Next Article in Journal
Solution Behavior near Envelopes of Characteristics for Certain Constitutive Equations Used in the Mechanics of Polymers
Next Article in Special Issue
Characterization of Glass-Ceramic Sealant for Solid Oxide Fuel Cells at Operating Conditions by Electrochemical Impedance Spectroscopy
Previous Article in Journal
Ablation Behavior of the SiC-Coated Three-Dimensional Highly Thermal Conductive Mesophase-Pitch-Based Carbon-Fiber-Reinforced Carbon Matrix Composite under Plasma Flame
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alanates, a Comprehensive Review

by
Karina Suárez-Alcántara
*,
Juan Rogelio Tena-Garcia
and
Ricardo Guerrero-Ortiz
Morelia Unit of Materials Institute Research, National Autonomus University of Mexico, 58190 Mexico City, Mexico
*
Author to whom correspondence should be addressed.
Materials 2019, 12(17), 2724; https://doi.org/10.3390/ma12172724
Submission received: 22 July 2019 / Revised: 14 August 2019 / Accepted: 21 August 2019 / Published: 25 August 2019
(This article belongs to the Collection Materials and Technologies for Hydrogen and Fuel Cells)

Abstract

:
Hydrogen storage is widely recognized as one of the biggest not solved problem within hydrogen technologies. The slow development of the materials and systems for hydrogen storage has resulted in a slow spread of hydrogen applications. There are many families of materials that can store hydrogen; among them, the alanate family can be of interest. Basic research papers and reviews have been focused on alanates of group 1 and 2. However, there are many alanates of transition metals, main group, and lanthanides that deserve attention in a review. This work is a comprehensive compilation of all known alanates. The approaches towards tuning the kinetics and thermodynamics of alanates are also covered in this review. These approaches are the formation of reactive composites, double cation alanates, or anion substitution. The crystallographic and X-ray diffraction characteristics of each alanate are presented along with this review. In the final sections, a discussion of the infrared, Raman, and thermodynamics was included.

1. Introduction

Hydrogen storage in solid materials is a relatively new branch of hydrogen technologies. It started during the ’60s of the last century with the systematic study of TiFe alloys and Mg [1,2,3]. The studies on hydrogen storage flourished with the spread of the use of mechanical milling to produce materials or precursors that exhibited improved properties regarding kinetics or thermodynamics [4,5,6]. Another breakthrough was the discovery that certain Ti-compounds made the hydrogen storage/release reversible in NaAlH4 [7,8]. Certainly, there are numerous materials that are potentially useful in hydrogen storage. Among them, the family of alanates stands out because of the high hydrogen content, rich chemistry, and the possibility of reversible storage [9]. Alanates (or aluminohydrides) are robust materials; some of them are so well known that prototypes of storage tanks had been constructed (i.e., NaAlH4) [10,11,12]. Others, such as Ti(AlH4)4 or Zr(AlH4)4, are barely known in terms of crystal structure or thermodynamics [13,14]. Figure 1 presents a “periodic table” of the known alanates with dehydrogenation temperatures.
Alanates are like other hydrogen storage materials, in the sense that no material fulfills all of the requirements of hydrogen capacity, dehydrogenation temperatures, or reversibility. The DOE (Department of Energy, USA [15]) had proposed along several decades the figures of merit for hydrogen storage materials and systems, specifying the type of applications (portable, light-duty vehicles, etc.). In general, high hydrogen storage capacity (6.5 wt.% [15]) and reversibility would prevail as the two fundamental characteristics of hydrogen storage materials. The exigencies of the DOE are very rigorous, particularly for light-duty vehicles applications [15], and they include (not limited to) the quantity of stored/released hydrogen (mass and volume of a complete system, 6.5 wt.% and 5 vol.%), reversibility, kinetics (optimum time to charge a hydrogen tank, 3–5 min), minimum number of cycles of hydrogen charge/discharge (1500), operational temperature (−40 to 85 °C), operational pressure (delivery pressure 5–12 bar), cost of the system (266 USD/kg H2), safety, etc. Other factors to be careful with are the thermodynamics (related to the dehydrogenation temperature, but also to the quantity of heat added/removed to/from the system), the onboard efficiency (90%), etc. Moreover, in the future, factors such as recyclability, sustainability, or alanate production from recycled materials [16,17] must also be included as critical factors. However, niche applications for different applications [18] could be developed while using different hydrogen storage materials, including the alanates. These niche applications must meet the particular characteristics of the hydrogen production type and the needs of the final user [18,19]. Nonetheless, the alanate family would allow for the development of new materials. The present work covers the general synthesis procedures, structure, thermodynamics, and hydrogen storage capacity of the known alanates (whenever available). Additionally, double cation alanates or anion substituted materials are also presented and discussed. In the last part of the work, we present a compilation of IR (Infrared) spectroscopy, Raman spectroscopy, and thermodynamics data, along with some general tendencies.

2. General Syntheses Procedures

In this section, the synthesis routes are enumerated, describing them in a general way. Further along in this review, more details are presented for each particular alanate. However, all of the alanates have the need for protective atmospheres during handling, synthesis, and actual hydrogenation or dehydrogenation reactions in common. All of the the alanates can be classified as dangerous materials due to their flammability when exposed to oxygen or humidity. Definitely, they ignite and release hydrogen in contact with water, some more violently than others. Thus, great precautions and security measures must be taken when working with alanates.

2.1. Syntheses in Organic Solvents

2.1.1. Direct Synthesis

Alanates are frequently synthesized by the reaction of metals or metallic hydrides (e.g., NaH) with Al, H2, and a catalyst in organic solvents, such as toluene, hexane, n-octane, ether, diglyme, ether, or tetrahydrofuran (THF) (Equations (1)–(4)) [20,21]. Frequently, a Ti-compound is used as a catalyst. Typically, an excess of Al is used. This method needs the use of moderate to high hydrogen pressure (100–150 bar) and moderate temperatures (120–150 °C); except for LiAlH4, which requires a higher pressure (350 bar) [21]. This method can be considered to be highly dangerous due to the explosive mixture of organic solvents, metal hydride, and Al with oxygen and humidity. The materials thus produced require further steps of purification and drying. Frequently, the alanates are kept and sold in THF solution.
M + ½ H2 → MH,
MH + Al +3/2 H2 ↔ MAlH4, M = Li (low conversion), Na, K, Cs
Frequently, Equation (2) is expanded as a two-step reaction with M3AlH6 as an intermediary [22]:
MH + 1/3 Al + 1/2 H2 ↔ 1/3 M3AlH6, and
1/3 M3AlH6 + 2/3 Al + H2 ↔ MAlH4

2.1.2. Reaction of Metal Hydrides and Aluminum Salts

Another example of lithium alanate synthesis is the reaction of LiH with AlCl3 in refluxing ether under an atmosphere of dry nitrogen [23]:
4LiH + AlCl3 → LiAlH4 + 3LiCl.
This type of reaction is known as “the Schlesinger method”. Despite the simplicity of this reaction, it requires the use of milled LiH (finer than 100 mesh). Additionally, this reaction requires an excess of LiH. Substitution of AlCl3 by AlBr3 can also be effective [24]. The same reaction outline of Equation (5) can be used with NaH or KH, and AlCl3, to produce NaAlH4 and KAlH4, respectively [24]. However, these reactions need the use of Al(C2H5)3 as a catalyst for the reaction with NaH, and C6H6-(C2H5)2O as the solvent; and Al(C2H5)3 or (i-C4H9)2AlH as a catalyst for the reaction with KH [24].
The same type of reaction can be applied to M+2 alanates, such as Mg(AlH4)2 (Equation (6)) [25,26,27] or Ca(AlH4)2 [28], for example:
4MgH2 + 2AlCl3 → Mg(AlH4)2 + 3MgCl2.
No catalyst is used in the last example.
Some materials are obtained rather as THF adducts when this solvent is used [29]. Frequently, the THF adducts cannot be purified (elimination of THF) without the decomposition of the alanate. The use of protective atmospheres during synthesis can improve the yield of the reactions [29]. A general reaction could be described as:
nMHx + xAlCl3 → M(AlH4)x+ (n − 1)MClx
Some of the references for this kind of synthesis are rather old. Initially, this synthesis procedure was not considered for hydrogen storage purposes.

2.1.3. Metathesis of Alanates

Several alanates having one cation or bi-cation have been produced by the metathesis reaction between NaAlH4 or LiAlH4 and metal halides in organic solvents, such as THF or Et2O [30]. One practical reason for this is that NaAlH4 and LiAlH4 are the only commercially available alanates. This type of reaction dates back from 1950, from the work of Wiberg and Bauer [27], and the reaction can be summarized as:
nM1AlH4 + M2Xn → M2(AlH4)n + nM1Xn,
where M1 = Na or Li, M2 = Mg, Ca, or other metals, and X = Cl, Br, I [27,30,31].
Reactions of this type normally are conducted under refluxing conditions from cryogenic to room temperature for several hours or even days. The products usually are adducts of the solvent used, and subsequent operations of purification and drying are required.

2.2. Syntheses Assisted by Mechanical Milling

During the 80s of the last century, the mechanical milling sped up the development of hydrogen storage. We refer both to the study of materials (number of new materials), and the materials themselves towards the storage/release of hydrogen (kinetics of reactions) [5,6]. There are many parameters of mechanical milling. Figure 2 summarizes some of the most important ideas around the mechanical milling that are relevant for the hydrogen storage.
By means of mechanical milling, the same reactions that are described in Section 2.1 can be performed. In most of the cases, the mechanically assisted reactions are faster than the same reactions in solvents. Additionally, the need for solvents is reduced or eliminated. However, some possible problems such as the elimination of side-products (purification) or the contamination by abrasion of balls and vials must be considered. The abrasion of the balls and vial can affect the performance of alanates. Although, an experienced “miller” will know that and would take actions to reduce contamination. These actions can be: (i) Not over-milling. Extended times of milling sometimes can be prejudicial by destroying the alanate, increasing the possibility of abrasion and is a waste of energy. (ii) Check the status of the balls and vial before every milling. (iii) Replace the balls and seals periodically. (iv) Keep the milling vial in good condition. (v) Use compatible materials; there are balls and vials of other materials beyond iron-alloys.

2.2.1. Direct Synthesis by Mechanical Milling

The mechanical milling of the corresponding metal hydride and aluminum and further ex-situ hydrogenation can produce some alanates. Or in-situ by means of a hydrogen atmosphere during the milling. In the ex-situ approach, the hydrogenation of the milled precursors is performed in a specialized reactor (Sieverts type apparatus) to complete Equations (1) and (2). The typical example is the NaAlH4 synthesis by means of mechanical milling of NaH and Al, or Na and Al, and further hydrogenation facilitated by additives, dopants, or catalysts. In the in-situ approach (or reactive mechanical milling), the production of alanates can be attained by the solid-gas reaction between the metal hydride, aluminum, additives, and hydrogen. Again, the typical example is the one-step synthesis of NaAlH4 [32]. The direct synthesis assisted by mechanical milling is an improvement towards “green chemistry”, including solvent free-synthesis [33,34]. Despite the relative simplicity of this method, it is usually performed only in lab-scale for studies of hydrogen storage.
A third approach is the use of solvents (i.e.; wet ball milling) to obtain a precursor mixture of the alanate [35] or the alanate of interest if a hydrogen atmosphere is used. This methodology requires a drying step.

2.2.2. Reaction of Metal Hydrides and Aluminum Salts under Mechanical Milling

Few examples of Equation (7) by mechanical milling have been reported. This absence of data can be related to the instability of some alanates and the consequent difficulty of their synthesis. Among the examples is the work of Hlova et al. [36,37,38]. They produced the reaction of LiH with AlCl3 or NaH with AlCl3 in several molar proportions, with the objective of forming AlH3. Besides forming AlH3, they formed mixtures of LiAlH4-LiAlCl4-Li3AlH6 or NaAlH4-NaAlCl4-Na3AlH6, respectively, under 345–350 bar of hydrogen pressure. In another example, Dymova et al. reacted 2MgH2-AlCl3 to form Mg(AlH4)2 and MgCl2 [39].
However, it must be considered that the mechanochemical version of Equation (5), and Equation (7), in general, or any similar reaction that involves hydrides or alanates plus aluminum salts, would compete with the formation of Al, for example [36]:
3MH + AlCl3 → Al + 3MCl + 1.5H2, (M = Li, Na).
Milling under cryogenic conditions, i.e., with liquid-nitrogen cooling, could be effective in reducing Al formation.

2.2.3. Reaction of Metal Hydrides and Alane under Mechanical Milling

Another disadvantage of the reaction of metal hydrides and aluminum salts is the loss in hydrogen capacity. This is the result of the formation of salts, such as LiCl, NaCl, or MgCl2, which usually are not separated from the products. Alternatively, to avoid the formation of these salts, a reaction of a metal hydride with AlH3 has been proposed [40,41,42], for example:
MgH2 + 2AlH3 → Mg(AlH4)2
In some cases, the reaction did not go to completion, and the formation of intermediaries, such as CaAlH5, was reported [41]. However, the general, complete, reaction would be:
MHx + xAlH3 → M(AlH4)x
The main drawback of this synthesis method is that alane is not a commercial reagent, as such it must be produced in a preliminary step.

2.2.4. Metathesis of Alanates under Mechanical Milling

Several alanates have been produced by metathesis promoted by mechanical milling, Equation (8). Once the milling parameters are well established, this method can be very simple and has many advantages. The main advantages include (i) total elimination of solvents and (ii) significant reduction of the reaction time [43]. However, the main disadvantage is that the produced alanate is impure; the product of milling is a mixture of the alanate and salt. The final result is a drastic reduction of the hydrogen capacity. Examples are the production of Mg(AlH4)2-2NaCl from 2NaAlH4-MgCl2 [43], Ca(AlH4)2-2LiCl from 2LiAlH4-CaCl2 [40], or Eu(AlH4)2-2NaCl from EuCl2-2NaAlH4 [44].

3. The “Single Metal” Alanates

In this section, experimental and theoretical “single metal” alanates are described. They are ordered in groups according to the periodic table. Alane was also included. At the end of this section, the binary (double cations) alanates are presented. Some of them are well-known, while others are barely developed. This section presents the essential characteristics of synthesis, dehydrogenation reactions, and temperatures, crystallographic data, crystal structures, expected X-ray diffraction patterns, and, in some cases, phase diagrams.

3.1. AlH3

The aluminum hydride or alane is a material with high hydrogen content (10 wt.%). A 10 wt.% of hydrogen is very attractive; it meets the DOE targets of 6.5 wt.% of hydrogen for mobile applications. Even more, the low dehydrogenation temperature makes the alane, in principle, compatible with polymer exchange membrane fuel cells (PEMFC) applications. AlH3 is typically produced by the reaction of LiAlH4 with AlCl3 in an organic solvent, such as THF or Et2O [45]:
3LiAlH4 + AlCl3 + nEt2O → 4AlH3·nEt2O + 3LiCl↓.
Instead of LiAlH4, LiH was used in the early studies of this reaction [45]. The product is an adduct that must be separated from the solvent. An excess of LiAlH4 or some LiBH4 is added to the reaction mixture to improve the time and temperature of desolvation [45,46]. The solvent-free mechanosynthesis of AlD3 was performed while using cryomilling 3LiAlD4 + AlCl3 at a low temperature (−196 °C). This conditions eliminated the competing reaction towards the formation of Al and LiCl [47,48]. This synthesis allowed for the determination of the structures of α-AlD3 and α’-AlD3. Mechanical milling of 3LiAlH4 + AlCl3 at room temperature also can produce the alane by using high pressures of hydrogen (210 bar) or inert gas (125 bar of He or 90 bar of Ar) [49]. Alternatively, the alane can be produced by the electrochemical reaction of LiAlH4 or NaAlH4 with or without LiCl as an electrocatalytic additive and with or without hydrogen atmosphere. The general reactions involved are [50,51,52]:
3AlH4 +Al + nTHF → 4AlH3·nTHF + 3e, anode of Al
3(M+ PdH + e→ MH + Pd), cathode of Pd, PdH2
According to reports, the alane has seven polymorphs, and here we present the four most frequently reported (Table 1) [46]. The energy of phase transition between these polymorphs is low: around −1 to −2 kJ/mol H2; thus, the phase transitions occur spontaneously at room temperature (adding complications to the crystal structure determination) [53]. The common structure of the alanes is corner-shared (AlH6) octahedra [54].
The reported formation enthalpy of alane is around −6 to −9 kJ/mol H2; thus, an equilibrium pressure of the order of 105 bar at room temperature is expected [53]. However, the minimum hydrogen pressure, experimentally observed and calculated, which is necessary for the formation of the alane from the elements is about 7000 bar at room temperature [55]. Thus, on-board regeneration of alane for hydrogen storage in automotive applications is definitely out of the picture. Recently, a report on nanoconfined AlH3 indicates partial re-hydrogenation at 150 °C and 60 bar [56]. Nanoconfinement reduces the hydrogen content; however, it must be explored as a way to reach reversibility.
Dehydrogenation enthalpies range from −5 to 6 kJ/mol H2 for the different polymorphs [57], thus near room temperature decomposition would be expected. Dehydrogenation temperatures are observed in the range of 150–200 °C [53,58]; however, ball milling has reduced the dehydrogenation temperature below 100 °C [59]. Alane is considered as a metastable hydride, due to the formation of surface oxides, which protect against to further oxidation or decomposition. The surface oxides impose a kinetic barrier to decomposition [58,60]. In particular, for the alane, the passivation is somehow beneficial, reducing decomposition during its storage and handle in the laboratory. However, in general, passivating surface oxidation is a problem. It is challenging to reduce the oxygen and humidity content of protective atmospheres (argon) until acceptable values (<10 ppm) for hydrogen storage applications. This means that a hydrogen storage system that is based on alanates (and hydrides in general) must have proper filtering, trapping, or regenerative systems to reduce oxygen and humidity content, which can be costly. Ball milling of alane exposes new, fresh, and non-oxidized surfaces that improve the kinetics of the dehydrogenation reaction [61]. The dehydrogenation pathways, as proposed by Sartory et al., are presented in Figure 3 [62].
The thermal dehydrogenation of alane was improved by the use of simple hydrides, such as LiH [63]; otherwise, AlH3 is useful in reducing dehydrogenation temperature or improving dehydrogenation kinetics when added to MgH2 or LiBH4 [64,65].
In the present review, the crystal structures and the calculated X-ray diffraction powder patterns (powder cell 2.3 and mercury software 3.8) are presented for visual comparison. The first of them correspond to the alane polymorphs (Figure 4).

3.2. Alanates of Group 1

3.2.1. Lithium Alanate

The LiAlH4 has the highest hydrogen content of all alanates, 10.6 wt.%; this is due to the lightness of Li atoms. LiAlH4 and NaAlH4 are the only commercially available alanates; their cost, of course, is not low enough for massive applications. Both of them are currently produced while using direct synthesis in an organic solvent. Mechanochemical production of LiAlH4 by the milling of LiH and Al under hydrogen atmosphere has given minimal results [69].
Pure and not milled LiAlH4 undergoes a melting transition, at 160–180 °C before undergoing a first dehydrogenation reaction to give Li3AlH6 and Al at 180–220 °C, Equation (15). A second dehydrogenation reaction is observed to occur at 228–282 °C to give LiH and Al, Equation (16) [70,71]:
LiAlH4 → 1/3 Li3AlH6 + 2/3 Al + H2
1/3 Li3AlH6 → LiH + 1/3Al + ½ H2
Global reaction: LiAlH4 → LiH + Al + 3/2 H2
Together, both reactions provide for a hydrogen release of 7.9 wt.%. The third dehydrogenation step, i.e., the LiH decomposition is beyond any practical hydrogen storage operational temperature. Ball milling and the use of additives have reduced the dehydrogenation temperature of LiAlH4 [72]. The list is extensive among the additives. However, the use of Ti-salts, TiCl3·1/3AlCl3, [73], or NbF5 [74] stands out. Data on apparent activation energies indicate an effective reduction of this parameter upon the use of additives [74]. Blanchard et al. proposed a reduction or elimination of an induction period (slow production rate of Al or Li3AlD6 nuclei) during the decomposition of LiAlD4 as the action mode of the additives [75].
A common characteristic of all alanates is the covalent character of the Al–H bond, while the interaction between [AlH4] or [AlH6]3− and Mn+ is ionic [76]. The crystal structure of α-LiAlH4 (α-LiAlD4) and Li3AlH6 (Li3AlD6) is well-known, as determined both experimentally and by first-principles (Table 2 and Figure 5) [77,78,79]. Additionally, two high-pressure phases, β-LiAlH4, and γ-LiAlH4, have been described [76,80]. The α-LiAlH4 to β-LiAlH4 transition is expected to occur between 26,000 [76] −71,500 [69] bar. The β-LiAlH4 to γ-LiAlH4 transition is expected at 338,000 bar [69]. These pressures are far away from any application in hydrogen storage.
LiAlH4 is a well-known hydrogen storage material due to its facile dehydrogenation, but practically impossible complete rehydrogenation at moderate conditions. Few examples of successful rehydrogenation were observed by transferring the dehydrogenated products to an organic solvent and then exposing them to a hydrogen atmosphere. Among the examples is the rehydrogenation in Me2O at room temperature, 100 bar hydrogen pressure, and 24 h stirring [82,83]. Another reported approach was performing the hydrogenation/dehydrogenation reactions in organic solvent [84]. The experiments and calculations indicate that the LiAlH4 rehydrogenation is thermodynamically restricted [85]. The theoretical (ab-inito) calculations indicate that the dehydrogenation products of LiAlH4 are thermodynamically favored [86]. Ke et al. give the (T, p) stability diagram of LiH and Al versus Li3AlH6; these data indicate the need for very high pressures to produce Li3AlH6 from 3LiH + Al + 3/2H2 (Figure 6). In a (T, p) phase diagram for LiH/Li3AlH6 and Li3AlH6/LiAlH4, Jang et al. demonstrated an equilibrium pressure of about 105 bar for Li3AlH6/LiAlH4 in a wide range of temperatures [87]. Unfortunately, no equation was given to reproduce that equilibrium line. On the other hand, the equilibrium pressure of the direct and reverse reaction in THF;
LiH + Al + 3 / 2 H 2   THF   LiAlH 4 · 4 THF ,  
is in the range of 1–2 bar at 80–90 °C [84]. This equilibrium has been studied in a very limited way. Perhaps, a liquid system of hydrogen storage based on LiAlH4 deserves more attention.

3.2.2. Reactive Mixtures (Composites) with LiAlH4

Reactive mixtures of hydrides have been proposed as a way to tailor the dehydrogenation temperature or improve rehydrogenation in borohydrides [89]. In this approach, two (or recently more) hydrides (simple or complex) are mixed; and, under suitable dehydrogenation conditions, they react with each other. The dehydrogenation is modified, including the dehydrogenation pathway, temperature, kinetics, and reversibility. Notably, the dehydrogenation temperature of composites is sensitive to the way of mixing of materials and the history of the composite; i.e., time and conditions of mixing, purity of reagents, cycling, etc. In the past decade, the research on LiAlH4 has extended, intentionally or not, to the formation of reactive mixtures (composites). Relevant published work is compiled in the next sections.

Composites of LiAlH4-MgH2

Along the last decade, several LiAlH4-MgH2 composites have been studied for hydrogen storage [90,91,92,93,94,95]. The main results coincide in that the dehydrogenation pathway is marked by three steps, the usual two of LiAlH4 and one of MgH2. The temperature of these dehydrogenation steps is somewhat reduced compared to the pure components. Even more, the use of additives, such as TiH2 [96], TiF3 [90], MnFe2O4 [91], or HfCl4 [93], reduced approximately up to 60 °C the dehydrogenation temperatures as compared to the mixture without additives. The role of the additives is to reduce the activation energy of dehydrogenation [93]. Other points of coincidence are the formation of Mg-Al and Li-Mg compounds of relatively varied stoichiometry after dehydrogenation and the occurrence of partial reversibility dominated by MgH2 rehydrogenation without indications of LiAlH4 rehydrogenation.

Composites of LiAlH4-LiBH4

LiBH4 is as a potential hydrogen storage material due to its high hydrogen content. However, the dehydrogenation/hydrogenation high temperature and pressure prevent its use in a pure form. Thus, LiBH4 has been mixed with a variety of chemicals, including LiAlH4, for the formation of binary composites [97,98,99,100]. Additionally, ternary composites of the type LiAlH4-LiBH4-MgF2 have been proposed [101]. In this regard, the possibilities of ternary composites are almost infinite. There are a lot of factors to consider, such as the selection of the composites, the relative composition, milling conditions, etc. Systematic studies are missing, noticeably by the difficulty and enormous of the task. The LiAlH4 did not survive the milling process in many catalyzed mixtures, resulting in a mixture of LiBH4, LiH, and Al [97]. Mao et al. proposed that LiAlH4-LiBH4 doped with TiF3 has a reduced dehydrogenation enthalpy as compared with pure LiBH4 [99]. The reported studies coincide in a two-step dehydrogenation pathway and a reduction of the dehydrogenation temperatures, especially if a catalyst, such as TiF3, is used [99]. The first reaction is the decomposition of LiAlH4 at temperatures around 100 °C. The second step is the decomposition of LiBH4. However, the presence of Al directs the formation of AlB2 [98]:
2LiBH4 +Al ↔ 2LiH + AlB2 + 3H2
The rehydrogenation of the LiAlH4-LiBH4 mixtures was proven to occur at various conditions of pressure and temperature, among them 40, 70, and 85 bar, and 350, 400, and 600 °C [97,98,99]. The rehydrogenation is directed to the formation of LiBH4, since no rehydrogenation of LiAlH4 has reported. While using NaBH4 instead of LiBH4 conduces to similar conclusions; a two-step dehydrogenation with reduced temperature as compared with pure materials, the presence of AlB2 after dehydrogenation, and partial hydrogenation due to the formation of NaBH4 [102].
However, Xia et al. [103] reported the formation of Li3AlH4 and LiBH4 in successive rehydrogenations of 2LiBH4-LiAlH4 confined in mesoporous carbon scaffolds (up to 8.5 wt.% content, rehydrogenation at 500 °C, 100 bar, 10 h, seven cycles). Confinement in meso or nanoporous materials is another strategy for reducing the dehydrogenation temperature and improving the reversibility. However, a reduction in the hydrogen capacity is expected. Other confinement effects are [104,105,106,107]: (i) The reduction or total elimination of the loss of critical elements, such as B in the borohydrides. (ii) Reduction of the diffusion pathways of several species. (iii) Interaction with the meso or nanomaterials supports (can be of catalytic type). (iv) Large surface areas. (v) Reduction of the activation energies. The confinement as a strategy for improving hydrogen storage properties depends on several factors, such as: (i) the material used for confinement (carbons, nanocarbons, zeolites, graphene, silica, etc.) (ii) The history of the confined material. (iii) The way of infiltration (and drying if necessary). (iv) The size of the porous. (v) Functionalization of the surface of the support material. Confinement is a universe of possibilities, and it deserves a mayor review that is beyond the scope of the present report on alanates.

Composites of LiAlH4-LiNH2

The LiAlH4-LiNH2 composites have also been studied [108,109,110,111,112,113]. The first dehydrogenation steps are the decomposition of LiAlH4 to Al and LiH. Then its dehydrogenation products react with LiNH2. Here, less consensus can be found (compared to the previous examples of LiAlH4 composites), and several reactions, mechanisms, and intermediaries have been proposed, for example:
Chen et al. proposed the reaction of LiNH2 with Al as [108]:
LiNH2 → ½ Li2NH + ½ NH3
Al + NH3 → AlN + 3/2 H2
Evidently, due to NH3 production, this method cannot be intended for proton-exchange membrane (PEM) fuel cells.
Dolotko et al. [111] indicated that reaction (21) has a minor contribution to the dehydrogenation reaction, instead, they proposed that LiNH2 reacts with both LiH and Al:
2LiNH2 + LiH + Al → Li3AlN2 + 5/2 H2,
and the overall reaction was proposed as:
LiAlH4 + LiNH2 → ½ Li3AlN2 + ½ Al + ½ LiH +11/4 H2
Lu et al. proposed that the overall reaction is [112]:
2LiAlH4 + LiNH2 → 2Al + Li2NH + LiH + 4H2
Jepsen et al. studied LiAlH4-LiNH2 composites in several molar proportions [113]. The intermediary Li4−xAlx(NH)2−2xN2 formed when the LiAlH4-LiNH2 ratio was 1:1.5, 1:2, and 1:2.5. This study supports that the LiNH2 reacts with LiH to form Li2NH and H2. The main differences among the studies are mainly the molar proportions and milling conditions. This last parameter ranged from some minutes of manual milling in a mortar to several hours of mechanical milling. The use of additives, such as transition metal chlorides reduced, approximately 30 °C, the dehydrogenation temperature [114]. Regarding the reversibility, partial reversibility was proven while using rather hard conditions, i.e., 180 bar and 275 °C [111] or 100 bar and 425 °C [113]. However, the reversibility does not rely on the formation of LiAlH4.

3.2.3. Sodium Alanate

NaAlH4 is the most important and studied alanate. NaAlH4 is used as a reducing agent in many reactions unrelated to the hydrogen storage. Due to the work of Bogdanović et al. on the use of catalysts or additives, the regeneration of NaAlH4 is possible in the solid-state. Thus, this material has been seriously considered for hydrogen storage [7,88,115]. The dehydrogenation and rehydrogenation reactions are [7,116]:
NaAlH4 → 1/3 Na3AlH6 + 2/3 Al + H2
1/3 Na3AlH6 → NaH + 1/3Al + ½ H2
Global reaction: NaAlH4 → NaH + Al + 3/2 H2
Reactions (25) and (26) account for 5.6 wt.% of reversible hydrogen storage. Uncatalyzed NaAlH4 experiences a solid to liquid phase transition before dehydrogenation. Meanwhile, in catalyzed NaAlH4, the dehydrogenation temperature is generally lower than the melting point [117]. The first dehydrogenation step occurs at 210–220 °C. Meanwhile, the second step occurs at approximately 250 °C [117].
The crystal structure of NaAlH4 was determined in 1979 (Table 3 and Figure 7) [118]. The NaAlH4 consists of [AlH4] tetrahedra, with the Na atoms that are surrounded by eight [AlH4] tetrahedra in a distorted square antiprism geometry [119,120].
The NaAlH4 and Na3AlH6 dehydrogenation enthalpies are well known (37 and 47 kJ/mol H2, respectively, Ti-doped) [88]. These values mainly indicate a kinetic restrain for hydrogenation/dehydrogenation reactions, rather than a thermodynamic difficulty (see Section 7 for details on dehydrogenation enthalpies). A phase diagram NaH + Al/Na3AlH4/NaAlH4 can be constructed from these data (Figure 8), [88] which indicates that equilibrium pressures at moderate temperatures are technically achievable, particularly if a catalyst is used. Since the work of Bogdanović, literally, thousands of papers have been published about different catalysts, variations in compositions or variations of the synthesis procedure [123]. NaAlH4 can be produced by all of the methods that are described in Section 2 in several conditions of pressure and temperature at laboratory scale by the use of a catalyst [32,35]. Among the catalysts, dopants, or additives, the list includes, but is not limited to: chlorides of the first and second row of transition metals [124], lanthanide-oxides, such as La2O3, CeO2, Sm2O3, and Gd2O3 [125], titanium compounds, such as Ti(OBu)4 [88], TiCl3 [7], TiF3, TiCl4 [117], TiB2 [126,127], TiN [128], TiCl3-1/3AlCl3 [129], chlorides of Sc and Ce [130], or carbon nanomaterials [131]. The effectiveness of these materials as reaction accelerators is related to the additive level, the addition technique (milling, impregnation with solvent, CVD, etc.), structure, and morphology [127,132].

Role of Catalyst

Among the extensive list of materials tested as catalysts, dopants, or additives for improving hydriding and dehydriding reactions of NaAlH4, the Ti, Sc, and Ce compounds stand out due to their effectiveness [132]. However, most of the theoretical and experimental studies to unravel the action mode of the catalyst have focused on Ti-compounds [134]. Nevertheless, after almost 20 years of the discovery of the benefits of using a catalyst, some fundamental questions are still not adequately addressed. Here are some points to consider:
  • Morphology/particle size effects. Beattie et al. demonstrated that Ti-doped NaAlH4 particles presented few morphological changes as compared with un-doped materials [135]. By-products of the addition of materials, such as TiCl3, i.e., Ti-Al alloys, and NaCl, can act as grain refiners for Al and NaH phases, keeping particle sizes small [136]. In general, much effort is put to reduce particle sizes and to avoid the sintering of particles, and thus maintaining the hydriding/dehydriding performance.
  • Location of Ti and substitution of atoms. The Ti atoms can be located in the bulk, in interstitial positions, at the subsurface, or the surface. The Ti preferred position depends on the doping level and synthesis technique (impregnation vs. ball milling), or in theoretical calculations, the choice of reference states. The Ti atoms can be located in NaH, Al, Na3AlH6, or NaAlH4 phases. Theoretical studies have been performed basically to include all of these possibilities. Some studies have unraveled the interactions of Ti (or Ti-compounds) with NaH and Al. Other reports indicated interactions of Ti (or Ti-compounds) with Na3AlH6 and NaAlH4. Contradictory results/conclusions frequently come across.
    Additionally, many studies point to atom substitution and formation of defects. The replacement of Al by Ti in NaAlH4 could be possible, yet this configuration is metastable [137,138]. Løvvik situates the substitution in the second metal layer from the surface [137,138]. Other DFT calculations suggest that the most frequent Ti-defect in NaAlH4 is a defect that is formed by the substitution of Al by Ti and the addition of two hydrogen ions; this defect has a −1 charge [139].
    The substitution of Na by Ti and other metal atoms also has been investigated. Marashdeh et al. classified the catalysts as “good” (Sc, Ti, Zr) and “bad” (Pt, Pd) according to their ability to exchange places with a Na atom on a (001) surface of NaAlH4 [140]. In the “zipper model”, Ti-species, at the surface or at a grain-boundary, displace subsurface Na atoms and eject them to the NaAlH4 surface. Subsequently, the Na atoms react quickly with other species and destabilize the surface, which returns the Ti-species to a surface location [134,140]. For Na3AlH6, Michel et al. found a competition between Ti substitution on the Na sites (+1 charge defect) and Ti substitution on the Al site, with an additional bound to H atom (neutral site) [139].
    For the hydrogenation reaction, the reports indicate that Ti near an Al surface (subsurface) promotes H2 dissociation and H spillover on the Al surface [141]. Wang et al. remind us, in favor of this role of subsurface Ti, that metallic aluminum does not absorb diatomic hydrogen from the gas phase by itself. Meanwhile, atomic hydrogen strongly reacts with aluminum surfaces to form alanes [142]. Thus, subsurface Ti would promote H2 dissociation and enhance H mobility and adsorption [142]. These effects constitute essentially the “hydrogen pump” action mechanism that was proposed for Ti [134]. Theoretical calculations of subsurface Sc, V or Nb substitution of Al indicate that these materials could also perform as a catalyst [143]. Wang et al. also remind us that Ti, Zr, V, Fe, Ni, Nb, Y, La, Ce, Pr, Nd, and Sm are expected to be good catalysts based on their ability to “mix” well with Al [142].
  • Progressive changes of the oxidation state of Ti-species. While Ti+3 species is the most recurrent initial oxidation state of the Ti-catalyst, several reports conclude that the oxidation state changes to Ti0, followed by the formation of Tix-Aly alloys, and finally the formation of Al3Ti [134,144,145,146]. However, Al3Ti seems to be an inefficient catalyst, as compared to other Ti or Ti-compounds [134,147]. Perhaps the formation of Tix-Aly alloys and Al3Ti is the reason for the long-term (after hydriding/dehydriding cycling) decay of hydrogen storage capacity [148].
  • Formation of Ti-Al-H complexes. Theoretical calculations suggest that the replacement of Na by Ti near o connected with [AlH4] would lead to the formation of Ti-Al-H complexes that can help during the dehydrogenation/rehydrogenation reactions [149,150,151]. TiAl2H7 and TiAl2H2 are two optimized structures of the Ti-Al-H complexes [150]. The effect of the Ti-Al-H complexes would be to reduce the desorption energy of hydrogen [149,151] and to break H-H and Al-H bonds as a result of balanced electron-accepting/back-donation [151].
  • Additional effects. Other effects, such as the formation of mobile species or vacancies, the changes in the Fermi level of reacting species, or the destabilization of Al–H bonds, can also influence the hydrogenation/dehydrogenation reactions [134].

3.2.4. Reactive Mixtures (Composites) with NaAlH4

Composites of NaAlH4-MgH2

Composites of NaAlH4 and MgH2 in several proportions (4:1, 2:1, and 1:1) have been studied in the past years [152,153]. In some cases, catalysts, such as TiF3 [154], TiO2 [155], or TiH2 [156], have been used. The composites in general present four dehydrogenation reactions [152,154]:
NaAlH4 + MgH2 → NaMgH3 +Al + 1.5H2 (170–212 °C)
17MgH2 + 12Al → Mg17Al12 + 17H2 (280–330 °C)
NaMgH3 → NaH + Mg + H2 (330–360 °C)
NaH → Na + ½ H2 (375 °C and higher)
Only the first three reactions are relevant for hydrogen storage purposes. The reported values of hydrogen released in the first cycle of dehydrogenation ranged between 6.7–7.2 wt.% [152,154,155]. However, these values consider the decomposition of NaH. Prolonged ball milling or the use of catalysts produced a decrement of the activation energy and dehydrogenation temperatures in all steps [152,153,154,155,156]. Nano-confinement in carbon aerogel scaffolds reduced the dehydrogenation steps from four to only two [157]. Regarding the reversibility, up to six hydrogenation/dehydrogenation cycles have been demonstrated when the composite is mixed with carbon nanotubes and graphene nanosheets. In this case, the hydrogen storage level is around 3.5 wt.% at 275 °C and 76 bar [158]. Reaction (28) occurs before NaAlH4 decomposes to Na3AlH6. Thus, a mutual destabilization between NaAlH4 and MgH2 was proposed as the reaction drive force [152,154]. Ismail et al. mixed MgH2 and Na3AlH6 (4:1) [159]; in this composite, the dehydrogenation pathway is initiated by the following reaction:
Na3AlH6 + 3MgH2 → 3NaMgH3 +Al + 3/2 H2 (120–250 °C)
The rest of the steps are similar to the reaction sequence (29)–(31).

Other Composites with NaAlH4

The LiBH4-NaAlH4 system was studied in two stoichiometric proportions, 1:0.5 and 1:1.15, with theoretical hydrogen storage capacity of 11.9 and 9.8 wt.%, respectively [160]. A metathesis reaction can occur during ball milling or during heating (~95 °C) depending on the amount of reactants and the energetics of the mixing (mortar vs. ball milling) [160]:
LiBH4 + NaAlH4 → LiAlH4 + NaBH4
The first dehydrogenation reaction is the decomposition of LiAlH4 to produce Li3AlH6, Al and H2 (~105–110 °C). The dehydrogenation pathway differs according to the excess of initial NaAlH4. If an excess of NaAlH4 is present, it reacts with Li3AlH6 to form LiNa2AH6, LiH, Al, and H2 (~200 °C). LiNa2AH6 decomposes at ~290 °C. Without excess of NaAlH4, Li3AlH6 decomposes at ~180 °C. NaBH4 (diffraction peaks) disappear at ~340 °C in both cases. Further heating can lead to the formation of Li-Al alloys and AlB2 phases [160].
Rehydrogenation was confirmed at ~110 bar hydrogen pressure and 400 °C. The rehydrogenation product was LiBH4, as confirmed by in-situ synchrotron radiation powder X-ray diffraction.

3.2.5. Potassium Alanate

KAlH4 has an acceptable total hydrogen content of 5.75 wt.% and a reversible hydrogen storage capacity of 4.3 wt.% (through reactions (34) and (35)). These values are comparable to NaAlH4 and, additionally, KAlH4 does not need a catalyst to undergo re-hydrogenation at a hydrogen pressure as low as 10 bar [161]. KAlH4 can be produced by direct synthesis in organic solvent from KH, Al, and hydrogen [21], or in powder form under high pressure of hydrogen (>175 bar) and heating [162], or by mechanical milling, followed by hydrogen exposure [161], or by the reactive mechanical milling in hydrogen atmosphere [163,164], or by the metathesis of NaAlH4 or LiAlH4 with KCl promoted by ball milling [165].
The dehydrogenation ad re-hydrogenation reactions most “commonly accepted” are [166]:
KAlH4 → 1/3 K3AlH6 + 2/3 Al + H2 (~250–330 °C)
1/3K3AlH6 → KH + 1/3Al + ½ H2 (~340 °C)
Global reaction: KAlH4 → KH + Al + 3/2H2
A third reaction is the decomposition of KH; however, this reaction is not of interest in hydrogen storage applications. An explanation of “commonly accepted” is required; for KAlH4 dehydrogenation and rehydrogenation reactions pathways are still not fully understood. Dehydrogenation pathway involving reactions (34) and (35) are similar to LiAlH4 and NaAlH4, and it is supported by pressure –composition isotherm (PCI) curves that present two plateaus (1 bar and 10 bar) at 355 °C [166,167]. Additionally, some DFT calculations indicate that K3AlH6 is sufficiently thermally stable to behave as an intermediary [168]. Santhanam et al. reported the synthesis of K3AlH6 by 12 h of the mechanical milling of KAlH4 and 2KH [169]. However, a number of experimental reports indicate the presence of other reaction intermediaries, such as KAlH2 [170], AlH3 [171], KxAlHy [167], or other phases with partially known crystallography [172]. Some of them were observed during the in-situ synchrotron radiation powder X-ray diffraction experiments; however, they have not been isolated [172]. The controversial results indicate a possible dependency of the dehydrogenation path of KAlH4 on the operating conditions, as pointed out by Ares et al. [164]. Additives, such as TiCl3 [164,167], or salts, such as NaCl and LiCl (the other product of the ball milling metathesis) [165], could modify the reaction kinetics.
The structure of KAlD4 was reported by Hauback et al. in 2005 (Table 4, Figure 9) [173]. KAlD4 takes the same structure as BaSO4 and KGaD4, i.e., the space group Pnma [173,174]. The experimental and theoretical studies coincide on a small distortion of the [AlH4] ion from the ideal tetrahedron [173,174]. More interesting is the case of the K3AlH6 structure; Vajeeston et al. reported three different K3AlH6 structures according to first-principles studies (Table 4, Figure 9) [175]. The α-K3AlH6 phase is isostructural with α-Na3AlF6, and it transforms into the high-pressure structures β-K3AlH6 and γ-K3AlH6:
α - K 3 AlH 6   534   k b a r   β - K 3 AlH 6   602   k b a r   γ - K 3 AlH 6
The experimental dehydrogenation enthalpies for reactions (34) and (35) are 70 ± 2 and 81 ± 2 kJ/mol H2, respectively [167]. A phase diagram was generated with these values (Figure 10). In this diagram, the feasibility of hydrogenation at low pressure is evident and it justifies the rehydrogenation without the need for a catalyst or additives.

3.2.6. Rubidium Alanate

RbAlH4 has a hydrogen content of 3.4 wt.%. If this material follows the group 1 tendency regarding dehydrogenation reactions, RbAlH4 could reach a 2.5 wt.% of reversible hydrogen storage. Weidenthaler et al. reported the synthesis of RbAlH4 from the metals Al, Rb, and with TiCl3 as an additive; milling was performed in a hydrogen atmosphere (200 bar) [176]. Adkis et al. reported the synthesis of RbAlH4 by the reaction between LiAlH4 and metallic Rb [177]. Bestide et al. reported the metathesis between LiAlH4 and rubidium halides that are assisted by triethylaluminum (AlEt3) in toluene, hexane, and diethyl ether [178]. A stoichiometric reaction (99% product) was almost obtained in the latter work. This reaction yield was explained by the formation of a complex between the halide salts and the triethylaluminum, i.e., a Ziegler-type complex. RbAlH4, and the deuterated species were also produced by the metathesis reaction between NaAlH4, LiAlH4, or LiAlD4 with RbCl or RbF promoted by ball milling [176]. RbAlH4 or RbAlD4 were further heated in an autoclave and then purified [176].
RbAlH4 decomposes in two steps at 300 °C and 350 °C (peak temperatures in TG-DCS curves) [176]. However, no complete dehydrogenation and full reversibility have been demonstrated. There is no consensus regarding the dehydrogenation pathway. Weidenthaler et al. proposed that the two dehydrogenation events are related to the formation of RbH plus Al, and the decomposition of RbH, respectively [176]. For its part, Dymova el at. proposed a first decomposition that is associated with the formation of Rb3AlH6 at 317–334 °C and a second dehydrogenation step by the formation of RbH at 390–417 °C [176,179]. Further confirmation of the dehydrogenation pathway and the formation of Rb3AlH6 is needed.
The structure of RbAlH4 was calculated by Vajeeston et al. [180] and then further confirmed by Weidenthaler et al. (Table 5 and Figure 11) [176]. By means of ab-initio calculations, two high-pressure RbAlH4 phases are anticipated [181]:
RbAlH 4   ( P m m a )   32   k b a r   RbAlH 4   ( I 41 / a )   68   k b a r   RbAlH 4   ( Cmc 21 )
However, no further details regarding the experimental crystallographic data were reported [181]. Ravindran et al. reported the structure of RbAlH4 obtained by theoretical calculations. This structure corresponds to a high-pressure phase above ~55 kbar [182].

3.2.7. Cesium Alanate

CsAlH4 has a hydrogen content of 2.4 wt.%; thus, the interest in CsAlH4 is pure chemistry research and is hardly relevant for hydrogen storage. CsAlH4 has been prepared by mechanical milling or solvent metathesis of NaAlH4 and CsCl, with subsequent purification [183,184]. Previously, Bestide et al. reported the metathesis between LiAlH4 and cesium halides assisted by triethylaluminum (AlEt3) in toluene, hexane, and diethyl ether [178]. CsAlH4 decomposition is marked by four endothermic events [180]:
  • 210–229 °C: polymorphic transition, the material gets an intense yellow color.
  • 280–302: hydrogen evolution due to the proposed reaction:
    3CsAlH4 → 2CsH + CsAl3H8 + H2
  • 454–485 °C: further decomposition reaction of 2CsH + CsAl3H8:
    2CsH + CsAl3H8 → 3Cs + 5H2 + 3Al
  • 666–672 °C: melting of Al. This reaction pathway does not follow the same decomposition and formation of intermediaries as the rest of the alanates of group 1. In-situ diffraction data is missing for further confirmation of this proposed decomposition pathway. Krech et al. [183] demonstrated a reversible polymorphic transformation between orthorhombic and tetragonal CsAlH4; the transformation can be activated by ball-milling or by thermal treatment:
CsAlH 4 ( o ) t h e r m a l   t r e a t m e n t   a t   200 ° C   b a l l m i l l i n g   a t   200   b a r   H 2 CsAlH 4 ( t )
Table 6 lists the collected experimental crystallographic data of cesium alanates (Figure 12).

3.3. Alanates of Group 2

In group 2, in principle, the expected alanates would be M(AlH4)2 and MAlH5. The alanates of group 2 will be discussed in the following sections.

3.3.1. Beryllium-Alanate

The existence of Be(AlH4)2 is questionable. Some reviews list the Be(AlH4)2 phase with a dehydrogenation temperature of 20 °C [186]. The cited reference of these reviews is a book of relatively difficult access [187], whih in turn refers to a series of published works on borohydrides and other boron compounds [188,189]. However, these references dealt with the synthesis of Be(BH4)2, not Be(AlH4)2 [188,189]. In 1973, Ashby et al. attempted to produce Be(AlH4)2 from LiAlH4, or NaAlH4, and BeCl2 in diethyl ether and THF without success [190]. In favor of the existence of Be(AlH4)2 is the report of Wiberg et al. (1951) [191]. In this work, the reaction between BeCl2 and LiAlH4 was proposed to produce Be(AlH4)2 in ether at 20 °C. However, no further details were presented.
Only a few theoretical works on BeAlH5 have been published. Klaveness et al. reported two calculated structures of BeAlH5; the low and high-pressure phases, namely, the α and β phases [192]. However, these calculations were estimated at 0 K, and it was not clear whether BeAlH5 could be stable at ambient conditions in that work. Later, Santhosh et al., also by first-principle calculations, found that the α-BeAlH5 phase could be stable at ambient (p and T) conditions [193]. The calculated α-BeAlH5 phase consisted of alternating sheets of corner-sharing (AlH6) octahedra and chains of corner-sharing (BeH4) tetrahedra. On the other hand, the calculated β-BeAlH5 phase only consisted of chains of corner-sharing (AlH6) octahedra (Table 7 and Figure 13) [192].

3.3.2. Magnesium Alanate

Mg(AlH4)2 has been known since the 1950s [25,27]. At that time, magnesium alanate was synthesized in an organic solvent by the reaction between magnesium hydride and aluminum tri-halides, Equation (6). After almost 50 years, the solid state version of reaction (6) was reported on by Dymova et al. [39] and others [194]. Additionally, roughly at the same time, the metathesis reaction between NaAlH4 and MgCl2 in organic solvent was reported [30]. In the synthesis that involves organic solvents, the formation of adducts, and the purification (drying without decomposing the alanates), is a frequent problem. Thus, more recently, the metathesis reaction between NaAlH4 (or LiAlH4) and MgCl2, as assisted by mechanical milling, was published and frequently used [43,195].
Mg(AlH4)2 has a hydrogen content of 9.3 wt.%; however, dehydrogenation studies report values in the range of 6–7 wt.% in the first dehydrogenation step [196]. The most accepted dehydrogenation pathway assumes that Mg(AlH4)2 decomposes in the temperature range of 110–200 °C, according to the reaction [196,197]:
Mg(AlH4)2 →MgH2 + 2Al + 3H2.
Subsequently, further dehydrogenation of MgH2 in the presence of Al leads to the formation of Mg-Al compounds of several reported stoichiometries [197,198]. Reports indicate that the dehydrogenation temperature can be reduced by additional milling [199], the addition of materials, such as TiF4, TiF3 [200], and TiCl3 [31], or the reduction of particle size [196,198]. Possibly, the other metathesis product, i.e., LiCl or NaCl, can also produce a change in the dehydrogenation temperatures [195]. Rehydrogenation is partially achieved by the formation of MgH2 instead of Mg(AlH4)2 [31,195,200]. However, Gremaud et al. reported the formation of Mg(AlH4)2 at 1 bar, and 100 °C from thin films of Mg-Al covered with a thin layer of Ti [201].
The crystal structure of Mg(AlH4)2 was determined by a combination of X-ray and neutron diffraction at 295 K (Table 8 and Figure 14.) [202]. The crystallographic information is consistent with other experimental and theoretical reports [203,204,205]. The structure consists of [AlH4] tetrahedra that formed double layers that were perpendicular to the c axis of the trigonal cell and alternating with Mg layers (Figure 14) [205].
MgAlH5 was originally proposed by Dymova et al. as a reaction intermediary of the decomposition of Mg(AlH4)2 [39]. However, further experimental reports did not confirm this. Other elements of group 2 (M) indeed form MAlH5 compounds. On the other hand, theoretical calculations indicate the possible crystals structures of MgAlH5 (Table 8).
The few available kinetic studies indicate that the dehydrogenation reaction is ruled by the diffusion of MgH2, Al, or hydrogen in the TiF4 doped samples [197,208]. In any case, the activation energy of dehydrogenation reaction (42) is high: 117.5 [206]–123 [197] kJ/mol. Theoretical studies indicate that Mg(AlH4)2 is metastable at room temperature with a formation enthalpy of −21 kJ/mol H2 [209]. By ab-initio calculations, Spanò et al. determined, that the dehydrogenation temperature at atmospheric pressure must be 111 °C [210]. Thus, despite the interesting high hydrogen content and low dehydrogenation temperatures, Mg(AlH4)2 can be classified as a one-way hydrogen storage material.

3.3.3. Reactive Mixtures (Composites) with Mg(AlH4)2

Few examples of the composites of Mg(AlH4)2 were found during the preparation of this review; they are, in summary: Mg(AlH4)2-NaAlH4 [211,212], Mg(AlH4)2-MgH2 [213], Mg(AlH4)2-LiBH4 [214,215], and Mg(AlH4)2-Ca(BH4)2 [216]. The original reports include several stoichiometries and preparation procedures. However, they have the reduction of dehydrogenation temperature as compared with the individual components and a relatively high amount of hydrogen released during the first dehydrogenation step in common. In many cases, the role of Mg(AlH4)2 is classified as a catalyst of the other components of the mixture. Usually, the first dehydrogenation step corresponds to the decomposition of Mg(AlH4)2, i.e., reaction (42). Afterwards, MgH2 reacts with other components of the mixture. For example, in the Mg(AlH4)2-NaAlH4 composite, NaMgH3 was formed, and the proposed reaction is [211]:
2Na3AlH6 + 6MgH2 → 6NaMgH3 + 2Al + 3H2
Further decomposition of NaMgH3 in the presence of Al leads to the formation of Mg-Al alloys.
For the Mg(AlH4)2-LiBH4 composite, Liu et al. proposed as a second step the formation of Mg2Al3 from the reaction of MgH2 and Al. Subsequently, Mg2Al3 reacts with LiBH4 [214]:
6LiBH4 + 0.5Mg2Al3 + 0.5Al → 6LiH + MgAlB4 +AlB2 + 9H2
The formation of MgAlB4 was also proposed by Pang et al. [215]. Two main drawbacks are observed for the composites of Mg(AlH4)2:
  • Despite the reduction in dehydrogenation temperatures, the “ideal” dehydrogenation temperature—compatible with PEM fuel cells—is not attained.
  • Re-hydrogenation is only partially achieved through the formation of MgH2, not Mg(AlH4)2.

3.3.4. Calcium Alanate

Reports on the synthesis of Ca(AlH4)2 dates from the 1950s [28]; back then, the synthesis was performed in an organic solvent by the reaction between CaH2 and AlCl3 [28] or AlBr3 [217]. As other alanates, the synthesis of Ca(AlH4)2 evolved towards the metathesis reaction between NaAlH4 or LiAlH4 and CaCl2 in an organic solvent [31], to finally take advantage of the use of mechanical milling to perform direct or metathesis synthesis. As for other alanates, the synthesis in organic solvents, such as THF, produced adducts of complicated purification without decomposition of the alanate. Thus, the synthesis that is assisted by mechanical milling is nowadays popular [218].
Calcium alanate has a total hydrogen content of 7.9 wt.%. Its complete decomposition occurs in four steps; the first two steps liberate 5.2–5.9 wt.% hydrogen of the theoretical 7.15 wt.% [165,219,220]. The dehydrogenation reactions are [220]:
Ca(AlH4)2 → CaAlH5 + Al + 3/2H2 (100–160° C)
CaAlH5 → CaH2 + Al + 3/2H2 (220–270 °C)
CaH2 + 4Al → Al4Ca + H2 (~350 °C)
CaH2 + Al4Ca → 2Al2Ca + H2 (~400 °C)
Adding TiF3 led to a decrease in the activation energy and the dehydrogenation temperature [219]. Li et al. suggest that the F atoms from TiF3 substitutes H atoms in Ca(AlH4)2 and that TiF3 initiates the decomposition of calcium alanate [219]:
Ca ( AlH 4 ) 2   +   x 3 TiF 3     CaAlF x H 5 x   +   Al   +   x 3 Ti   +   x + 3 2 H 2 .  
The crystal structure of Ca(AlH4)2 and CaAlH5 were determined in 2009 (Table 9 and Figure 15) [221]. However, theoretical predictions and partial experimental reports were published as early as 2005–2006 [222,223,224]. Ca(AlD4)2 takes an orthorhombic Ca(BF4)2-type structure [221]. Meanwhile, the crystal structure of CaAlD5 was found to be a monoclinic α-SrAlF5-type structure [221]. CaAlH5 consists of corner-sharing (AlH6) octahedra [224].
Ca(AlH4)2 decomposition is slightly exothermic [224], with the enthalpy of reaction (45) being about −7 [220] to −9 kJ/mol H2 [224]. The second dehydrogenation step (reaction (46)) is endothermic with an enthalpy of 32 kJ/mol H2 [224]. The reversibility of Equations (45) and (46) was not reported. The enthalpy values indicate that the first reaction is not suitable for hydrogen storage for mobile applications. However, the second reaction, in principle, could be suitable for mobile hydrogen storage. The enthalpy value of reaction (46) was used to generate the phase diagram of Figure 16. The diagram indicates that CaAlH5 could be produced at temperatures and pressures that are compatible with fuel cells, perhaps with the help of a proper catalyst. In this scenario, the reversible hydrogen storage capacity would be 4.19 wt.%.

3.3.5. Reactive Mixtures (Composites) with Ca(AlH4)2

Scarce examples of reactive mixtures with Ca(AlH4)2 were found during the redaction of this review. One of them was the mixture of LiBH4 and Ca(AlH4)2, giving the best results with a molar ratio of 6:1, respectively [225]. In that system, the released hydrogen was 8.2 wt.% up to 450 °C. Reactions (45) and (46) initiate the dehydrogenation pathway. Subsequently, LiBH4 reacts as: [225]
8LiBH4 + CaH2 + Al → CaB6 + AlB2 + 8LiH + 13H2.
The last step is the reaction (47) of the remaining materials. Rehydrogenation was demonstrated at 450 °C and 40 bar to produce LiBH4 and Ca(BH4)2 and 4.5 wt.% hydrogen storage.
Hanada et al. reported the dehydrogenation of Ca(AlH4)2 + Si, Ca(AlH4)2 + 2MgH2, Ca(AlH4)2 + 2LiH, and Ca(AlH4)2 + 2LiNH2 mixtures that were produced by manual or ball milling [226]. In their work, Ca(AlH4)2 was produced by a metathesis reaction and it was used without purifying, i.e., with the load of NaCl. The weight losses were 6.1 wt.% for Ca(AlH4)2 + 2MgH2 and 5.5 wt.% for manually milled Ca(AlH4)2 + 2LiNH2. These values were reported without taking the load of NaCl into account. The rest of the hydrogen release values were not clearly specified. For the Ca(AlH4)2 + Si mixture, the first two reactions are the usual Ca(AlH4)2 dehydrogenation reactions, Si does not react with CaH2 or Ca-containing phases [226]. For the Ca(AlH4)2 + 2MgH2 mixture, after the usual first two dehydrogenation reactions, MgH2 decomposes at around 270–350 °C and then reacts with Al to form Al12Mg17 [226]. Alapati et al., by means of first-principle calculations, arrived at the same reactions, plus a high-temperature reaction [227]:
6CaH2 + Al12Mg17 → 17Mg + 6Al2Ca + 6H2 (~600 °C)
For the Ca(AlH4)2 + 2LiH mixture, CaAlH5 is generated during ball milling due to the solid-state reaction between Ca(AlH4)2 and LiH [226]. Meanwhile, Li3AlH6 appears after heating to 150 °C under 3 bar of He. Subsequently, at 250 °C, the CaH2 and Al phases arise and Li3AlH6 disappears [226]. For the Ca(AlH4)2 + 2LiNH2 mixture, a reaction of decomposition of Ca(AlH4)2 with LiNH2 occurs during ball milling. A similar hand-milled mixture produced the same two dehydrogenation reactions of Ca(AlH4)2, plus the reaction:
CaH2 + 2LiNH2 → Li2NH + CaNH + 2H2
The last reaction is reported to simultaneously occur with Equation (47) in this system [226]. The re-hydrogenation reactions are not reported.

3.3.6. Strontium Alanates

The system Sr-Al-H presents a richness of chemistry and compounds. Here, we present the most representative characteristics reported for them. Sr(AlH4)2 has a hydrogen content of 5.3 wt%. It was first produced by the reaction between SrH2 and AlH3 by mechanochemical activation in 2000 [228]. After that, Sr(AlH4)2 was produced by the metathesis reaction between SrCl2 and 2NaAlH4, being assisted by mechanical milling [44]. The decomposition of Sr(AlH4)2 initiated at about 130 °C. Subsequently, a second dehydrogenation step occurred at about 240 °C to achieve a total of 2.1 wt.% of released hydrogen with both reactions (0.8 and 1.3 wt.%, respectively, including the NaCl load) [44]. SrAlH5 (4.21 wt.% of total hydrogen content) is proposed as a reaction intermediary of the decomposition of Sr(AlH4)2 [192,228]:
2Sr(AlH4)2 → 2SrAlH5 + 2Al + 3H2 (145–165 °C)
2SrAlH5 → 2SrH2 + 2Al + 3H2 (220–320 °C)
2SrH2 + 4Al → Al4Sr + SrH2 + H2 (355–390 °C)
SrH2 → Sr + H2 (890–950 °C)
Partial rehydrogenation was achieved by (re)milling at high hydrogen pressure (300 bar). Further dehydrogenation demonstrated a drastic reduction of the hydrogen release (about 0.8 wt.%) [44].
SrAlD5 was produced by the mechanical milling of SrD2 and AlD3 and further heating at 154 °C for 1 h in Ar [229]. SrAlD5 was studied by synchrotron and neutron diffraction in detail; the resolved structure consists of (AlD6) octahedra that share corner D atom forming a chain (Figure 17) [229]. This was the first complete experimental report on the crystallography of SrAlD5 (Table 10). Previously, the partial crystal structure (no H positions given) [44] and first-principle crystallographic data of SrAlH5 were reported [192]. The calculated and the experimental data appreciably differ (Figure 17).
Sr2AlH7 (3.37 wt.% of hydrogen content) was produced by the mechanical milling of SrAl2 and further hydrogenation at 70 bar and 270 °C for ten days. The arc melting of Sr and Al previously prepared SrAl2 [230]. Zhang et al. reported that the crystal structure of Sr2AlD7 consisted of isolated (AlD6) units and one-dimensional chains of edge-sharing (DSr4) tetrahedra [230].
The proposed formation pathway is [231,232,233]:
SrAl2 + H2 → SrAl2H2 (190 °C, 50 bar)
4SrAl2H2 + 3H2 → 2Sr2AlH7 + 6Al (240 °C, 70 bar)
The milling of SrAl2 in hydrogen atmosphere led to the formation of SrH2 and Al. The milled materials were further hydrogenated at 260 °C (no pressure indicated) for two days to give Sr2AlH7 [231]:
4SrH2 + 2Al +3H2 → 2Sr2AlH7.
On the other hand, Sr2AlH7 decomposes to SrH2, Al, and H2 at 290 °C [231,232]. However, attempts to hydrogenate mixtures of 4SrH2 + 2Al (70 bar, 260 °C, two days) did not succeed. In this last case, stearic acid was used as a process control agent (PCA) to avoid the cold welding of Al during mechanical milling. Possibly, another PCA might lead to successful hydrogenation. Unfortunately, the dehydrogenation curves of Sr2AlH7 were not presented in these studies.
Table 10 lists the collected crystallographic information of Sr-Al-H compounds.

3.3.7. Barium Alanates

For Barium, two barium-aluminum-hydride compounds have been reported: BaAlH5 (2.97 wt.% hydrogen content) and Ba2AlH7 (2.28 wt.% hydrogen content). They have been prepared from Ba7Al13 or Ba4Al5 alloys, followed by ball-milling and several days in hydrogenation conditions (70 bar, ~200 °C). The Ba7Al13 and Ba4Al5 alloys were previously prepared by arc melting [234,235,236]. Alternatively, the reactive ball milling of the mixture of BaH2 and Al can produce the BaAlH5 and Ba2AlH7 [237]. The formation of BaAlH5 or Ba2AlH7 can be directed by the choice of precursor or by the selection of the temperature (Figure 18) [234,235,236]. BaAlH5 and Al were produced by the hydrogenation of Ba7Al13 (dark pink reaction, Figure 18). Meanwhile, BaAlH5, BaAl4, and BaH2 were produced by the hydrogenation of Ba4Al5 (green reaction, Figure 18). The further heating of BaAlH5 (black reaction, Figure 18) or high-temperature synthesis from Ba7Al13 (blue reaction, Figure 18) can produce Ba2AlH7 along with some by-products [234,235,236].
Liu et al. reported a clear effect of the initial stoichiometry of the mixture on the product when a mixture of BaH2 and Al was used as the precursor. The 2:1 and 1:1 mixtures directed the product to Ba2AlH7. Meanwhile, a 1:2 mixture directed the product to BaAlH5 [237]. However, none of the mixtures produced a complete reaction.
Liu et al. proposed the following reactions for the decomposition of BaAlH5 and Ba2AlH7 [236]:
5BaAlH5 → Ba2AlH7 + 2BaH2 + BaAl4 + 7H2 (280 °C, Argon)
4BaAlH5 → 3BaH2 + BaAl4 + 7H2 (350 °C, Vacuum)
4Ba2AlH7 → 7BaH2 + BaAl4 + 7H2 (350 °C, Vacuum)
Table 11 and Figure 19 present the crystal structures of the barium-aluminum hydrides. The crystal structure of BaAlH5 is composed of corner-sharing (AlH6) octahedra that form zigzag chains along the crystallographic c axis [207]. Meanwhile, Ba2AlD7 is composed of isolated (AlD6) octahedra and infinite one-dimensional chains of edge-sharing (DBa4) tetrahedra [235].

3.4. Alanates of Transition Metals

The alanates of the transition metals date from the 1950s–1960s. Although most of them have decomposition temperatures too low for hydrogen storage purposes, some of them can be of interest. However, almost all of the reported materials have been poorly characterized. Normally, the old reports did not present the basic characterization of materials, for example, X-ray diffraction or infrared spectroscopy. On the other hand, some of them have only been theoretically discussed. In the following sections, the most important (experimental and/or theoretical) characteristics of this family of alanates are presented.

3.4.1. Scandium Alanate

Charkin et al., in a theoretical study, proposed the decomposition of a hypothetical Sc(AlH4)3 to provide the following products: a) HSc(AlH4)2 + AlH3, b) H2Sc(AlH4) + 2AlH3, or c) ScH3 + 3AlH3. This latter route will give the highest dissociation energy [238]. Sc(AlH4)3 deserves more research to estimate, for example, formation energy or crystal structure. Experimentally, Sc(AlH4)3 has not been synthesized, despite a possible and interesting 8.7 wt.% of hydrogen content.

3.4.2. Yttrium Alanate

Y(AlH4)3 was first described by Kost et al. in 1978 [239]. Later, in 2017, Cao et al. demonstrated the partial reversibility of hydrogen storage [240]. Y(AlH4)3 has a hydrogen content of 6.6 wt.% and it can be produced by the metathesis reaction between 3LiAlH4 + YCl3 [240].
Kost et al. reported the beginning of decomposition of Y(AlH4)3 at 50 °C [239]. However, they did not present additional details. Cao et al., based on different characterization techniques, proposed that the decomposition of Y(AlH4)3 occurs as:
Y(AlH4)3 → YAlH6 + 2Al +3H2 (80–170 °C)
YAlH6 → YH3 + Al + 1.5H2 (170–250 °C)
YH3 → YH2 + 0.5H2 (250–350 °C)
YH2 + 3Al → YAl3 + H2 (>350 °C)
In reaction (63), at 140 °C, 3.4 wt.% of hydrogen was released. 2.6 wt.% of hydrogen was re-adsorbed at 145 °C and 100 bar. However, no direct hydrogenation from YH3 +Al at 145 °C and 100 bar occurred [240]. Y(AlH4)3, and YAlH6 are reported as amorphous materials [240]. However, no direct evidence of YAlH6 was presented [240]; thus, further characterizations of these materials are needed.

3.4.3. Titanium Alanate

Wiberg et al. reported the formation of Ti(AlH4)4 (11.1 wt.% hydrogen content) in 1951 [14]. The synthetic route was the metathesis reaction between TiCl4 and LiAlH4 in ether at −110 °C [14]. Later, in 1975, Kost et al. reported a similar synthesis while using LiAlH4 and TiBr4 or TiCl4. The product was separated from the solution in a filter cooled with dry ice [241]. The reported stoichiometries indicted that the metathesis reaction was not completed or that partial substitution of Cl by [AlH4] was achieved [241]. Wiberg reported that Ti(AlH4)4 was decomposed at −85 °C [14]; for its part, Kost reported the evolution of “two g-atom of H per g-atom of Ti” at −70 °C [241]. The decomposition of Ti(AlH4)4 was proposed as [241]:
Ti(AlH4)4 → TiH2 + 4AlH3 + H2.
Further decomposition of AlH3 was observed at 110 °C [241]. No more characteristics of this material have been reported. However, Ti(AlH4)4 can be a very interesting material in regards to its hydrogen content, perhaps tailoring the dehydrogenation temperature with some structural or chemical modification could be explored. Another point to discuss is that Ti can work in other oxidation states besides Ti4+; for example, Ti3+ or Ti2+. The Ti3+ and Ti2+ compounds are generally more stable than the Ti4+ compounds, i.e., the liquid and volatile TiCl4 versus solid TiF3 or TiCl2. Alternatively, other compositions of Ti-Al alloys or intermetallics could be explored. For example, Ramzan et al. explored employing DFT calculations, the structural stability, and other properties of Ti4AlH3 and Ti3AlH2 phases [242]. Maeland et al., some time ago, reported the reversible hydrogenation of Ti3Al at 9.2 bar of deuterium pressure and 200 °C to form Ti3AlDx (x = 5.9–8) [243].

3.4.4. Zirconium Alanate

The first report on Zr(AlH4)4 was the work of Reid et al. in 1957 [13]. Zr(AlH4)4 (7.49 wt.% hydrogen content) was produced by the metathesis reaction between Zr(BH4)4 and LiAlH4 in ether solution and He atmosphere [13]. Zr(BH4)4 was formerly prepared by metathesis of LiBH4 and ZrCl4 [13]. In 2008, Zr(AlH4)4 was produced by the reaction between LiAlH4 and ZrCl4 in ether solution [244]. No clear indication of the reaction temperature was found in this work. No reports regarding the characteristics of dehydrogenation or on the characterization of this material were found. Other compositions of the Zr-Al-H system deserve further research; for example, Matsubara et al. achieved the hydrogenation of the intermetallic Zr3Al to give Zr3AlH4 [245].

3.4.5. Vanadium Alanate

Charkin et al. also proposed the decomposition of a hypothetical V(AlH4)3 to provide the following products: (a) HV(AlH4)2 + AlH3, (b)H2V(AlH4) + 2AlH3, or (c) VH3 + 3AlH3 [238]. Experimental confirmation of the existence of V(AlH4)3 is missing.

3.4.6. Niobium Alanates

Wiberg et al., in 1965, reported the reaction between NbCl5 and LiAlH4 in several proportions and temperatures in ether at low temperature [246]. Wiberg et al. concluded that the products were a function of the temperature and the excess of LiAlH4 used; the first family of products was [246]:
NbCl 5 + 5 LiAlH 4     Nb ( AlH 4 ) n + ( 5     n ) AlH 3 + 5 n 2 H 2 + 5 LiCl ,
when n = 3.5 at −70 °C the product was Nb2(AlH4)7, for n = 3.0 at −40 °C the product was Nb2(AlH4)6, and for n = 2.5 at 20 °C the product was Nb2(AlH4)5.
The other family of products was:
NbCl 5 + ( 5 + m ) LiAlH 4     Li m Nb ( AlH 4 ) n + m + ( 5     n ) AlH 3 + 5 n 2 H 2 + 5 LiCl ,
LiNb2(AlH4)7 was formed at −70 °C; meanwhile, LiNb2(AlH4)5 and LiNb(AlH4)3 were formed at 25 °C [246]. Wiberg et al. wonderfully described the synthesis procedure and the changes in the color that are associated with each Nb or LiNb- alanates. However, a detailed characterization is needed, particularly the characterization of the material obtained at room temperature Nb2(AlH4)5 (5.9 wt.% hydrogen content).

3.4.7. Tantalum Alanates

TaH2(AlH4)2 was reported by Kost et al. in 1978 [239]. The compound has a hydrogen content of 4.11 wt.%. It was produced in cold ether by the reaction between LiH, Al and a metal halide. Kost et al. reported that TaH2(AlH4)2 is a red powder that decomposes at 130 °C. TaH2(AlH4)2 and AlH3 are the decomposition products of a very unstable Ta(AlH4)n [239].

3.4.8. Manganese Alanate

The reports on Mn(AlH4)2 are rather diffuse, as in the case of Be(AlH4)2. The first compilation where Mn(AlH4)2 appeared, is the book of Mackay [187]. In that book, Mn(AlH4)2 was reported to be prepared from a halide complex (no mention of which halide) and LiAlH4 in Et2O, and to decompose at 25 °C. The book refers, in turn, to two reports of Monnier et al. [247,248]. No further reports on Mn(AlH4)2 were found. Mn(AlH4)2 would have a hydrogen content of 6.89 wt.%.

3.4.9. Iron Alanate

Fe(AlH4)2 can be an interesting material for hydrogen storage, due to the 6.84 wt.% of hydrogen content. However, contradictory reports on the decomposition temperature are published. In favor of the near-room temperature stability of Fe(AlH4)2 is the report of Neumaier et al. [249]. Fe(AlH4)2 was prepared by means of metathesis of FeCl3 + 3LiAlH4 in ether at low temperature (−116 °C) [249]. Once formed, the iron easily decomposed. Neumaier et al. presented a p-T diagram of the decomposition reaction; around 20 °C a continuous partial decomposition was observed. Meanwhile, a fast decomposition was observed at 90–100 °C. Two comments can be mentioned: (1) The quantity of released hydrogen was not reported despite a detailed thermolysis study being presented. (2) The fast decomposition at 90–100 °C is near to the temperature of α-, and α’-alane decomposition [186], which is one by-product of iron alanate formation. This leave doubts about who is decomposing Fe(AlH4)2 or AlH3. Despite that, Neumaier et al. considered Fe(AlH4)2 to be stable at room temperature. The proposed reactions of formation and decomposition are [249]:
FeCl3 + 3LiAlH4 → Fe(AlH4)2 +AlH3 + 3LiCl +0.5H2,
Fe(AlH4)2 → Fe + 2Al + 4H2
Against the near-room temperature stability of Fe(AlH4)2 is the report of Schaeffer et al. [250]. They also produced Fe(AlH4)2 by means of metathesis of FeCl3 and an excess of LiAlH4. However, Schaeffer et al. considered Fe(AlH4)2 to be unstable at room temperature.

3.4.10. Copper Alanate

CuAlH4 (4.2 wt.% hydrogen content) was reported as a product of the reaction between CuI and LiAlH4 in ether at −78 °C by Ashby et al. [251]. CuAlH4 is unstable and it reacts quickly, with the proposed product being Cu3AlH6 [251]:
CuAlH4 → CuH + AlH3
2CuH + CuAlH4 → Cu3AlH6
Both of the alanates decomposed with a slight heating. Wiberg et al. reported that the reaction between CuI and LiAlH4 in pyridine at room temperature did not produce Cu-alanates; it produced LiI, AlI3, and CuH [252].

3.4.11. Silver Alanate

AgAlH4 (2.9 wt.% hydrogen content) was produced by the following reaction in ether at −80 °C [253]:
AgClO4 + LiAlH4 → AgAlH4 +LiClO4
AgAlH4 decomposed at −50 °C to the elements Ag, Al, and H2 [253].

3.4.12. Zinc Alanate

Zhizhin et al. (and references wherein) summarized the production of ZnH2; one of the reactions is [254]:
2LiAlH4 + ZnI2 → 2LiI + 2AlH3 + ZnH2
However, depending on the reaction conditions (solvent composition, mainly), admixtures of Zn-AlH4 can be present:
LiAlH 4   ZnI 2   Zn ( AlH 4 ) 2   ZnI 2 , LiAlH 4   Zn [ ZnI 2 ( AlH 4 ) 2 ]   ZnI 2 , LiAlH 4   Zn [ ZnI 3 ( AlH 4 ) ]

3.5. Alanates of the Main Group

As in the case of transition metals alanates, the alanates of the main group elements are scarce, with most of them being unstable, even at low temperatures.

3.5.1. Gallium Alanate

Ga(AlH4)3 (7.4 wt.% hydrogen content) was produced by the reaction between GaCl3 and LiAlH4 in ether at 0 °C [255,256]:
GaCl3 + 3LiAlH4 → Ga(AlH4)3 + 3LiCl
However, at 35 °C, the Ga(AlH4)3 decomposes into GaH3 and AlH3 [256].

3.5.2. Indium Alanate

In(AlH4)3 (5.8 wt.% hydrogen content) was produced by the reaction between InCl3 and LiAlH4 in ether at −70 °C [256]:
InCl3 + 3LiAlH4 → In(AlH4)3 + 3LiCl
The In(AlH4)3 decomposed at −40 °C. However, in a similar reaction at room temperature:
InCl3 + LiAlH4 → InCl2(AlH4) + LiCl,
the product InCl2(AlH4) is stable up to 100 °C [256].

3.5.3. Thallium Alanate

The synthesis of TlAlH4 was reported in 1967, the reaction was performed in ether at −100 °C [257,258]:
TlClO4 + LiAlH4 → TlAlH4 + LiClO4.
TlAlH4 decomposed at −80 °C (1.9 wt.% hydrogen content). Wiberg et al. tried the metathesis reaction between TiCl3 and LiAlH4 in ether at −115 °C [258]. However, no Tl+3-alanate could be isolated from the reaction of TlCl3 and LiAlH4, with the product spontaneously decomposing at −110 °C [258,259]. A marginal stabilization was achieved when a Cl substituted an [AlH4] ion: TlCl(AlH4)2 was produced by the reaction between TlCl3 and AlH3 in ether at −115 °C in the presence of AlH3·AlCl3 [259]. TlCl(AlH4)2 decomposed at −95 °C [259].

3.5.4. Tin Alanate

Sn(AlH4)4 (6.6 wt.% hydrogen content) was produced by the reaction between SnCl4 and LiAlH4 in ether at −80 °C [260]:
SnCl4 + 4LiAlH4 → Sn(AlH4)4 + 4LiCl.
Sn(AlH4)4 decompose at −40 °C. The decomposition products are Sn, Al and H2.

3.6. Alanates of Lanthanides and Actinides

3.6.1. Lanthanum, Cerium, Praseodymium and Neodymium Alanates

La, Ce, Pr, and Nd alanates were produced by metathesis that was assisted by mechanical milling of the corresponding trichlorides and NaAlH4 (in excess 1:3) under hydrogen pressure (1–15 bar) [261]. The expected products, M(AlH4)3, M = La, Ce, Pr, and Nd, are unstable and decompose during ball milling. Instead of M(AlH4)3, alumino-hydrides of stoichiometry MAlxHy were obtained (very close to MAlH6 stoichiometry). Thermolysis of the MAlH6 (M = Ce, Pr, and Nd) materials demonstrated two-steps of decomposition, except for LaAlH6 [261]. The first step is associated with the decomposition of the alanate. Meanwhile, the second step can be associated with the decomposition of the corresponding metal hydride and the formation of M-Al alloys. Although the decomposition pathway was proposed for Nd, based on the in-situ X-ray diffraction data that were presented by Weidenthaler et al., the reaction can be extrapolated for Ce and Pr [261]:
NdAlH6 → NdH3 + Al + 3/2H2
NdH3 +4Al → NdAl4 +3/2H2
Table 12 summarizes the hydrogen content, hydrogen released, decomposition temperatures, and crystal structure data [261]. The experimental X-ray diffraction patterns of MAlxHy were compared to the DFT calculations of hypothetical MAlH6 materials. Figure 20 presents the expected X-ray diffraction patterns and the structures.

3.6.2. Europium Alanate

Eu(AlH4)2 was produced by the metathesis reaction of EuCl2 + 2NaAlH4 or EuCl3 + 3NaAlH4. The reaction was performed by means of mechanical milling in a hydrogen atmosphere (1–15 bar) and different milling times (180 min seems enough time) [44]. Independently of the initial oxidation state of Eu ion, Eu2+, or Eu3+, the final alanate was Eu2+, i.e., Eu(AlH4)2. Additionally, NaEu2Cl6 was observed as an intermediary. Eu(AlH4)2 has a hydrogen content of 3.76 wt.%. Pommerin et al. demonstrated a hydrogen release of about 1.8 wt.% (including the NaCl load) in two steps [44]. The first step occurred at about 100–125 °C with the formation of EuAlH5. The second step occurred at about 200–225 °C. Further heating led to the formation of EuAl4. Rehydrogenation was achieved by milling at high hydrogen pressure (50, 200, or 300 bar). Unfortunately, the rehydrogenation was not achieved under 1000 bar of static H2 pressure; i.e., the temperature of rehydrogenation was not clearly indicated without milling. Further dehydrogenation demonstrated that the two-step reactions and temperature range are kept. However, a drastic reduction of the hydrogen release was found (about 0.8 wt.%) [44]. Partial crystallographic information was reported, i.e., no H position was determined (Table 13) [44]. Figure 20 presents the expected X-ray diffraction patterns and structures.

3.6.3. Ytterbium Alanate

Yb(AlH4)2 was reported by Kost et al. in 1978 [239]. The compound has a hydrogen content of 3.43 wt.%. It was produced in cold ether by the metathesis reaction between LiH, Al, and a metal halide. Kost et al. reported that Yb(AlH4)2 is a yellow powder that decomposes at 70 °C. The decomposition products of Yb(AlH4)2 are the hydrides of Al and Yb [239]. The YbH2 is metastable at room temperature [262].

3.6.4. Thorium-Aluminum Hydride

No records of thorium alanate were found; however, an intermetallic hydride of Th was found: Th2AlH4. The thorium-aluminum hydride can be easily obtained by the hydrogenation of the intermetallic Th2Al [263]. Th2Al needs activation at 450 °C in vacuum, followed by deuterium absorption at 0.15 bar and iced-water cooling [264]. The products were Th2AlDx, x = 3.9 ± 0.1, 2.7, and 2.3 [264]. Experimental and theoretical crystal structure of Th2AlH4 reasonably agreed on a I4/mcm space group with lattice parameters a = 7.626 Å, and c = 6.515 Å, and atomic positions Th (0.1656, 0.6656, 0), Al (0, 0, 0.25), and H (0.377, 0.8707, 0.1512) [265].

4. Cation-Mixed Alanates

Cation substitution has demonstrated utility in the tailoring of the thermodynamic and kinetic properties in borohydrides [22,266]. A similar approach has been applied to alanates, for which LiAlH4 or NaAlH4 are frequently used as starting materials due to their reactivity. These alanates react with other metal hydrides to form mixed cation alanates. The reactions can be generalized as [267]:
M’AlH4 + 2MH → M2M’AlH6,
MAlH4 + MH + M´H → M2M´AlH6, M≠M´.
Theoretical calculations had predicted the stability of alanates, such as LiNa2AlH6, K2LiAlH6, K2NaAlH6, K2.5Na0.5AlH6, LiMgAlH6, LiCaAlH6, NaCaAlH6, and KCaAlH6 [268,269]. Some of them have been successfully synthesized, as presented below.

4.1. Li-Na Alanates

Na2LiAlH6 can be obtained by the reaction of 2NaH and LiAlH4 in an organic solvent [270], in the solid-state at very high hydrogen pressure [270], by means of a mechanically activated reaction between NaH, LiH, and NaAlH4 (Equation (86)) [271], or 2NaH + LiAlH4 [272,273,274], or NaH + LiAlH4 [275], or 2NaAlH4 + LiH [276], or by the reactive mechanical milling of 2NaH+LiH+Al that was catalyzed with TiF3 under 30 bar of hydrogen pressure [277]. Wang et al. produced Na2LiAlH6 by the reaction between 2NaH and LiAH4 [274]. However, detailed study of the synthesis reaction pathway by X-ray diffraction demonstrated that, during mechanical milling, a metathesis reaction occurred to produce a mixture of LiH, NaAlH4, and residual NaH, i.e., the same reactants of Equation (86).
NaH + LiH + NaAlH4 → Na2LiAlH6.
LiNa2AlH6 was also observed during the electrochemical decomposition of NaAlH6 in the presence of Li [278]:
NaAlH4 + 3/2 Li → ½ Na2LiAlH6 + ½ Al + LiH.
Na2LiAlH6 has a total hydrogen content of 7.03 wt.% and a theoretical reversible hydrogen storage of 3.51 wt.% (Equation (88)). Na2LiAlH6 has demonstrated reversibility (Equation (88)) [274,277], which is enhanced by the use of a catalysts, such as TiF3, TiFe3, TiCl3, CeO2, ZrCl4, TiBr4, CrCl3, AlCl3, TiO2, Y2O3, or MnCl2 [276,277,279,280].
Na2LiAlH6 ↔ 2NaH + LiH + Al + 3/2 H2.
Dehydrogenation reaction (Equation (88)) without additives occurs between 190–250 °C and it releases about 3.35 wt.%. Further reactions involve NaH decomposition at 320–380 °C and finally the reaction of LiH with Al at 380–480 °C, with the formation of LiAl and H2. Wang et al. demonstrated a release of 6.73 wt.% and a re-hydrogenation level of 6.6 wt.% when heating up to 530 °C under vacuum, and 285 °C and 135 bar, respectively. Small amounts of Na3AlH6 have been observed during the dehydrogenation of Na2LiAlH6 [274,281]. Additives, such as TiF3, resulted in a low-temperature beginning of Na2LiAlH6 decomposition (~50 °C) [276,277,280]. Additionally, Al3Ti was found after dehydrogenation when Na2LiAlH6 is mixed with TiF3 [279,280].
First principle studies (before experimentation, i.e., synthesis and crystal structure determination) indicated that Na2LiAlH6 would have P 21/n [282] or P 21/c [283] symmetry, which is very close to Fm-3m symmetry [283]. Brinks et al. determined the group symmetry of Na2LiAlD6 as Fm-3m. This material consists of corning-sharing (AlD6) and (LiD6) octahedra, where each octahedron is surrounded by six octahedra (Table 14 and Figure 21) [284]. The deuterated Na2LiAlD6 was produced by the ball milling of NaAlD4 and LiAlD4 [284].
The research group of Prof. Q. Wang performed a complete study regarding the determination of the (p, T) equilibrium of reaction (88) with and without TiF4 as a catalyst (Figure 22) [274,285]. The results indicate that the catalyst moves to higher pressure the equilibrium towards Na2LiAlH6 formation at a given temperature; or conversely a reduction of the equilibrium temperature at a given pressure. Fonneløp et al. revealed that the addition of 10 mol% of TiF3 to Na2LiAlH6 induced hydrogen release at temperatures as low as 50 °C [281]. In such a case, the dehydrogenation pathway changes from a one-step process (Equation (88)) to a two-step process, with the formation of Na3AlH6 as the intermediary. Between 50–180 °C, the decomposition reaction was described as:
Na2LiAlH6 → 2/3 Na3AlH6 + LiH + 1/3 Al + 1/2 H2.
Further heating (180–225 °C) leads to the usual decomposition reaction of Na3AlH6.
Finally, the other possible combination of Li, Na, Al, and H would be as Li2NaAlH6. However, attempts to synthesize this material have been unsuccessful. The attempts involve the synthesis in organic solvents, such as Me2O (160 °C, 12 h), or by ball-milling [267]. As proposed by Santhanam et al. [169], Li2NaAlH6 is not formed at all under the tested conditions, or it disproportionates Na2LiAlH6, LiH and LiAlH4.

4.2. Li-K Alanates

K2LiAlH6 was reported in 2005 by Graetz et al. [267]. K2LiAlH6 was produced by the ball-milling of 2KH + LiAlH4 [267]. Graetz et al. determined an Fm-3m structure for K2LiAlH6. However, in their paper, they recognized that the diffraction pattern was not suitable for Rietveld analysis [267]. Briefly, after that, Rönnebro et al. performed the mechanical milling of the same precursors followed by a heating treatment of the pelletized sample at 320–330 °C and 700 bar for 1–2 days. By doing this, K2LiAlH6 was crystallized, and its crystal structure was determined to have R3m symmetry (Table 15) [286]. As in the case of Na2LiAlH6, theoretical calculations (predating synthesis and crystal structure determination) predicted that K2LiAlH6 would have P 21/n symmetry (Table 15 and Figure 23) [282,283]. The differences between the calculated and the experimental data could be related to the temperature of calculation (0 K) versus the temperature of synthesis and testing (near room temperature).
K2LiAlH6 has a total hydrogen content of 5.11 wt.% and a possible reversible hydrogen storage of 2.56 wt.%. The dehydrogenation of K2LiAlH6 was performed at 227 °C, while rehydrogenation was performed at 300 °C and up to 10 bar [267]. The rehydrogenation achieved 2.3 wt.% hydrogen storage, i.e., approximately 90% of the theoretical value. However, the reaction time was very long, around 280 h; and, perhaps a higher hydrogenation pressure would improve kinetics.
Regarding other Li-K alanates and similar to the Li2NaAlH6 case, no Li2KAlH6 has been produced so far [169].

4.3. Li-Mg Alanates

The mixed alanate LiMg(AlH4)3 has a hydrogen content of 9.7 wt.%; LiMg(AlH4)3 is known since 1979 by the work of Bulychev et al. [287]. It can be produced by the metathesis reaction between LiAlH4 and MgCl2 [165,220,288]:
3LiAlH4 + MgCl2 → LiMg(AlH4)3 + 2LiCl.
Reaction (90) can be performed in an organic solvent or assisted by mechanical milling. The decomposition of LiMg(AlH4)3 is a two-step process [289,290]:
LiMg(AlH4)3 → LiMgAlH6 + 2Al + 3H2 (100–130 °C)
LiMgAlH6 → LiH + MgH2 + Al + 3/2 H2 (150–180 °C)
The addition of graphitic nanofibers can reduce the dehydrogenation temperatures [291]. Addition of TiF3 leads to the decomposition of the mixed alanate even during ball-milling [290]. Attempts of re-hydrogenation were unsuccessful, even at high pressures [289,290]. The structure of LiMg(AlH4)3 consists of a corner-sharing network of alternating [AlH4] tetrahedra and (LiH6) and (MgH6) octahedra (Table 16 and Figure 24) [288]. The structure of LiMgAlH6 consists of alternating AlMg3 and Al2Li3 layers; in the Al2Li3 layer, the [AlH6] octahedra share edges with three (LiD6) octahedra [206,290].

4.4. Li-Ca Alanates

LiCa(AlH4)3 has a total hydrogen content of 8.6 wt.%; thus, it appears as a very attractive hydrogen storage material. LiCa(AlH4)3 was produced by the metathesis reaction between LiAlH4 and CaCl2, utilizing mechanical milling [292]:
3LiAlH4 + CaCl2 → LiCa(AlH4)3 + 2LiCl.
LiCa(AlH4)3 (plus LiCl) starts decomposing at 120 °C and it ends at about 180 °C. Liu et al. proposed the formation of LiCaAlH6 in the first dehydrogenation step [292]. In the second step (180–300 °C), LiCaAlH6 decomposed to form Al, CaH2, and LiH. The two steps released 6 wt.% of hydrogen [292]:
LiCa(AlH4)3 → LiCaAlH6 + 2Al + 3H2
LiCaAlH6 → CaH2 + LiH + Al + 3/2 H2
In the second step, some CaH2−xClx was detected. No information regarding possible re-hydrogenation was found. The crystal structure of LiCa(AlH4)3 was experimentally determined as the space group P63/m (Table 17 and Figure 25) [292]. Theoretical research confirmed this symmetry and contributed to determining the hydrogen atomic positions (Table 17) [293]. The complete crystal structure of LiCaAlH6 was predicted from the theoretical calculations [294].

4.5. Na-K Alanates

K2NaAlH6 is the only reported mixed Na-K alanate. This material has a total hydrogen content of 4.46 wt.%. K2NaAlH6 can be produced by the reaction assisted by ball-milling between KH and NaAlH4 in a 2:1 molar relation, with or without hydrogen pressure (10 bar) [295,296]. K2NaAlH6 decomposes into simple hydrides, Al and hydrogen gas at ~352 °C [296,297]:
K2NaAlH6 → 2KH + NaH + Al + 3/2 H2
The addition of TiCl3, TiF3, graphene, or carbon nanotubes slightly reduced the dehydrogenation temperature, with TiF3 being the most effective material [296]. K2NaAlH6 is reported to store hydrogen reversible; however, full capacity was not recovered [295]. K2NaAlH6 is reported as a cubic close-packed structure of isolated [AlH6]3− octahedra; the octahedral interstices are occupied by Na+ ions, while the tetrahedral interstices are filled with K+ ions (Table 18, Figure 26) [295].

5. Anion Substitution

Ion size and oxidation state make, in principle, F ions suitable for substituting H ions in some hydrogen storage compounds, such as hydrides [298], borohydrides, or alanates [299]. The substitution could tune the thermodynamics, with the goal being to reduce the dehydrogenation temperature [299]. Perhaps the clearest example of this is the production of Na3AlH6−xFx from NaF and Al [300]. However, despite reducing the enthalpy of the first dehydrogenation, the reversibility of the system was compromised [300]. Other examples of anion substitution, despite being less studied, included K3AlH6−xFx [301] and CaAlFxH5−x [219]. Unfortunately, limited information regarding these systems can be found, thus experimental and/or theoretical studies should be performed in the future.

6. Techniques of Characterization of Alanates

The most common physicochemical characterization techniques for hydrogen storage materials, and thus alanates, are X-ray diffraction (in-situ, ex-situ, with synchrotron or conventional X-ray sources), and spectroscopies, such as Infrared and Raman. Other vibrational spectroscopy techniques, such as Inelastic Neutron Scattering (INS), Nuclear Resonant Inelastic X-ray Scattering Spectroscopy (NRIXS), or Photoacoustic (PA) Infrared Spectroscopy are far less widespread. The main results of X-ray diffraction studies were presented along with the description of each alanate. Thus, we did not include a special section for it. On the other hand, the characterization of alanates by IR and Raman Spectroscopies is also frequently used due to the relatively low cost of equipment and the relative simplicity of sample preparation for such tests. Therefore, we present IR and Raman spectroscopies in this review.

Fourier Transformed Infrared Spectroscopy (IR) and Raman Spectroscopy

Vibrational transitions can be observed as infrared or Raman spectra. Although frequently, these two techniques are complementary, their physical origins are different [302]. IR absorption spectra originate from photons in the infrared region that are absorbed by transitions between two vibrational levels of the molecule in the electronic ground state. Raman spectra have their origin in the electronic polarization that is caused by ultraviolet, visible, and near-IR light [302]. The observed vibration modes depend on factors, such as the molecular symmetry, identity of atoms, and bond energies, i.e., the kinetic and potential energies of the system. The kinetic energy is determined by the masses of the individual atoms and their geometrical arrangement in the molecule. On the other hand, the potential energy arises from the interaction between the individual atoms and it is described in terms of the force constants [302]. For the alanates, the common structures are the tetrahedral [AlH4] and octahedral [AlH6]3− units. Figure 27 illustrates the four normal modes of vibration of a tetrahedral [AlH4]. All four vibrations are Raman-active, whereas only ν3 and ν4 are infrared active [302]. Octahedral molecules have six normal modes of vibration; of these, vibrations ν1, ν2, and ν5 are Raman-active, whereas only ν3 and ν4 are infrared-active (Figure 28) [302].
The vibrational spectra of alanates are frequently classified as external and internal. The external vibrations are due to the vibration of the whole crystal structure. Meanwhile, the internal vibrations are due to the [AlH4] ion, which has four active vibrational modes in Raman and only two in infrared [303]. Some of these features are shared with other materials of similar structure, for example, the borohydrides [304]. The infrared active modes of the [AlH4] ion are the asymmetric stretching modes in the region 1600–2000 cm−1 and the bending modes in the region 700–900 cm−1 [305]. Some representative data are collected in Table 19 and Table 20. As a generally accepted trend of infrared vibrations in the alanates of group 1, the stretching modes, in wavenumbers, roughly decrease with increasing mass of the cation [306]. Meanwhile, the bending modes are unaffected by the counter-ion [305,306]. Other correlations between the stretching and bending peaks (or regions) versus ionization energy, electronegativities, or bond distance have been proposed [302]. Indeed, we tried to find correlations with these parameters. However, we obtained the best results by using the difference in the electronegativities between Al and the counter-cation or the counter-cation ion size. In Figure 29 we present a correlation between the most intense stretching and bending IR peak of MAlH4 (M = group 1 metals, [AlH4] tetrahedra) versus the difference in electronegativities of Al and the metal. The electronegativity scale was the Allred–Rochow [307]. The IR data that were obtained by Adicks et al. in pure crystalline materials [177] were complemented by data published in several experimental and theoretical reports compiled in this review [167,308,309,310,311,312,313,314,315,316,317,318,319,320]. The data reflects the significant dispersion of results. The NaAlH4 data are the most common and particularly disperse, which is probably due to the diversity in the material history, such as milling, doping, or cycling [318]. The quantity of available IR data on K, Rb, and Cs- alanates is rather scarce. Still, some tendencies were found; there is a bell-shape dispersion of the Al-H stretching frequency (most intense peak) versus the difference of electronegativity between Al and the group 1 metal. Meanwhile, there is an almost linear increase of the Al-H bending frequency (most intense peak). This can be related to the changes in the geometry of the alanates, along with the group.
The octahedral ion [AlH6]3− that is present in the so-called intermediaries of alanates also shows infrared and Raman active modes. From the 15 normal vibration modes of a group with octahedral symmetry, two modes are active in the infrared, and three modes are active in the Raman [321]. In Figure 30, we present a correlation between the stretching IR most intense peak of M3AlH6 (M = group 1, [AlH6]3− octahedra) versus the effective ionic radii [307]. The available data for the so-called intermediaries of alanates of group 1 (M3AlH6) are scarcer than for the tetrahedral alanates, i.e., MAlH4. Thus, the correlation was constructed with data of Li, Na, and K [167,307,318,322,323,324,325,326,327]. The red dots of Rb3AlH6 and Cs3AlH6 are an extrapolation based on the fitted curve.
In Figure 31, we present a correlation between the stretching and bending Raman most intense peak versus the difference in electronegativities between Al and the metal of MAlH4 (M = group 1). In general, there are less Raman data available than IR data. In both stretching and bending Raman modes, the correlation with the difference in electronegativity is not linear. The reported data were found only for Li, Na, and K-alanates. Thus, the Rb and Cs-alanates data are an extrapolation, pending future reports to corroborate this forecast.
Not enough IR or Raman data are available for group 2 (apart from Mg and Ca) and the rest of alanates of the periodic table. Additionally to the Figure 29, Figure 30 and Figure 31, an attempt to find trends that include the double-metal alanates of groups 1 and 2 was performed; no clear trends were found. This can open the possibility of theoretical and experimental studies to obtain these missing data and to obtain general rules that correlate structure and spectroscopic properties.

7. Thermodynamics

A dehydrogenation enthalpy of about 40 kJ/mol is required in order to meet the dehydrogenation temperature compatible with PEMFCs [328]. This enthalpy value roughly means an equilibrium pressure of 1 bar at room temperature. The equilibrium pressure is a function of the temperature, the dehydrogenation enthalpy, and entropy. It is described by the Van’t Hoff equation [328]:
ln ( p e q p e q 0 ) = Δ H R × 1 T Δ S R
ΔS mostly corresponds to the change from molecular hydrogen gas to dissolved solid hydrogen [328]. It amounts approximately to the standard entropy of hydrogen (130 J·K−1mol−1) and is, therefore, frequently taken as a constant for all metal-hydrogen systems [328]. ΔH must be the dehydrogenation reaction enthalpy and each material must report it. However, the formation enthalpy is sometimes used instead, particularly if the material is a metal and its hydride. The enthalpies of formation and dehydrogenation have been related, directly or indirectly, to the bond energy, i.e., the stability of the compound [328,329]. The reported dehydrogenation enthalpies were used to construct the phase diagrams that are presented in this review. The representative values are condensed in Table 21, altogether with formation enthalpies and the activation energies. The thermodynamic data is concentrated mainly in the alanates of group 1, a lot of data is missing on other alanates. The calculated and experimental data of formation enthalpy and dehydrogenation enthalpy show good correlation. However, an in-deep comment on the dispersion of the thermodynamic data is needed. Along with the several consulted papers, different experimental techniques and conditions were used to determine the thermodynamic data. The most used techniques are the differential scanning calorimetry (with variations, such as high-pressure, with or without hydrogen flow, different of values of flows, etc.), pressure-composition isotherms, and theoretical calculations (different levels of theory, programs, basis sets, etc.). Thus, the natural result is the dispersion of data. Perhaps, a standard method will be advisable. Meanwhile, the activation energies present the most disperse values, which is due to the additive and the history of the materials (mechanical milling, purification, recrystallization, cycling, etc.). Additionally, some of the original data are explicitly related to the released mol of H2, meanwhile, other data is not clearly reported of mol of which compound is related.

8. Conclusions and Perspectives

NaAlH4 and KAlH4 stand out among all of the alanates due to their acceptable hydrogen content and reversibility. Perhaps for light-duty vehicles applications, an option will be the NaAlH4, where the catalyst performance is essential. In that subject, along with the consulted papers, the Ti-based catalyst could be limited in the long-term because of the progressive change in the oxidation state of Ti, associated with the decay of performance. Perhaps, lanthanide-metals compounds could be the solution. However, more research on extensive cycling must be done: There is not enough data up to now on the long-term performance of Ce-catalysts on NaAlH4. On the other hand, KAlH4 can be suitable for niche applications where the high-temperature dehydrogenation is not an issue. However, there is no data regarding extensive cycling.
During the preparation of this review, the compilation of alanates beyond the group 1 and 2 was a good surprise. Many of them have a reasonable good dehydrogenation temperature and hydrogen content. Others can be viewed just as a chemical curiosity. In general, the reports of the alanates of transition metals and main group are very old. Perhaps, re-visiting and updating the information of these alanates with new synthesis and characterization techniques could provide new approaches for solving the hydrogen storage problem.
Despite that the formation of reactive composite materials has proven useful in other hydrogen storage materials, this approach seems not so useful in the alanate family. However, the formation of double cation alanates seems to be attractive for improving the dehydrogenation temperature without the sacrifice of the hydrogen content. The anion substitution is explored to a limited extent in the alanates family, and this modification should be studied deeply.

Author Contributions

Conceptualization, writing—review and editing, K.S.-A.; Crystal structures, X-ray diffraction and Infrared data compilement, J.R.T.-G.; Thermodinamic data and figures, R.G.-O.

Funding

This research was funded by CONACyT, Ciencia Básica 251347-ALANATOS CONVENCIONALES Y NO CONVENCIONALES PARA ALMACENAMIENTO DE HIDROGENO, grant number 251347.

Acknowledgments

All authors are very gratefully to Teresa Vásquez Mejía for the support during bibliographic compilation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sandrock, G.A. Panoramic overview of hydrogen storage alloys from a gas reaction point of view. J. Alloys Compd. 1999, 293–295, 877–888. [Google Scholar] [CrossRef]
  2. Wiswall, R.H., Jr.; Reilly, J.J. Metal Hydrides for Energy Storage. Office of Scientific & Technical Information Technical Reports, Brookhaven National Laboratory, Upton, N.Y. 1972. Available online: https://www.osti.gov/servlets/purl/6051964 (accessed on 2 July 2019).
  3. Vigeholm, B.; Kjøller, J.; Larsen, B. Magnesium for hydrogen storage. J. Less Common Met. 1980, 74, 341–350. [Google Scholar] [CrossRef]
  4. Zhang, D.L. Processing of advanced materials using high-energy mechanical milling. Prog. Mater. Sci. 2004, 49, 537–560. [Google Scholar] [CrossRef]
  5. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  6. Yu, X.; Tang, T.; Sun, D.; Ouyang, L.; Zhu, O. Recent advances and remaining challenges of nanostructured materials for hydrogen storage applications. Prog. Mater. Sci. 2017, 88, 1–48. [Google Scholar] [CrossRef]
  7. Bogdanović, B.; Schwickardi, M. Ti-doped alkali metal aluminium hydrides as potential novel reversible hydrogen storage materials. J. Alloys Compd. 1997, 253, 1–9. [Google Scholar] [CrossRef]
  8. Gross, K.; Majzoub, E. Direct Synthesis of Catalyzed Hydride Compounds. U.S. Patent No. US2003 0143154 A1, 31 July 2003. [Google Scholar]
  9. Milanese, C.; Garroni, S.; Gennari, F.; Marini, A.; Klassen, T.; Dornheim, M.; Pistidda, C. Solid state hydrogen storage in alanates and alanate-based compounds: A review. Metals 2018, 8, 567. [Google Scholar] [CrossRef]
  10. Na Ranong, C.; Hoehne, M.; Franzen, J.; Hapke, J.; Fieg, G.; Dornheim, M.; Eigen, N.; Bellosta von Colbe, J.M.; Metz, O. Concept, design and manufacture of a prototype hydrogen storage tank based on sodium alanate. Chem. Eng. Technol. 2009, 32, 1154–1163. [Google Scholar] [CrossRef]
  11. Ley, M.B.; Meggouh, M.; Moury, R.; Peinecke, K.; Felderhoff, M. Development of hydrogen storage tank systems based on complex metal hydrides. Materials 2015, 8, 5891–5921. [Google Scholar] [CrossRef]
  12. Johnson, T.A.; Jorgensen, S.W.; Dedrick, D.E. Performance of a full-scale hydrogen-storage tank based on complex hydrides. Faraday Discuss. 2011, 151, 327–352. [Google Scholar] [CrossRef]
  13. Reid, W.E.; Bish, J.M.; Brenner, A. Electrodeposition of metals from organic solutions III. Preparation and electrolysis of titanium and zirconium compounds in nonaqueous media. J. Electrochem. Soc. 1957, 104, 21–29. [Google Scholar] [CrossRef]
  14. Wiberg, E.; Usón, R. Zur Kenntnis eines Titan-aluminium-wassertsoffs Ti(AlH4)4. Z. Nat. 1951, 6, 392–393. [Google Scholar]
  15. DOE technical targets for onboard hydrogen storage for ligh-duty vehicles. Available online: https://www.energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogen-storage-light-duty-vehicles (accessed on 12 August 2019).
  16. Bergemann, N.; Pistidda, C.; Milanese, C.; Girella, A.; Hanse, B.R.S.; Wurr, J.; Bellosta von Colbe, J.M.; Jepsen, J.; Jensen, T.R.; Marini, A.; et al. NaAlH4 production from waste aluminum by reactive ball milling. Int. J. Hydrogen Energy 2014, 39, 9877–9882. [Google Scholar] [CrossRef]
  17. Guerrero-Ortiz, R.; Tena-Garcia, J.R.; Flores-Jacobo, A.; Suarez-Alcantara, K. From the can to the tank: NaAlH4 from recycled aluminum. Int. J. Hydrogen Energy 2019, 44, 20183–20190. [Google Scholar] [CrossRef]
  18. Bellosta von Colbe, J.; Ares, J.R.; Barale, J.; Baricco, M.; Buckley, C.; Capurso, G.; Gallandat, N.; Grant, D.M.; Guzik, M.N.; Jacob, I.; et al. Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. Int. J. Hydrogen Energy 2019, 44, 7780–7808. [Google Scholar] [CrossRef]
  19. Abdalla, A.M.; Hossaina, S.; Nisfindya, O.B.; Azad, A.T.; Dawood, M.; Azada, A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018, 165, 602–627. [Google Scholar] [CrossRef]
  20. Sanjeev, L. Process for Preparing Dry Sodium Aluminum Hydride. U.S. Patent No. 5,295,581, 22.03, 22 March 1994. [Google Scholar]
  21. Ashby, E.C.; Brendel, G.J.; Redman, H.E. Direct synthesis of complex metal hydrides. Inorg. Chem. 1963, 2, 499–504. [Google Scholar] [CrossRef]
  22. Møller, K.T.; Sheppard, D.; Ravnsbæk, D.B.; Buckley, C.E.; Akiba, E.; Li, H.W.; Jensen, T.J. Complex metal hydrides for hydrogen, thermal and electrochemical energy storage. Energies 2017, 10, 1645. [Google Scholar] [CrossRef]
  23. Finholt, A.E.; Bond, A.C., Jr.; Schlesinger, H.I. Lithium aluminum hydride, aluminum hydride and lithium gallium hydride, and some of their applications in organic and inorganic chemistry. J. Am. Chem. Soc. 1947, 69, 1199–1203. [Google Scholar] [CrossRef]
  24. Barton, L. 1.7.4.2. of Compounds of Aluminum. In Inorganic Reactions and Methods: The Formation of the Bond to Hydrogen (Part 2), Volume 2; Zuckerman, J.J., Hagen, A.P., Eds.; VCH Publishers, Inc.: Oklahoma, OK, USA, 1987; pp. 136–137. [Google Scholar]
  25. Wiberg, E.; Bauer, R. Zur Kenntnis eines Magnesium-aluminium-wasserstoffs Mg(AlH4)2. Z. Nat. 1950, 5, 397–398. [Google Scholar]
  26. Jain, I.P.; Jain, P.; Jain, A. Novel hydrogen storage materials: A review of lightweight complex hydrides. J. Alloys Compd. 2010, 503, 303–339. [Google Scholar] [CrossRef]
  27. Wiberg, E.; Bauer, R. Neues zur Kenntnis des Magnesium-aluminium-wasserstoffs Mg(AlH4)2. Z. Nat. 1952, 7, 131–132. [Google Scholar] [CrossRef]
  28. Schwab, W.; Wintersberger, K. Über Darstellung und Eigenschaften von Calciumaluminiumhydrid Ca(AlH4)2. Z. Nat. 1953, 8, 690–691. [Google Scholar] [CrossRef]
  29. Schwarz, M.; Haiduc, A.; Stil, H.; Paulus, P.; Geerlings, H. The use of complex metal hydrides as hydrogen storage materials: Synthesis and XRD-studies of Ca(AlH4)2 and Mg(AlH4)2. J. Alloys Compd. 2005, 404–406, 762–765. [Google Scholar] [CrossRef]
  30. Fichtner, M.; Fuhr, O. Synthesis and structures of magnesium alanate and two solvent adducts. J. Alloys Compd. 2002, 345, 286–296. [Google Scholar] [CrossRef]
  31. Komiya, K.; Morisaku, N.; Shinzato, Y.; Ikeda, K.; Orimo, S.; Ohki, Y.; Tatsumi, K.; Yukawa, H.; Morinaga, M. Synthesis and dehydrogenation of M(AlH4)2 (M = Mg, Ca). J. Alloys Compd. 2007, 446–447, 237–241. [Google Scholar] [CrossRef]
  32. Bellosta von Colbe, J.M.; Felderhoff, M.; Bogdanović, B.; Schüth, F.; Weidenthaler, C. One-step direct synthesis of a Ti-doped sodium alanate hydrogen storage material. Chem. Commun. 2005, 41, 4732–4734. [Google Scholar] [CrossRef] [PubMed]
  33. Bogdanović, B.; Schwickardi, M. Method for Reversibly Storing Hydrogen on the Basis of Alkali Metals and Aluminum. U.S. Patent No. US 2003/0053948 A1, 20 March 2003. [Google Scholar]
  34. Huot, J.; Ravnsbæk, D.B.; Zhang, J.; Cuevas, F.; Latroche, M.; Jensen, T.R. Mechanochemical synthesis of hydrogen storage materials. Prog. Mater. Sci. 2013, 58, 30–75. [Google Scholar] [CrossRef]
  35. Eigen, N.; Gosch, F.; Dornheim, M.; Klassen, T.; Bormann, R. Improved hydrogen sorption of sodium alanate by optimized processing. J. Alloys Compd. 2008, 465, 310–316. [Google Scholar] [CrossRef]
  36. Hlova, I.; Goldston, J.F.; Gupta, S.; Kobayashi, T.; Pruski, M. A benign synthesis of alane by the composition-controlled mechanochemical reaction of sodium hydride and aluminum chloride. J. Mater. Sci. 2017, 52, 11900–11910. [Google Scholar] [CrossRef] [Green Version]
  37. Hlova, I.; Gupta, S.; Goldston, J.E.; Kobayashi, T.; Pruski, M.; Pecharsky, V.K. Dry mechanochemical synthesis of alane from LiH and AlCl3. Faraday Discuss. 2014, 170, 137–153. [Google Scholar] [CrossRef] [PubMed]
  38. Hlova, I. Mechanochemical Synthesis of Hydrogen-Storage Materials Based on Aluminum, Magnesium and Their Compounds. Ph.D. Thesis, Iowa State University, Ames, IA, USA, 2015. [Google Scholar]
  39. Dymova, T.N.; Mal’Tseva, N.N.; Konoplev, V.N.; Golovanova, A.I.; Aleksandrov, D.P.; Sizareva, A.S. Solid-phase solvate-free formation of magnesium hydroaluminates Mg(AlH4)2 and MgAlH5 upon mechanochemical activation or heating of magnesium hydride and aluminum chloride mixtures. Russ. J. Coord. Chem. 2003, 29, 385–398. [Google Scholar] [CrossRef]
  40. Kabbour, H.; Ahn, C.C.; Hwang, S.-J.; Bowman, C.R.; Graetz, J. Direct synthesis and NMR characterization of calcium alanate. J. Alloys Compd. 2007, 445–447, 264–266. [Google Scholar] [CrossRef]
  41. Iosub, V.; Matsunaga, T.; Tange, K.; Ishikiriyama, M. Direct synthesis of Mg(AlH4)2 and CaAlH5 crystalline compounds by ball milling and their potential as hydrogen storage materials. Int. J. Hydrogen Energy 2009, 34, 906–912. [Google Scholar] [CrossRef]
  42. Sato, T.; Ikeda, K.; Li, H.-W.; Yukawa, H.; Morinaga, M.; Orimo, S. Direct dry syntheses and thermal analyses of a series of aluminum complex hydrides. Mater. Trans. 2009, 50, 182–196. [Google Scholar] [CrossRef]
  43. Kim, Y.; Lee, E.-K.; Shim, J.-H.; Cho, Y.W.; Yoon, K.B. Mechanochemical synthesis and thermal decomposition of Mg(AlH4)2. J. Alloys Compd. 2006, 422, 283–287. [Google Scholar] [CrossRef]
  44. Pommerin, A.; Wosylus, A.; Felderhoff, M.; Schüth, F.; Weidenthaler, C. Synthesis, crystal structures, and hydrogen-storage properties of Eu(AlH4)2 and Sr(AlH4)2 and their decomposition intermediates, EuAlH5 and SrAlH5. Inorg. Chem. 2012, 51, 4143–4150. [Google Scholar] [CrossRef]
  45. Graetz, J.; Reilly, J.J.; Yartys, V.A.; Maehlen, J.P.; Bulychev, B.M.; Antonov, V.E.; Tarasov, B.P.; Gabis, I.E. Aluminum hydride as a hydrogen and energy storage material: Past, present and future. J. Alloys Compd. 2011, 509, S517–S528. [Google Scholar] [CrossRef]
  46. Brower, F.M.; Matzek, N.E.; Reigler, P.F.; Rinn, H.W.; Roberts, C.B.; Schmidt, D.L.; Snover, J.A.; Terada, K. Preparation and properties of aluminum hydride. J. Am. Chem. Soc. 1976, 98, 2450–2454. [Google Scholar] [CrossRef]
  47. Brinks, H.W.; Istad-Lem, A.; Hauback, B.C. Mechanochemical synthesis and crystal structure of α’-AlD3 and α-AlD3. J. Phys. Chem. B 2006, 110, 25833–25837. [Google Scholar] [CrossRef]
  48. Paskevicius, M.; Sheppard, D.A.; Buckley, C.E. Characterization of mechanochemically synthesized alane (AlH3) nanoparticles. J. Alloys Compd. 2009, 487, 370–376. [Google Scholar] [CrossRef]
  49. Gupta, S.; Kobayashi, T.; Hlova, I.Z.; Goldston, J.F.; Pruski, M.; Pecharsky, V.K. Solvent-free mechanochemical synthesis of alane, AlH3: Effect of pressure on the reaction pathway. Green Chem. 2014, 16, 4378–4388. [Google Scholar] [CrossRef]
  50. Martínez-Rodríguez, M.J.; García-Díaz, B.L.; Teprovich, J.A., Jr.; Knight, D.A.; Zidan, R. Advances in the electrochemical regeneration of aluminum hydride. Appl. Phys. A 2012, 106, 545–550. [Google Scholar] [CrossRef]
  51. Zidan, R. Electrochemical Process and Production of Novel Complex Hydrides. U.S. Patent No. US 8,470,156 B2, 25 June 2013. [Google Scholar]
  52. Zidan, R. Enhancing Electrochemical Methods for Producing and Regenerating Alane by Using Electrochemical Catalytic Additive. U.S. patent No. US 9, 850,585 B1, 26 December 2017. [Google Scholar]
  53. Graetz, J.; Reilly, J.J. Thermodynamics of the α, β and γ polymorphs of AlH3. J. Alloys Compd. 2006, 424, 262–265. [Google Scholar] [CrossRef]
  54. Graetz, J.; Hauback, B.C. Recent developments in aluminum-based hydrides for hydrogen storage. MRS Bull. 2013, 38, 473–479. [Google Scholar] [CrossRef]
  55. Graetz, J.; Chaudhuri, S.; Lee, Y.; Vogt, T.; Muckerman, J.T.; Reilly, J.J. Pressure-induced structural and electronic changes in α-AlH3. Phys. Rev. B 2006, 74, 214114. [Google Scholar] [CrossRef]
  56. Wang, L.; Rawal, A.; Aguey-Zinsou, K.F. Hydrogen storage properties of nanoconfined aluminium hydride (AlH3). Chem. Eng. Sci. 2019, 194, 64–70. [Google Scholar] [CrossRef]
  57. Orimo, S.; Nakamori, Y.; Kato, T.; Brown, C.; Jensen, C.M. Intrinsic and mechanical modified thermal stabilities of α-, β- and γ-aluminum trihydries AlH3. Appl. Phys. A 2006, 83, 5–8. [Google Scholar] [CrossRef]
  58. Sandrock, G.; Reilly, J.; Graetz, J.; Zhou, W.-M.; Johnson, J.; Wegrzyn, J. Alkali metal hydride doping of α-AlH3 for enhanced H2 desorption kinetics. J. Alloys Compd. 2006, 421, 185–189. [Google Scholar] [CrossRef]
  59. Graetz, J.; Reilly, J.J. Decomposition kinetics of the AlH3 polymorphs. J. Phys. Chem. B 2005, 47, 22181. [Google Scholar] [CrossRef]
  60. Ikeda, K.; Muto, S.; Tatsumi, K.; Menjo, M.; Kato, S.; Bielmann, M.; Züttel, A.; Jensen, C.M.; Orimo, S. Dehydriding reaction of AlH3: In situ microscopic observations combined with thermal and surface analyses. Nanotechnology 2009, 20, 204004. [Google Scholar] [CrossRef] [PubMed]
  61. Liu, H.; Wang, X.; Dong, Z.; Cao, G.; Liu, Y.; Chen, L.; Yan, M. Dehydriding properties of γ-AlH3. Int. J. Hydrogen Energy 2013, 38, 10851–10865. [Google Scholar] [CrossRef]
  62. Sartori, S.; Opalka, S.M.; Løvvik, O.M.; Guzik, M.N.; Tang, X.; Hauback, B.C. Experimental studies of α-AlD3 and α’-AlD3 versus first-principles modeling of the alane isomorphs. J. Mater. Chem. 2008, 18, 2361–2370. [Google Scholar] [CrossRef]
  63. Sandrock, G.; Reilly, J.; Graetz, J.; Zhou, W.-M.; Johnson, J.; Wegrzyn, J. Accelerated thermal decomposition of AlH3 for hydrogen-fueled vehicles. Appl. Phys. A. 2005, 80, 687–690. [Google Scholar] [CrossRef]
  64. Liu, H.; Wang, X.; Liu, Y.; Dong, Z.; Ge, H.; Li, S.; Yan, M. Hydrogen desorption properties of the MgH2–AlH3 composites. J. Phys. Chem. C 2014, 118, 37–45. [Google Scholar] [CrossRef]
  65. Liu, H.; Wang, X.; Zhou, H.; Gao, S.; Ge, H.; Li, S.; Yan, M. Improved hydrogen desorption properties of LiBH4 by AlH3 addition. Int. J. Hydrogen Energy 2016, 41, 22118–22127. [Google Scholar] [CrossRef]
  66. Goncharenko, I.N.; Glazkov, V.P.; Irodova, A.V.; Somenkov, V.A. Neutron diffraction study of crystal structure and equation of state AID3 up to the pressure of 7.2 GPa. Physica B 1991, 174, 117–120. [Google Scholar] [CrossRef]
  67. Brinks, H.W.; Brown, C.; Jensen, C.M.; Graetz, J.; Reilly, J.J.; Hauback, B.C. Synthesis and crystal structure of β-AlD3. J. Alloys Compd. 2007, 433, 180–183. [Google Scholar] [CrossRef]
  68. Brinks, H.W.; Langley, W.; Jensen, C.M.; Graetz, J.; Reilly, J.J.; Hauback, B.C. The crystal structure of γ-AlD3. J. Alloys Compd. 2007, 441, 364–367. [Google Scholar] [CrossRef]
  69. Kojima, Y.; Kawai, Y.; Haga, T.; Matsumoto, M.; Koiwai, A. Direct formation of LiAlH4 by a mechanochemical reaction. J. Alloys Compd. 2007, 441, 189–191. [Google Scholar] [CrossRef]
  70. Block, J.; Gray, A.P. The thermal decomposition of lithium aluminum hydride. Inorg. Chem. 1965, 4, 304–305. [Google Scholar] [CrossRef]
  71. 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]
  72. Ares, J.R.; Aguey-Zinsou, F.; Porcu, M.; Sykes, J.M.; Dornheim, M.; Klassen, T.; Bormann, R. Thermal and mechanically activated decomposition of LiAlH4. Mater. Res. Bull. 2008, 43, 1263–1275. [Google Scholar] [CrossRef]
  73. 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]
  74. Ismail, M.; Zhao, Y.; Yu, X.B.; Dou, S.X. Effects of NbF5 addition on the hydrogen storage properties of LiAlH4. Int. J. Hydrogen Energy 2010, 35, 2361–2367. [Google Scholar] [CrossRef]
  75. Blanchard, D.; Brinks, H.W.; Hauback, B.C.; Norby, P.; Muller, J. Isothermal decomposition of LiAlD4 with and without additives. J. Alloys Compd. 2005, 404–406, 743–747. [Google Scholar] [CrossRef]
  76. Vajeeston, P.; Ravindran, P.; Vidya, R.; Fjellvåg, H.; Kjekshus, A. Huge-pressure-induced volume collapse in LiAlH4 and its implications to hydrogen storage. Phys. Rev. B 2003, 68, 212101. [Google Scholar] [CrossRef]
  77. Sklar, N.; Post, B. The crystal structure of lithium aluminum hydride. Inorg. Chem. 1967, 6, 669–671. [Google Scholar] [CrossRef]
  78. Hauback, B.C.; Brinks, H.W.; Fjellvåg, H. Accurate structure of LiAlD4 studied by combined powder neutron and X-ray diffraction. J. Alloys Compd. 2002, 346, 184–189. [Google Scholar] [CrossRef]
  79. Sato, T.; Tomiyasu, K.; Ikeda, K.; Otomo, T.; Feygenson, M.; Neuefeind, J.; Yamada, K.; Orimo, S.I. Local atomic structural investigations of precursory phenomenon of the hydrogen release from LiAlD4. J. Alloys Compd. 2014, 586, 244–247. [Google Scholar] [CrossRef]
  80. Pitt, M.P.; Blanchard, D.; Hauback, B.C.; Fjellvåg, H.; Marshall, W.G. Pressure-induced phase transitions of the LiAlD4 system. Phys. Rev. B 2005, 72, 214113. [Google Scholar] [CrossRef]
  81. Løvvik, O.; Opalka, S.; Brinks, W.H.; Hauback, B.C. Crystal structure and thermodynamic stability of the lithium alanates LiAlH4 and Li3AlH6. Phys. Rev. B 2004, 69, 134117. [Google Scholar] [CrossRef]
  82. Liu, X.; McGrady, G.S.; Langmi, H.W.; Jensen, C.M. Facile cycling of Ti-doped LiAlH4 for high performance hydrogen storage. J. Am. Chem. Soc. 2009, 131, 5032–5033. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, X.; Langmi, H.W.; Beattie, S.D.; Azenwi, F.F.; McGradyc, S.; Jensen, C.M. Ti-doped LiAlH4 for hydrogen storage: Synthesis, catalyst loading and cycling performance. J. Am. Chem. Soc. 2011, 133, 15593–15597. [Google Scholar] [CrossRef] [PubMed]
  84. Graetz, J.; Wegrzyn, J.; Reilly, J.J. Regeneration of lithium aluminum hydride. J. Am. Chem. Soc. 2008, 130, 17790–17794. [Google Scholar] [CrossRef] [PubMed]
  85. Brinks, H.W.; Fossdal, A.; Fonneløp, J.E.; Hauback, B.C. Crystal structure and stability of LiAlD4 with TiF3 additive. J. Alloys Compd. 2005, 397, 291–295. [Google Scholar] [CrossRef]
  86. Ke, X.; Chen, C. Thermodynamic functions and pressure-temperature phase diagram of lithium alanates by ab initio calculations. Phys. Rev. B 2007, 76, 024112. [Google Scholar] [CrossRef]
  87. Jang, J.-W.; Shim, J.-H.; Cho, Y.C.; Lee, B.-J. Thermodynamic calculation of LiH ↔ Li3AlH6 ↔ LiAlH4 reactions. J. Alloys Compd. 2006, 420, 286–290. [Google Scholar] [CrossRef]
  88. Bogdanović, B.; Brand, R.A.; Marjanović, A.; Schwickardi, M.; Tölle, J. Metal-doped sodium aluminium hydrides as potential new hydrogen storage materials. J. Alloys Compd. 2000, 302, 36–58. [Google Scholar] [CrossRef]
  89. Vajo, J.J.; Skeith, S.L.; Mertens, F. Reversible storage of hydrogen in destabilized LiBH4. J. Phys. Chem. B 2005, 109, 3719–3722. [Google Scholar] [CrossRef]
  90. Mao, J.; Guo, Z.; Yu, X.; Ismail, M.; Liu, H. Enhanced hydrogen storage performance of LiAlH4–MgH2–TiF3 composite. Int. J. Hydrogen Energy 2011, 36, 5369–5374. [Google Scholar] [CrossRef]
  91. Wan, Q.; Li, P.; Li, Z.; Zhai, F.; Qu, X.; Volinsky, A.V. Improved hydrogen storage performance of MgH2-LiAlH4 composite by the addition of MnFe2O4. J. Phys. Chem. C 2013, 117, 26940–26947. [Google Scholar] [CrossRef]
  92. Lin, I.C.; Tsai, W.-T. In situ synchrotron X-ray diffraction study on the rehydrogenation behavior of MgH2-LiAlH4 composites. In Proceedings of the 2017 IEEE International Conference on Applied System Innovation: Applied System Innovation for Modern Technology, ICASI, Sapporo, Japan, 13–17 May 2017; Volume 2017, pp. 1918–1921. [Google Scholar]
  93. Ismail, M.; Zhao, Y.; Yu, X.B.; Dou, S.X. Effect of different additives on the hydrogen storage properties of the MgH2-LiAlH4 destabilized system. RSC Adv. 2011, 1, 408–414. [Google Scholar] [CrossRef]
  94. Chen, R.; Wang, X.; Xu, L.; Chen, L.; Li, S.; Chen, C. An investigation on the reaction mechanism of LiAlH4–MgH2 hydrogen storage system. Mater. Chem. Phys. 2010, 124, 83–87. [Google Scholar] [CrossRef]
  95. Mustafa, N.S.; Ismail, M. Enhanced hydrogen storage properties of 4MgH2-LiAlH4 composite system by doping with Fe2O3 nanopowder. Int. J. Hydrogen Energy 2014, 39, 7834–7841. [Google Scholar] [CrossRef]
  96. Sun, W.; Zhang, Y.; Zhu, Y.; Zhuang, X.; Dong, J.; Qu, Y.; Guo, X.; Chen, J.; Wang, Z.; Li, L. The hydrogen storage performance of a 4MgH2-LiAlH4-TiH2 composite system. J. Alloys Compd. 2016, 676, 557–564. [Google Scholar] [CrossRef]
  97. Meggouh, M.; Grant, D.M.; Walker, G.S. Optimizing the destabilization of LiBH4 for hydrogen storage and the effect of different Al sources. J. Phys. Chem. C 2011, 115, 22054–22061. [Google Scholar] [CrossRef]
  98. Jin, A.-S.; Shim, J.-H.; Cho, Y.W.; Yi, K.-W.; Zabara, O.; Fichtner, M. Reversible hydrogen storage in LiBH4–Al–LiH composite powder. Scr. Mater. 2008, 58, 963–965. [Google Scholar] [CrossRef]
  99. Mao, J.F.; Guo, Z.P.; Liu, H.K.; Yu, X.B. Reversible hydrogen storage in titanium-catalyzed LiAlH4–LiBH4 system. J. Alloys Compd. 2009, 487, 434–438. [Google Scholar] [CrossRef]
  100. Sridechprasat, P.; Phuitot, L.; Rangsunvigit, P.; Kitiyanna, B.; Kulprathipanja, S. A revisit to the hydrogen desorption/absorption behaviors of LiAlH4/LiBH4: Effects of catalysts. Energies 2012, 5, 3691–3700. [Google Scholar] [CrossRef]
  101. Gu, R.; Wang, C.Y.; Liu, D.M.; Gao, C.; Li, Y.T.; Si, T.Z. De-/rehydrogenation properties and reaction mechanisms of 4LiBH4–LiAlH4–MgF2 system. Int. J. Hydrogen Energy 2015, 40, 10536–10541. [Google Scholar] [CrossRef]
  102. Mao, J.F.; Yu, X.B.; Guo, Z.P.; Poh, C.K.; Liu, H.K.; Wu, Z.; Ni, J. Improvement of the LiAlH4−NaBH4 system for reversible hydrogen storage. J. Phys. Chem. C 2009, 113, 10813–10818. [Google Scholar] [CrossRef]
  103. Xia, G.; Meng, Q.; Guo, Z.; Gud, Q.; Liu, H.; Liu, Z.; Yu, X. Nanoconfinement significantly improves the thermodynamics and kinetics of co-infiltrated 2LiBH4–LiAlH4 composites: Stable reversibility of hydrogen absorption/resorption. Acta Mater. 2013, 61, 6882–6893. [Google Scholar] [CrossRef]
  104. Huang, Y.; Xia, G.; Zhang, J.; Guo, Z.; Yu, X. Graphene-tailored molecular bonds for advanced hydrogen and lithium storage performance. Energy Storage Mater. 2019, 17, 178–185. [Google Scholar] [CrossRef]
  105. Carr, C.L.; Jayawardana, W.; Zou, H.; White, J.L.; Gabaly, F.E.; Corandi, M.S.; Stavila, V.; Allendorf, M.D.; Majzoub, E.H. Anomalous H2 desorption rate of NaAlH4 confined in nitrogen-doped nanoporous carbon frameworks. Chem. Mater. 2018, 30, 2930–2938. [Google Scholar] [CrossRef]
  106. De Jongh, P.E.; Adelhelm, P. Nanosizing and nanoconfinement: New strategies towards meeting hydrogen storage goals. ChemSusChem 2010, 3, 1332–1348. [Google Scholar] [CrossRef] [PubMed]
  107. Fichtner, M. Nanoconfinement effects in energy storage materials. Phys. Chem. Chem. Phys. 2011, 13, 21186–21195. [Google Scholar] [CrossRef]
  108. Xiong, Z.; Wu, G.; Hu, J.; Chen, P. Investigation on chemical reaction between LiAlH4 and LiNH2. J. Power Sources 2006, 159, 167–170. [Google Scholar] [CrossRef]
  109. Varin, R.A.; Zbroniec, L.; Jang, M. Mechano-chemical synthesis of nanostructured hydride composites based on Li-Al-N-Mg for solid state hydrogen storage. Eng. Rev. 2011, 31, 111–123. [Google Scholar]
  110. Chen, R.; Wang, X.; Chen, L.; Li, S.; Ge, H.; Lei, Y.; Chen, C. An investigation on the reaction pathway between LiAlH4 and LiNH2 via gaseous ammonia. J. Alloys Compd. 2010, 495, 17–22. [Google Scholar] [CrossRef]
  111. Dolotko, O.; Kobayashi, T.; Wiench, J.W.; Pruski, M.; Pecharsky, V. Investigation of the thermochemical transformations in the LiAlH4-LiNH2 system. Int. J. Hydrogen Energy 2011, 36, 10626–10634. [Google Scholar] [CrossRef]
  112. Lu, J.; Fang, Z.Z. Dehydrogenation of a combined LiAlH4/LiNH2 system. J. Phys. Chem. B 2005, 109, 20830–20834. [Google Scholar] [CrossRef] [PubMed]
  113. Jepsen, L.H.; Ranvsbæk, D.B.; Grundlach, C.; Besenbacher, F.; Skibsted, J.; Jensen, T.R. A novel intermediate in the LiAlH4–LiNH2 hydrogen storage system. Dalton Trans. 2014, 43, 3095–3103. [Google Scholar] [CrossRef] [PubMed]
  114. Naik, M.; Rather, S.; So, C.S.; Hwang, S.W.; Kim, A.R.; Nahm, K.S. Thermal decomposition of LiAlH4 chemically mixed with lithium amide and transition metal chlorides. Int. J. Hydrogen Energy 2009, 34, 8937–8943. [Google Scholar] [CrossRef]
  115. Bogdanović, B.; Felderhoff, M.; Kaskel, S.; Pommerin, A.; Schlichte, K.; Schüth, F. Improved hydrogen storage properties of Ti-doped sodium alanate using titanium nanoparticles as doping agents. Adv. Mater. 2003, 15, 1012–1015. [Google Scholar] [CrossRef]
  116. Gross, K.J.; Guthrie, S.; Takara, S.; Thomas, G. In-situ X-ray diffraction study of the decomposition of NaAlH4. J. Alloys Compd. 2000, 297, 270–281. [Google Scholar] [CrossRef]
  117. Jensen, C.; Wang, Y.; Chou, M.Y. Alanates as hydrogen storage materials. In Solid-State Hydrogen Storage: Materials and Chemistry; Walker, G., Ed.; Woodhead Publishing: Boca Raton, FL, USA, 2008; pp. 381–419. [Google Scholar]
  118. Lauher, J.W.; Dougherty, D.; Herley, P.J. Sodium tetrahydroaluminate. Acta Crystallogr. Sect. B 1979, 35, 1454–1456. [Google Scholar] [CrossRef]
  119. Hauback, B.C.; Brinks, H.W.; Jensen, C.M.; Murphy, K.; Maeland, A.J. Neutron diffraction structure determination of NaAlD4. J. Alloys Compd. 2003, 358, 142–145. [Google Scholar] [CrossRef]
  120. Vajeeston, P.; Fjellvåg, H. Crystal structures of aluminum-based hydrides. Emerg. Mater. Res. 2015, 4, 192–217. [Google Scholar] [CrossRef] [Green Version]
  121. Canton, P.; Fichtner, M.; Frommen, C.; Léon, A. Synchrotron X-ray studies of Ti-doped NaAlH4. J. Phys. Chem. B 2006, 110, 3051–3054. [Google Scholar] [CrossRef]
  122. Rönnebro, E.; Noréus, D.; Kadir, K.; Reiser, A.; Bogdanović, B. Investigation of the perovskite related structures of NaMgH3, NaMgF3 and Na3AlH6. J. Alloys Compd. 2000, 299, 101–106. [Google Scholar] [CrossRef]
  123. Li, L.; Xu, C.; Chen, C.; Wang, Y.; Jiao, L.; Yuan, H. Sodium alanate system for efficient hydrogen storage. Int. J. Hydrogen Energy 2013, 38, 8798–8812. [Google Scholar] [CrossRef]
  124. Pitt, M.P.; Vullum, P.E.; Sørby, M.H.; Emerich, H.; Paskevicius, M.; Buckley, C.E.; Walmsley, J.C.; Holmestad, R.; Hauback, B.C. Hydrogen absorption kinetics of the transition-metal-chloride-enhanced NaAlH4 system. J. Phys. Chem. C 2012, 116, 14205–14217. [Google Scholar] [CrossRef]
  125. Lee, G.-J.; Shim, J.-H.; Whan, Y.C.; Lee, K. Reversible hydrogen storage in NaAlH4 catalyzed with lanthanide oxides. Int. J. Hydrogen Energy 2007, 32, 1911–1915. [Google Scholar] [CrossRef]
  126. Li, L.; Wang, Y.; Wang, Y.; Liu, G.; Han, Y.; Qiu, F.; Yan, C.; Song, D.; Jiao, L.; Yuan, H. Direct synthesis of sodium alanate with novel catalytic TiB2. J. Alloys Compd. 2011, 509S, S747–S749. [Google Scholar] [CrossRef]
  127. Liu, C.; Jiang, J.; Huang, S.; Wang, P.; Tian, H. Electronic and dehydrogenation properties of TiB2 cluster-doped NaAlH4 (101) surface: A first-principle approach. Int. J. Hydrogen Energy 2014, 39, 14178–14183. [Google Scholar] [CrossRef]
  128. Li, L.; Wang, Y.; Qiu, F.; Wang, Y.; Xu, Y.; An, C.; Jiao, L.; Yuan, H. Reversible hydrogen storage properties of NaAlH4 enhanced with TiN catalyst. J. Alloys Compds. 2013, 566, 137–141. [Google Scholar] [CrossRef]
  129. Mosher, D.A.; Tang, X.; Brown, R.J.; Saitta, S.; Laube, B.L.; Dold, R.H.; Anton, D.L. High Density Hydrogen Storage System Demonstration Using NaAlH4 Based Complex Compound Hydrides. 1–161; United Technologies Research Center: East Hartford, CT, USA, 2007. [Google Scholar]
  130. Bogdanović, B.; Felderhoff, M.; Pommerin, A.; Schüth, F.; Spielkamp, N.; Stark, A. Cycling properties of Sc-and Ce-doped NaAlH4 hydrogen storage materials prepared by the one-step direct synthesis method. J. Alloys Compd. 2009, 471, 383–386. [Google Scholar] [CrossRef]
  131. Berseth, P.A.; Harter, A.G.; Zidan, R.; Blomqvist, A.; Araújo, C.M.; Scheicher, R.H.; Ahuja, R.; Jena, P. Carbon nanomaterials as catalysts for hydrogen uptake and release in NaAlH4. Nano Lett. 2009, 4, 1501–1505. [Google Scholar] [CrossRef]
  132. Rongeat, C.; Scheerbaum, N.; Schultz, L.; Gutfleisch, O. Catalysis of H2 sorption in NaAlH4: General description and new insights. Acta Mater. 2011, 59, 1725–1733. [Google Scholar] [CrossRef]
  133. Lozano, G.A.; Na Ranong, C.; Bellosta von Colbe, J.M.; Bormann, R.; Fieg, G.; Hapke, J.; Dornheim, M. Empirical kinetic model of sodium alanate reacting system (I). Hydrogen absorption. Int. J. Hydrogen Energy 2010, 35, 6763–6772. [Google Scholar] [CrossRef] [Green Version]
  134. Frankcombe, T.J. Proposed mechanisms for the catalytic activity of Ti in NaAlH4. Chem. Rev. 2012, 112, 2164–2178. [Google Scholar] [CrossRef] [PubMed]
  135. Beattie, S.D.; Mcgrady, G.S. Hydrogen desorption studies of NaAlH4 and LiAlH4 by in situ heating in an ESEM. Int. J. Hydrogen Energy 2009, 34, 9151–9156. [Google Scholar] [CrossRef]
  136. Singh, S.; Eijt, S.W.H.; Huot, J.; Kockelmann, W.A.; Wagemaker, M.; Mulder, F.M. The TiCl3 catalyst in NaAlH4 for hydrogen storage induces grain refinement and impacts on hydrogen vacancy formation. Acta Mater. 2007, 55, 5549–5557. [Google Scholar] [CrossRef]
  137. Løvvik, O.M.; Opalka, S.M. Density functional calculations of Ti-enhanced NaAlH4. Phys. Rev. B 2005, 71, 054103. [Google Scholar] [CrossRef]
  138. Løvvik, O.M.; Opalka, S.M. Stability of Ti in NaAlH4. Appl. Phys. Lett. 2006, 88, 161917. [Google Scholar] [CrossRef]
  139. Michel, K.J.; Ozoliņš, V. Site substitution of Ti in NaAlH4 and Na3AlH6. J. Phys. Chem C 2011, 115, 21454–21464. [Google Scholar] [CrossRef]
  140. Marashdeh, A.; Versluis, J.-W.I.; Valdés, A.; Olsen, R.A.; Løvvik, O.M.; Kroes, G.-J. Effect of transition metal dopants on initial mass transport in the dehydrogenation of NaAlH4: Density functional theory study. J. Phys. Chem. C 2013, 117, 3–14. [Google Scholar] [CrossRef]
  141. Al-Mahboob, A.; Muller, E.; Karim, A.; Muckerman, J.T.; Ciobanu, C.V.; Sutter, P. Site-dependent activity of atomic Ti catalysts in Al-based hydrogen storage materials. J. Am. Chem. Soc. 2012, 134, 10381–10384. [Google Scholar] [CrossRef]
  142. Wang, Y.; Zhang, F.; Stumpf, R.; Lin, P.; Chou, M.Y. Catalytic effect of near-surface alloying on hydrogen interaction on the aluminum surface. Phys. Rev. B 2011, 83, 195419. [Google Scholar] [CrossRef]
  143. Zhang, F.; Wang, Y.; Chou, M.Y. Hydrogen interaction with the Al surface promoted by subsurface alloying with transition metals. J. Phys. Chem. C 2012, 116, 18663–18668. [Google Scholar] [CrossRef]
  144. Xiong, R.; Sang, G.; Zhang, G.; Yan, X.; Li, P.; Yao, Y.; Luo, D.; Chen, C.; Tang, T. Evolution of the active species and catalytic mechanism of Ti doped NaAlH4 for hydrogen storage. Int. J. Hydrogen Energy 2017, 42, 6088–6095. [Google Scholar] [CrossRef]
  145. Léon, A.; Kircher, O.; Rothe, J.; Fichtner, M. Chemical state and local structure around titanium atoms in NaAlH4 doped with TiCl3 using X-ray absorption spectroscopy. J. Phys. Chem. B 2004, 108, 16372–16376. [Google Scholar] [CrossRef]
  146. Haiduc, A.G.; Stil, H.A.; Schwarz, M.A.; Paulus, P.; Geerlings, J.J.C. On the fate of the Ti catalyst during hydrogen cycling of sodium alanate. J. Alloys Compd. 2005, 393, 252–263. [Google Scholar] [CrossRef]
  147. Kang, X.D.; Wang, P.; Song, X.P.; Yao, X.D.; Lu, G.Q.; Cheng, H.M. Catalytic effect of Al3Ti on the reversible dehydrogenation of NaAlH4. J. Alloys Compd. 2006, 424, 365–369. [Google Scholar] [CrossRef]
  148. Brinks, H.W.; Hauback, B.C.; Srinivasan, S.S.; Jensen, C.M. Synchrotron X-ray studies of Al1−yTiy formation and re-hydriding inhibition in Ti-enhanced NaAlH4. J. Phys. Chem. B 2005, 109, 15780–15785. [Google Scholar] [CrossRef] [PubMed]
  149. Liu, J.; Ge, Q. A precursor state for formation of TiAl3 complex in reversible hydrogen desorption/adsorption from Ti-doped NaAlH4. Chem. Commun. 2006, 42, 1822–1824. [Google Scholar] [CrossRef]
  150. Dathara, G.K.P.; Mainardi, D.S. Structure and dynamics of Ti-Al-H compounds in Ti-doped NaAlH4. Mol. Simul. 2008, 34, 201–210. [Google Scholar] [CrossRef]
  151. Liu, J.; Han, Y.; Ge, Q. Effect of doped transition metal on reversible hydrogen release/uptake from NaAlH4. Chem. Eur. J. 2009, 15, 1685–1695. [Google Scholar] [CrossRef]
  152. Ismail, M.; Zhao, Y.; Yu, X.B.; Mao, J.F.; Dou, S.X. The hydrogen storage properties and reaction mechanism of the MgH2-NaAlH4 composite system. Int. J. Hydrogen Energy 2011, 36, 9045–9050. [Google Scholar] [CrossRef]
  153. Bendyna, J.K.; Dyjak, S.; Notten, P.H.L. The influence of ball-milling time on the dehydrogenation properties of the NaAlH4-MgH2 composite. Int. J. Hydrogen Energy 2015, 40, 4200–4206. [Google Scholar] [CrossRef]
  154. Ismail, M.; Zhao, Y.; Yu, X.B.; Dou, S.X. Improved hydrogen storage performance of MgH2-NaAlH4 composite by the addition of TiF3. Int. J. Hydrogen Energy 2012, 37, 8395–8401. [Google Scholar] [CrossRef]
  155. Rafi-ud-Din; Qu, X.; Zahid, G.H.; Asghar, Z.; Shahzad, M.; Iqbal, M.; Ahmad, E. Improved hydrogen storage performances of MgH2-NaAlH4 system catalyzed by TiO2 nanoparticles. J. Alloys Compd. 2014, 604, 317–324. [Google Scholar] [CrossRef]
  156. Cheng, H.; Chen, Y.; Sun, W.; Lou, H.; Liu, Y.; Qi, Q.; Zhang, J.; Liu, J.; Yan, K.; Jin, H.; et al. The enhanced de/re-hydrogenation performance of 4MgH2-NaAlH4 composite by doping with TiH2. J. Alloys Compd. 2017, 698, 1002–1008. [Google Scholar] [CrossRef]
  157. Plerdseanoy, P.; Meethon, S.; Utke, R. Dehydrogenation kinetics, reversibility, and reaction mechanisms of reversible hydrogen storage material based on nanoconfined MgH2-NaAlH4. J. Phys. Chem. Solids 2015, 87, 16–22. [Google Scholar] [CrossRef]
  158. Bhatnagar, A.; Pandey, S.K.; Dixit, V.; Shukla, V.; Shahi, R.R.; Shaz, M.A.; Srivastava, O.N. Catalytic effect of carbon nanostructures on the hydrogen storage properties of MgH2-NaAlH4 composite. Int. J. Hydrogen Energy 2014, 39, 14240–14246. [Google Scholar] [CrossRef]
  159. Ismail, M.; Zhao, Y.; Dou, S.X. An investigation on the hydrogen storage properties and reaction mechanism of the destabilized MgH2-Na2AlH6 (4:1) system. Int. J. Hydrogen Energy 2013, 38, 1478–1483. [Google Scholar] [CrossRef]
  160. Ranvsbæk, D.B.; Jensen, T.R. Tuning the hydrogen storage properties and reactivity: Investigation of the LiBH4-NaAlH4 system. J. Phys. Chem. Solids 2010, 71, 144–1149. [Google Scholar] [CrossRef]
  161. Morioka, H.; Kakizaki, K.; Chung, S.C.; Yamada, A. Reversible hydrogen decomposition of KAlH4. J. Alloys Compd. 2003, 353, 310–314. [Google Scholar] [CrossRef]
  162. Dymova, T.N.; Eliseeva, N.G.; Bakum, S.I.; Dergachev, Y.M. Direct synthesis of alkali metal aluminum hydrides in the melt. Dokl. Akad. Nauk SSSR 1974, 215, 1369–1372. [Google Scholar]
  163. Pukazhselvan, D.; Fagg, D.P.; Srivastava, O.N. On step high pressure mechanochemical synthesis of reversible alanates NaAlH4 and KAlH4. Int. J. Hydrogen Energy 2015, 40, 4916–4924. [Google Scholar] [CrossRef]
  164. Ares, J.R.; Zhang, J.; Charpentier, T.; Cuevas, F.; Latroche, M. Asymmetric reaction path and hydrogen sorption mechanism in mechanochemically synthesized potassium alanate (KAlH4). J. Phys. Chem. C 2016, 120, 21299–21308. [Google Scholar] [CrossRef]
  165. Mamatha, M.; Weidenhaler, C.; Pommerin, A.; Felderhoff, M.; Schüth, F. Comparative studies of the decomposition of alanates followed by in situ XRD and DSC methods. J. Alloys Compd. 2006, 416, 303–314. [Google Scholar] [CrossRef]
  166. Matysina, Z.A.; Zaginaichenko, S.Y.; Schur, D.V.; Zolotarenko, A.D.; Zolotarenko, A.D.; Gabdulin, M.T. Hydrogen sorption properties of potassium alanate. Russ. Phys. J. 2018, 61, 253–263. [Google Scholar] [CrossRef]
  167. Ares, J.R.; Aguey-Zinsou, K.-F.; Leardini, F.; Jímenez Ferrer, I.; Fernandez, J.-F.; Guo, Z.-X.; Sánchez, C. Hydrogen absorption/desorption mechanism in potassium alanate (KAlH4) and enhancement by TiCl3 doping. J. Phys. Chem. C 2009, 113, 6845–6851. [Google Scholar] [CrossRef]
  168. Arroyo y de Dompablo, M.E.; Ceder, G. First principles investigations of complex hydrides AMH4 and A3MH6 (A = Li, Na, K, M = B, Al, Ga) as hydrogen storage systems. J. Alloys Compd. 2004, 364, 6–12. [Google Scholar] [CrossRef]
  169. Santhanam, R.; McGrady, G.S. Synthesis of alkali metal hexahydroaluminate complexes using dimethyl ether as a reaction medium. Inorg. Chim. Acta 2008, 361, 473–478. [Google Scholar] [CrossRef]
  170. Dymova, T.N.; Selivokhina, M.S.; Eliseeva, N.G. Thermal stability of potassium alumohydride. Dokl. Akad. Nauk 1963, 153, 1330–1332. [Google Scholar]
  171. Sorte, E.G.; Emery, S.B.; Majzoub, E.H.; Ellis-Caleo, T.; Ma, Z.L.; Hammann, B.A.; Hayes, S.E.; Bowman, R.C., Jr.; Conradi, M.S. NMR study of anion dynamics in solid KAlH4. J. Phys. Chem. C 2014, 118, 5725–5732. [Google Scholar] [CrossRef]
  172. Arnbjerg, L.M.; Jensen, T.R. New compounds in the potassium-aluminium-hydrogen system observed during release and uptake of hydrogen. Int. J. Hydrogen Energy 2012, 37, 345–356. [Google Scholar] [CrossRef]
  173. Hauback, B.C.; Brinks, H.W.; Heyn, R.H.; Blom, R.; Fjellvag, H. The crystal structure of KAlD4. J. Alloys Compd. 2005, 394, 35–38. [Google Scholar] [CrossRef]
  174. Vajeeston, P.; Ravindran, P.; Kjekshus, A.; Fjellvag, H. Crystal structure of KAlH4 from first principle calculations. J. Alloys Compds. 2004, 363, L7–L11. [Google Scholar] [CrossRef]
  175. Vajeeston, P.; Ravindran, P.; Kjekhus, A.; Fjellvåg, H. First-principles investigations of aluminum hydrides: M3AlH6 (M = Na, K). Phys. Rev. B 2005, 71, 092103. [Google Scholar] [CrossRef]
  176. Weindenthaler, C.; Felderhoff, M.; Bernert, T.; Sørby, M.H.; Hauback, B.C.; Krech, D. Synthesis, crystal structure analysis and decomposition of RbAlH4. Crystals 2018, 8, 103. [Google Scholar] [CrossRef]
  177. Adkis, T.G.; Gavrilenko, V.V.; Zahartin, I.I.; Ignat’eva, I.A. Study of the infrared spectra of alkali metal aluminum hydrides. Zhural Prikl. Spektrosk. 1967, 6, 806–812. [Google Scholar] [CrossRef]
  178. Bestide, J.P.; Hajri, J.E.; Claudy, P.; Hajbi, A.E. A new route to alkali metal aluminum hydrides MAlH4 with M = Na, K, Rb, Cs and structural features for the whole family with M = Li to Cs. Synth. React. Inorg. Met. Org. Chem. 1995, 25, 1037–1047. [Google Scholar] [CrossRef]
  179. Dymova, T.N.; Bakum, S.I.; Mirsaidov, U. Phase States of Alkaly Aluminum Hydrides. Dokl. Akad. Nauk SSSR 1974, 216, 87–90. [Google Scholar]
  180. Vajeeston, P.; Ravindran, P.; Vidya, R.; Fjellvåg, H.; Kjeshus, A. Design of potential hydrogen-storage materials using first-principle density-functional calculations. Cryst. Growth Des. 2004, 4, 471–477. [Google Scholar] [CrossRef]
  181. Adimi, S.; Arabi, H.; Ghorbani, S.R.; Pourarin, F. AB-initio study of pressure-induced aluminum hydrides AAlH4 (A = Li, Na, K, Rb, Cs). Int. J. Hydrogen Energy 2017, 42, 25303–25309. [Google Scholar] [CrossRef]
  182. Ravindan, P.; Vajeeston, P.; Vidya, R.; Fjellvåg, H.; Kjekshus, A. Modeling of hydrogen storage materials by density-functional calculations. J. Power Sources 2006, 159, 88–99. [Google Scholar] [CrossRef]
  183. Krech, D.; Zibrowius, B.; Weidenthaler, C.; Felderhoff, M. On the preparation and structure of cesium aluminum tetrahydride. Eur. J. Inorg. Chem. 2014, 2014, 5683–5688. [Google Scholar] [CrossRef]
  184. Bakum, S.I.; Kuznetsova, S.F.; Tarasov, V.P. CsAlH4 thermolysis. Russ. J. Inorg. Chem. 2013, 58, 1547–1549. [Google Scholar] [CrossRef]
  185. Bernert, T.; Krech, D.; Kockelmann, W.; Felderhoff, M.; Frankcombe, T.J.; Weidenthaler, C. Crystal structure relation between tetragonal and orthorhombic CsAlD4: DFT and time-of-flight neutron powder diffraction studies. Eur. J. Inorg. Chem. 2015, 2015, 5545–5550. [Google Scholar] [CrossRef]
  186. Graetz, J. Metastable metal hydrides for hydrogen storage. ISRN Mater. Sci. 2012, 2012, 863025. [Google Scholar] [CrossRef]
  187. Mackay, K.M. Hydrogen Compounds of the Metallic Elements; E. & F. N. Spon Ltd.: London, UK, 1966; p. 177. [Google Scholar]
  188. Schlesinger, H.I.; Brown, H.C.; Abraham, B.; Bond, A.C.; Davidson, N.; Finholt, A.E.; Gilbreath, J.R.; Hoekstra, H.; Horvitz, L.; Hyde, E.K.; et al. New developments in the chemistry of diborane and the borohydrides. I. General summary. J. Am. Chem. Soc. 1953, 75, 186–190. [Google Scholar] [CrossRef]
  189. Schlesinger, H.I.; Brown, H.C.; Hyde, E.K. The preparation of other borohydrides by metathetical reactions utilizing the alkali metal borohydrides. J. Am. Chem. Soc. 1953, 75, 209–213. [Google Scholar] [CrossRef]
  190. Ashby, E.C.; Sanders, J.R.; Claudy, P.; Schwartz, R.D. A study of the reactions of lithium aluminum hydride and sodium aluminum hydride with beryllium chloride in diethyl ether and tetrahydrofuran. A report questioning the existence of Be(AlH4)2 in solution. Inorg. Chem. 1973, 12, 2860–2868. [Google Scholar] [CrossRef]
  191. Wiberg, E.; Bauer, R. Zur Kenntnis eines Berylliumwasserstoffs BeH2. Z. Nat. 1951, 6, 171. [Google Scholar] [CrossRef]
  192. Klaveness, A.; Vajeeston, P.; Ravindran, P.; Fjellvåg, H.; Kjekshus, A. Structure and bonding in BAlH5 (B = Be, Ca, Sr) from first-principle calculations. J. Alloys Compd. 2007, 443, 225–232. [Google Scholar] [CrossRef]
  193. Santhosh, M.; Rajeswarapalanichamy, R.; Priyanga, G.S.; Kanagaprabha, S.; Cinthia, A.J.; Iyakutti, K. A first principles study of structural stability, electronic structure and mechanical properties of beryllium alanate BeAlH5. AIP Conf. Proc. 2015, 1665, 090032. [Google Scholar] [CrossRef]
  194. He, L.; Wang, S.; Li, Z.; Liu, X.; Jlang, L. Synthesis of magnesium alanate by ball milling MgH2 and AlCl3 mixtures. Rare Met. 2011, 30, 55–58. [Google Scholar] [CrossRef]
  195. Xiao, X.; Qin, T.; Jiang, Y.; Jiang, F.; Li, M.; Fan, X.; Li, S.; Ge, H.; Wang, Q.; Chen, L. Significantly enhanced hydrogen desorption properties of Mg(AlH4)2 nanoparticles synthesized using solvent free strategy. Prog. Nat. Sci. 2017, 27, 112–120. [Google Scholar] [CrossRef]
  196. Fichtner, M.; Fuhr, O.; Kircher, O. Magnesium alanate—A material for reversible hydrogen storage? J. Alloys Compd. 2003, 356–357, 418–422. [Google Scholar] [CrossRef]
  197. Liu, Y.; Pang, Y.; Zhang, X.; Zhou, Y.; Gao, M.; Pan, H. Synthesis and hydrogen storage thermodynamics and kinetics of Mg(AlH4)2 submicron rods. Int. J Hydrog. Energy 2012, 37, 18148–18154. [Google Scholar] [CrossRef]
  198. Fossdal, A.; Brinks, H.W.; Fichtner, M.; Hauback, B.C. Thermal decomposition of Mg(AlH4)2 studied by in situ synchrotron X-Ray diffraction. J. Alloys Compd. 2005, 404–406, 752–756. [Google Scholar] [CrossRef]
  199. Pang, Y.; Liu, Y.; Zhang, X.; Gao, M.; Pan, H. Role of particle size, grain size, microstrains and lattice distortion in improved dehydrogenation properties of the ball-milled Mg(AlH4)2. Int. J. Hydrogen Energy 2013, 38, 1460–1468. [Google Scholar] [CrossRef]
  200. Pang, Y.; Liu, Y.; Zhang, X.; Gao, M.; Pan, H. TiF4-doped Mg(AlH4)2 with significantly improved dehydrogenation properties. Int. J. Hydrogen Energy 2013, 38, 13343–13351. [Google Scholar] [CrossRef]
  201. Gremaud, R.; Borgschulte, A.; Lohstroh, W.; Schreuders, H.; Züttel, A.; Dam, B.; Griessen, R. Ti-Catalyzed Mg(AlH4)2: A reversible hydrogen storage material. J. Alloys Compd. 2005, 404–406, 775–778. [Google Scholar] [CrossRef]
  202. Fossdal, A.; Brinks, H.W.; Fichtner, M.; Hauback, B.C. Determination of the crystal structure of Mg(AlH4)2 by combined X-ray and neutron diffraction. J. Alloys Compd. 2005, 387, 47–51. [Google Scholar] [CrossRef]
  203. Fichtner, M.; Engel, J.; Fuhr, O.; Glöss Rubner, O.; Ahlrichs, R. The structure of magnesium alanate. Inorg. Chem. 2003, 42, 7060–7066. [Google Scholar] [CrossRef]
  204. Vajeeston, P.; Ravindran, P.; Kjekshus, A.; Fjellvag, H. High hydrogen content complex hydrides: A density-functional study. Appl. Phys. Lett. 2006, 89, 071906. [Google Scholar] [CrossRef] [Green Version]
  205. Van Setten, M.J.; de Wijs, G.A.; Popa, V.A.; Brocks, G. Ab initio study of Mg(AlH4)2. Phys. Rev. B 2005, 72, 073107. [Google Scholar] [CrossRef]
  206. Akbarzadeh, A.R.; Wolverton, C.; Ozolins, V. First-principles determination of crystal structures, phase stability and reaction thermodynamics in the Li-Mg-Al-H hydrogen storage system. Phys. Rev. B 2009, 79, 184102. [Google Scholar] [CrossRef]
  207. Klaveness, A.; Vajeeston, P.; Ravindran, P.; Fjellvåg, H.; Kjekshus, A. Structural phase stability and bonding behavior of BAlH5 (B = Mg, Ba) from first-principles calculations. Phys. Rev. B 2006, 73, 094122. [Google Scholar] [CrossRef]
  208. Pang, Y.; Li, Q. Insight into the kinetic mechanism of the first-step dehydrogenation of Mg(AlH4)2. Scr. Mater. 2017, 130, 223–228. [Google Scholar] [CrossRef]
  209. Løvvik, O.M.; Molin, P.N. Density-functional band-structure calculations of magnesium alanate Mg(AlH4)2. Phys. Rev. B 2005, 72, 073201. [Google Scholar] [CrossRef]
  210. Spanò, E.; Bernasconi, M. Ab initio of the vibrational properties of Mg(AlH4)2. Phys. Rev. B 2005, 71, 174301. [Google Scholar] [CrossRef]
  211. Yang, C.H.; Chen, T.T.; Tsai, W.T.; Liu, B.H. In situ synchrotron X-ray diffraction study on the improved dehydrogenation performance of NaAlH4-Mg(AlH4)2 mixture. J. Alloys Compd. 2013, 577, 6–10. [Google Scholar] [CrossRef]
  212. Hudson, M.S.L.; Pukazhselvan, D.; Scheeja, I.G.; Srivastava, O.N. Studies on synthesis and dehydrogenation behavior of magnesium alanate and magnesium-sodium alanate mixture. Int. J. Hydrogen Energy 2007, 32, 4933–4938. [Google Scholar] [CrossRef]
  213. Wang, Y.; Wang, Y.; Wang, X.; Zhang, H.; Jiao, L.; Yuan, H. Destabilization effects of Mg(AlH4)2 on MgH2: Improved desorption performance and its reaction mechanism. Int. J. Hydrogen Energy 2014, 39, 17747–17753. [Google Scholar] [CrossRef]
  214. Liu, D.; Liu, Q.; Si, T.; Zhang, Q.; Fang, F.; Sun, D.; Ouyang, L.; Zhu, M. Superior hydrogen storage properties of LiBH4 catalyzed by Mg(AlH4)2. Chem. Commun. 2011, 47, 5741–5743. [Google Scholar] [CrossRef] [PubMed]
  215. Pang, Y.; Liu, Y.; Zhang, X.; Li, Y.; Gao, M.; Pan, H. New insights into the effects of NaCl and LiCl on the hydrogen storage behaviors of a 6LiBH4-Mg(AlH4)2 composite. RSC Adv. 2015, 5, 12144–12151. [Google Scholar] [CrossRef]
  216. Huang, J.; Gao, M.; Li, Z.; Cheng, X.; Gu, J.; Liu, Y. Destabilization of combined Ca(BH4)2 and Mg(AlH4)2 for improved hydrogen storage properties. J. Alloys Compd. 2016, 670, 135–143. [Google Scholar] [CrossRef]
  217. Finholt, E.; Barbaras, G.D.; Barbaras, G.K.; Urry, G.; Wartik, T.; Schlesinger, H.I. The preparation of sodium and calcium aluminum hydrides. J. Inorg. Nucl. Chem. 1955, 1, 317–325. [Google Scholar] [CrossRef]
  218. Fichtner, M.; Frommen, C.; Fuhr, O. Synthesis and properties of calcium alanate and two solvent adducts. Inorg. Chem. 2005, 44, 3479–3484. [Google Scholar] [CrossRef]
  219. Li, C.; Xiao, X.; Ge, P.; Xue, J.; Li, S.; Ge, H. Investigation on synthesis, structure and catalytic modification of Ca(AlH4)2 complex hydride. Int. J. Hydrogen Energy 2012, 37, 936–941. [Google Scholar] [CrossRef]
  220. Mamatha, M.; Bogdanović, B.; Felderhoff, M.; Pommerin, A.; Schmidt, W.; Schüth, F.; Weidenthaler, C. Mechanochemical preparation and investigation of properties of magnesium, calcium and lithium-magnesium alanates. J. Alloys Compds. 2006, 407, 78–86. [Google Scholar] [CrossRef]
  221. Sato, T.; Sørby, M.H.; Ikeda, K.; Sato, S.; Hauback, B.C.; Orimo, S. Syntheses, crystal structures, and thermal analyses of solvent-free Ca(AlD4)2 and CaAlD5. J. Alloys Compd. 2009, 487, 472–478. [Google Scholar] [CrossRef]
  222. Weidenthaler, C.; Frankcombe, T.J.; Felderhoff, M. First crystal structure studies of CaAlH5. Inorg. Chem. 2006, 45, 3849–3851. [Google Scholar] [CrossRef]
  223. Løvvik, O.M. Crystal structure of Ca(AlH4)2 predicted from density-functional band-structure calculations. Phys. Rev. B 2005, 71, 144111. [Google Scholar] [CrossRef]
  224. Wolverton, C.; Ozoliņš, V. Hydrogen storage in calcium alanate: First-principles thermodynamics and crystal structures. Phys. Rev. B 2007, 75, 064101. [Google Scholar] [CrossRef]
  225. Liu, D.M.; Gao, C.; Qian, Z.X.; Si, T.Z.; Zhang, Q.A. Reversible hydrogen storage in LiBH4/Ca(AlH4)2 systems. Int. J. Hydrogen Energy 2013, 38, 3291–3295. [Google Scholar] [CrossRef]
  226. Hanada, N.; Lohstroh, W.; Fichtner, M. Comparison of the calculated and experimental scenarios for solid-state reactions involving Ca(AlH4)2. J. Phys. Chem. C 2008, 112, 131–138. [Google Scholar] [CrossRef]
  227. Alapati, S.V.; Johnson, J.K.; Sholl, D.S. Large-scale screening of metal hydride mixtures for high-capacity hydrogen storage from first-principles calculations. J. Phys. Chem. C 2008, 112, 5258–5262. [Google Scholar] [CrossRef]
  228. Dymova, T.N.; Aleksandrov, D.P.; Konoplev, V.N.; Silina, T.A.; Sizareva, A.S. Peculiarities of the solid-phase formation of strontium pentahydroaluminate from binary hydrides by mechanochemical activation and modeling of theprocess on the basis of thermo gasovolumetry data. Russ. J. Coord. Chem. 2000, 26, 531–537. [Google Scholar]
  229. Sato, T.; Takagi, S.; Sørby, M.H.; Deledda, S.; Hauback, B.C.; Orimo, S.I. Crystal structural determination of SrAlD5 with corner-sharing AlD6 octahedron chains by X-ray and neutron diffraction. Crystals 2018, 8, 89. [Google Scholar] [CrossRef]
  230. Zhang, Q.A.; Makamura, Y.; Oikawa, K.I.; Kamiyama, T.; Akiba, E. Synthesis and crystal structure of Sr2AlH7: A new structural type of alkaline earth aluminum hydride. Inorg. Chem. 2002, 41, 6547–6549. [Google Scholar] [CrossRef]
  231. Zhang, Q.A.; Akiba, E. Synthesis of Sr2AlH7 by ball milling followed by hydrogenation. J. Alloys Compd. 2005, 394, 308–311. [Google Scholar] [CrossRef]
  232. Zhang, Q.A.; Enoki, H.; Akiba, E. Formation mechanism of Sr2AlH7. J. Alloys Compd. 2007, 427, 153–159. [Google Scholar] [CrossRef]
  233. Gingl, F.; Vogt, T.; Akiba, E. Trigonal SrAl2H2: The first zintl phase hydride. J. Alloys Compod. 2000, 306, 127–132. [Google Scholar] [CrossRef]
  234. Zhang, Q.A.; Nakamura, Y.; Oikawa, K.I.; Kamiyama, T.; Akiba, E. New alkaline earth aluminum hydride with one-dimensional zigzag chains of [AlH6]: Synthesis and crystal structure of BaAlH5. Inorg. Chem. 2002, 41, 6941–6943. [Google Scholar] [CrossRef] [PubMed]
  235. Zhang, Q.A.; Nakamura, Y.; Oikawa, K.; Kamiyama, T.; Akiba, E. Hydrogen-induced phase decomposition of Ba7Al13 and the crystal structure of Ba2AlH7. J. Alloys Compd. 2003, 316, 180–186. [Google Scholar] [CrossRef]
  236. Liu, X.; Asano, K.; Sakaki, K.; Nakamura, Y.; Enoki, H.; Akiba, E. Investigations on the formation and decomposition behaviors of BaAlH5 and Ba2AlH7. J. Phys. Chem. C 2008, 112, 17423–17426. [Google Scholar] [CrossRef]
  237. Liu, X.; Asano, K.; Akiba, E. Direct synthesis of BaAlH5 and Ba2AlH7 from BaH2 and Al system and their hydriding/dehydriding characteristics. J. Alloys Compd. 2009, 477, 744–748. [Google Scholar] [CrossRef]
  238. Chahrkin, O.P. Theoretical study of metal tetrahydroborates and alanates L(MH4)3, HL(MH4)2, and H2L(MH4) (L = Be, Mg, Al, Sc, Ti, V, Zn, M = B, Al). Russ. J. Inorg. Chem. 2008, 53, 1910–1919. [Google Scholar] [CrossRef]
  239. Kost, M.E.; Golovanova, A.I. Interaction of LiH and Al with transition metal halogens. Izv. Akad. Nauk SSSR 1978, 14, 1732–1734. [Google Scholar]
  240. Cao, Z.; Ouyang, L.; Wang, H.; Liu, J.; Felderhoff, M.; Zhu, M. Reversible hydrogen storage in yttrium aluminum hydride. J. Mater. Chem. A 2017, 5, 6042–6046. [Google Scholar] [CrossRef] [Green Version]
  241. Kost, M.E.; Golovanova, A.L. Interaction of titanium and iron halides with lithium aluminum hydride in diethyl ether. Translated from Izv. Akad. Nauk SSSR 1975, 5, 991–994. [Google Scholar] [CrossRef]
  242. Ramzan, M.; Ahuja, R. MN+1AXN (M = Ti, A = Al. X = H) phase class materials with hydrogen: Ti4AlH3 and Ti3AlH2. Appl. Phys. Lett. 2010, 96, 261906. [Google Scholar] [CrossRef]
  243. Maeland, A.J.; Hauback, B.C.; Fjellvåg, H.; Sørby, M.H. The structures of hydride phases in the Ti3Al/H system. Int. J. Hydrogen Energy 1999, 24, 163–168. [Google Scholar] [CrossRef]
  244. Epshteyn, A.; Miller, J.B.; Pettigrew, K.A.; Stroud, R.M.; Purdy, A.P. Surface passivated air and moisture stable mixed zirconium aluminum metal-hydride nanoparticles. Mater. Res. Soc. Symp. Proc. 2008, 1056, HH03–HH16. [Google Scholar] [CrossRef]
  245. Matsubara, E.; Waseda, Y.; Li, X.G.; Aoki, K.; Masumoto, T. X-Ray diffraction study of amorphous Zr3InH4 and Zr3AlH4 alloys produced by hydrogen-induced amorphization method. J. Mater. Sci. Lett. 1990, 9, 1017–1019. [Google Scholar] [CrossRef]
  246. Wiberg, E.; Neumaier, H. Zur Kenntnis von Doppel-und Tripelhydriden der Übergangsmetalle. I. Über die Umsetzung von Niob(v)-chlorid mit Lithiumalanat. Z. Anorg. Allg. Chem. 1965, 340, 189–200. [Google Scholar] [CrossRef]
  247. Aubry, J.; Monnier, G.; Hackspill, M.J. Sur la preparation de quelques aluminohydrures métalliques. Comptes Rendus 1954, 238, 2534–2535. [Google Scholar]
  248. Monnier, G. Contribution a l’étude de l´éther oxyde d´éthyle, milieu réactionel, en chimie minérale. Ann. Chim. 1957, 13, 14–57. [Google Scholar]
  249. Neumaier, H.; Büchel, D.; Ziegelmaier, G. Über die Umsetzung von Eisen(III)-chlorid mit Lithiumalanat bei tiefen Temperaturen. Z. Anorg. Allg. Chem. 1996, 345, 46–51. [Google Scholar] [CrossRef]
  250. Scheaffer, G.W.; Roscoe, J.S.; Stewart, A.C. The reduction of iron(III) chloride with lithium aluminohydride and lithium borohydride: Iron(II) borohydride. J. Am. Chem. Soc. 1956, 78, 729–733. [Google Scholar] [CrossRef]
  251. Ashby, E.C.; Kovar, R.A. Reaction of lithium aluminum hydride with copper(I) and mercury(II) salts. Nature of the reactive species in the conjugate reducing agent LiAlH4-CuI. Inorg. Chem. 1977, 16, 1437–1440. [Google Scholar] [CrossRef]
  252. Wiberg, E.; Henle, W. Über eine neue Darstellungsweise des Kupferwasserstoffs CuH. Zeitschrift für Naturforschung B 1952, 7, 250. [Google Scholar] [CrossRef]
  253. Wiberg, E.; Henle, W. Zur Kenntnis eines Silber-aluminium-wasserstoffs AgAlH4. Z. Nat. 1952, 7, 250–251. [Google Scholar] [CrossRef]
  254. Zhizhin, K.Y.; Mal´tseva, N.N.; Buzanov, G.A.; Kunetsov, N.T. Hydride compounds of zinc. Russ. J. Inorg. Chem. 2014, 59, 1665–1678. [Google Scholar] [CrossRef]
  255. Wiberg, E.; Schmidt, M. Zur Kenntnis eines Gallium-aluminium-wasserstoffs Ga(AlH4)3 und eines Galliumwasserstoff-Ätherats GaH3*OR2. Z. Nat. 1951, 6, 172. [Google Scholar]
  256. Wiberg, E. Neuere Ergebnisse der präparativen Hydrid-Forschung. Angew. Chem. 1953, 65, 16–34. [Google Scholar] [CrossRef]
  257. Wiberg, E.; Dittmann, O.; Nöth, H.; Schmidt, M. Über Wasserstoff-Verbindungen des Thalliums, V. Zur Kenntnis eines Thallium(I)-boranats TlBH4 und Thallium(I)-alanats TlAlH4. Z. Nat. 1957, 12, 62–63. [Google Scholar]
  258. Wiberg, E.; Nöth, H. Über Wasserstoff-Verbindungen des Thalliums VI. Zur Kenntnis eines Thallium(III)-boranats TlCl(BH4)2. Z. Nat. 1957, 12, 63–65. [Google Scholar]
  259. Wiberg, E.; Schmidt, M. Zur Kenntnis eines Thallium-aluminium-wasserstoffs TlCl(AlH4)2. Z. Nat. 1951, 6, 334–335. [Google Scholar] [CrossRef]
  260. Wiberg, E.; Bauer, R. Zur Kenntnis eines Zinn-aluminium-wasserstoffs Sn(AlH4)4. Z. Nat. 1951, 6, 392. [Google Scholar] [CrossRef]
  261. Weidenthaler, C.; Pommerin, A.; Felderhoff, M.; Sun, W.; Wolverton, C.; Bogdanović, B.; Schüth, F. Complex rare-earth aluminum hydrides: Mechanochemical preparation, crystal structure and potential for hydrogen storage. J. Am. Chem. Soc. 2009, 131, 16735–16743. [Google Scholar] [CrossRef]
  262. Mueller, W.M.; Blackledge, J.P.; Libowitz, G.G. Metal Hydrides; Academic Press: Moscow, Russia, 1968; pp. 435–436. [Google Scholar]
  263. Bergsma, J.; Goedkoop, J.A.; Vucht, J.H.N. Neutron diffraction investigation of solid solutions AlTh2Dn. Acta Cryst. 1961, 14, 223–228. [Google Scholar] [CrossRef]
  264. Sørby, M.H.; Fjellvåg, H.; Hauback, B.C.; Maeland, A.J.; Yartys, V.A. Crystal structure of Th2Al deuterides. J. Alloys Compd. 2000, 309, 154–164. [Google Scholar] [CrossRef]
  265. Vajeeston, P.; Vidya, R.; Ravindran, P.; Fjellvåg, H.; Kjekshus, A.; Skjeltorp, A. Electronic structure, phase stability, and chemical bonding in Th2Al and Th2AlH4. Phys. Rev. B 2002, 65, 075101. [Google Scholar] [CrossRef]
  266. Rude, L.H.; Nielsen, T.K.; Ravnsbæk, D.B.; Bösenberg, U.; Ley, M.B.; Richter, B.; Arnbjerg, L.M.; Dornheim, M.; Filinchuk, Y.; Besenbacher, F.; et al. Tailoring properties of borohydrides for hydrogen storage: A review. Phys. Status Solidi A 2011, 208, 1754–1773. [Google Scholar] [CrossRef]
  267. Graetz, J.; Lee, Y.; Reilly, J.J.; Park, S.; Vogt, T. Structures and thermodynamics of the mixed alkali alanates. Phys. Rev. B 2005, 71, 184115. [Google Scholar] [CrossRef] [Green Version]
  268. Løvvik, O.M.; Swang, O.; Opalka, S.M. Modeling alkali alanates for hydrogen storage by density-functional band-structure calculations. J. Mater. Res. 2005, 20, 3199–3213. [Google Scholar] [CrossRef] [Green Version]
  269. Majzoub, E.H.; Ozoliņš, V. Prototype electrostatic ground state approach to predicting crystal structures of ionic compounds: Application to hydrogen storage materials. Phys. Rev. B 2008, 77, 104115. [Google Scholar] [CrossRef]
  270. Claudy, P.; Bonnetot, B.; Bastide, J.P.; Letoffe, J.M. Reactions of lithium and sodium aluminum hydride with sodium or lithium hydride. Preparation of a new alumino-hydride of lithium and sodium LiN2AlH6. Mater. Res. Bull. 1982, 17, 1499–1504. [Google Scholar] [CrossRef]
  271. Huot, J.; Boily, S.; Güther, V.; Schulz, R. Synthesis of Na3AlH6 and Na2LiAlH6 by mechanical alloying. J. Alloys Compd. 1999, 383, 304–306. [Google Scholar] [CrossRef]
  272. Okada, N.; Genma, R.; Nishi, Y.; Uchida, H.H. RE-oxide doped alkaline hydrogen storage materials prepared by mechanical activation. J. Mater. Sci. 2004, 39, 5503–5506. [Google Scholar] [CrossRef]
  273. Genma, R.; Uchida, H.H.; Okada, N.; Nishi, Y. Hydrogen reactivity of Li-containing hydrogen storage materials. J. Alloys Compd. 2003, 356, 358–362. [Google Scholar] [CrossRef]
  274. Wang, F.; Liu, Y.; Gao, M.; Luo, K.; Pan, H.; Wang, Q. Formation reactions and the thermodynamics and kinetics of dehydrogenation reaction of mixed alanate Na2LiAlH6. J. Phys. Chem. C 2009, 113, 7978–7984. [Google Scholar] [CrossRef]
  275. Genma, R.; Okada, N.; Sobue, T.; Uchida, H.H. Mechanically milled alanates as hydrogen storage materials. Int. J. Hydrogen Energy 2006, 31, 309–311. [Google Scholar] [CrossRef]
  276. Fossdal, A.; Brinks, H.W.; Fonneløp, J.E.; Hauback, B.C. Pressure-composition isotherms and thermodynamic properties of TiF3-enhanced Na2LiAlH6. J. Alloys Compd. 2005, 397, 135–139. [Google Scholar] [CrossRef]
  277. Fan, X.; Xiao, X.; Chen, L.; Li, S.; Ge, H.; Wang, Q. Direct synthesis and hydrogen storage behaviors of nanocrystalline Na2LiAlH6. J. Mater. Sci. 2011, 46, 3314–3318. [Google Scholar] [CrossRef]
  278. Teprovich, J.A., Jr.; Zhang, J.; Colón-Mercado, H.; Cuevas, F.; Peters, B.; Greenway, S.; Zidan, R.; Latroche, M. Li-driven electrochemical conversion reaction of AlH3, LiAlH4, and NaAlH4. J. Phys. Chem. C 2015, 119, 4666–4674. [Google Scholar] [CrossRef]
  279. Nakamura, Y.; Fossdal, A.; Brinks, H.W.; Hauback, B.C. Characterization of Al–Ti phases in cycled TiF3-enhanced Na2LiAlH6. J. Alloys Compd. 2006, 416, 274–278. [Google Scholar] [CrossRef]
  280. Ma, X.Z.; Martinez-Franco, E.; Dornheim, M.; Klassen, T.; Bormann, R. Catalyzed Na2LiAlH6 for hydrogen storage. J. Alloys Compd. 2005, 404–406, 771–774. [Google Scholar] [CrossRef]
  281. Fonneløp, J.E.; Løvvik, O.M.; Sørby, M.H.; Brinks, H.W.; Hauback, B.C. Adjustment of the decomposition path for Na2LiAlH6 by TiF3 addition. Int. J. Hydrogen Energy 2011, 36, 12279–12285. [Google Scholar] [CrossRef]
  282. Løvvik, O.M.; Swang, O. Crystal structures and electronic structures of alkali aluminohexahydrides from density functional calculations. J. Alloys Compd. 2005, 404–406, 757–761. [Google Scholar] [CrossRef]
  283. Løvvik, O.M.; Swang, O. Structure and stability of possible new alanates. Europhys. Lett. 2004, 67, 607–613. [Google Scholar] [CrossRef]
  284. Brinks, H.W.; Huback, B.C.; Jensen, C.M.; Zidan, R. Synthesis and crystal structure of Na2LiAlD6. J. Alloys Compd. 2005, 392, 27–30. [Google Scholar] [CrossRef]
  285. Liu, Y.; Wang, F.; Cao, Y.; Gao, M.; Pan, H.; Wang, Q. Mechanisms for the enhanced hydrogen desorption performance of the TiF4-catalyzed Na2LiAlH6 used for hydrogen storage. Energy Environ. Sci. 2010, 3, 645–653. [Google Scholar] [CrossRef]
  286. Rönnebro, E.; Majzoub, E.H. Crystal structure, Raman spectroscopy and Ab initio calculations of a new bialkali alanate K2LiAlH6. J. Phys. Chem. B 2006, 110, 25686–25691. [Google Scholar] [CrossRef] [PubMed]
  287. Bulychev, B.M.; Semenenko, K.N.; Bitcoev, K.B. Synthesis and investigation of complex compounds of magnesium alanate. Koord Khim 1978, 4, 374–380. [Google Scholar]
  288. Grove, H.; Brinks, H.W.; Heyn, R.H.; Wu, F.-J.; Opalka, S.M.; Tang, X.; Laube, B.L.; Hauback, B.C. The structure of LiMg(AlD4)3. J. Alloys Compd. 2008, 455, 249–254. [Google Scholar] [CrossRef]
  289. Grove, H.; Løvvik, O.M.; Huang, W.; Opalka, S.M.; Heyn, R.H.; Hauback, B.C. Decomposition of lithium magnesium aluminum hydride. Int. J. Hydrogen Energy 2011, 36, 7602–7611. [Google Scholar] [CrossRef]
  290. Grove, H.; Brinks, H.W.; Løvvik, O.M.; Heyn, R.H.; Hauback, B.C. The crystal structure of LiMgAlD6 from combined neutron and synchrotron X-ray powder diffraction. J. Alloys Compds. 2008, 460, 64–68. [Google Scholar] [CrossRef]
  291. Hudson, M.S.L.; Raghubanshi, H.; Pukazhselvan, D.; Srivastava, O.N. Effects of helical GNF on improving the dehydrogenation behavior of LiMg(AlH4)3 and LiAlH4. Int. J. Hydrogen Energy 2010, 35, 2083–2090. [Google Scholar] [CrossRef]
  292. Liu, D.M.; Qian, Z.X.; Si, T.Z.; Zhang, Q.A. Synthesis, crystal structure and thermal decomposition of LiCa(AlH4)3. J. Alloys. Compd. 2012, 520, 202–206. [Google Scholar] [CrossRef]
  293. Wang, H.C.; Zheng, J.; Wu, D.H.; Wei, L.T.; Tang, B.Y. Crystal feature and electronic structure of novel mixed alanate LiCa(AlH4)3: A density functional theory investigation. RSC Adv. 2015, 5, 16439–16445. [Google Scholar] [CrossRef]
  294. Sato, T.; Takagi, S.; Deledda, S.; Hauback, B.C.; Orimo, S.I. Extending the applicability of the Goldschmidt tolerance factor to arbitrary ionic compounds. Sci. Rep. 2016, 6, 23592. [Google Scholar] [CrossRef]
  295. Sørby, M.H.; Brinks, H.W.; Fossdal, A.; Thorshaung, K.; Hauback, B.C. The crystal structure and stability of K2NaAlH6. J. Alloys Compd. 2006, 415, 284–287. [Google Scholar] [CrossRef]
  296. Bhatnagar, A.; Pandey, S.K.; Shahi, R.R.; Hudson, M.S.L.; Shaz, M.A.; Srivastava, O.N. Synthesis, characterization and hydrogen sorption studies of mixed sodium-potassium alanate. Cryst. Res. Technol. 2013, 48, 520–531. [Google Scholar] [CrossRef]
  297. Pan, R.K.; Yao, J.G.; Ji, R.L.; Liu, W.W.; Yin, D.F. First principles study on elastic and electronic properties of bialkali alanates M2M’AlH6. Int. J. Hydrogen Energy 2018, 43, 3862–3870. [Google Scholar] [CrossRef]
  298. Fonneløp, J.E.; Corno, M.; Pinatel, E.R.; Sørby, M.H.; Ugliengo, P.; Baricco, M.; Hauback, B.C. Experimental and computational investigations on the AlH3/AlF3 system. J. Alloys Compd. 2011, 509, 10–14. [Google Scholar] [CrossRef]
  299. Yin, L.C.; Wang, P.; Kang, X.D.; Sun, C.H.; Cheng, H.M. Functional anion concept: Effect of fluorine anion on hydrogen storage of sodium alanate. Phys. Chem. Chem. Phys. 2007, 9, 1499–1502. [Google Scholar] [CrossRef] [PubMed]
  300. Eigen, N.; Bösenberg, U.; Bellosta von Colbe, J.; Jensen, T.R.; Cerenius, Y.; Dornheim, M.; Klassen, T.; Bormann, R. Reversible hydrogen storage in NaF-Al composites. J. Alloys Compd. 2009, 477, 76–80. [Google Scholar] [CrossRef]
  301. Brinks, H.W.; Fossdal, A.; Hauback, B.C. Adjustment of the stability of complex hydrides by anion substitution. J. Phys. Chem. C 2008, 112, 5658–5661. [Google Scholar] [CrossRef]
  302. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds. Part A: Theory and Applications in Inorganic Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009; pp. 192–236. [Google Scholar]
  303. Biliskov, N. Infrared spectroscopy as a convenient tool for investigation of hydrogen sorption mechanisms and bonding in complex hydrides. In Infrared Spectroscopy: Theory, Advances and Development; Cozzolino, D., Ed.; Nova Science Publishers: Hauppauge, NY, USA, 2014; pp. 1–42. [Google Scholar]
  304. Zavorotynska, O.; Corno, M.; Damin, A.; Spoto, G.; Ugliengo, P.; Barico, M. Vibrational properties of MBH4 and MBF4 crystals (M = Li, Na, K): A combined DFT, infrared, and Raman study. J. Phys. Chem. C 2011, 115, 18890–18900. [Google Scholar] [CrossRef]
  305. Tsumuraya, T.; Shishidou, T.; Oguchi, T. Ab initio study on the electronic structure and vibration modes of alkali and alkaline-earth amides and alanates. J. Phys. 2009, 21, 185501. [Google Scholar] [CrossRef]
  306. Parker, S.F. Spectroscopy and bonding in ternary metal hydride complexes-potential hydrogen storage media. Coord. Chem. Rev. 2010, 254, 215–234. [Google Scholar] [CrossRef]
  307. Huheey, J.E.; Keiter, E.A.; Keiter, R.L. Inorganic Chemistry. Principles of Structure and Reactivity; Harper Collins: New York, NY, USA, 1993; pp. 187–190. [Google Scholar]
  308. Li, L.; Qiu, F.; Wang, Y.; Xu, Y.; An, C.; Liu, G.; Jiao, L.; Yuan, H. Enhanced hydrogen storage properties of TiN–LiAlH4 composite. Int. J. Hydrogen Energy 2013, 38, 3695–3701. [Google Scholar] [CrossRef]
  309. Zang, L.; Cai, J.; Zhao, L.; Gao, W.; Liu, J.; Wang, Y. Improved hydrogen storage properties of LiAlH4 by mechanical milling with TiF3. J. Alloys Compd. 2015, 647, 756–762. [Google Scholar] [CrossRef]
  310. Ares Fernandez, J.R.; Aguey-Zinsou, F.; Elsaesser, M.; Ma, X.Z.; Dornheim, M.; Klassen, T.; Bormann, R. Mechanical and thermal decomposition of LiAlH4 with metal halides. Int. J. Hydrogen Energy 2007, 32, 1033–1040. [Google Scholar] [CrossRef]
  311. Liu, S.S.; Sun, L.-X.; Zhang, Y.; Xu, F.; Zhang, J.; Chu, H.-L.; Fan, M.-Q.; Zhang, T.; Song, X.-Y.; Grolier, J.P.E. Effect of ball milling time on the hydrogen storage properties of TiF3-doped LiAlH4. Int. J. Hydrogen Energy 2009, 34, 8079–8085. [Google Scholar] [CrossRef]
  312. Li, Z.; Zhai, F.; Wan, Q.; Liu, Z.; Shan, J.; Li, P.; Volinsky, A.A.; Qu, X. Enhanced hydrogen storage properties of LiAlH4 catalyzed by CoFe2O4 nanoparticles. RSC Adv. 2014, 4, 18989–18997. [Google Scholar] [CrossRef]
  313. Li, Z.; Li, P.; Wan, Q.; Zhai, F.; Liu, Z.; Zhao, K.; Wang, L.; Lü, S.; Zou, L.; Qu, X.; et al. Dehydrogenation improvement of LiAlH4 catalyzed by Fe2O3 and Co2O3 nanoparticles. J. Phys. Chem. C 2013, 117, 18343–18352. [Google Scholar] [CrossRef]
  314. Cheng, H.; Xu, L.; Fan, X.; Huang, X.; Liu, J.; Yan, K.; Zhang, Y. Synergistic effects played by CMK-3 and NbF5 Co-additives on de/re-hydrogenation performances of NaAlH4. Int. J. Hydrogen Energy 2018, 43, 9705–9712. [Google Scholar] [CrossRef]
  315. Huang, Y.; Li, P.; Wan, Q.; Zhang, J.; Li, Y.; Li, R.; Dong, X.; Qu, X. Improved dehydrogenation performance of NaAlH4 using NiFe2O4 nanoparticles. J. Alloys Compd. 2017, 709, 850–856. [Google Scholar] [CrossRef]
  316. Kumar, S.; Kain, V.; Kojima, Y. Remarkably improved dehydrogenation of ZrCl4 doped NaAlH4 for hydrogen storage application. Int. J. Hydrogen Energy 2017, 42, 15299–15307. [Google Scholar] [CrossRef]
  317. Wan, Q.; Li, P.; Li, Z.; Zhao, K.; Liu, Z.; Wang, L.; Zhai, F.; Qu, X.; Volinsay, A.A. NaAlH4 dehydrogenation properties enhanced by MnFe2O4 nanoparticles. J. Power Sources 2014, 248, 388–395. [Google Scholar] [CrossRef]
  318. Gomes, S.; Renaudin, G.; Hagemann, H.; Yvon, K.; Sulic, M.P.; Jensen, C.M. Effects of milling, doping and cycling of NaAlH4 studied by vibrational spectroscopy and X-ray diffraction. J. Alloys Compd. 2005, 390, 305–313. [Google Scholar] [CrossRef]
  319. Rafi-Ud-Din; Qu, X.; Li, P. Superior catalytic effects of Nb2O5, TiO2, and Cr2O3 nanoparticles in improving the hydrogen sorption properties of NaAlH4. J. Phys. Chem. 2012, 116, 11924–11938. [Google Scholar] [CrossRef]
  320. Franke, I.; Flick, T.; Bauer, H. Hydrogen desorption kinetics of CeCl3-doped sodium aluminum hydride compacts measured by parallel in-situ FTIR-ATR-spectroscopy and gravimetry. Int. J. Hydrogen Energy 2015, 40, 4175–4183. [Google Scholar] [CrossRef]
  321. Bureau, J.C.; Amri, Z.; Létoffé, J.M. Etude comparative des hexahydrido-et des hexadeuteridoaluminates de lithium et de sodium. I-spectres Raman et infrarouge de Li3-et Na3AlH6, et Li3AlD6. Mat. Res. Bull. 1989, 24, 23–31. [Google Scholar] [CrossRef]
  322. Amama, P.B.; Grant, J.T.; Shamberger, P.J.; Voevodin, A.A.; Fisher, T.S. Improved dehydrogenation properties of Ti-doped LiAlH4: Role of Ti precursors. J. Phys. Chem. C 2012, 116, 21886–21894. [Google Scholar] [CrossRef]
  323. Ismail, N.; Aboud, A.A.; Hamdel-Din, A.; Farghali, A.A.; Khedr, M.H. Influence of LiH and Ti metal additives on milling LiAlH4 compound. Int. J. Adv. Res. 2014, 2, 307–316. [Google Scholar]
  324. Wu, X.; Wang, X.; Cao, G.; Li, S.; Ge, H.; Chen, L.; Yan, M. Hydrogen storage properties of LiBH4–Li3AlH6 composites. J. Alloys Compd. 2012, 517, 127–131. [Google Scholar] [CrossRef]
  325. Ismail, M.; Zhao, Y.; Yu, X.; Dou, S.X. Improved hydrogen storage property of LiAlH4 by milling with carbon based additives. Int. J. Electroact. Mater. 2013, 1, 13–22. [Google Scholar]
  326. Halim Yap, F.A.; Ali, N.A.; Idris, N.H.; Ismail, M. Catalytic effect of MgFe2O4 on the hydrogen storage properties of Na3AlH6–LiBH4 composite system. Int. J. Hydrogen Energy 2018, 43, 20882–20891. [Google Scholar] [CrossRef]
  327. Franke, I.; Bauer, H.D.; Scheppat, B. OCM 2013. Optical Characterization of Materials. Conference Proceedings; Beyerer, J., Puente León, F., Längle, T., Eds.; KIT Scientific Publishing: Karlsruhe, Germany, 2013; pp. 78–88. [Google Scholar]
  328. Züttel, A. Hydrogen Storage Methods. Naturwissenschaften 2004, 91, 157–172. [Google Scholar] [CrossRef]
  329. Yoshino, M.; Komiya, K.; Takahasi, Y.; Shinzato, Y.; Yukawa, H.; Morinaga, M. Nature of the chemical bond in complex hydrides, NaAlH4, LiAlH4, LiBH4 and LiNH2. J. Alloys Compds. 2005, 404–406, 185–190. [Google Scholar] [CrossRef]
  330. Claudy, P.; Bonnetot, B.; Letoffe, M.J.; Turk, G. Determination des constantes thermodynamiques des hydrures simples et complexes de l’aluminium. IV. Enthalpie de formation de LiAIH2 et Li3AIH6. Thermochim. Acta. 1978, 27, 213–221. [Google Scholar] [CrossRef]
  331. Smith, M.B.; Bass, G.E., Jr. Heats and free energies of formation of the alkali aluminum hydrides and of cesium hydride. J. Chem. Eng. Data 1963, 8, 342–346. [Google Scholar] [CrossRef]
  332. Sulaiman, N.N.; Ismail, M. Catalytic effect of SrFe12O19 on the hydrogen storage properties of LiAlH4. Int. J. Hydrogen Energy 2017, 42, 19126–19134. [Google Scholar] [CrossRef]
  333. Cai, J.; Zang, L.; Zhao, L.; Liu, J.; Wang, Y. Dehydrogenation characteristics of LiAlH4 improved by in-situ formed catalysts. J. Energy Chem. 2016, 25, 868–873. [Google Scholar] [CrossRef]
  334. Opalka, S.M.; Anton, D.L. First principles study of sodium–aluminum–hydrogen phases. J. Alloys Compd. 2003, 356–357, 486–489. [Google Scholar] [CrossRef]
  335. Lee, B.-M.; Jang, J.-W.; Shim, J.-H.; Cho, Y.W.; Lee, B.-J. Thermodynamic assessment of the NaH ↔ Na3AlH6 ↔ NaAlH4 hydride system. J. Alloys Compd. 2006, 424, 370–375. [Google Scholar] [CrossRef]
  336. Fan, X.; Xiao, X.; Chen, L.; Han, L.; Li, S.; Ge, H.; Wang, Q. Hydriding-dehydriding kinetics and the microstructure of La-and Sm-doped NaAlH4 prepared via direct synthesis method. Int. J. Hydrogen Energy 2011, 36, 10861–10869. [Google Scholar] [CrossRef]
  337. Grochala, W.; Edwards, P.P. Thermal decomposition of the non-interstitial hydrides for the storage and production of hydrogen. Chem. Rev. 2004, 104, 1283–1315. [Google Scholar] [CrossRef]
  338. Palumbo, M.; Torres, F.J.; Ares, J.R.; Pisani, C.; Fernandez, J.F.; Baricco, M. Thermodynamic and ab initio investigation of the Al–H–Mg system. Comput. Coupling Phase Diagr. Thermochem. 2007, 31, 457–467. [Google Scholar] [CrossRef]
  339. Cuevas, F. Advanced Materials and Technologies; Cuevas, F., Burzo, E., Eds.; Springer: Berlin/Heidelberg, Germany, 2018; pp. 252–260. [Google Scholar]
Figure 1. Periodic table of alanates. The reported alanates were collected in this “periodic table”.
Figure 1. Periodic table of alanates. The reported alanates were collected in this “periodic table”.
Materials 12 02724 g001
Figure 2. Mechanical milling main concepts.
Figure 2. Mechanical milling main concepts.
Materials 12 02724 g002
Figure 3. Dehydrogenation pathway of several deuterated alanes (adapted from [62]).
Figure 3. Dehydrogenation pathway of several deuterated alanes (adapted from [62]).
Materials 12 02724 g003
Figure 4. Crystal structure of several alanes and their calculated diffraction patterns (λ = Cukα1).
Figure 4. Crystal structure of several alanes and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g004
Figure 5. Crystal structure of lithium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 5. Crystal structure of lithium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g005
Figure 6. Phase diagram of LiH/Al/H2 and Li3AlH6. The blue line represents the equilibrium. Data adapted from reference [86]: ln ( p ) = 0.22 R T + 13.89 ; where (in the original formula) p is in atm, T in Kelvin and Δ H R = 0.22   eV . For visual reference (bottom and right) the equilibrium of Ti-doped Na3AlH6 (blue zone) and NaH + Al (yellow zone) phases were included [88].
Figure 6. Phase diagram of LiH/Al/H2 and Li3AlH6. The blue line represents the equilibrium. Data adapted from reference [86]: ln ( p ) = 0.22 R T + 13.89 ; where (in the original formula) p is in atm, T in Kelvin and Δ H R = 0.22   eV . For visual reference (bottom and right) the equilibrium of Ti-doped Na3AlH6 (blue zone) and NaH + Al (yellow zone) phases were included [88].
Materials 12 02724 g006
Figure 7. Crystal structure of sodium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 7. Crystal structure of sodium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g007
Figure 8. Phase diagram of Ti-doped (Ti(OBu)4) NaAlH4, Na3AlH6, and NaH + Al. Na3AlH6/NaAlH4: ln ( p e q p ) = 37   kJ · mol 1 R T + 122   J · mol 1   K 1 R . NaH and Al/Na3AlH4: ln ( p e q p ) = 47   kJ · mol 1 R T + 126   J · mol 1   K 1 R [88,133].
Figure 8. Phase diagram of Ti-doped (Ti(OBu)4) NaAlH4, Na3AlH6, and NaH + Al. Na3AlH6/NaAlH4: ln ( p e q p ) = 37   kJ · mol 1 R T + 122   J · mol 1   K 1 R . NaH and Al/Na3AlH4: ln ( p e q p ) = 47   kJ · mol 1 R T + 126   J · mol 1   K 1 R [88,133].
Materials 12 02724 g008
Figure 9. Crystal structure of potassium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 9. Crystal structure of potassium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g009
Figure 10. Phase diagram of KAlH4, K3AlH6, and KH + Al. Constructed with data of reference [167] K3AlH6/KAlH4: ln ( p e q p ) = 70   kJ · mol 1 R T + 130   J · mol 1   K 1 R . KH and Al/K3AlH4: ln ( p e q p ) = 81   kJ · mol 1 R T + 130   J · mol 1   K 1 R .
Figure 10. Phase diagram of KAlH4, K3AlH6, and KH + Al. Constructed with data of reference [167] K3AlH6/KAlH4: ln ( p e q p ) = 70   kJ · mol 1 R T + 130   J · mol 1   K 1 R . KH and Al/K3AlH4: ln ( p e q p ) = 81   kJ · mol 1 R T + 130   J · mol 1   K 1 R .
Materials 12 02724 g010
Figure 11. Crystal structure of rubidium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 11. Crystal structure of rubidium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g011
Figure 12. Crystal structure of cesium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 12. Crystal structure of cesium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g012
Figure 13. Crystal structure of beryllium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 13. Crystal structure of beryllium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g013
Figure 14. Crystal structure of magnesium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 14. Crystal structure of magnesium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g014
Figure 15. Crystal structure of calcium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 15. Crystal structure of calcium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g015
Figure 16. Phase diagram of CaAlH5, CaH2, and Al. CaH2 and Al/CaAlH5: ln ( p e q p ) = 32   kJ · mol 1 R T + 130   J · mol 1   K 1 R [224].
Figure 16. Phase diagram of CaAlH5, CaH2, and Al. CaH2 and Al/CaAlH5: ln ( p e q p ) = 32   kJ · mol 1 R T + 130   J · mol 1   K 1 R [224].
Materials 12 02724 g016
Figure 17. Crystal structure of strontium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 17. Crystal structure of strontium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g017
Figure 18. Production of BaAlH4 or B2AlH7 from hydrogenation of Ba7Al13 or Ba4Al5.
Figure 18. Production of BaAlH4 or B2AlH7 from hydrogenation of Ba7Al13 or Ba4Al5.
Materials 12 02724 g018
Figure 19. Crystal structure of barium alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 19. Crystal structure of barium alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g019
Figure 20. Crystal structure of lanthanides alanates and their calculated diffraction patterns (λ = Cukα1).
Figure 20. Crystal structure of lanthanides alanates and their calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g020
Figure 21. Crystal structure of Na-Li alanate and its calculated diffraction patterns (λ = Cukα1).
Figure 21. Crystal structure of Na-Li alanate and its calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g021
Figure 22. Phase diagram of 2NaH + LiH + Al vs. Na2LiAlH6. Data adapted from references [274,285], ln ( p ) = 7685.3 T + 18.3 for un-catalyzed material, and ln ( p ) = 6894.9 T + 17.0 for material catalyzed with TiF4. In the original formulae, p is in atm, and T in Kelvin.
Figure 22. Phase diagram of 2NaH + LiH + Al vs. Na2LiAlH6. Data adapted from references [274,285], ln ( p ) = 7685.3 T + 18.3 for un-catalyzed material, and ln ( p ) = 6894.9 T + 17.0 for material catalyzed with TiF4. In the original formulae, p is in atm, and T in Kelvin.
Materials 12 02724 g022
Figure 23. Crystal structure of Li-K alanate and its calculated diffraction patterns (λ = Cukα1).
Figure 23. Crystal structure of Li-K alanate and its calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g023
Figure 24. Crystal structure of Li-Mg alanates and its calculated diffraction patterns (λ = Cukα1)
Figure 24. Crystal structure of Li-Mg alanates and its calculated diffraction patterns (λ = Cukα1)
Materials 12 02724 g024
Figure 25. Crystal structure of Li-Ca mixted alanate and its calculated diffraction patterns (λ = Cukα1).
Figure 25. Crystal structure of Li-Ca mixted alanate and its calculated diffraction patterns (λ = Cukα1).
Materials 12 02724 g025
Figure 26. Crystal structure of Na-K mixed alanate and its calculated diffraction pattern (λ = Cukα1).
Figure 26. Crystal structure of Na-K mixed alanate and its calculated diffraction pattern (λ = Cukα1).
Materials 12 02724 g026
Figure 27. Normal modes of vibration of tetrahedral [AlH4]. Adapted from reference [302].
Figure 27. Normal modes of vibration of tetrahedral [AlH4]. Adapted from reference [302].
Materials 12 02724 g027
Figure 28. Normal modes of vibration of octahedral [AlH6]3−. Adapted from reference [302].
Figure 28. Normal modes of vibration of octahedral [AlH6]3−. Adapted from reference [302].
Materials 12 02724 g028
Figure 29. Most intense peak of infrared vibrations in the group 1 alanates, MAlH4. (a) Stretching, (b) Bending.
Figure 29. Most intense peak of infrared vibrations in the group 1 alanates, MAlH4. (a) Stretching, (b) Bending.
Materials 12 02724 g029
Figure 30. Most intense peak of infrared vibrations in the group 1 intermediaries, M3AlH6. The red dots are an extrapolation based on the fitted curve.
Figure 30. Most intense peak of infrared vibrations in the group 1 intermediaries, M3AlH6. The red dots are an extrapolation based on the fitted curve.
Materials 12 02724 g030
Figure 31. Most intense Raman peak in the group 1 alanates. (a) Stretching and (b) Bending modes. The red dots are an extrapolation based on the fitted curve.
Figure 31. Most intense Raman peak in the group 1 alanates. (a) Stretching and (b) Bending modes. The red dots are an extrapolation based on the fitted curve.
Materials 12 02724 g031
Table 1. Crystallographic data of alanes.
Table 1. Crystallographic data of alanes.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
α-AlD3(R-3c) No. 167 [66]
a = 4.227; b = 4.227; c = 11.244
α = β = 90; γ = 120
Al: 0, 0, 0
D: 0.63, 0, ¼
α’-AlD3(Cmcm) No. 63 [47]
a = 6.470(3); b = 11.117(5); c = 6.562(2)
α = β = 90; γ = 90
Al1: 0, ½, 0
Al2: ¼, ¼, 0
D1: 0, 0.197(2), 0.451(4)
D2: 0.312(2), 0.1000(14), 0.047(3)
D3: 0, 0.465(3), ¼
D4: 0.298(4), 0.277(2), ¼
β-AlD3(Fd-3m) No. 227 [67]
a = 9.0037(1); b = 9.0037(1); c = 9.0037(1)
α = β = γ = 90
Al1: ½, 0, 0
D1: 0.4301(1); 0.125; 0.125
γ-AlD3(Pnnm) No. 58 [68]
a = 7.3360(3); b = 5.3672(2); c = 5.7562(1)
α = β = γ = 90
Al1: 0, 0, 0
Al2: 0.4174(5), 0.7127(6), 0
D1: 0.2044(9), 0.8269(11), 0
D2: 0.3668(10), 0.3931(13), 0
D3: 0, ½, ½
D4: 0.4174(6), 0.7038(8), 0.3009(6)
Table 2. Crystallographic data of Li-alanates.
Table 2. Crystallographic data of Li-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
α-LiAlD4(P21/c) No. 14 [78]
a = 4.8254(1); b = 7.8040(1); c = 7.8968(1)
α = 90; β = 112.268(1); γ = 90
Al: 0.1428(2), 0.2013(1), 0.9311(1)
Li: 0.5601(12), 0.4657(6), 0.8236(6)
D1: 0.1902(10), 0.0933(8), 0.7710(6)
D2: 0.3526(10), 0.3726(7), 0.9769(6)
D3: 0.2384(11), 0.0840(7), 0.1141(7)
D4: 0.8024(14), 0.2644(7), 0.8689(8)
β-LiAlH4(I 41/a) No. 88 [76]
a = 4.6611; b = 4.6611; c = 10.5219
α = β = γ = 90
Li: 0, ¼, 0.625
Al: 0, ¼, 0.125
H: 0.2527, 0.4237, 0.5413
γ-LiAlH4(Pnma) No. 62 [76]
a = 6.4667; b = 5.3478; c = 6.5931
α = β = γ = 90
Li: 0.2428, ¼, 0.2467
Al: 0.513, ¼, 0.8221
H1: 0.3067, ¼, 0.9617
H2: 0.7162, ¼, 0.9631
H3:0.4889, 0.9833, 0.2943
Li3AlH6(R-3) No. 148 [81]
a = 8.0389(2); b = 8.0389(2); c = 9.4755(5)
α = β = 90, γ = 120
Al1: 0, 0, 0
Al2: 0, 0, ½,
Li: 0.966(2), 0.236(3), 0.3007(17)
D1: 0.8325(11), 0.8030(7), 0.1008(6)
D2: 0.1593(10), 0.1799(8), 0.3884(6)
Table 3. Crystallographic data of Na-alanates.
Table 3. Crystallographic data of Na-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
NaAlH4(I 41/a) No. 88 [118]
a = b = 5.020(2); c = 11.330(3)
α = β = γ = 90
Al: 0, 0, 0
Na: 0, 0, ½
H: 0.228(1) 0.117(2) 0.838(9)
Na3AlH6(P 21/c) No. 14 [121]
a = 5.4145(3); b = 5.5402(3); c = 7.7620(4)
α = 90; β = 89.871(4), γ = 90
Al: 0, 0, 0
Na1: 0, 0, ½
Na2: −0.00129(5), 0.46129(4), 0.25008(4)
H1: 0.0918, 0.0352, 0.2207
H2: 0.222, 0.3283, 0.5454
H3:0.1649, 0.2689, 0.95
Na3AlD6(P 21/c) No. 14 [122]
a = 5.390(2); b = 5.514(2); c = 7.725(3)
α = 90; β = 89.86(3), γ = 90
Na1: 0, 0, ½
Na2: −0.006(5), 0.461(4), 0.252(5)
Al: 0, 0, 0
D1:0.091(3), 0.041(3), 0.215(3)
D2: 0.234(3), 0.328(3), 0.544(3)
D3: 0.165(3), 0.266(3), 0.944(3)
Table 4. Crystallographic data of K-alanates.
Table 4. Crystallographic data of K-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
KAlD4(Pnma) No. 62 [173]
a = 8.8514(14); b = 5.8119(8); c = 7.3457(11)
α = β = γ = 90
K: 0.1839(12), ¼, 0.1522(17)
Al: 0.5578(11), ¼, 0.8209(13)
D1: 0.4018(10), ¼, 0.9156(9)
D2: 0.7050(9), ¼, 0.9630(12)
D3: 0.4209(6), 0.9741(8), 0.3098(7)
α-K3AlH6(P 21/c) No. 14 [175]
a = 6.1771; b = 5.8881; c = 8.6431
α = 90; β = 89.3, γ = 90
K1: 0, 0, ½
K2: −0.0058, 0.4828, 0.2544
Al: 0, 0, 0
H1: 0.0617, 0.0089, 0.2042
H2: 0.2799, 0.3136, 0.5349
H3: 0.1786, 0.2281, 0.9652
β-K3AlH6(I 4/mmm) No. 139 [175]
a = b = 4.4441; c = 7.8098
α = β = γ = 90
K1: 0, 0, ½
K2: 0, ½, ¼
Al1: 0, 0, 0
H1: 0, 0, 0.2128
H2: 0.3429, 0, 0
γ-K3AlH6(Pnnm) No. 58 [175]
a = 10.8885; b = 10.2576; c = 2.5538
α = β = γ = 90
K1: 0.2347, 0.03444, 0
K2: 0.55047, ¾, 0
K3: 0.691, 0.2178, 0
Al1: ½, ½, 0
Al2: 0, ½, 0
H1: 0.9388, 0.0715, 0
H2: 0.5928, 0.3931, 0
H3: 0.3085, 0.3814, 0
H4: 0.0632, 0.3708, 0
H5: 0.4194, 0.0352, 0
H6: 0.8387, 0.3512, 0
Table 5. Crystallographic data of Rb-alanates.
Table 5. Crystallographic data of Rb-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
α-RbAlD4(Pnma) No. 62 [176]
a = 9.2862(6); b = 5.9392(3); c = 7.5784(6)
α = β = γ = 90
Rb: 0.1813(4), ¼, 0.1574(7)
Al: 0.5639(6), ¼, 0.8121(7)
D1: 0.4045(7), ¼, 0.9073(7)
D2: 0.6884(7), ¼, 0.9615(8)
D3: 0.4204(4), 0.9691(6), 0.3080(6)
α-RbAlD4(Cmc21) No. 36 [176]
a = 3.9933; b = 14.6472; c = 6.4933
α = β = γ = 90
Rb: ½, 0.6206, 0.2833
Al: ½, 0.1154, 0.7607
D1: ½, 0.7996, 0.0670
D2: ½, 0.1717, 0.9990
D3: ½, 0.5992, 0.7814
D4: ½, 0.9888, 0.1074
Table 6. Crystallographic data of Cs-alanates.
Table 6. Crystallographic data of Cs-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
CsAlD4 (o)(Pnma) No. 62 [185]
a = 9.8847(5); b = 6.15949(29); c = 7.9182(4)
α = β = γ = 90
Al: 0.55462(33), ¼, 0.80887(30)
D1: 0.5755(4), 0.04042(31), 0.69536(32)
D2: 0.6641(6), ¼, 0.9620(5)
D3: 0.4017(5), ¼, 0.8868(6)
Cs: 0.1847(4), ¼, 0.1652(8)
CsAlD4 (t)(I 41/a) No. 88 [185]
a = b = 5.67231(9); c = 14.2823(5)
α = β = γ = 90
Al: 0, ¾, 0.875
Cs: ½, ¾, 0.125
D1: 0.19658(31), 0.7115(9), 0.95567(12)
D2:0.25993(26), 0.7644(19), 0.92159(17)
Table 7. Calculated crystallographic data of Be-alanates.
Table 7. Calculated crystallographic data of Be-alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
α-BeAlH5(P21) No. 4 [192]
a = 4.790; b = 4.324; c = 6.227
α = γ = 90; β = 89.408
Be: 0.002, 0.230, 0.623
Al: 0.243, 0.990, 1.000
H1: 0.247, 0.162, 0.749
H2: 0.001, 0.740, 0.902
H3: 0.501, 0.740, 0.914
H4: 0.240, 0.821, 0.251
H5: 0.890, 0.965, 0.515
β-BeAlH5(C2/c) No. 15 [192]
a = 5.959; b = 7.008; c = 6.241
α = γ = 90; β = 116.205
Be: 0, 0.333, 0.250
Al: 0,0,0
H1: 0, 0.904, 0.250
H2: 0.902, 0.777, 0.881
H3: 0.688, 0.044; 0.913
Table 8. Crystallographic data of magnesium alanates.
Table 8. Crystallographic data of magnesium alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
Mg(AlH4)2(P −3 m 1) No. 164 [202]
a = b = 5.1949(2); c = 5.8537(2)
α = 90; β = 90; γ = 120
Mg: 0, 0, 0
Al: 0.3333, 0.6667, 0.7057(5)
H1: 0.3333, 0.6667, 0.439(2)
H2: 0.1589(14), −0.1589(14), 0.804(2)
MgAlH5(P 21 21 21) No. 19 [206]
a = 4.55; b = 4.26; c = 13.024
α = 90; β = 90; γ = 90
Mg: −0.2504, −0.2466, −0.3204
Al: 0.2486, 0.2528, −0.4083
H1: −0.4756, −0.0559, 0.4069
H2: −0.03, 0.0912, 0.3051
H3: 0.4719, −0.0516, −0.4063
H4: 0.0284, 0.0975, −0.3045
H5: −0.0024, 0.0916, −0.4994
α-MgAlH5(P 21 /c) No. 14 [207]
a = 4.7499; b = 8.8127; c = 6.6281
α = 90; β = 90; γ = 109.75
Mg: 0.527, 0.985, 0.253
Al: 0.092, 0.245, 0.395
H1: 0.400, 0.121, 0.444
H2: 0.349, 0.390, 0.495
H3: 0.121, 0.592, 0.201
H4: 0.197, 0.862, 0.142
H5: 0.130, 0.305, 0.156
β-MgAlH5(Cc) No. 9 [207]
a = 7.8033; b = 5.7251; c = 6.7393
α = 90; β = 90; γ = 115.39
Mg: 0.542, 0.025, 0.257
Al: 0.000, 0.000, 0.000
H1: 0.008, 0.924, 0.256
H2: 0.201, 0.289, 0.034
H3: 0.771, 0.969, 0.882
H4: 0.027, 0.299, 0.979
H5: 0.246, 0.031, 0.130
Table 9. Crystallographic data of calcium alanates.
Table 9. Crystallographic data of calcium alanates.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
Ca(AlD4)2(Pbca) No. 61 [221]
a = 13.4491(27); b = 9.5334(19); c = 9.0203(20)
α = β = γ=90
Ca: 0.8958(1), 0.4662(2), 0.2818(3)
Al1: 0.4389(3), 0.7757(5), −0.0011(8)
Al2: 0.8460(3), 0.1060(4), 0.1839(5)
D1: 0.3710(9), 0.6842(11), 0.1087(12)
D2: 0.5280(8), 0.8546(12), 0.0825(14)
D3: 0.4877(9), 0.6706(12), −0.1183(13)
D4: 0.3647(8), 0.8817(11), −0.0835(13)
D5: 0.8264(10), 0.0829(11), 0.0086(8)
D6: 0.8094(8), 0.2610(8), 0.2337(14)
D7: 0.9590(5), 0.0702(12), 0.2407(16)
D8: 0.7762(9), −0.0075(10), 0.2636(16)
CaAlD5(P 21/c) No. 14 [221]
a = 9.8000(19); b = 6.9081(13); c = 12.4503(23)
α = 90; β = 137.936(4); γ = 90
Ca1: 0.7845(16), 0.2166(19), 0.7382(13)
Ca2: 0.3275(14), 0.2676(16), 0.1816(11)
Al1: 0.8017(15), 0.3097(16), 0.4907(12)
Al2: 0.2071(14), 0.2175(14), 0.8706(11)
D1: 0.0058(17), 0.3009(19), 0.5190(14)
D2: 0.6406(16), 0.4242(18), 0.3076(12)
D3: 0.6070(14), 0.2725(17), 0.4696(13)
D4: 0.7010(18), 0.3865(14), 0.8592(15)
D5: 0.9589(14), 0.1915(15), 0.6767(10)
D6: 0.1259(17), 0.0329(14), 0.9070(13)
D7: 0.1154(19), 0.3773(14), 0.9139(15)
D8: 0.2848(16), 0.0634(15), 0.8156(14)
D9: 0.2612(19), 0.4064(13), 0.8154(13)
D10: 0.4470(13), 0.1884(16), 0.0707(12)
Table 10. Crystallographic data of strontium-aluminum hydrides.
Table 10. Crystallographic data of strontium-aluminum hydrides.
CompoundSpace Group, Cell Dimensions [Å] and Angles [°]Atomic Coordinates
Sr(AlH4)2Pmmn (No. 59) [44]
a = 9.1165(18); b = 5.2164(11); c = 4.3346(8)
α = β = γ = 90
Sr: 0.1958(3), ¼, ¾
Al1: 0.9665(11), ¼, ¼
Al2: 0.37309(11), ¾, ¼
SrAlD5 (experimental)Pbcm (No. 57) [229]
a = 4.6226(10); b = 12.6213(30); c = 5.0321(10)
α = β = γ = 90
Sr: 0.2532(7), 0.8925(3), ¼
Al: 0.3296(11), 0.1597(3), ¼
D1: 0.4366(13), ¼, 0
D2: 0.3461(13), 0.5790(5), ¼
D3: 0.0311(13), 0.7146(3), ¼
D4: 0.1914(7), 0.0718(3), 0.4986(9)
SrAlH5
(theoretical)
P 212121 (No. 19) [192]
a = 12.679; b = 5.200; c = 4.508
α = β = γ =90
Sr: 0.908, 0.104, 0.036
Al: 0.165, 0.117, 0.071
H1: 0.763, 0.859, 0.278
H2: 0.078, 0.337, 0.918
H3: 0.093, 0.860, 0.945
H4: 0.079, 0.114, 0.374
H5: 0.254, 0.116, 0.768
Sr2AlD7I2 (No. 5) [230]
a = 12.552(1); b = 9.7826(8); c = 7.9816(7)
α = γ = 90; β = 100.286(4)
Sr1: 0.0935(3), 0.3289(4), 0.3195(6)
Sr1´: 0.9065(3), 0.6711(4), 0.3195(6)
Sr2: 0.8609(4), 0.0684(4), 0.0882(6)
Sr2´: 0.1391(4), 0.9316(4), 0.4118(6)
Al1: 0.671(1), 0.847(1), 0.232(2)
Al1´: 0.329(1), 0.153(1), 0.268(2)
D1: 0.7494(7), 0.8594(7), 0.077(1)
D1´: 0.2506(7), 0.1406(7), 0.423(1)
D2: 0.6014(7), 0.7106(7), 0.117(1)
D2´: 0.3986(7), 0.2894(7), 0.383(1)
D3: 0.7658(6), 0.7378(8), 0.341(1)
D3´: 0.2342(6), 0.2622(8), 0.159(1)
D4: 0.5885(6), 0.8298(8), 0.379(1)
D4´: 0.4115(6), 0.1702(8), 0.121(1)
D5: 0.7395(6), 0.9919(7), 0.3291(9)
D5´: 0.2605(6), 0.0081(7), 0.1709(9)
D6: 0.5748(6), 0.9558(7), 0.1157(8)
D6´: 0.4252(6), 0.0442(7), 0.3843(8)
D7: 0.4375(6), 0.6037(7), 0.3189(9)
D7´: 0.5625(6), 0.3963(7), 0.1811(9)
SrAl2D2P-3m1 (No. 164) [233]
a = b = 4.5253(1); c = 4.7214(2)
α = γ = 90; β = 120
Sr: 0,0,0
Al: 0.3333, 0.6667, 0.4589(7)
D: 0.3333, 0.6667, 0.0976(4)
Table 11. Crystallographic data of barium-aluminum hydrides.
Table 11. Crystallographic data of barium-aluminum hydrides.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
BaAlH5Pna21 (No. 33) [207]
a = 9.1568; b = 7.0718; c = 5.1039
α = β = γ = 90
Ba: 0.686, 0.156, 0.256
Al: 0.041, 0.846, 0.229
H1: 0.008, 0.946, 0.905
H2: 0.584, 0.844, 0.025
H3: 0.578, 0.786, 0.504
H4: 0.357, 0.695, 0.233
H5: 0.708; 0.545, 0.214
Ba2AlD7I2/a (No. 15) [235]
a = 13.197(3); b = 10.237(2); c = 8.509(2)
α = γ = 90; β = 101.290(9)
Ba1: 0.3459, 0.5848, 0.3249
Ba2: 0.1084, 0.3247, 0.0852,
Al1: 0.927, 0.096, 0.235
D1: 0.004(1), 0.116(1), 0.077(2)
D2: 0.846(1), 0.974(1), 0.135(2)
D3: 0.023(1), 0.999(2), 0.325(2)
D4: 0.844(1), 0.104(2), 0.387(2)
D5: 0.983(1), 0.249(2), 0.324(2)
D6: 0.832(1), 0.207(1), 0.115(2)
D7: 0.693(1), 0.864(1), 0.322(2)
Table 12. Lanthanides-Aluminum Hydrides (MAlH6) relevant data [261].
Table 12. Lanthanides-Aluminum Hydrides (MAlH6) relevant data [261].
MaterialHydrogen ContentHydrogen Release *Decomposition Temperature [°C]Crystal Structure
(R-3m, No.166) [Å]
[wt.%]ExperimentalDFT **
LaAlH63.510.98Beginning 100, ending 240a = 6.4732
c = 6.2765
a = 6.5272(4)
c = 6.3212(7)
H: 0.2149, 0.7851, 0.4904
CeAlH63.490.80First step: Beginning 100, ending 170
Second step: Beginning 180, ending 270
a = 6.4711
c = 6.2527
a = 6.4637(4)
c = 6.2609(7)
H: 0.2147, 0.7853, 0.4910
PrAlH63.470.78a = 6.4217
c = 6.2028
a = 6.4106(7)
c = 6.2118(11)
H: 0.2139, 0.7861, 0.4894
NdAlH63.410.78a = 6.3796
c = 6.1616
a = 6.3846(7)
c = 6.1741(10)
H: 0.2132, 0.7868, 0.4883
* (including NaCl load) ** M = La, Ce, Pd, Nd on 0, 0, 1/2, and Al on 0, 0, 0.
Table 13. Crystallographic data of Europium alanates [44].
Table 13. Crystallographic data of Europium alanates [44].
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
Eu(AlH4)2Pmmn (No. 59)
a: 9.1003(13); b: 5.1912(8); c: 4.2741(5)
Eu: 0.1966(3), 0.25, 0.75
Al: 0.9625(12), 0.25, 0.25
Al: 0.3821(5), 0.75, 0.25
EuAlH5Pnma (No. 62)
a: 12.481(3); b: 5.0103(12); c: 4.5887(11)
Eu: 0.6517(3), 0.25, 0.2016(12)
Al: 0.4105(14), 0.25, 0.586(4)
Table 14. Crystallographic data of Li-Na mixed alanates.
Table 14. Crystallographic data of Li-Na mixed alanates.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
Na2LiAlD6
(experimental)
Fm-3m (No. 225) [284]
a: 7.38484 (5)
Na: 0.25, 0.25, 0.25
Li: 0.5, 0.5, 0.5
Al: 0, 0, 0
D: 0.238(4), 0, 0
Na2LiAlH6
(calculated)
P 21/c (No. 14) [282]
a = 5.165; b = 5.251; c = 7.339
α = 90, β = 90.03, γ = 90
Li: 0, 0, 0.5
Na: 0.99, 0.47, 0.25
Al: 0, 0, 0
H: 0.07, 0.02, 0.23
H: 0.23, 0.3, 0.53
H: 0.2, 0.27, 0.96
Table 15. Crystallographic data of Li-K mixed alanates.
Table 15. Crystallographic data of Li-K mixed alanates.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
K2LiAlH6R-3m (No. 166) [286]
a: 5.62068(8)
c: 27.3986(6)
Li: 0, 0, 0.4036(8)
Al: 0, 0, 0
Al: 0, 0, ½
K: 0, 0, 0.1270(1)
K: 0, 0, 0.2853 (1)
H: 0.096(7), −0.096(7), 0.466(3)
H: 0.205(5), −0.205(5), 0.638(2)
K2LiAlH6Fm-3m (225) [267]
a = 7.9383
K: ¼, ¼, ¼
Li: ½, ½, ½
Al: 0, 0, 0
H: 0.216, 0, 0
K2LiAlH6
(calculated)
P 21/n (No. 14) [282]
a = 5.528
b = 5.536
c = 7.832
α = 90, β = 90.03, γ = 90
K: 0, ½, ¼
Li: 0, 0, ½
Al: 0, 0, 0
H: 0, 0, 0.23
H: 0.27, 0.27, ½
H: 0.23, 0.23, 0
Table 16. Crystallographic data of Li-Mg mixed alanates.
Table 16. Crystallographic data of Li-Mg mixed alanates.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
LiMg(AlD4)3P 21/c (No. 14) [288]
a = 8.37113(16)
b = 8.73910(17)
c = 14.3012(3)
α = γ = 90, β = 124.8308(8)
Mg: 0.6305(6), 0.5292(4), 0.8833(3)
Li: 0.127(3), 0.4720(19), 0.3822(14)
Al1: 0.7615(5), 0.6282(4), 0.1512(3)
Al2: 0.4745(5), 0.8809(4), 0.8581(3)
A13: 0.9593(5), 0.2510(4), 0.4986(3)
D1: 0.6057(14), 0.5722(12), 0.1782(9)
D2: 0.6523(14), 0.5907(11), 0.0190(6)
D3: 0.7843(17), 0.8088(9), 0.1721(10)
D4: 0.9475(12), 0.5201(10), 0.2158(9)
D5: 0.4888(15), 0.7127(10), 0.8153(9)
D6: 0.6918(11), 0.9294(11), 0.9554(8)
D7: 0.3783(15), 0.9895(12), 0.7474(8)
D8: 0.3312(15), 0.8752(13), 0.8981(10)
D9: 0.9500(15), 0.3124(13), 0.3908(8)
D10: 0.7599(14), 0.1597(12), 0.4549(10)
D11: 1.1293(13), 0.1222(10), 0.5635(8)
D12: 0.9941(14), 0.3727(11), 0.5902(7)
LiMgAlD6P321 (No. 150) [290]
a = b = 7.985550(2)
c = 4.378942(7)
α = β = 90, γ = 120
Mg: 1, 0.3570(13), 0
Li: 0, 0.686(6), ½
Al1: 0, 0, 0
Al2: 1/3, 2/3, 0.492(10)
D1: 0.540(3), 0.763(2), 0.278(3)
D2: 0.119(3), 0.576(2), 0.734(3)
D3: 0.904(2), 0.117(2), 0.228(3)
Table 17. Crystallographic data of Li-Ca mixed alanates.
Table 17. Crystallographic data of Li-Ca mixed alanates.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
LiCa(AlH4)3
(experimental)
P63/m (No. 176) [292]
a = b = 8.91978(12); c = 5.8887(7)
α = γ = 90, β = 120
Li: 0, 0, 0
Ca: 2/3, 1/3, ¼
Al: 0.2805(3), 0.9027(4), ¼
LiCa(AlH4)3
(theoretical)
P63/m (No. 176) [293]
a = b = 9.093
c = 5.996
α = γ = 90, β = 120
Li: 0, 0, 0
Ca: 2/3, 1/3, ¼
Al: 0.3, 0.9, ¼
H1: 0.544, 0.501, ¼
H2: 0.807, 0.815, ¼
H3: 0.535, 0.754, 0.029
LiCaAlH6
(theoretical)
P–4 (No. 81) [294]
a = b = 6.6652
c = 16.5607
α = γ = β= 90
Li1: 0, 0, 0
Li2: 0, 0, ½
Li3: ½, ½, 0
Li4: ½, ½, ½
Li5: 0, ½, 0.4843
Li6: 0, ½, 0.0085
Ca1: 0.3119, 0.2730, 0.1937
Ca2: 0.2380, 0.1803, 0.6978
Al1: 0.2812, 0.2452, 0.3777
Al2: 0.2812, 0.2398, 0.8264
H1: 0.4729, 0.2445, 0.3245
H2: 0.2947, 0.0386, 0.4346
H3: 0.1523, 0.0958, 0.2964
H4: 0.1623, 0.4222, 0.3053
H5: 0.2861, 0.4340, 0.4450
H6: 0.2449, 0.0032, 0.5840
H7: 0.0609, 0.3014, 0.8263
H8: 0.2532, 0.4729, 0.9283
H9: 0.3881, 0.3663, 0.7959
H10: 0.2891, 0.0307, 0.8161
H11: 0.2130, 0.1039, 0.9599
H12: 0.2158, 0.4740, 0.0859
Table 18. Crystallographic data of Na-K mixed alanates.
Table 18. Crystallographic data of Na-K mixed alanates.
CompoundSpace Group, Cell Dimensions [Å]Atomic Coordinates
K2NaAlD6Fm-3m (No. 225) [295]
a = b = c = 8.118(1)
α = β = γ = 90
K: ¼, ¼, ¼
Na: ½, ½, ½
Al: 0, 0, 0
D: 0.2167(8), 0, 0
Table 19. Representative infrared frequencies of Al-H bonds reported for different alanates.
Table 19. Representative infrared frequencies of Al-H bonds reported for different alanates.
AlanateMode/Peak Position [cm−1]Comments/Reference
StretchingBendingLibrational
LiAlH41779, 1642885, 811, 715465Pure crystalline material [177]
1800, 1780, 1645890, 810, 700 [306] and Refs. within
1757, 1615900, 830 [310] and Refs. within
Li3AlH61410, 13001000, 960, 854 [321]
1386, 12761000, 950, 850 [310] and Refs. within
Li3AlD61020, 915740, 700, 635 [321]
NaAlH41680900, 811, 730, 680 Pure crystalline material [177]
1680900, 800, 735, 690 [306] and Refs. within
Na3AlH61440, 1290930, 842, 690 [321]
KAlH41715811, 729 Pure crystalline material [177]
RbAlH41715811, 763, 739 Pure crystalline material [177]
1715811, 769, 729 [306] and Refs. within
CsAlH41711741 Pure crystalline material [177], Ref. [306] and Refs. within
Mg(AlH4)21935800, 625 Ref. [306] and Refs. within
642, 1937 [43]
1620, 1700–1800 [39]
2013, 1905, 1850, 716, 663, 620, 360, 302, 282[210]
Ca(AlH4)2600, 1780 [40]
1788816, 653482[306] and Refs. within
Table 20. Representative Raman frequencies of Al-H bonds reported for different alanates.
Table 20. Representative Raman frequencies of Al-H bonds reported for different alanates.
AlanateAssignation/Peaks Position [cm−1] Comments/Reference
CombinationStretchingBendingLibrationalTranslational
LiAlH4 1837, 1762, 1722950, 882, 830, 780, 690510, 438, 322220, 165, 151, 143, 112, 95[306]
Li3AlH62090, 19741604, 13111014, 975577, 510 [321]
Li3AlD61478, 13971137, 940730, 686412, 360 [321]
NaAlH4 1762, 1681, 848, 817, 770521, 429180, 125, 117[306]
Na3AlH6 1556, 1465, 1152, 1070990, 815, 760560, 480 [321]
KAlH4 1779, 1711790 [306]
Mg(AlH4)2 1969, 1944, 1808824, 768, 736 [306]
2077, 1852, 1845, 812, 758, 742, 298, 232, 87[310]
Table 21. Thermodynamic data of alanates.
Table 21. Thermodynamic data of alanates.
AlanateFormation Enthalpy Δ H f 0   [ kJ / mol ] Dehydrogenation Reaction/Dehydrogenation Enthalpy [kJ/mol]Apparent Activation Energy [kJ/mol]
LiAlH4−107.1 [330]
−113.42 [81]
−114.8 [87]
−118.9 [331]
−119 [71]
(15)−10 [26]
−9.79 [81]
102 [75], 103 [332] (pure)
42.6 [73] (TiCl3-1/3AlCl3 2 mol%)
67 [74] (NbF3 1 mol%)
81.5 [333] (FeCl2 2 mol%)
87.4 [308] (TiN 2 mol%)
Li3AlH6−310.89 [81]
−298.5 to −311.0 [81,87,330]
(16)15.72 [81]
25 [26]
54.8 [73] (TiCl3-1/3AlCl3 2 mol%)
77 [74] (NbF3 1 mol%)
NaAlH4−78.9 [334]
−105.6 [268]
−113.0 [331]
−116.3 [335]
(25)36.7 [335]
37 [88] ѳ
36–40.9 [71] ֎ Materials 12 02724 i001
114.2 [336] (pure)
113.8 (NiFe2O4 3 mol%) [315], and (MnFe2O4) [317]
86.4 [336] (LaCl3 2 mol%)
Na3AlH6−238.8 [335]
−172.8 [334]
−260 [337]
(26)69.6 [335]
47 [88] ѳ
46.8–47 [71] ֎ Materials 12 02724 i001
162.6 [336] (pure)
86.4 [336] (LaCl3 2 mol%)
KAlH4−166.6 [331]
−183.7 [161] ֎
−128 [175]
(34)70 [167] Materials 12 02724 i001
~55 [168] ‖ Materials 12 02724 i001
140 [164] (pure) Materials 12 02724 i001
80 [164] (TiCl3 2% mol) Materials 12 02724 i001
K3AlH6−224.7 [175] (35)81 [167] Materials 12 02724 i001?
CsAlH4−164.9 [330](39)??
Mg(AlH4)2−79 [338] (“assessed value”)(42)20.4 [43] (at 0 K, ab-initio)123.8 [199] (pure not milled)
123.6 [195] (with LiCl2)
123 [197] (submicron rods)
82.3 [200], 85.5 [208] (TiF4 doped)
Ca(AlH4)2−214 [192] (45)−7 [224] Materials 12 02724 i001, −7.4 [220]?
CaAlH5−224 [192] (46)26 [41]
32 [224] Materials 12 02724 i001, 31.1 [220]
161 [41], 153.4 [219] (pure)
57.4 [219] (TiF3 10 wt.%)
SrAlH5−248 [192] (54)??
BaAlH5−224 [207] --??
LaAlH6 [261]??~30 ‖, Materials 12 02724 i001?
MAlH6, M=Ce, Pr, Nd [261]?(82)~28-32 ‖, Materials 12 02724 i001?
Eu(AlH4)2 [44]??−4.4 and 57 (for 2 consecutive reactions of hydrogen release, Materials 12 02724 i001).?
Na2LiAlH6−84.5 [268] Materials 12 02724 i001
−55.26 [297] Materials 12 02724 i001 (, 300 K)
−53.5 [267] Materials 12 02724 i001
(88)53.5 [267] Materials 12 02724 i001, 63.8 [274] Materials 12 02724 i001
56.4 [276] Materials 12 02724 i001 TiF3 doped
57.3 [285] TiF4 doped
173 [274]
143.6 [285] TiF4 doped
K2LiAlH6−100.5 [268] Materials 12 02724 i001
−102.42 [297] Materials 12 02724 i001 (, 300 K)
−82 [267] Materials 12 02724 i001
?82 [267] Materials 12 02724 i001?
LiMg(AlH4)3−192.6 [206] (, 0 K)(91)−4.16 [289] Materials 12 02724 i001, −5 [339] Materials 12 02724 i001~66 277
LiMgAlH6−184.8 [206] (, 0 K)(92)8.89 [289] Materials 12 02724 i001, 9 [339] Materials 12 02724 i001?
K2NaAlH6−107.66 [297] Materials 12 02724 i001 (, 300 K)
−97 [267] Materials 12 02724 i001
(96)97 [267] Materials 12 02724 i001
98 [295] TiF3 doped Materials 12 02724 i001
124.43 [296]
88.05 TiF3 catalyzed [296]
CALPHAD DFT Ѳ Ti-doped ֎ (and references within) Materials 12 02724 i001 Explicitly reported per mole of H2, i.e., kJ/mol H2, ? Unknown.

Share and Cite

MDPI and ACS Style

Suárez-Alcántara, K.; Tena-Garcia, J.R.; Guerrero-Ortiz, R. Alanates, a Comprehensive Review. Materials 2019, 12, 2724. https://doi.org/10.3390/ma12172724

AMA Style

Suárez-Alcántara K, Tena-Garcia JR, Guerrero-Ortiz R. Alanates, a Comprehensive Review. Materials. 2019; 12(17):2724. https://doi.org/10.3390/ma12172724

Chicago/Turabian Style

Suárez-Alcántara, Karina, Juan Rogelio Tena-Garcia, and Ricardo Guerrero-Ortiz. 2019. "Alanates, a Comprehensive Review" Materials 12, no. 17: 2724. https://doi.org/10.3390/ma12172724

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