Next Article in Journal
Techno-Economic Comparison of Onshore and Offshore Underground Coal Gasification End-Product Competitiveness
Previous Article in Journal
Diesel/CNG Mixture Autoignition Control Using Fuel Composition and Injection Gap
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage

1
Interdisciplinary Nanoscience Center (iNANO) and Department of Chemistry, Aarhus University, DK-8000 Aarhus, Denmark
2
Department of Physics and Astronomy, Fuels and Energy Technology Institute, Curtin University, GPO Box U1987, Perth, WA 6845, Australia
3
Department of Physics, Chemistry and Pharmacy, University of Southern Denmark, Campusvej 55, 5230 Odense M, Denmark
4
International Research Center for Hydrogen Energy, Kyushu University, Fukuoka 819-0395, Japan
5
WPI International Institute for Carbon Neutral Energy Research (WPI-I2CNER), Kyushu University, Fukuoka 819-0395, Japan
6
Department of Mechanical Engineering, Faculty of Engineering, Kyushu University, Fukuoka 819-0395, Japan
7
Kyushu University Platform of Inter/Transdisciplinary Energy Research, Fukuoka 819-0395, Japan
*
Authors to whom correspondence should be addressed.
Energies 2017, 10(10), 1645; https://doi.org/10.3390/en10101645
Submission received: 18 September 2017 / Revised: 5 October 2017 / Accepted: 12 October 2017 / Published: 18 October 2017

Abstract

:
Hydrogen has a very diverse chemistry and reacts with most other elements to form compounds, which have fascinating structures, compositions and properties. Complex metal hydrides are a rapidly expanding class of materials, approaching multi-functionality, in particular within the energy storage field. This review illustrates that complex metal hydrides may store hydrogen in the solid state, act as novel battery materials, both as electrolytes and electrode materials, or store solar heat in a more efficient manner as compared to traditional heat storage materials. Furthermore, it is highlighted how complex metal hydrides may act in an integrated setup with a fuel cell. This review focuses on the unique properties of light element complex metal hydrides mainly based on boron, nitrogen and aluminum, e.g., metal borohydrides and metal alanates. Our hope is that this review can provide new inspiration to solve the great challenge of our time: efficient conversion and large-scale storage of renewable energy.

1. Introduction

The ultimate challenge of our time is the development of an environmentally friendly energy system based on sustainable renewable energy to replace our fossil fuel dependency. Fossil fuels are burnt and energy release is accompanied by release of carbon dioxide (CO2), nitrogen oxides (NOx), particles etc. and may locally lead to extreme air pollution and smog while climate changes are observed globally. Hence, the great challenge of our time is efficient conversion and large-scale storage of renewable energy for days, weeks, and maybe months [1]. However, renewable energy sources generally vary significantly over time and place [2]. Thus, the realization of this scenario calls for a range of novel ideas, technologies, and a paradigm shift in design and development of novel energy materials with new functionalities. Incremental improvement of known materials never changes the world, but just one new material may, since novel materials often form the basis for technological paradigm shifts. The chemistry of hydrogen is very diverse and a wide range of novel hydrides have been discovered during the past decade with a rich and fascinating structural chemistry and a range of useful properties, e.g., energy storage, which is the topic of this review.
Renewable energy occurs mainly as heat from the sun or electricity from sun or wind. Electricity can be used directly, stored in a battery, or used to split water to hydrogen and oxygen, see Figure 1.
Hydrogen (H2), is the lightest element of all and also has some unique properties among known substances, e.g., the highest gravimetric energy density, the fastest diffusion speed in air, and a low density both as a gas and liquid. The latter makes hydrogen challenging to store in a dense manner. Figure 1 illustrates that maybe in the future, carbon dioxide from the atmosphere is reacted with hydrogen to create ‘carbon neutral’ hydrocarbons or carbohydrates. Oxidation of such hydrocarbons/carbohydrates will not increase the carbon dioxide level in the atmosphere. Sustainability means that all material cycles are closed, including our energy system. Therefore, all materials that we use must be reused as illustrated in Figure 1.
This review reveals metal hydrides as a diverse class of materials with a range of compositions, structures and properties towards multi-functionality. The focus is on light elemental hydrides, consisting mainly of boron, nitrogen or aluminum, which may be used for storage of hydrogen, solar heat or as novel battery materials. It is hoped that this review will provide new inspiration to create new materials and technologies for efficient conversion and large-scale storage of renewable energy.

2. Complex Metal Hydrides for High-Density Hydrogen Storage

Complex metal hydrides (CMHs) formed by light elements, such as boron, nitrogen, or aluminum often have extreme hydrogen densities but also poor thermodynamic and kinetic properties and limited reversibility [1,2,3,4]. Initially, the discovery of reversibility in titanium-catalyzed NaAlH4 initiated a paradigm shift in hydrogen storage research towards complex anions [5]. Hence, research was extended to include metal borohydrides, e.g., LiBH4 and nitrogen-based complex hydrides, e.g., LiNH2 [6,7,8]. The complex hydride anions consist mainly of covalent bonds with well-defined directionality, whereas ionic bonding dominates in the solid state between the complex anions and counter cations, e.g., in LiBH4, NaBH4, and NaAlH4. Solid state CMHs have high volumetric and gravimetric hydrogen densities (see Table 1), which makes them interesting as hydrogen storage materials. However, the poor thermodynamic and kinetic properties only allow hydrogen release and uptake at elevated temperatures and pressures, which makes them unsuitable for portable applications. The hydrogen release pathway from complex hydrides is complicated compared to metallic and ionic hydrides, and thus remains not fully understood.
Amide-hydride composite systems, e.g., LiNH2-LiH, has been investigated as possible candidates for hydrogen storage [7]. The LiNH2-LiH composite has a high calculated hydrogen content of ρm = 10.5 wt % H2 and releases hydrogen according to reaction scheme (1).
LiNH2(s) + 2LiH(s) ↔ Li2NH(s) + LiH(s) + H2(g) ↔ Li3N(s) + 2H2(g)
The decomposition reaction initiates at T = 180 °C, which is lower than those of the individual reactants, i.e., 300 °C for LiNH2 and 600 °C for LiH [7]. Thus, similar composite systems have been considered [9,10,11,12,13], e.g., Mg(NH2)2-2LiH (5.6 wt % H2) with suitable thermodynamic properties (∆Hdes = 39 kJ mol−1 H2, ∆Sdes = 112 J K−1 mol−1 H2), i.e., dehydrogenation may occur below 90 °C at p(H2) = 1 bar [14]. However, experimental results reveal that T > 180 °C is required to achieve hydrogen release [9]. Addition of CaH2 (8 mol %) catalyzes the reaction and decreases the activation energy, i.e., the decomposition begins at 78 °C [15]. Generally, the dehydrogenation of amide-hydride systems suffer from sluggish kinetics that might be caused by interface reactions, nucleation/nuclei growth, and/or diffusion processes. Mechanical milling was found to be effective in increasing the dehydrogenation kinetics, due to the reduced grain size and increased homogeneous distribution of amide and hydride particles [16,17]. An effective catalyst may significantly improve the kinetics. Recently, K2[Zn(NH2)4] was synthesized by mechanochemical milling and the composite K2[Zn(NH2)4]-8LiH showed release of 4.2 wt % H2 below 400 °C [18,19,20]. Remarkably, the composite may be fully hydrogenated in 30 s at 230 °C and p(H2) = 50 bar [18,19,20]. The LiAl(NH2)4-4LiH desorbs 5.0 wt % H2 at 130 °C [21]. However, the instability of LiAl(NH)2, one of the desorption products, prevents rehydrogenation [21]. Metal amides and metal hydroxides may also form solid solutions, such as in the NaNH2-NaOH system [22].

2.1. Complex Aluminum Hydrides

2.1.1. Aluminum Hydride

Aluminum hydride (AlH3, alane), is a promising hydrogen-storage material based on thegravimetric and volumetric hydrogen density of ρm = 10.1 wt % H2 and ρV = 149 g H2/L, respectively. Additionally, alane has a low hydrogen desorption temperature and fast desorption kinetics, whilst aluminum is abundant [32]. Alane is thermodynamically unstable at room temperature due to a positive Gibbs free energy of formation, ΔGf° = 46.5 kJ/mol, however, the crystalline α-AlH3 is kinetically stable at room temperature [33,34,35]. Alane was first synthesized in its pure form in 1947 [36]. Since then, a variety of six polymorphs of alane has been discovered (α, α′, β, γ, δ and ε) [37]. The obtained polymorph is entirely dependent on the chosen synthesis route [38]. The structures of α-, α′-, β-, and γ-AlH3 consist of corner-sharing AlH6 octahedra whereas γ-AlH3 also possesses edge-sharing AlH6 octahedra [37,39,40,41,42,43].
Alane can be produced directly from the elements by hydrogenation of aluminum. However, this synthesis method is impractical as it requires a hydrogen pressure above 25 kbar at room temperature [44]. Another synthesis method is by mechanochemical treatment at liquid nitrogen temperature (cryomilling, T = −196 °C) where a thermodynamically spontaneous metathesis reaction, ΔGr° = −191 kJ/mol, between AlCl3 and LiAlH4 occurs (see reaction scheme (2)), producing several polymorphs of alane (α, α′, β, and γ-AlH3) [40,45,46].
AlCl3(s) + 3LiAlH4(s) → 3LiCl(s) + 4AlH3(s)
Furthermore, ball milling of AlCl3 and LiAlH4 at room temperature under hydrogen pressure (p(H2) > 210 bar) only leads to formation of α-AlH3 [47]. A solid state reaction is facilitated by thermal treatment at T ~ 75 °C followed by solvent extraction of LiCl [48]. The monomeric alane can also be stabilized as a solvate R∙AlH3 (R = Et2O, amines or tetrahydrofuran (THF)) by reaction in donating solvents such as ethers or amines [49,50]. Subsequently, the solvent can be removed from the AlH3 adduct by heating in vacuum. Alane readily reacts with metal hydrides to form other complex hydrides [51], e.g., NaAlH4, which will be discussed in the next section.

2.1.2. Metal Alanates

The potential of complex aluminum hydrides was seriously established with the discovery of titanium-catalyzed NaAlH4 by Bogdanović in 1997 [5]. The system shows a reversible capacity approximately twice of any of the conventional metal hydrides [52]. The complex aluminum hydrides are based on the tetrahydridoaluminate(1−) or hexahydridoaluminate(3−) complex anion, AlH4 or AlH63−, respectively, which is commonly denoted alanates.
In the 1960s it was discovered that a range of alkali metal alanates can be prepared through direct hydrogenation of the alkali metal hydride (MH, M = Li, Na, K, Cs) and aluminum either in the solid state or in a solvent-mediated reaction [53,54]. The reaction is commonly described as a two-step reaction [55,56], see reaction schemes (3) and (4)
3MH(s) + Al(s) + 3/2H2(g) ↔ M3AlH6(s)
M3AlH6(s) + 2Al(s) + 3H2(g) ↔ 3MAlH4(s)
However, the detailed reaction mechanism for different metal alanates is debatable and other intermediates may be involved [55,57].
LiAlH4 (ρm = 10.5 wt % H2) was synthesized in the late 1940s and is widely used as a reducing agent in organic chemistry [58], but it attained increased attention as a hydrogen storage material with the discovery of the titanium-catalyzed NaAlH4. Thermolysis of LiAlH4 is initiated by the melting of LiAlH4, Tmp(LiAlH4) = 125 °C [23], which subsequently decomposes into solid Li3AlH6 and Al accompanied by an exothermic release of hydrogen gas in the temperature range 150–220 °C [59,60]. The exothermic decomposition in this first step hinders the full reversibility of the system. The second decomposition step occurs around 190–260 °C as an endotherm [61,62]. Ball milling metal alanates tends to lower the decomposition temperature as particle size is reduced and defects are introduced. However, LiAlH4 often decomposes during milling with additives, e.g., TiCl3, VCl3, and Fe [63,64,65,66]. Furthermore, addition of TiCl3 only reduces the decomposition temperature but does not enable rehydrogenation [67]. A way to overcome the energy barrier preventing rehydrogenation of LiAlH4 is to use a solvent-mediated synthesis route, hence LiH and Al stirred in, e.g., THF or Me2O and apply hydrogen pressure as low as p(H2) > 10 bar [68,69].
Sodium alanate, NaAlH4 (ρm = 7.4 wt % H2) is the most investigated metal alanate with moderate rehydrogenation conditions of titanium-catalyzed NaAlH4 (at T = 170 °C, p(H2) = 150 bar) [5]. However, the useful capacity is low (5.6 wt % H2) and not ideal for consideration in mobile applications. Thus, a tremendous effort has been carried out in research to tailor the system, e.g., by addition of dopants or nanoconfinement [70,71,72,73,74,75,76,77]. Furthermore, NaAlH4 has also been widely used as a precursor in metathesis reactions to synthesize new metal alanate compounds, e.g., Mg(AlH4)2, Ca(AlH4)2, Sr(AlH4)2, and Eu(AlH4)2 [78,79,80].
Potassium alanate, KAlH4 (ρm = 5.7 wt % H2), has a less understood mechanism of decomposition as compared to LiAlH4 and NaAlH4. The formation of an intermediate has been observed experimentally and a crystal structure and composition of K3AlH6 has been suggested by density functional theory (DFT) calculations [37,81,82,83]. In contrast, experimental observations suggest a KyAlHx compound with 1 ≤ y ≤ 3 and 4 ≤ x ≤ 6 [57]. The main advantage of KAlH4 over LiAlH4 and NaAlH4 is the reversible hydrogen sorption at low pressure (<10 bar) and moderate temperature (250–300 °C) without the need of additives [84].
In general the alkali metal alanates are less stable than the alkali metal borohydrides with decomposition temperatures of 60 to 300 °C as compared to T > 350 °C, respectively [2,67]. Hence, they may more easily become a practical hydrogen storage material, but generally with lower gravimetric hydrogen storage capacity.

2.2. Metal Borohydrides

In 2003, lithium borohydride, LiBH4, was proposed as a hydrogen storage material. Since then, a wide range of novel metal borohydrides has been discovered, due to the great coordination flexibility of the complex BH4 anion, which can act as a counter ion in the solid state, or coordinate to a metal via corner sharing (η1), edge sharing (η2), or face sharing (η3) [3,85]. Metal borohydride structures appear to be related to structures of metal oxides, e.g., polymorphs of Ca(BH4)2 are isostructural to polymorphs of TiO2 [85].

2.2.1. Monometallic Borohydrides

Lithium is one of the lightest elements and LiBH4 has a high gravimetric and volumetric hydrogen density of ρm = 18.4 wt % H2 and ρV = 122.5 g H2/L, respectively [6,8,86,87]. The structure of LiBH4 is orthorhombic (Pnma), o-LiBH4, at room temperature whereas a first-order polymorphic transition into a hexagonal (P63mc) structure, denoted h-LiBH4, occurs at T ~ 108 °C [8,86,88]. Both polymorphs contain relatively rigid, regular, and undistorted BH4 tetrahedra [89]. Furthermore, LiBH4 melts at T ~ 275 °C and eventually decomposes at T > 350 °C into LiH and elemental B where the enthalpy change amounts to ∆Hdec = 74 kJ/mol H2 [25]. The other alkali metal borohydrides MBH4, M = Na, K, Rb and Cs have higher thermal stability and all crystallize in the Rock salt structure. Sodium borohydride is also stable in basic aqueous solutions, which has been used as a hydrogen storage media [90]. Alkali metal borohydrides are known to have the highest thermal stability, which hampers the possible application as hydrogen storage materials.
Only recently, the full series of alkaline earth metal borohydrides was completed with the discovery of Ba(BH4)2, Sr(BH4)2, and their chloride derivatives [91,92,93]. The alkaline earth metal borohydrides show decreasing degree of covalent and directional bonding in the series, Mg > Ca > Sr > Ba. Magnesium borohydride is known in seven different polymorphs and, remarkably, one of these, γ-Mg(BH4)2, has permanent porosity and 30 % of open space in the structure [94,95,96,97,98,99,100,101,102,103]. The heavier alkali earth metal borohydrides form three to four polymorphs [3].
A few transition metal borohydrides have also been described and those that are stable at room temperature have electron configurations d0, d5 or d10 [2], i.e., containing the metal ions, Sc3+, Y3+, Zr4+, Mn2+, Zn2+ or Cd2+. A few less stable transition metal borohydrides are also described, e.g., Ti(BH4)3 (d1) which sublimes at T ~ 0 °C, or Fe(BH4)2 which is stable in solution (at T ~ −30 °C) and can be stabilized as [Fe(NH3)6](BH4)2 [104,105].
Some monometallic borohydrides form molecular solids and sublime at low temperatures, e.g., Al(BH4)3 and Zr(BH4)4, where the latter is the most stable, Tsub = 29 °C and may be prepared by vapour deposition [106].

2.2.2. Bimetallic Borohydrides

Lithium potassium borohydride, LiK(BH4)2, was the first discovered bimetallic borohydride [107,108]. Later, a range of lithium, rubidium, and caesium borohydrides were described while also a solid solution of Na1−xKxBH4 was discovered [85,109]. These compounds have structures similar to their parent monometallic borohydrides.
Bimetallic borohydrides, where the metals have increasing difference in Pauling electronegativity, have structures constructed of larger metal complexes with the most electronegative metal as the central ion and the less electronegative metal with dominantly ionic coordination. Thus, lithium scandium borohydride, LiSc(BH4)4, consists of [Sc(BH4)4] complexes and Li+ ions. Additionally, a relationship between the Pauling electronegativity of the metal cation and the thermal stability of the metal borohydride has been established as higher electronegativity results in lower decomposition temperature, see Figure 2 [2,110,111].
Sodium zinc borohydrides are known in two compositions, NaZn(BH4)3 and NaZn2(BH4)5, which contain the complex ions [Zn(BH4)3] and [Zn2(BH4)5], respectively. The latter can also be described as two identical interpenetrated three-dimensional frameworks [112,113,114]. Strong and directional Zn-BH4 bonding by edge sharing (η2) is observed, which defines the structural architecture. The compound LiCe(BH4)3Cl, has a very unique structure consisting of tetranuclear anionic clusters [Ce4Cl4(BH4)12]4 with a distorted cubane Ce4Cl4 core [3,115].
Perovskite structures have long been known for metal hydrides, e.g., CsCaH3 and NaMgH3 containing the anion H [117,118,119,120,121]. However, only recently the first perovskite-type metal borohydride, KMn(BH4)3, was reported, though followed up by extensive research reporting multiple new perovskite-type metal borohydrides containing the complex anion BH4, see also Figure 3 [116,122,123,124]. Structurally, the metal borohydride perovskites differ from regular oxide and halide ABX3 perovskites as their polymorphic transitions are not always intuitively towards higher symmetry with increasing temperature. H-H repulsion between the BH4 ligands may in some cases result in lower symmetry.[123,125].

2.2.3. Trimetallic Borohydrides

The first trimetallic borohydrides, Li3MZn5(BH4)15, M = Mg and Mn, presented in 2013 [126], were prepared by partial cation substitution in LiZn2(BH4)5 using the chemical knowledge that the ions, Li+, Mg2+, Mn2+, and Zn2+ share some chemical properties and may substitute for each other in the solid state. The two compounds, Li3MZn5(BH4)15, M = Mg and Mn, are isostructural with cation disorder in the hexagonal (P63/mcm) structure related to bimetallic LiZn2(BH4)5 [126]. The five trimetallic compounds ALiM(BH4)4 (A = K or Rb; M = Mg or Mn) and K3Li2Mg2(BH4)9 have negatively charged Li-Mg/Mn layered structures [127]. These trimetallic compounds have interesting structural relationships to aluminosilicates, e.g., KAlSi3O8 (feldspar), and may open new research directions, e.g., preparation of zeolite-like metal borohydrides. Similarly, zeolites contain three-dimensional negatively charged Al-Si-O networks [127]. Double perovskites LiA2M(BH4)6 have been observed for large alkali metals A = Rb, Cs and M = Y, Ce, Gd while the smaller alkali metal A = K stabilizes the garnet-type structure Li3K3M2(BH4)12 for M = La, Ce [128,129].

2.2.4. Metal Borohydrides Modified by Neutral Molecules

Metal borohydrides have been modified with neutral molecules in order to improve hydrogen release and uptake properties [3]. Nitrogen-based molecules have attracted significant attention, e.g., ammonia, NH3, hydrazine, N2H4, and ammonia borane, NH3BH3, in order to introduce dihydrogen bonding via partly positively charged hydrogen atoms, e.g., B–Hδ−H–N in the solid state. Dihydrogen bonding is explored for tailoring thermal stability and decomposition. Trends in composition, structure, and properties for a range of new ammine metal borohydrides, M(BH4)mnNH3, have been investigated recently. The interest in ammine metal borohydrides has increased tremendously during the past ~5 years [3,4]. The number of BH4 anions (m) in the compounds is fixed and determined by the oxidation state of the metal, whereas the number of NH3 ligands (n) may range from 1 to 8, e.g., Ca(BH4)2∙NH3 and Zr(BH4)4∙8NH3 [130,131]. In all cases the ammonia molecule, NH3, coordinates via the lone pair on N to the metal, and the BH4 anions coordinate in a more flexible way (η0–η3). The most extensive series of crystalline ammine metal borohydride is formed by Y(BH4)3nNH3 (n = 1, 2, 4, 5, 6 and 7), which illustrates the great structural flexibility ranging from cation complexes (n = 6 and 7), molecular neutral complexes (n = 4 and 5), one-dimensional chain like (n = 2), and two-dimensional layered structure (n = 1) [132].
Ammine metal borohydrides have been considered promising hydrogen storage materials, e.g., Al(BH4)3∙6NH3, Li2Al(BH4)5∙6NH3, and Zn(BH4)2∙2NH3 release 9 to 12 wt % H2 in the temperature range 115 to 170 °C with traces of NH3 [133,134,135].
Interestingly, ammonia has a destabilizing effect for metal borohydrides with low metal electronegativity (χp < ~1.6) and a stabilizing effect for those with χp > ~1.6. Thus the latter are often more thermally stable than their respective metal borohydride [104,130]. The stabilization may be due to complex formation and a shielding effect of the metal cation, which would otherwise be reduced to the metallic state, e.g., Zn(BH4)2∙4NH3 and Zn(BH4)2∙2NH3 [135].
Previously, a low NH3/BH4 ratio (n/m ~ 1) and strong dihydrogen bonds were assumed to provide H2 rich (and NH3 poor) gas release during thermolysis. However, recent experimental results disagree with this hypothesis [130,132,136]. The composition of the released gas appears to depend on the thermal stability of the ammine metal borohydride, i.e., if it is similar to or higher than the thermal stability of the corresponding metal borohydride, then H2 is released. In contrast, mainly NH3 is released if the thermal stability of the ammine metal borohydride is significantly lower than the corresponding metal borohydride. For instance, the compounds, M(BH4)mnNH3, M = Al, Zn, Zr, and V, release H2 (and maybe some NH3), e.g., Al(BH4)3∙6NH3 releases H2 and a small amount of NH3 at T ~ 165 °C [131,133,135,137].
Ammonia borane metal borohydride compounds have also been investigated, M(BH4)mnNH3BH3, M = Li, Mg, Ca, or Al [138,139,140,141,142]. They tend to separate during thermal treatment and NH3BH3 decomposes at T > 200 °C and the release of the toxic gasses is not suppressed, as observed for metal amidoboranes, M(NH2BH3)m [143]. An interesting amidoborane is the complex Na[Al(NH2BH3)4] which was found to reversibly absorb 1.7 mole of H2, although between two amorphous decomposition states [144]. Replacement of the Na+ ion with a K+ ion results in similar thermal decomposition temperatures. However, the reversibility seem to be disabled by this cation substitution [145].

3. Complex Metal Hydrides for Electrochemical Applications

Ni-metal hydride (Ni-MH) batteries have been used commercially in portable electrical devices since the early 1990’s. The cathode consists of Ni(OH)2, while the anode is a MH and the electrolyte is an alkaline solution [146]. The overall reaction is:
MH + NiOOH ↔ M + Ni(OH)2M: Hydrogen storage alloy
The LaNi5-based AB5 type alloy was modified and commercially used as negative electrodes for Ni-metal hydride batteries [147]. Even though the energy densities of these batteries are smaller than those of Li-ion batteries, they still play an important role in the market, especially for hybrid and fuel cell vehicles. Furthermore, several metal hydrides have been proposed as conversion type electrodes and others as electrolytes for complete solid state batteries. Different properties of metal hydrides have also been highlighted elsewhere [148].

3.1. Metal Hydrides as Electrode Materials

In 2008 use of metal hydrides as conversion type anodes in Li-ion batteries was proposed for the first time by Oumellal et al. [149]. In this concept, hydrides store lithium through the following general conversion reaction:
MHn + nLi → nLiH + M  M: Metal or intermetallic hydride
The first investigation in 2008 focused on MgH2 as the anode material, but the study also showed successful conversion reactions for TiH2, NaH, LaNi4.25Mn0.75H5 and Mg2NiH3.7. The high theoretical capacities of the hydrides (e.g., 2038 mAh/g for MgH2, i.e., more than five times that of graphite) naturally caused significant interest for this new class of anodes. Furthermore, the investigation revealed that magnesium hydride exhibits smaller cell polarization (ΔV between charge and discharge) than any of the other previously investigated conversion type electrodes (e.g., electrodes based on nitrides, sulfides and oxides) [149]. Unfortunately, hydride materials show limitations in their initial capacity (relative to the theoretical values) and large capacity loss upon cycling, which is subscribed to poor electronic conductivity and the large volume change occurring in the electrode during cycling, which may lead to poor contact. The capacity of MgH2 can be improved by limiting the amount of inserted lithium [149], through careful formulation (e.g., choice of binder) and choice of electrolyte. Hereby, a capacity retention for MgH2 of 542 mAh/g over 40 cycles has been achieved [150]. Subsequently, several studies have investigated Li-ion storage in MgH2 [151,152,153,154,155,156,157,158], TiH2 [159,160], AlH3 [151,161], M-LiH (M = Mg, Ti) [162] and TiH2-MgH2 [163] composites as well as a series of Mg-based complex hydrides such as Mg2FeH6, Mg2CoH5 and Mg2NiH4 [164]. For detailed descriptions of these results we refer to other reviews [165,166].
Li-ion storage through conversion reactions in lithium and sodium alanate, LiAlH4 and NaAlH4, were investigated for the first time in 2015 [161,167,168]. The theoretical capacity for these materials is ~2000 mAh/g when considering reaction with 3 Li equivalents and full reduction to LiH and metallic aluminium. Silvestri et al. [168] obtained an initial discharge capacity of 1180 mAh/g for LiAlH4, corresponding to storage of 1.67 Li equivalent, at discharge potentials dominated by extended plateaus at 0.78, 0.26 and 0.16 V versus Li. The conversion reaction has been shown by powder X-ray diffraction to occur through the following steps:
2LiAlH4 + 3Li+ + 3e → Li3AlH6 + Al + 2LiH
Li3AlH6 + 3Li+ + 3e → Al + 6LiH
The formation of Al results in formation of the LiAl alloy according to the following reaction, which is predicted to occur below 0.29 V [168].
Li+ + e + Al → LiAl
The reversible capacity in the first charge is less than 1/3 of the capacity in the first discharge with the alloying/de-alloying reaction being responsible for the majority of the reversible capacity.
For NaAlH4 anodes initial capacities of 1700–1800 mAh/g can be achieved according to Reale et al. [167] with the majority of the Li storage reaction taking place at potentials <0.45 V. High energy ball milling with conductive carbon (Super P) for 15 h improves the reversible capacity in the first cycle from less than 30% to almost 70%. The discharge mechanism is still not completely agreed upon. In general, multiple reaction steps are reported and the bi-alkali hexa-alanate, LiNa2AlH6 is a common intermediate [161,167].
NaAlH4 + 3/2Li+ + 3/2 e → 1/2LiNa2AlH6 + 1/2Al + LiH
LiNa2AlH6 + 5Li+ + 5 e → 2Na + Al + 6LiH
However, Reale et al. [167] also observed formation of the sodium hexa-alanate, Na3AlH6 (Equation (12)), which is suggested to be a competitive process affected by kinetic limitations. Formation of Na3AlH6 during discharge was also confirmed in a separate study through 23Na and 27Al solid state NMR [169].
3NaAlH4 + 6Li + 6 e → Na3AlH6 + 6LiH + 2Al
Reale et al. also used the hexa-alanates Na3AlH6 and LiNa2AlH6 directly as anodes, which were found to behave similarly to the alanates.
Recently, attempts were made to improve the reversibility of NaAlH4 by nanoconfinement in mesoporous carbon scaffolds [170,171]. Nanoconfinement seems to alter the conversion mechanism by enhancing the conversion from NaAlH4 to Na3AlH6. Also it appears to increase the capacity retention of the system, however the results suggests that the majority of the long term capacity stems from the carbon scaffolds [171].
Metal borohydrides have very high theoretical capacities, e.g., 4992 mAh/g for LiBH4 and thus have great potential as future anode materials. However, only a few studies of borohydride-based anodes exists [172,173]. Theoretical calculations predict that Li-conversion reactions are thermodynamically feasible for a long list of metal borohydrides both through a one-step mechanism with direct formation of the metal, elemental boron and lithium hydride, or through a two-step reaction with formation of a metal hydride as an intermediate. In the same investigation, five mono-metallic borohydrides, M(BH4)n (M = Li, Na, K, Mg and Ca) were tested as anodes in Li-ion batteries. Ca(BH4)2 appears to be electrochemically inactive, and for LiBH4 and KBH4 capacities below 75 mAh/g were achieved in the first discharge. Mg(BH4)2 and NaBH4 exhibited initial capacities of ~540 and 250 mAh/g, which is well below their theoretical capacities of 3971 and 2834 mAh/g, respectively. Furthermore, the reversible capacity in the second discharge is less than 50%. Thus, it appears that metal borohydride based systems require further improvements in order to yield results comparable to those obtained for the metal alanates.

3.2. Complex Metal Hydrides as Electrolytes

Metal borohydrides have recently been employed as a new class of solid-state electrolytes for batteries [174,175] and the main advantages of light-weight CMHs are: (i) low material density compared to similar oxides and chalcogenides (can be two to three times lower); (ii) thinner solid state electrolytes, hence (i) and (ii) provide higher battery energy density; (iii) safer batteries because flammable organic ionic liquids and polymeric electrolytes are avoided; and (iv) easier assembly of all-solid-state batteries [176].
Research on metal borohydrides and derivatives, oxides and chalcogenides as solid state electrolytes have been mainly focused on monovalent alkali metals, Li+ and Na+. Indeed, multivalent cations, e.g., Mg2+, Ca2+, and Al3+, may further increase the energy density of solid state batteries. However, development of fast ion conductors for these cations is challenging due to much larger charge density as compared to mono-valent cations. Therefore, very large, low charge density anions, such as closo-boranes, B12H122−, or closo-carboranes, CB11H12, may prove useful for design of new electrolytes.
The high-temperature polymorph h-LiBH4 was among the first metal hydrides to be discovered as fast ion conductors with Li+ conductivity of the order of 10−3 S cm−1 at T > 115 °C [177]. Attempts to stabilize the high-temperature polymorph at room temperature were successful by halide-substitution of the borohydride group, i.e., h-Li(BH4)1−xIx [178,179,180]. Reversible Mg stripping/plating has been demonstrated using Mg(BH4)2 as an electrolyte in THF and dimethoxyethane solutions [181]. However, the conductivity of solid state Mg(BH4)2 is very low (<10−12 S cm−1 at 30 °C) [182] but may be enhanced in double anion compounds, e.g., Mg(BH4)(NH2) has a Mg2+ conductivity of 1 × 10−6 S cm−1 at T = 150 °C [183]. Recently, the addition of ethylene diamine, NH2(CH2)2NH2 (en), to Mg(BH4)2 resulted in the formation of cis-Mg(en)(BH4)2 which shows ion conductivity between 5 × 10−8 and 6 × 10−5 cm−1 in the temperature range 30 to 70 °C, respectively [182]. So far, the most promising compounds are Na2B10H10 and Na2B12H12, which show superionic conductivity in the high-temperature polymorphs (phase transition at 87 and 207 °C, respectively) with Na ion conductivity of ~0.01 and 0.1 S cm−1, respectively [184,185].
Intermolecular anion substitution in metal borohydrides has been explored as a valuable technique for tailoring physical and chemical properties [3]. The heavier halides have provided a wide range of metal borohydride halides with either fully ordered, e.g., KZn(BH4)2Cl or Sr(BH4)Cl [92,186], partly ordered, e.g., NaY(BH4)2−xCl2+x [187,188], or disordered structures, e.g., K2Zn(BH4)xCl4−x [189]. The smaller fluoride ion can substitute for the hydride ion, i.e., intramolecular anion substitution, F → H, in the BH4 complex and the composite NaBH4-NaBF4 provided the first fluorine-substituted borohydride, NaBH2.1F1.9, observed in the temperature range of 200–215 °C [190,191].
Recently, a new class of Li ion conductors was discovered, LiRe(BH4)3Cl, Re = La, Ce, Pr, Nd, Sm, Gd, with an interesting new structure type [115,192,193,194]. The structure reveals disordered Li+ ions that occupy only 2/3 of the 12d Wyckoff sites, but a fully ordered anion lattice. The Li+ ion conductivity for LiCe(BH4)3Cl was measured to be 1.03 × 10−4 S cm−1, at 20 °C [115]. Additionally, solid-state 1H, 11B, and 7Li NMR measurements of spin-lattice relaxation rates of LiLa(BH4)3Cl reveal two types of dynamics on the same frequency scale, i.e., Li-ion diffusion and reorientational motion of BH4 groups. Therefore the Li-ion diffusion and the dynamics of BH4 complexes appear to be correlated, which suggests a paddle wheel mechanism is responsible for fast ionic mobility [195,196]. Anion substitution in this class of materials is also possible, i.e., LiLa(BH4)3X, X = Cl, Br, I, which may open new possibilities for improvement of ion conductivity by tailoring the structural aperture for cation diffusion in the solid materials [197,198].
New types of double anion hydride-oxides, i.e., containing BH4 and BO33−, have also been discovered recently [199,200], which may suggest formation of new types of hydrides containing, also SO42−, PO43−, or PS43−, which may be compatible with known electrode materials.
A new class of silver closo-boranes was discovered recently, Ag2B12H12 and Ag2B10H10, with high ion conductivities. Anion substitution with iodine stabilizes a high temperature polymorph with ion conductivity up to 3.2 mS cm−1 at room temperature [201]. Partly or fully halogenated closo-boranes may in future be useful to create new materials with higher conductivities [202,203].
Substitution of a boron atom with a carbon atom in closo-boranes leads to the formation of closo-carboranes, e.g., CB11H12 with even lower charge density. Particularly, LiCB11H12, and NaCB11H12 have lower polymorphic transition temperatures (127 and 107 °C, respectively) compared to the analogues Li2B12H12 and Na2B12H12 (342 and 256 °C, respectively) while showing some of the most promising Li and Na ion conductivities of >0.1 S cm−1 yet observed [204].

4. Complex Metal Hydrides for Thermal Energy Storage

There are thousands of metals, metal alloys and compounds that can reversibly react with gaseous hydrogen [205] at temperatures as low as −100 °C (TiCr1.9H3.5 [206]) and as high as 1100 °C (LaHx [207]). The absorption of hydrogen is an exothermic process that releases heat while the desorption of hydrogen is an endothermic process that absorbs heat and the direction of the reaction can be controlled by changing either the hydride temperature or hydrogen pressure. This property of metal hydrides means that they can be exploited for a wide range of closed-loop energy storage and energy transformation applications including as; H2 compressors; hydride heat engines, actuators and temperature sensors; thermochemical heat storage; heat pumping and thermally driven refrigeration. In fact, many of the relevant engineering principles were developed and proven during the 1970s and 1980s based on the high-cost, low capacity intermetallic hydrides known at the time. Extensive reviews on the progress of metal hydrides for these applications have been published [208,209,210,211]. However, a consistent road block to the commercialisation of metal hydrides for these applications has been the relatively low H2 capacity and high cost of traditional intermetallic hydrides based on transition metals [210].
With the commercialization of concentrating solar power plants (CSPs) incorporating molten nitrate salts heat storage and, in conjunction with the U.S. Department of Energy SunShot Initiative to drive down the cost of solar electricity [212], there has been a renewed focus on the potential of low-cost, high-temperature metal hydrides to be the second generation of CSP heat storage materials that can operate at temperatures above 600 °C [213,214,215,216,217]. This is because the thermochemical heat storage capacity of metal hydrides can exceed the sensible heat storage capacity of molten salts by a factor of >10 [216]. In fact, the theoretical thermochemical heat storage capacity of some metal hydrides, such as LiH and CaH2, are only exceeded by methane reforming reactions [214,218]. Specific details about the energy cycle for using metal hydrides as high temperature heat storage materials for CSP can be found elsewhere [67,214,219,220,221,222]. Besides the intermetallic hydrides first considered for heat storage in the 1970s and 1980s, most of the research on metal hydrides for high temperature heat storage (T > 400 °C) has focused on a few particular sub-sets. These include: simple binary/ternary hydrides and their partially fluorinated analogues [216], such as MgH2 [223,224,225,226,227], NaH/NaH1−xFx [215,228], NaMgH3/NaMgH2F [229,230], and TiHx [213]; destabilized binary hydrides, such as LiH-Al [231], xLiH-Si [232,233], and CaH2-Al [234] and; complex transition metal hydrides, such as Mg2NiH4 [235,236], Mg2FeH6 [236,237], Mg2CoH5, and Mg6CoH11 [236]. The high thermodynamic stability in conjunction with their high hydrogen capacity make CMHs intriguing candidates as thermochemical heat storage materials for CSP and for industrial waste heat recovery and utilization. Despite this, CMHs have rarely been considered for heat storage applications with examples in the literature limited to: the Li2NH/Li4NH system [238]; Na3AlH6, Na2LiAlH6 and K2NaAlH6 [215] and; 6LiBH4 + CaH2 [239].
For this review, we have used the thermodynamic data reported in the literature for the H2 absorption/desorption from CMHs to calculate properties relevant for heat storage, Figure 4a,b and Table 2. Figure 4a shows the theoretical gravimetric and volumetric heat storage capacities of 15 CMH systems compared to other well-known: (1) thermochemical heat storage materials (TCMs) based on metal hydrides, carbonates, hydroxides and oxides; (2) phase change materials (PCMs) and; (3) sensible heat storage materials (SHMs). Another important factor in determining the suitability of a metal hydride for a particular application is the H2 equilibrium pressure which increases exponentially with temperature. Figure 4b shows the temperature dependent H2 equilibrium pressure for the CMH systems, with NaH, MgH2, Mg2NiH4, Mg2FeH6 and TiH1.0–1.6H included as comparisons. The H2 equilibrium pressures have been plotted up to a value of 150 bar except for where thermodynamic data is not available, i.e., above the melting point of NaBH4.
Figure 4a shows that the theoretical gravimetric and volumetric heat storage capacities of LiBH4 (4936 kJ/kg and 3296 kJ/L at 727 °C [240]), NaBH4 (5176 kJ/kg and 5559 kJ/L at 507.5 °C [241]) and KBH4 (4248 kJ/kg and 4986 kJ/L at 615.5 °C [241]) are some of the highest known for a gas-solid thermochemical reaction. Only LiH (8389 kJ/kg and 6543 kJ/L at 938 °C [241]) and CaH2 (3857 kJ/kg and 7374 kJ/L at 1018 °C [241]) have theoretical gravimetric and/or volumetric heat storage capacities that are higher. However, the theoretical operating temperatures for LiBH4, NaBH4 and KBH4 are well within the target window of 600–800 °C for next generation CSP [212], whereas LiH and CaH2 require operating temperatures of >900 °C and 1000 °C, respectively. One issue with utilizing the alkali metal borohydrides for thermal heat storage is their reversibility. Both NaBH4 and KBH4 release hydrogen and decompose directly to the respective molten alkali metals that easily vaporize and segregate from the other decomposition product, boron. Pure LiBH4 also has problems with full reversibility but progress has recently been made by starting from the 0.68LiBH4-0.32Ca(BH4)2 eutectic composition [242].

Complex Metal Hydrides and Fuel Cell Applications

An emerging research direction is that of CMHs as H2 storage materials for stationary fuel cells (FCs) for clean power generation in seasonal storage, remote area and off-grid applications [261,262,263,264]. Unlike the closed-loop applications discussed in Section 4, this application utilizes an electrolyser to produce H2, which is stored (e.g., in a metal hydride) until the consumption of hydrogen by the fuel cell is required for electricity production. Interest in the use of classic intermetallic hydrides for this role began in the 1990s [261,262] and an in-depth review of the principles and progress on the use of metal hydrides in fuel cell applications has recently been published [264]. Some of the advantages of energy storage based on a photo voltaic (PV)-electrolyser-H2 storage-fuel cell configuration over a PV-Li ion battery system include:
Self-discharge: All batteries steadily lose charge over time and the self-discharge rate of Li-ion is highly dependent on temperature and the state-of-charge (SOC). The higher the temperature/SOC, the higher self-discharge rate [265]. The self-discharge rate of Li-ion batteries is usually 3–5% a month [265]. However, ageing of the battery may influence the self-discharge rate significantly [266]. In comparison, metal hydrides do not undergo “self-discharge” and can store their “charge” indefinitely.
Safety: Li-ion batteries suffer from lithium dendrite formation which may lead to short-circuit of the battery. Additionally, the organic electrolyte is hazardous in the presence of an oxidizing agent, which may result in runaway reactions and the battery catching fire or exploding [267]. The classic intermetallic hydrides first studied at near-ambient temperatures in combination with FCs have intrinsic safety due to their modest H2 pressures and operating temperatures combined with the endothermic nature of the H2 release process [264]. A more complicated scenario is presented by CMHs due to the fact that they are, in general, strong reducing agents, and many have the potential to be flammable or pyrophoric upon contact with air or water. However, in practical applications CMHs are typically used in the form of dense compacts and research on these has shown that they only undergo a mild temperature increase on exposure to air, and that direct immersion in water results in a mild temperature increase combined with slow release of H2 [268,269]. A comprehensive determination of the safety of CMHs would require performing reactivity testing at their practical operating temperatures.
Deep discharge: Unlike rechargeable batteries [270], metal hydride systems do not, in general, suffer capacity loss from being fully discharged [263].
Longevity: A battery’s cycle life is defined as the number of cycles until the capacity reaches 80% of its initial reversible value [271]. Commonly, Li-ion batteries have a cycle life between 1000 and 4500 cycles, i.e., a lifetime between 7 and 20 years [272]. Li-ion batteries with a 95% retention after 30,000 cycles have been discovered, however, at the expense of energy density [273], which is a severe drawback. In comparison, LaNi5 showed no capacity loss over 3300 cycles when using a H2 purity above 99.9999% [274]. Furthermore, TiFe0.8Ni0.2 showed only a 16% capacity loss after 65,000 cycles: equivalent to a ~178 year product lifetime based on a daily cycling regime [275]. The biggest issue so far for the CMHs is the reversibility and cyclic stability and much less long term cycling has been performed.
System size: As the energy density of CMHs is much higher than for Li-ion batteries, the system size of a stationary energy storage system will be much denser [263]. Additionally, the mass of an extensive Li-ion battery capable of storing large amounts of energy is considerable [276].
An additional benefit of thermally integrating CMHs with moderate- and high-temperature FCs is that, by consuming a significant fraction of the waste heat generated, the CMHs could actually improve the thermal control of high-temperature FC stacks [264].
One of the original driving forces for reducing the operating temperature of high capacity CMHs was that they could be used in mobile applications/passenger vehicles, where the waste heat (~80 °C) of Low Temperature Proton Exchange Membrane Fuel Cells (LT-PEMFCs) could be used to release H2 for FC consumption. However, H2 storage in complex hydrides with high thermodynamic stability is less of an issue for stationary applications as a number of different types of fuel cells are available that operate in different temperature ranges (AFC = Alkaline Fuel Cell, HT-PEMFC = High-Temperature Proton Exchange Membrane Fuel Cell, PAFC = Phosphoric Acid Fuel Cell, MCFC = Molten Carbonate Fuel Cell, SOFC = Solid Oxide Fuel Cell). Consequently, using complex hydride systems such as NaAlH4 [277,278,279,280,281,282] and Mg(NH2)2 + 2LiH [283] in conjunction with HT-PEMFCs is now being investigated. The integration of SOFCs with MgH2 [284,285,286,287,288], a metal hydride that historically suffered from high thermodynamic stability and poor kinetics, highlights the potential of CMHs for this application.
The different types of FCs are characterized by different operating temperatures as well as different efficiencies. The efficiency of the FC determines the amount of waste heat generated, Qloss, which can be used to release H2 from the metal hydride provided it exceeds the enthalpy of desorption, ΔHdes. The potential for thermal integration of a CMH with a particular type of FC can be assessed by considering their respective operating temperatures and the amount of waste heat available in comparison to that required to release H2 from the CMH. Figure 5 presents this comparison for the CMH systems and NaH, MgH2, Mg2NiH4, Mg2FeH6 and TiH1.0–1.6. It also includes the general range that applies to intermetallic hydrides, as a comparison, since most fuel cell integration research has been performed with these types of hydrides. Not all metal hydrides are suitable as H2 storage materials when paired with FCs. As an example, TiH1.0–1.6 would not be appropriate for use with any of the fuel cells, regardless of type, as the thermal energy required for H2 release, exceeds the waste heat available from the fuel cell.
The choice of CMH for H2 storage when paired with a particular FC will be dependent on the application under consideration. Some of the potential deciding factors include: CMH cost; cost and complexity of thermal integration between the CMH and FC; CMH footprint/volume; CMH operating temperature range; H2 equilibrium pressure of the CMH; H2 sorption kinetics of the CMH as a function of temperature and; the amount of FC waste heat available. The supply of H2 from CMHs with operating temperatures below ~250 °C requires only modest energy input where simple heat transfer solutions with low-efficiency, but low-cost, can be implemented. In principle, metal hydrides that operate at either low- or high-temperature are suitable for use with high-temperature fuel cells (MCFCs and SOFCs).
Figure 5 shows that, in general, the classic intermetallic hydrides are suitable for use with low- and moderate-temperature FCs (i.e., ~250 °C and below), including LT- and HT-PEMFCs, AFCs and PAFCs. In contrast, of all the CMH systems, only NaAlH4 and the Mg(NH2)2 + 2LiH system can theoretically operate at low enough temperatures for use with a LT-PEMFC. However, in practice, these systems only display sufficient H2 sorption kinetics above ~100 °C, which excludes their practical use with LT-PEMFCs. The CMH systems based on Na3AlH6, Na2LiAlH6 and KAlH4 + LiCl can also, in theory, operate with the low- and moderate-temperature FCs other than LT-PEMFCs. Of these systems, uncatalyzed KAlH4-LiCl also has kinetic limitations with the maximum rate of H2 release only achieved at ~230 °C [247]. Figure 5 also shows that all of the CMHs can potentially be used for H2 storage when paired with MCFCs and SOFCs. The choice of CMH would then be determined by the best match of the FC operating conditions, CMH properties and the application requirements.
To summarize, the high thermal stability of CMHs makes them attractive as heat storage materials as the energy density may be tenfold higher than current state-of-the-art materials. Secondly, CMHs in combination with a fuel cell for stationary energy storage, have several advantages compared to commercially available Li-ion batteries, e.g., no self-discharge, higher safety and energy density.

5. Conclusions

The chemistry of complex metal hydrides is extremely diverse, and has provided a wide range of novel materials in the past decade. Especially metal borohydrides with gravimetric hydrogen density of >10 mass %, have received increasing interest as solid state hydrogen storage materials. Furthermore, these materials have extremely rich chemistry based on structural flexibility and a wide range of elemental compositions.
The potential of complex metal hydrides for use in a range of new applications have also been highlighted in this review, see Figure 6. Examples are: as electrodes or electrolytes in solid state batteries, e.g., the high-temperature polymorph of Na2B12H12 which has a high ionic conductivity; thermal energy storage, e.g., at 507.5 °C NaBH4 has theoretical gravimetric and volumetric heat storage capacities of 5176 kJ/kg and 5559 kJ/L, respectively; and fuel cell applications. Indeed, many other properties, e.g., optical and magnetic properties are also promising and have been discovered in the metal borohydrides CsEu(BH4)3 and K2Gd(BH4)5 as fluorescent and magnetocaloric properties, respectively [123,289]. Furthermore, the metal closo-boranes, Ag2B10H10 and Ag2B12H12, display photosensitivity analogous to silver halides and forms silver nanofilaments upon electron beam exposure [201].
Hopefully, new metal borohydrides will be possible to design with desired chemical composition, atomic coordination, and exciting properties, based on the extensive knowledge about novel compounds, e.g., new combinations of light elements and boron, nitrogen or aluminum in complexes with hydrogen may lead to new interesting possibilities in applications.

Acknowledgments

The work was supported by the Aarhus University Research Foundations (AUFF), the Danish National Research Foundation, Center for Materials Crystallography (DNRF93), The Innovation Fund Denmark (project HyFill-Fast), Danish Council for Independent Research, DFF—4181-00462 (HyNanoBorN), Energistyrelsen, EUDP (64013-0136), the Danish Research Council for Nature and Universe (Danscatt), and by JSPS KAKENHI Grant Number 15KK0212 and the International Institute for Carbon Neutral Energy Research (WPI-I2CNER), sponsored by the Japanese Ministry of Education, Culture, Sports, Science and Technology. CEB and DS acknowledge the financial support of the Australian Research Council (ARC) for ARC Linkage grant LP120101848 and LP150100730. Finally, the authors are grateful to the Carlsberg Foundation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ley, M.B.; Jepsen, L.H.; Lee, Y.-S.; Cho, Y.W.; Bellosta von Colbe, J.M.; Dornheim, M.; Rokni, M.; Jensen, J.O.; Sloth, M.; Filinchuk, Y.; et al. Complex hydrides for hydrogen storage—New perspectives. Mater. Today 2014, 17, 122–128. [Google Scholar] [CrossRef] [Green Version]
  2. 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. Tailoring properties of borohydrides for hydrogen storage: A review. Phys. Status Solidi A 2011, 208, 1754–1773. [Google Scholar] [CrossRef]
  3. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Černý, R.; Ravnsbæk, D.B.; Filinchuk, Y.; Dornheim, M.; Besenbacher, F.; Jensen, T.R. Metal borohydrides and derivatives—Synthesis, structure and properties. Chem. Soc. Rev. 2017, 46, 1565–1634. [Google Scholar] [CrossRef] [PubMed]
  4. Jepsen, L.H.; Ley, M.B.; Lee, Y.-S.; Cho, Y.W.; Dornheim, M.; Jensen, J.O.; Filinchuk, Y.; Jørgensen, J.E.; Besenbacher, F.; Jensen, T.R. Boron–nitrogen based hydrides and reactive composites for hydrogen storage. Mater. Today 2014, 17, 129–135. [Google Scholar] [CrossRef] [Green Version]
  5. Bogdanović, B.; Schwickardi, M. Ti-doped alkali metal aluminium hydrides as potential novel reversible hydrogen storage materials. J. Alloys Compd. 1997, 253–254, 1–9. [Google Scholar] [CrossRef]
  6. Züttel, A.; Wenger, P.; Rentsch, S.; Sudan, P.; Mauron, P.; Emmenegger, C. LiBH4 a new hydrogen storage material. J. Power Sources 2003, 118, 1–7. [Google Scholar] [CrossRef]
  7. Chen, P.; Xiong, Z.; Luo, J.; Lin, J.; Tan, K.L. Interaction of hydrogen with metal nitrides and imides. Nature 2002, 420, 302–304. [Google Scholar] [CrossRef] [PubMed]
  8. Soulié, J.-P.; Renaudin, G.; Černý, R.; Yvon, K. Lithium boro-hydride LiBH4: I. Crystal structure. J. Alloys Compd. 2002, 346, 200–205. [Google Scholar] [CrossRef]
  9. Wang, J.; Li, H.-W.; Chen, P. Amides and borohydrides for high-capacity solid-state hydrogen storage—Materials design and kinetic improvements. MRS Bull. 2013, 38, 480–487. [Google Scholar] [CrossRef]
  10. Nakamori, Y.; Kitahara, G.; Orimo, S. Synthesis and dehydriding studies of Mg–N–H systems. J. Power Sources 2004, 138, 309–312. [Google Scholar] [CrossRef]
  11. Leng, H.Y.; Ichikawa, T.; Hino, S.; Hanada, N.; Isobe, S.; Fujii, H. New Metal–N–H System Composed of Mg(NH2)2 and LiH for Hydrogen Storage. J. Phys. Chem. B 2004, 108, 8763–8765. [Google Scholar] [CrossRef]
  12. Luo, W. (LiNH2–MgH2): A viable hydrogen storage system. J. Alloys Compd. 2004, 381, 284–287. [Google Scholar] [CrossRef]
  13. Paik, B.; Li, H.-W.; Wang, J.; Akiba, E. A Li–Mg–N–H composite as H2 storage material: A case study with Mg(NH2)2–4LiH–LiNH2. Chem. Commun. 2015, 51, 10018–10021. [Google Scholar] [CrossRef] [PubMed]
  14. Xiong, Z.; Hu, J.; Wu, G.; Chen, P.; Luo, W.; Gross, K.; Wang, J. Thermodynamic and kinetic investigations of the hydrogen storage in the Li–Mg–N–H system. J. Alloys Compd. 2005, 398, 235–239. [Google Scholar] [CrossRef]
  15. Torre, F.; Valentoni, A.; Milanese, C.; Pistidda, C.; Marini, A.; Dornheim, M.; Enzo, S.; Mulas, G.; Garroni, S. Kinetic improvement on the CaH2-catalyzed Mg(NH2)2 + 2LiH system. J. Alloys Compd. 2015, 645, S284–S287. [Google Scholar] [CrossRef]
  16. Shaw, L.L.; Ren, R.; Markmaitree, T.; Osborn, W. Effects of mechanical activation on dehydrogenation of the lithium amide and lithium hydride system. J. Alloys Compd. 2008, 448, 263–271. [Google Scholar] [CrossRef]
  17. Liu, Y.; Zhong, K.; Luo, K.; Gao, M.; Pan, H.; Wang, Q. Size-Dependent Kinetic Enhancement in Hydrogen Absorption and Desorption of the Li–Mg–N–H System. J. Am. Chem. Soc. 2009, 131, 1862–1870. [Google Scholar] [CrossRef] [PubMed]
  18. Cao, H.; Santoru, A.; Pistidda, C.; M. Richter, T.M.; Chaudhary, A.-L.; Gizer, G.; Niewa, R.; Chen, P.; Klassen, T.; Dornheim, M. New synthesis route for ternary transition metal amides as well as ultrafast amide–hydride hydrogen storage materials. Chem. Commun. 2016, 52, 5100–5103. [Google Scholar] [CrossRef] [PubMed]
  19. Cao, H.; Richter, T.M.M.; Pistidda, C.; Chaudhary, A.-L.; Santoru, A.; Gizer, G.; Niewa, R.; Chen, P.; Klassen, T.; Dornheim, M. Ternary Amides Containing Transition Metals for Hydrogen Storage: A Case Study with Alkali Metal Amidozincates. ChemSusChem 2015, 8, 3777–3782. [Google Scholar] [CrossRef] [PubMed]
  20. Cao, H.; Pistidda, C.; Richter, T.M.M.; Santoru, A.; Milanese, C.; Garroni, S.; Bednarcik, J.; Chaudhary, A.-L.; Gizer, G.; Liermann, H.-P.; et al. In situ X-ray diffraction studies on the de/rehydrogenation processes of the K2[Zn(NH2)4]-8LiH system. J. Phys. Chem. C 2017, 121, 1546–1551. [Google Scholar] [CrossRef]
  21. Janot, R.; Eymery, J.-B.; Tarascon, J.-M. Decomposition of LiAl(NH2)4 and Reaction with LiH for a Possible Reversible Hydrogen Storage. J. Phys. Chem. C 2007, 111, 2335–2340. [Google Scholar] [CrossRef]
  22. Jepsen, L.H.; Wang, P.; Wu, G.; Xiong, Z.; Besenbacher, F.; Chen, P.; Jensen, T.R. Thermal decomposition of sodium amide, NaNH2, and sodium amide hydroxide composites, NaNH2-NaOH. Phys. Chem. Chem. Phys. 2016, 18, 25257–25264. [Google Scholar] [CrossRef] [PubMed]
  23. Lide, D.R. CRC Handbook of Chemistry and Physics, 88th ed.; CRC Press: Boca Raton, FL, USA, 25 June 2007; ISBN 0-8493-0488-1. [Google Scholar]
  24. 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]
  25. Mauron, P.; Buchter, F.; Friedrichs, O.; Remhof, A.; Bielmann, M.; Zwicky, C.N.; Zuttel, A. Stability and Reversibility of LiBH4. J. Phys. Chem. B 2008, 112, 906–910. [Google Scholar] [CrossRef] [PubMed]
  26. Martelli, P.; Caputo, R.; Remhof, A.; Mauron, P.; Borgschulte, A.; Züttel, A. Stability and Decomposition of NaBH4. J. Phys. Chem. C 2010, 114, 7173–7177. [Google Scholar] [CrossRef]
  27. Dornheim, M. Thermodynamics of Metal Hydrides: Tailoring Reaction Enthalpies of Hydrogen Storage Materials; InTech: Rijeka, Croatia, 2011; ISBN 978-953-307-563-1. [Google Scholar]
  28. Na Ranong, C.; Höhne, 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]
  29. Isobe, S.; Ichikawa, T.; Tokoyoda, K.; Hanada, N.; Leng, H.; Fujii, H.; Kojima, Y. Evaluation of enthalpy change due to hydrogen desorption for lithium amide/imide system by differential scanning calorimetry. Thermochim. Acta 2008, 468, 35–38. [Google Scholar] [CrossRef]
  30. Kojima, Y.; Kawai, Y. IR characterizations of lithium imide and amide. J. Alloys Compd. 2005, 395, 236–239. [Google Scholar] [CrossRef]
  31. Ichikawa, T.; Isobe, S.; Hanada, N.; Fujii, H. Lithium nitride for reversible hydrogen storage. J. Alloys Compd. 2004, 365, 271–276. [Google Scholar] [CrossRef]
  32. Paskevicius, M.; Webb, J.; Pitt, M.P.; Blach, T.P.; Hauback, B.C.; Gray, E.; Buckley, C.E. Mechanochemical synthesis of aluminium nanoparticles and their deuterium sorption properties to 2kbar. J. Alloys Compd. 2009, 481, 595–599. [Google Scholar] [CrossRef]
  33. 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 modelling of the alane isomorphs. J. Mater. Chem. 2008, 18, 2361. [Google Scholar] [CrossRef]
  34. 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]
  35. Ikeda, K.; Ohshita, H.; Kaneko, N.; Zhang, J.; Yonemura, M.; Otomo, T.; Suzuya, K.; Yukawa, H.; Morinaga, M.; Li, H.-W.; et al. Structural and Hydrogen Desorption Properties of Aluminum Hydride. Mater. Trans. 2011, 52, 598–601. [Google Scholar] [CrossRef]
  36. Finholt, A.E.; Bond, A.C.; Schlesinger, H.I. Lithium Aluminum Hydride, Aluminum Hydride and Lithium Gallium Hydride, and Some of their Applications in Organic and Inorganic Chemistry1. J. Am. Chem. Soc. 1947, 69, 1199–1203. [Google Scholar] [CrossRef]
  37. Hauback, B.C. Structures of aluminium-based light weight hydrides. Z. Krist. 2009, 223, 636–648. [Google Scholar] [CrossRef]
  38. 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, 2450–2453. [Google Scholar] [CrossRef]
  39. Turley, J.W.; Rinn, H.W. Crystal structure of aluminum hydride. Inorg. Chem. 1969, 8, 18–22. [Google Scholar] [CrossRef]
  40. 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] [PubMed]
  41. Brinks, H.W.; Langley, W.; 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]
  42. Brinks, H.W.; Brown, C.; 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]
  43. Yartys, V.A.; Denys, R.V.; Maehlen, J.P.; Frommen, C.; Fichtner, M.; Bulychev, B.M.; Emerich, H. Double-Bridge Bonding of Aluminium and Hydrogen in the Crystal Structure of γ-AlH3. Inorg. Chem. 2007, 46, 1051–1055. [Google Scholar] [CrossRef] [PubMed]
  44. Konovalov, S.K.; Bulychev, B.M. The P,T-State Diagram and Solid Phase Synthesis of Aluminum Hydride. Inorg. Chem. 1995, 34, 172–175. [Google Scholar] [CrossRef]
  45. Sartori, S.; Istad-Lem, A.; Brinks, H.W.; Hauback, B.C. Mechanochemical synthesis of alane. Int. J. Hydrogen Energy 2009, 34, 6350–6356. [Google Scholar] [CrossRef]
  46. Paskevicius, M.; Sheppard, D.A.; Buckley, C.E. Characterisation of mechanochemically synthesised alane (AlH3) nanoparticles. J. Alloys Compd. 2009, 487, 370–376. [Google Scholar] [CrossRef]
  47. 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]
  48. Dinh, L.V.; Knight, D.A.; Paskevicius, M.; Buckley, C.E.; Zidan, R. Novel methods for synthesizing halide-free alane without the formation of adducts. Appl. Phys. A 2012, 107, 173–181. [Google Scholar] [CrossRef]
  49. 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]
  50. Graetz, J.; Reilly, J.J. Decomposition Kinetics of the AlH3 Polymorphs. J. Phys. Chem. B 2005, 109, 22181–22185. [Google Scholar] [CrossRef] [PubMed]
  51. 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, 182–186. [Google Scholar] [CrossRef]
  52. Graetz, J. New approaches to hydrogen storage. Chem. Soc. Rev. 2009, 38, 73. [Google Scholar] [CrossRef] [PubMed]
  53. Ashby, E.C.; Brendel, G.J.; Redman, H.E. Direct Synthesis of Complex Metal Hydrides. Inorg. Chem. 1963, 2, 499–504. [Google Scholar] [CrossRef]
  54. Clasen, H. Alanat-Synthese aus den Elementen und ihre Bedeutung. Angew. Chem. 1961, 73, 322–331. [Google Scholar] [CrossRef]
  55. Balema, V.P.; Balema, L. Missing pieces of the puzzle or about some unresolved issues in solid state chemistry of alkali metal aluminohydrides. Phys. Chem. Chem. Phys. 2005, 7, 1310–1314. [Google Scholar] [CrossRef] [PubMed]
  56. Dilts, J.A.; Ashby, E.C. Thermal decomposition of complex metal hydrides. Inorg. Chem. 1972, 11, 1230–1236. [Google Scholar] [CrossRef]
  57. 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]
  58. Ares, J.R.; Aguey-Zinsou, K.-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]
  59. Block, J.; Gray, A.P. The Thermal Decomposition of Lithium Aluminum Hydride. Inorg. Chem. 1965, 4, 304–305. [Google Scholar] [CrossRef]
  60. 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]
  61. Dymova, T.N.; Aleksandrov, D.P.; Konoplev, V.N.; Silina, T.A.; Sizareva, A.S. Spontaneous and thermal-decomposition of Lithium Tetrahydroaluminate LiAlH4-the promoting effect of mechanochemical action on the process. Koord. Khimiya 1994, 20, 279–285. [Google Scholar]
  62. Andreasen, A.; Vegge, T.; Pedersen, A.S. Dehydrogenation kinetics of as-received and ball-milled. J. Solid State Chem. 2005, 178, 3672–3678. [Google Scholar] [CrossRef]
  63. Balema, V.P.; Dennis, K.W.; Pecharsky, V.K. Rapid solid-state transformation of tetrahedral [AlH4] into octahedral [AlH6]3 in lithium aluminohydride. Chem. Commun. 2000, 1665–1666. [Google Scholar] [CrossRef]
  64. Balema, V.P.; Wiench, J.W.; Dennis, K.W.; Pruski, M.; Pecharsky, V.K. Titanium catalyzed solid-state transformations in LiAlH4 during high-energy ball-milling. J. Alloys Compd. 2001, 329, 108–114. [Google Scholar] [CrossRef]
  65. Balema, V.P.; Pecharsky, V.K.; Dennis, K.W. Solid state phase transformations in LiAlH4 during high-energy ball-milling. J. Alloys Compd. 2000, 313, 69–74. [Google Scholar] [CrossRef]
  66. Blanchard, D.; Brinks, H.W.; Hauback, B.C.; Norby, P. Desorption of LiAlH4 with Ti- and V-based additives. Mater. Sci. Eng. B 2004, 108, 54–59. [Google Scholar] [CrossRef]
  67. Lai, Q.; Paskevicius, M.; Sheppard, D.A.; Buckley, C.E.; Thornton, A.W.; Hill, M.R.; Gu, Q.; Mao, J.; Huang, Z.; Liu, H.K.; et al. Hydrogen Storage Materials for Mobile and Stationary Applications: Current State of the Art. ChemSusChem 2015, 8, 2789–2825. [Google Scholar] [CrossRef] [PubMed]
  68. 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]
  69. Graetz, J.; Wegrzyn, J.; Reilly, J.J. Regeneration of Lithium Aluminum Hydride. J. Am. Chem. Soc. 2008, 130, 17790–17794. [Google Scholar] [CrossRef] [PubMed]
  70. 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]
  71. Nielsen, T.K.; Javadian, P.; Polanski, M.; Besenbacher, F.; Bystrzycki, J.; Skibsted, J.; Jensen, T.R. Nanoconfined NaAlH4: Prolific effects from increased surface area and pore volume. Nanoscale 2014, 6, 599–607. [Google Scholar] [CrossRef] [PubMed]
  72. Nielsen, T.K.; Javadian, P.; Polanski, M.; Besenbacher, F.; Bystrzycki, J.; Jensen, T.R. Nanoconfined NaAlH4: Determination of distinct prolific effects from pore size, crystallite size, and surface interactions. J. Phys. Chem. C 2012, 116, 21046–21051. [Google Scholar] [CrossRef]
  73. Nielsen, T.K.; Polanski, M.; Zasada, D.; Javadian, P.; Besenbacher, F.; Bystrzycki, J.; Skibsted, J.; Jensen, T.R. Improved Hydrogen Storage Kinetics of Nanoconfined NaAlH4 Catalyzed with TiCl3 Nanoparticles. ACS Nano 2011, 5, 4056–4064. [Google Scholar] [CrossRef] [PubMed]
  74. Jensen, C.M.; Gross, K.J. Development of catalytically enhanced sodium aluminum hydride as a hydrogen-storage material. Appl. Phys. A 2001, 72, 213–219. [Google Scholar] [CrossRef]
  75. Zidan, R.A.; Takara, S.; Hee, A.G.; Jensen, C.M. Hydrogen cycling behavior of zirconium and titanium–zirconium-doped sodium aluminum hydride. J. Alloys Compd. 1999, 285, 119–122. [Google Scholar] [CrossRef]
  76. Bogdanović, B.; Felderhoff, M.; Pommerin, A.; Schüth, F.; Spielkamp, N. Advanced Hydrogen-Storage Materials Based on Sc-, Ce-, and Pr-Doped NaAlH4. Adv. Mater. 2006, 18, 1198–1201. [Google Scholar] [CrossRef]
  77. Pitt, M.P.; Paskevicius, M.; Webb, C.J.; Sørby, M.H.; Delleda, S.; Jensen, T.R.; Hauback, B.C.; Buckley, C.E.; Gray, E.M. Nanoscopic Al1−xCex phases in the NaH + Al + 0.02CeCl3 system. Int. J. Hydrogen Energy 2011, 36, 8403–8411. [Google Scholar] [CrossRef]
  78. Fichtner, M.; Fuhr, O. Synthesis and structures of magnesium alanate and two solvent adducts. J. Alloys Compd. 2002, 345, 286–296. [Google Scholar] [CrossRef]
  79. 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] [PubMed]
  80. 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 of Their Decomposition Intermediates, EuAlH5 and SrAlH5. Inorg. Chem. 2012, 51, 4143–4150. [Google Scholar] [CrossRef] [PubMed]
  81. 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]
  82. 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]
  83. Ares, J.R.; Aguey-Zinsou, K.-F.; Leardini, F.; Ferrer, I.J.; 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]
  84. Morioka, H.; Kakizaki, K.; Chung, S.-C.; Yamada, A. Reversible hydrogen decomposition of KAlH4. J. Alloys Compd. 2003, 353, 310–314. [Google Scholar] [CrossRef]
  85. Černý, R.; Schouwink, P. The crystal chemistry of inorganic metal borohydrides and their relation to metal oxides. Acta Crystallogr. B 2015, 71, 619–640. [Google Scholar] [CrossRef] [PubMed]
  86. Züttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage properties of LiBH4. J. Alloys Compd. 2003, 356–357, 515–520. [Google Scholar] [CrossRef]
  87. Nakamori, Y.; Orimo, S. Destabilization of Li-based complex hydrides. J. Alloys Compd. 2004, 370, 271–275. [Google Scholar] [CrossRef]
  88. Hartman, M.R.; Rush, J.J.; Udovic, T.J.; Bowman, R.C., Jr.; Hwang, S.-J. Structure and vibrational dynamics of isotopically labeled lithium borohydride using neutron diffraction and spectroscopy. J. Solid State Chem. 2007, 180, 1298–1305. [Google Scholar] [CrossRef]
  89. Filinchuk, Y.; Chernyshov, D.; Cerny, R. Lightest Borohydride Probed by Synchrotron X-ray Diffraction: Experiment Calls for a New Theoretical Revision. J. Phys. Chem. C 2008, 112, 10579–10584. [Google Scholar] [CrossRef]
  90. Eberle, U.; Felderhoff, M.; Schüth, F. Chemical and Physical Solutions for Hydrogen Storage. Angew. Chem. Int. Ed. 2009, 48, 6608–6630. [Google Scholar] [CrossRef] [PubMed]
  91. Grube, E.; Olesen, C.H.; Ravnsbæk, D.B.; Jensen, T.R. Barium borohydride chlorides: Synthesis, crystal structures and thermal properties. Dalton Trans. 2016, 45, 8291–8299. [Google Scholar] [CrossRef] [PubMed]
  92. Ravnsbæk, D.B.; Nickels, E.A.; Černý, R.; Olesen, C.H.; David, W.I.F.; Edwards, P.P.; Filinchuk, Y.; Jensen, T.R. Novel Alkali Earth Borohydride Sr(BH4)2 and Borohydride-Chloride Sr(BH4)Cl. Inorg. Chem. 2013, 52, 10877–10885. [Google Scholar] [CrossRef] [PubMed]
  93. Sharma, M.; Didelot, E.; Spyratou, A.; Lawson Daku, L.M.; Černý, R.; Hagemann, H. Halide Free M(BH4)2 (M = Sr, Ba, and Eu) Synthesis, Structure, and Decomposition. Inorg. Chem. 2016, 55, 7090–7097. [Google Scholar] [CrossRef] [PubMed]
  94. Filinchuk, Y.; Richter, B.; Jensen, T.R.; Dmitriev, V.; Chernyshov, D.; Hagemann, H. Porous and Dense Magnesium Borohydride Frameworks: Synthesis, Stability, and Reversible Absorption of Guest Species. Angew. Chem. Int. Ed. 2011, 50, 11162–11166. [Google Scholar] [CrossRef] [PubMed]
  95. Filinchuk, Y.; Černý, R.; Hagemann, H. Insight into Mg(BH4)2 with Synchrotron X-ray Diffraction: Structure Revision, Crystal Chemistry, and Anomalous Thermal Expansion. Chem. Mater. 2009, 21, 925–933. [Google Scholar] [CrossRef]
  96. Her, J.-H.; Stephens, P.W.; Gao, Y.; Soloveichik, G.L.; Rijssenbeek, J.; Andrus, M.; Zhao, J.-C. Structure of unsolvated magnesium borohydride Mg(BH4)2. Acta Crystallogr. B 2007, 63, 561–568. [Google Scholar] [CrossRef] [PubMed]
  97. Pitt, M.P.; Webb, C.J.; Paskevicius, M.; Sheptyakov, D.; Buckley, C.E.; Gray, E.M. In Situ Neutron Diffraction Study of the Deuteration of Isotopic Mg11B2. J. Phys. Chem. C 2011, 115, 22669–22679. [Google Scholar] [CrossRef]
  98. Paskevicius, M.; Pitt, M.P.; Webb, C.J.; Sheppard, D.A.; Filsø, U.; Gray, E.M.; Buckley, C.E. In-Situ X-ray Diffraction Study of γ-Mg(BH4)2 Decomposition. J. Phys. Chem. C 2012, 116, 15231–15240. [Google Scholar] [CrossRef]
  99. David, W.I.F.; Callear, S.K.; Jones, M.O.; Aeberhard, P.C.; Culligan, S.D.; Pohl, A.H.; Johnson, S.R.; Ryan, K.R.; Parker, J.E.; Edwards, P.P.; et al. The structure, thermal properties and phase transformations of the cubic polymorph of magnesium tetrahydroborate. Phys. Chem. Chem. Phys. 2012, 14, 11800–11807. [Google Scholar] [CrossRef] [PubMed]
  100. Amieiro-Fonseca, A.; Ellis, S.R.; Nuttall, C.J.; Hayden, B.E.; Guerin, S.; Purdy, G.; Soulié, J.-P.; Callear, S.K.; Culligan, S.D.; David, W.I.F.; et al. A multidisciplinary combinatorial approach for tuning promising hydrogen storage materials towards automotive applications. Faraday Discuss. 2011, 151, 369. [Google Scholar] [CrossRef] [PubMed]
  101. Li, H.-W.; Kikuchi, K.; Nakamori, Y.; Miwa, K.; Towata, S.; Orimo, S. Effects of ball milling and additives on dehydriding behaviors of well-crystallized Mg(BH4)2. Scr. Mater. 2007, 57, 679–682. [Google Scholar] [CrossRef]
  102. Li, H.-W.; Miwa, K.; Ohba, N.; Fujita, T.; Sato, T.; Yan, Y.; Towata, S.; Chen, M.W.; Orimo, S. Formation of an intermediate compound with a B12H12 cluster: Experimental and theoretical studies on magnesium borohydride Mg(BH4)2. Nanotechnology 2009, 20, 204013. [Google Scholar] [CrossRef] [PubMed]
  103. Li, H.-W.; Matsunaga, T.; Yan, Y.; Maekawa, H.; Ishikiriyama, M.; Orimo, S. Nanostructure-induced hydrogenation of layered compound MgB2. J. Alloys Compd. 2010, 505, 654–656. [Google Scholar] [CrossRef]
  104. Roedern, E.; Jensen, T.R. Ammine-Stabilized Transition-Metal Borohydrides of Iron, Cobalt, and Chromium: Synthesis and Characterization. Inorg. Chem. 2015, 54, 10477–10482. [Google Scholar] [CrossRef] [PubMed]
  105. Callini, E.; Szilágyi, P.Á.; Paskevicius, M.; Stadie, N.P.; Réhault, J.; Buckley, C.E.; Borgschulte, A.; Züttel, A. Stabilization of volatile Ti(BH4)3 by nano-confinement in a metal–organic framework. Chem. Sci. 2015, 7, 666–672. [Google Scholar] [CrossRef] [PubMed]
  106. Rude, L.H.; Corno, M.; Ugliengo, P.; Baricco, M.; Lee, Y.-S.; Cho, Y.W.; Besenbacher, F.; Overgaard, J.; Jensen, T.R. Synthesis and Structural Investigation of Zr(BH4)4. J. Phys. Chem. C 2012, 116, 20239–20245. [Google Scholar] [CrossRef]
  107. Nickels, E.A.; Jones, M.O.; David, W.I.F.; Johnson, S.R.; Lowton, R.L.; Sommariva, M.; Edwards, P.P. Tuning the Decomposition Temperature in Complex Hydrides: Synthesis of a Mixed Alkali Metal Borohydride. Angew. Chem. Int. Ed. 2008, 47, 2817–2819. [Google Scholar] [CrossRef] [PubMed]
  108. Roedern, E.; Hansen, B.R.S.; Ley, M.B.; Jensen, T.R. Effect of Eutectic Melting, Reactive Hydride Composites, and Nanoconfinement on Decomposition and Reversibility of LiBH4–KBH4. J. Phys. Chem. C 2015, 119, 25818–25825. [Google Scholar] [CrossRef]
  109. Jensen, S.R.H.; Jepsen, L.H.; Skibsted, J.; Jensen, T.R. Phase Diagram for the NaBH4–KBH4 System and the Stability of a Na1−xKxBH4 Solid Solution. J. Phys. Chem. C 2015, 119, 27919–27929. [Google Scholar] [CrossRef]
  110. Nakamori, Y.; Miwa, K.; Ninomiya, A.; Li, H.; Ohba, N.; Towata, S.; Züttel, A.; Orimo, S. Correlation between thermodynamical stabilities of metal borohydrides and cation electronegativites: First-principles calculations and experiments. Phys. Rev. B 2006, 74, 45126. [Google Scholar] [CrossRef]
  111. Li, H.-W.; Orimo, S.; Nakamori, Y.; Miwa, K.; Ohba, N.; Towata, S.; Züttel, A. Materials designing of metal borohydrides: Viewpoints from thermodynamical stabilities. J. Alloys Compd. 2007, 446–447, 315–318. [Google Scholar] [CrossRef]
  112. Ravnsbæk, D.; Filinchuk, Y.; Cerenius, Y.; Jakobsen, H.J.; Besenbacher, F.; Skibsted, J.; Jensen, T.R. A Series of Mixed-Metal Borohydrides. Angew. Chem. Int. Ed. 2009, 48, 6659–6663. [Google Scholar] [CrossRef] [PubMed]
  113. Ravnsbæk, D.B.; Frommen, C.; Reed, D.; Filinchuk, Y.; Sørby, M.; Hauback, B.C.; Jakobsen, H.J.; Book, D.; Besenbacher, F.; Skibsted, J.; et al. Structural studies of lithium zinc borohydride by neutron powder diffraction, Raman and NMR spectroscopy. J. Alloys Compd. 2011, 509, S698–S704. [Google Scholar] [CrossRef]
  114. Černý, R.; Kim, K.C.; Penin, N.; D’Anna, V.; Hagemann, H.; Sholl, D.S. AZn2(BH4)5 (A = Li, Na) and NaZn(BH4)3: Structural Studies. J. Phys. Chem. C 2010, 114, 19127–19133. [Google Scholar] [CrossRef]
  115. Ley, M.B.; Ravnsbæk, D.B.; Filinchuk, Y.; Lee, Y.-S.; Janot, R.; Cho, Y.W.; Skibsted, J.; Jensen, T.R. LiCe(BH4)3Cl, a New Lithium-Ion Conductor and Hydrogen Storage Material with Isolated Tetranuclear Anionic Clusters. Chem. Mater. 2012, 24, 1654–1663. [Google Scholar] [CrossRef]
  116. Møller, K.T.; Ley, M.B.; Schouwink, P.; Černý, R.; Jensen, T.R. Synthesis and thermal stability of perovskite alkali metal strontium borohydrides. Dalton Trans. 2016, 45, 831–840. [Google Scholar] [CrossRef] [PubMed]
  117. Park, H.H.; Pezat, M.; Darriet, B. A new ternary hydride: CsCaH3. Rev. Chim. Minérale 1986, 23, 323–328. [Google Scholar]
  118. Bouamrane, A.; Laval, J.P.; Soulie, J.-P.; Bastide, J.P. Structural characterization of NaMgH2F and NaMgH3. Mater. Res. Bull. 2000, 35, 545–549. [Google Scholar] [CrossRef]
  119. Pottmaier, D.; Pinatel, E.R.; Vitillo, J.G.; Garroni, S.; Orlova, M.; Baró, M.D.; Vaughan, G.B.M.; Fichtner, M.; Lohstroh, W.; Baricco, M. Structure and Thermodynamic Properties of the NaMgH3 Perovskite: A Comprehensive Study. Chem. Mater. 2011, 23, 2317–2326. [Google Scholar] [CrossRef]
  120. Ikeda, K.; Kogure, Y.; Nakamori, Y.; Orimo, S. Reversible hydriding and dehydriding reactions of perovskite-type hydride NaMgH3. Scr. Mater. 2005, 53, 319–322. [Google Scholar] [CrossRef]
  121. Ikeda, K.; Kato, S.; Shinzato, Y.; Okuda, N.; Nakamori, Y.; Kitano, A.; Yukawa, H.; Morinaga, M.; Orimo, S. Thermodynamical stability and electronic structure of a perovskite-type hydride, NaMgH3. J. Alloys Compd. 2007, 446–447, 162–165. [Google Scholar] [CrossRef]
  122. Schouwink, P.; D’Anna, V.; Ley, M.B.; Daku, L.M.L.; Richter, B.; Jensen, T.R.; Hagemann, H.; Černý, R. Bimetallic Borohydrides in the System M(BH4)2–KBH4 (M = Mg, Mn): On the Structural Diversity. J. Phys. Chem. C 2012, 116, 10829–10840. [Google Scholar] [CrossRef]
  123. Schouwink, P.; Ley, M.B.; Tissot, A.; Hagemann, H.; Jensen, T.R.; Smrčok, Ľ.; Černý, R. Structure and properties of complex hydride perovskite materials. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  124. Møller, K.T.; Jørgensen, M.; Fogh, A.S.; Jensen, T.R. Perovskite alkali metal samarium borohydrides: Crystal structures and thermal decomposition. Dalton Trans. 2017, 46, 11905–11912. [Google Scholar] [CrossRef] [PubMed]
  125. Schouwink, P.; Hagemann, H.; Embs, J.P.; D’Anna, V.; Černý, R. Di-hydrogen contact induced lattice instabilities and structural dynamics in complex hydride perovskites. J. Phys. Condens. Matter 2015, 27, 265403. [Google Scholar] [CrossRef] [PubMed]
  126. Černý, R.; Schouwink, P.; Sadikin, Y.; Stare, K.; Smrčok, L.; Richter, B.; Jensen, T.R. Trimetallic Borohydride Li3MZn5(BH4)15 (M = Mg, Mn) Containing Two Weakly Interconnected Frameworks. Inorg. Chem. 2013, 52, 9941–9947. [Google Scholar] [CrossRef] [PubMed]
  127. Schouwink, P.; Ley, M.B.; Jensen, T.R.; Černý, R. Borohydrides: From sheet to framework topologies. Dalton Trans. 2014, 43, 7726–7733. [Google Scholar] [CrossRef] [PubMed]
  128. Brighi, M.; Schouwink, P.; Sadikin, Y.; Černý, R. Fast ion conduction in garnet-type metal borohydrides Li3K3Ce2(BH4)12 and Li3K3La2(BH4)12. J. Alloys Compd. 2016, 662, 388–395. [Google Scholar] [CrossRef]
  129. Payandeh GharibDoust, S.; Heere, M.; Sørby, M.H.; Ley, M.B.; Ravnsbæk, D.B.; Hauback, B.C.; Černý, R.; Jensen, T.R. Synthesis, structure and properties of new bimetallic sodium and potassium lanthanum borohydrides. Dalton Trans. 2016, 45, 19002–19011. [Google Scholar] [CrossRef] [PubMed]
  130. Jepsen, L.H.; Ley, M.B.; Filinchuk, Y.; Besenbacher, F.; Jensen, T.R. Tailoring the Properties of Ammine Metal Borohydrides for Solid-State Hydrogen Storage. ChemSusChem 2015, 8, 1452–1463. [Google Scholar] [CrossRef] [PubMed]
  131. Huang, J.; Tan, Y.; Su, J.; Gu, Q.; Černý, R.; Ouyang, L.; Sun, D.; Yu, X.; Zhu, M. Synthesis, structure and dehydrogenation of zirconium borohydride octaammoniate. Chem. Commun. 2015, 51, 2794–2797. [Google Scholar] [CrossRef] [PubMed]
  132. Jepsen, L.H.; Ley, M.B.; Černý, R.; Lee, Y.-S.; Cho, Y.W.; Ravnsbæk, D.; Besenbacher, F.; Skibsted, J.; Jensen, T.R. Trends in Syntheses, Structures, and Properties for Three Series of Ammine Rare-Earth Metal Borohydrides, M(BH4)3·nNH3 (M = Y, Gd, and Dy). Inorg. Chem. 2015, 54, 7402–7414. [Google Scholar] [CrossRef] [PubMed]
  133. Guo, Y.; Yu, X.; Sun, W.; Sun, D.; Yang, W. The Hydrogen-Enriched Al–B–N System as an Advanced Solid Hydrogen-Storage Candidate. Angew. Chem. Int. Ed. 2011, 50, 1087–1091. [Google Scholar] [CrossRef] [PubMed]
  134. Guo, Y.; Wu, H.; Zhou, W.; Yu, X. Dehydrogenation Tuning of Ammine Borohydrides Using Double-Metal Cations. J. Am. Chem. Soc. 2011, 133, 4690–4693. [Google Scholar] [CrossRef] [PubMed]
  135. Gu, Q.; Gao, L.; Guo, Y.; Tan, Y.; Tang, Z.; Wallwork, K.S.; Zhang, F.; Yu, X. Structure and decomposition of zinc borohydride ammonia adduct: Towards a pure hydrogen release. Energy Environ. Sci. 2012, 5, 7590–7600. [Google Scholar] [CrossRef]
  136. Jepsen, L.H.; Lee, Y.-S.; Černý, R.; Sarusie, R.S.; Cho, Y.W.; Besenbacher, F.; Jensen, T.R. Ammine Calcium and Strontium Borohydrides: Syntheses, Structures, and Properties. ChemSusChem 2015, 8, 3472–3482. [Google Scholar] [CrossRef] [PubMed]
  137. Guo, Y.; Jiang, Y.; Xia, G.; Yu, X. Ammine aluminium borohydrides: An appealing system releasing over 12 wt % pure H2 under moderate temperature. Chem. Commun. 2012, 48, 4408–4410. [Google Scholar] [CrossRef] [PubMed]
  138. Wu, H.; Zhou, W.; Pinkerton, F.E.; Meyer, M.S.; Srinivas, G.; Yildirim, T.; Udovic, T.J.; Rush, J.J. A new family of metal borohydride ammonia borane complexes: Synthesis, structures, and hydrogen storage properties. J. Mater. Chem. 2010, 20, 6550–6556. [Google Scholar] [CrossRef]
  139. Jepsen, L.H.; Ban, V.; Møller, K.T.; Lee, Y.-S.; Cho, Y.W.; Besenbacher, F.; Filinchuk, Y.; Skibsted, J.; Jensen, T.R. Synthesis, Crystal Structure, Thermal Decomposition, and 11B MAS NMR Characterization of Mg(BH4)2(NH3BH3)2. J. Phys. Chem. C 2014, 118, 12141–12153. [Google Scholar] [CrossRef]
  140. Dovgaliuk, I.; Le Duff, C.S.; Robeyns, K.; Devillers, M.; Filinchuk, Y. Mild Dehydrogenation of Ammonia Borane Complexed with Aluminum Borohydride. Chem. Mater. 2015, 27, 768–777. [Google Scholar] [CrossRef]
  141. Luo, J.; Wu, H.; Zhou, W.; Kang, X.; Fang, Z.; Wang, P. LiBH4·NH3BH3: A new lithium borohydride ammonia borane compound with a novel structure and favorable hydrogen storage properties. Int. J. Hydrogen Energy 2012, 37, 10750–10757. [Google Scholar] [CrossRef]
  142. Chen, X.; Yuan, F.; Gu, Q.; Yu, X. Synthesis, structures and hydrogen storage properties of two new H-enriched compounds: Mg(BH4)2(NH3BH3)2 and Mg(BH4)2·(NH3)2(NH3BH3). Dalton Trans. 2013, 42, 14365–14368. [Google Scholar] [CrossRef] [PubMed]
  143. Chua, Y.S.; Chen, P.; Wu, G.; Xiong, Z. Development of amidoboranes for hydrogen storage. Chem. Commun. 2011, 47, 5116–5129. [Google Scholar] [CrossRef] [PubMed]
  144. Dovgaliuk, I.; Jepsen, L.H.; Safin, D.A.; Łodziana, Z.; Dyadkin, V.; Jensen, T.R.; Devillers, M.; Filinchuk, Y. A Composite of Complex and Chemical Hydrides Yields the First Al-Based Amidoborane with Improved Hydrogen Storage Properties. Chem. Eur. J. 2015, 14562–14570. [Google Scholar] [CrossRef] [PubMed]
  145. Møller, K.T.; Jørgensen, M.; Andreasen, J.G.; Skibsted, J.; Łodziana, Z.; Filinchuk, Y.; Jensen, T.R. Synthesis and Thermal Decomposition of Potassium Tetraamidoboranealuminate, K[Al(NH2BH3)4]. Int. J. Hydrogen Energy 2017. accepted. [Google Scholar]
  146. Ikoma, M. Ni-Metal Hydride Batteries. In Encyclopedia of Applied Electrochemistry; Kreysa, G., Ota, K., Savinell, R.F., Eds.; Springer: New York, NY, USA, 2014; pp. 1363–1366. ISBN 978-1-4419-6995-8. [Google Scholar]
  147. Sakai, T.; Miyamura, H.; Kuriyama, N.; Kato, A.; Oguro, K.; Ishikawa, H. Metal Hydride Anodes for Nickel-Hydrogen Secondary Battery. J. Electrochem. Soc. 1990, 137, 795–799. [Google Scholar] [CrossRef]
  148. Møller, K.T.; Jensen, T.R.; Akiba, E.; Li, H. Hydrogen—A sustainable energy carrier. Prog. Nat. Sci. Mater. Int. 2017, 27, 34–40. [Google Scholar] [CrossRef]
  149. Oumellal, Y.; Rougier, A.; Nazri, G.A.; Tarascon, J.-M.; Aymard, L. Metal hydrides for lithium-ion batteries. Nat. Mater. 2008, 7, 916–921. [Google Scholar] [CrossRef] [PubMed]
  150. Zaïdi, W.; Oumellal, Y.; Bonnet, J.-P.; Zhang, J.; Cuevas, F.; Latroche, M.; Bobet, J.-L.; Aymard, L. Carboxymethylcellulose and carboxymethylcellulose-formate as binders in MgH2–carbon composites negative electrode for lithium-ion batteries. J. Power Sources 2011, 196, 2854–2857. [Google Scholar] [CrossRef]
  151. Hanada, N.; Kamura, A.; Suzuki, H.; Takai, K.; Ichikawa, T.; Kojima, Y. Electrochemical charge and discharge properties for the formation of magnesium and aluminum hydrides. J. Alloys Compd. 2011, 509, S584–S587. [Google Scholar] [CrossRef]
  152. Brutti, S.; Mulas, G.; Piciollo, E.; Panero, S.; Reale, P. Magnesium hydride as a high capacity negative electrode for lithium ion batteries. J. Mater. Chem. 2012, 22, 14531–14537. [Google Scholar] [CrossRef]
  153. Ikeda, S.; Ichikawa, T.; Kawahito, K.; Hirabayashi, K.; Miyaoka, H.; Kojima, Y. Anode properties of magnesium hydride catalyzed with niobium oxide for an all solid-state lithium-ion battery. Chem. Commun. 2013, 49, 7174–7176. [Google Scholar] [CrossRef] [PubMed]
  154. Oumellal, Y.; Zlotea, C.; Bastide, S.; Cachet-Vivier, C.; Léonel, E.; Sengmany, S.; Leroy, E.; Aymard, L.; Bonnet, J.-P.; Latroche, M. Bottom-up preparation of MgH2 nanoparticles with enhanced cycle life stability during electrochemical conversion in Li-ion batteries. Nanoscale 2014, 6, 14459–14466. [Google Scholar] [CrossRef] [PubMed]
  155. Ikeda, S.; Ichikawa, T.; Goshome, K.; Yamaguchi, S.; Miyaoka, H.; Kojima, Y. Anode properties of Al2O3-added MgH2 for all-solid-state lithium-ion batteries. J. Solid State Electrochem. 2015, 12, 3639–3644. [Google Scholar] [CrossRef]
  156. Meggiolaro, D.; Gigli, G.; Paolone, A.; Reale, P.; Doublet, M.L.; Brutti, S. Origin of the Voltage Hysteresis of MgH2 Electrodes in Lithium Batteries. J. Phys. Chem. C 2015, 119, 17044–17052. [Google Scholar] [CrossRef]
  157. Zeng, L.; Ichikawa, T.; Kawahito, K.; Miyaoka, H.; Kojima, Y. Bulk-Type All-Solid-State Lithium-Ion Batteries: Remarkable Performances of a Carbon Nanofiber-Supported MgH2 Composite Electrode. ACS Appl. Mater. Interfaces 2017, 9, 2261–2266. [Google Scholar] [CrossRef] [PubMed]
  158. Peng, X.; Wang, H.; Hu, R.; Ouyang, L.; Liu, J.; Zhu, M. Electrochemical performances of MgH2 and MgH2-C films for lithium ion battery anode. J. Alloys Compd. 2017, 711, 473–479. [Google Scholar] [CrossRef]
  159. Oumellal, Y.; Zaïdi, W.; Bonnet, J.-P.; Cuevas, F.; Latroche, M.; Zhang, J.; Bobet, J.-L.; Rougier, A.; Aymard, L. Reactivity of TiH2 hydride with lithium ion: Evidence for a new conversion mechanism. Int. J. Hydrogen Energy 2012, 37, 7831–7835. [Google Scholar] [CrossRef]
  160. Kawahito, K.; Zeng, L.; Ichikawa, T.; Miyaoka, H.; Kojima, Y. Electrochemical Performance of Titanium Hydride for Bulk-Type All-Solid-State Lithium-Ion Batteries. Mater. Trans. 2016, 57, 755–757. [Google Scholar] [CrossRef]
  161. Teprovich, J.A.; 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]
  162. Oumellal, Y.; Rougier, A.; Tarascon, J.-M.; Aymard, L. 2LiH + M (M = Mg, Ti): New concept of negative electrode for rechargeable lithium-ion batteries. J. Power Sources 2009, 192, 698–702. [Google Scholar] [CrossRef]
  163. Huang, L.; Aymard, L.; Bonnet, J.-P. MgH2–TiH2 mixture as an anode for lithium-ion batteries: Synergic enhancement of the conversion electrode electrochemical performance. J. Mater. Chem. A 2015, 3, 15091–15096. [Google Scholar] [CrossRef]
  164. Zaïdi, W.; Bonnet, J.-P.; Zhang, J.; Cuevas, F.; Latroche, M.; Couillaud, S.; Bobet, J.-L.; Sougrati, M.T.; Jumas, J.-C.; Aymard, L. Reactivity of complex hydrides Mg2FeH6, Mg2CoH5 and Mg2NiH4 with lithium ion: Far from equilibrium electrochemically driven conversion reactions. Int. J. Hydrogen Energy 2013, 38, 4798–4808. [Google Scholar] [CrossRef]
  165. Aymard, L.; Oumellal, Y.; Bonnet, J.-P. Metal hydrides: An innovative and challenging conversion reaction anode for lithium-ion batteries. Beilstein J. Nanotechnol. 2015, 6, 1821–1839. [Google Scholar] [CrossRef] [PubMed]
  166. Sartori, S.; Cuevas, F.; Latroche, M. Metal hydrides used as negative electrode materials for Li-ion batteries. Appl. Phys. A 2016, 2, 1–7. [Google Scholar] [CrossRef]
  167. Silvestri, L.; Forgia, S.; Farina, L.; Meggiolaro, D.; Panero, S.; La Barbera, A.; Brutti, S.; Reale, P. Lithium Alanates as Negative Electrodes in Lithium-Ion Batteries. ChemElectroChem 2015, 2, 877–886. [Google Scholar] [CrossRef]
  168. Silvestri, L.; Farina, L.; Meggiolaro, D.; Panero, S.; Padella, F.; Brutti, S.; Reale, P. Reactivity of Sodium Alanates in Lithium Batteries. J. Phys. Chem. C 2015, 119, 28766–28775. [Google Scholar] [CrossRef]
  169. Cirrincione, L.; Silvestri, L.; Mallia, C.; Stallworth, P.E.; Greenbaum, S.; Brutti, S.; Panero, S.; Reale, P. Investigation of the Effects of Mechanochemical Treatment on NaAlH4 Based Anode Materials for Li-Ion Batteries. J. Electrochem. Soc. 2016, 163, A2628–A2635. [Google Scholar] [CrossRef]
  170. Silvestri, L.; Paolone, A.; Cirrincione, L.; Stallworth, P.; Greenbaum, S.; Panero, S.; Brutti, S.; Reale, P. NaAlH4 Nanoconfinement in a Mesoporous Carbon for Application in Lithium Ion Batteries. J. Electrochem. Soc. 2017, 164, A1120–A1125. [Google Scholar] [CrossRef]
  171. Huen, P.; Peru, F.; Charalambopoulou, G.; Steriotis, T.A.; Jensen, T.R.; Ravnsbæk, D.B. Nanoconfined NaAlH4 Conversion Electrodes for Li Batteries. ACS Omega 2017, 2, 1956–1967. [Google Scholar] [CrossRef]
  172. Meggiolaro, D.; Farina, L.; Silvestri, L.; Panero, S.; Brutti, S.; Reale, P. Lightweight Borohydrides Electro-Activity in Lithium Cells. Energies 2016, 9, 238. [Google Scholar] [CrossRef] [Green Version]
  173. Mason, T.H.; Liu, X.; Hong, J.; Graetz, J.; Majzoub, E.H. First-Principles Study of Novel Conversion Reactions for High-Capacity Li-Ion Battery Anodes in the Li–Mg–B–N–H System. J. Phys. Chem. C 2011, 115, 16681–16687. [Google Scholar] [CrossRef]
  174. Unemoto, A.; Matsuo, M.; Orimo, S. Complex Hydrides for Electrochemical Energy Storage. Adv. Funct. Mater. 2014, 24, 2267–2279. [Google Scholar] [CrossRef]
  175. Matsuo, M.; Orimo, S. Lithium Fast-Ionic Conduction in Complex Hydrides: Review and Prospects. Adv. Energy Mater. 2011, 1, 161–172. [Google Scholar] [CrossRef]
  176. De Jongh, P.E.; Blanchard, D.; Matsuo, M.; Udovic, T.J.; Orimo, S. Complex hydrides as room-temperature solid electrolytes for rechargeable batteries. Appl. Phys. A 2016, 122, 251. [Google Scholar] [CrossRef] [Green Version]
  177. Matsuo, M.; Nakamori, Y.; Orimo, S.; Maekawa, H.; Takamura, H. Lithium superionic conduction in lithium borohydride accompanied by structural transition. Appl. Phys. Lett. 2007, 91, 224103. [Google Scholar] [CrossRef]
  178. Maekawa, H.; Matsuo, M.; Takamura, H.; Ando, M.; Noda, Y.; Karahashi, T.; Orimo, S. Halide-Stabilized LiBH4, a Room-Temperature Lithium Fast-Ion Conductor. J. Am. Chem. Soc. 2009, 131, 894–895. [Google Scholar] [CrossRef] [PubMed]
  179. Rude, L.H.; Groppo, E.; Arnbjerg, L.M.; Ravnsbæk, D.B.; Malmkjær, R.A.; Filinchuk, Y.; Baricco, M.; Besenbacher, F.; Jensen, T.R. Iodide substitution in lithium borohydride, LiBH4-LiI. J. Alloys Compd. 2011, 509, 8299–8305. [Google Scholar] [CrossRef]
  180. Mosegaard, L.; Møller, B.; Jørgensen, J.-E.; Filinchuk, Y.; Cerenius, Y.; Hanson, J.C.; Dimasi, E.; Besenbacher, F.; Jensen, T.R. Reactivity of LiBH4:  In Situ Synchrotron Radiation Powder X-ray Diffraction Study. J. Phys. Chem. C 2008, 112, 1299–1303. [Google Scholar] [CrossRef]
  181. Mohtadi, R.; Matsui, M.; Arthur, T.S.; Hwang, S.-J. Magnesium Borohydride: From Hydrogen Storage to Magnesium Battery. Angew. Chem. Int. Ed. 2012, 51, 9780–9783. [Google Scholar] [CrossRef] [PubMed]
  182. Roedern, E.; Kühnel, R.-S.; Remhof, A.; Battaglia, C. Magnesium Ethylenediamine Borohydride as Solid-State Electrolyte for Magnesium Batteries. Sci. Rep. 2017, 7, 46189. [Google Scholar] [CrossRef] [PubMed]
  183. Higashi, S.; Miwa, K.; Aoki, M.; Takechi, K. A novel inorganic solid state ion conductor for rechargeable Mg batteries. Chem. Commun. 2014, 50, 1320–1322. [Google Scholar] [CrossRef] [PubMed]
  184. Udovic, T.J.; Matsuo, M.; Tang, W.S.; Wu, H.; Stavila, V.; Soloninin, A.V.; Skoryunov, R.V.; Babanova, O.A.; Skripov, A.V.; Rush, J.J.; et al. Exceptional Superionic Conductivity in Disordered Sodium Decahydro-closo-decaborate. Adv. Mater. 2014, 26, 7622–7626. [Google Scholar] [CrossRef] [PubMed]
  185. Udovic, T.J.; Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Rush, J.J.; Takamura, H.; Orimo, S. Sodium superionic conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750. [Google Scholar] [CrossRef] [PubMed]
  186. Ravnsbæk, D.B.; Sørensen, L.H.; Filinchuk, Y.; Reed, D.; Book, D.; Jakobsen, H.J.; Besenbacher, F.; Skibsted, J.; Jensen, T.R. Mixed-Anion and Mixed-Cation Borohydride KZn(BH4)Cl2: Synthesis, Structure and Thermal Decomposition. Eur. J. Inorg. Chem. 2010, 2010, 1608–1612. [Google Scholar] [CrossRef]
  187. Ravnsbæk, D.B.; Ley, M.B.; Lee, Y.-S.; Hagemann, H.; D’Anna, V.; Cho, Y.W.; Filinchuk, Y.; Jensen, T.R. A mixed-cation mixed-anion borohydride NaY(BH4)2Cl2. Int. J. Hydrogen Energy 2012, 37, 8428–8438. [Google Scholar] [CrossRef]
  188. Jaroń, T.; Grochala, W. Probing Lewis acidity of Y(BH4)3 via its reactions with MBH4 (M = Li, Na, K, NMe4). Dalton Trans. 2011, 40, 12808–12817. [Google Scholar] [CrossRef] [PubMed]
  189. Černý, R.; Ravnsbæk, D.B.; Schouwink, P.; Filinchuk, Y.; Penin, N.; Teyssier, J.; Smrčok, Ľ.; Jensen, T.R. Potassium Zinc Borohydrides Containing Triangular [Zn(BH4)3] and Tetrahedral [Zn(BH4)xCl4−x]2 Anions. J. Phys. Chem. C 2012, 116, 1563–1571. [Google Scholar] [CrossRef]
  190. Rude, L.H.; Filsø, U.; D’Anna, V.; Spyratou, A.; Richter, B.; Hino, S.; Zavorotynska, O.; Baricco, M.; Sørby, M.H.; Hauback, B.C.; et al. Hydrogen–fluorine exchange in NaBH4–NaBF4. Phys. Chem. Chem. Phys. 2013, 15, 18185–18194. [Google Scholar] [CrossRef] [PubMed]
  191. Richter, B.; Ravnsbæk, D.B.; Sharma, M.; Spyratou, A.; Hagemann, H.; Jensen, T.R. Fluoride substitution in LiBH4: Destabilization and decomposition. Phys. Chem. Chem. Phys. 2017, in press. [Google Scholar] [CrossRef]
  192. Ley, M.B.; Boulineau, S.; Janot, R.; Filinchuk, Y.; Jensen, T.R. New Li Ion Conductors and Solid State Hydrogen Storage Materials: LiM(BH4)3Cl, M = La, Gd. J. Phys. Chem. C 2012, 116, 21267–21276. [Google Scholar] [CrossRef]
  193. Frommen, C.; Sørby, M.H.; Ravindran, P.; Vajeeston, P.; Fjellvåg, H.; Hauback, B.C. Synthesis, Crystal Structure, and Thermal Properties of the First Mixed-Metal and Anion-Substituted Rare Earth Borohydride LiCe(BH4)3Cl. J. Phys. Chem. C 2011, 115, 23591–23602. [Google Scholar] [CrossRef] [Green Version]
  194. Olsen, J.E.; Frommen, C.; Jensen, T.R.; Riktor, M.D.; Sørby, M.H.; Hauback, B.C. Structure and thermal properties of composites with RE-borohydrides (RE = La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, Er, Yb or Lu) and LiBH4. RSC Adv. 2013, 4, 1570–1582. [Google Scholar] [CrossRef]
  195. Skripov, A.V.; Soloninin, A.V.; Ley, M.B.; Jensen, T.R.; Filinchuk, Y. Nuclear Magnetic Resonance Studies of BH4 Reorientations and Li Diffusion in LiLa (BH4)3Cl. J. Phys. Chem. C 2013, 117, 14965–14972. [Google Scholar] [CrossRef]
  196. Lundén, A. On the Paddle -Wheel Mechanism for Cation Conduction in Lithium Sulphate. Z. Naturforschung A 2014, 50, 1067–1076. [Google Scholar] [CrossRef]
  197. Payandeh GharibDoust, S.; Brighi, M.; Sadikin, Y.; Ravnsbæk, D.B.; Černý, R.; Skibsted, J.; Jensen, T.R. Synthesis, Structure, and Li-Ion Conductivity of LiLa(BH4)3X, X = Cl, Br, I. J. Phys. Chem. C 2017, 121, 19010–19021. [Google Scholar] [CrossRef]
  198. Lee, Y.-S.; Ley, M.B.; Jensen, T.R.; Cho, Y.W. Lithium Ion Disorder and Conduction Mechanism in LiCe(BH4)3Cl. J. Phys. Chem. C 2016, 120, 19035–19042. [Google Scholar] [CrossRef]
  199. Lee, Y.-S.; Filinchuk, Y.; Lee, H.-S.; Suh, J.-Y.; Kim, J.W.; Yu, J.-S.; Cho, Y.W. On the Formation and the Structure of the First Bimetallic Borohydride Borate, LiCa3(BH4)(BO3)2. J. Phys. Chem. C 2011, 115, 10298–10304. [Google Scholar] [CrossRef]
  200. Riktor, M.D.; Filinchuk, Y.; Vajeeston, P.; Bardají, E.G.; Fichtner, M.; Fjellvåg, H.; Sørby, M.H.; Hauback, B.C. The crystal structure of the first borohydride borate, Ca3(BD4)3(BO3). J. Mater. Chem. 2011, 21, 7188. [Google Scholar] [CrossRef]
  201. Paskevicius, M.; Hansen, B.R.S.; Jørgensen, M.; Richter, B.; Jensen, T.R. Multifunctionality of Silver closo-Boranes. Nat. Commun. 2017. [Google Scholar] [CrossRef] [PubMed]
  202. Hansen, B.R.S.; Paskevicius, M.; Jørgensen, M.; Jensen, T.R. Halogenated Sodium-closo-Dodecaboranes as Solid-State Ion Conductors. Chem. Mater. 2017, 29, 3423–3430. [Google Scholar] [CrossRef]
  203. Bukovsky, E.V.; Peryshkov, D.V.; Wu, H.; Zhou, W.; Tang, W.S.; Jones, W.M.; Stavila, V.; Udovic, T.J.; Strauss, S.H. Comparison of the Coordination of B12F122−, B12Cl122−, and B12H122− to Na+ in the Solid State: Crystal Structures and Thermal Behavior of Na2(B12F12), Na2(H2O)4(B12F12), Na2(B12Cl12), and Na2(H2O)6(B12Cl12). Inorg. Chem. 2017, 56, 4369–4379. [Google Scholar] [CrossRef] [PubMed]
  204. Tang, W.S.; Unemoto, A.; Zhou, W.; Stavila, V.; Matsuo, M.; Wu, H.; Orimo, S.; Udovic, T.J. Unparalleled lithium and sodium superionic conduction in solid electrolytes with large monovalent cage-like anions. Energy Environ. Sci. 2015, 8, 3637–3645. [Google Scholar] [CrossRef] [PubMed]
  205. Sandrock, G.; Thomas, G. The IEA/DOE/SNL on-line hydride databases. Appl. Phys. A 2001, 72, 153–155. [Google Scholar] [CrossRef]
  206. Johnson, J.R. Reaction of hydrogen with the high temperature (C14) form of TiCr2. J. Common Met. 1980, 73, 345–354. [Google Scholar] [CrossRef]
  207. Manchester, F.D. Phase Diagrams of Binary Hydrogen Alloys; Monograph Series on Alloy Phase Diagrams, 13; ASM International: Materials Park, OH, USA, 2000; ISBN 0-87170-587-7. [Google Scholar]
  208. Sandrock, G.; Suda, S.; Schlapbach, L. Hydrogen in Intermetallic Compounds II, Topics in Applied Physics; Springer: Berlin/Heidelberg, Germany, 1992; Volume 67. [Google Scholar]
  209. Dantzer, P. Properties of intermetallic compounds suitable for hydrogen storage applications. Mater. Sci. Eng. A 2002, 329–331, 313–320. [Google Scholar] [CrossRef]
  210. Sandrock, G.; Bowman, R.C. Gas-based hydride applications: Recent progress and future needs. J. Alloys Compd. 2003, 356, 794–799. [Google Scholar] [CrossRef]
  211. Lototskyy, M.V.; Yartys, V.A.; Pollet, B.G.; Bowman, R.C., Jr. Metal hydride hydrogen compressors: A review. Int. J. Hydrogen Energy 2014, 39, 5818–5851. [Google Scholar] [CrossRef]
  212. SunShot Vision Study. Chapter 5: Concentrating Solar Power: Technologies, Cost, and Performance; US Department of Energy: Washington, DC, USA, 2012. [Google Scholar]
  213. Rönnebro, E.C.E.; Whyatt, G.; Powell, M.; Westman, M.; Zheng, F.; Fang, Z.Z. Metal Hydrides for High-Temperature Power Generation. Energies 2015, 8, 8406–8430. [Google Scholar] [CrossRef]
  214. Sheppard, D.A.; Paskevicius, M.; Humphries, T.D.; Felderhoff, M.; Capurso, G.; von Colbe, J.B.; Dornheim, M.; Klassen, T.; Ward, P.A.; Teprovich, J.A.; et al. Metal hydrides for concentrating solar thermal power energy storage. Appl. Phys. A 2016, 122, 395. [Google Scholar] [CrossRef]
  215. Sheppard, D.A.; Humphries, T.D.; Buckley, C.E. Sodium-based hydrides for thermal energy applications. Appl. Phys. A 2016, 122, 406. [Google Scholar] [CrossRef]
  216. Fellet, M.; Buckley, C.E.; Paskevicius, M.; Sheppard, D.A. Research on metal hydrides revived for next-generation solutions to renewable energy storage. MRS Bull. 2013, 38, 1012–1013. [Google Scholar] [CrossRef]
  217. Ward, P.A.; Corgnale, C.; Teprovich, J.A.; Motyka, T.; Hardy, B.; Sheppard, D.; Buckley, C.; Zidan, R. Technical challenges and future direction for high-efficiency metal hydride thermal energy storage systems. Appl. Phys. A 2016, 122, 462. [Google Scholar] [CrossRef]
  218. Pardo, P.; Deydier, A.; Anxionnaz-Minvielle, Z.; Rougé, S.; Cabassud, M.; Cognet, P. A review on high temperature thermochemical heat energy storage. Renew. Sustain. Energy Rev. 2014, 32, 591–610. [Google Scholar] [CrossRef] [Green Version]
  219. Corgnale, C.; Hardy, B.; Motyka, T.; Zidan, R.; Teprovich, J.; Peters, B. Screening analysis of metal hydride based thermal energy storage systems for concentrating solar power plants. Renew. Sustain. Energy Rev. 2014, 38, 821–833. [Google Scholar] [CrossRef]
  220. Harries, D.N.; Paskevicius, M.; Sheppard, D.A.; Price, T.E.C.; Buckley, C.E. Concentrating Solar Thermal Heat Storage Using Metal Hydrides. Proc. IEEE 2012, 100, 539–549. [Google Scholar] [CrossRef]
  221. Sheppard, D.A.; Humphries, T.D.; Buckley, C.E. What is old is new again. Mater. Today 2015, 8, 414–415. [Google Scholar] [CrossRef]
  222. Felderhoff, M.; Urbanczyk, R.; Peil, S. Thermochemical heat storage for high temperature applications—A review. Green 2013, 3, 113–123. [Google Scholar] [CrossRef]
  223. Paskevicius, M.; Sheppard, D.A.; Williamson, K.; Buckley, C.E. Metal hydride thermal heat storage prototype for concentrating solar thermal power. Energy 2015, 88, 469–477. [Google Scholar] [CrossRef]
  224. Bogdanović, B.; Spliethoff, B.; Ritter, A. The magnesium hydride system for heat storage and cooling. Z. Phys. Chem. 1989, 164, 1497–1508. [Google Scholar] [CrossRef]
  225. Groll, M.; Isselhorst, A.; Wierse, M. Metal hydride devices for environmentally clean energy technology. Int. J. Hydrogen Energy 1994, 19, 507–515. [Google Scholar] [CrossRef]
  226. Bogdanović, B.; Ritter, A.; Spliethoff, B.; Straβburger, K. A process steam generator based on the high temperature magnesium hydride/magnesium heat storage system. Int. J. Hydrogen Energy 1995, 20, 811–822. [Google Scholar] [CrossRef]
  227. Dong, D.; Humphries, T.D.; Sheppard, D.A.; Stansby, B.; Paskevicius, M.; Sofianos, M.V.; Chaudhary, A.-L.; Dornheim, M.; Buckley, C.E. Thermal optimisation of metal hydride reactors for thermal energy storage applications. Sustain. Energy Fuels 2017. [Google Scholar] [CrossRef]
  228. Humphries, T.D.; Sheppard, D.A.; Rowles, M.R.; Sofianos, M.V.; Buckley, C.E. Fluoride substitution in sodium hydride for thermal energy storage applications. J. Mater. Chem. A 2016, 4, 12170–12178. [Google Scholar] [CrossRef]
  229. Sheppard, D.A.; Paskevicius, M.; Buckley, C.E. Thermodynamics of Hydrogen Desorption from NaMgH3 and Its Application as a Solar Heat Storage Medium. Chem. Mater. 2011, 23, 4298–4300. [Google Scholar] [CrossRef]
  230. Sheppard, D.A.; Corgnale, C.; Hardy, B.; Motyka, T.; Zidan, R.; Paskevicius, M.; Buckley, C.E. Hydriding characteristics of NaMgH2F with preliminary technical and cost evaluation of magnesium-based metal hydride materials for concentrating solar power thermal storage. RSC Adv. 2014, 4, 26552–26562. [Google Scholar] [CrossRef]
  231. Javadian, P.; Sheppard, D.A.; Jensen, T.R.; Buckley, C.E. Destabilization of lithium hydride and the thermodynamic assessment of the Li–Al–H system for solar thermal energy storage. Rsc Adv. 2016, 6, 94927–94933. [Google Scholar] [CrossRef]
  232. Doi, K.; Hino, S.; Miyaoka, H.; Ichikawa, T.; Kojima, Y. Hydrogen storage properties of lithium silicon alloy synthesized by mechanical alloying. J. Power Sources 2011, 196, 504–507. [Google Scholar] [CrossRef]
  233. Vajo, J.J.; Mertens, F.; Ahn, C.C.; Robert, C.B.; Fultz, B. Altering Hydrogen Storage Properties by Hydride Destabilization through Alloy Formation: LiH and MgH2 Destabilized with Si. J. Phys. Chem. B 2004, 108, 13977–13983. [Google Scholar] [CrossRef]
  234. Veleckis, E. Application of the hydrogen titration method to a thermodynamic investigation of solid Al-Ca alloys. J. Common Met. 1981, 80, 241–255. [Google Scholar] [CrossRef]
  235. Reilly, J.J.; Wiswall, R.H. Reaction of hydrogen with alloys of magnesium and nickel and the formation of Mg2NiH4. Inorg. Chem. 1968, 7, 2254–2256. [Google Scholar] [CrossRef]
  236. Reiser, A.; Bogdanović, B.; Schlichte, K. The application of Mg-based metal-hydrides as heat energy storage systems. Int. J. Hydrogen Energy 2000, 25, 425–430. [Google Scholar] [CrossRef]
  237. Bogdanović, B.; Reiser, A.; Schlichte, K.; Spliethoff, B.; Tesche, B. Thermodynamics and dynamics of the Mg–Fe–H system and its potential for thermochemical thermal energy storage. J. Alloys Compd. 2002, 345, 77–89. [Google Scholar] [CrossRef]
  238. Nguyen, T.T.; Sheppard, D.A.; Buckley, C.E. Lithium imide systems for high temperature heat storage in concentrated solar thermal systems. J. Alloys Compd. 2017, 716, 291–298. [Google Scholar] [CrossRef]
  239. Li, Y.; Li, P.; Qu, X. Investigation on LiBH4-CaH2 composite and its potential for thermal energy storage. Sci. Rep. 2017, 7, 41754. [Google Scholar] [CrossRef] [PubMed]
  240. El Kharbachi, A.; Pinatel, E.; Nuta, I.; Baricco, M. A thermodynamic assessment of LiBH4. Calphad 2012, 39, 80–90. [Google Scholar] [CrossRef]
  241. HSC Chemistry, version 6.12; Outotec Research Oy: Pori, Finland, 2006.
  242. Javadian, P.; GharibDoust, S.P.; Li, H.-W.; Sheppard, D.A.; Buckley, C.E.; Jensen, T.R. Reversibility of LiBH4 Facilitated by the LiBH4-Ca(BH4)2 Eutectic. J. Phys. Chem. C 2017. [Google Scholar] [CrossRef]
  243. 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]
  244. 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]
  245. 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]
  246. Sørby, M.H.; Brinks, H.W.; Fossdal, A.; Thorshaug, K.; Hauback, B.C. The crystal structure and stability of K2NaAlH6. J. Alloys Compd. 2006, 415, 284–287. [Google Scholar] [CrossRef]
  247. Mamatha, M.; Weidenthaler, 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]
  248. Varin, R.A.; Jang, M. The effects of graphite on the reversible hydrogen storage of nanostructured lithium amide and lithium hydride (LiNH2 + 1.2LiH) system. J. Alloys Compd. 2011, 509, 7143–7151. [Google Scholar] [CrossRef]
  249. Paskevicius, M.; Sheppard, D.A.; Buckley, C.E. Thermodynamic Changes in Mechanochemically Synthesized Magnesium Hydride Nanoparticles. J. Am. Chem. Soc. 2010, 132, 5077–5083. [Google Scholar] [CrossRef] [PubMed]
  250. Kenisarin, M.M. High-temperature phase change materials for thermal energy storage. Renew. Sustain. Energy Rev. 2010, 14, 955–970. [Google Scholar] [CrossRef]
  251. Wagman, D.D.; Evans, W.H.; Parker, V.B.; Schumm, R.H.; Halow, I.; Bailey, S.M.; Churney, K.L.; Nuttall, R.L. The NBS tables of chemical thermodynamic properties. Selected values for inorganic and C1 and C2 organic substances in SI units. J. Phys. Chem. Ref. Data 1982, 11 (Suppl. 2). [Google Scholar] [CrossRef]
  252. Gavrichev, K.S. Heat Capacity and Thermodynamic Properties of Inorganic Compounds Containing Tetrahedral Anions (BH-4, AlH-4, GaH-4, BF-4, ClO-4, BrO-4, and IO-4). Inorg. Mater. 2003, 39, S89–S112. [Google Scholar] [CrossRef]
  253. Juza, R.; Opp, K. Metallamide und Metallnitride, 24. Mitteilung. Die Kristallstruktur des Lithiumamides. Z. Anorg. Allg. Chem. 1951, 266, 313–324. [Google Scholar] [CrossRef]
  254. Stasinevich, D.S.; Egorenko, G.A. Thermographic investigation of alkali metal and magnesium tetrahydroborates at pressures up to 10 atm. Russ. J. Inorg. Chem. 1968, 13, 341–343. [Google Scholar]
  255. Paskevicius, M.; Ley, M.B.; Sheppard, D.A.; Jensen, T.R.; Buckley, C.E. Eutectic melting in metal borohydrides. Phys. Chem. Chem. Phys. 2013, 15, 19774–19789. [Google Scholar] [CrossRef] [PubMed]
  256. Liu, Z.-K.; Zhong, Y.; Schlom, D.G.; Xi, X.X.; Li, Q. Computational thermodynamic modeling of the Mg-B system. Calphad 2001, 25, 299–303. [Google Scholar] [CrossRef]
  257. Bogdanović, B.; Bohmhammel, K.; Christ, B.; Reiser, A.; Schlichte, K.; Vehlen, R.; Wolf, U. Thermodynamic investigation of the magnesium–hydrogen system. J. Alloys Compd. 1999, 282, 84–92. [Google Scholar] [CrossRef]
  258. Bogdanović, B.; Hofmann, H.; Neuy, A.; Reiser, A.; Schlichte, K.; Spliethoff, B.; Wessel, S. Ni-doped versus undoped Mg–MgH2 materials for high temperature heat or hydrogen storage. J. Alloys Compd. 1999, 292, 57–71. [Google Scholar] [CrossRef]
  259. Marcus, Y. Chapter 3—High Melting Salts in Ionic Liquid Properties: From Molten Salts to RTILs; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar]
  260. Sohal, M.S.; Ebner, M.A.; Sabharwall, P.; Sharpe, P. Engineering Database of Liquid Salt Thermophysical and Thermochemical Properties; Idaho National Laboratory (INL): Idaho Falls, ID, USA, 2010. [Google Scholar]
  261. Lehman, P.A.; Chamberlin, C.E. Design of a photovoltaic-hydrogen-fuel cell energy system. Int. J. Hydrogen Energy 1991, 16, 349–352. [Google Scholar] [CrossRef]
  262. Vanhanen, J.P.; Lund, P.D.; Hagström, M.T. Feasibility study of a metal hydride hydrogen store for a self-sufficient solar hydrogen energy system. Int. J. Hydrogen Energy 1996, 21, 213–221. [Google Scholar] [CrossRef]
  263. Gray, E.M.; Webb, C.J.; Andrews, J.; Shabani, B.; Tsai, P.J.; Chan, S.L.I. Hydrogen storage for off-grid power supply. Int. J. Hydrogen Energy 2011, 36, 654–663. [Google Scholar] [CrossRef]
  264. Lototskyy, M.V.; Tolj, I.; Pickering, L.; Sita, C.; Barbir, F.; Yartys, V. The use of metal hydrides in fuel cell applications. Prog. Nat. Sci. Mater. Int. 2017, 27, 3–20. [Google Scholar] [CrossRef]
  265. Lu, L.; Han, X.; Li, J.; Hua, J.; Ouyang, M. A review on the key issues for lithium-ion battery management in electric vehicles. J. Power Sources 2013, 226, 272–288. [Google Scholar] [CrossRef]
  266. Vetter, J.; Novák, P.; Wagner, M.R.; Veit, C.; Möller, K.-C.; Besenhard, J.O.; Winter, M.; Wohlfahrt-Mehrens, M.; Vogler, C.; Hammouche, A. Ageing mechanisms in lithium-ion batteries. J. Power Sources 2005, 147, 269–281. [Google Scholar] [CrossRef]
  267. Armand, M.; Tarascon, J.-M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  268. Khalil, Y.F.; Opalka, S.M.; Laube, B.L. Experimental and theoretical investigations for mitigating NaAlH4 reactivity risks during postulated accident scenarios involving exposure to air or water. Process Saf. Environ. Prot. 2013, 91, 463–475. [Google Scholar] [CrossRef]
  269. Chaise, A.; de Rango, P.; Marty, P.; Fruchart, D.; Miraglia, S.; Olivès, R.; Garrier, S. Enhancement of hydrogen sorption in magnesium hydride using expanded natural graphite. Int. J. Hydrogen Energy 2009, 34, 8589–8596. [Google Scholar] [CrossRef]
  270. Maleki, H.; Howard, J.N. Effects of overdischarge on performance and thermal stability of a Li-ion cell. J. Power Sources 2006, 160, 1395–1402. [Google Scholar] [CrossRef]
  271. Goodenough, J.B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  272. Soloveichik, G.L. Battery Technologies for Large-Scale Stationary Energy Storage. Annu. Rev. Chem. Biomol. Eng. 2011, 2, 503–527. [Google Scholar] [CrossRef] [PubMed]
  273. Zaghib, K.; Dontigny, M.; Guerfi, A.; Charest, P.; Rodrigues, I.; Mauger, A.; Julien, C.M. Safe and fast-charging Li-ion battery with long shelf life for power applications. J. Power Sources 2011, 196, 3949–3954. [Google Scholar] [CrossRef]
  274. Uchida, H.; Naragaki, Y. Influence of the Cyclic Hydriding-Dehydriding Treatment on Pressure-Composition-Temperature Relations of the LaNi5—H System. Z. Phys. Chem. 1993, 179, 93–101. [Google Scholar] [CrossRef]
  275. Bershadsky, E.; Josephy, Y.; Ron, M. Investigation of kinetics and structural changes in TiFe0.8Ni0.2 after prolonged cycling. J. Common Met. 1991, 172, 1036–1043. [Google Scholar] [CrossRef]
  276. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the development of advanced Li-ion batteries: A review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar] [CrossRef]
  277. Urbanczyk, R.; Bathen, D.; Felderhoff, M.; Burfeind, J. HT-PEM fuel cell system with integrated complex hydride storage tank. In Parallel Sessions Book 4: Storage Systems/Policy Perspectives, Initiatives and Co-operations, Proceedings of the 18th World Hydrogen Energy Conference 2010 (WHEC2010), Essen, Germany, 16–21 May 2010; Stolten, D., Grube, T., Eds.; Forschungszentrum Jülich GmbH: Jülich, Germany, 2010; pp. 177–183. [Google Scholar]
  278. Farnes, J.; Vik, A.; Bokach, D.; Svendsen, T.; Schautz, M.; Geneste, X. Recent developments of regenerative fuel cell systems for satellites. In Proceedings of the 10th European Space Power Conference, Noordwijkerhout, The Netherlands, 13–17 April 2014; Available online: http://www.esa-tec.eu/workspace/assets/files/tdo0157-paper-regenerative-f-5504a3929eb6a.pdf (accessed on 16 October 2017).
  279. Reissner, A.; Pawelke, R.H.; Hummel, S.; Cabelka, D.; Gerger, J.; Farnes, J.; Vik, A.; Wernhus, I.; Svendsen, T.; Schautz, M.; et al. Metal hydride hydrogen and heat storage systems as enabling technology for spacecraft applications. J. Alloys Compd. 2015, 645, S9–S13. [Google Scholar] [CrossRef]
  280. Pfeifer, P.; Wall, C.; Jensen, O.; Hahn, H.; Fichtner, M. Thermal coupling of a high temperature PEM fuel cell with a complex hydride tank. Int. J. Hydrogen Energy 2009, 34, 3457–3466. [Google Scholar] [CrossRef]
  281. Nasri, M.; Dickenson, D. Thermal Management of Fuel Cell-driven Vehicles using HT-PEM and Hydrogen Storage. In Proceedings of the Ninth International Conference on Ecological Vehicles and Renewable Energies, Monte-Carlo, Monaco, 25–27 March 2014; pp. 1–6. [Google Scholar]
  282. Reddy, E.H.; Jayanti, S. Thermal Coupling Studies of a High Temperature Proton Exchange Membrane Fuel Cell Stack and a Metal Hydride Hydrogen Storage System. Energy Procedia 2012, 29, 254–264. [Google Scholar] [CrossRef]
  283. Hu, J.; Fichtner, M.; Baricco, M. Preparation of Li-Mg-NH hydrogen storage materials for an auxiliary power unit. Int. J. Hydrogen Energy 2017, 42, 17144–17148. [Google Scholar] [CrossRef]
  284. Delhomme, B.; Lanzini, A.; Ortigoza-Villalba, G.A.; Nachev, S.; de Rango, P.; Santarelli, M.; Marty, P.; Leone, P. Coupling and thermal integration of a solid oxide fuel cell with a magnesium hydride tank. Int. J. Hydrogen Energy 2013, 38, 4740–4747. [Google Scholar] [CrossRef]
  285. Garrier, S.; Chaise, A.; de Rango, P.; Marty, P.; Delhomme, B.; Fruchart, D.; Miraglia, S. MgH2 intermediate scale tank tests under various experimental conditions. Int. J. Hydrogen Energy 2011, 36, 9719–9726. [Google Scholar] [CrossRef]
  286. Parra, D.; Gillott, M.; Walker, G.S. Design, testing and evaluation of a community hydrogen storage system for end user applications. Int. J. Hydrogen Energy 2016, 41, 5215–5229. [Google Scholar] [CrossRef]
  287. Gkanas, E.I.; Makridis, S.S. Effective thermal management of a cylindrical MgH2 tank including thermal coupling with an operating SOFC and the usage of extended surfaces during the dehydrogenation process. Int. J. Hydrogen Energy 2016, 41, 5693–5708. [Google Scholar] [CrossRef]
  288. Yiotis, A.G.; Kainourgiakis, M.E.; Kosmidis, L.I.; Charalambopoulou, G.C.; Stubos, A.K. Thermal coupling potential of Solid Oxide Fuel Cells with metal hydride tanks: Thermodynamic and design considerations towards integrated systems. J. Power Sources 2014, 269, 440–450. [Google Scholar] [CrossRef]
  289. Schouwink, P.; Didelot, E.; Lee, Y.-S.; Mazet, T.; Černý, R. Structural and magnetocaloric properties of novel gadolinium borohydrides. J. Alloys Compd. 2016, 664, 378–384. [Google Scholar] [CrossRef]
  290. Hansen, B.R.S.; Paskevicius, M.; Li, H.-W.; Akiba, E.; Jensen, T.R. Metal boranes: Progress and applications. Coord. Chem. Rev. 2016, 323, 60–70. [Google Scholar] [CrossRef]
  291. McPhy Energy McPhy Energy. Available online: www.mcphy.com (accessed on 16 October 2017).
  292. Huiberts, J.N.; Griessen, R.; Rector, J.H.; Wijngaarden, R.J.; Dekker, J.P.; de Groot, D.G.; Koeman, N.J. Yttrium and lanthanum hydride films with switchable optical properties. Nature 1996, 380, 231–234. [Google Scholar] [CrossRef]
Figure 1. Illustration of a sustainable global energy system based on energy storage in either batteries, as heat storage, using hydrogen as a carrier (left) or a carbon neutral cycle (right). Modified figure from [1].
Figure 1. Illustration of a sustainable global energy system based on energy storage in either batteries, as heat storage, using hydrogen as a carrier (left) or a carbon neutral cycle (right). Modified figure from [1].
Energies 10 01645 g001
Figure 2. Thermal decomposition temperature of metal borohydrides as a function of the central metals electronegativity.
Figure 2. Thermal decomposition temperature of metal borohydrides as a function of the central metals electronegativity.
Energies 10 01645 g002
Figure 3. (Left) Structure of the perovskite-type CsSr(BH4)3 seen along the c-axis consisting of [SrB6] octahedra and Cs+ cations in interstitial positions; (Right) the similarity to a perovskite structure is highlighted with [BH4] groups on cell edge centres, Sr occupying the edge corners and the monovalent metal in the body centre [116].
Figure 3. (Left) Structure of the perovskite-type CsSr(BH4)3 seen along the c-axis consisting of [SrB6] octahedra and Cs+ cations in interstitial positions; (Right) the similarity to a perovskite structure is highlighted with [BH4] groups on cell edge centres, Sr occupying the edge corners and the monovalent metal in the body centre [116].
Energies 10 01645 g003
Figure 4. (a) Theoretical volumetric heat storage capacities versus gravimetric heat storage capacities for complex hydrides with a number of other heat storage materials included for comparison; (b) Hydrogen absorption equilibrium pressures of complex hydrides versus temperature. Data for some well-known heat storage metal hydrides are included as a comparison. Open symbols refer to data for complex hydrides. TCM = thermochemical heat storage material, PCM = phase change heat storage material, SHM = sensible heat storage material.
Figure 4. (a) Theoretical volumetric heat storage capacities versus gravimetric heat storage capacities for complex hydrides with a number of other heat storage materials included for comparison; (b) Hydrogen absorption equilibrium pressures of complex hydrides versus temperature. Data for some well-known heat storage metal hydrides are included as a comparison. Open symbols refer to data for complex hydrides. TCM = thermochemical heat storage material, PCM = phase change heat storage material, SHM = sensible heat storage material.
Energies 10 01645 g004
Figure 5. Potential of various metal hydrides, based on their H2 enthalpy of desorption (ΔHdes), and the associated heat loss (Qloss) of different types of fuel cells. The operating temperature range of the metal hydrides corresponds to H2 absorption equilibrium pressures of between 1 and 150 bar unless otherwise noted in Table 2. The step changes that occur for the NaH and LiBH4 curves are a result of a phase change for either one of the reactants or products. This figure is based on that from reference [264].
Figure 5. Potential of various metal hydrides, based on their H2 enthalpy of desorption (ΔHdes), and the associated heat loss (Qloss) of different types of fuel cells. The operating temperature range of the metal hydrides corresponds to H2 absorption equilibrium pressures of between 1 and 150 bar unless otherwise noted in Table 2. The step changes that occur for the NaH and LiBH4 curves are a result of a phase change for either one of the reactants or products. This figure is based on that from reference [264].
Energies 10 01645 g005
Figure 6. State-of-the-art applications for complex metal hydrides [1,123,290,291,292].
Figure 6. State-of-the-art applications for complex metal hydrides [1,123,290,291,292].
Energies 10 01645 g006
Table 1. Properties of selected light element complex metal hydrides [23,24]. M = molar mass; ρ = volumetric mass density; ρm = gravimetric hydrogen density; ρV = volumetric hydrogen density; ΔHdec = decomposition enthalpy; T(1 bar) = equilibrium temperature at p(H2) = 1 bar; Tdec = decomposition temperature.
Table 1. Properties of selected light element complex metal hydrides [23,24]. M = molar mass; ρ = volumetric mass density; ρm = gravimetric hydrogen density; ρV = volumetric hydrogen density; ΔHdec = decomposition enthalpy; T(1 bar) = equilibrium temperature at p(H2) = 1 bar; Tdec = decomposition temperature.
M (g/mol)ρ (g/mL)ρm (wt % H2)ρV (g H2/L)ΔHdec (kJ/mol)T(1 bar) (°C)Tdec (°C)Ref.
LiBH421.780.6618.4122.574370~400[25]
NaBH437.831.0710.8115.6108534~500[26]
LiAlH437.950.9210.697.5−10-~150 a[27]
Li3AlH653.851.0211.2114.225−81 c~200 a[27]
NaAlH454.001.287.393.433.118~230 a[28]
Na3AlH6102.001.455.985.649.0103~275 a[28]
LiNH222.961.188.8103.667 b-~300[29,30,31]
a Not catalysed. Decomposition temperatures strongly depend on the physical conditions for the measurement and the published data scatter significantly; b Reported for the LiNH2-LiH system; c Calculated based on the van’t Hoff equation using ΔS = 130 J/(mol K).
Table 2. Properties of complex metal hydrides compared to selected heat storage materials. * Volumetric heat storage capacity (in kJ/L) calculated based on the room-temperature crystalline density of materials unless otherwise noted. # The operating temperature ranges correspond to H2 equilibrium pressures that range between 1 bar and 150 bar unless otherwise noted.
Table 2. Properties of complex metal hydrides compared to selected heat storage materials. * Volumetric heat storage capacity (in kJ/L) calculated based on the room-temperature crystalline density of materials unless otherwise noted. # The operating temperature ranges correspond to H2 equilibrium pressures that range between 1 bar and 150 bar unless otherwise noted.
Hydride MaterialsTheoretical H2 Capacity (wt %)ΔHdesHabs (kJ/mol·H2)kJ/kg* kJ/L# Operating Temperature Range (°C)
NaAlH4 ↔ ⅓Na3AlH6 + ⅔Al + H2(g)3.7338.4/−35.2 [243]651.8808.325 a–202 b
Na3AlH6 ↔ 3NaH + Al + 3/2H2(g)2.9647.6/−46.1 [243]678.0983.0100–290
LiNa2AlH6 ↔ 2NaH + LiH + Al + 3/2H2(g)3.5254.95/n.a. [244,245]959.01371.4135–315
NaK2AlH6 ↔ 2KH + NaH + Al + 3/2H2(g)2.2598.0/−98.0 [246]1095.31818.1380–600 c
KAlH4 + LiCl↔KCl + LiH + Al + 3/2H2(g) [247]2.6937.6/−37.6 d501.2728.2111–396
KAlH4 + NaCl ↔ KCl + NaH + Al + 3/2H2(g) [247]2.3562.3/−62.3 d726.61111.8334–574 e
Mg(NH2)2 + 2LiH ↔ Li2Mg(NH)2 + 2H2(g)5.5838.9/n.a. [14]2086.32166.575–280
LiNH2 + LiH ↔ Li2NH + H2(g)6.5264.5/n.a. [30,248]912.62086.1270–375 f
CaNH + CaH2 ↔ Ca2NH + H2(g)2.07n.a./−88.7 [7]1077.81257.8590–780 g
LiBH4(l) ↔ LiH + B + 3/2H2(g)13.8857.3/−57.3[240]3945.72634.9460–688 h
LiBH4(l) ↔ LiH(l) + B + 3/2H2(g)13.8871.7/−71.7 [240]4936.23296.3688 h–1000+
NaBH4 ↔ Na(l) + B + 2H2(g)10.6697.9/−97.9 [241]5176.05559.1~507.5 i
KBH4 ↔ K(l) + B + 2H2(g)7.47114.6/−114.6 [241]4250.14985.9~615.5 j
2LiBH4 + MgH2 ↔ 2LiH + MgB2 + 4H2(g)11.5448.3/−48.3 k2766.62308.0205–467 l
NaH ↔ Na(l) + 1/2H2(g)4.20116.8/−116.8 m2433.52355.7427–638 m
MgH2 ↔ Mg + H2(g)7.6674.1/−74.1 [249]2813.23994.7282–534 n
Mg2NiH4 ↔ Mg2Ni + 2H2(g)3.6264.6/n.a. [235]1159.73142.7253–523
Mg2FeH6 ↔ 2Mg + Fe + 3H2(g)5.4777.4/−77.4 [237]2101.15757.0300–566
TiH1.6 ↔ TiH1.0 + 0.3H2(g)1.22n.a./−165.5 o1003.23772.0645–921
Other Thermochemical Materials (TCM) ΔHdesHabs (kJ/mol Gas Species)kJ/kgkJ/LTemperature Range (°C)
CaCO3 ↔ CaO + CO2(g) 178/−178 [218]17644982.4700–1000 p
Ca(OH)2 ↔ CaO + H2O(g) 104/−104 [218]14043146.4350–900 q
2Co3O4 ↔ 6CoO + O2(g) 205/−205 [218]8642124700–850 r
2BaO2 ↔ 2BaO + O2(g) 77/−77 [218]4682361.6400–1025 s
Phase Change Materials (PCM) ΔHmeltHfusion (kJ/mol)kJ/kgkJ/LMelting Point (°C)
Al ↔ Al(l) 10.7/−10.7 [241]3971071.9660
LiF ↔ LiF(l) 27.0/−27.0 [250]10412747.2849
LiCl ↔ LiCl(l) 19.9/−19.9 [241]469.3969.6610
NaCl ↔ NaCl(l) 28.2/−28.2 [250]4821033.4801
Sensible Heat Materials (SHM) Specific Heat (kJ/kg·K)kJ/kgkJ/LTemperature Range (°C)
60 wt % NaNO3(l), 40 wt % KNO3(l) [212] 1.59 t436.3958.6 (802.8 u)290–565
60 wt % NaNO3(l), 40 wt % KNO3(l) [212] 1.59 t174.5383.4 (321.1 u)290–400
“n.a.” = data not available. a The lower temperature limit is based on the thermodynamics of the system. In practice, however, temperatures above ~100 °C are required for reasonable kinetics. b NaAlH4 is molten above 183 °C [243]. c Upper temperature potentially limited by melting point of NaH, ~638 °C [241], and KH, ~619 °C [241], decomposition products. There are some indications that NaH and KH may also form a eutectic melt at a temperature as low as 280 °C [246]. d There are some discrepancies in the reported thermodynamics for KAlH4. The enthalpy of formation of KAlH4, ∆Hf° = −175.4 kJ/mol, was taken as the average of the values reported in [241,251] while the entropy of KAlH4, S° = 120.9 J/mol·H2·K, was taken as the average of the values reported in [241,252]. e At ~574 °C, decomposition directly to liquid sodium metal rather than solid NaH becomes thermodynamically preferred according to the reaction KAlH4 + NaCl ↔ KCl + Na(l) + Al + 2H2(g). The H2 equilibrium pressure at this temperature is ~33.2 bar. f Limited to melting point of LiNH2 = 375 °C [253]. g Calculated up to the α-CaH2 to β-CaH2 phase transition at 780 °C [241]. At this temperature, the H2 equilibrium pressure for the CaNH + CaH2 ↔ Ca2NH + H2(g) system is ~10.0 bar. The H2 equilibrium pressure for the 6LiBH4(l) + CaH2 ↔ 6LiH(l) + CaB6 + 10H2(g) system is ~132.4 bar. h The melting point of LiH is ~688 °C. The enthalpy and entropy of melting are ∆Hm = 21.8 kJ/mol and ∆Sm = 22.6 J/mol.K, respectively [241]. i The melting point of NaBH4, 507.5 °C, is taken as the average of reference values [254,255]. This temperature corresponds to an H2 equilibrium pressure of ~0.26 bar [241]. The thermodynamics of molten NaBH4 have not been experimentally determined. j The melting point of KBH4, 615.5 °C, is taken as the average of reference [254,255]. This temperature corresponds to an H2 equilibrium pressure of ~0.23 bar [241]. The thermodynamics of molten KBH4 have not been experimentally determined. k Thermodynamic data source for: (1) LiBH4 is reference [240]; (2) Mg is reference [241]; (3) MgH2 is reference [256] and; (4) MgB2 is reference [257]. l At 467 °C, the H2 equilibrium pressure is ~57.8 bar for the reaction between 2LiBH4 and MgH2. Above this temperature the direct decomposition of MgH2 becomes thermodynamically preferred which is then followed by the reaction of metallic Mg with LiBH4(l). m Data is for solid NaH up to its melting point ~638 °C [207,241]. The H2 equilibrium pressure at this temperature is ~106.5 bar. n Practical maximum operating temperature limited to ~400 °C due Mg sintering and capacity loss above this temperature [258]. o Enthalpy and entropy for the reaction TiH1.6−TiH-0.3H2(g) calculated using data extracted for 527 °C ≤ T ≤ 707 °C from Figure 2 on p 241 of [207]. p The operating temperature range corresponds to a CO2(g) partial pressure of between 0 and 10 bar [218]. q The operating temperature range corresponds to a H2O(g) partial pressure of between 0 and 2 bar [218]. r The operating temperature range corresponds to a O2(g) partial pressure of between 0 and 1 bar [218]. s The operating temperature range corresponds to a O2(g) partial pressure of between 0 and 10 bar [218]. t Weighted average of the values for NaNO3(l) (1.83 kJ/kg·K), and KNO3(l) (1.22 kJ/kg·K), respectively [259]. u Due to the substantial volume change upon melting, the volumetric heat storage capacity calculated at 400 °C, ρ = 1.84 g/cm3 [260], has also been included for 60 wt % NaNO3/40 wt % KNO3.

Share and Cite

MDPI and ACS Style

Møller, K.T.; Sheppard, D.; Ravnsbæk, D.B.; Buckley, C.E.; Akiba, E.; Li, H.-W.; Jensen, T.R. Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage. Energies 2017, 10, 1645. https://doi.org/10.3390/en10101645

AMA Style

Møller KT, Sheppard D, Ravnsbæk DB, Buckley CE, Akiba E, Li H-W, Jensen TR. Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage. Energies. 2017; 10(10):1645. https://doi.org/10.3390/en10101645

Chicago/Turabian Style

Møller, Kasper T., Drew Sheppard, Dorthe B. Ravnsbæk, Craig E. Buckley, Etsuo Akiba, Hai-Wen Li, and Torben R. Jensen. 2017. "Complex Metal Hydrides for Hydrogen, Thermal and Electrochemical Energy Storage" Energies 10, no. 10: 1645. https://doi.org/10.3390/en10101645

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