason for the loss in hydrogen storage capacity is attributed to the non-homogeneous dispersion of the additive over the host material (Li-RHC), as reported previously [50]. Particularly the hydrogenation kinetics of the Li-RHC + AlTi3 is faster during the heating and the beginning of the isothermal period at 350 °C, compared with the neat Li-RHC, suggesting that the presence of AlTi3 nanoparticles improves the hydrogenation rate of the material. The fact that the onset temperature of hydrogenation of the Li-RHC + AlTi3 is higher than that of the pristine Li-RHC sample might be related to the experimental fluctuations during the acquisition of the initial points of the measurement.
Similar to hydrogenation, the dehydrogenation kinetic behavior was also analyzed. Figure 5 presents the dehydrogenation properties of the Li-RHC + AlTi3 sample over four cycles compared with the undoped Li-RHC sample. The dehydrogenation of the first reaction step, which relates to the decomposition of MgH2, are almost the same for all samples and a corresponding amount of hydrogen release is about 2.2 wt. %. However, for the second dehydrogenation step, the dehydrogenation kinetic behavior of the Li-RHC composite is significantly improved and the dehydrogenation is more stable by the presence of the dopant (i.e., AlTi3). The dehydrogenation kinetics of the Li-RHC + AlTi3 sample (Figure 5b), for the investigated cycles, are faster than those of the pristine Li-RHC sample (Figure 5a). More specifically, the first dehydrogenation of the Li-RHC + AlTi3 sample requires ~1 h, whereas almost 9 h are necessary for the neat Li-RHC sample. The following dehydrogenation cycles of the Li-RHC + AlTi3 sample are faster than the first dehydrogenation and those of the Li-RHC sample. Furthermore, the stability upon the successive hydrogenation—dehydrogenation cycles also improves markedly for the added AlTi3 material (Figure 5b).
Compared with the pristine material Li-RHC, the addition of AlTi3 nanoparticles significantly improves the hydrogenation–dehydrogenation kinetics of the Li-RHC. However, compared with the material-doped 3TiCl3·AlCl3 [50], the kinetic performance of Li-RHC + AlTi3 leads to similar capacities and slightly slower rates. In this regard, the effective amount of AlTi3 added to the Li-RHC accounts for the observed kinetic performance of the Li-RHC + AlTi3. In this study, the powder obtained after washing and drying contains a mixture of AlTi3 and Al2O3 that appears to be the predominant phases (Figure 2 and Figure 3). Therefore, the used additives are not composed of 100% of AlTi3. The presence of Al2O3 might lead to a reduced number of additional nucleation sites (i.e., AlTi3) for the formation of MgB2, resulting in a slower hydrogen kinetic behavior of the Li-RHC + AlTi3 sample compared with the already reported Li-RHC + 3TiCl3·AlCl3 sample [39].
In the case of these kinds of materials, the crystallographic similarity between the newly forming phase and its parental phase can be evaluated by their interfacial energy. A minimized interfacial energy prompts a proper crystallographic matching between the newly grown phase and the parental phase. Therefore, the promoted nucleation of a product, such as MgB2, results in faster dehydrogenation kinetic behavior.
According to the edge-to-edge matching model [63], the crystallographic matching is determined with small interplanar spacing mismatch fd of close-packed planes (below 6%) and interatomic spacing misfit fs along with the matching close-packed directions (less than 10%) between two phases. In addition, the matching directions have to lie in matching planes. The close-packed plane is commonly based on the structure factor (Fhkl) and the interplanar spacing dhkl. The large values of Fhkl and dhkl result in the smaller area per atom in the plane (hkl), leading to the nearer planes being close-packed [72]. These values (i.e., Fhkl and dhkl) depend on the lattice parameters of the unit cell. In this study, MgB2, AlTi3, and Al2O3 are described in terms of the Miller–Bravais indices and crystal structures with different lattice parameters, leading to different close-packed planes observed on these phases. For MgB2, three possible close-packed planes are { 10 1 ¯ 1 }   { 10 1 ¯ 0 } , and { 0002 } , while the possible close-packed planes for AlTi3 are { 20 2 ¯ 1 } , { 20 2 ¯ 2 } , and { 20 2 ¯ 0 } , and the close-packed plane of Al2O3 is { 11 2 ¯ 0 } . Accordingly, there are a total of nine possible matching plane pairs between MgB2 and AlTi3 and three possible matching plane pairs between MgB2 and Al2O3. The calculation of interplanar spacing mismatch (fd) of these possible matching plane pairs between the MgB2, AlTi3, and Al2O3 are shown in Supplementary Materials ESI_Table S1. As seen there, two possible matching plane pairs for MgB2 and AlTi3 with a mismatch less than 6% are { 10 1 ¯ 1 } MgB 2 // { 20 2 ¯ 1 } AlTi 3 and { 0002 } MgB 2 // { 20 2 ¯ 0 } AlTi 3 . In addition, the close-packed plane pair { 10 1 ¯ 0 } MgB 2 // { 20 2 ¯ 0 } AlTi 3 has a d-value mismatch of 6.3%. This plane pair could also be considered a potential matching plane if the close-packed directions on these matching planes satisfy the interatomic spacing mismatch below 10% [63]. In addition, there are no matching planes between MgB2 and Al2O3 associated with the lattice mismatches of close-packed planes over 6% between these phases. However, the low lattice mismatch of close-packed planes is insufficient to predict orientation relationships between the two MgB2 and AlTi3 phases, particularly for a structure system as complex as the hexagonal one. Within this regard, the close-packed atom directions in two phases have to be matched with an interatomic spacing misfit of less than 10%, and the matching directions must lie on the matching planes. MgB2 and AlTi3 are hexagonal close-packed structures (hcp) in which there are four possible close-packed directions: straight atom row 11 2 ¯ 0 and zigzag atom rows 10 1 ¯ 0 , 11 2 ¯ 3 , and 0001 . These close-packed directions are also reported in literature [73,74,75], while Al2O3 has one close-packed direction 10 1 ¯ 0 (straight row) [76]. As the result, the close-packed directions on the closed packed planes for MgB2: 11 2 ¯ 0 , 11 2 ¯ 3 , and 0001 on { 10 1 ¯ 1 } ; 11 2 ¯ 0 , 10 1 ¯ 0 , and 0001 on { 10 1 ¯ 0 } ; and 11 2 ¯ 0 and 10 1 ¯ 0 on { 0002 } . Similarly, the close-packed directions for AlTi3 can be found: 11 2 ¯ 0 , 11 2 ¯ 3 , and 0001 on { 20 2 ¯ 1 } ; 11 2 ¯ 0 , 10 1 ¯ 0 , and 11 2 ¯ 3 on { 20 2 ¯ 2 } , and 11 2 ¯ 0 , 0001 , and 10 1 ¯ 0 on { 20 2 ¯ 0 } . Following the rule of the edge-to-edge matching model [63], to maximize the matching of atoms across the interface, a straight atom row in one phase matches with straight atom rows in the other phase, and a zigzag row matches with zigzag rows. Considering AlTi3 as a parental phase, the interatomic spacing misfit of direction matching pairs (fs) was calculated, as shown in Supplementary Materials ESI-Table S2 and the mismatch with less than 10% belongs to the matching direction pair 11 2 ¯ 3 MgB 2 // 1101 AlTi 3 (fs = 0.73%). As concluded, there is one matching direction pair 11 2 ¯ 3 MgB 2 // 1101 AlTi 3 on one matching plane { 10 1 ¯ 1 } MgB 2 // { 20 2 ¯ 1 } AlTi 3 . This implies that AlTi3 could act as heterogeneous nucleation sites for nucleating of MgB2. The calculation also points out that Al2O3 does not have a good crystallographic matching with MgB2. The large lattice mismatch between MgB2 and Al2O3 is also found in [77,78]. Therefore, it is not expected that the presence of Al2O3 in the composite improves the reaction kinetics of the Li-RHC system. In addition, during the measurement at high temperature (i.e., 400 °C), Mg might have reacted with Al2O3 to form MgO and Al (3 Mg + Al2O3 = 3 MgO + 2 Al, ΔG400 °C= −122.4 kJ) [79]. The formation of MgO confirmed by XRD, shown in Supplementary Materials ESI-Figure S3, can also negatively react to the hydrogen uptake-release kinetic behavior of the Li-RHC system. The location of nanostructured additives such as AlTi3 at the grain boundaries is crucial for the nucleation of the phase that assured the reversibility of the system, i.e., MgB2. In this regard, the in situ formation of AlTi3 from the precursor 3TiCl3∙AlCl3 during ball milling leads to its better distribution at the grain boundaries. The addition of previously synthesized AlTi3 nanoparticles to Li-RHC might result in an inhomogeneous distribution of this species and, hence, the weaker performance of the system.
To evaluate the rate-limiting step of the first hydrogen absorption process, the gas–solid models described in [70] were applied. Particular emphasis was placed on understanding whether the improved kinetic behavior of the Li-RHC material by the addition of AlTi3 nanoparticles is related to a change in the rate-limiting step. The fitting procedure is based on Sharp’s and Jone’s method [69,71], in which it is assumed that the most suitable reaction model provides a linear fitting with the highest correlation coefficient R2 and a straight line through the origin with a slope close to 1. In this investigation, different reaction kinetic equations (Supplementary Materials ESI-Table S3) [36] were applied to the experimental data of the hydrogenation for the Li-RHC + AlTi3 sample measured at 400 °C and 50 bar of hydrogen. Then, when a good fit of the experimental data concerning a specific kinetic equation can be achieved, the rate-limiting step of the kinetics can be determined. For the fitting, only the converted fraction (α) between 0.1 and 0.9 is considered. This procedure is applied to prevent the uncertainty associated with the acquisition of the initial and final points of the reactions. As an example of the fitting process, the hydrogenation behavior of the Li-RHC + AlTi3 is best described by the three-dimensional contracting volume reaction model (CV, n = 3) with the highest correlation coefficient R2 of 0.997, the slope close to 1, and the intercept equal to 0, amongst others, as shown in Figure 6. Comparison of this result with the data for the pristine Li-RHC and Li-RHC + 3TiCl3∙AlCl3 samples given in [50] proves that the addition of AlTi3 nanoparticles has changed the reaction mechanism for the hydrogenation of the Li-RHC system from a one-dimensional interface-controlled reaction to a three-dimensional interface-controlled reaction.
The activation energy for the hydrogenation process was determined by fitting the kinetic curves with the selected reaction model. The 2nd hydrogenation kinetic curves of the Li-RHC + AlTi3 in the range of temperature between 380 °C and 410 °C were fitted with the CV (n = 3) model. The fitting results and the calculated activation energy for the hydrogenation are presented in Figure 7. As seen in Figure 7a, the measured and the fitted data of the Li-RHC + AlTi3 sample show a good matching with correlation coefficients R2 close to 1. A kinetic rate constant (k) is obtained from the fitting of each hydrogenation curve (Supplementary Materials ESI-Table S4) for each hydrogenation curve fitted. As seen in Figure 7b, the activation energy of the hydrogenation is about 123 ± 3 kJ/mol H2. This value is significantly lower than that of the pristine Li-RHC (Ea = 184 ± 6 kJ/mol H2), and similar to that of the Li-RHC + 3TiCl3·AlCl3 sample (Ea = 124 ± 6 kJ/mol H2) given in [50]. This result indicates that the presence of AlTi3 nanoparticles changed the rate-limiting step and reduced the activation energy of the hydrogenation process, leading to the faster kinetic properties of the Li-RHC + AlTi3 compared with the neat Li-RHC system.
Similar to the hydrogenation process, the activation energy for the dehydrogenation was also determined. Since the dehydrogenation behavior of the Li-RHC + AlTi3 material takes place in two consecutive steps, similar to the dehydrogenation of the Li-RHC and the Li-RHC + 3TiCl3·AlCl3, thus the fitting procedure of this material is equivalent to the ones described for the pristine Li-RHC and Li-RHC + 3TiCl3·AlCl3 materials, in which the combination of two kinetic reaction models are used [50,68]. The dehydrogenation kinetic curves of the Li-RHC + AlTi3 sample measured in the range of 370 °C and 400 °C under three bar of hydrogen were then fitted. The results obtained from the fitting, as presented in Figure 8a, show a good agreement between the measured and fitted data with all correlation coefficients (R2) near 1 (Supplementary Materials ESI-Table S5). Figure 8b shows the plot of ln(k) versus 1/T where the activation energy of the 1st and the 2nd reaction step for the Li-RHC + AlTi3 sample are determined about 150 ± 7 kJ/mol H2 and 270 ± 5 kJ/mol H2, respectively. These values are lower by ~20 (kJ/mol H2) compared with the values of the neat Li-RHC sample (e.g., the 1st reaction step: 172 ± 19 kJ/mol H2 and the 2nd reaction step: 293 ± 12 kJ/mol H2) [50]. However, compared with the sample plus 3TiCl3∙AlCl3 additive also given in [51], the activation energies for dehydrogenation of the Li-RHC + AlTi3 sample are slightly higher. These findings are in good accordance with the kinetic behavior reported above for the investigated materials [50].
Based on the results obtained in this work, the possibility of synthesizing nanometric AlTi3 particles and using those as an additive for the system Li-RHC was investigated. AlTi3 was synthesized via BM of a stoichiometric mixture of NaH and 3TiCl3·AlCl3. XRD and HR-TEM have verified the formation of hexagonal AlTi3 nanoparticles (10 nm). Apart from the existence of a hexagonal AlTi3 phase, nanoparticles of Al2O3 were also observed among the reaction products after the washing and drying procedures. The volumetric results show that upon cycling, the presence of AlTi3 nanoparticles enhances the hydrogenation–dehydrogenation kinetics of the Li-RHC system significantly. However, in comparison with the Li-RHC + 3TiCl3·AlCl3 material specified in [50], the hydrogenation–dehydrogenation kinetic behaviors of the Li-RHC + AlTi3 material are slightly slower. This behavior can be attributed to the lower number of nucleation sites (i.e., AlTi3 species compared with the reference system doped with 3TiCl3·AlCl3) present in the added material.

4. Conclusions

The results reported in this study show that the synthesis and the use of nanoparticles (i.e., AlTi3 nanoparticles) pave the way for a possible solution to improve the sorption kinetics of the RHC systems avoiding considerable hydrogen capacity loss (due to side reactions of additives with the hydride matrix). The synthesis performed in this work can be optimized by making the washing and drying processes in an inert atmosphere so that the formation of oxide compounds such as Al2O3 and MgO is minimized. Further investigations on possible synthetic pathways for producing nanostructured TM-based nanoparticle compounds are part of ongoing research. This approach can lead to an optimal tuning of the dynamic properties of hydride complex systems for efficient hydrogen storage.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/en14237853/s1, Figure S1: Gibbs minimization equilibrium calculations for the interaction between AlTi3 and water, Figure S2: TEM diffraction of the mixture NaH+ 3TiCl3·AlCl3 after washing and drying and possible phases, Figure S3: XRD diffractograms of Li-RHC + AlTi3 sample after 4 cycles, Table S1: Interplanar spacing mismatch (%) of possible close-packed plane pairs between MgB2, AlTi3, and Al2O3, Table S2: Interatomic spacing misfit (%) of possible matching directions between MgB2, AlTi3, and Al2O3, Table S3: Gas–solid reaction models and their integral expressions, Table S4: Fitting parameters of the hydrogenation kinetic curves, Table S5: Fitting parameters of the dehydrogenation kinetic curves.

Author Contributions

Conceptualization, T.-T.L. and C.P.; methodology, T.-T.L.; investigation, T.-T.L.; resources, T.-T.L., D.M.D., J.P. and M.V.C.R.; writing—original draft preparation, T.-T.L.; writing—review and editing, T.-T.L., C.P., M.D., J.P. and T.K.; supervision, C.P., M.D. and T.K.; funding acquisition, T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by DFG (Deutsche Forschungsgemeinschaft), grant number PU 131/16-1.

Data Availability Statement

Data available on request due to restrictions e.g., privacy or ethical. The data presented in this study are available on request from the corresponding author. The data are not publicly available due to the regulations of data protection of the responsible institute.

Acknowledgments

The authors acknowledge Metal Physics Group from HZG (currently Hereon) for the use of the microscope facilities.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353. [Google Scholar] [CrossRef] [PubMed]
  2. Züttel, A.; Borgschulte, A.; Schlapbach, L. Hydrogen as a Future Energy Carrier. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2007, 368, 3329–3342. [Google Scholar] [CrossRef] [PubMed]
  3. Global Warming of 1.5 °C. Summary for Policymakers. Available online: https://www.ipcc.ch/sr15/ (accessed on 30 October 2021).
  4. Hassanpouryouzband, A.; Joonaki, E.; Edlmann, K.; Haszeldine, R.S. Offshore Geological Storage of Hydrogen: Is This Our Best Option to Achieve Net-Zero? ACS Energy Lett. 2021, 6, 2181–2186. [Google Scholar] [CrossRef]
  5. Klebanoff, L. Hydrogen Storage Technology: Materials and Applications; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar]
  6. Hirscher, M. Handbook of Hydrogen Storage: New Material for Future Energy Storage; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010. [Google Scholar]
  7. Felderhoff, M.; Weidenthaler, C.; von Helmolt, R.; Eberle, U. Hydrogen storage: The remaining scientific and technological challenges. Phys. Chem. Chem. Phys. 2007, 9, 2643–2653. [Google Scholar] [CrossRef] [PubMed]
  8. Züttel, A. Materials for Hydrogen Storage. Mater. Today 2003, 6, 24–33. [Google Scholar] [CrossRef]
  9. Broom, D.P. Hydrogen Storage Materials: The Characterisation of Their Storage Properties. In Green Energy and Technology; Springer: London, UK, 2011. [Google Scholar]
  10. Broom, D.P.; Webb, C.J.; Fanourgakis, G.S.; Froudakis, G.E.; Trikalitis, P.N.; Hirscher, M. Concepts for improving hydrogen storage in nanoporous materials. Int. J. Hydrogen Energy 2019, 44, 7768–7779. [Google Scholar] [CrossRef]
  11. Broom, D.P.; Webb, C.J.; Hurst, K.E.; Parilla, P.A.; Gennett, T.; Brown, C.M.; Zacharia, R.; Tylianakis, E.; Klontzas, E.; Froudakis, G.E.; et al. Outlook and challenges for hydrogen storage in nanoporous materials. Appl. Phys. A 2016, 122, 151. [Google Scholar] [CrossRef] [Green Version]
  12. 2025 DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles. Available online: https://www.energy.gov/ (accessed on 30 October 2021).
  13. Lee, H.; Lee, J.-w.; Kim, D.Y.; Park, J.; Seo, Y.-T.; Zeng, H.; Moudrakovski, I.L.; Ratcliffe, C.I.; Ripmeester, J.A. Tuning clathrate hydrates for hydrogen storage. Nature 2005, 434, 743–746. [Google Scholar] [CrossRef] [PubMed]
  14. Florusse Louw, J.; Peters Cor, J.; Schoonman, J.; Hester Keith, C.; Koh Carolyn, A.; Dec Steven, F.; Marsh Kenneth, N.; Sloan, E.D. Stable Low-Pressure Hydrogen Clusters Stored in a Binary Clathrate Hydrate. Science 2004, 306, 469–471. [Google Scholar] [CrossRef]
  15. Jepsen, L.H.; Ley, M.B.; Lee, Y.-S.; Cho, Y.W.; Dornheim, M.; Jensen, J.O.; Filinchuk, Y.; Jørgensen, J.E.; Besenbacher, F.; Jensen, T.R. Boron–nitrogen based hydrides and reactive composites for hydrogen storage. Mater. Today 2014, 17, 129–135. [Google Scholar] [CrossRef] [Green Version]
  16. Ley, M.B.; Jepsen, L.H.; Lee, Y.-S.; Cho, Y.W.; Bellosta von Colbe, J.M.; Dornheim, M.; Rokni, M.; Jensen, J.O.; Sloth, M.; Filinchuk, Y.; et al. Complex hydrides for hydrogen storage—New perspectives. Mater. Today 2014, 17, 122–128. [Google Scholar] [CrossRef] [Green Version]
  17. Buchner, H.; Schmidt.-Ihn, E.; Lang, U. Stationäre Hydridspeicher als Ausgleichsspeicher elektrischer Spitzenenergien. In Kommission der Europäischen Gemeinschaften; European Commission: Luxembourg, 1981. [Google Scholar]
  18. Metallurgy, G.P. GKN Powder Metallurgy. Available online: https://www.gknpm.com/en/innovation/hydrogen-technology/hy2green/ (accessed on 30 October 2021).
  19. Psoma, A.; Sattler, G. Fuel cell systems for submarines: From the first idea to serial production. J. Power Sources 2002, 106, 381–383. [Google Scholar] [CrossRef]
  20. 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]
  21. Orimo, S.; Nakamori, Y.; Kitahara, G.; Miwa, K.; Ohba, N.; Towata, S.; Züttel, A. Dehydriding and rehydriding reactions of LiBH4. J. Alloys Compd. 2005, 404–406, 427–430. [Google Scholar] [CrossRef]
  22. Orimo, S.-I.; Nakamori, Y.; Eliseo, J.R.; Züttel, A.; Jensen, C.M. Complex Hydrides for Hydrogen Storage. Chem. Rev. 2007, 107, 4111–4132. [Google Scholar] [CrossRef] [PubMed]
  23. Orimo, S.-I.; Nakamori, Y.; Ohba, N.; Miwa, K.; Aoki, M.; Towata, S.-I.; Züttel, A. Experimental studies on intermediate compound of LiBH4. Appl. Phys. Lett. 2006, 89, 021920. [Google Scholar] [CrossRef]
  24. Puszkiel, J.; Garroni, S.; Milanese, C.; Gennari, F.; Klassen, T.; Dornheim, M.; Pistidda, C. Tetrahydroborates: Development and Potential as Hydrogen Storage Medium. Inorganics 2017, 5, 74. [Google Scholar] [CrossRef] [Green Version]
  25. Züttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage properties of LiBH4. J. Alloys Compd. 2003, 356–357, 515–520. [Google Scholar] [CrossRef]
  26. Züttel, A.; Wenger, P.; Rentsch, S.; Sudan, P.; Mauron, P.; Emmenegger, C. LiBH4 a new hydrogen storage material. J. Power Sources 2003, 118, 1–7. [Google Scholar] [CrossRef]
  27. Milanese, C.; Jensen, T.R.; Hauback, B.C.; Pistidda, C.; Dornheim, M.; Yang, H.; Lombardo, L.; Zuettel, A.; Filinchuk, Y.; Ngene, P.; et al. Complex hydrides for energy storage. Int. J. Hydrogen Energy 2019, 44, 7860–7874. [Google Scholar] [CrossRef] [Green Version]
  28. Reilly, J.J.; Wiswall, R.H. Reaction of hydrogen with alloys of magnesium and copper. Inorg. Chem. 1967, 6, 2220–2223. [Google Scholar] [CrossRef]
  29. Chen, P.; Xiong, Z.; Luo, J.; Lin, J.; Tan, K.L. Interaction of hydrogen with metal nitrides and imides. Nature 2002, 420, 302. [Google Scholar] [CrossRef]
  30. Barkhordarian, G.; Klassen, T.; Dornheim, M.; Bormann, R. Unexpected kinetic effect of MgB2 in reactive hydride composites containing complex borohydrides. J. Alloys Compd. 2007, 440, L18–L21. [Google Scholar] [CrossRef]
  31. Vajo, J.J.; Skeith, S.L. Reversible Storage of Hydrogen in Destabilized LiBH4. Phys. Chem. B Lett. 2005, 109, 3719–3722. [Google Scholar] [CrossRef] [PubMed]
  32. Garroni, S.; Pistidda, C.; Brunelli, M.; Vaughan, G.B.M.; Suriñach, S.; Baró, M.D. Hydrogen desorption mechanism of 2NaBH4+MgH2 composite prepared by high-energy ball milling. Scr. Mater. 2009, 60, 1129–1132. [Google Scholar] [CrossRef]
  33. Czujko, T.; Varin, R.A.; Wronski, Z.; Zaranski, Z.; Durejko, T. Synthesis and hydrogen desorption properties of nanocomposite magnesium hydride with sodium borohydride (MgH2+NaBH4). J. Alloys Compd. 2007, 427, 291–299. [Google Scholar] [CrossRef]
  34. Xiong, Z.; Hu, J.; Wu, G.; Chen, P.; Luo, W.; Gross, K.; Wang, J. Thermodynamic and kinetic investigations of the hydrogen storage in the Li–Mg–N–H system. J. Alloys Compd. 2005, 398, 235–239. [Google Scholar] [CrossRef]
  35. 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]
  36. Gizer, G.; Puszkiel, J.; Cao, H.; Pistidda, C.; Le, T.T.; Dornheim, M.; Klassen, T. Tuning the reaction mechanism and hydrogenation/dehydrogenation properties of 6Mg(NH2)2-9LiH system by adding LiBH4. Int. J. Hydrogen Energy 2019, 44, 11920–11929. [Google Scholar] [CrossRef]
  37. Bérubé, V.; Radtke, G.; Dresselhaus, M.; Chen, G. Size effects on the hydrogen storage properties of nanostructured metal hydrides: A review. Int. J. Energy Res. 2007, 31, 637–663. [Google Scholar] [CrossRef]
  38. Berlouis, L.E.A.; Cabrera, E.; Hall-Barientos, E.; Hall, P.J.; Dodd, S.; Morris, S.; Imam, M.A. A thermal analysis investigation of the hydriding properties of nanocrystalline Mg–Ni based alloys prepared by high energy ball milling. J. Alloys Compd. 2000, 305, 82–89. [Google Scholar] [CrossRef]
  39. Gertsman, V.Y.; Birringer, R. On the room-temperature grain growth in nanocrystalline copper. Scr. Metall. Mater. 1994, 30, 577–581. [Google Scholar] [CrossRef]
  40. Suzuki, Y.; Haraki, T.; Uchida, H. Effect of LaNi5H6 hydride particles size on desorption kinetics. J. Alloys Compd. 2002, 330–332, 488–491. [Google Scholar] [CrossRef]
  41. Vigeholm, B.; Kjøller, J.; Larsen, B.; Schrøder Pedersen, A. Hydrogen sorption performance of pure magnesium during continued cycling. Int. J. Hydrogen Energy 1983, 8, 809–817. [Google Scholar] [CrossRef]
  42. Vega, L.E.R.; Leiva, D.R.; Leal Neto, R.M.; Silva, W.B.; Silva, R.A.; Ishikawa, T.T.; Kiminami, C.S.; Botta, W.J. Mechanical activation of TiFe for hydrogen storage by cold rolling under inert atmosphere. Int. J. Hydrogen Energy 2018, 43, 2913–2918. [Google Scholar] [CrossRef]
  43. Strozi, R.B.; Ivanisenko, J.; Koudriachova, N.; Huot, J. Effect of HPT on the First Hydrogenation of LaNi5 Metal Hydride. Energies 2021, 14, 6710. [Google Scholar] [CrossRef]
  44. Huot, J. Enhancing Hydrogen Storage Properties of Metal Hybrides Enhancement by Mechanical Deformations. In Briefs in Applied Sciences and Technology; Springer: Cham, Switzerland, 2016. [Google Scholar]
  45. Bösenberg, U.; Kim, J.W.; Gosslar, D.; Eigen, N.; Jensen, T.R.; von Colbe, J.M.B.; Zhou, Y.; Dahms, M.; Kim, D.H.; Günther, R. Role of additives in LiBH4–MgH2 reactive hydride composites for sorption kinetics. Acta Mater. 2010, 58, 3381–3389. [Google Scholar] [CrossRef] [Green Version]
  46. Bosenberg, U.; Vainio, U.; Pranzas, P.K.; von Colbe, J.M.; Goerigk, G.; Welter, E.; Dornheim, M.; Schreyer, A.; Bormann, R. On the chemical state and distribution of Zr- and V-based additives in reactive hydride composites. Nanotechnology 2009, 20, 204003. [Google Scholar] [CrossRef] [PubMed]
  47. Deprez, E.; Justo, A.; Rojas, T.C.; López-Cartés, C.; Bonatto Minella, C.; Bösenberg, U.; Dornheim, M.; Bormann, R.; Fernández, A. Microstructural study of the LiBH4–MgH2 reactive hydride composite with and without Ti-isopropoxide additive. Acta Mater. 2010, 58, 5683–5694. [Google Scholar] [CrossRef] [Green Version]
  48. Fan, M.; Sun, L.; Zhang, Y.; Xu, F.; Zhang, J.; Chu, H. The catalytic effect of additive Nb2O5 on the reversible hydrogen storage performances of LiBH4–MgH2 composite. Int. J. Hydrogen Energy 2008, 33, 74–80. [Google Scholar] [CrossRef]
  49. Gizer, G.; Puszkiel, J.; Riglos, M.V.C.; Pistidda, C.; Ramallo-López, J.M.; Mizrahi, M.; Santoru, A.; Gemming, T.; Tseng, J.-C.; Klassen, T.; et al. Improved kinetic behaviour of Mg(NH2)2-2LiH doped with nanostructured K-modified-LixTiyOz for hydrogen storage. Sci. Rep. 2020, 10, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Le, T.-T.; Pistidda, C.; Puszkiel, J.; Castro Riglos, M.V.; Karimi, F.; Skibsted, J.; GharibDoust, S.P.; Richter, B.; Emmler, T.; Milanese, C.; et al. Design of a Nanometric AlTi Additive for MgB2-Based Reactive Hydride Composites with Superior Kinetic Properties. J. Phys. Chem. C 2018, 122, 7642–7655. [Google Scholar] [CrossRef] [Green Version]
  51. Liu, B.H.; Zhang, B.J.; Jiang, Y. Hydrogen storage performance of LiBH4+1/2MgH2 composites improved by Ce-based additives. Int. J. Hydrogen Energy 2011, 36, 5418–5424. [Google Scholar] [CrossRef]
  52. Santoru, A.; Garroni, S.; Pistidda, C.; Milanese, C.; Girella, A.; Marini, A.; Masolo, E.; Valentoni, A.; Bergemann, N.; Le, T.T.; et al. A new potassium-based intermediate and its role in the desorption properties of the K–Mg–N–H system. Phys. Chem. Chem. Phys. 2016, 18, 3910–3920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sridechprasat, P.; Suttisawat, Y.; Rangsunvigit, P.; Kitiyanan, B.; Kulprathipanja, S. Catalyzed LiBH4 and MgH2 mixture for hydrogen storage. Int. J. Hydrogen Energy 2011, 36, 1200–1205. [Google Scholar] [CrossRef]
  54. Wang, P.; Ma, L.; Fang, Z.; Kang, X.; Wang, P. Improved hydrogen storage property of Li–Mg–B–H system by milling with titanium trifluoride. Energy Environ. Sci. 2009, 2, 120–123. [Google Scholar]
  55. Bonatto Minella, C.; Garroni, S.; Pistidda, C.; Baró, M.D.; Gutfleisch, O.; Klassen, T.; Dornheim, M. Sorption properties and reversibility of Ti(IV) and Nb(V)-fluoride doped-Ca(BH4)2–MgH2 system. J. Alloys Compd. 2015, 622, 989–994. [Google Scholar] [CrossRef] [Green Version]
  56. Karimi, F.; Pranzas, P.K.; Pistidda, C.; Puszkiel, J.A.; Milanese, C.; Vainio, U.; Paskevicius, M.; Emmler, T.; Santoru, A.; Utke, R.; et al. Structural and kinetic investigation of the hydride composite Ca(BH4)2 + MgH2 system doped with NbF5 for solid-state hydrogen storage. Phys. Chem. Chem. Phys. 2015, 17, 27328–27342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bonatto Minella, C.; Pellicer, E.; Rossinyol, E.; Karimi, F.; Pistidda, C.; Garroni, S.; Milanese, C.; Nolis, P.; Baro, M.D.; Gutfleisch, O.; et al. Chemical State, Distribution, and Role of Ti- and Nb-Based Additives on the Ca(BH4)2 System. J. Phys. Chem. C 2013, 117, 4394–4403. [Google Scholar] [CrossRef] [Green Version]
  58. Xian, K.; Gao, M.; Li, Z.; Gu, J.; Shen, Y.; Wang, S.; Yao, Z.; Liu, Y.; Pan, H. Superior Kinetic and Cyclic Performance of a 2D Titanium Carbide Incorporated 2LiH + MgB2 Composite toward Highly Reversible Hydrogen Storage. ACS Appl. Energy Mater. 2019, 2, 4853–4864. [Google Scholar] [CrossRef]
  59. Le, T.T.; Pistidda, C.; Abetz, C.; Georgopanos, P.; Garroni, S.; Capurso, G.; Milanese, C.; Puszkiel, J.; Dornheim, M.; Abetz, V.; et al. Enhanced Stability of Li-RHC Embedded in an Adaptive TPX™ Polymer Scaffold. Materials 2020, 13, 991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Karimi, F.; Pranzas, P.K.; Puszkiel, J.A.; Castro Riglos, M.V.; Milanese, C.; Vainio, U.; Pistidda, C.; Gizer, G.; Klassen, T.; Schreyer, A.; et al. A comprehensive study on lithium-based reactive hydride composite (Li-RHC) as a reversible solid-state hydrogen storage system toward potential mobile applications. RSC Adv. 2021, 11, 23122–23135. [Google Scholar] [CrossRef]
  61. Karimi, F.; Börries, S.; Pranzas, P.K.; Metz, O.; Hoell, A.; Gizer, G.; Puszkiel, J.A.; Riglos, M.V.C.; Pistidda, C.; Dornheim, M.; et al. Characterization of LiBH4–MgH2 Reactive Hydride Composite System with Scattering and Imaging Methods Using Neutron and Synchrotron Radiation. Adv. Eng. Mater. 2021, 23, 2100294. [Google Scholar] [CrossRef]
  62. Karimi, F.; Riglos, M.V.C.; Santoru, A.; Hoell, A.; Raghuwanshi, V.S.; Milanese, C.; Bergemann, N.; Pistidda, C.; Nolis, P.; Baro, M.D.; et al. In Situ Formation of TiB2 Nanoparticles for Enhanced Dehydrogenation/Hydrogenation Reaction Kinetics of LiBH4–MgH2 as a Reversible Solid-State Hydrogen Storage Composite System. J. Phys. Chem. C 2018, 122, 11671–11681. [Google Scholar] [CrossRef]
  63. Zhang, M.X.; Kelly, P.M. Edge-to-edge matching model for predicting orientation relationships and habit planes—the improvements. Scr. Mater. 2005, 52, 963–968. [Google Scholar] [CrossRef]
  64. Moon, K.I.; Lee, K.S. A study of the microstructure of nanocrystalline Al–Ti alloys synthesized by ball milling in a hydrogen atmosphere and hot extrusion. J. Alloys Compd. 1999, 291, 312–321. [Google Scholar] [CrossRef]
  65. HSC Chemistry. Available online: http://www.hsc-chemistry.net/index.html (accessed on 30 October 2021).
  66. Kattner, U.R.; Lin, J.C.; Chang, Y.A. Thermodynamic Assessment and Calculation of the Ti-Al System. Metall. Trans. A 1992, 23, 2081–2090. [Google Scholar] [CrossRef]
  67. Bösenberg, U.; Doppiu, S.; Mosegaard, L.; Barkhordarian, G.; Eigen, N.; Borgschulte, A.; Jensen, T.R.; Cerenius, Y.; Gutfleisch, O.; Klassen, T.; et al. Hydrogen sorption properties of MgH2–LiBH4 composites. Acta Mater. 2007, 55, 3951–3958. [Google Scholar] [CrossRef]
  68. Puszkiel, J.A.; Castro Riglos, M.V.; Ramallo-Lopez, J.M.; Mizrahi, M.; Karimi, F.; Santoru, A.; Hoell, A.; Gennari, F.C.; Larochette, P.A.; Pistidda, C.; et al. A novel catalytic route for hydrogenation-dehydrogenation of 2LiH + MgB2 via in situ formed core-shell LixTiO2 nanoparticles. J. Mater. Chem. A 2017, 5, 12922–12933. [Google Scholar] [CrossRef]
  69. Jones, L.F.; Dollimore, D.; Nicklin, T. Comparison of experimental kinetic decomposition data with master data using a linear plot method. Thermochim. Acta 1975, 13, 240–245. [Google Scholar] [CrossRef]
  70. Khawam, A.; Flanagan, D.R. Solid-State Kinetic Models:  Basics and Mathematical Fundamentals. J. Phys. Chem. B 2006, 110, 17315–17328. [Google Scholar] [CrossRef] [PubMed]
  71. Sharp, J.H.; Brindley, G.W.; Achar, B.N.N. Numerical Data for Some Commonly Used Solid State Reaction Equations. J. Am. Ceram. Soc. 1966, 49, 379–382. [Google Scholar] [CrossRef]
  72. Kelly, P.M.; Ren, H.P.; Qiu, D.; Zhang, M.X. Identifying close-packed planes in complex crystal structures. Acta Mater. 2010, 58, 3091–3095. [Google Scholar] [CrossRef]
  73. Yang, J.; Wang, J.L.; Wu, Y.M.; Wang, L.M.; Zhang, H.J. Extended application of edge-to-edge matching model to HCP/HCP (α-Mg/MgZn2) system in magnesium alloys. Mater. Sci. Eng. A 2007, 460–461, 296–300. [Google Scholar] [CrossRef]
  74. Zhang, M.X.; Kelly, P.M. Edge-to-edge matching and its applications: Part I. Application to the simple HCP/BCC system. Acta Mater. 2005, 53, 1073–1084. [Google Scholar] [CrossRef]
  75. 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]
  76. Lee, W.E.; Lagerlof, K.P.D. Structural and electron diffraction data for sapphire (α-Al2O3). J. Electron. Microsc. Tech. 1985, 2, 247–258. [Google Scholar] [CrossRef]
  77. Won Kyung, S.; Sangjun, O.; Won Nam, K. Perfect Domain-Lattice Matching between MgB2 and Al2O3: Single-Crystal MgB2 Thin Films Grown on Sapphire. Jpn. J. Appl. Phys. 2012, 51, 083101. [Google Scholar]
  78. Tian, W.; Pan, X.Q.; Bu, S.D.; Kim, D.M.; Choi, J.H.; Patnaik, S.; Eom, C.B. Interfacial structure of epitaxial MgB2 thin films grown on (0001) sapphire. Appl. Phys. Lett. 2002, 81, 685–687. [Google Scholar] [CrossRef] [Green Version]
  79. Hallstedt, B.; Liu, Z.-K.; Ågren, J. Reactions in Al2O3-Mg metal matrix composites during prolonged heat treatment at 400, 550 and 600 °C. Mater. Sci. Eng. A 1993, 169, 149–157. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the composites: (a) mixture of NaH + 3TiCl3∙AlCl3 after milling; (b) mixture of NaH + 3TiCl3∙AlCl3 after washing and drying; and (c) the Li-RHC + AlTi3 sample after milling.
Figure 1. XRD patterns of the composites: (a) mixture of NaH + 3TiCl3∙AlCl3 after milling; (b) mixture of NaH + 3TiCl3∙AlCl3 after washing and drying; and (c) the Li-RHC + AlTi3 sample after milling.
Energies 14 07853 g001
Figure 2. HR-TEM characterization of AlTi3 nanoparticles in NaH + 3TiCl3∙AlCl3 material after washing: (a) Bright field TEM image captured with original width; (b) Bright field TEM image cropped ¼ of the original width; (c) FFT + simul. DP + particle in zone A; and (d) FFT + simul. DP + particle in zone B. FFT was calculated in each region and compared with simulated diffraction patterns (DPs) in the adequate orientation; the width of the FFT and corresponding simulations are 18 nm−1.
Figure 2. HR-TEM characterization of AlTi3 nanoparticles in NaH + 3TiCl3∙AlCl3 material after washing: (a) Bright field TEM image captured with original width; (b) Bright field TEM image cropped ¼ of the original width; (c) FFT + simul. DP + particle in zone A; and (d) FFT + simul. DP + particle in zone B. FFT was calculated in each region and compared with simulated diffraction patterns (DPs) in the adequate orientation; the width of the FFT and corresponding simulations are 18 nm−1.
Energies 14 07853 g002
Figure 3. HR-TEM characterization of Al2O3 nanoparticles in NaH + 3TiCl3∙AlCl3 material after washing: (a) Bright field TEM image captured with original width; (b) FFT + simul. DP + particle in zone C; and (c) FFT + simul. DP + particle in zone D. FFT was calculated in each region and compared with simulated diffraction patterns (DPs) in the adequate orientation; the width of the FFT and corresponding simulations are 18 nm−1.
Figure 3. HR-TEM characterization of Al2O3 nanoparticles in NaH + 3TiCl3∙AlCl3 material after washing: (a) Bright field TEM image captured with original width; (b) FFT + simul. DP + particle in zone C; and (c) FFT + simul. DP + particle in zone D. FFT was calculated in each region and compared with simulated diffraction patterns (DPs) in the adequate orientation; the width of the FFT and corresponding simulations are 18 nm−1.
Energies 14 07853 g003
Figure 4. (A) The first hydrogenation reaction of the Li-RHC composites with and without additive, measured from RT to 350 °C, under 50 bar of hydrogen pressure, and (B) magnification of first few hours of hydrogenation of all investigated samples presented in (A).
Figure 4. (A) The first hydrogenation reaction of the Li-RHC composites with and without additive, measured from RT to 350 °C, under 50 bar of hydrogen pressure, and (B) magnification of first few hours of hydrogenation of all investigated samples presented in (A).
Energies 14 07853 g004
Figure 5. Dehydrogenation of the pristine Li-RHC (a) and Li-RHC + AlTi3 (b).
Figure 5. Dehydrogenation of the pristine Li-RHC (a) and Li-RHC + AlTi3 (b).
Energies 14 07853 g005
Figure 6. Plots of (t/t0.5)experimental versus (t/t0.5)theoretical for the 2nd hydrogenation kinetics of the Li-RHC + AlTi3 sample at 400 °C and under 50 bar of H2.
Figure 6. Plots of (t/t0.5)experimental versus (t/t0.5)theoretical for the 2nd hydrogenation kinetics of the Li-RHC + AlTi3 sample at 400 °C and under 50 bar of H2.
Energies 14 07853 g006
Figure 7. (a) The hydrogenation kinetic curves of the Li-RHC + AlTi3 sample at different temperatures fitted with CV (n = 3) model; and (b) calculated hydrogenation activation energy of the Li-RHC + AlTi3 sample.
Figure 7. (a) The hydrogenation kinetic curves of the Li-RHC + AlTi3 sample at different temperatures fitted with CV (n = 3) model; and (b) calculated hydrogenation activation energy of the Li-RHC + AlTi3 sample.
Energies 14 07853 g007
Figure 8. (a) The dehydrogenation kinetics and their fitted curves of Li-RHC + AlTi3 sample at different temperatures; and (b) dehydrogenation activation energy of the Li-RHC + AlTi3 sample.
Figure 8. (a) The dehydrogenation kinetics and their fitted curves of Li-RHC + AlTi3 sample at different temperatures; and (b) dehydrogenation activation energy of the Li-RHC + AlTi3 sample.
Energies 14 07853 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Le, T.-T.; Pistidda, C.; Puszkiel, J.; Castro Riglos, M.V.; Dreistadt, D.M.; Klassen, T.; Dornheim, M. Enhanced Hydrogen Storage Properties of Li-RHC System with In-House Synthesized AlTi3 Nanoparticles. Energies 2021, 14, 7853. https://doi.org/10.3390/en14237853

AMA Style

Le T-T, Pistidda C, Puszkiel J, Castro Riglos MV, Dreistadt DM, Klassen T, Dornheim M. Enhanced Hydrogen Storage Properties of Li-RHC System with In-House Synthesized AlTi3 Nanoparticles. Energies. 2021; 14(23):7853. https://doi.org/10.3390/en14237853

Chicago/Turabian Style

Le, Thi-Thu, Claudio Pistidda, Julián Puszkiel, María Victoria Castro Riglos, David Michael Dreistadt, Thomas Klassen, and Martin Dornheim. 2021. "Enhanced Hydrogen Storage Properties of Li-RHC System with In-House Synthesized AlTi3 Nanoparticles" Energies 14, no. 23: 7853. https://doi.org/10.3390/en14237853

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