Next Article in Journal
Preparation and Properties of Electrodeposited Ni-B-Graphene Oxide Composite Coatings
Previous Article in Journal
Influence of ZnF2 and WO3 on Radiation Attenuation Features of Oxyfluoride Tellurite WO3-ZnF2-TeO2 Glasses Using Phy-X/PSD Software
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Complex Metal Borohydrides: From Laboratory Oddities to Prime Candidates in Energy Storage Applications

by
Cezar Comanescu
1,2,3
1
National Institute of Materials Physics, 405A Atomiștilor St., 077125 Magurele, Romania
2
Inorganic Chemistry Department, Politehnica University of Bucharest, 1 Polizu St., 011061 Bucharest, Romania
3
Faculty of Physics, University of Bucharest, 405, Atomiștilor St., 077125 Magurele, Romania
Materials 2022, 15(6), 2286; https://doi.org/10.3390/ma15062286
Submission received: 31 January 2022 / Revised: 26 February 2022 / Accepted: 11 March 2022 / Published: 19 March 2022
(This article belongs to the Special Issue Novel Nanomaterials for Energy Storage)

Abstract

:
Despite being the lightest element in the periodic table, hydrogen poses many risks regarding its production, storage, and transport, but it is also the one element promising pollution-free energy for the planet, energy reliability, and sustainability. Development of such novel materials conveying a hydrogen source face stringent scrutiny from both a scientific and a safety point of view: they are required to have a high hydrogen wt.% storage capacity, must store hydrogen in a safe manner (i.e., by chemically binding it), and should exhibit controlled, and preferably rapid, absorption–desorption kinetics. Even the most advanced composites today face the difficult task of overcoming the harsh re-hydrogenation conditions (elevated temperature, high hydrogen pressure). Traditionally, the most utilized materials have been RMH (reactive metal hydrides) and complex metal borohydrides M(BH4)x (M: main group or transition metal; x: valence of M), often along with metal amides or various additives serving as catalysts (Pd2+, Ti4+ etc.). Through destabilization (kinetic or thermodynamic), M(BH4)x can effectively lower their dehydrogenation enthalpy, providing for a faster reaction occurring at a lower temperature onset. The present review summarizes the recent scientific results on various metal borohydrides, aiming to present the current state-of-the-art on such hydrogen storage materials, while trying to analyze the pros and cons of each material regarding its thermodynamic and kinetic behavior in hydrogenation studies.

Graphical Abstract

1. Introduction

A world with ever more scarce fossil fuels must transition to alternative energy storing materials. Fossil fuels, such as coal, gas, and oil, are finite, non-renewable resources, and a current estimation based on their ongoing burning rate points to a complete depletion by the end of 2060. While they are more reliable than currently engineered materials, they also pollute the environment and are responsible in great part for the greenhouse effect [1,2,3,4,5,6,7,8,9]. While the risk of running out of fossil fuels has been advocated for decades, the timeline is approaching and humanity needs to make a change. Therefore, alternative energy carriers are long sought-after; among them, metal hydrides and complex borohydrides have a special role [10,11,12,13,14,15,16,17,18]. Reaching or exceeding the U.S. Department of Energy (DOE) target is no easy task; the specific target for 2020 was 1.5 kWh/kg (4.5 wt.% hydrogen) and 1.0 kWh/L (0.03 Kg hydrogen/L), costing more than USD 10 /kWh (or USD 333/Kg of stored hydrogen capacity) [19,20]. A series of options have been investigated for hydrogen storage applications: cryo-compression (7–27 Kg/m3), metal hydrides (40–70 Kg/m3), chemical hydrides (50–120 Kg/m3), carbon sorption (20–50 Kg/m3), complex hydrides, and liquified dihydrogen (70 Kg/m3) (Figure 1).
A special role is held by complex metal hydrides and borohydrides in particular, given their high gravimetric energy content (Figure 2).
Computing the theoretical hydrogen storage of a complex metal borohydride yields the following:
wt .   %   hydrogen   in   M ( BH 4 ) x = 4   x   A M + 14.842   x × 100 = 400 A M x + 14.842   %
This relation would imply that, in order to reach a minimum storage capacity set by the DOE of 4.5 wt.%, the following relation should hold:
400 A M x + 14.842 4.5     A M x 74.406
One could argue that the DOE target could be met, looking at Equations (1) and (2), by any mono-valent metal having AM  74.406 g mol , or a divalent metal with AM    148.812   g mol , or a trivalent metal with AM  223.218   g mol a.s.o. Indeed, mTHRost known borohydrides fall into this category: they afford a theoretical hydrogen storage capacity superior to that imposed by the DOE (4.5 wt.%). Equation (2) also implies that the hydrogen storage capacity will be the highest for the most lightweight borohydrides, thus corresponding to low AM values: second, third, and fourth period metals are the borohydrides most used to date [21,22,23,24,25,26,27]. A summary of known borohydrides with specific hydrogen storage capacity is presented in Figure 2, which are arranged by metal valency and atomic weight of the metal, in decreasing order of their theoretical hydrogen storage capacity (wt.%). The maximum hydrogen storage capacity (denoted herein as: H, wt.%) is computed for a series of mono-, di-, tri-, and tetravalent metals; in some cases, only hypothetic formulae were used since the actual borohydride’s existence is unclear (AuBH4, Ni(BH4)2, Co(BH4)2, Hg(BH4)2, Sc(BH4)3, Nb(BH4)3, U(BH4)3, Ge(BH4)4, Sn(BH4)4), while for others that exist as adducts with various stabilizing solvents/ligands (like Cr(BH4)2), the capacity is reported with respect to the theoretical, anhydrous form [28,29].
Many of the formulated borohydrides from Figure 2 comply with the DOE’s theoretical minimum hydrogen storage capacity (4.5 wt.%), yet very few of them can actually be used in storage materials. However, there are quite a few aspects plaguing metal borohydrides: high dehydrogenation enthalpies, high dehydrogenation temperature onset, slow dehydrogenation/rehydrogenation kinetics, and side-reactions occurring during dehydrogenation leading to boron-loss and thus incomplete/impossible rehydrogenation, but also very high temperature and H2 pressures required to rehydrogenate them.
From a structural point of view, borohydrides M(BH4)x are ionic/covalent compounds formed by a metal center (Mx+) that binds accordingly to the negatively charged borohydride tetrahedra BH4. The lightest member of the borohydride family is LiBH4, depicted in Figure 3 [30,31,32,33,34,35,36,37,38]. Established by synchrotron XRD at RT, LiBH4 shows at normal conditions an orthorhombic symmetry (Pnma), with the BH4 tetrahedra aligned along two orthogonal directions. The [BH4] units are surrounded by four Li+ cations, and each Li+ by four [BH4].
More important for hydrogenation studies, the B-H bond is of a covalent nature (rB-H = 119 pm), thus considerable energy is required to break that bond (ΕB-H = 389 kJ/mol). It follows that the dehydrogenation enthalpy ΔHdehydrogenation has a large value, an aspect that must be overcome using thermodynamical considerations. First (IA) and second group (IIA) metal borohydrides are not only known, but also commercially available. Various reviews have tackled different areas regarding complex borohydrides and hydrogen, focusing either on current technologies [1,2,6] and conceptual design of storage materials [3,4,5,15], or on various forms of hydrogen carriers [7,8] utilized to eventually switch the aging and depleting conventional fuel to a “green” fuel [9]. Other reviews evaluate the role of nanostructured metal hydride species for hydrogen storage [10,17], and usually focus on the potential presented by light metal borohydrides [10,16,18]. In the latter, sodium borohydride is actually presented as the fuel for the future [18] and judging by the emerging reports dating from 2021–2022 one could argue that this is true. Synthetic strategies and structure characterization [21] have been complemented by in-depth studies on hydrolysis process of light borohydrides [22,24,25,27] or by investigation on the possible intermediates that dehydrogenation could entail [12]. Destabilization strategies employed to tame the rigid behavior of tetrahydroborates during hydrogenation cycles have been reviewed [30,31,32,33,34], and the field of ionic conductivity has been explored with aplomb especially after the discovery of LT(low temperature)–HT(high temperature) phase transition in lithium borohydride [36,37,38]. The current review aims to gather most areas of interest regarding the fascinating chemistry of complex borohydrides and bring them under one umbrella, considering the historical evolution of the field, important milestones, and current and emerging trends.

2. Hydrogen Storage Options: Physical vs. Chemical Storage

Hydrogen storage methods can basically be divided in two groups: physical (liquid H2, cryo-compressed and compressed gas, and physically adsorbed as in metal-organic frameworks of MOF-5 type), and chemical storage (organic liquid: BN-methyl cyclopentane; interstitial hydrides: LaNi5H6; complex hydrides: NaAlH4; or in ammonia–borane adducts: NH3·BH3) (Figure 4).

2.1. Physical Storage of Hydrogen

Hydrogen has an atomic weight of 1.0079 g/mol, and is thus the lightest element known. Under normal conditions (standard temperature and pressure), hydrogen is an odorless, nontoxic but combustible gas comprising two hydrogen atoms covalently bonded by a single σ-bond: H2. Hydrogen (1H, most abundant isotope called protium) has no neutrons, one proton in the nucleus and one electron in the outer shell, and a covalent radius of only 0.315 Å. The common and perhaps the most important physical properties of hydrogen related to its storage applications are the melting point (m.p. = −259.14 °C) and boiling point (b.p. = −252.87 °C), which, coupled with its low density (0.0898 g/L) and high diffusivity, make hydrogen storage a difficult task from an economic and technological point of view.
Hydrogen has three OS (oxidation states): −1 (hydrides), 0 (molecular dihydrogen), and +1 (typical OS for hydrogen, as found in most compounds, as for instance in hydrocarbons or the ubiquitous H2O); it is therefore a versatile element, as it can act as both an oxidizing and a reducing reagent. Its single valence electron makes hydrogen a very reactive element; it is therefore usually found in nature in a variety of inorganic and organic compounds.
Being a fuel and highly flammable (low ignition energy, low explosion limit in air at 4% vol.), hydrogen is regulated through a series of EU safety datasheets: S9 (containers must be kept in well-ventilated areas), S16 (no smoking allowed in its proximity, and ignition sources are forbidden nearby) and S33 (danger due to electrostatic discharge).
Combustion of hydrogen produces the highest gravimetric energy known (33.3–39.4 Wh/g), depending on the state in which water is produced—gaseous water yields the lower value (33.6 Wh/g), while combustion to liquid water produces 39.4 Wh/g:
H 2 ( g ) + 1 2 O 2 ( g )     H 2 O ( g )                                       Δ H 1 = 242 kJ mol ;   33.3 Wh g
H 2 ( g ) + 1 2 O 2 ( g )     H 2 O ( l )                                       Δ H 2 = 285 kJ mol ;   39.4 Wh g
Hydrogen can be stored either as a liquid (cryogenic temperatures, due to low b.p.) or as a gas (at high pressures, in the range 350–700 atm). Physically sorbed hydrogen is also known: adsorption at surface of solids (MOF-5) or absorption (within solids).

2.2. Chemical Storage of Hydrogen

Facing the obvious difficulties of cryogenic temperatures and/or very high pressures, the physical storage of hydrogen has inherent safety issues that have swayed attention of researchers towards chemical storage of hydrogen. Solid-state hydrogen storage can be achieved in metal hydrides, complex metal hydrides, ammonia–borane adducts, or organic liquids (BN-methylcyclopentane) (Figure 4).
Among these, complex hydrides have gained more popularity due to their salt-like nature, high hydrogen storage wt.%, and involvement in a variety of advantageous chemical systems (M(BH4)x-MHy, M(BH4)x-M(NH2)y, etc). The general formula of a complex hydride can be written as AxDyHz, with A being usually group IA and IIA elements, and D either boron or aluminum. Complex hydrides contain hydrogen covalently bonded to D, forming the complex anion [DyHz]r−. These compounds have been known for many years, but they were revisited by Bogdanovic and Schwickardy, who discovered that addition of 2 mol % β-TiCl3 or Ti(OBu)4 catalyst to NaAlH4 would bring down the kinetic barrier of dehydrogenation to technologically feasible levels for hydrogenation studies. The novel material had a tentative formula NaAlH3.9Ti0.02Cl0.06 and showed a thermogravimetric curve featuring a reduction in the dehydrogenation temperature of 80–85 °C with respect to undoped NaAlH4, while proving the reversible character of the prepared sample up to the 100th cycle. However, the dehydrogenation–rehydrogenation rate was still rather low, requiring high temperatures (>150 °C) and pressures (60–150-bar H2) (5) [39].
N a A l H 4   1 3   N a 3 A l H 6 + 2 3   A l + H 2   N a H + A l + 3 2   H 2
This pioneering work opened the door to further investigation of complex hydrides as new motifs for hydrogen storage materials; understanding the role of the catalyst and the exact nature of dehydrogenation and rehydrogenation reactions was the main aim. Other complex hydrides such as Al(BH4)3 or Mg2FeH6 have almost double volumetric hydrogen density compared to liquid hydrogen (150 kg/m3 vs. 70 kg/m3), which stimulated research in this intriguing field of hydrogen storage systems.

3. General Synthesis Strategies for Metal Borohydrides M(BH4)x

There are many known complex metal borohydrides (Figure 2 summarizes most of them). The first known was aluminum borohydride Al(BH4)3, synthesized in 1939 by Schlesinger et al. by mistake, followed by discovery of NaBH4 in 1943 [40]. Out of all known borohydrides, Be(BH4)2 remains the one providing the highest gravimetric hydrogen content (20.7 wt.%, Figure 2), but it is also very toxic, which limits its usability as hydrogen storage material. While using them as raw materials for hydrogen storage is impractical due to poor reversibility and very high (or low) decomposition temperature, the high amount of hydrogen bound to boron in complex anion [BH4] or [B2H7] (the latter is identified only in solution) makes them very interesting starting points for novel energy storage systems.
The general strategies used for metal borohydride synthesis are: solid-state reactions (mechano-chemical), wet chemistry (solvent-assisted), nanoconfined hydrides, and formation of adducts of borohydrides. The mechano-chemical synthesis route is perhaps the most-used and best-known method to obtain metal borohydrides; it is also the only one able to afford mixed-cation borohydrides (like LiZn2(BH4)5) [41]. The main drawback of ball-milling / the mechano-chemical route is the impurity of the resulting borohydride with metal halides, which are by-products of the metathesis reaction [42,43,44,45,46]. By contrast, wet-chemistry methods can yield phase-pure metal borohydrides, and they afford far easier separation procedures.

3.1. Solid-State—Mechanochemical Synthesis

Ball-milling has a series of advantages: it allows for good homogenization of the sample, reduces the size of the particle, and introduces defects which are highly reactive, inclusive in hydrogenation studies [47]. The mechano-chemical route is a highly energetic process carried out usually in a planetary ball mill [47,48,49]. The milling program typically consists in 1–5 min ball-milling followed by a 1–5 min break [50], such that the chemical activation observed is not due to local temperature increase, which is estimated to remain around 60 °C [51,52,53], but rather due to high pressure caused by collisions between ball, powder, and vial which produce material shear stress and deformations [54]. The mechano-chemical approach can increase the reactivity of materials by size-reduction of the particles, which increases their surface area and thereby provides better interaction between particles needed for a complete reaction to occur.
A double substitution reaction (metathesis) occurs when a metal halide (such as MgCl2, or more recently MgBr2) reacts with NaBH4, producing Mg(BH4)2 in its β-phase polymorph:
M g B r 2 + 2   N a B H 4   β M g ( B H 4 ) 2 + 2   N a B r
Using a bromide source rather than chloride yields the expected β-Mg(BH4)2 faster (6 h), when the reactants are introduced in a near-stoichiometric ratio (MgBr2:NaBH4 = 1:2 or 1:2.15). The usual workup procedure consists in Soxhlet extraction with Et2O, followed by a two-step evaporation under vacuum (150 °C, 24 h and 190 °C, 5 h) [55,56,57,58,59]. The observed reactivity enhancement mirrors the halide reactivity trend, which increases down the 17 group in the order Cl < Br < I.
The seemingly first candidate for hydrogen storage (highest gravimetric hydrogen content, 20.7 wt.%), Be(BH4)2, was first synthesized by a metathesis reaction, at 145 °C [60]:
B e C l 2 + 2   L i B H 4   B e ( B H 4 ) 2 + 2   L i C l
Schlesinger et al. made a great contribution to the development of novel metal borohydrides, after the discovery of Al(BH4)3 in 1939, and studied the synthesis of borohydrides by metathetical reactions using alkali metal borohydrides [60], alternative syntheses of NaBH4 and LiBH4 [61], synthesis of NaBH4 from NaH and borate esters [62], and synthesis of U(BH4)4 [63,64]. The metathetical reaction leading to Be(BH4)2 is not a proper mechano-chemical reaction since it is a solid-state reaction but performed by vigorous shaking of the reagents rather than in a planetary mill [60]. In fact, due to the requirement to continuously remove the volatile (and toxic) Be(BH4)2 from the system, the reaction cannot be performed in a closed system (like the vial of a planetary mill).
The metathesis is best carried out using LiBH4, since it showed considerably faster reactivity towards MClx than NaBH4, having also an added reaction drive due to the formation of stable alkali salt, LiCl [60]:
Y C l 3 + 3   L i B H 4   Y ( B H 4 ) 3 + 3   L i C l
Synthesis of Y(BH4)3 starts from yttrium(III) chloride and LiBH4, producing in high yield the expected borohydride of Y(III) and LiCl as by-product, in an all-solid metathesis reaction [65,66,67,68,69,70,71,72,73,74,75]. Yan et al. have also tried ball-milling the reagents according to Equation (8) with the aim of obtaining pure Y(BH4)3, then solvent metathesis of the grinded mixture; however, the obtained yttrium(III) borohydride was only ~85% pure (impurity: LiCl, 15 wt.%) [76]. Moreover, when the stoichiometric mixture YCl3: 3 LiBH4 was employed for ball-milling at RT or cryo-ball-milling, the reaction mixture showed incomplete conversion to Y(BH4)3, while YCl3: 4 LiBH4 yielded almost completely the expected borohydride [67,77,78,79,80]. Usage of LiBH4 was compulsory, as Y(BH4)3 was not obtained when replacing the borohydride source with NaBH4 [74].
Zr(BH4)4 can be prepared by a series of solid-state reactions, for instance between NaZrF5 + 2 Al(BH4)3, ZrCl4 + 2 Al(BH4)3, or ZrCl4 + 4 LiBH4 [77,78,79,80].
It has also been suggested that, in the case of mechano-chemical synthesis, the need to evacuate the volatile borohydride species (Be(BH4)2 or Al(BH4)3 being prime examples) advocates for an equilibrium reaction that is shifted towards products with removal of the volatile product, in line with Le Chatelier’s principle [81,82]. Al(BH4)3 has a theoretical hydrogen storage capacity of 16.8 wt.%, but is a liquid at RT (mp = −64 °C) due to weak intermolecular forces between its molecules; it is fairly unstable and begins to decompose at ~40 °C, which is too low for use in a hydrogen production mobile tank. Its crystal structure presents [BH4] units that are covalently bonded by two H-bridges to the Al3+ center. A ball-milling reaction of AlCl3 with NaBH4 in a stoichiometric molar ratio results in good yields of Al(BH4)3 [60].
A l C l 3 + 3   N a B H 4   A l ( B H 4 ) 3 + 3   N a C l
The solid-state Al(BH4)3 has two polymorphs (transition temperature −93 to −78 °C) [83,84,85,86,87,88]. Due to its liquid nature, toxicity, and moisture and air sensitivity, Al(BH4)3 rather serves as a starting point for synthesis of other, more stable borohydrides, and it is not currently the focus of research on novel energy storage materials because of these limitations.
Ball-milling is also the only way to produce mixed-cation or mixed-anion borohydrides, with reaction conditions (milling time, temperature, pressure) depending ultimately on the metal electronegativity and d-electron configuration in the case of transition metal systems. For instance, mixing ZnCl2 and KBH4 in a 1:1 ratio produces a chloro-borohydride of K+ and Zn2+. The complexity of the ball-milling process and its high dependence on reagent molar ratio is depicted below, showing a family of products all starting from ZnCl2 and a metal alkali borohydride source (Figure 5) [41].
Besides the formation of LiCl (or NaCl) as by-products, the metathetic route has another drawback: some amounts of this salt can dissolve in LiBH4, yielding a solid solution of approximate formula Li(BH4)1-pClp (p is an integer number, 0 < p < 1), making further purification even more difficult [89].
Synthesis of borohydrides can also be accomplished from starting elements as raw materials, but under extreme conditions caused by the very low reactivity of boron, as shown by Friedrichs et al. (700 °C, 150 bar) [90,91]:
L i + B + 2   H 2   L i B H 4
Other high-pressure reaction routes starting from solid metal hydride or metal boride have been investigated. For instance, reacting metal hydride MHx with MgB2 and hydrogen H2 leads to corresponding M(BH4)x and MgH2 [92,93].
M H x + x 2   M g B 2 + 2 x   H 2   M ( B H 4 ) x + x 2   M g H 2
  f o r   x = 1 ,   M = L i ,   N a ;   f o r   x = 2 ,   M = C a
Regeneration of Ca(BH4)2 during hydrogenation studies was shown to proceed at 350 °C and 150-bar H2, with a faster dehydrogenation/rehydrogenation kinetics when a Ti-catalyst was employed [94].
Ca(BH4)2 can be obtained in good yield (60%) by rehydrogenation of calcium hexaboride CaB6 and CaH2 under harsh conditions (690–700-bar H2, 400–460 °C) [95]:
2   C a H 2 + C a B 6 + 10   H 2   3   α C a ( B H 4 ) 2
Interesting to note in Equation (12) is the formation of α-Ca(BH4)2 at temperatures above 400 °C; typically, when synthesized at lower temperature, α-Ca(BH4)2 undergoes an irreversible phase transformation to β-Ca(BH4)2 when heated above 180 °C; this synthesis route thus has a main advantage in affording the α-phase at high temperatures (Equation (32)). The overall hydrogenation capacity of Ca(BH4)2 based on Equation (12) is 9.6 wt.%. Formation of CaB6 instead of the unreactive B or high boranes like CaB12H12 in the dehydrogenation process confers to the reaction a real potential for reversibility [96,97,98,99]. The formation of Ca(BH4)2 can also take place in modest yields (~20%) by ball-milling at RT for 24 h and under 140-bar H2, when using TiF3 as catalyst; in fact, many rehydrogenation reactions benefit from using Ti-based catalysts [100]. Ball-milling under Ar backpressure (1 atm) between CaCl2 and MBH4 (M = Li, Na) was reported to also produce Ca(BH4)2, albeit made impure with chloride by-products (Equation (13)) [101].
C a C l 2 + 2 M B H 4 A r ,   1   a t m α C a ( B H 4 ) 2 + 2   M C l   ; M = L i ,   N a
Reaction between sodium hydride NaH and boric oxide have produced sodium borohydride in good yield (60%) after ball-milling at 330–350 °C for 20–48 h [62]:
4   N a H + 2   B 2 O 3   N a B H 4 + 3   N a B O 2
In a similar fashion, but using methyl borate B(OCH3)3 instead of boric oxide, LiBH4 was obtained with an overall yield of about 70% [62]:
4   L i H + B ( O C H 3 ) 3   L i B H 4 + 3   L i O C H 3
However, the economical aspect of using mechano-chemical synthesis may tip the balance, because large-scale facilities for grinding are known and operable, so that cost-effectiveness is insured for the preparation of light complex hydrides via the ball-milling technique.

3.2. Wet Chemistry—Solvent-Assisted Synthesis

The first borohydride, Al(BH4)3, was synthesized by Schlesinger by accident in 1939 [40] (Equation (16)). He was hoping to obtain AlH3 by mixing Al(CH3)3 and B2H6, when the following reaction actually occurred, which has further been expanded to the synthesis of Be(BH4)2 (Equation (17)) and LiBH4 (using an alkyl-lithium in the latter case). LiBH4 is the second most used borohydride today, but it was first synthesized also in the 1940s by Schlesinger et al. according to Equation (18) [102].
A l ( C H 3 ) 3 + 2   B 2 H 6   A l ( B H 4 ) 3 + 3   ( C H 3 ) 3 B
3   B e ( C H 3 ) 2 + 4   B 2 H 6   3   B e ( B H 4 ) 2 + 2   ( C H 3 ) 3 B
3   L i C 2 H 5 + 2   B 2 H 6   3   L i B H 4 + ( C 2 H 5 ) 3 B
At present, the usual procedure for the synthesis of M(BH4)x follows the general reactivity scheme (19) and (20), and it can be successfully applied for divalent borohydride synthesis (M = Be, Mg, Ca, Sr, Ba) or trivalent borohydrides (M = Al):
L i X + N a B H 4   L i B H 4 +   N a X                 ( X = C l ,   B r ,   I )
M C l x + x   L i B H 4 M ( B H 4 ) x + x   L i C l
Al(BH4)3 was obtained by Kollonitsch and Fuchs and reported back in 1955, by reacting AlCl3 with Ca(BH4)2 in a 2:3 molar ratio [103].
2 A l C l 3 + 3 C a ( B H 4 ) 2 2 A l ( B H 4 ) 3 + 3 C u C l 2
Synthesis of LiBH4 according to Equation (19) takes place in NaH-dried isopropylamine iPrNH2 [60,104], and usage of toxic diborane B2H6 is now avoided as in its first synthesis by Schlesinger and Brown [102]. The metathesis reactions (19) and (20) are well-documented, and proceed with a reasonable rate for LiBH4 compared to slower alkali borohydride starting materials (i.e., NaBH4, and ever slower with KBH4) [60], and afford the complex borohydride M(BH4)x in near-quantitative yield. NaBH4 remains the most wide-spread borohydride in terms of synthesis methods and uses, including but not limited to organic synthesis, reduction reactions, and metathesis [105].
Mg(BH4)2 could be synthesized according to Equation (20), using diethyl ether Et2O as solvent for the LiBH4, but the product obtained contains impurities of Li and Cl [106,107,108]. At modest temperatures (34–60 °C), reaction (22) yields some Mg(BH4)2 (overall yield: 0–11.5%); different conditions must therefore be used if NaBH4 is used as borohydride source [107,108]. Even when conducted in recommended solvent iPrNH2, the final mixture could not surpass 40 wt.% of Mg(BH4)2, with the impurities being mainly represented by unreacted NaBH4; this aspect could be proof of an equilibrium reaction being reached under those specific reaction conditions, but it could also originate from poor solubility of the intermediate magnesium amine complex formed in the reaction mixture [106]. Phase-pure β-Mg(BH4)2 was obtained by replacing highly reactive LiBH4 with NaBH4, using a combined mechano-chemical (2 h) double exchange reaction (60 h, refluxing Et2O solvent), followed by drying at 240 °C [109].
M g C l 2 + 2   N a B H 4   M g ( B H 4 ) 2 + 2   N a C l
Organometallic compounds can also be used to synthesize metal borohydrides; reaction of n-butyl-magnesium with Al(BH4)3 in non-coordinating solvent (toluene) yields a solvent-free Mg(BH4)2 in very good yield (85%) (Equation (23)) [110].
3   M g ( C 4 H 9 ) 2 + 2   A l   ( B H 4 ) 3   3   M g ( B H 4 ) 2 + 2   A l   ( C 4 H 9 ) 3
Reaction of CaH2 or Ca(OC2H5)2 with diborane B2H6 in THF also affords the adduct of Ca(BH4)2 in good yield [111,112,113,114,115,116,117]. The reaction of ball-milled activated CaH2 (1 h, under Ar, 600 rpm) and triethylamine borazane complex Et3N·BH3 produces Ca(BH4)2 in 85% yield (unreacted CaH2 as main impurity). The reaction is carried out at 140 °C for 5–6 h and the subsequent workup consists in n-hexane washing, filtering, and drying at 200 °C for 12 h [118].
Zn(BH4)2 could also be prepared by wet-chemistry at room temperature (RT), using tetrahydrofuran (THF) as solvent and ZnCl2 and NaBH4 as starting materials (Figure 5) [119,120,121,122]. With a hydrogen storage capacity of 8.5 wt.%, Zn(BH4)2 starts to decompose at ~85 °C and eliminates up to 140 °C hydrogen and diborane B2H6, making it for the time being too unstable for vehicular applications. On the other hand, formation of toxic B2H6(g) should urge researchers to look for alternative decomposition pathways that avoid diborane as by-product, since current dehydrogenation will not be reversible in the absence of the boron source [119,120,121,122].
Potassium borohydride KBH4 (also RbBH4 and CsBH4) could be synthesized by a precipitation reaction of NaBH4 with a concentrated solution of a strong base, KOH; the reaction could be carried out in aqueous or CH3OH media, with an overall yield of 75.5% (Equation (24)). The reagents are independently dissolved in MeOH, cooled, and then mixed, when the borohydride of interest precipitates and is obtained in pure form after a subsequent vacuum drying step [123,124].
N a B H 4 + M O H   M B H 4 + N a O H           ( M = K ,   R b ,   C s )
LiBH4 can also be produced by reacting LiH with diborane at mild temperatures (120 °C) [125]. However, no reaction takes place between LiH and CaB2; Ca(BH4)2 therefore cannot be obtained via this route (Equation (25)) [92].
2   L i H + B 2 H 6     2   L i B H 4
A similar reaction between the metal hydride MHx (M = Li, Mg, Ca) and B2H6 in a solid–gas mechanochemical process was proposed by Friedrichs et al., who obtained borohydride products in good-to-high yield, but containing some impurities (unreacted starting metal hydride and Fe-Ni phases from vial abrasion during milling): LiBH4 (η = 94%), Mg(BH4)2 (η = 91%) and Ca(BH4)2 (η = 73%). Diborane B2H6 was also proposed as a key intermediate in hydrogenation studies under moderate conditions [93].
Neutron diffraction studies have been facilitated by a hydrogen-deuterium H-D exchange reaction taking place at increasing temperatures for Li, Na, and K (200 °C, 350 °C, and 500 °C, respectively) (Equation (26)) [126].
  M B H 4 + 2 D 2       M B D 4 + 2 H 2                       ( M = L i ,   N a ,   K )

3.3. Nanoconfined Hydrides

The ball-milling technique is usually the go-to process for obtaining nanosized powders of metal hydrides, to sizes often less than 100 nm [127]. Ball-milling implies milling in a vial using balls of special alloys; however, the milling process could induce impurities in the sample from the materials of the vial and balls. However, these nanoparticles could also sinter into larger particles during hydrogenation–dehydrogenation cycles, which negatively impacts gravimetric hydrogen uptake and release [128,129,130,131]. Using specially engineered nano-scaffolds, complex hydrides can be obtained in the nanometer range, far below the average sizes afforded by the mechano-chemical approach.
Among nanoconfinement methods, the most widely used are the direct synthesis of nanoconfined complex metal (boro)hydride, and infiltration methods (melt-infiltration and wet-infiltration) [132].
Nanoconfinement of metal hydrides synthesized in situ has been extensively studied for magnesium hydrides in carbon aerogels produced by the resorcinol–formaldehyde reaction (RF-CA), using a 1 M solution of Mg(nC4H9)2 in n-heptane [133,134]. Nielsen et al. synthesized by this method MgH2@RF-CA of pore sizes 7 and 22 nm [133]. This is a typical case of soft-nanocasting, as the initial porous support is not removed from the final composite material (Figure 6).
The excess n-butyl magnesium can be removed manually after the reaction from the surface of RF-CA (Equation (27)).
M g ( C 4 H 9 ) 2 + 2   H 2   Δ   M g H 2 + 2   C 4 H 10
Melt infiltration requires a very inert scaffold, with a melting point higher than that of the respective metal hydride used. The driving force of this method is the lower interfacial energy metal hydride/scaffold, as reviewed by de Gennes [135,136]. This method has the advantage of completely removing the necessity of a solvent, therefore avoiding post-synthesis steps of the hydride@scaffold composite. In addition, the melting point of the hydride must not coincide with the onset of decomposition of the said material. However, sodium tetrahydroaluminate NaAlH4 (mp = 183 °C) has been utilized for melt infiltration, despite decomposition onset into Al and Na3AlH6 upon melting [137].
The decomposition of NaAlH4 must be considered a partial equilibrium (Equations (28) and (29)), because when working around the borohydride melting point and using high H2 pressure (160–190 bar), the decomposition is avoided and melt infiltration has been successful [137,138,139,140,141,142].
3   N a A l H 4     Δ   N a 3 A l H 6 + 2   A l + 3   H 2
  N a 3 A l H 6     Δ 3   N a H + A l + 3 2   H 2
Solvent infiltration requires a suitable solvent that affords complete metal hydride dissolution in order to be successful. A series of physical properties of metal borohydrides (including knowledge of select solubility data) are a prerequisite, as limited solubility in commonly used solvents is the main limitation of this method. Sometimes even weakly coordinating solvents could be used, but these solvents should be removed under moderate conditions (T, p). Following the rather poor solubility of complex metal hydrides, consequent steps of impregnation–evaporation may be required to reach full (or the desired degree of) scaffold pore filling [132]. This technique—incipient wetness method—can be advantageous as it avoids crystallization of active hydride species outside of mesopores; complex metal borohydrides like Ca(BH4)2 have been infiltrated in chemically activated micro-mesoporous carbons of very high surface area to produce composites exhibiting reversible hydrogen storage under more modest conditions [143].
Various scaffolds can be used; inert, mesoporous silica of 2D-ordered type (SBA-15) could serve such a purpose. Computational studies have shown that molecules of NHxBHx type (x = 1–4) display thermodynamically accessible dehydrogenation steps of less than 40 kJ/mol, and relatively low dehydrogenation temperature onset (100–150 °C); they can thus be considered candidates for hydrogen storage reactions [144]. When a 5.4 M solution of ammonia–borane NH3.BH3 in CH3OH infiltrated the ~7.5 nm pores of SBA-15 silica, a total loading of roughly 50 wt.% NH3BH3@SBA-15 was obtained after solvent evaporation [145]. A clear advantage of solvent-mediated impregnation is that the experimental procedure requires less energy, since it can be carried out even at room temperature.

3.4. Derivatives—Formation of Adducts of M(BH4)x

Solvent metathesis reactions often lead to the formation of desired borohydrides as solvates: M(BH4)x·p Et2O (p is an integer number), as is the actual case for M = Mg and Y. Both Mg(BH4)2·p Et2O [106,107,108] and Y(BH4)3·p Et2O [76] could be obtained as pure, unsolvated solids after heating at 150 °C, as shown by IR and TG data (Equation (30)).
M ( B H 4 ) x · p   E t 2 O   150   ° C M ( B H 4 ) x + p   E t 2 O             ( M = M g ,   Y )
Other borohydrides are obtained as tetrahydrofuran adducts, such as Ca(BH4)2·p THF, which is also commercially available. Heating this adduct at 160 °C for 1 h leads to phase-pure, α-Ca(BH4)2 (Equation (31)) [146,147,148,149,150,151]. Careful control of the temperature profile is advised, as from 180 °C the conversion to β-Ca(BH4)2 occurs (Equation (32)).
C a ( B H 4 ) 2 · p   T H F   160   ° C ,   1 h α C a ( B H 4 ) 2 + p   T H F
α C a ( B H 4 ) 2   180 200   ° C     β C a ( B H 4 ) 2
As is typical with organometallic synthesis, purification of metal borohydrides can be achieved using recrystallization from a concentrated solution (usually the synthesis solvent: iPrNH2, MeOH or even H2O). NaBH4 could be recrystallized from iPrNH2 [152] or H2O [153]. In the latter case, single crystals could be obtained corresponding to dihydrate sodium borohydride, NaBH4·2H2O.
Triethylamine (Et3N)-solvated metal borohydrides: NaBH4 or partially deuterated NaBD3H [154,155], Mg(BH4)2 [156], and Ca(BH4)2 [157], could be obtained by reacting the corresponding premilled (and thus highly reactive) metal hydride with an excess of aminoborane (Equations (33) and (34)).
N a H + E t 3 N · B D 3         N a B D 3 H + E t 3 N  
M H 2 + 2   E t 3 N · B H 3         M ( B H 4 ) 2   ·   2   E t 3 N   Δ     α M ( B H 4 ) 2 + 2   E t 3 N     ( M = M g ,   C a )
Dimethyl sulfide S(CH3)2-stabilized magnesium borohydride can also be obtained, provided that a large excess of the tioborane is used. Solvent-free Mg(BH4)2 can be obtained by working up the reaction mixture at low pressures (10−1–10−5 mbar, for 19 h in total), a step which removes coordinating solvent S(CH3)2 (Equations (35) and (36)) [110].
3   M g ( C 4 H 9 ) 2 + 8   B H 3 · S ( C H 3 ) 2   3   M g ( B H 4 ) 2 · 2   S ( C H 3 ) 2 + 2   B ( C 4 H 9 ) 3 · S ( C H 3 ) 2
M g ( B H 4 ) 2 · 2   S ( C H 3 ) 2 Δ     α M ( B H 4 ) 2 + 2   S ( C H 3 ) 2
Desolvation from a dimethylsulfide borane complex solvate led instead to a new, nanoporous polymorph, γ-Mg(BH4)2 (Equation (37)) [158].
S ( C H 3 ) 2 · B H 3   M g ( C 4 H 9 ) 2   M g ( B H 4 ) 2 · 1 2 S ( C H 3 ) 2   v a c u u m ;   R T 80   ° C     γ M g ( B H 4 ) 2
The structural variety of polymorphs obtained in the case of magnesium borohydride (six in total) (Equations (35)–(37)) suggests that novel, nanoporous metal hydrides could be potentially accessible via this desolvation route.
Thermal analysis coupled with IR spectroscopy data has pointed out two possible formulations of solvated complexes of closely related tetrahydroaluminates, when coordinated to a tertiary amine “L” (Figure 7). If a weak coordination is expected, then structural formula I is favorable, and it should resemble that of an uncoordinated complex metal hydride, as pointed out by Schlesinger and Brown. The complex should exhibit high thermal stability—which is actually the case for I: L → M+ [AlH4], due to the increased size of the cation and the reduced distortion of the AlH4 tetrahedron by alkali metal interaction [102]. By contrast, formula II: H-M…AlH3L is expected to have a lower thermal stability, resulting in a complex mixture of products upon decomposition due to the complex bond amine-alanate R3N: →AlH3 [159].
Full decomposition of solvate LiBH4·L could proceed through either pathway (38) or (39) as evidenced by the ambiguous DTA-TGA curve, with both leading to loss of coordinating molecule L:
L i A l H 4 · L         L i H + A l + 3 2 H 2 + L
3   L i A l H 4 · L           L i 3 A l H 6 + 2   A l + 6   H 2 + 3   L
Mg(BH4)2·THF adducts decompose violently [160] due to the formation of alkoxy species by the THF ring-opening reaction, as is the case with aluminum hydride-THF adducts [161]. However, Ca(BH4)2·THF can be desolvated with relative ease, without such strong exothermic effects, as evidenced by Equation (31) [146,147,148,149,150,151,162].
A metathesis reaction is currently the preferred method of synthesis for metal borohydrides because they can be obtained in pure form. There are recent reports expanding the recognized reducing character of metal borohydrides. For instance, Ca(BH4)2 can be successfully used in advanced organic synthesis, such as polymerization of ε-caprolactine and L-lactide [163].

4. Structural Considerations of M(BH4)x

There is an intricate connection between structure and solid-state chemistry related to borohydride materials. Among most important physical properties affected by lattice structure are the thermal stability and the accessible decomposition pathways.

4.1. Framework and Crystal Structure

Metal borohydrides are typically ionic compounds, but covalent borohydrides are also known. Table 1 summarizes some of the novel metal borohydrides and mixed-cation borohydrides and their structural architecture; many present multiple polymorphism (LT, HT, metastable).
The structural diversity of metal borohydrides is stunning, and a glimpse of such diversity is provided in Figure 8, with a few representative examples. For instance, CsSr(BH4)3 exhibits a structure similar to that of perovskite, which on the one hand explains its high stability with a decomposition onset at 398 °C [185]. Potassium octahydridotriborate β-KB3H8, on the other hand, exhibits discrete units implying a high mobility of boron tetrahedra during chemical transformations [186].
Perhaps one of the most studied metal borohydrides of the main groups, LiBH4, (1 in Figure 8) crystallizes in the orthorhombic Pnma space group, with cell dimensions a = 7.1785 Å, b = 4.4368 Å, c = 6.8032 Å, and V = 216.685(3) Å3 at 20 °C. While early diffraction data of lower resolution suggesting strong distortion of the BH4 tetrahedra, more recent determinations show that B-H bonds and H-B-H angles practically describe perfect tetrahedrons [30,31,32,33,34,35,36,37,38]. Each Li+ cation coordinates four [BH4] tetrahedra, and each [BH4] tetrahedron coordinates four Li+ cations. The LT polymorph of LiBH4 shows the Li+ cation in four-coordinate tetrahedral sites between anionic layers of [BH4] that are hexagonally stacked. When heated to ~110 °C, the LT-LiBH4 (o-LiBH4) undergoes a phase transition to a hexagonal polymorph, HT-LiBH4 (h-LiBH4), which crystallizes in hexagonal space group P63mc resembling the wurtzite (ZnS) structure, with the following unit cell parameters: a = 4.2763 Å, c = 6.9484 Å, and V = 110.041(4) Å3 at 135 °C (2—in Figure 8). The o-LiBH4 to h-LiBH4 polymorphic transformation taking place at ~110 °C can be seen as a contraction along orthorhombic axis a (hexagonal axis c) and an expansion of the [b*c] orthorhombic plane, and it is accompanied by a slight endothermic enthalpy of transition (4.18 J/mol).
KBH4 (3 in Figure 8) in its β-polymorph, crystallizes in the tetragonal P42/nmc space group. The 3D structure shows a 12-coordinate geometry around the alkali metal, in an equivalent H-surrounding. The K-H bond lengths average 2.815 Å, while the B-H bonds in the [BH4] unit measure 1.23 Å. The structure of β-KBH4 bears structural similarities to tetragonal NaBH4, but it is refined to a higher symmetry group P42/nmc. As the size of the cation increases in the alkali group (from Li to Rb), the cubic MBH4 (M = alkali metal) unit cell expands and the [BH4]…[BH4] repulsions are reduced, such that the cubic-to-tetragonal transition temperature is reduced in heavier alkali metal borohydrides.
RbBH4 (4 in Figure 8) crystallizes in the cubic Fm 3 ¯ m space group (a = 6.9633 Å at 0.5 GPa). The Rb+ cations and BH4 anions have a NaCl structural prototype and this room-temperature polymorph is isostructural to NaBH4, with the BH4 groups orientationally disordered over the two positions around the inversion center, similar to the sodium tetrahydroborate case.
Be(BH4)2 (5 in Figure 8) presents only one structurally characterized form which crystallizes in the tetragonal I41/cd space group. It contains helical polymeric chains with the independent Be2+ coordinated in a trigonal-planar geometry to two bridging [BH4] tetrahedra via face coordination (Be-B bond distance 2.00 Å) and one terminal [BH4] tetrahedron via the tetrahedral edge (Be-B bond distance 1.92 A). This originates presumably from the repulsive [BH4]…[BH4] forces due to very short H…H distances (ranging from 2.24 Å to 2.31 Å). The bridging [BH4] group shows a linear coordination to the two Be2+ cations (Be-B-Be angle close to 180°), leading to a 1D polymeric structure.
α-Ca(BH4)2 (7 in Figure 8) can be obtained by desolvation of the THF adduct, and crystallizes in the orthorhombic Fddd space group, being one of the four possible polymorphs (α, α’, β and γ) and bearing structural similarities to TiO2 polymorphs [187,188]. The orthorhombic structure has been further amended to better be represented by a lower symmetry (F2dd or Fdd2). All four polymorphs show Ca2+ octahedrally coordinated to six equivalent [BH4] tetrahedra, with the Ca-B bond distances falling in the range 2.82–2.97 Å. The divalent calcium cations form in α-Ca(BH4)2 a close-packed, diamond-type structure where the tetrahydroborate groups exhibit T-shape coordination [188].
Three polymorphs of Mg(BH4)2 are depicted (9: α-, 10: β-, 11: γ-), showing increased structural complexity, but also polymeric properties [106,158,171,172,173,189,190,191,192,193,194,195]. The α-Mg(BH4)2 represents the most stable polymorph and its structure determination was a result of combined synchrotron (most information data) and neutron powder diffraction data (orientation of BH4 tetrahedra). Single crystal diffraction data collected at 100 K established the space group to be the hexagonal P6122, containing three symmetry-independent Mg2+ ions and six independent [BH4] units. In this novel 3D arrangement, the Mg-Mg distances lie in the range 4.6–4.9 Å, while the Mg2+ center is tetrahedrally coordinating four [BH4] tetrahedra. The Mg-B bond lengths are in the range 2.432(a)–2.434(4) Å, and there is a dispersity of B-Mg-B angles of 92.1(2)–108.80(13)° accounting for a distorted tetrahedral coordination around the metal center. The porous γ-Mg(BH4)2 has a cubic lattice with inherited porous nature (33% open space), high surface area (in excess of 1100 m2/g as deduced from N2 sorption isotherms by BET method) and a partially covalent framework. The stability of Mg(BH4)2 can be related to the Mg-BH4-Mg linear units, with BH4 acting as directional ligands, and stable MgH8 polyhedra. This complex structural behavior resembles that of coordination polymers like MOFs. The γ-polymorph is particularly interesting due to large empty spaces available for gas entrapment [158].
One of the main dehydrogenation borohydride species found during hydrogen desorption studies is the [B12H12]2− anion, as found for instance in BaB12H12 (12 in Figure 8). These dodecahydro-closo-dodecaborate anions [B12H12]2− have a role in the poor recyclability of metal tetrahydroborates from hydrogenation studies; their formation and possible side-reactions are therefore of clear importance. Moreover, it has been hypothesized that these species appear amorphous in XRD diffraction data, and only show short-range ordering. BaB12H12 crystallizes in the P31c space group (trigonal symmetry) of wurtzite structural type, and shows Ba2+ cation surrounded by four [B12H12]2− anions. Each closo-borate [B12H12]2− anion is tetrahedrally surrounded by four Ba2+ cations (mutual tetrahedral arrangement of cations and anions), with distances from the center of the anion to the Ba2+ of 4.40 and 4.54 Å at 295 K. Since each of the four [B12H12]2− anions is facing the Ba2+ metal center with three H atoms, the cation is surrounded overall by twelve H atoms (Ba–H bond distances range from 2.77–3.02 Å). The rich boron–hydrogen speciation, open channels, and convenient molecular size (5.8 Å) have made closo-borate salts very interesting candidates for ionic conductivity studies.
Al(BH4)3 (1 in Figure 9) is a liquid under normal conditions, but low temperature single-crystal XRD data shows two possible polymorphs with a transition temperature at ~190 K [83]. Single-crystal diffraction data by Aldridge et al. from 1997 shows that Al(BH4)3 crystallizes at 195 K in the orthorhombic Pna21 space group with the following unit cell parameters: a = 18.021(3) Å, b = 6.138(2) Å, and c = 6.199(1) Å. It consists of discrete Al(BH4)3 units, with a trigonal-planar geometry around the Al3+ center which coordinates the [BH4] tetrahedra via the tetrahedral edges. The Al-B bond lengths are in the range of 2.10(2)–2.14(2) Å, while the longer B-H distances for the B-H…Al bridges of 1.12(3)–1.14(4) Å are consistent with Al-H coordination [83].
α-Y(BH4)3, one of the most promising hydrogen storage candidates among transition metal borohydrides (2 in Figure 9) was synthesized by cryo-milling of LiBH4 and YCl3. It crystallizes in the cubic Pa 3 ¯ space group with a lattice constant a = 10.8522(7)Å [65,68,69,70,71,72,73]. The structure has a Y3+ center in a highly distorted octahedral geometry being surrounded by six [BD4] tetrahedra (B-Y-B angles in the range 78.6(2)–160.9(1)°). The α-Y(BH4)3 converts at 10 MPa D2 and 475 K to the fcc β-Y(BH4)3 polymorph in the Fm 3 ¯ c space group.
α-Mn(BH4)2 (3 in Figure 9) crystallizes in the P3112 space group and is isostructural with ζ-Mg(BH4)2, having the unit cell parameters a = 10.435(1)Å and c = 10.835(2)Å [189,190,191,192,193,194,195]. This represents the first crystal structure of a 3d-metal tetrahydroborate. The α-Mn(BH4)2 polymorph has a stability range between 90 to 450 K. The structure bears structural similarities to Mg(BH4)2; the two independent Mn2+ centers are each tetrahedrally surrounded by four [BH4] units with Mn-B distances of 2.39–2.52 Å.
LiSc(BH4)4 can best be described as a Li+ salt of complex anion [Sc(BH4)4] that was evidenced by vibrational spectroscopy studies. It crystallizes in the tetragonal P 4 ¯ 2c space group, with a CuAlCl4 structural prototype and the following unit cell parameters: a = 6.076 Å and c = 12.034 Å (5 in Figure 9). More specifically, it was obtained by Hageman et al. in 2008 by ball-milling processing of a 4:1 molar ratio LiBH4 and ScCl3 [181]. The central cation Sc3+ of the complex anion [Sc(BH4)4] is surrounded by four BH4 tetrahedra in a distorted tetrahedral coordination, with a Sc-B bond distance of 2.28(1)Å, which compares favorably to the DFT value predictions of 2.33 Å. The B-Sc-B angle has a spread distribution between 96.5(5)° and 124.4(5)°, suggestive of the deformed tetrahedral coordination. Each of the four [BH4] tetrahedra has three H atoms oriented towards the Sc center, so that the global coordination is 8 + 4: 8 Sc-H distances in the range 2.11–2.15 A and 4 Sc-H distances of 2.31 A, while the remaining B-H bond pointing outwards is responsible for the high B-H stretching frequencies of 2485–2498 cm−1. Moreover, the crystal structure contains disordered Li+ cations along the z-axis of the tetragonal cell, which in turn could make this mixed borohydride a suitable candidate for ionic conductivity applications [181].
NaZn2(BH4)5 (7 in Figure 9) crystallizes in the monoclinic P21/c space group, similar to the Li+ counterpart LiZn2(BH4)5, and it has no analogues to known inorganic structures. The unit cell parameters are: a = 9.397(2) A, b = 16.635(3)A, c = 9.136(2)A, and β = 112.66(2)° [41]. The structure consists of two independent Zn2+ cations that have almost a trigonal-planar coordination to three [BH4] tetrahedra (coordination number of Zn is 3), while the Na+ cation has a new, saddle-like coordination (coordination number of Na is 4) [41]. Interestingly, NaZn2(BH4)5 consists of two doubly penetrating 3D frameworks, with no covalent bonds between them—a structural motif found in coordination polymers. All [BH4] units are nearly linearly coordinated by two metal atoms, and the coordination occurs via the two opposite tetrahedral edges bridging two Zn2+ cations, or a Zn2+ and a Na+ cation.
The hexagonal structure observed for many metal borohydrides can be attributed in part to an entropic term brought about by the vibrational amplitudes in BH4 units and the shortening of B-H bonds from 1.22 to under 1.1 Å [196,197]. Mixed-cation borohydrides give rise to a fascinating family of crystal structures, and have many times exhibited promising results regarding hydrogen release and reversible hydrogen uptake.

4.2. Stability of MBH

The most ionic and stable borohydrides have been studied since the 1940s, namely NaBH4, KBH4, RbBH4, and CsBH4; they are often stable in basic aqueous solution [165,166,167,168,169,198]. Closo-boranes are stable under neutral and acidic solutions. Being typically air- and moisture-sensitive compounds, their synthesis should be performed using Schlenk techniques or in the glove box, under inert (N2, Ar) atmosphere. Metal borohydrides are prone to adsorbing water, with which they react upon heating to release hydrogen. Some borohydrides of low electronegative elements are stable at RT, and even more so in the form of their hydrates. High electronegative metals forming metal borohydrides are highly reactive even below RT, and produce a highly exothermic reaction with water (some even explode upon contact with moisture). The layered structure of LiBH4 is non-trivial and shows local distortion of the polymorphs, which defines their stability and plays a role in corresponding phase transitions [199]. Theoretical computations based on first-principles study shed new light on the stability of the family of possible borohydride intermediates from LiBH4, namely Li2BnHn (n = 5–12) [200].
Transition-metal borohydrides prepared by mechano-chemical methods are only stable when the ion itself is stable (d0, d5 and d10), that is, when it has no electrons in the d-orbital, at half-occupation, or when the d-shell is completely occupied [201]. However, lowering synthesis temperature to below −30 °C and using ammonia as stabilizing ligand, stable M(BH4)2(NH3)6 have been prepared in the case of M = Cr2+ (d4), Fe2+ (d6), and Co2+(d7). These borohydrides are only stable in solution at low temperature, and they have not been isolated in pure, solid form, with a d1 exception to the rule coming from a molecular borohydride of Ti(BH4)3, which has limited stability below 0°C [202]. Some borohydrides like Zr(BH4)4 (sublimes at 29 °C) form quality single-crystals by CVD when stored in a closed vessel, due to its molecular structure [203,204,205,206,207,208,209,210].
Bi- or tri-metallic borohydrides can usually be prepared by a mechano-chemical route; however, some diethyl ether Et2O solvates have been prepared and are stable in solvated form, such as NaMn(BH4)3·0.5 Et2O [211]. Dimethyl sulfide S(CH3)2 has recently been used as a weak-coordinating solvent which eases borohydride removal from the products, avoiding formation of ternary chlorides/bimetallic borohydrides. This strategy allowed for the synthesis and stabilization of some rare earth metal borohydrides M(BH4)3·S(CH3)2, where M is Y, La, Ce, or Gd [72,212,213,214,215]. Formation of ternary chlorides leads to incomplete reaction in mechanochemical synthesis; for instance, if Zn(BH4)2 is targeted, a mixture of NaBH4 and ZnCl2 used as reagents can often lead to the formation of Na2ZnCl4 by-product, which blocks Zn2+ cation availability for the metathesis reaction (Equation (40)).
Z n C l 2 + 2   N a C l   N a 2 Z n C l 4
Other approaches utilize bulky-cation-stabilized precursors like [nBu4N][Y(BH4)4] to produce less-stable borohydrides like LiZn2(BH4)5 [216,217].

4.3. Multi-Cationic Borohydrides

The first reported mixed alkali metal borohydride, LiK(BH4)2, was observed to form when a mixture of LiBH4 and KBH4 was mixed and heated to 125 °C [218]. This intermediate compound was fully characterized, including by crystal structure determination: Li+ coordinates to four [BH4] units, while K+ coordinates seven [BH4] tetrahedral unit. The decomposition temperature of LiK(BH4)2 is almost half of that of individual parent alkali-metal borohydrides.
Other mixed cation borohydrides are metastable, such as NaK(BH4)2; upon cooling to RT, the mixed metal borohydride separates into the monomeric borohydrides within one day [219]. The certainty of its formation by heating the system NaBH4–KBH4 was confirmed by an increase in the lattice parameter a, and the higher potassium content [220]. However, it was later proved that the mixed borohydride NaK(BH4)2 is not stoichiometric, but a solid solution of approximate formula NapK1-pBH4 (0 < p < 1), and only stable at elevated temperatures (200–450 °C) [221].
Recently, a series of bimetallic borohydrides of alkali metal with alkaline earth, d- or f-block metals has been reported, containing the more electropositive alkali metal with a countercharge brought about by the complex anion [21,22,23,24,25,26,27,222,223]. K2Mg(BH4)4 and K3Mg(BH4)5 have isolated [Mg(BH4)4]2− tetrahedra serving as complex anions, while the metal alkali K+ ensures the charge neutrality of the framework [224]. MIMII(BH4)3 (MI = K, Rb, Cs; MII = Ca, Sr, Sm) of perovskite-type 3D-networks have the divalent metal in octahedral coordination via κ2-bridging BH4 units, with the alkali metal cuboctahedrally surrounded by twelve BH4 units [185,225,226,227]. This perovskite-type structure is best stabilized by alkali-earth metals of smaller radii (Ca2+, 1.14 Å) than heavier analogues (Sr2+, 1.32 Å), a trend which is more apparent during heating [21,22,23,24,25,26,27,225,226].
Given the size-match between ammonium cation NH4+ (1,48 Å) and alkaline metals ionic radii K+ (1.38 Å) and Rb (1.52 Å), a family of bi-cationic borohydrides (NH4)xM(BH4)y was produced when NH4BH4 was mixed with M(BH4)p (M = Li, Na, Mg, Al, Y, Mn, Gd). These novel borohydrides exhibit high energy densities of up to 24.5 wt.% hydrogen [222].

4.4. Anion Substitution of MBH: From Light to Heavy Halides and Pseudo-Halide Substitution

Substitution of BH4 was reported for LiBH4-LiCl [228], with this mixture actually forming a partial solid solution of anticipated formula Li(BH4)1-pClp (0 < p < 1), an aspect which hindered further purification of the borohydride. The superionic HT-LiBH4 is formed at ~110 °C (108 °C), but this phase transition could occur at even lower temperatures possibly due to the formation of Li(BH4)1-pClp during ball-milling [89,228]. Easier substitution has been confirmed for halide anions (F, Cl, Br, I) and it allowed for fine-tuning of the unit cell parameters and eventually to physical properties related to the crystallographic system and strain [229]. The shrinkage or enlargement of the unit cell volume mirrors the size of the anions: F (1.33 Å) < Cl (1.81 Å) < Br (1.96 Å) < BH4 (2.05 Å) < I (2.20 Å) [153,230].
Depending on the crystallographic system and local environment, some polymorphs undergo anion substitution more easily than others, as shown by HT-LiBH4 and β-Mg(BH4)2 [231,232]. If both components of the system MBH4-MX (M-alkali metal, X- halide) are isostructural, anion substitution may occur both ways, such that two solid solutions can form: LiBH4–LiBr, LiBH4-LiI, or NaBH4-NaCl. While heavier halides (Br, I) stabilize the hydride, hindering hydrogen release at lower temperature, they lower the energy of hydrogen uptake during rehydrogenation [233,234,235,236]. While o-LiBH4 (orthorhombic) shows no conduction at RT, its HT-polymorph, HT-LiBH4 (hexagonal, stable at t > 110 °C), exhibits fast lithium-ion conductivity that is enhanced upon iodide anion I substitution of 25% with respect to starting BH4 units, in Li[BH4]0.75I0.25 (Equation (41)) [237,238]. J. Rude et al. showed through usage of synchrotron radiation powder X-ray diffraction (SR-PXD) and attenuated total reflectance infrared spectroscopy (ATR-IR) that two solid solutions of hexagonal structure (just like HT-LiBH4 and β-LiI) form in the system LiBH4-LiI, which combine into one upon heating. The solid solutions of composition Li(BH4)1-pIp (0 < p < 0.62) have been found to be stable over a 300-degree temperature range, and after four hydrogen release/uptake cycles they retained 68% of expected hydrogen capacity, in contrast to only 25% retained by pristine LiBH4.
L T L i B H 4   Δ   H T L i B H 4 L i I L i ( B H 4 ) 0.75 I 0.25
Anion substitution has also been expanded to include amide -NH2- anions, and these boron–nitrogen species are promising from a hydrogen storage viewpoint and for properties of fast ion conductivities, as seen above in the case of halide substitution.

4.5. Stabilization of M(BH4)x by Coordination of Neutral Molecules (NH3, N2H4, H2O, (CH3)2S)

While exhibiting a rich and fascinating structural diversity akin to that of metal organic frameworks (MOF), the metal borohydrides do often run into stability problems: they can be obtained solely by mechano-chemical synthesis, are stable only at low temperature, are only stable in solution or as solid solutions. These shortcomings can be overcome by neutral molecule coordination, enhancing their stability. Among these, N-containing ligands are particularly interesting because they can bind hydrogen, and could interact with the Lewis acidic boron through a frustrated Lewis pair (FLP)-type interaction: B-Hγ−γ+H-N. In fact, the shortest distance of 1.85 A has been reported for Y(BH4)3·7NH3, between the complex cation [Y(NH3)7]3+ and the borohydride anion BH4. These interactions facilitate heterolytic E-H bond cleavage (E = B, N) to produce molecular H2, thus becoming attractive options for hydrogenation studies. Moreover, this inherent instability caused by H…H interactions gives rise to structural flexibility and variability. The solvent should behave as a Lewis base, containing highly electronegative atoms: N, O, S. In this regard, the most studied ligands are N-based ligands, such as ammonia NH3, hydrazine H2N-NH2, and ammonia–borane NH3BH3. S-based ligands (such as S(CH3)2, dimethyl sulfide) are more desirable than O-based ones (diethyl ether (C2H5)2O, tetrahydrofuran C4H8O, and water H2O), because M-S bonds are longer than M-O bonds and thus the S-ligand could potentially be easier to remove.
A dihydrate adduct has been isolated in the case of NaBH4·2H2O crystallohydrate, where Na+ has octahedral coordination to four [BH4] units and two H2O molecules bridging two Na+ (space group: orthorhombic, Pbca) [153,239]. NaBH4·2H2O loses the two water molecules upon mild heating at 40 °C, and then slowly starts to release hydrogen (Equation (42)).
N a B H 4 · 2 H 2 O 40   ° C   N a B H 4 + 2   H 2 O 40   ° C N a B O 2 + 4   H 2
Monohydrates of lithium and calcium borohydrides have also been reported: LiBH4·H2O and Ca(BH4)2·H2O [240].
A large class of ammine single-metal borohydrides M(BH4)x·pNH3 and ammine bimetallic borohydrides MIMII(BH4)3·pNH3 have been synthesized and structurally characterized, containing variable solvent molecules, in the range p = 1–8 (Equation (43)) [202,241,242].
M ( B H 4 ) x + p   N H 3   M ( B H 4 ) x · p N H 3    
For instance, LiBH4·NH3, the only alkali metal borohydride stable at RT as an ammine complex [243], Mg(BH4)2·pNH3 with p = 1,2,3, and 6 [244,245,246], as well as octa-ammine complex Zr(BH4)2·8NH3 [247], have been prepared and studied. For instance, the di-solvate Mg(BH4)2·2NH3 can be written in its molecular formula [Mg(BH4)2(NH3)2], with Mg2+ coordinating in a tetrahedral fashion two BH4 units and two ammonia molecules [245]. Moreover, whereas only ions with electron configuration d0, d5, and d10 (with very few exceptions) could be obtained via the ball-milling technique, various other TM borohydride configurations (TM = transition metal) could be obtained as adducts due to enhanced stability brought about by NH3 coordination: d1 (Ti2+), d2 (V3+), d6 (Fe2+), or d7 (Co2+) [202].
Depending on the number of coordinating ligands, the thermal decomposition of ammine metal borohydrides can follow different pathways, as shown in Equation (44) [245].
M g ( B H 4 ) 2 · 6 N H 3   Δ , 4   N H 3     M g ( B H 4 ) 2 · 2 N H 3   Δ   M g H 2 + 2   B N + 6   H 2
The hexaamine complex Mg(BH4)2·6NH3 produces by thermal decomposition Mg(BH4)2·2NH3, which upon further heating releases irreversibly H2 and produces traces of BN as the sole crystalline phase. Rehydrogenation attempts at 250 °C for 60 h led to no amounts of Mg(BH4)2 [248].
A similar hexaammine complex Ca(BH4)2·6NH3 was produced by solid–gas reaction between Ca(BH4)2 and NH3(g) [249], and it is different from the solution synthesis used to prepare Mg(BH4)2·6NH3 (Equations (45)–(48)) [245].
C a ( B H 4 ) 2 ( s ) + 6   N H 3 ( g ) C a ( B H 4 ) 2 · 6 N H 3   Δ , 20   m i n ,   2   N H 3 ,     A r     C a ( B H 4 ) 2 · 4 N H 3    
C a ( B H 4 ) 2 · 4 N H 3 87   ° C , 2 N H 3 ,     A r C a ( B H 4 ) 2 · 2 N H 3 162   ° C , N H 3 ,     A r C a ( B H 4 ) 2 · N H 3  
C a ( B H 4 ) 2 · N H 3   230   ° C , N H 3 ,     A r C a ( B H 4 ) 2
C a ( B H 4 ) 2 · 2 N H 3 190 500   ° C   1 4     C a ( B H 4 ) 2 + 1 4   C a 3 ( B N 2 ) 2 + B N + 6 H 2
Reactions (45)–(47) were performed under Ar flow, when NH3 was released. Running the reaction in a closed vessel, no NH3 evolution was detected and instead a sudden dehydrogenation reaction with onset at 190 °C took place (Equation (48)). The final dehydrogenated species was assigned based on 11B NMR spectra, consistent with an overall hydrogen storage capacity of Ca(BH4)2·2NH3 of 12.3 wt.%. This diammoniate system was not reversible, as no rehydrogenation occurred during attempts at 50-bar H2 and 20–300 °C [249].
Coordination of H2N-NH2 (hydrazine) to metal borohydrides produces M(BH4)x·nN2H4 (M = Li, Na, Mg), out of which the LiBH4.N2H4 shows a remarkably high 13 wt.% H2 released at 140 °C when using Fe-B catalyst [250,251].
Ca(BH4)2 produces hydrazinates of composition Ca(BH4)2·nN2H4 (n = 1, 4/3, 2, 3) that store 8.8 wt.% H2 (n = 2), 9.2 wt.% H2 (n = 1) and 10.8 wt.% H2 (n = 4/3) (Figure 10). Z. Li et al. propose a dehydrogenation pathway involving H(δ+)…H(δ-) interaction [249] and NH3-mediated mechanisms [252]. The dehydrogenation temperature was reduced from 240 °C to 140 °C by utilizing an Fe-based catalyst (2–5 mol% FeCl3); however, the Fe3+ seemed to partially catalyze decomposition of hydrazine as well, as NH3 was found among the gaseous products (Equation (49)).
C a C l 2 + 2   N a B H 4 B M ,     T H F , 2 N a C l C a ( B H 4 ) 2 n   N 2 H 4 C a ( B H 4 ) 2 · n N 2 H 4 ; n = 1 , 4 3 , 2 ,   3      
Hydrazine itself (H2N-NH2) has two competitive decomposition pathways, leading to either (2 H2 + N2), or (4/3 NH3 + 1/3 N2); at least one side-reaction thus leads to enhanced hydrogen storage capacity (Figure 10). However, this system has not achieved reversibility yet and research is due to regenerate Ca(BH4)2 from spent Ca3(BN2)2.
Ammonium-substituted MBH were also produced via the cryo-mechanochemical route for a series of alkali metal (Li+, Na+), alkali-earth metals (Mg2+, Ca2+, Sr2+), and TM- and RE-borohydrides (Mn2+, Y3+, La3+, Gd3+) (Equation (50)) [222].
x   N H 4 B H 4 + y   M ( B H 4 ) m   196   ° C     ( N H 4 ) x M y ( B H 4 ) x + m y
These ammonium borohydrides show decomposition in the temperature range 30–60 °C and usually release toxic diborane B2H6, along with dihydrogen H2 (Equation (51)).
( N H 4 ) 3 M g ( B H 4 ) 5 35   ° C , B 2 H 6 . 2 H 2     ( N H 4 ) M g ( B H 4 ) 3 · 2 N H 3 40   ° C . H 2 M g ( B H 4 ) 2 · 2 N H 3 · N H 3 B H 3
However, not all ammino-stabilized ammonium borohydrides release diborane upon heating; (NH4)Ca(BH4)3 decomposes to release an ammonia–borane adduct C a ( B H 4 ) 2 · N H 3 B H 3 and hydrogen in the first step, producing β-Ca(BH4)2 upon further heating and loss of a molecule of ammonia borane NH3BH3. Interestingly, H2 is only released during the first, lower-temperature decomposition step; (NH4)Li(BH4)2 follows a similar decomposition pathway to release hydrogen (Equation (52)) [222].
( N H 4 ) C a ( B H 4 ) 3 111   ° C ,   H 2     C a ( B H 4 ) 2 · N H 3 B H 3 162   ° C . N H 3 B H 3 β C a ( B H 4 ) 3

5. Physical and Chemical Properties of M(BH4)x

Ionic and covalent borohydrides will intrinsically have different properties. For a rough estimation, Pauling’s empirical rule and Hannay–Smith formulae have been employed to estimate the ionicity percentage in known and reported borohydrides. As the electropositivity of the metal increases, the bonding character is more ionic, such that the most ionic borohydrides are those of alkali metal [253]. While the main properties of metal borohydrides have been polarized by their high and potentially reversible hydrogen content, other studies have expanded their field of interest, including optical, magnetic, and semiconductor materials.

5.1. Electrochemistry of Metal Hydrides and M(BH4)x: Electrodes, Electrolytes (Li+, Mg2+, Ca2+), Complex Metal Hydrides

Due to their strong reducing character, metal borohydrides have recently been investigated as electrolytes in liquid and solid form, showing conductivities σ higher than 10−4 S/cm [21,22,23,24,25,26,27,254,255]. Solid-state electrolyte research reached new heights after the discovery that o-LiBH4 (Pnma, orthorhombic, ionic conductivity σLi+ = 10−8 S/cm) transforms at 108–110 °C into another high temperature polymorph h-LiBH4 (P63mc, hexagonal, σLi+ = 10−3 at 120 °C, five orders of magnitude higher than that of o-LiBH4) (Table 2) [256,257,258,259,260,261]. As shown previously, several strategies have been employed to obtain RT-stable borohydrides with comparable ion conductivity: halide or halide-alike substitution in the complex anion (Li4(BH4)3I, Li2(BH4)(NH2)), nanoconfinement in mesoporous silica of LiBH4 or Li4(BH4)3I, or formation of rare-earth (RE)-containing bimetallic borohydrides Li(RE)(BH4)3Cl (RE = La, Ce, etc.) [262,263]. Ionic conductivity has been expanded by utilizing neutral ligands or larger closo-borate salts which prove to ease the cation migration and represent the state-of-the-art in research on borohydride materials as electrolytes.
Most conductivity studies have been carried out regarding LiBH4 and Mg(BH4)2 complexes, especially on neutral ligand-stabilized solvates of these borohydrides. It becomes apparent that ionic radii increase must be an obstacle in front of ion migration, as Ca2+ in Ca(BH4)2 shows slow conductivity (rLi+ = 0.9 Å; rMg2+ = 0.86 Å; rCa2+ = 1.14 Å). It seems that ~26% increase in ionic radii from Li+ to Ca2+ would inhibit ion migration in the framework. The interaction charge framework is important and increases with the charge of the cation, but it is not essential for conductivity. Although they have the same charge (+2) of Mg2+ and Ca2+, respective borohydrides are different in terms of conductivity, with solvated Mg(BH4)2 showing real promise as an electrolyte candidate with an oxidation stability of 1.2 V (by CV, cyclic voltammetry) (Table 2) [272,275,276]. Recent trends in improving ionic conductivity have explored complexes with weaker coordinating ligands/counteranions, as is the case for CaB12H12 [277]. Despite their low electrochemical stability, borohydrides give rise to rich speciation: nido- (typical tetrahydroborate BH4), planar- (B6H66−), and closo (B10H102−; B12H122−), making their alkali (Li+, Na+, K+) and alkali-earth (Mg2+, Ca2+) salts particularly appealing in ion conduction applications [278,279,280,281]. The ionic conductivity trend correlates well with the size of the hydroborate species (Figure 11).
The closo-borates are currently under active research due to their rapid ion conduction, due in part to the open structure with channels for ionic conduction [281]. Both Li2B12H12 and Na2B12H12 exhibit high ionic conductivities of the order of magnitude of 10−1 S cm−1 which are beneficial for Na-battery electrolytes [257,258,259]. The superionicity found in these closo-salts could also originate from the high rotational mobility of the large B12H122− anion [259].

5.2. Optical and Magnetic Properties

As expected for RE (RE = rare earth) complexes, Eu(BH4)2·2THF has blue luminescence and an impressive quantum yield of ~75%, due to charge separation of Eu2+ centers by borohydride BH4 anions, which avoid quenching. Similar favorable luminescent properties were found in Y(BH4)2·2THF [282]. The ionic bond Eu2+–BH4 allows d → f emission in the blue region of the spectrum. A series of mixed metal borohydrides of perovskite-type structure, such as KYb(BH4)3 and CsEu(BH4)3, have also been investigated, with a red shift in the latter compared to parent Eu(BH4)2·2THF [225,226]. Avoidance of quenching has been proven by addition of a 5% Eu2+ in CsEu(BH4)3, which showed similar luminescence properties regardless of [Eu2+] doping concentration [225,226].

5.3. M(BH4)x as Semi- and Superconductors

Borohydrides of perovskite structure could be fine-tuned for photovoltaics, as their typical band-gap (>5 eV) could be lowered to ~1.5 eV at RT in the particular case of CsPb(BH4)3, which has semiconductor behavior. The reasoning for this seems to lie in the partial covalent bonding in the Pb(BH4)3 framework, supported by the p(Pb) states present in the conduction band edge [225].
Superconductivity has been achieved for metal hydrides at very high pressures, exceeding 100 GPa, but providing a solid proof-of-concept. While the lanthanum-based borohydride LaBH8 is not yet confirmed experimentally, computations predict it to be a superconductor and stable at a considerably lower pressure of about 40 GPa [283,284,285]. However, recent reports (2021) suggest that under pressure (12 GPa) a ternary alkali borohydride KB2H8 shows phonon-mediated high-temperature superconductivity [226].

5.4. CO2 Capture

Reaction of MBH with CO2(g) can be used for carbon dioxide sequestration/removal, or to produce derivatives like formates (Equation (53)) [286,287].
M B H ( H C O O ) 3 ( D M E ) 3   C O 2 ,   D M E     M B H 4 Δ ,       2   C O 2   M B O ( H C O O ) ( O C H 3 ) ;     M = L i ,   N a
The reaction with CO2 can follow different pathways, depending on temperature and solvent used (DME, dimethoxyethane), which adds to the growing interest in CO2 capture globally [288,289]. LiBH4 and NaBH4 reaction with CO2 was also monitored in non-catalytic, solvent-free conditions, when IRGA (infrared gas analysis), FTIR, 11B NMR, and 13C NMR confirmed the presence, respectively, of borate, formate, and methoxy species [290].
Reaction of KBH4 under solvent-free ball-milling with solid CO2 leads to a formylhydroborate species K[HxB(OCHO)4-x] (x = 1–3) [291]. The synthesis of this interesting substituted borate species follows a three-step, consecutive nucleophilic attachment of the borohydride H on the electron-deficient carbon of CO2 molecules (Equation (54)).
    K B H 4   C O 2 K [ H 3 B ( O C H O ) ]   C O 2 K [ H 2 B ( O C H O ) 2 ] C O 2 K [ H B ( O C H O ) 3 ]
While reaction of MBH with H2O usually follows the pattern seen in Equations (42) and (55), variations including running the reactions under small CO2 amounts have proven fruitful in synthesizing novel substituted metal borohydrides [292].
M B H 4 + ( m + 2 )   H 2 O M B O 2 · m   H 2 O + 4   H 2
Contrary to the Li+ and Na+ counterparts, KBH4 borohydride produces under CO2-promoted hydrolysis a metal complex still retaining the [BH4] moiety: K9[B4O5(OH)4]3(CO3)(BH4)·7H2O. This complex was investigated by XRD, DSC, TGA, Raman, and IR spectroscopies and fully characterized by solving its single crystal structure, showing a hexagonal crystal structure belonging to the P-62c space group, and high hydrogen gravimetric content. The reaction proposed by Filinchuk et al. is complex and leads to potassium carbonate as by-product (Equation (56)), along with a large amount of hydrogen, so that it can be regarded as having a double utility: removal of CO2 from car exhaust gases, and efficient conversion of CO2 to a clean fuel, H2 [286]:
  13 K B H 4 + 37 H 2 O + 3 C O 2   K 9 [ B 4 O 5 ( O H ) 4 ] 3 ( C O 3 ) ( B H 4 ) · 7 H 2 O + 2 K 2 C O 3 + 48 H 2
The porous polymorph of γ-Mg(BH4)2 adsorbs CO2 under very mid, near-RT conditions and also yields formate and methoxy products, proving suitable for CO2 sequestration. The highest CO2 uptake of 12 mol/kg was recorded after 7 days of exposure at 1-bar CO2 and 30 °C, and was attributed to the very porous nature of magnesium borohydride polymorph, which features a very large specific surface area [293].

6. M(BH4)x as Hydrogen Storage Materials

6.1. Thermodynamic Properties of MBH

Thermal decomposition of metal borohydrides is directly related to the electronegativity of the M metal, which roughly displays a clear decrease in decomposition temperature with the increase in M electronegativity [294].
Pauling hypothesized that the energy E(A-B) of a heteroatomic bond (A ≠ B) relates to the difference in electron affinities of A and B by Equation (57), where the second term can be approximated in the case of compounds MRn (R-anion) by ΔHf/n.
χ A χ B E A B E A A + E B B 2         Δ H f n
The decomposition temperature (Td, K, and td, °C; Td[K] = td(°C) + 273.15) can be correlated to the Pauling electronegativity of the cation [295,296,297,298,299]. The empirical formula (Equation (58)) used to determine Td was found to be [300]:
T d [ K ] = 175.914 ( χ P 3.11 ) 2     o r   t d ( ° C ) = 175.914   χ P 2 1094.2   χ P + 1428.3  
The van’t Hoff equation adapted for decomposition of metal borohydrides (Equation (58)) contains the factor 3.11, which can be regarded as the electron attractive coefficient of borohydride anion [BH4], while 175.914 is a coefficient containing entropy values of decomposition (oversimplified as being constant for all compounds, in a first approximation) (Figure 12).
First-principles calculations led to the estimation of formation enthalpies for metal borohydrides, with an absolute error of only 10.4 kJ/mol (Equation (33)) [295,296,297,298,299].
Δ H b o r o = 248.7   χ P 390.8               ;   T d = 175.914 ( χ P 3.11 ) 2
A representative plot for metal borohydrides presented in Figure 2 is given in Figure 13, where a linear correlation between enthalpy and electronegativity χ P is confirmed.
Pauling electronegativities are preferred because when inferring Mulliken scaling ( χ M ), the values are fitted for ΔHboro with a higher error (13.1 kJ/mol) and moreover the enthalpies for Cu+ and Zn2+ borohydrides are almost the same (Mulliken electronegativities: 4.48 eV and 4.45 eV, respectively), which cannot explain the differences observed experimentally between the two [294,301,302]. Based on Equation (59), it can be inferred that for metal borohydrides having ΔH0f ≥ 0, that is for χ P 1.57 , the borohydride would have an endothermic formation enthalpy and could be considered unstable (Table 3). For selected M(BH4)x the free Gibbs energy ΔG has also been computed. Group I and II metal borohydrides have been the most investigated regarding thermodynamic data, and the following suggestive formulae were found (in kJ/mol formula unit); Δ G L i B H 4 = 222466.273 + 751.660356   T 124.7   T l n T [296], Δ G N a B H 4 = 217735 + 693   T 119.233   T l n T [296], Δ G N a B H 4 = 16926 21.756   T [297], Δ G K B H 4 = 19176 21.841   T [296], Δ G α M g ( B H 4 ) 2 = 222624.9 + 158.46145   T 35.22138   T l n T 0.035975   T 2 [298], Δ G β M g ( B H 4 ) 2 = 12954.437 26.4266   T [298].
Screening the known borohydrides, one might expect the following elements to form unstable borohydrides: Cu, Ag, (Be), Cr, Fe, Ni, Co, Zn, Cd, Hg, Al, V, Ga, Nb, In, Ge, and Sn. Many of these are known, but either they are stable at very low temperatures or they need other techniques for stabilizations, i.e., formation of adducts with neutral molecules (Table 2).
Many borohydrides follow a decomposition scheme leading to production of B2H6 and H2 in a 1:1 molar ratio (Equation (60)). These pathways that lead to production of diborane (toxic gas) will be, unfortunately, not useful for expanding the class of hydrogen storage material because they lead to a loss of the boron source [300].
C d ( B H 4 ) 2   75   ° C   C d + B 2 H 6 + H 2

6.2. Destabilizing Methods for Complex Metal Borohydrides

Nanoconfinement is a valuable method for the synthesis of nano-sized MBH. This route can be utilized to lower energy barriers and speed up kinetics. Nanoscaffolds can be engineered to have highly advantageous textural properties: tunable pore size (Dp), as well as high pore volume (Vp) and specific surface area (SSA), which in turn will greatly enhance the kinetics of hydrogen release/uptake. These recyclability enhancements will be described for representative metal borohydride systems and their composites. Anion substitution in complex metal hydrides can lead to system destabilization: TiF3 and TiCl3 were investigated in NaBH4 and showed a 6.6 kJ/mol decrease in decomposition enthalpy, by altering hydrogen release and uptake [303]. The replacement of H by F was investigated in LiBH4 as well, showing significant theoretical thermodynamic improvements [304]. Nanoconfinement improvements brought about by size-restriction are discussed in relation to their bulk state in the following chapter (Section 7). Some volatile borohydrides such as Zr(BH4)4 could be embedded into mesoporous silica (MCM-48 type) to show enhanced stability towards air and moisture, and more importantly, to retain Zr(BH4)4 inside siloxanic pores, thus mitigating the high volatility issue of neat Zr(IV) borohydride.
Thermal destabilization by utilizing mixtures of metal borohydrides has proven to be an effective technique [305]. Coupled dehydrogenation of a physical mixture (1 − x)LiBH4 + xCa(BH4)2 comprised of o-LiBH4 and α-/γ-Ca(BH4)2 (synchrotron XRD patterns) also led to enhanced ionic conductivity of the resulting cubic-CaH2. Moreover, it seems that the ionic conductivity may be solely due to Li+ and irrespective of Ca2+ content, as shown by samples (1 − x)LiBH4+xCa(BH4)2 with similar ionic conductivity to that of pure ball-milled LiBH4 [305]. Further investigations by first-principles computations of a similarly coupled system by Ozolins et al. have shown that the decomposition reaction yields the complex intermediate Li2B12H12 alongside the anticipated CaH2 and hydrogen H2, with an endothermic enthalpy of 37.9 kJ/mol and 6.7   wt . %   H 2 (Equation (61)) [306].
  5 C a ( B H 4 ) 2 + 2 L i B H 4   83   ° C L i 2 B 12 H 12 + 5 C a H 2 + 13 H 2 ;                                 Δ H 300 K = 37.9 kJ mol
Using Mg(BH4)2 with either LiBH4 (Equation (62); 8.4 wt.%H2) or Ca(BH4)2 (Equation (63); 7.7 wt.%H2) produces similar chemical reactions, but at negative temperatures which therefore would hinder their usage under ambient conditions.
  5 M g ( B H 4 ) 2 + 2 L i B H 4   29   ° C L i 2 B 12 H 12 + 5 M g H 2 + 13 H 2 ;                           Δ H 300 K = 24.4 kJ mol
  5 M g ( B H 4 ) 2 + C a ( B H 4 ) 2   18   ° C C a B 12 H 12 + 5 M g H 2 + 13 H 2 ;                   Δ H 300 K = 25.7 kJ mol
Decomposition of group II metal dodecaborohydrides MB12H12 seems to occur around 300 °C (309 °C for MgB12H12 and 306 °C for CaB12H12), but the group I counterpart shows unfavorable thermodynamics (LiB12H12, ΔH = 116.7 kJ/mol, td = 696 °C) (Equations (64) and (65)).
  L i 2 B 12 H 12   696   ° C 2 L i H + 12 B + 5 H 2 ;                                                                       Δ H 300 K = 116.7 kJ mol  
  C a B 12 H 12   306   ° C C a B 6 + 6 B + 6 H 2 ;                                                                               Δ H 300 K = 73.7 kJ mol
Further hydrogen release by reaction of [B12H12]2− complex anion with the metal hydrides MH2 should lead to additional hydrogen release (Equations (66) and (67)). It is thermodynamically more favored the addition of CaH2 rather than MgH2 to CaB12H12, as the switch from MgH2 to CaH2 used for destabilization could lower the operating temperature of the system even more, from 144 °C to 86 °C [306].
  C a B 12 H 12 + 3 M g H 2   144   ° C 3 M g B 2 + C a B 6 + 9 H 2 ;                               Δ H 300 K = 53.2 kJ mol
  C a B 12 H 12 + C a H 2   86   ° C 2 C a B 6 + 7 H 2 ;                                                                         Δ H 300 K = 47.0 kJ mol
Ca(BH4)2-M(NH2)2 destabilized system (M = Mg, Ca) is a typical and representative system producing significant thermodynamic alteration for decomposition of Ca(BH4)2. Destabilization of Ca(BH4)2 was attempted with various group II amides (Mg(NH2)2, Ca(NH2)2) leading to mixed-metal-/metal nitridoborates [ C a 3 M g 6 ( B N 2 ) 6 ] and C a 9 ( B N 2 ) 6 , respectively. The reasoning for such an outcome lies in the interaction of [BH4] and [NH2] anions, and more precisely in the B-H…H-N interaction of dihydrogen bonding type (Figure 14) [307,308].
While the dehydrogenation onset of bulk Ca(BH4)2 is roughly 320° (290onset-361peak), the binary system Ca(BH4)2-2Mg(NH2)2 (8.3 wt.% H2) and Ca(BH4)2-2Ca(NH2)2 (6.8 wt.% H2) starts to release hydrogen at 220 °C, so that a significant ΔT of 100 degrees is produced. Although some NH3 release was recorded, XRD study of the final dehydrogenated binary systems indicates Equations (68) and (69) as taking place when heating the samples below 480 °C [309].
C a ( B H 4 ) 2 + 2 M g ( N H 2 ) 2     220 480   ° C 1 3 [ C a 3 M g 6 ( B N 2 ) 6 ] + 8 H 2
C a ( B H 4 ) 2 + 2 C a ( N H 2 ) 2     220 480   ° C 1 3   C a 9 ( B N 2 ) 6 + 8 H 2
The binary systems were prepared by BM (ball-milling technique, 200 rpm, 5 h, Ar) while the nature of the evolved gases was assessed by MS-coupled TPD (temperature-programmed desorption) [309]. This pattern of destabilization has been reported as well for LiBH4-LiNH2 [310], NaBH4-NaNH2 [311], Ca(BH4)2-LiNH2 [312,313], and Ca(BH4)2-NaNH2 [313] systems, and they all led to new complex hydrides.
Table 4 summarizes the main destabilization systems known for metal borohydrides; these include destabilization by metal, metal oxide, metal chloride, amides, and even metal borohydrides. The decomposition pathway is often altered by the molar ratio M(BH4)x: destabilizing agent. For instance, a clear dehydrogenation proposal (Equation (70)) was advanced by Aoki et al. for LiBH4-2LiNH2 [314], while others reported the dehydrogenation for LiBH4-LiNH2 without an explicit mechanism.
  L i B H 4 + 2 L i N H 2   249   ° C   L i 3 B N 2 + 4 H 2   ;                 Δ H = 23 kJ mol H 2 ;         11.9   wt . %   H 2
Somer et al. propose two possible reactions in the NaBH4:NaNH2 1:1 and 1:2 systems (Equations (71) and (72)), with very different outcomes. The 1:1 system leads to the production of amorphous boron B and Na metal, whereas the 1:2 mixture leads to hydrogen evolution and a unique solid product, trisodium dinitridoborate Na2BN2 [311].
N a B H 4 + N a N H 2   400   ° C   N a H + B + N a + 1 2 N 2 + 3 2 H 2
N a B H 4 + 2 N a N H 2   297   ° C   N a 3 B N 2 + 4 H 2
When LiH was mixed with γ-Mg(BH4)2 in various molar ratios (1:0.5, 1:0.97, and 1:1.89), it was found that the decomposition follows the one of pure products resulting from a metathetical reaction (Equation (73)). The sample 1 γ-Mg(BH4)2: 0.5 LiH showed the highest hydrogen storage capacity, essentially equal to the theoretical value [315].
M g ( B H 4 ) 2 + 2 L i H   380 420   ° C   2 L i B H 4 + M g H 2
Destabilization of Ca(BH4)2 with LiNH2 in a 1:4 molar ratio yields a cubic phase of an anion-substituted borohydride [Ca(BH4)(NH2)], alongside a hydro-nitridoborate phase of Li4BN3H10. Upon heating to either 288 °C (5 h, undoped sample) or 178 °C (5 h, 5 wt.% CoCl2), the system 1 Ca(BH4)2: 4 LiNH2 releases 8 wt.% and 7 wt.% hydrogen, respectively (Equation (74)).
C a ( B H 4 ) 2 + 4 L i N H 2 B M L i 4 B N 3 H 10 + [ C a ( B H 4 ) ( N H 2 ) ] 288   ° C ,   5   h 1 4 L i C a 4 ( B N 2 ) 3 + 5 4 L i 3 B N 2 + 8 H 2
DFT computation by Ozolins et al. showed various dehydrogenation pathways for binary borohydride–borohydride, borohydride–amide and borohydride–metal systems (Equation (75)) [306]. The system 5Mg(BH4)2 + 2LiBH4 (8 wt.% H2) has a decomposition temperature of −29 °C and is therefore ruled out from usage in mobility applications; however, the Ca2+-counterpart 5Ca(BH4)2 + 2LiBH4 (6.7 wt.% H2) has a reasonably low hydrogen release temperature of 83 °C [306].
5 M ( B H 4 ) 2 + 2 L i B H 4 L i 2 B 12 H 12 + 5 M H 2 + 13 H 2   ;           M = M g ,   C a              
Similarly, the system 5Mg(BH4)2 + Ca(BH4)2 with 7.7 wt.% H2 (theoretical) was predicted by first-principles computations to release H2 at −18 °C [306], but was actually found experimentally to only dehydrogenate around 150 °C, and the reaction was not reversible (Equation (76)) [316,317].
5 M g ( B H 4 ) 2 + C a ( B H 4 ) 2 C a B 12 H 12 + 5   M g H 2 + 13 H 2
Assuming that in composites of type 2Mg(BH4)2 + Ca(BH4)2 a temperature high enough is reached (above 306–309 °C, the decomposition temperature for plausible intermediates CaB12H12 and MgB12H12), one may consider that final decomposition products are those expected from individual borohydrides, so that the overall decomposition can be formulated as in Equations (77) and (78), with a theoretical hydrogen storage capacity of 10.5–12.75 wt.% [316,317,318].
2 M g ( B H 4 ) 2 + C a ( B H 4 ) 2   1 3 C a B 6 + 2 3 C a H 2 + 4 B + 2 M g H 2 + 28 3 H 2   ; 10.5 wt . % H 2            
2 M g ( B H 4 ) 2 + C a ( B H 4 ) 2   1 3 C a B 6 + 2 3 C a H 2 + 2 M g B 2 + 34 3 H 2   ;       12.75 wt . % H 2                        
Table 4. Representative examples of destabilized borohydride systems: binary borohydride–amide systems M(BH4)x-RNH2, borohydride–borohydride systems M(BH4)x-R(BH4)y, borohydride–chloride MIBH4-MgCl2, borohydride–TM oxide, borohydride–metal.(td: decomposition temperature).
Table 4. Representative examples of destabilized borohydride systems: binary borohydride–amide systems M(BH4)x-RNH2, borohydride–borohydride systems M(BH4)x-R(BH4)y, borohydride–chloride MIBH4-MgCl2, borohydride–TM oxide, borohydride–metal.(td: decomposition temperature).
SystemCatalysttd (°C)Obs.Reference
LiBH4–LiNH2 -
via metastable Li2[BH4][NH2] and Li4[BH4][NH2]3
95melt-160onset-315peak, 230mean5.8 wt.%H2;
H2 major; small traces of NH3 (TPD-MS data)
[310]
LiBH4-2LiNH2-
(LiH possible intermediate)
2497.8 wt.%H2
LiBH4: 75 kJ/molH2;
LiBH4-2LiNH2: 23 kJ/molH2
[314]
x NaBH4–NaNH2 (x = 1,2,3,4)-
via α-/β- Na2[BH4][NH2]
265onset-350peak,297 (2:1); 400 (1:1)8 wt.%[311]
γ- Mg(BH4)2-0.5LiH-380–42015.5 wt.%exp., 14.78 wt.% (theoretical) *
* higher wt.% exp. due to usage of solvated γ- Mg(BH4)2
[315]
Ca(BH4)2–4 LiNH2-250onset-320peak (288mean)8 wt.%[312,313]
Ca(BH4)2–4LiNH2-5 wt.% CoCl2Co2+ (CoCl2)150onset-207peak (178mean)>7 wt.%[312]
Ca(BH4)2-2Mg(NH2)2-220–480
(270, 290 and 310–multistep)
8.3 (8.8 wt.% theoretical)
not reversible (50-bar H2, 20–300 °C)
[309]
Ca(BH4)2-2Ca(NH2)2-220–480 (270, 290 and 310–multistep)6.8 (7.5 wt.% theoretical)
not reversible (50-bar H2, 20–300 °C)
[309]
5Ca(BH4)2-2LiBH4-836.7 wt.% *
* theoretical
[306]
5Mg(BH4)2-2LiBH4-−298.4 wt.%*
* theoretical
[306]
2Mg(BH4)2-Ca(BH4)2-272, 326, 346, 398
(rehydrogenation at 288onset, 273peak)
10.5 wt.% *
* theoretical
[316,317,318]
5Mg(BH4)2-Ca(BH4)2-−18 °C (p = 1atm) *
* predicted based on DFT calculations;
150 °C
7.73 wt.%H2 *
* predicted based on DFT calculations
[306]* [316,317]
LiBH4 + 0.2 MgCl2 + 0.1 TiCl3;
LiBH4 + 0.076 MgCl2 + 0.047 TiCl3
MgCl2, TiCl360onset-400peak5 wt.%H2[319,320]
LiBH4 + 0.09 TM oxides (TiO2, V2O3)TiB2-possible active intermediate2007–9 wt.%H2;
reversible
[319]
LiBH4 + 0.2 M (M = Mg, Al)Mg/Al60–300fast-6009 wt.%H2;
Al-based yields material probably volatile, only recharges to 3.5 wt.% capacity
[320]
LiBH4 + 0.0897 AlAl45012.4 wt.%, partially reversible[321]
Combined Kissinger plot and Arrhenius equation allowed establishing the first-order kinetic parameters for dehydrogenation reaction of binary borohydride–amine systems (Table 5).
The specific rate for Ca(BH4)2-2Ca(NH2)2 shows a 3.5-fold increase in reaction rate, when compared to Ca(BH4)2-2Mg(NH2)2 system. For comparison, pristine Ca(BH4)2 has slow desorption kinetics that onsets at 290° and peaks at 361 °C [309].

7. Model Systems for Hydrogen Storage of MBH and Other RMH (Reactive Metal Hydrides): Nanoconfinement vs. Bulk Behavior for Improved Thermodynamics and Kinetics

7.1. LiBH4

Being the lightest metal borohydride and potentially having some of the highest hydrogen storage capacity (18.4 wt.%), LiBH4 was first prepared in 1940 by Schlesinger and Brown by an organolithium compound and diborane B2H6 (Equation (18)) [102], and afterwards investigated in detail by many researchers [125,322,323,324,325]. Transition to a h-LiBH4 phase at 108–112°C exhibiting unexpectedly fast Li+ ion conductivity drove even further the pace of MBH research. Following melting (275 °C), decomposition occurs above 400 °C with an endothermic reaction enthalpy of ΔHd = 74 kJ/mol [326], and leads to formation of LiH, B, and H2, which account for a real practical hydrogen storage of 13.8 wt.% H2 (Equations (79)–(81)) [327].
        o L i B H 4   ( s ) 108 112   ° C h L i B H 4   ( s ) 275   ° C L i B H 4   ( l ) 400 600   ° C   L i H ( s ) + B ( s ) + 3 2 H 2
L i B H 4   ( s ) Δ 5 6   L i H ( s ) + 1 12   L i 2 B 12 H 12   ( s ) + 13 12   H 2   ;                     Δ H = 56 kJ mol
L i B H 4   ( s ) Δ   L i H ( s ) + 1 2   B 2 H 6   ( g )   L i H ( s ) + B ( s ) + 3 2   H 2         Δ H = 125 kJ mol        
Synchrotron radiation XRD (SR-PXD) showed the presence of many intermediate phases during decomposition of LiBH4 [322]. These intermediates (Equations (80) and (81)) are supported by Raman spectroscopy and 11B-NMR experiments [328,329]. The data provided by Zuttel et al. show that the two-step dehydrogenation of LiBH4 proceeds via Li2B12H12 formation (10 wt.% H2 release, ΔH = 56 kJ/mol, Equation (80)), and a highly endothermic second step leading to LiH and B, releasing 4 wt.% H2 (ΔH = 125 kJ/mol) [200]. A thermodynamic data scheme is pictured by Y. Yan et al., who assigned enthalpy effects to the intermediate species formation during o-LiBH4 decomposition [164].
The 11B NMR data showed B2H6 to be an impurity during dehydrogenation processes. Therefore, running the reaction of LiBH4 with B2H6 confirmed the formation of the hypothesized Li2B12H12 during the dehydrogenation pathway (Equation (82)) [330].
2   L i B H 4   + 5   B 2 H 6 Δ   L i 2 B 12 H 12   + 13   H 2
Attempts to lower the hydrogen desorption to temperatures below 400 °C have been successful when using SiO2 powder addition (LiBH4:SiO2 = 1:3) [331]. Three important peaks in TPD spectra of LiBH4 corroborated the NMR studies, Raman spectroscopy, and XRD data, all of which point to the formation of dodecahydro-closo-dodecaborate species [B12H12]2−, and possibly other boron hydrides as intermediates [328,329].
Rehydrogenation to reform LiBH4 is possible, but it requires extreme conditions (70–350-bar H2, t > 600 °C) and will not proceed to completion, thus leading to continual decrease in hydrogen capacity with the number of hydrogen release–uptake cycles (Equations (10) and Equation (83)) [48,326,332,333].
L i H ( s ) + B ( s ) + 3 2 H 2   690   ° C ,     197.4   a t m   H 2 L i B H 4   ( s )
Nanoconfinement of LiBH4 was investigated in a variety of supports (2–25 nm), such as mesoporous silica (MSU-H, SBA-15 etc.) or mesoporous carbon (carbon replica of MSU-H silica: C-MSU-H, or of SBA-15 silica: C-SBA-15 i.e., CMK-3) (Figure 15) [334,335,336]. A pertinent overview of C-replica synthesis of mesoporous silicas comes from M. Kruk et al. [337] and R. Ryoo et al. [338]. Mesoporous SiO2, however, is to be avoided, since an irreversible reaction with LiBH4 during dehydrogenation could lead to lithium silicates (Figure 16) [335,339]. Attempts to improve hydrogen uptake by using Mo-decorated mesoporous silica (MSU-H type) increased the reversible hydrogen capacity from 3.2 wt.% (pristine LiBH4) to 5.2 wt.% H2 in the nanocomposite LiBH4-Mo:MSU-H [335].
Formation of lithium silicates (Li2SiO3, Li4SiO4) was confirmed by XRD and synchrotron powder XRD, but only after LiBH4 decomposed in the as-synthesized composite LiBH4@meso-SiO2; the nanocomposite seems otherwise stable at temperatures t < 275 °C (Equation (84)) [89,335,340].
4 L i B H 4   + 3 S i O 2 Δ   2 L i 2 S i O 3 + S i + 4 B + 8 H 2   2   L i B H 4 3 2 L i 4 S i O 4 + 1 2 S i + 2 B + 4 H 2
Nanoconfinement can be performed by melt infiltration or solvent infiltration, both of which lead to a decrease in the crystallite size of the LiBH4 [341,342]. Incipient wetness technique has been used to infiltrate MTBE (methyl tert-butyl ether) solution of LiBH4 in mesoporous carbon (C-MSU-H). The maximum loading attempted was about 40 wt.% LiBH4, but the XRD data confirms that even for a 20 wt.% loading some diffraction peaks of LiBH4 appear. This is indicative of the borohydride species crystallizing outside of the mesopores [334].
Interestingly, the lowest onset for the as-prepared nanocomposites LiBH4@C-MSU-H was recorded at 150 °C for the lowest investigated borohydride loading (8 wt.%), while the others showed decomposition starting at 170 °C (20 wt.% LiBH4) and 200 °C (40 wt.% LiBH4), with effect ascribed to the better dispersion of the borohydride inside the mesoporous support at lower loadings (Figure 17) [334,335].
The influence of carbon porosity was investigated showing that a proper impregnation of LiBH4 inside 1.8–3.2 nm activated carbon (SBET = 1341 m2/g; Vp = 0.87 cm3/g) is accompanied by reduction of both surface area (to 245 m2/g) and pore volume (to 0.2 cm3/g) [336,343]. The size of the scaffold alters the activation energy Ea of dehydrogenation: non-porous graphite (146 kJ/mol LiBH4), 24 nm RF-CA (111 kJ/mol LiBH4), and 13 nm RF-CA (103 kJ/mol LiBH4), consistent with inverse proportionality between nanoscaffold size and activation energy Ea [344,345]. The Arrhenius Equation (85) shows that, taking solely into account the Ea of these three supports, one might expect the dehydrogenation rates to follow this order: kgraphite:k24nm:k13nm = 1:e14.03:e17.20 (at 298.15 K), and kgraphite:k24nm:k13nm = 1:e11.28:e13.86 (at 373.15 K).
k = A   e E a R T
This approximation was performed assuming the same pre-exponential Arrhenius factor A (an oversimplification), but this shows that nonporous carbon barely has any benefit from confinement, whereas the rate in nanoconfined composites is ~106 times faster. Moreover, reducing the nanoscaffold from 24 nm to 13 nm affords a rate that is 23.8 times higher (at 25 °C) or 13.2 times higher (at 100 °C, a temperature considered optimal for operation of fuel cell operation). Nanoconfined LiBH4 shows a decrease in melting temperature of 30 °C (in 13 nm RF-CA) or 23 °C (in 24 nm RF-CA), but only when melt infiltration was employed; a modest 1 °C decrease was recorded for the solvent infiltration technique [341,346].
When a 33.67 wt.% loading of LiBH4 was achieved in C-SBA-15 (CMK-3 carbon, average diameter 4 nm) prepared by the concentrated solution infiltration technique, a clear reduction in temperature onset of dehydrogenation was recorded, and the composite LiBH4@CMK-3 exhibited complete dehydrogenation up to 400 °C [336]. The solvent infiltration approach can alter the dehydrogenation pathway, as well as the reaction kinetics; for comparison, a ball-milled sample containing LiBH4 affords full dehydrogenation above 450 °C [347].
While bulk dehydrogenated LiBH4 (LiH and B) can be rehydrogenated under harsh conditions (p, t, Equation (83)), the nanoconfined borohydride can be regenerated using 100-bar H2 at below 400 °C within 2 h when microporous carbon was used as scaffold. The improved recyclability of such a composite points to better reversibility when pore size decreases to a microporous range (Dpore < 2 nm) [341]. A comparison between nanoconfinement of LiBH4 in carbon and silica supports of pore radii 8–20 Å showed improvement in desorption kinetics, reversibility, and Li+ ionic conductivity with decreasing pore size. It was shown that similarly sized silica and carbon scaffolds produce different effects upon the melt infiltration of LiBH4, with silica-based support showing obvious improvements compared to carbon scaffolds, an aspect attributed to increased LiBH4 interfacial layer thickness “t” for SiO2 (1.94 ± 0.13 nm) compared to C-based support (1.41 ± 0.16 nm). In this regard, the Gibbs–Thomson equation correlates with the shift (depression) in melting temperature (ΔT) to interfacial effects, namely, it increases with the interface energy (Δγ: 0.053 J/m2 for LiBH4/SiO2 and 0.033 J/m2 for LiBH4/C system) and decreases with pore radius (rp) (Equation (86)) [348,349].
Δ T T 0 = 2 Δ γ V m Δ H ( r p 1 )  
The phase transition shift for LiBH4 (~114 °C for bulk) was observed at 65 °C for LiBH4/SiO2 (3.3 nm) and at 19 °C for LiBH4/Cm-4.9 (of high microporosity). P. Ngene et al. assumed a core-shell model and found a direct relation for computing the enthalpy for the confined material, which could allow fine-tuning of thermodynamic parameter by adjusting pore size and interfacial thickness (Equation (87)) [350].
Δ H c o n f i n e d Δ H b u l k = ( 1 t r ) 2
Pore size distribution (PSD) and pore geometry should always play a decisive role in nanoconfinement results, especially for C-nanoscaffolds where microporosity percentage could be dominant [351]. The nano-synergy effect has recently made possible a reversible hydrogen capacity of 9.2 wt.% [352]. Investigation of possible side-pathways in a dehydrogenation reaction and particularly establishing the nature of intermediate species (such as Li2B12H12) could trigger new directions in nanoconfined borohydride materials [41,323,324,325].

7.2. LiBH4 + MgH2

Magnesium-based materials (alloys, hydrides, alanates, borohydrides, etc.) have sparked interest as promising hydrogen storage materials due to their low cost, high wt.% hydrogen storage, and interesting properties, including the much sought-after recyclability/reversibility condition for any future energy storage material [353]. Among all hydrides, MgH2 is very promising due to high abundance of magnesium, low cost, and reversibility behavior in hydrogenation studies, along with high hydrogen uptake capacity (7.7 wt.%) and very high energy density (9 kJ/g Mg) [354,355,356,357,358,359]. However, the desorption kinetics are slow at 300 °C and 1-bar H2, while the reactivity towards air (oxygen in particular) and the hydrogen desorption has a high enthalpy—thus requiring operation at high temperature [355,360]. Lowering desorption enthalpy can be achieved by introducing high-reactivity defects in the structure (high-energy ball-milling) reaction with a destabilizing element or with a catalyst [361].
Given the advantages of using the Mg-based hydrides mentioned above, inclusion of MgH2 in various reactive hydride composites (RHCs) has been attempted. The most studied such reactive composite is 2LiBH4 + MgH2 (11.6 wt.% hydrogen storage capacity, according to the decomposition pathway in Equation (88). While taken individually, both LiBH4 and MgH2 have unfavorably high thermal stability and slow kinetics, and their 2:1 mixture is found to produce MgB2 instead of the rock-stable boron B, which brought new hopes in pursuit of a viable energy-storage material (Equation (88)) [92,306,362].
2   L i B H 4 ( l ) + M g H 2 ( s )   2   L i H ( s ) + M g B 2 ( s ) + 4   H 2     ;         Δ H = 45.7 kJ mol  
This RHC has favorable thermodynamics and the formation of MgB2 reduces the dehydrogenation enthalpy to ΔH(88) = −46 kJ/mol H2, which makes the reaction reversible under mild conditions (3–20-bar H2, 315–450 °C) [92,362,363,364,365,366,367,368,369]. Similar destabilization has been observed for the (LiBH4 + 2CaH2) system, with an endothermic enthalpy slightly less than 41.3 kJ/mol and 11.7 wt.% H2 storage (Equation (89)) [306].
2   L i B H 4 ( l ) + C a H 2 ( s )   2   L i H ( s ) + C a B 2 ( s ) + 4   H 2     ;         Δ H = 41.3 kJ mol
Nanoconfinement of the (2LiBH4 + MgH2) system was attempted in 21 nm RF-CA porous scaffolds showing 4 wt.% reversible hydrogen storage capacity (92% of theoretical storage of the composite RHC@RF-CA, after three cycles), and it was found that decomposition in the nanoconfined state mirrors that from the bulk (Equation (88)) [346]. The hydrogen capacity is higher than that of nanoconfined MgH2@RF-CA22nm (1.4 wt.%) [133].
DSC analysis of the composite 2LiBH4-MgH2 shows four endothermic peaks: 113, 267, 332, and 351 °C for the nanoconfined state in 21 nm RF-CA vs. 117, 290, 364, and 462 °C in bulk (Table 6) [346]. The hydrogen desorption kinetics are clearly improved by nanoconfinement, with MgH2 registering a 32 °C lowering of dehydrogenation temperature, while dehydrogenation of LiBH4 is lowered by a significant 111 °C.
Regardless of the kinetic improvements during recyclability studies, the high thermal stability of bulk metal borohydrides remains an obstacle in their use as hydrogen storage materials.

7.3. NaBH4

With a hydrogen storage capacity of 10.6 wt.%, NaBH4 is a valuable reducing reagent used in organic synthesis and decomposes at a high temperature typical of ionic compounds like metal salts (534 °C) [370,371,372,373,374]. The synthesis was reported first by Schlesinger, who reacted trimethyl borate with sodium hydride in hydrocarbon oil, at boiling point (250 °C) (Equation (90)) [375]. The sodium methoxide is hydrolyzed (NaOCH3 + H2O → NaOH + CH3OH) to form CH3OH, which allows separation of the reactive mixture, and a sodium hydroxide solution of NaBH4 that allows borohydride extraction using isopropylamine.
4   N a H + B ( O C H 3 ) 3   N a B H 4 + 3   N a O C H 3
An alternative method (Bayer process) utilizes Na2B4O7·7SiO2 with Na and 3 atm H2 at 400–500 °C in a one-pot synthesis [376,377]. This process allows for obtaining large quantities of NaBH4, but it does present the risk of explosion (the working temperature is above decomposition of NaBH4) and it produces large quantities of Na2SiO3 of lower market demand or value (Equation (91)).
N a 2 B 4 O 7 + 16   N a + 8   H 2 + 7   S i O 2   700   ° C 4   N a B H 4 + 7   N a 2 S i O 3
Given the lower cost of Mg as compared to Na, a modified Bayer process uses MgH2 as reducing reagent (Equation (92)) [378,379,380].
1 4 N a 2 B 4 O 7 + 2   M g H 2 + 1 4 N a 2 C O 3   2 M g O , 1 4 C O 2   N a B H 4   2   M g O   N a B O 2 + 2 M g H 2
The decomposition is different to that of LiBH4 analogue, due to the lower stability of NaH compared to LiH, and it leads to the formation of the parent elements (Equation (93)). The decomposition process of NaBH4 starts at ~240 °C and releases the greatest amount of hydrogen above 450 °C.
N a B H 4 534   ° C   N a + B + 2 H 2

7.4. Mg(BH4)2

With a high theoretical gravimetric energy storage capacity (14.8 wt.%), and reasonably low dehydrogenation enthalpy (ΔH = −39.3 kJ/mol H2), Mg(BH4)2 remains an attractive candidate for hydrogen storage. First-principles calculations on the ground-state crystal structure of Mg(BH4)2 confirm these energetic characteristics [381]. Decomposition shows several steps in order to finally produce MgB2 and H2 (Equation (94)). Formation of intermediate species such as MgB12H12 is also implied [382,383,384,385,386].
α M g ( B H 4 ) 2 190   ° C   β M g ( B H 4 ) 2 3   H 2 ( 11.1   w t . % ) M g H 2 + 2 B   H 2   ( 4.2   w t . % )   M g B 2
In fact, an alternative reaction mechanism has been proposed concerning formation of the MgB12H12 species in a first step (−15 ppm, singlet, in 11B-MAS NMR spectrum), dehydrogenation of MgH2 in a second step, and a final reaction of Mg and MgB12H12 forming MgB2—which we know is the final dehydrogenation “resting-state” of magnesium (Figure 18).
The decomposition temperature could be lowered to 100°C by the addition of TiCl3 catalyst [109]. Rehydrogenation of MgB2 under 100-bar H2 at 350 °C showed ~3 wt.% hydrogen capacity, confirming the reversibility of this dehydrogenation step [156,385]. The intermediacy of MgB12H12 was investigated, as in the case of other metal borohydrides, and is considered plausible [387].
The first decomposition step (277 °C) has an enthalpy of −40 …57 kJ/mol, depending on experimental data [385,387], and a computed value of 38 kJ/mol (DFT computations) [388]. In situ synchrotron radiation SR-PXD data show as final dehydrogenation products a mixture of Mg, MgO, and MgH2, but these results should be considered with care given the high affinity of Mg for oxygen and proneness to subsequent oxidation [118].
Regeneration of Mg(BH4)2 by hydrogenation of MgB2 requires high H2 pressure and temperature (950-bar H2, 400 °C), but also prolonged time (108 h), which makes the bulk material impractical for use in a tank for mobile applications [386]. When starting from nano-sized MgB2 synthesized at RT, a partial rehydrogenation to Mg(BH4)2 occurred at 300°C and 30-bar H2, leading to a material with 2.9 wt.% hydrogen storage, but quite far off the theoretical value for magnesium borohydride (14.8 wt.%) [389,390].
Another polymorph of magnesium borohydride, γ-Mg(BH4)2, has recently shown promising results when Co-based additives were used (2 mol% CoX2, X = F, Cl), desorbing 4 wt.% H2 at 285 °C and 2.5-bar H2 backpressure. The reverse reaction took place at 285 °C, reforming the more stable β-Mg(BH4)2, but using 120-bar H2 was needed to yield a material with 2 wt.% hydrogen storage capacity. In addition, the γ → ε phase transition temperature of Mg(BH4)2 was reduced using CoX2 by ~50 °C [391,392].
Mg-based materials greatly benefit from nanoconfinement in carbon nanotubes, which reduces the enthalpy of dehydrogenation even by 50% (for MgH2, a reduction from 36 kJ/molH in bulk, to 17 kJ/molH in 9-unit cell thick MgH2 thin film slab) [393]. Nanoconfinement of Mg(BH4)2 in microporous carbon AC resulted in nanocomposites with loadings as high as 44 wt.%, albeit at the cost of partially crystallizing on the outer surface of AC (as seen by powder-XRD and EDS data analysis) [382,383]. The total weight loss of 6 wt.% corresponds to 44% borohydride loading in the nanocomposite [382]. Small angle neutron scattering (SANS) is a technique allowing for the study of particle size and distribution. Using isotopically labeled magnesium borohydride, Mg(11BD4)2@AC, Sartori et al. observed nanoconfined borohydride particles of less than 4 nm [383]. Improved hydrogen release is obvious from the shifting of the endothermic peak events from 269 °C to 257 °C (−12 °C decrease) and from 376 °C to 317 °C (−50 °C decrease), and it can be ascribed to the nanoconfinement effect.

7.5. Ca(BH4)2

Ca(BH4)2 has a hydrogen storage capacity of 11.5 wt.% (maximum theoretical capacity), but the real capacity is only 9.6 wt.% considering the actual dehydrogenation reaction (Equations (12) and (95)). Equation (95) is essentially the same as Equation (12), but written for one mol Ca(BH4)2, as the reaction is reversible and decomposition and rehydrogenation processes follow the same reaction, having a ΔH = 32 kJ/mol [188].
  α C a ( B H 4 ) 2 2 3   C a H 2 + 1 3 C a B 6 + 10 3   H 2   ;           Δ H = 32 kJ mol   H 2
The orthorhombic (α-phase) can be obtained free of THF solvent by evacuation under vacuum of the commercially available THF adduct. There is another low-temperature orthorhombic phase proposed for Ca(BH4)2 (γ-phase polymorph) [118]. The decomposition temperatures are higher than those anticipated based on thermodynamic data, partly due to sluggish kinetics [394]. Heating α-Ca(BH4)2 to 170 °C leads to a phase transition to β-Ca(BH4)2 and upon further heating the DSC scan data shows two endothermic peaks at 360 °C and 500 °C [395,396]. The rehydrogenation of this reversible system can proceed at 400 °C and 690 atm H2, but these conditions are too harsh to warrant the use of bulk Ca(BH4)2 in vehicular applications [95]. The influence of using catalysts (Nb, Ti) showed beneficial reduction in hydrogen desorption temperature, but the rehydrogenation conditions are far from ideal (350 °C, 150-bar H2, 12–24 h) and the re-hydrogenated material exhibits 4.5 wt.% H2 capacity [397]. Using additives like TiCl3 could lower even further re-hydrogenation kinetics, making it possible when dehydrogenated material was subjected at 350 °C for 24 h at a lower pressure of 90-bar H2 [394]. Ball-milling Ca(BH4)2 with 5 mol%TiCl3 (Ca(BH4)2:TiCl3 = 1:0.05) leads to hydrogen evolution and formation of CaCl2, according to Equation (96) [394].
  7 C a ( B H 4 ) 2 + 4 T i C l 3   6 C a C l 2 + C a B 6 + 4 T i B 2 + 28   H 2
This is also an example where the additive (TiCl3) alters the dehydrogenation pathway (commonly accepted as described by Equation (95)), as the proposed weight-loss based on DSC and TGA curves conform to about 7.1 to 8.7% reduction. This is consistent with the following decomposition pathway, producing CaH2, B, and H2 (7.2 wt.% hydrogen storage capacity based on Equation (97)) [394]. The hydrogen storage capacity after rehydrogenation is only 3.8 wt.%; complete rehydrogenation was thus not achieved. It is, however, apparent that TiCl3 additive plays an important role in lowering activation energy for the hydrogenation reaction.
  C a ( B H 4 ) 2   C a H 2 + 2   B + 3   H 2
A proof-of-concept process for complete hydrogen release/uptake was shown for a ball-milled mixture CaB6 + 2 CaH2 + 8 wt.% TiCl3/Pd, but the reaction conditions were harsh (700-bar H2, 400–440 °C); however, almost complete recyclability with 9.6 wt.% hydrogen storage has been demonstrated [95]. A key role in this achievement is thought to be played by the deposition of TiCl3 on porous Pd, which is itself known for absorbing record amounts of hydrogen and giving interstitial hydrides, while other TM halides (RuCl3) led to polymorphs of Ca(BH4)2 and other unknown phases, confirming the alteration in the rehydrogenation pathway [95,398].
Nanoconfined Ca(BH4)2 in microporous carbon showed tangible improvements in the recyclability behavior of as-prepared nanocomposites [143]. Using carefully engineered chemically activated micro-mesoporous carbon as scaffolds (specific surface area SSA = 1780 m2/g, Vpore~1 cm3/g and 60% microporosity), Comanescu et al. obtained by the incipient wetness method Ca(BH4)2@MC-a nanocomposites with 36 wt.% borohydride loading (Figure 19).
These Ca(BH4)2@MC-a composites started the hydrogen release step at about 100 °C and managed to retain 2.4 wt.% hydrogen capacity even after 18 cycles of hydrogen release/uptake. The rehydrogenation of the desorbed composites needed rather mild conditions (25–40 atm H2, 6.5 h), confirming the important role of the nanoporous scaffold of choice (Figure 20) [143].
The result obtained for the composite Ca.(BH4)/2 = MC 650-a (containing 36 wt.% Ca(BH4)2) is an improvement over previously reported data on reversibility studies on Ca(BH4)2, confirming that about 68% of Ca(BH4)2 behaves reversibly [143].

7.6. Ammonia Borane NH3BH3

A special place is reserved to ammonia–borane, an interesting adduct with 19.6 wt.% hydrogen. NH3BH3 (ammonia–borane, AB) and ammonium borohydride (NH4BH4) have been reported from 1958, but NH4BH4 is far too unstable releasing H2 at even −40 °C to form AB [399]. NH3BH3 is a stable solid, water-sensitive, and susceptible to oxidation or attack by acidic reagents, having a high solubility in water (33.6 g/100 g H2O). The AB molecules present dipole–dipole interactions due to the dihydrogen bonding, i.e., interaction between a hydridic hydrogen (-NH3) and an acidic one (BH3) [400], and organizes in an orthorhombic phase at low temperature (<225 K) due to these interactions. While it is not a proper metal borohydride, NH3BH3 has the advantageous property of being RT-stable, exhibiting a two-step dehydrogenation at very reasonable temperatures: 114 °C (melting point, 6.5 wt.% H2 release, forming polyaminoboranes (NH2BH2)n), and 150 °C (remaining H2 release, formation of polyiminoborane (NHBH)n), with the downsides of irreversibility, slow kinetics, and formation of “inorganic benzene” as by-product during dehydrogenation (borazine, c-(NHBH)3) (Equation (98)). The N-B (nitrogen–boron) units are isoelectronic to C-C (carbon–carbon) moieties, and they can therefore be regarded as hydrocarbon analogues [145,401,402].
n   N H 3 B H 3 70 120   ° C , n H 2 ( N H 2 B H 2 ) n   120 200   ° C , n H 2 ( N H B H ) n > 500   ° C , n H 2   ( B N ) n
Mesoporous silica or carbon have served as scaffolds for NH3BH3 impregnation, and this nano-reduction strategy bypassed borazine formation [145,402,403]. Even NH3 release is suppressed when Li+-based dopants are used [402]. Some key thermodynamic aspects connected to NH3BH3 confinement in nanoporous substrates are summarized in Table 7. A common denominator is that using lower-size pores, the decomposition temperature is decreased due to the lowered activation energy Ea caused by nanoconfinement [404,405].
The porosity of the MOF scaffolds (IRMOF-1, IRMOF-10, UiO-66, UiO-67, MIL-52(Al)) for AB nanoconfinement (8.38 wt.%, 3.51 wt.%, 12.95 wt.%, 14.70 wt.%, and 11.26 wt.%, respectively, computed from SI data) has been investigated, showing that AB impregnation does not change MOF crystallinity. A low desorption temperature of 64°C was recorded for AB@UiO-66 (12.95 wt.% loading), which also featured the smallest pore size among investigated MOFs, of only 1.17 nm [411].
The reduction in decomposition temperature of AB@MOF was found to be linearly proportional to the reciprocal of AB size; in other words, the expectations are that a lower pore size nano-restriction would yield the best destabilization in ammonia borane nanocomposites [411].
Another advantage of utilizing SiO2-based scaffolds in AB nanoconfinement is that this does not lead to a reaction with the framework, as was the case for typical M(BH4)x. It is also apparent from Table 7 that 1D-structured mesosilica (MCM-41) are less attractive as scaffolds than 2D-mesostructured silica (SBA-15) of hexagonal pore arrangement, although the pore size might suggest otherwise (4 nm for MCM-41 vs. 7.5 nm for SBA-15). These conclusions could be extended to other borohydride@silica nanosystems as well, and could provide a starting point for further nanoscaffold engineering research. Nanoconfinement alters the dehydrogenation pathway in AB, as utilizing mesoporous carbons like CMK-3 manages to avoid toxic gas evolution; borazine or ammonia are not detected in that case. It can be hypothesized that nanoconfinement might help rehydrogenation by ammonia borane restructuring inside scaffold nanopores, and by lowering activation energies to improve the thermodynamics of hydrogen release/uptake. Some advances have been made by suggesting a methanolysis cycle, where [NH4][B(OCH3)4] and B(OCH3)3 are key intermediates [412]. Complete elucidation of a yet-unclear dehydrogenation mechanism, combined with collaborative theoretical/experimental research efforts, could pave the way to the ultimate hydrogen-storage material.

8. Conclusions and Outlook

It has been almost half a century since hydrogen was first considered a renewable energy carrier, aiming at replacing fossil fuel and building up towards a “hydrogen economy” goal. While not without its pitfalls, solid-state hydrogen storage exhibits obvious advantages compared to the physical storage of H2 gas: it is safer, requires less extreme conditions for storage and regeneration, and occupies less space.
Among various metal hydrides, metal borohydrides have inherited structural flexibility, accommodating both mixed cations and anion substitutions, and can be tailored towards multi-functional materials. Borohydrides can be engineered to form novel composites that reversibly store hydrogen chemically, reaching very good hydrogen storage capacities, in line with the DOE’s expectations for the near future. Recent studies conducted under high temperature and pressure (hydrogen, diborane) have shed new light on mechanistic aspects regarding formation and decomposition pathways of complex metal borohydrides.
Very recent progress on alkali metal borohydrides is directed towards controlling the kinetics of hydrogen production by utilizing nanoporous silica and carbon, Co-based catalysts, bimetallic catalysts (NiPt, NiCo9, RuPd, Co-Ru) and nano-catalyst composites, or harnessing the latent potential in alcoholysis (methanolysis, in particular) of said metal borohydrides. Kinetic studies on metal borohydrides currently target PEM fuel cell applications, and have shown that the hydrolysis mechanism has as a rate-limiting step the splitting of the H-OH bond. Transition metal nanoparticles (Ni, Pt, Zr, Co, Mn, Cu), mixed-valence oxides (Co3O4), transition metal complexes (of Ru3+, for instance), and MOFs have shown controlled hydrolysis of alkali metal borohydrides. Engineered catalysts (Mg95Ni5) have managed to suppress ammonia release from borohydride ammoniates by improving hydrogen release at very reasonable temperatures (70–80°C). Exploring new decomposition pathways of such ligand-stabilized borohydrides has also gained new momentum. Divalent metal borohydrides stabilized by neutral molecules show high ionic conductivity and a real promise for use in solid-state batteries. The rich boron–hydrogen chemistry, together with the open channels of closo-structured borohydrides (B10H102−, B12H122−), currently make research on closo-borate salts of alkali and alkali-earth metals a hot topic in ionic conductivity for solid-state electrolytes. Direct borohydride fuel cell (DBFC) and direct borohydride–hydrogen peroxide fuel cell (DBHPFC) systems research has expanded recently towards high-performance applications aimed at space exploration, by harnessing the power of nanoparticle catalysis (Ni, Cu, or bimetallic PdFe, PdAg, PdAu) to control the resultant power density. Various destabilization strategies of borohydride compounds have now included novel catalyst systems and complementary in-depth mechanistic insights. Nanostructuring of employed catalysts for hydrolysis reactions or de-/redrydrogenation studies has confirmed a positive effect by lowering activation energies associated with individual mechanistic steps. The porous polymorph of Mg(BH4)2 has gained attention recently by involvement in a core-shell, oxidation-resistant nanostructure (γ-Mg(BH4)2@MgCl2) to afford dehydrogenation onset as low as 100 °C. Theoretical computations on ternary borohydride materials (K2B2H8) predict superconducting properties at high pressures and around 140 K. Fundamental research is constantly expanding the class of structurally characterized metal borohydrides with new members (Th(BH4)4 is a new addition from 2021), while more cost-effective and high-yielding synthesis routes emerge—the synthesis of NaBH4 was reported by milling sodium tetraborate with Al and NaH as additive.
Meeting targets related to performance range, refueling time, car performance, and passenger space and safety are at the forefront of development of novel hydrogen storage materials and material-based storage system technologies. Mobile applications and vehicular applications in particular could receive a new breath of life when a proper hydride-based material is engineered by scientists, in line with safety, performance, and cost estimation goals.

Funding

This work was supported by the Romanian Ministry of Research and Innovation through the Core Program PN19-03 (contract no. PN21N/2019), PN-III-P2-2.1-PED-2019-4816 within PNCD III, and Project No. PN-III-P1-1.1-TE-2021-1657.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares that there are no commercial or financial relationships that could be considered as a potential conflict of interest.

References

  1. Yue, M.; Lambert, H.; Pahon, E.; Roche, R.; Jemei, S.; Hissel, D. Hydrogen energy systems: A critical review of technologies, applications, trends and challenges. Renew. Sustain. Energ. Rev. 2021, 146, 111180. [Google Scholar] [CrossRef]
  2. Abe, J.O.; Popoola, A.P.I.; Ajenifuja, E.; Popoola, O.M. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrogen Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  3. Lai, Q.W.; Sun, Y.H.; Wang, T.; Modi, P.; Cazorla, C.; Demirci, U.B.; Ares Fernandez, J.R.; Leardini, F.; Aguey-Zinsou, K.F. How to design hydrogen storage materials? Fundamentals, synthesis, and storage tanks. Adv. Sustain. Syst. 2019, 3, 1900043. [Google Scholar] [CrossRef]
  4. Ong, B.C.; Kamarudin, S.K.; Basri, S. Direct liquid fuel cells: A review. Int. J. Hydrogen Energy 2017, 42, 10142–10157. [Google Scholar] [CrossRef]
  5. Mohtadi, R.; Orimo, S.I. The renaissance of hydrides as energy materials. Nat. Rev. Mater. 2017, 2, 16091. [Google Scholar] [CrossRef]
  6. Veras, T.D.; Mozer, T.S.; dos Santos, D.D.R.M.; Cesar, A.D. Hydrogen: Trends, production and characterization of the main process worldwide. Int. J. Hydrogen Energy 2017, 42, 2018–2033. [Google Scholar] [CrossRef]
  7. He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen carriers. Nat. Rev. Mater. 2016, 1, 16067. [Google Scholar] [CrossRef]
  8. Lai, Q.W.; Paskevicius, M.; Sheppard, D.A.; Buckley, C.E.; Thornton, A.W.; Hill, M.R.; Gu, Q.F.; Mao, J.F.; Huang, Z.G.; 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]
  9. Huang, Y.J.; Cheng, Y.H.; Zhang, J.Y. A review of high density solid hydrogen storage materials by pyrolysis for promising mobile applications. Ind. Eng. Chem. Res. 2021, 60, 2737–2771. [Google Scholar] [CrossRef]
  10. Schüth, F.; Bogdanović, B.; Felderhoff, M. Light metal hydrides and complex hydrides for hydrogen storage. Chem. Commun. 2004, 20, 2249–2258. [Google Scholar] [CrossRef]
  11. Schneemann, A.; White, J.L.; Kang, S.Y.; Jeong, S.; Wan, L.F.; Cho, E.S.; Heo, T.W.; Prendergast, D.; Urban, J.J.; Wood, B.C.; et al. Nanostructured metal hydrides for hydrogen storage. Chem. Rev. 2018, 118, 10775–10839. [Google Scholar] [CrossRef]
  12. Grahame, A.; Aguey-Zinsou, K.F. properties and applications of metal (M) dodecahydro-closo-dodecaborates (Mn=1,2B12H12) and their implications for reversible hydrogen storage in the borohydrides. Inorganics 2018, 6, 106. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, K.; Pan, Z.X.; Yu, X.B. Metal B-N-H hydrogen-storage compound: Development and perspectives. J. Alloys Compd. 2019, 794, 303–324. [Google Scholar] [CrossRef]
  14. Kumar, R.; Karkamkar, A.; Bowden, M.; Autrey, T. Solid-state hydrogen rich boron-nitrogen compounds for energy storage. Chem. Soc. Rev. 2019, 48, 5350–5380. [Google Scholar] [CrossRef]
  15. Ley, M.B.; Jepsen, L.H.; Lee, Y.S.; Cho, Y.W.; von Colbe, J.M.B.; 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]
  16. Ouyang, L.Z.; Chen, K.; Jiang, J.; Yang, X.S.; Zhu, M. Hydrogen storage in light-metal based systems: A review. J. Alloys Compd. 2020, 829, 154597. [Google Scholar] [CrossRef]
  17. Singh, R.; Altaee, A.; Gautam, S. Nanomaterials in the advancement of hydrogen energy storage. Heliyon 2020, 6, e04487. [Google Scholar] [CrossRef]
  18. Santos, D.M.F.; Sequeira, C.A.C. Sodium borohydride as a fuel for the future. Renew. Sustain. Energ. Rev. 2011, 15, 3980–4001. [Google Scholar] [CrossRef]
  19. Available online: https://www.energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogen-storage-light-duty-vehicles (accessed on 31 January 2022).
  20. Available online: https://www.energy.gov/sites/default/files/2017/05/f34/fcto_targets_onboard_hydro_storage_explanation.pdf (accessed on 31 January 2022).
  21. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Cerny, 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]
  22. Abdelhamid, H.N. A review on hydrogen generation from the hydrolysis of sodium borohydride. Int. J. Hydrogen Energy 2021, 46, 726–765. [Google Scholar] [CrossRef]
  23. Liu, X.; Zhang, X.Y.; Li, D.S.; Zhang, S.Q.; Zhang, Q.C. Recent advances in the “on-off” approaches for on-demand liquid-phase hydrogen evolution. J. Mater. Chem. A 2021, 9, 18164–18174. [Google Scholar] [CrossRef]
  24. Nunes, H.X.; Silva, D.L.; Rangel, C.M.; Pinto, A.M.F.R. Rehydrogenation of sodium borates to close the NaBH4-H2 cycle: A review. Energies 2021, 14, 3567. [Google Scholar] [CrossRef]
  25. Gras, M.; Lota, G. Control of hydrogen release during borohydride electrooxidation with porous carbon materials. RSC Adv. 2021, 11, 15639–15655. [Google Scholar] [CrossRef]
  26. Yao, Q.L.; Ding, Y.Y.; Lu, Z.H. Noble-metal-free nanocatalysts for hydrogen generation from boron- and nitrogen-based hydrides. Inorg. Chem. Front. 2020, 7, 3837–3874. [Google Scholar] [CrossRef]
  27. Deng, J.F.; Chen, S.P.; Wu, X.J.; Zheng, J.; Li, X.G. Recent progress on materials for hydrogen generation via hydrolysis. J. Inorg. Mater. 2021, 36, 1–8. [Google Scholar] [CrossRef]
  28. Dionne, M.; Hao, S.; Gambarotta, S. Preparation and characterization of a new series of Cr(ll) hydroborates. Can. J. Chem. 1995, 73, 1126–1134. [Google Scholar] [CrossRef]
  29. Nöth, H.; Seitz, M. Preparation and characterization of chromium(ii) tetrahydroborate-tetrahydrofuran (1/2). J. Chem. Soc. Chem. Commun. 1976, 24, 1004a. [Google Scholar] [CrossRef]
  30. Soulie, J.P.; Renaudin, G.; Cerny, R.; Yvon, K. Lithium borohydride LiBH4 I. Crystal structure. J. Alloys Compd. 2002, 346, 200–205. [Google Scholar] [CrossRef]
  31. Puszkiel, J.; Gasnier, A.; Amica, G.; Gennari, F. Tuning LiBH4 for hydrogen storage: Destabilization, additive, and nanoconfinement approaches. Molecules 2020, 25, 163. [Google Scholar] [CrossRef] [Green Version]
  32. Skripov, A.V.; Soloninin, A.V.; Babanova, O.A.; Skoryunov, R.V. Anion and cation dynamics in polyhydroborate salts: NMR studies. Molecules 2020, 25, 2940. [Google Scholar] [CrossRef] [PubMed]
  33. Le, T.T.; Pistidda, C.; Nguyen, V.H.; Singh, P.; Raizada, P.; Klassen, T.; Dornheim, M. Nanoconfinement effects on hydrogen storage properties of MgH2 and LiBH4. Int. J. Hydrogen Energy 2021, 46, 23723–23736. [Google Scholar] [CrossRef]
  34. Skoryunov, R.V.; Babanova, O.A.; Soloninin, A.V.; Grinderslev, J.B.; Skripov, A.V.; Jensen, T.R. Dynamical properties of lithium borohydride-ammine composite LiBH4.NH3: A nuclear magnetic resonance study. J. Alloys Compd. 2022, 894, 162446. [Google Scholar] [CrossRef]
  35. Ye, J.K.; Xia, G.L.; Yu, X.B. In-situ constructed destabilization reaction of LiBH4 wrapped with graphene toward stable hydrogen storage reversibility. Mater. Today Energy 2021, 22, 100885. [Google Scholar] [CrossRef]
  36. Yang, G.Y.; Xie, C.; Li, Y.T.; Li, H.W.; Liu, D.M.; Chen, J.G.; Zhang, Q.G. Enhancement of the ionic conductivity of lithium borohydride by silica supports. Dalton Trans. 2021, 50, 15352–15358. [Google Scholar] [CrossRef] [PubMed]
  37. De Kort, L.M.; Gulino, V.; de Jongh, P.E.; Ngene, P. Ionic conductivity in complex metal hydride-based nanocomposite materials: The impact of nanostructuring and nanocomposite formation. J. Alloys Compd. 2022, 901, 163474. [Google Scholar] [CrossRef]
  38. Zhang, Q.; Gao, Z.Q.; Shi, X.M.; Zhang, C.; Liu, K.; Zhang, J.; Zhou, L.; Ma, C.J.; Du, Y.P. Recent advances on rare earths in solid lithium ion conductors. J. Rare Earths 2021, 39, 1–10. [Google Scholar] [CrossRef]
  39. Bogdanovic, B.; Schwickardi, M. Ti-doped alkali metal aluminium hydrides as potential novel reversible hydrogen storage materials. J. Alloys Compd. 1997, 253, 1–9. [Google Scholar] [CrossRef]
  40. Schlesinger, H.I.; Sanderson, R.T.; Burg, A.B. A volatile compound of aluminum, boron and hydrogen. J. Am. Chem. Soc. 1939, 61, 536. [Google Scholar] [CrossRef]
  41. Ravnsbæk, D.B.; 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]
  42. Hagemann, H.; Cerny, R. Synthetic approaches to inorganic borohydrides. Dalton Trans. 2010, 39, 6006–6012. [Google Scholar] [CrossRef]
  43. Wang, T.; Aguey-Zinsou, K.F. Synthesis of borohydride nanoparticles at room temperature by precipitation. Int. J. Hydrogen Energy 2021, 46, 24286–24292. [Google Scholar] [CrossRef]
  44. Suarez-Alcantara, K.; Garcia, J.R.T. Metal borohydrides beyond groups I and II: A review. Materials 2021, 14, 2561. [Google Scholar] [CrossRef] [PubMed]
  45. Cerny, R.; Brighi, M.; Murgia, F. The crystal chemistry of inorganic hydroborates. Chemistry 2020, 2, 53. [Google Scholar] [CrossRef]
  46. Gu, T.T.; Gu, J.; Zhang, Y.; Ren, H. Metal borohydride-based system for solid-state hydrogen storage. Prog. Chem. 2020, 32, 665–686. [Google Scholar] [CrossRef]
  47. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  48. Dornheim, M.; Eigen, N.; Barkhordarian, G.; Klassen, T.; Bormann, R. Tailoring hydrogen storage materials towards application. Adv. Eng. Mater 2006, 8, 377–385. [Google Scholar] [CrossRef]
  49. Eigen, N.; Keller, C.; Dornheim, M.; Klassen, T.; Bormann, R. Industrial production of light metal hydrides for hydrogen storage. Scr. Mater. 2007, 56, 847–851. [Google Scholar] [CrossRef] [Green Version]
  50. Hassan, I.A.; Ramadan, H.S.; Saleh, M.A.; Hissel, D. Hydrogen storage technologies for stationary and mobile applications: Review, analysis and perspectives. Renew. Sustain. Energ. Rev. 2021, 149, 111311. [Google Scholar] [CrossRef]
  51. Takacs, L. Self-sustaining reactions induced by ball milling. Prog. Mater. Sci. 2002, 47, 355–414. [Google Scholar] [CrossRef]
  52. Balema, V.P.; Wiench, J.W.; Pruski, M.; Pecharsky, V.K. Solvent-free mechanochemical synthesis of phosphonium salts. Chem. Commun. 2002, 7, 724–725. [Google Scholar] [CrossRef]
  53. Pistidda, C.; Santhosh, A.; Jerabek, P.; Shang, Y.Y.; Girella, A.; Milanese, C.; Dore, M.; Garroni, S.; Bordignon, S.; Chierotti, M.R.; et al. Hydrogenation via a low energy mechanochemical approach: The MgB2 case. J. Phys. Energy 2021, 3, 044001. [Google Scholar] [CrossRef]
  54. Balema, V.P. Mechanical processing in hydrogen storage research and development. Mater. Matter. 2007, 2, 16–18. Available online: https://www.sigmaaldrich.com/deepweb/assets/sigmaaldrich/marketing/global/documents/190/337/material-matters-v2n2.pdf (accessed on 31 January 2022). [CrossRef]
  55. Bateni, A.; Scherpe, S.; Acar, S.; Somer, M. Novel approach for synthesis of Magnesium Borohydride, Mg(BH4)2. Energy Procedia 2012, 29, 26–33. [Google Scholar] [CrossRef]
  56. Xu, D.Y.; Zhang, Y.; Guo, Q.J. Research progress on catalysts for hydrogen generation through sodium borohydride alcoholysis. Int. J. Hydrogen Energy 2022, 47, 5929–5946. [Google Scholar] [CrossRef]
  57. Ouyang, L.Z.; Jiang, J.; Chen, K.; Zhu, M.; Liu, Z.W. Hydrogen production via hydrolysis and alcoholysis of light metal-based materials: A review. Nano-Micro Lett. 2021, 13, 134. [Google Scholar] [CrossRef] [PubMed]
  58. Al-Msrhad, T.M.H.; Devrim, Y.; Uzundurukan, A.; Budak, Y. Investigation of hydrogen production from sodium borohydride by carbon nano tube-graphene supported PdRu bimetallic catalyst for PEM fuel cell application. Int. J. Energy Res. 2021. [Google Scholar] [CrossRef]
  59. Faghihi, M.; Akbarbandari, F.; Zabihi, M.; Pazouki, M. Synthesis and characterization of the magnetic supported metal-organic framework catalysts (CuCoBTC@MAC and CuBTC@MAC) for the hydrogen production from sodium borohydride. Mater. Chem. Phys. 2021, 267, 124599. [Google Scholar] [CrossRef]
  60. Schlesinger, H.I.; Brown, H.C.; Hyde, E.K. The preparation of other borohydrides by metathetical reactions utilizing the alkali metal borohydrides. J. Am. Chem. Soc. 1953, 75, 209–213. [Google Scholar] [CrossRef]
  61. Schlesinger, H.I.; Brown, H.C.; Hoekstra, H.R.; Rapp, L.R. Reactions of diborane with alkali metal hydrides and their addition compounds. New syntheses of borohydrides. Sodium and potassium borohydrides. J. Am. Chem. Soc. 1953, 75, 199–204. [Google Scholar] [CrossRef]
  62. Schlesinger, H.I.; Brown, H.C.; Finholt, A.E. The preparation of sodium borohydride by the high temperature reaction of sodium hydride with borate esters. J. Am. Chem. Soc. 1953, 75, 205–209. [Google Scholar] [CrossRef]
  63. Schlesinger, H.I.; Brown, H.C. Uranium(IV) Borohydride. J. Am. Chem. Soc. 1953, 75, 219–221. [Google Scholar] [CrossRef]
  64. Payandeh Gharibdoust, S.; Heere, M.; Nervi, C.; Sørby, M.H.; Hauback, B.C.; Jensen, T.R. Synthesis, structure, and polymorphic transitions of praseodymium(III) and neodymium(III) borohydride, Pr(BH4)3 and Nd(BH4)3. Dalt. Trans. 2018, 47, 8307–8319. [Google Scholar] [CrossRef] [PubMed]
  65. Frommen, C.; Aliouane, N.; Deledda, S.; Fonneløp, J.E.; Grove, H.; Lieutenant, K.; Llamas-Jansa, I.; Sartori, S.; Sørby, M.H.; Hauback, B.C. Crystal structure, polymorphism, and thermal properties of yttrium borohydride Y(BH4)3. J. Alloys Compd. 2010, 496, 710–716. [Google Scholar] [CrossRef]
  66. Yang, C.H.; Tsai, W.T.; Chang, J.K. Hydrogen desorption behavior of vanadium borohydride synthesized by modified mechanochemical process. Int. J. Hydrogen Energy 2011, 36, 4993–4999. [Google Scholar] [CrossRef]
  67. Ravnsbæk, D.B.; Filinchuk, Y.; Cerny, R.; Ley, M.B.; Haase, D.; Jakobsen, H.J.; Skibsted, J.; Jensen, T.R. Thermal polymorphism and decomposition of Y(BH4)3. Inorg. Chem. 2010, 49, 3801–3809. [Google Scholar] [CrossRef] [Green Version]
  68. Sato, T.; Miwa, K.; Nakamori, Y.; Ohoyama, K.; Li, H.; Noritake, T.; Aoki, M.; Towata, S.I.; Orimo, S.I. Experimental and computational studies on solvent-free rare-earth metal borohydrides R(BH4)3 (R=Y, Dy, and Gd). Phys. Rev. B 2008, 77, 104114. [Google Scholar] [CrossRef]
  69. Wegner, W.; Jaron, T.; Grochala, W. Polymorphism and hydrogen discharge from holmium borohydride, Ho(BH4)3, and KHo(BH4)4. Int. J. Hydrogen Energy 2014, 39, 20024–20030. [Google Scholar] [CrossRef]
  70. Wegner, W.; Jaron, T.; Grochala, W. Preparation of a series of lanthanide borohydrides and their thermal decomposition to refractory lanthanide borides. J. Alloys Compd. 2018, 744, 57–63. [Google Scholar] [CrossRef]
  71. Andrade-Gamboa, J.; Puszkiel, J.A.; Fernández-Albanesi, L.; Gennari, F.C. A novel polymorph of gadolinium tetrahydroborate produced by mechanical milling. Int. J. Hydrogen Energy 2010, 35, 10324–10328. [Google Scholar] [CrossRef]
  72. Ley, M.B.; Paskevicius, M.; Schouwink, P.; Richter, B.; Sheppard, D.A.; Buckley, C.E.; Jensen, T.R. Novel solvates M(BH4)3S(CH3)2 and properties of halide-free M(BH4)3 (M = Y or Gd). Dalton Trans. 2014, 43, 13333–13342. [Google Scholar] [CrossRef] [Green Version]
  73. Grinderslev, J.B.; Møller, K.T.; Bremholm, M.; Jensen, T.R. Trends in synthesis, crystal structure, and thermal and magnetic properties of rare-earth metal borohydrides. Inorg. Chem. 2019, 58, 5503–5517. [Google Scholar] [CrossRef] [PubMed]
  74. Jaron, T.; Grochala, W. Y(BH4)3—An old–new ternary hydrogen store akalearning from a multitude of failures. Dalton Trans. 2010, 39, 160–166. [Google Scholar] [CrossRef] [PubMed]
  75. Lee, Y.-S.; Shim, J.-H.; Cho, Y.W. Polymorphism and thermodynamics of Y(BH4)3 from first principles. J. Phys. Chem. C 2010, 114, 12833–12837. [Google Scholar] [CrossRef]
  76. Yan, Y.; Li, H.W.; Sato, T.; Umeda, N.; Miwa, K.; Towata, S.; Orimo, S. Dehydriding and rehydriding properties of yttrium borohydride Y(BH4)3 prepared by liquid-phase synthesis. Int. J. Hydrogen Energy 2009, 34, 5732–5736. [Google Scholar] [CrossRef]
  77. Haaland, A.; Shorokhov, D.J.; Tutukin, A.V.; Volden, H.V.; Swang, O.; McGrady, G.S.; Kaltsoyannis, N.; Downs, A.J.; Tang, C.Y.; Turner, J.F.C. Molecular structures of two metal tetrakis(tetrahydroborates), Zr(BH4)4 and U(BH4)4: Equilibrium conformations and barriers to internal rotation of the triply bridging BH4 groups. Inorg. Chem. 2002, 41, 6646–6655. [Google Scholar] [CrossRef] [PubMed]
  78. Banks, R.H.; Edelstein, N.M. Synthesis and Characterization of Protactinium(IV), Neptunium(IV), and Plutonium (IV) Borohydrides. In Lanthanide and Actinide Chemistry and Spectroscopy; American Chemical Society Publications: Washington, DC, USA, 1980; pp. 331–348. [Google Scholar] [CrossRef]
  79. Rajnak, K.; Gamp, E.; Shinomoto, R.; Edelstein, N. Optical and magnetic properties of uranium borohydride and tetrakismethylborohydride. J. Chem. Phys. 1984, 80, 5942–5950. [Google Scholar] [CrossRef] [Green Version]
  80. Banks, R.H.; Edelstein, N.M.; Rietz, R.R.; Templeton, D.H.; Zalkin, A. Preparation and properties of the actinide borohydrides: Protactinium(IV), neptunium(IV), and plutonium(IV) borohydrides. J. Am. Chem. Soc. 1978, 100, 1957–1958. [Google Scholar] [CrossRef]
  81. Desch, C.H. Henry Louis Le Chatelier, 1850–1936; Royal Society: London, UK, 1938; Volume 2, pp. 250–259. [Google Scholar] [CrossRef]
  82. Atkins, P.; de Paula, J. Physical Chemistry, 8th ed.; Chemical Equilibrium; Oxford University Press: Oxford, UK, 2006; ISBN 0716787598. [Google Scholar]
  83. Aldridge, S.; Blake, A.J.; Downs, A.J.; Gould, R.O.; Parsons, S.; Pulham, C.R. Some tetrahydroborate derivatives of aluminium: Crystal structures of dimethylaluminium tetrahydroborate and the α and β phases of aluminium tris(tetrahydroborate) at low temperature. J. Chem. Soc. Dalton Trans. 1997, 6, 1007–1012. [Google Scholar] [CrossRef]
  84. Downs, A.J.; Thomas, P.D.P. Gallium borohydrides: The synthesis and properties of HGa(BH4)2. J. Chem. Soc. Chem. Commun. 1976, 20, 825–827. [Google Scholar] [CrossRef]
  85. Pulham, C.R.; Brain, P.T.; Downs, A.J.; Rankin, D.W.H.; Robertson, H.E. Gallaborane, H2Ga(µ-H)2BH2: Synthesis, properties, and structure of the gaseous molecule as determined by electron diffraction. J. Chem. Soc. Chem. Commun. 1990, 2, 177–178. [Google Scholar] [CrossRef]
  86. Wiberg, E.; Nöth, H. Notizen: Über Wasserstoff-Verbindungen des Indiums. Z. Naturforsch. B 1957, 12, 59–60. [Google Scholar] [CrossRef]
  87. Wiberg, E.; Dittmann, O.; Nöth, H.; Schmidt, M. Notizen: Über Wasserstoff-Verbindungen des Thalliums. V. Zur Kenntnis eines Thallium(I)-boranats TlBH4 und Thallium(I)-alanats TlAlH4. Z. Naturforsch. B 1957, 12, 62–63. [Google Scholar] [CrossRef]
  88. Wiberg, E.; Nöth, H. Notizen: Über Wasserstoff-Verbindungen des Thalliums. VI. Zur Kenntnis eines Thallium(III)-boranats TlCl(BH4)2. Z. Naturforsch. B 1957, 12, 63–65. [Google Scholar]
  89. Mosegaard, L.; Møller, B.; Jørgensen, J.; 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] [Green Version]
  90. Goerrig, D. Verfahren zur Herstellung von Boranaten. German Patent No. DE 000001077644 A, 17 March 1960. [Google Scholar]
  91. Friedrichs, O.; Buchter, F.; Zwicky, C.N.; Borgschulte, A.; Remhof, A.; Mauron, P.; Bielmann, M.; Zuttel, A. Direct synthesis of Li[BH4] and Li[BD4] from the elements. Acta Mater. 2008, 56, 949–954. [Google Scholar] [CrossRef]
  92. Barkhodarian, G.; Klassen, T.; Dornheim, M.; Bormann, R. Unexpected kinetic effect of MgB2 in reactive hydride composites containing complex borohydrides. J. Alloys Compd. 2007, 440, L18–L21. [Google Scholar] [CrossRef]
  93. Friedrichs, O.; Remhof, A.; Borgschulte, A.; Buchter, F.; Orimo, S.I.; Züttel, A. Breaking the passivation—the road to a solvent free borohydride synthesis. Phys. Chem. Chem. Phys. 2010, 12, 10919–10922. [Google Scholar] [CrossRef] [PubMed]
  94. Barkhordarian, G.; Jensen, T.R.; Doppiu, S.; Bosenberg, U.; Borgschulte, A.; Gremaud, R.; Cerenius, Y.; Dornheim, M.; Klassen, T.; Bormann, R. Formation of Ca(BH4)2 from Hydrogenation of CaH2+MgB2 Composite. J. Phys. Chem. C 2008, 112, 2743–2749. [Google Scholar] [CrossRef]
  95. Ronnebro, E.; Majzoub, E.H. Calcium borohydride for hydrogen storage:  Catalysis and reversibility. J. Phys. Chem. B 2007, 111, 12045–12047. [Google Scholar] [CrossRef]
  96. Yan, Y.; Rentsch, D.; Remhof, A. Controllable decomposition of Ca(BH4)2 for reversible hydrogen storage. Phys. Chem. Chem. Phys. 2017, 19, 7788–7792. [Google Scholar] [CrossRef]
  97. Pinkerton, F.E.; Meyer, M.S. Reversible hydrogen storage in the lithium borohydride—Calcium hydride coupled system. J. Alloys Compd. 2008, 464, L1–L4. [Google Scholar] [CrossRef]
  98. Sahle, C.J.; Sternemann, C.; Giacobbe, C.; Yan, Y.; Weis, C.; Harder, M.; Forov, Y.; Spiekermann, G.; Tolan, M.; Krisch, M.; et al. Formation of CaB6 in the thermal decomposition of the hydrogen storage material Ca(BH4)2. Phys. Chem. Chem. Phys. 2016, 18, 19866–19872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Castilla-Martinez, C.A.; Moury, R.; Ould-Amara, S.; Demirci, U.B. Destabilization of boron-based compounds for hydrogen storage in the solid-state: Recent advances. Energies 2021, 14, 7003. [Google Scholar] [CrossRef]
  100. Rongeat, C.; D’Anna, V.; Hagemann, H.; Borgschulte, A.; Zuttel, A.; Schultz, L.; Gutfleisch, O. Effect of additives on the synthesis and reversibility of Ca(BH4)2. J. Alloys Compd. 2010, 493, 281–287. [Google Scholar] [CrossRef]
  101. Nakamori, Y.; Li, H.W.; Kikuchi, K.; Aoki, M.; Miwa, K.; Towata, S.; Orimo, S. Thermodynamical stabilities of metal-borohydrides. J. Alloys Compd. 2007, 446–447, 296–300. [Google Scholar] [CrossRef]
  102. Schlesinger, H.I.; Brown, H.C. Metallo Borohydrides. III. Lithium borohydride. J. Am. Chem. Soc. 1940, 62, 3429–3435. [Google Scholar] [CrossRef]
  103. Kollonitsch, J.; Fuchs, O. Preparation of aluminium borohydride and its applications in organic reductions. Nature 1955, 176, 1081. [Google Scholar] [CrossRef]
  104. Brown, H.C.; Choi, Y.M.; Narasimhan, S. Addition compounds of alkali metal hydrides. 22. Convenient procedures for the preparation of lithium borohydride from sodium borohydride and borane-dimethyl sulfide in simple ether solvents. Inorg. Chem. 1982, 21, 3657–3661. [Google Scholar] [CrossRef]
  105. Rittmeyer, P.; Wietelmann, U. Ullmann’s Encyclopaedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2000. [Google Scholar]
  106. Cerny, R.; Filinchuk, Y.; Hagemann, H.; Yvon, K. Magnesium borohydride: Synthesis and crystal structure. Angew. Chem. Int. Ed. 2007, 46, 5765–5767. [Google Scholar] [CrossRef] [Green Version]
  107. Soloveichik, G.L.; Andrus, M.; Gao, Y.; Zhao, J.-C.; Kniajanski, S. Magnesium borohydride as a hydrogen storage material: Synthesis of unsolvated Mg(BH4)2. Int. J. Hydrogen Energy 2009, 34, 2144–2152. [Google Scholar] [CrossRef]
  108. Huang, H.X.; Liu, B.G.; Lv, Y.J.; Lv, W.; Yuan, J.G.; Wu, Y. Double regulation of Mg95Ni5 hydride in suppressing ammonia and promoting hydrogen evolution for Mg(BH4)2.2NH3. J. Alloys Compd. 2022, 901, 163468. [Google Scholar] [CrossRef]
  109. 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]
  110. Zanella, P.; Crociani, L.; Masciocchi, N.; Giunchi, G. Facile high-yield synthesis of pure, crystalline Mg(BH4)2. Inorg. Chem. 2007, 46, 9039–9041. [Google Scholar] [CrossRef] [PubMed]
  111. Wiberg, E.; Hartwimmer, R.Z. Zur Kenntnis von Erdalkaliboranaten Me[BH4]2 III. Synthese aus Erdalkalihydriden und Diboran. Z. Naturforsch. B Chem. Sci. 1955, 10, 295–296. [Google Scholar] [CrossRef]
  112. Wiberg, E.; Noth, H.; Hartwimmer, R.Z. Zur Kenntnis von Erdalkaliboranaten Me[BH4]2 I. Darstellung aus Erdalkali-tetramethoxoboraten und Diboran. Z. Naturforsch. B Chem. Sci. 1955, 10, 292–294. [Google Scholar] [CrossRef]
  113. Wiberg, E. Zur kenntnis eines kupfer-bor-wasserstoffs CuBH4. Z. Naturforsch. B 1952, 7b, 582. [Google Scholar]
  114. Wiberg, E.; Henle, W. Notizen: Zur Kenntnis eines Silber-bor-wasserstoffs AgBH4. Z. Naturforsch. B 1952, 7, 575–576. [Google Scholar] [CrossRef]
  115. Wiberg, E. Neuere Ergebnisse der praparativen Hydrid-Forschung. Angew. Chem. 1953, 65, 16–33. [Google Scholar] [CrossRef]
  116. Musaev, D.G.; Morokuma, K. Does the Tetrahydroborate Species AuBH4 Exist? Ab Initio MO Study of the Structure and Stability of CuBH4, AgBH4, and AuBH4. Organometallics 1995, 14, 3327–3334. [Google Scholar] [CrossRef]
  117. Jain, I.P.; Jain, P.; Jain, A. Novel hydrogen storage materials: A review of lightweight complex hydrides. J. Alloys Compd. 2010, 503, 303–339. [Google Scholar] [CrossRef]
  118. Riktor, M.D.; Sørby, M.H.; Chlopek, K.; Fichtner, M.; Buchter, F.; Zuttel, A.; Hauback, B.C. In situ synchrotron diffraction studies of phase transitions and thermal decomposition of Mg(BH4)2 and Ca(BH4)2. J. Mater. Chem. 2007, 17, 4939–4942. [Google Scholar] [CrossRef]
  119. Jeon, E.; Cho, Y.W. Mechanochemical synthesis and thermal decomposition of zinc borohydride. J. Alloys Compd. 2006, 422, 273–275. [Google Scholar] [CrossRef]
  120. Srinivasan, S.; Escobar, D.; Jurczyk, M.; Goswami, Y.; Stefanakos, E. Nanocatalyst doping of Zn(BH4)2 for on-board hydrogen storage. J. Alloys Compd. 2008, 462, 294–302. [Google Scholar] [CrossRef]
  121. Wiberg, E.; Henle, W. Notizen: Zur Kenntnis eines ätherlöslichen Zink-bor-wasser-stoffs Zn(BH4)2. Z. Naturforsch. B 1952, 7, 579–580. [Google Scholar] [CrossRef] [Green Version]
  122. Nöth, H.; Wiberg, E.; Winter, L.P. Boranate und Boranato-metallate. I. Zur Kenntnis von Solvaten des Zinkboranats. Z. Anorg. Allg. Chem. 1969, 370, 209–223. [Google Scholar] [CrossRef]
  123. Banus, M.D.; Bragdon, R.W.; Hinckley, A.A. Potassium, rubidium and cesium borohydrides. J. Am. Chem. Soc. 1954, 76, 3848–3849. [Google Scholar] [CrossRef]
  124. Hagemann, H.; Gomes, S.; Renaudin, G.; Yvon, K. Raman studies of reorientation motions of [BH4] anions in alkali borohydrides. J. Alloys Compd. 2004, 363, 129–132. [Google Scholar] [CrossRef] [Green Version]
  125. Friedrichs, O.; Borgschulte, A.; Kato, S.; Buchter, F.; Gremaud, R.; Remhof, A.; Zuttel, A. Low-temperature synthesis of LiBH4 by gas-solid reaction. Chem. Eur. J. 2009, 15, 5531–5534. [Google Scholar] [CrossRef]
  126. James, B.D.; Wallbridge, M.G.H. Metal tetrahydroborates. Prog. Inorg. Chem. 1970, 11, 99–231. [Google Scholar]
  127. 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]
  128. Berube, V.; Radtke, G.; Dresselhaus, M.; Chen, G. Size effects on the hydrogen storage properties of nanostructured metal hydrides: A review. Int. J. Energy Res 2007, 31, 637–663. [Google Scholar] [CrossRef]
  129. Gertsman, V.; Birringer, R. On the room-temperature grain growth in nanocrystalline copper. Scr. Metall. Mater. 1994, 30, 577–581. [Google Scholar] [CrossRef]
  130. Vigeholm, B.; Kjoller, J.; Larsen, B.; Schroder Pedersen, A. Hydrogen sorption performance of pure magnesium during continued cycling. Int. J. Hydrogen Energy 1983, 8, 809–817. [Google Scholar] [CrossRef]
  131. Berlouis, L.; Cabrera, E.; Hall-Barientos, E.; Hall, P.; Dodd, S.; Morris, S.; Imam, M. Thermal analysis investigation of hydriding properties of nanocrystalline Mg–Ni- and Mg–Fe-based alloys prepared by high-energy ball milling. J. Mater. Res. 2001, 16, 45–57. [Google Scholar] [CrossRef]
  132. Nielsen, T.K.; Besenbacher, F.; Jensen, T.R. Nanoconfined hydrides for energy storage. Nanoscale 2011, 3, 2086–2098. [Google Scholar] [CrossRef]
  133. Nielsen, T.K.; Manickam, K.; Hirscher, M.; Besenbacher, F.; Jensen, T.R. Confinement of MgH2 Nanoclusters within Nanoporous Aerogel Scaffold Materials. ACS Nano 2009, 3, 3521–3528. [Google Scholar] [CrossRef]
  134. Zhang, S.; Gross, A.F.; Atta, S.L.; Lopez, M.; Liu, P.; Ahn, C.C.; Vajo, J.J.; Jensen, C.M. The synthesis and hydrogen storage properties of a MgH2 incorporated carbon aerogel scaffold. Nanotechnology 2009, 20, 204027. [Google Scholar] [CrossRef] [PubMed]
  135. de Gennes, P.G. Wetting: Statics and dynamics. Rev. Mod. Phys. 1985, 57, 827–863. [Google Scholar] [CrossRef]
  136. Dujardin, E.; Ebbesen, T.W.; Hiura, H.; Tanigaki, K. Capillarity and wetting of carbon nanotubes. Science 1994, 265, 1850–1852. [Google Scholar] [CrossRef] [PubMed]
  137. Gao, J.; Adelhelm, P.; Verkuijlen, M.H.W.; Rongeat, C.; Herrich, M.; van Bentum, P.J.M.; Gutfleisch, O.; Kentgens, A.P.M.; de Jong, K.P.; de Jongh, P.E. Confinement of NaAlH4 in nanoporous carbon: Impact on h2 release, reversibility, and thermodynamics. J. Phys. Chem. C. 2010, 114, 4675–4682. [Google Scholar] [CrossRef]
  138. Dilts, J.A.; Ashby, E.C. A study of the thermal decomposition of complex metal hydrides. Inorg. Chem. 1972, 11, 1230–1236. [Google Scholar] [CrossRef]
  139. Bogdanović, B.; Brand, R.A.; Marjanovic, 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]
  140. Verkuijlen, M.H.W.; Gao, J.; Adelhelm, P.; van Bentum, P.J.M.; de Jongh, P.E.; Kentgens, A.P.M. Solid-state NMR studies of the local structure of NaAlH4/C nanocomposites at different stages of hydrogen desorption and rehydrogenation. J. Phys. Chem. C 2010, 114, 4683–4692. [Google Scholar] [CrossRef]
  141. Lohstroh, W.; Roth, A.; Hahn, H.; Fichtner, M. Thermodynamic effects in nanoscale NaAlH4. ChemPhysChem 2010, 11, 789–792. [Google Scholar] [CrossRef] [PubMed]
  142. Stephens, R.D.; Gross, A.F.; Atta, S.L.; Vajo, J.J.; Pinkerton, F.E. The kinetic enhancement of hydrogen cycling in NaAlH4 by melt infusion into nanoporous carbon aerogel. Nanotechnology 2009, 20, 204018. [Google Scholar] [CrossRef]
  143. Comanescu, C.; Capurso, G.; Maddalena, A. Nanoconfinement in activated mesoporous carbon of calcium borohydride for improved reversible hydrogen storage. Nanotechnology 2012, 23, 385401. [Google Scholar] [CrossRef]
  144. Gutowski, M.; Autrey, T. Computational studies of boron/nitrogen and aluminum/nitrogen compounds for chemical hydrogen storage. Prepr. Pap. Am. Chem. Soc. Div. Fuel Chem. 2004, 49, 275–276. Available online: http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.459.4955&rep=rep1&type=pdf (accessed on 31 January 2022).
  145. Gutowska, A.; Li, L.; Shin, Y.; Wang, C.M.; Li, X.S.; Linehan, J.C.; Smith, R.S.; Kay, B.D.; Schmid, B.; Shaw, W.; et al. Nanoscaffold mediates hydrogen release and the reactivity of ammonia borane. Angew. Chem. Int. Ed. 2005, 44, 3578–3582. [Google Scholar] [CrossRef] [PubMed]
  146. Filinchuk, Y.; Ronnebro, E.; Chandra, D. Crystal structures and phase transformations in Ca(BH4)2. Acta Mater. 2009, 57, 732–738. [Google Scholar] [CrossRef]
  147. Ravnsbæk, D.B.; Nickels, E.A.; Cerny, 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]
  148. Sharma, M.; Didelot, E.; Spyratou, A.; Daku, L.M.L.; Cerny, 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]
  149. Ravnsbæk, D.B.; Sørensen, L.H.; Filinchuk, Y.; Besenbacher, F.; Jensen, T.R. Screening of Metal Borohydrides by Mechanochemistry and Diffraction. Angew. Chem. Int. Ed. 2012, 51, 3582–3586. [Google Scholar] [CrossRef]
  150. Wiberg, E.; Henle, W. Zur Kenntnis eines Cadmium-bor-wasserstoffs Cd(BH4)2. Z. Naturforschg. B 1952, 7, 582. [Google Scholar] [CrossRef] [Green Version]
  151. Nöth, H.; Winter, L.P. Metallboranate und Boranato-metallate. V. Solvate des Cadmiumboranats und Boranatocadmate. Z. Anorg. Allg. Chem. 1972, 389, 225–234. [Google Scholar] [CrossRef]
  152. Stockmayer, W.H.; Rice, D.W.; Stephenson, C.C. Thermodynamic properties of sodium borohydride and aqueous borohydride ion. J. Am. Chem. Soc. 1955, 77, 1980–1983. [Google Scholar] [CrossRef]
  153. Filinchuk, Y.; Hagemann, H. Structure and properties of NaBH4·2H2O and NaBH4. Eur. J. Inorg. Chem. 2008, 20, 3127–3133. [Google Scholar] [CrossRef] [Green Version]
  154. Hagemann, H.; D’Anna, V.; Carbonniere, P.; Bardaji, E.G.; Fichtner, M. Synthesis and characterization of NaBD3H, a potential structural probe for hydrogen storage materials. J. Phys. Chem. A 2009, 113, 13932–13936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Davis, R.E.; Kenson, R.E. Boron Hydrides. XII. The synthesis and infrared spectra of NaBH3D and NaBD3H. Proc. Indiana Acad. Sci. 1966, 76, 236–240. [Google Scholar]
  156. Chłopek, K.; Frommen, C.; Leon, A.; Zabara, O.; Fichtner, M. Synthesis and properties of magnesium tetrahydroborate, Mg(BH4)2. J. Mater. Chem. 2007, 17, 3496–3503. [Google Scholar] [CrossRef] [Green Version]
  157. Buchter, F.; Łodziana, Z.; Remhof, A.; Friedrichs, O.; Borgschulte, A.; Mauron, P.; Zuttel, A.; Sheptyakov, D.; Barkhordarian, G.; Bormann, R.; et al. Structure of Ca(BD4)2 β-phase from combined neutron and synchrotron x-ray powder diffraction data and density functional calculations. J. Phys. Chem. B 2008, 112, 8042–8048. [Google Scholar] [CrossRef] [PubMed]
  158. 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]
  159. Ehrlich, R.; Rice, G. The chemistry of alane. XII. The lithium tetrahydroalanate-triethylamine complex. Inorg. Chem. 1966, 5, 1284–1286. [Google Scholar] [CrossRef]
  160. Plešek, J.; Heřmánek, S. Chemistry of boranes. IV. On preparation, properties, and behavior towards Lewis bases of magnesium borohydride. Collect. Czech. Chem. Commun. 1966, 31, 3845–3858. [Google Scholar] [CrossRef]
  161. Brown, H.C.; Yoon, N.M. Selective reductions. X. reaction of aluminum hydride with selected organic compounds containing representative functional groups. Comparison of the reducing characteristics of lithium aluminum hydride and its derivatives. J. Am. Chem. Soc. 1966, 88, 1464–1472. [Google Scholar] [CrossRef]
  162. Ruff, J.K.; Hawthorne, M.F. The amine complexes of aluminum hydride. II. J. Am. Chem. Soc. 1961, 83, 535–538. [Google Scholar] [CrossRef]
  163. Kuzdrowska, M.; Annunziata, L.; Marks, S.; Schmid, M.; Jaffredo, C.G.; Roesky, P.W.; Guillaume, S.M.; Maron, L. Organometallic calcium and strontium borohydrides as initiators for the polymerization of epsilon-caprolactone and L-lactide: Combined experimental and computational investigations. Dalton Trans. 2013, 42, 9352–9360. [Google Scholar] [CrossRef] [PubMed]
  164. Yan, Y.; Remhof, A.; Hwang, S.-J.; Li, H.-W.; Mauron, P.; Orimo, S.-I.; Zuttel, A. Pressure and temperature dependence of the decomposition pathway of LiBH4. Phys. Chem. Chem. Phys. 2012, 14, 6514–6519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Abrahams, S.C.; Kalnajs, J. The lattice constants of the alkali borohydrides and the low-temperature phase of sodium borohydride. J. Chem. Phys. 1954, 22, 434–436. [Google Scholar] [CrossRef]
  166. Kumar, R.S.; Kim, E.; Cornelius, A.L. Structural Phase Transitions in the Potential Hydrogen Storage Compound KBH4 under Compression. J. Phys. Chem. C 2008, 112, 8452–8457. [Google Scholar] [CrossRef]
  167. Filinchuk, Y.; Talyzin, A.V.; Hagemann, H.; Dmitriev, V.; Chernyshov, D.; Sundqvist, B. Cation Size and Anion Anisotropy in Structural Chemistry of Metal Borohydrides. The Peculiar Pressure Evolution of RbBH4. Inorg. Chem. 2010, 49, 5285–5292. [Google Scholar] [CrossRef]
  168. Babanova, O.A.; Soloninin, A.V.; Stepanov, A.P.; Skripov, A.V.; Filinchuk, Y. Structural and Dynamical Properties of NaBH4 and KBH4: NMR and Synchrotron X-ray Diffraction Studies. J. Phys. Chem. C 2010, 114, 3712–3718. [Google Scholar] [CrossRef]
  169. Filinchuk, Y.; Talyzin, A.V.; Chernyshov, D.; Dmitriev, V. High-pressure phase of NaBH4: Crystal structure from synchrotron powder diffraction data. Phys. Rev. B Condens. Matter Mater. Phys. 2007, 76, 092104. [Google Scholar] [CrossRef] [Green Version]
  170. Marynick, D.S.; Lipscomb, W.N. Crystal structure of beryllium borohydride. Inorg. Chem. 1972, 11, 820–823. [Google Scholar] [CrossRef]
  171. Filinchuk, Y.; Cerny, 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] [Green Version]
  172. 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. Sect. B Struct. Sci. 2007, 63, 561–568. [Google Scholar] [CrossRef] [PubMed]
  173. 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]
  174. Cerny, R.; Penin, N.; Hagemann, H.; Filinchuk, Y. The first crystallographic and spectroscopic characterization of a 3d-metal borohydride: Mn(BH4)2. J. Phys. Chem. C 2009, 113, 9003–9007. [Google Scholar] [CrossRef] [Green Version]
  175. Tumanov, N.A.; Roedern, E.; Łodziana, Z.; Nielsen, D.B.; Jensen, T.R.; Talyzin, A.V.; Černy, R.; Chernyshov, D.; Dmitriev, V.; Palasyuk, T.; et al. High-Pressure study of Mn(BH4)2 reveals a stable polymorph with high hydrogen density. Chem. Mater. 2016, 28, 274–283. [Google Scholar] [CrossRef]
  176. Łodziana, Z. Multivalent metal tetrahydroborides of Al, Sc, Y., Ti, and Zr. Phys. Rev. B 2010, 81, 144108. [Google Scholar] [CrossRef]
  177. Dain, C.J.; Downs, A.J.; Goode, M.J.; Evans, D.G.; Nicholls, K.T.; Rankin, D.W.H.; Robertson, H.E. Molecular structure of gaseous titanium tris(tetrahydroborate), Ti(BH4)3: Experimental determination by electron diffraction and molecular orbital analysis of some Ti(BH4)3 derivatives. J. Chem. Soc. Dalt. Trans. 1991, 967–977. [Google Scholar] [CrossRef]
  178. Franz, K.; Fusstetter, H.; Nöth, H. Metallboranate und Boranatometallate. IX [1]; Äther-Addukte von Tris(boranato)-titan(III) und dimere Alkoxy-bis(boranato)-titan(III)-Verbindungen. Z. Anorg. Allg. Chem. 1976, 427, 97–113. [Google Scholar] [CrossRef]
  179. Fang, Z.Z.; Ma, L.P.; Kang, X.D.; Wang, P.J.; Wang, P.; Cheng, H.M. In situ formation and rapid decomposition of Ti(BH4)3 by mechanical milling LiBH4 with TiF3. Appl. Phys. Lett. 2009, 94, 1–4. [Google Scholar] [CrossRef]
  180. Nöth, H. Anorganische Reaktionen der Alkaliboranate. Angew. Chem. 1961, 73, 371–383. [Google Scholar] [CrossRef]
  181. Hagemann, H.; Longhini, M.; Kaminski, J.W.; Wesolowski, T.A.; Cerny, R.; Penin, N.; Sørby, M.H.; Hauback, B.C.; Severa, G.; Jensen, C.M. LiSc(BH4)4: A novel salt of Li+ and dIscrete Sc(BH4)4 complex anions. J. Phys. Chem. A 2008, 112, 7551–7555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Kim, C.; Hwang, S.-J.; Bowman, R.C.J.; Reiter, J.W.; Zan, J.A.; Kulleck, J.G.; Kabbour, H.; Majzoub, E.H.; Ozolins, V. LiSc(BH4)4 as a hydrogen storage material: Multinuclear high-resolution solid-state nmr and first-principles density functional theory studies. J. Phys. Chem. C 2009, 113, 9956–9968. [Google Scholar] [CrossRef] [Green Version]
  183. Morris, J.H.; Smith, W.E. Synthesis and characterisation of a tetrahydrofuran derivative of scandium tetrahydroborate. J. Chem. Soc. D Chem. Commun. 1970, 245a. [Google Scholar] [CrossRef]
  184. Cerny, 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]
  185. 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]
  186. Grinderslev, J.B.; Møller, K.T.; Yan, Y.; Chen, X.M.; Li, Y.; Li, H.W.; Zhou, W.; Skibsted, J.; Chen, X.; Jensen, T.R. Potassium octahydridotriborate: Diverse polymorphism in a potential hydrogen storage material and potassium ion conductor. Dalton Trans. 2019, 48, 8872–8881. [Google Scholar] [CrossRef]
  187. Fischer, P.; Zuttel, A. Order-disorder phase transition in NaBD4. Mater. Sci. Forum 2004, 443, 287–290. [Google Scholar] [CrossRef]
  188. Miwa, K.; Aoki, M.; Noritake, T.; Ohba, N.; Nakamori, Y.; Towata, S.; Zuttel, A.; Orimo, S. Thermodynamical stability of calcium borohydride Ca(BH4)2. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 74, 155122. [Google Scholar] [CrossRef]
  189. Richter, B.; Ravnsbaek, D.B.; Tumanov, N.; Filinchuk, Y.; Jensen, T.R. Manganese borohydride; synthesis and characterization. Dalton Trans. 2015, 44, 3988–3996. [Google Scholar] [CrossRef]
  190. Severa, G.; Hagemann, H.; Longhini, M.; Kaminski, J.W.; Wesolowski, T.A.; Jensen, C.M. Thermal desorption, vibrational spectroscopic, and DFT computational studies of the complex manganese borohydrides Mn(BH4)2 and [Mn(BH4)4]2−. J. Phys. Chem. C 2010, 114, 15516–15521. [Google Scholar] [CrossRef]
  191. Makhaev, V.D.; Borisov, A.P.; Gnilomedova, T.P.; Lobkovskii, É.B.; Chekhlov, A.N. Production of manganese borohydride complexes of manganese solvated with THF, and the structure of Mn(BH4)2 (THF)3. Bull. Acad. Sci. USSR Div. Chem. Sci. 1987, 36, 1582–1586. [Google Scholar] [CrossRef]
  192. Schaeffer, G.W.; Roscoe, J.S.; Stewart, A.C. The Reduction of Iron(III) Chloride with Lithium Aluminohydride and Lithium Borohydride: Iron(II) Borohydride. J. Am. Chem. Soc. 1956, 78, 729–733. [Google Scholar] [CrossRef]
  193. Varin, R.A.; Bidabadi, A.S. Rapid, ambient temperature hydrogen generation from the solid state Li-B-Fe-H system by mechanochemical activation synthesis. J. Power Sources 2015, 284, 554–565. [Google Scholar] [CrossRef]
  194. Stewart, A.C.; Schaeffer, G.W. The reaction of cobalt(II) bromide with lithium borohydride and lithium aluminohydride. J. Inorg. Nucl. Chem. 1956, 3, 194–197. [Google Scholar] [CrossRef]
  195. Raje, S.; Angamuthu, R. Solvent-free synthesis and reactivity of nickel(II) borohydride and nickel(II) hydride. Green Chem. 2019, 21, 2752–2758. [Google Scholar] [CrossRef]
  196. 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]
  197. Filinchuk, Y.; Chernyshov, D. Looking at hydrogen atoms with X-rays: Comprehensive synchrotron diffraction study of LiBH4. Acta Crystallogr. Sect. A Found. Crystallogr. 2007, 63, s240–s241. Available online: http://scripts.iucr.org/cgi-bin/paper?a38061 (accessed on 31 January 2022). [CrossRef] [Green Version]
  198. Soldate, A.M. Crystal structure of sodium borohydride. J. Am. Chem. Soc. 1947, 69, 987–988. [Google Scholar] [CrossRef]
  199. Dmitriev, V.; Filinchuk, Y.; Chernyshov, D.; Talyzin, A.V.; Ilewski, A.; Andersson, O.; Sundqvist, B.; Kurnosov, A. Pressure-temperature phase diagram of LiBH4: Synchrotron x-ray diffraction experiments and theoretical analysis. Phys. Rev. B Condens. Matter Mater. Phys. 2008, 77, 174112. [Google Scholar] [CrossRef] [Green Version]
  200. Ohba, N.; Miwa, K.; Aoki, M.; Noritake, T.; Towata, S.-i.; Nakamori, Y.; Orimo, S.-i.; Züttel, A. First-principles study on the stability of intermediate compounds of LiBH4. Phys. Rev. B 2006, 74, 075110. [Google Scholar] [CrossRef] [Green Version]
  201. Rude, L.H.; Nielsen, T.K.; Ravnsbæk, D.B.; Bosenberg, U.; Ley, M.B.; Richter, B.; Arnbjerg, L.M.; Dornheim, M.; Filinchuk, Y.; Besenbacher, F.; et al. Tailoring properties of borohydrides for hydrogen storage: A review. Phys. Status Solidi A 2011, 208, 1754–1773. [Google Scholar] [CrossRef]
  202. 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]
  203. 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]
  204. Bird, P.H.; Churchill, M.R. The crystal structure of zirconium(IV) borohydride (at –160°). Chem. Commun. (London) 1967, 403. [Google Scholar] [CrossRef]
  205. Broach, R.W.; Chuang, I.S.; Marks, T.J.; Williams, J.M. Metrical characterization of tridentate tetrahydroborate ligation to a transition-metal ion. Structure and bonding in Hf(BH4)4 by single-crystal neutron diffraction. Inorg. Chem. 1983, 22, 1081–1084. [Google Scholar] [CrossRef]
  206. Hoekstra, H.R.; Katz, J.J. The Preparation and Properties of the Group IV-B Metal Borohydrides. J. Am. Chem. Soc. 1949, 71, 2488–2492. [Google Scholar] [CrossRef]
  207. Dunbar, A.C.; Wright, J.C.; Grant, D.J.; Girolami, G.S. X-ray Crystal Structure of Thorium Tetrahydroborate, Th(BH4)4, and Computational Studies of An(BH4)4 (An = Th, U). Inorg. Chem. 2021, 60, 12489–12497. [Google Scholar] [CrossRef]
  208. Bernstein, E.R.; Hamilton, W.C.; Keiderling, T.A.; La Placa, S.J.; Lippard, S.J.; Mayerle, J.J. 14-Coordinate uranium(IV). Structure of uranium borohydride by single-crystal neutron diffraction. Inorg. Chem. 1972, 11, 3009–3016. [Google Scholar] [CrossRef]
  209. Charpin, P.; Nierlich, M.; Vigner, D.; Lance, M.; Baudry, D. Structure of the second crystalline form of uranium(IV) tetrahydroborate. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1987, 43, 1465–1467. [Google Scholar] [CrossRef]
  210. Banks, R.H.; Edelstein, N.M.; Spencer, B.; Templeton, D.H.; Zalkin, A. Volatility and molecular structure of neptunium(IV) borohydride. J. Am. Chem. Soc. 1980, 102, 620–623. [Google Scholar] [CrossRef] [Green Version]
  211. Tumanov, N.A.; Safin, D.A.; Richter, B.; Lodziana, Z.; Jensen, T.R.; Garcia, Y.; Filinchuk, Y. Challenges in the synthetic routes to Mn(BH4)2: Insight into intermediate compounds. Dalton Trans. 2015, 44, 6571–6580. [Google Scholar] [CrossRef] [PubMed]
  212. Ley, M.B.; Jørgensen, M.; Cerny, R.; Filinchuk, Y.; Jensen, T.R. From M(BH4)3 (M = La, Ce) borohydride frameworks to controllable synthesis of porous hydrides and ion conductors. Inorg. Chem. 2016, 55, 9748–9756. [Google Scholar] [CrossRef] [PubMed]
  213. Humphries, T.D.; Ley, M.B.; Frommen, C.; Munroe, K.T.; Jensen, T.R.; Hauback, B.C. Crystal structure and in situ decomposition of Eu(BH4)2 and Sm(BH4)2. J. Mater. Chem. A 2015, 3, 691–698. [Google Scholar] [CrossRef] [Green Version]
  214. Olsen, J.E.; Frommen, C.; Jensen, T.R.; Riktor, M.D.; Sorby, 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. 2014, 4, 1570–1582. [Google Scholar] [CrossRef]
  215. Olsen, J.E.; Frommen, C.; Sorby, M.H.; Hauback, B.C. Crystal structures and properties of solvent-free LiYb(BH4)4−xClx, Yb(BH4)3 and Yb(BH4)2−xClx. RSC Adv. 2013, 3, 10764–10774. [Google Scholar] [CrossRef]
  216. Jaron, T.; Orłowski, P.A.; Wegner, W.; Fijałkowski, K.J.; Leszczynski, P.J.; Grochala, W. Hydrogen storage materials: Room-temperature wet-chemistry approach toward mixed-metal borohydrides. Angew. Chem. Int. Ed. 2015, 54, 1236–1239. [Google Scholar] [CrossRef]
  217. Jaron, T.; Wegner, W.; Fijałkowski, K.J.; Leszczynski, P.J.; Grochala, W. Facile formation of thermodynamically unstable novel borohydride materials by a wet chemistry route. Chem. Eur. J. 2015, 21, 5689–5692. [Google Scholar] [CrossRef]
  218. 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]
  219. Seballos, L.; Zhang, J.Z.; Ronnebro, E.; Herberg, J.L.; Majzoub, E.H. Metastability and crystal structure of the bialkali complex metal borohydride NaK(BH4)2. J. Alloys Compd. 2009, 476, 446–450. [Google Scholar] [CrossRef]
  220. Semenenko, K.N.; Chavgun, A.P.; Surov, V.N. Interaction of sodium tetrahydroborate with potassium and lithium tetrahydroborates. Russ. J. Inorg. Chem. 1971, 16, 271–273. [Google Scholar]
  221. 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]
  222. Grinderslev, J.B.; Jepsen, L.H.; Lee, Y.-S.; Møller, K.T.; Cho, Y.W.; Černý, R.; Jensen, T.R. Structural diversity and trends in properties of an array of hydrogen-rich ammonium metal borohydrides. Inorg. Chem. 2020, 59, 12733–12747. [Google Scholar] [CrossRef]
  223. Černý, R.; Schouwink, P. The crystal chemistry of inorganic metal borohydrides and their relation to metal oxides. Acta Crystallogr. Sect. B 2015, B71, 619–640. [Google Scholar] [CrossRef]
  224. 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]
  225. 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, 5706. [Google Scholar] [CrossRef]
  226. Gao, M.; Yan, X.W.; Lu, Z.Y.; Xiang, T. Phonon-mediated high-temperature superconductivity in the ternary borohydride KB2H8 under pressure near 12 GPa. Phys. Rev. B 2021, 104, L100504. [Google Scholar] [CrossRef]
  227. 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]
  228. Matsuo, M.; Takamura, H.; Maekawa, H.; Li, H.-W.; Orimo, S.-I. Stabilization of lithium superionic conduction phase and enhancement of conductivity of LiBH4 by LiCl addition. Appl. Phys. Lett. 2009, 94, 084103. [Google Scholar] [CrossRef]
  229. D’Anna, V.; Daku, L.M.L.; Hagemann, H.; Kubel, F. Ionic layered BaFCl and Ba1−xSrxFCl compounds: Physical- and chemical-pressure effects. Phys. Rev. B 2010, 82, 024108. [Google Scholar] [CrossRef]
  230. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  231. Arnbjerg, L.M.; Ravnsbæk, D.B.; Filinchuk, Y.; Vang, R.T.; Cerenius, Y.; Besenbacher, F.; Jørgensen, J.-E.; Jakobsen, H.J.; Jensen, T.R. Structure and dynamics for LiBH4−LiCl solid solutions. Chem. Mater. 2009, 21, 5772–5782. [Google Scholar] [CrossRef] [Green Version]
  232. Blanchard, D.; Maronsson, J.B.; Riktor, M.D.; Kheres, J.; Sveinbjörnsson, D.; Bardají, E.G.; Léon, A.; Juranyi, F.; Wuttke, J.; Lefmann, K.; et al. Hindered rotational energy barriers of BH4 tetrahedra in β-Mg(BH4)2 from quasielastic neutron scattering and DFT calculations. J. Phys. Chem. C 2012, 116, 2013–2023. [Google Scholar] [CrossRef]
  233. Rude, L.H.; Zavorotynska, O.; Arnbjerg, L.M.; Ravnsbæk, D.B.; Malmkjær, R.A.; Grove, H.; Hauback, B.C.; Baricco, M.; Filinchuk, Y.; Besenbacher, F.; et al. Bromide substitution in lithium borohydride, LiBH4–LiBr. Int. J. Hydrogen Energy 2011, 36, 15664–15672. [Google Scholar] [CrossRef] [Green Version]
  234. Ravnsbæk, D.B.; Rude, L.H.; Jensen, T.R. Chloride substitution in sodium borohydride. J. Solid State Chem. 2011, 184, 1858–1866. [Google Scholar] [CrossRef]
  235. Llamas-Jansa, I.; Aliouane, N.; Deledda, S.; Fonneløp, J.E.; Frommen, C.; Humphries, T.; Lieutenant, K.; Sartori, S.; Sørby, M.H.; Hauback, B.C. Chloride substitution induced by mechano-chemical reactions between NaBH4 and transition metal chlorides. J. Alloys Compd. 2012, 530, 186–192. [Google Scholar] [CrossRef]
  236. Olsen, J.E.; Sørby, M.H.; Hauback, B.C. Chloride-substitution in sodium borohydride. J. Alloys Compd. 2011, 509, L228–L231. [Google Scholar] [CrossRef]
  237. 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] [Green Version]
  238. Borgschulte, A.; Gremaud, R.; Kato, S.; Stadie, N.P.; Remhof, A.; Züttel, A.; Matsuo, M.; Orimo, S.-I. Anharmonicity in LiBH4–LiI induced by anion exchange and temperature. Appl. Phys. Lett. 2010, 97, 031916. [Google Scholar] [CrossRef]
  239. Custelcean, R.; Jackson, J.E. Dihydrogen bonding:  Structures, energetics, and dynamics. Chem. Rev. 2001, 101, 1963–1980. [Google Scholar] [CrossRef]
  240. Noritake, T.; Aoki, M.; Matsumoto, M.; Miwa, K.; Towata, S.; Li, H.W.; Orimo, S. Crystal structure and charge density analysis of Ca(BH4)2. J. Alloys Compd. 2010, 491, 57–62. [Google Scholar] [CrossRef]
  241. Grinderslev, J.B.; Ley, M.B.; Lee, Y.-S.; Jepsen, L.H.; Jørgensen, M.; Cho, Y.W.; Skibsted, J.; Jensen, T.R. Ammine lanthanum and cerium borohydrides, M (BH4)3·NH3; trends in synthesis, structures, and thermal properties. Inorg. Chem. 2020, 59, 7768–7778. [Google Scholar] [CrossRef]
  242. Grinderslev, J.B.; Jensen, T.R. Trends in the series of ammine rare-earth-metal borohydrides: Relating structural and thermal properties. Inorg. Chem. 2021, 60, 2573–2589. [Google Scholar] [CrossRef]
  243. Johnson, S.R.; David, W.I.F.; Royse, D.M.; Sommariva, M.; Tang, C.Y.; Fabbiani, F.P.A.; Jones, M.O.; Edwards, P.P. The monoammoniate of lithium borohydride, Li(NH3)BH4: An effective ammonia storage compound. Chem. Asian J. 2009, 4, 849–854. [Google Scholar] [CrossRef] [PubMed]
  244. 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]
  245. Soloveichik, G.; Her, J.-H.; Stephens, P.W.; Gao, Y.; Rijssenbeek, J.; Andrus, M.; Zhao, J.C. Ammine magnesium borohydride complex as a new material for hydrogen storage: Structure and properties of Mg(BH4)2·2NH3. Inorg. Chem. 2008, 47, 4290–4298. [Google Scholar] [CrossRef]
  246. Yang, Y.; Liu, Y.; Li, Y.; Gao, M.; Pan, H. Synthesis and thermal decomposition behaviors of magnesium borohydride ammoniates with controllable composition as hydrogen storage materials. Chem. Asian J. 2013, 8, 476–481. [Google Scholar] [CrossRef]
  247. Huang, J.; Tan, Y.; Su, J.; Gu, Q.; Cerny, 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]
  248. Konoplev, V.N.; Silina, T.A. Ammonia derivatives of magnesium borohydride. Zh. Neorg. Khim. 1985, 30, 1125–1128. [Google Scholar]
  249. Chu, H.; Wu, G.; Xiong, Z.; Guo, J.; He, T.; Chen, P. Structure and hydrogen storage properties of calcium borohydride diammoniate. Chem. Mater. 2010, 22, 6021–6028. [Google Scholar] [CrossRef]
  250. He, T.; Wu, H.; Wu, G.; Wang, J.; Zhou, W.; Xiong, Z.; Chen, J.; Zhang, T.; Chen, P. Borohydride hydrazinates: High hydrogen content materials for hydrogen storage. Energy Environ. Sci. 2012, 5, 5686–5689. [Google Scholar] [CrossRef]
  251. He, T.; Wu, H.; Chen, J.; Zhou, W.; Wu, G.; Xiong, Z.; Zhang, T.; Chen, P. Alkali and alkaline-earth metal borohydride hydrazinates: Synthesis, structures and dehydrogenation. Phys. Chem. Chem. Phys. 2013, 15, 10487–10493. [Google Scholar] [CrossRef] [PubMed]
  252. Li, Z.; He, T.; Wu, G.; Ju, X.; Chen, P. The synthesis, structure and dehydrogenation of calcium borohydride hydrazinates. Int. J. Hydrogen Energy 2015, 40, 5333–5339. [Google Scholar] [CrossRef]
  253. Shriver, D.; Weller, M.; Overton, T.; Rouke, J.; Armstrong, F. Inorganic Chemistry, 6th ed.; Oxford University Press: Oxford, UK, 2014; ISBN-10: 1–4292–9906–1. [Google Scholar]
  254. Unemoto, A.; Matsuo, M.; Orimo, S. Complex hydrides for electrochemical energy storage. Adv. Funct. Mater. 2014, 24, 2267–2279. [Google Scholar] [CrossRef]
  255. 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. [Google Scholar] [CrossRef] [Green Version]
  256. 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]
  257. Tang, W.S.; Unemoto, A.; Zhou, W.; Stavila, V.; Matsuo, M.; Wu, H.; Orimo, S.-I.; 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] [Green Version]
  258. Lu, Z.; Ciucci, F. Metal Borohydrides as Electrolytes for Solid-State Li, Na, Mg, and Ca Batteries: A First-Principles Study. Chem. Mater. 2017, 29, 9308–9319. [Google Scholar] [CrossRef]
  259. Udovic, T.J.; Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Rush, J.J.; Takamura, H.; Orimo, S.-I. Sodium superionic conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750–3752. [Google Scholar] [CrossRef]
  260. Das, S.; Ngene, P.; Norby, P.; Vegge, T.; De Jongh, P.E.; Blanchard, D. All-Solid-State Lithium-Sulfur Battery Based on a Nanoconfined LiBH4 Electrolyte. J. Electrochem. Soc. 2016, 163, A2029–A2034. [Google Scholar] [CrossRef] [Green Version]
  261. Unemoto, A.; Ikeshoji, T.; Yasaku, S.; Matsuo, M.; Stavila, V.; Udovic, T.J.; Orimo, S.-I. Stable Interface Formation between TiS2 and LiBH4 in Bulk-Type All-Solid-State Lithium Batteries. Chem. Mater. 2015, 27, 5407–5416. [Google Scholar] [CrossRef]
  262. Matsuo, M.; Orimo, S. Lithium fast-ionic conduction in complex hydrides: Review and prospects. Adv. Energy Mater. 2011, 1, 161–172. [Google Scholar] [CrossRef]
  263. Guzik, M.N.; Mohtadi, R.; Sartori, S. Lightweight complex metal hydrides for Li-, Na-, and Mg-based batteries. J. Mater. Res. 2019, 34, 877–904. [Google Scholar] [CrossRef] [Green Version]
  264. Takano, A.; Oikawa, I.; Kamegawa, A.; Takamura, H. Enhancement of the lithium-ion conductivity of LiBH4 by hydration. Solid State Ion. 2016, 285, 47–50. [Google Scholar] [CrossRef]
  265. Zhang, T.; Wang, Y.; Song, T.; Miyaoka, H.; Shinzato, K.; Miyaoka, H.; Ichikawa, T.; Shi, S.; Zhang, X.; Isobe, S.; et al. Ammonia, a switch for controlling high ionic conductivity in lithium borohydride ammoniates. Joule 2018, 2, 1522–1533. [Google Scholar] [CrossRef] [Green Version]
  266. Yan, Y.; Grinderslev, J.B.; Lee, Y.-S.; Lee, M.; Cho, Y.W.; Černý, R.; Jensen, T.R. Ammonia-assisted fast Li-Ion conductivity in a new hemiammine lithium borohydride, LiBH4·1/2NH3. Chem. Commun. 2020, 56, 3971–3974. [Google Scholar] [CrossRef]
  267. Zhao, W.; Zhang, R.; Li, H.; Zhang, Y.; Wang, Y.; Wu, C.; Yan, Y.; Chen, Y. Li-Ion conductivity enhancement of LiBH4·xNH3 with in situ formed Li2O nanoparticles. ACS Appl. Mater. Interfaces 2021, 13, 31635–31641. [Google Scholar] [CrossRef]
  268. Liu, H.; Ren, Z.; Zhang, X.; Hu, J.; Gao, M.; Pan, H.; Liu, Y. Incorporation of ammonia borane groups in the lithium borohydride structure enables ultrafast lithium ion conductivity at room temperature for solid-state batteries. Chem. Mater. 2020, 32, 671–678. [Google Scholar] [CrossRef]
  269. Mohtadi, R.; Matsui, M.; Arthur, T.S.; Hwang, S.-J. Magnesium borohydride: From hydrogen storage to magnesium battery. Angew. Chem. Int. Ed. Engl. 2012, 51, 9780–9783. [Google Scholar] [CrossRef] [PubMed]
  270. 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]
  271. Le Ruyet, R.; Fleutot, B.; Berthelot, R.; Benabed, Y.; Hautier, G.; Filinchuk, Y.; Janot, R. Mg3(BH4)4(NH2)2 as inorganic solid electrolyte with high Mg2+ ionic conductivity. ACS Appl. Energy Mater. 2020, 3, 6093–6097. [Google Scholar] [CrossRef]
  272. 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] [Green Version]
  273. Burankova, T.; Roedern, E.; Maniadaki, A.E.; Hagemann, H.; Rentsch, D.; Lodziana, Z.; Battaglia, C.; Remhof, A.; Embs, J.P. Dynamics of the coordination complexes in a solid-state Mg electrolyte. J. Phys. Chem. Lett. 2018, 9, 6450–6455. [Google Scholar] [CrossRef] [PubMed]
  274. Yan, Y.; Dononelli, W.; Jørgensen, M.; Grinderslev, J.B.; Lee, Y.-S.; Cho, Y.W.; Černý, R.; Hammer, B.; Jensen, T.R. The mechanism of Mg2+ conduction in ammine magnesium borohydride promoted by a neutral molecule. Phys. Chem. Chem. Phys. 2020, 22, 9204–9209. [Google Scholar] [CrossRef]
  275. Yan, Y.; Grinderslev, J.B.; Jørgensen, M.; Skov, L.N.; Skibsted, J.; Jensen, T.R. Ammine magnesium borohydride nanocomposites for all-solid-state magnesium batteries. ACS Appl. Energy Mater. 2020, 3, 9264–9270. [Google Scholar] [CrossRef]
  276. Kisu, K.; Kim, S.; Inukai, M.; Oguchi, H.; Takagi, S.; Orimo, S. Magnesium borohydride ammonia borane as a magnesium ionic conductor. ACS Appl. Energy Mater. 2020, 3, 3174–3179. [Google Scholar] [CrossRef]
  277. Kisu, K.; Kim, S.; Shinohara, T.; Zhao, K.; Züttel, A.; Orimo, S. Monocarborane cluster as a stable fluorine-free calcium battery electrolyte. Sci. Rep. 2021, 11, 7563. [Google Scholar] [CrossRef]
  278. Luo, X.X.; Rawal, A.; Salman, M.S.; Aguey-Zinsou, K.F. Core-Shell NaBH4@Na2B12H12 Nanoparticles as Fast Ionic Conductors for Sodium-Ion Batteries. ACS Appl. Nano Mater. 2022, 5, 373–379. [Google Scholar] [CrossRef]
  279. Bao, K.; Pang, Y.; Yang, J.; Sun, D.L.; Fang, F.; Zheng, S.Y. Modulating composite polymer electrolyte by lithium closo-borohydride achieves highly stable solid-state battery at 25°C. Sci. China Mater. 2022, 65, 95–104. [Google Scholar] [CrossRef]
  280. Tran, B.L.; Allen, T.N.; Bowden, M.E.; Autrey, T.; Jensen, C.M. Effects of Glymes on the Distribution of Mg(B10H10) and Mg(B12H12) from the Thermolysis of Mg(BH4)2. Inorganics 2021, 9, 41. [Google Scholar] [CrossRef]
  281. Dobbins, T.A. Overview of the Structure-Dynamics-Function Relationships in Borohydrides for Use as Solid-State Electrolytes in Battery Applications. Molecules 2021, 26, 3239. [Google Scholar] [CrossRef] [PubMed]
  282. Marks, S.; Heck, J.G.; Habicht, M.H.; Oña-Burgos, P.; Feldmann, C.; Roesky, P.W. [Ln (BH4)2(THF)2] (Ln = Eu, Yb)—A highly luminescent material. synthesis, properties, reactivity, and NMR studies. J. Am. Chem. Soc. 2012, 134, 16983–16986. [Google Scholar] [CrossRef] [PubMed]
  283. Drozdov, A.P.; Eremets, M.I.; Troyan, I.A.; Ksenofontov, V.; Shylin, S.I. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature 2015, 525, 73–76. [Google Scholar] [CrossRef] [PubMed]
  284. Somayazulu, M.; Ahart, M.; Mishra, A.K.; Geballe, Z.M.; Baldini, M.; Meng, Y.; Struzhkin, V.V.; Hemley, R.J. Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures. Phys. Rev. Lett. 2019, 122, 027001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Drozdov, A.P.; Kong, P.P.; Minkov, V.S.; Besedin, S.P.; Kuzovnikov, M.A.; Mozaffari, S.; Balicas, L.; Balakirev, F.F.; Graf, D.E.; Prakapenka, V.B.; et al. Superconductivity at 250 K in lanthanum hydride under high pressures. Nature 2019, 569, 528–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Dovgaliuk, I.; Hagemann, H.; Leyssens, T.; Devillers, M.; Filinchuk, Y. CO2-promoted hydrolysis of KBH4 for efficient hydrogen co-generation. Int. J. Hydrogen Energy 2014, 39, 19603–19608. [Google Scholar] [CrossRef]
  287. Lombardo, L.; Ko, Y.; Zhao, K.; Yang, H.; Züttel, A. Direct CO2 capture and reduction to high-end chemicals with tetraalkylammonium borohydrides. Angew. Chem. 2021, 133, 9666–9675. [Google Scholar] [CrossRef]
  288. Knopf, I.; Cummins, C.C. Revisiting CO2 reduction with NaBH4 under aprotic conditions: Synthesis and characterization of sodium triformatoborohydride. Organometallics 2015, 34, 1601–1603. [Google Scholar] [CrossRef]
  289. Grice, K.A.; Groenenboom, M.C.; Manuel, J.D.A.; Sovereign, M.A.; Keith, J.A. Examining the selectivity of borohydride for carbon dioxide and bicarbonate reduction in protic conditions. Fuel 2015, 150, 139–145. [Google Scholar] [CrossRef]
  290. Zhu, W.; Zhao, J.; Wang, L.; Teng, Y.-L.; Dong, B.-X. Mechanochemical reactions of alkali borohydride with CO2 under ambient temperature. J. Solid State Chem. 2019, 277, 828–832. [Google Scholar] [CrossRef]
  291. Picasso, C.V.; Safin, D.A.; Dovgaliuk, I.; Devred, F.; Debecker, D.; Li, H.-W.; Proost, J.; Filinchuk, Y. Reduction of CO2 with KBH4 in solvent-free conditions. Int. J. Hydrogen Energy 2016, 41, 14377–14386. [Google Scholar] [CrossRef]
  292. Laversenne, L.; Goutaudier, C.; Chiriac, R.; Sigala, C.; Bonnetot, B. Hydrogen storage in borohydrides comparison of hydrolysis conditions of LiBH4, NaBH4 and KBH4. J. Therm. Anal. Calorim. 2008, 94, 785–790. [Google Scholar] [CrossRef]
  293. Vitillo, J.G.; Groppo, E.; Bardají, E.G.; Baricco, M.; Bordiga, S. Fast carbon dioxide recycling by reaction with γ-Mg(BH4)2. Phys. Chem. Chem. Phys. 2014, 16, 22482–22486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Zuttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage properties of LiBH4. J. Alloys Compd. 2003, 356, 515–520. [Google Scholar] [CrossRef]
  295. Nakamori, Y.; Miwa, K.; Ninomiya, A.; Li, H.; Ohba, N.; Towata, S.-I.; Züttel, A.; Orimo, S.-I. Correlation between thermodynamical stabilities of metal borohydrides and cation electronegativites: First-principles calculations and experiments. Phys. Rev. B 2006, 74, 045126. [Google Scholar] [CrossRef] [Green Version]
  296. Kharbachi, E.A.; Pinatel, E.; Nuta, I.; Baricco, M. A thermodynamic assessment of LiBH4. Calphad Comput. Coupling Phase Diagrams Thermochem. 2012, 39, 80–90. [Google Scholar] [CrossRef] [Green Version]
  297. Dematteis, E.M.; Pinatel, E.R.; Corno, M.; Jensen, T.R.; Baricco, M. Phase diagrams of the LiBH4-NaBH4-KBH4 system. Phys. Chem. Chem. Phys. 2017, 19, 25071–25079. [Google Scholar] [CrossRef] [PubMed]
  298. Pinatel, E.R.; Albanese, E.; Civalleri, B.; Baricco, M. Thermodynamic modelling of Mg(BH4)2. J. Alloys Compd. 2015, 645 (Suppl. 1), S64–S68. [Google Scholar] [CrossRef] [Green Version]
  299. Błoński, P.; Łodziana, Z. Correlation between the ionic potential and thermal stability of metal borohydrides: First-principles investigations. Phys. Rev. B 2014, 90, 054114. [Google Scholar] [CrossRef]
  300. Blanchard, D.; Zatti, M.; Vegge, T. Analysis of the decomposition gases from α and β-Cd(BH4)2 synthesized by temperature controlled mechanical milling. J. Alloys Compd. 2013, 547, 76–80. [Google Scholar] [CrossRef] [Green Version]
  301. Mulliken, R.S. A new electroaffinity scale; together with data on valence states and on valence ionization potentials and electron affinities. J. Chem. Phys. 1934, 2, 782–793. [Google Scholar] [CrossRef]
  302. Parr, R.G.; Pearson, R.G. Absolute hardness: Companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. [Google Scholar] [CrossRef]
  303. Kang, X.; Wang, P.; Cheng, H. Advantage of TiF3 over TiCl3 as a dopant precursor to improve the thermodynamic property of Na3AlH6. Scr. Mater. 2007, 56, 361–364. [Google Scholar] [CrossRef]
  304. Yin, L.C.; Wang, P.; Fang, Z.; Cheng, H.M. Thermodynamically tuning LiBH4 by fluorine anion doping for hydrogen storage: A density functional study. Chem. Phys. Lett. 2008, 450, 318–321. [Google Scholar] [CrossRef]
  305. Sveinbjornsson, D.; Blanchard, D.; Myrdal, J.S.G.; Younesi, R.; Viskinde, R.; Riktor, M.D.; Norby, P.; Vegge, T. Ionic conductivity and the formation of cubic CaH2 in the LiBH4-Ca(BH4)2 composite. J. Solid State Chem. 2014, 211, 81–89. [Google Scholar] [CrossRef] [Green Version]
  306. Ozolins, V.; Majzoub, E.H.; Wolverton, C. First-principles prediction of thermodynamically reversible hydrogen storage reactions in the Li-Mg-Ca-B-H system. J. Am. Chem. Soc. 2009, 131, 230–237. [Google Scholar] [CrossRef] [PubMed]
  307. Xiong, Z.T.; Hu, J.J.; Wu, G.T.; Chen, P.; Luo, W.F.; 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]
  308. Gibb, B.C. The centenary (maybe) of the hydrogen bond. Nat. Chem. 2020, 12, 665–667. [Google Scholar] [CrossRef]
  309. Chu, H.; Wu, G.; Zhang, Y.; Xiong, Z.; Guo, J.; He, T.; Chen, P. Improved dehydrogenation properties of calcium borohydride combined with alkaline-earth metal amides. J. Phys. Chem. C 2011, 115, 18035–18041. [Google Scholar] [CrossRef]
  310. Chater, P.A.; David, W.I.F.; Anderson, P.A. Synthesis and structure of the new complex hydride Li2BH4NH2. Chem. Commun. 2007, 45, 4770–4772. [Google Scholar] [CrossRef]
  311. Somer, M.; Acar, S.; Koz, C.; Kokal, I.; Hohn, P.; Cardoso-Gil, R.; Aydemir, U.; Akselrud, L. α- and β-Na2[BH4][NH2]: Two modifications of a complex hydride in the system NaNH2–NaBH4; syntheses, crystal structures, thermal analyses, mass and vibrational spectra. J. Alloys Compd. 2010, 491, 98–105. [Google Scholar] [CrossRef]
  312. Chu, H.L.; Xiong, Z.T.; Wu, G.T.; Guo, J.P.; Zheng, X.L.; He, T.; Wu, C.Z.; Chen, P. Hydrogen Storage Properties of Ca(BH4)2–LiNH2 System. Chem. Asian J. 2010, 5, 1594–1599. [Google Scholar] [CrossRef] [PubMed]
  313. Poonyayant, N.; Stavila, V.; Majzoub, E.H.; Klebanoff, L.E.; Behrens, R.; Angboonpong, N.; Ulutagay-Kartin, M.; Pakawatpanurut, P.; Hecht, E.S.; Breit, J.S. An investigation into the hydrogen storage characteristics of Ca(BH4)2/LiNH2 and Ca(BH4)2/NaNH2: Evidence of intramolecular destabilization. J. Phys. Chem. C 2014, 118, 14759–14769. [Google Scholar] [CrossRef]
  314. Aoki, M.; Miwa, K.; Noritake, T.; Kitahara, G.; Nakamori, Y.; Orimo, S.; Towata, S. Destabilization of LiBH4 by mixing with LiNH2. Appl. Phys. A Mater. Sci. Process. 2005, 80, 1409–1412. [Google Scholar] [CrossRef]
  315. Grube, E.; Jensen, S.R.H.; Nielsen, U.G.; Jensen, T.R. Reactivity of magnesium borohydride—Metal hydride composites, γ-Mg(BH4)2-MHx, M = Li, Na, Mg, Ca. J. Alloys Compd. 2019, 770, 1155–1163. [Google Scholar] [CrossRef]
  316. Ibikunle, A.A.; Goudy, A.J. Kinetics and modeling study of a Mg(BH4)2/Ca(BH4)2 destabilized system. Int. J. Hydrogen Energy 2012, 37, 12420–12424. [Google Scholar] [CrossRef]
  317. Durojaiye, T.; Ibikunle, A.; Goudy, A.J. Hydrogen storage in destabilized borohydride materials. Int. J. Hydrogen Energy 2010, 35, 10329–10333. [Google Scholar] [CrossRef]
  318. Dematteis, E.M.; Baricco, M. Hydrogen desorption in Mg(BH4)2-Ca(BH4)2 system. Energies 2019, 12, 3230. [Google Scholar] [CrossRef] [Green Version]
  319. Au, M.; Jurgensen, A. Modified lithium borohydrides for reversible hydrogen storage. J. Phys. Chem. B 2006, 110, 7062–7067. [Google Scholar] [CrossRef] [PubMed]
  320. Au, M.; Jurgensen, A.; Zeigler, K. Modified lithium borohydrides for reversible hydrogen storage (2). J. Phys. Chem. B 2006, 110, 26482–26487. [Google Scholar] [CrossRef] [PubMed]
  321. Muller, A.; Havre, L.; Mathey, F.; Petit, V.I.; Bensoam, J. Production of Hydrogen. U.S. Patent 4193978 A, 18 March 1980. [Google Scholar]
  322. Mosegaard, L.; Møller, B.; Jørgensen, J.E.; Bosenberg, U.; Dornheim, M.; Hanson, J.C.; Cerenius, Y.; Walker, G.S.; Jakobsen, H.J.; Besenbacher, F.; et al. Intermediate phases observed during decomposition of LiBH4. J. Alloys Compd. 2007, 446, 301–305. [Google Scholar] [CrossRef]
  323. Her, J.H.; Yousufuddin, M.; Zhou, W.; Jalisatgi, S.S.; Kulleck, J.G.; Zan, J.A.; Hwang, S.J.; Bowman, R.C., Jr.; Udovic, T.J. Crystal structure of Li2B12H12: A possible intermediate species in the decomposition of LiBH4. Inorg. Chem. 2008, 47, 9757–9759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Her, J.H.; Wu, H.; Verdal, N.; Zhou, W.; Stavila, V.; Udovic, T.J. Structures of the strontium and barium dodecahydro-closo-dodecaborates. J. Alloys Compd. 2012, 514, 71–75. [Google Scholar] [CrossRef]
  325. Friedrichs, O.; Remhof, A.; Hwang, S.; Zuttel, A. Role of Li2B12H12 for the formation and decomposition of LiBH4. Chem. Mater. 2010, 22, 3265–3268. [Google Scholar] [CrossRef]
  326. 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]
  327. Orimo, S.I.; Nakamori, Y.; Kitahara, G.; Miwa, K.; Ohba, N.; Towata, S.; Zuttel, A. Dehydriding and rehydriding reactions of LiBH4. J. Alloys Compd. 2005, 404, 427–430. [Google Scholar] [CrossRef]
  328. Orimo, S.; Nakamori, Y.; Ohba, N.; Miwa, K.; Aoki, M.; Towata, S.; Zuttel, A. Experimental studies on intermediate compound of LiBH4. Appl. Phys. Lett. 2006, 89, 021920. [Google Scholar] [CrossRef]
  329. Hwang, S.J.; Bowman, R.C., Jr.; Reiter, J.W.; Rijssenbeck, J.; Soloveichik, G.L.; Zhao, J.C.; Kabbour, H.; Ahn, C.C. NMR confirmation for formation of [B12H12]2− complexes during hydrogen desorption from metal borohydrides. J. Phys. Chem. C 2008, 112, 3164–3169. [Google Scholar] [CrossRef]
  330. Kato, S.; Bielmann, M.; Borgschulte, A.; Zakaznova-Herzog, V.; Remhof, A.; Orimo, S.; Zuttel, A. Effect of the surface oxidation of LiBH4 on the hydrogen desorption mechanism. Phys. Chem. Chem. Phys. 2010, 12, 10950–10955. [Google Scholar] [CrossRef] [PubMed]
  331. Zuttel, A.; Borgschulte, A.; Orimo, S.I. Tetrahydroborates as new hydrogen storage materials. Scr. Mater. 2007, 56, 823–828. [Google Scholar] [CrossRef]
  332. Schlapbach, L.; Zuttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  333. Orimo, S.I.; Nakamori, Y.; Eliseo, J.R.; Zuttel, A.; Jensen, C.M. Complex hydrides for hydrogen storage. Chem. Rev. 2007, 107, 4111–4132. [Google Scholar] [CrossRef] [PubMed]
  334. Palade, P.; Comanescu, C.; Mercioniu, I. Improvements of hydrogen desorption of lithium borohydride by impregnation onto MSH-H carbon replica. J. Ovonic Res. 2012, 8, 155–160. [Google Scholar]
  335. Comanescu, C.; Guran, C.; Palade, P. Improvements of kinetic properties of LiBH4 by supporting on MSU-H type mesoporous silica. Optoelectron. Adv. Mater. Rapid Commun. 2010, 4, 705–708. [Google Scholar]
  336. Cahen, S.; Eymery, J.B.; Janot, R.; Tarascon, J.M. Improvement of the LiBH4 hydrogen desorption by inclusion into mesoporous carbons. J. Power Sources 2009, 189, 902–908. [Google Scholar] [CrossRef]
  337. Kruk, M.; Dufour, B.; Celer, E.B.; Kowalewski, T.; Jaroniec, M.; Matyjaszewski, K. Synthesis of mesoporous carbons using ordered and disordered mesoporous silica templates and polyacrylonitrile as carbon precursor. J. Phys. Chem. B 2005, 109, 9216–9225. [Google Scholar] [CrossRef]
  338. Ryoo, R.; Joo, S.H.; Kruk, M.; Jaroniec, M. Ordered mesoporous carbons. Adv. Mater. 2001, 13, 677–681. [Google Scholar] [CrossRef]
  339. Ngene, P.; Adelhelm, P.; Beale, A.M.; de Jong, K.P.; de Jongh, P.E. LiBH4/SBA-15 Nanocomposites prepared by melt infiltration under hydrogen pressure: Synthesis and hydrogen sorption properties. J. Phys. Chem. C 2010, 114, 6163–6168. [Google Scholar] [CrossRef]
  340. Zhang, Y.; Zhang, W.Z.; Fan, M.Q.; Liu, S.S.; Chu, H.L.; Zhang, Y.H.; Gao, X.Y.; Sun, L.X. Enhanced hydrogen storage performance of LiBH4−SiO2−TiF3 composite. J. Phys. Chem. C 2008, 112, 4005–4010. [Google Scholar] [CrossRef]
  341. Gross, A.F.; Vajo, J.J.; Van Atta, S.L.; Olson, G.L. Enhanced hydrogen storage kinetics of LiBH4 in nanoporous carbon scaffolds. J. Phys. Chem. C 2008, 112, 5651–5657. [Google Scholar] [CrossRef]
  342. David, R.L. Handbook of Chemistry and Physics, 90th ed.; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  343. Fang, Z.Z.; Wang, P.; Rufford, T.E.; Kang, X.D.; Lu, G.Q.; Cheng, H.M. Kinetic- and thermodynamic-based improvements of lithium borohydride incorporated into activated carbon. Acta Mater. 2008, 56, 6257–6263. [Google Scholar] [CrossRef]
  344. Ozawa, T. A New method of analyzing thermogravimetric data. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886. [Google Scholar] [CrossRef] [Green Version]
  345. Ozawa, T. Kinetic analysis of derivative curves in thermal analysis. J. Therm. Anal. 1970, 2, 301–324. [Google Scholar] [CrossRef]
  346. Nielsen, T.K.; Bosenberg, U.; Gosalawit, R.; Dornheim, M.; Cerenius, Y.; Besenbacher, F.; Jensen, T.R. A Reversible nanoconfined chemical reaction. ACS Nano 2010, 4, 3903–3908. [Google Scholar] [CrossRef]
  347. Au, M.; Spencer, W.; Jurgensen, A.; Zeigler, C. Hydrogen storage properties of modified lithium borohydrides. J. Alloys Compd. 2008, 462, 1–2. [Google Scholar] [CrossRef]
  348. Faivre, C.; Bellet, D.; Dolino, G. Phase transitions of fluids confined in porous silicon: A differential calorimetry investigation. Eur. Phys. J. B 1999, 7, 19–36. [Google Scholar] [CrossRef]
  349. Petrov, O.; Furo, I. Curvature-dependent metastability of the solid phase and the freezing-melting hysteresis in pores. Phys. Rev. E 2006, 73, 011608. [Google Scholar] [CrossRef]
  350. Verdal, N.; Udovic, T.J.; Rush, J.J.; Liu, X.; Majzoub, E.H.; Vajo, J.J.; Gross, A.F. Dynamical perturbations of tetrahydroborate anions in LiBH4 due to nanoconfinement in controlled-pore carbon scaffolds. J. Phys. Chem. C 2013, 117, 17983–17995. [Google Scholar] [CrossRef]
  351. Ngene, P.; Nale, A.; Eggenhuisen, T.M.; Oschatz, M.; Embs, J.P.; Remhof, A.; de Jongh, P.E. Confinement effects for lithium borohydride: Comparing silica and carbon scaffolds. J. Phys. Chem. C 2017, 121, 4197–4205. [Google Scholar] [CrossRef]
  352. Zhang, X.; Zhang, L.; Zhang, W.; Ren, Z.; Huang, Z.; Hu, J.; Gao, M.; Pan, H.; Liu, Y. Nano-synergy enables highly reversible storage of 9.2 wt% hydrogen at mild conditions with lithium borohydride. Nano Energy 2021, 83, 105839. [Google Scholar] [CrossRef]
  353. Grochala, W.; Edwards, P.P. Thermal decomposition of the non-interstitial hydrides for the storage and production of hydrogen. Chem. Rev. 2004, 104, 1283–1316. [Google Scholar] [CrossRef] [PubMed]
  354. Zaluska, A.; Zaluski, L.; Ström-Olsen, J.O. Structure, catalysis and atomic reactions on the nano-scale: A systematic approach to metal hydrides for hydrogen storage. Appl. Phys. A 2001, 72, 783. [Google Scholar] [CrossRef]
  355. Imamura, H.; Masanari, K.; Kusuhara, M.; Katsumoto, H.; Sumi, T.; Sakata, Y. High hydrogen storage capacity of nanosized magnesium synthesized by high energy ball-milling. J. Alloys Compd. 2005, 386, 211–216. [Google Scholar] [CrossRef]
  356. Zaluski, L.; Zaluska, A.; Ström-Olsen, J.O. Nanocrystalline metal hydrides. J. Alloys Compd. 1997, 253, 70–79. [Google Scholar] [CrossRef]
  357. Zhu, M.; Wang, H.; Ouyang, L.Z.; Zeng, M.Q. Composite structure and hydrogen storage properties in Mg-base alloys. Int. J. Hydrogen Energy 2006, 31, 251–257. [Google Scholar] [CrossRef]
  358. Wiswall, R. Hydrogen Storage in Metals. In Hydrogen Metals II; Springer: Berlin, Germany, 1978; Volume 29, pp. 201–242. [Google Scholar] [CrossRef]
  359. Fukai, Y. The Metal–Hydrogen System, Basic Bulk Properties; Springer Series in Materials Science; Springer: Berlin, Germany, 1993; ISBN 978-3-540-28883-1. [Google Scholar] [CrossRef]
  360. Bogdanovic, B.; Bohmhamme, 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]
  361. Reule, H.; Hirscher, M.; Weißhardt, A.; Krönmuller, H. Hydrogen desorption properties of mechanically alloyed MgH2 composite materials. J. Alloys Compd. 2000, 305, 246–252. [Google Scholar] [CrossRef]
  362. Vajo, J.J.; Skeith, S.L.; Mertens, F. Reversible storage of hydrogen in destabilized LiBH4. J. Phys. Chem. B 2005, 109, 3719–3722. [Google Scholar] [CrossRef]
  363. Bosenberg, U.; Doppiu, S.; Mosegaard, L.; Barkhordarian, G.; Eigen, N.; Borgschulte, A.; Jensen, T.R.; Cerenius, Y.; Gutfleisch, O.; Klassen, T.; et al. Hydrogen sorption properties of MgH2–LiBH4 composites. Acta Mater. 2007, 55, 3951–3958. [Google Scholar] [CrossRef]
  364. Pinkerton, F.E.; Meyer, M.S.; Meisner, G.P.; Balogh, M.P.; Vajo, J.J. Phase boundaries and reversibility of LiBH4/MgH2 hydrogen storage material. J. Phys. Chem. C 2007, 111, 12881–12885. [Google Scholar] [CrossRef]
  365. Walker, G.S.; Grant, D.M.; Price, T.C.; Yu, X.; Legrand, V. High capacity multicomponent hydrogen storage materials: Investigation of the effect of stoichiometry and decomposition conditions on the cycling behaviour of LiBH4–MgH2. J. Power Sources 2009, 194, 1128–1134. [Google Scholar] [CrossRef] [Green Version]
  366. Bosenberg, U.; Kim, J.W.; Gosslar, D.; Eigen, N.; Jensen, T.R.; von Colbe, J.M.B.; Zhou, Y.; Dahms, M.; Kim, D.H.; Gunther, R.; et al. Role of additives in LiBH4–MgH2 reactive hydride composites for sorption kinetics. Acta Mater. 2010, 58, 3381–3389. [Google Scholar] [CrossRef] [Green Version]
  367. Wan, X.F.; Markmaitree, T.; Osborn, W.; Shaw, L.L. Nanoengineering-enabled solid-state hydrogen uptake and release in the LiBH4 Plus MgH2 System. J. Phys. Chem. C 2008, 112, 18232–18243. [Google Scholar] [CrossRef]
  368. Bosenberg, U.; Ravnsbæk, D.B.; Hagemann, H.; D’Anna, V.; Minella, C.B.; Pistidda, C.; van Beek, W.; Jensen, T.R.; Bormann, R.; Dornheim, M. Pressure and temperature influence on the desorption pathway of the LiBH4−MgH2 composite system. J. Phys. Chem. C 2010, 114, 15212–15217. [Google Scholar] [CrossRef]
  369. Price, T.E.C.; Grant, D.M.; Legrand, V.; Walker, G.S. Enhanced kinetics for the LiBH4:MgH2 multi-component hydrogen storage system—The effects of stoichiometry and decomposition environment on cycling behaviour. Int. J. Hydrogen Energy 2010, 35, 4154–4161. [Google Scholar] [CrossRef] [Green Version]
  370. Martelli, P.; Caputo, R.; Remhof, A.; Mauron, P.; Borgschulte, A.; Zuttel, A. Stability and decomposition of NaBH4. J. Phys. Chem. C 2010, 114, 7173–7177. [Google Scholar] [CrossRef]
  371. Ugale, A.D.; Ghodke, N.P.; Kang, G.S.; Nam, K.B.; Bhoraskar, S.V.; Mathe, V.L.; Yoo, J.B. Cost-effective synthesis of carbon loaded Co3O4 for controlled hydrogen generation via NaBH4 hydrolysis. Int. J. Hydrogen Energy 2022, 47, 16–29. [Google Scholar] [CrossRef]
  372. Kaya, S.; Yilmaz, Y.; Er, O.F.; Alpaslan, D.; Ulas, B.; Dudu, T.E.; Kivrak, H. Highly Active RuPd Bimetallic Catalysts for Sodium Borohydride Electrooxidation and Hydrolysis. J. Electron. Mater. 2022, 51, 403–411. [Google Scholar] [CrossRef]
  373. Yu, S.S.; Lee, T.H.; Oh, T.H. Ag-Ni nanoparticles supported on multiwalled carbon nanotubes as a cathode electrocatalyst for direct borohydride-hydrogen peroxide fuel cells. Fuel 2022, 315, 123151. [Google Scholar] [CrossRef]
  374. Oh, T.H. Effect of cathode conditions on performance of direct borohydride-hydrogen peroxide fuel cell system for space exploration. Renew. Energy 2021, 178, 1156–1164. [Google Scholar] [CrossRef]
  375. Schlesinger, H.J.; Brown, H.C.; Abraham, B.; Bond, A.C.; Davidson, N.; Finholt, A.E.; Gilbreath, J.R.; Hoekstra, H.; Horvitz, L.; Hyde, E.K.; et al. New Developments in the chemistry of diborane and the borohydrides. I. General SUmmary. J. Am. Chem. Soc. 1953, 75, 186–190. [Google Scholar] [CrossRef]
  376. Broja, G.; Schlabacher, W. Process for the Production of Alkali Metal Borohydrides. DE Patent DE 000001108670 A, 15 June 1961. [Google Scholar]
  377. Schubert, F.; Lang, K.; Schlabacher, W. Process for the Production of Borohydrides. DE Patent DE 000001067005 A, 15 October 1959. [Google Scholar]
  378. Haga, T.; Kojima, Y. Method for Manufacturing Metal Borohydride. JP Patent P 2002-241109, 28 August 2002. [Google Scholar]
  379. Kojima, Y.; Haga, T. Recycling process of sodium metaborate to sodium borohydride. Int. J. Hydrogen Energy 2003, 28, 989–993. [Google Scholar] [CrossRef]
  380. Li, Z.P.; Morigazaki, N.; Liu, B.H.; Suda, S. Preparation of sodium borohydride by the reaction of MgH2 with dehydrated borax through ball milling at room temperature. J. Alloys Compd. 2003, 349, 232–236. [Google Scholar] [CrossRef]
  381. Ozolins, V.; Majzoub, E.H.; Wolverton, C. First-principles prediction of a ground state crystal structure of magnesium borohydride. Phys. Rev. Lett. 2008, 100, 135501. [Google Scholar] [CrossRef] [PubMed]
  382. Fichtner, M.; Zhao-Karger, Z.; Hu, J.; Roth, A.; Weidler, P. The kinetic properties of Mg(BH4)2 infiltrated in activated carbon. Nanotechnology 2009, 20, 204029. [Google Scholar] [CrossRef]
  383. Sartori, S.; Knudsen, K.D.; Zhao-Karger, Z.; Bardaij, E.G.; Fichtner, M.; Hauback, B.C. Small-angle scattering investigations of Mg-borohydride infiltrated in activated carbon. Nanotechnology 2009, 20, 505702. [Google Scholar] [CrossRef]
  384. George, L.; Saxena, S.K. Structural stability of metal hydrides, alanates and borohydrides of alkali and alkali-earth elements: A review. Int. J. Hydrogen Energy 2010, 35, 5454–5470. [Google Scholar] [CrossRef]
  385. Matsunaga, T.; Buchter, F.; Mauron, P.; Bielman, A.; Nakamori, Y.; Orimo, S.; Ohba, N.; Miwa, K.; Towata, S.; Zuttel, A. Hydrogen storage properties of Mg[BH4]2. J. Alloys Compd. 2008, 459, 583–588. [Google Scholar] [CrossRef]
  386. Severa, G.; Ronnebro, E.; Jensen, C.M. Direct hydrogenation of magnesium boride to magnesium borohydride: Demonstration of >11 weight percent reversible hydrogenstorage. Chem. Commun. 2010, 46, 421–423. [Google Scholar] [CrossRef] [PubMed]
  387. Li, H.W.; Kikuchi, K.; Nakamori, Y.; Ohba, N.; Miwa, K.; Towata, S.; Orimo, S. Dehydriding and rehydriding processes of well-crystallized Mg(BH4)2 accompanying with formation of intermediate compounds. Acta Mater. 2008, 56, 1342–1347. [Google Scholar] [CrossRef]
  388. Van Setten, M.J.; de Wijs, G.A.; Fichtner, M.; Brocks, G. A Density Functional Study of α-Mg(BH4)2. Chem. Mater. 2008, 20, 4952–4956. [Google Scholar] [CrossRef]
  389. Cansizoglu, M.F.; Karabacak, T. Enhanced hydrogen storage properties of magnesium nanotrees with nanoleaves. Mat. Res. Soc. Symp. Proc. 2009, 1216, 503. [Google Scholar] [CrossRef]
  390. Bayca, S.U.; Cansizoglu, M.F.; Biris, A.S.; Watanabe, F.; Karabacak, T. Enhanced oxidation resistance of magnesium nanorods grown by glancing angle deposition. Int. J. Hydrogen Energy 2011, 36, 5998–6004. [Google Scholar] [CrossRef]
  391. Zavorotynska, O.; Deledda, S.; Vitillo, J.; Saldan, I.; Guzik, M.; Baricco, M.; Walmsley, J.C.; Muller, J.; Hauback, B.C. Combined X-ray and raman studies on the effect of cobalt additives on the decomposition of magnesium borohydride. Energies 2015, 8, 9173–9190. [Google Scholar] [CrossRef] [Green Version]
  392. Zavorotynska, O.; Saldan, I.; Hino, S.; Humphries, T.D.; Deledda, S.; Hauback, B.C. Hydrogen cycling in γ-Mg(BH4)2 with cobalt-based additives. J. Mater. Chem. A 2015, 3, 6592–6602. [Google Scholar] [CrossRef] [Green Version]
  393. Liang, J.J. Theoretical insight on tailoring energetics of Mg hydrogen absorption/desorption through nano-engineering. Appl. Phys. A Mater. Sci. Process. 2005, 80, 173–178. [Google Scholar] [CrossRef]
  394. Kim, J.-H.; Jin, S.-A.; Shim, J.-H.; Cho, Y.W. Reversible hydrogen storage in calcium borohydride Ca(BH4)2. Scr. Mater. 2008, 58, 481–483. [Google Scholar] [CrossRef]
  395. Aoki, M.; Miwa, K.; Noritake, T.; Ohba, N.; Matsumoto, M.; Li, H.W.; Nakamori, Y.; Towata, S.; Orimo, S. Structural and dehydriding properties of Ca(BH4)2. Appl. Phys. A 2008, 92, 601–605. [Google Scholar] [CrossRef]
  396. Kim, J.H.; Jin, S.A.; Shim, J.H.; Cho, Y.W. Thermal decomposition behavior of calcium borohydride Ca(BH4)2. J. Alloys Compd. 2008, 461, L20–L22. [Google Scholar] [CrossRef]
  397. Kim, J.H.; Shim, J.H.; Cho, Y.W. On the reversibility of hydrogen storage in Ti- and Nb-catalyzed Ca(BH4)2. J. Power Sources 2008, 181, 140–143. [Google Scholar] [CrossRef]
  398. Manchester, F.D.; San-Martin, A.; Pitre, J.M. The H-Pd (hydrogen-palladium) System. J. Phase Equilibria JPE 1994, 15, 62–83. [Google Scholar] [CrossRef]
  399. Parry, R.W.; Schultz, D.R.; Girardot, P.R. The Preparation and properties of hexamminecobalt(III) borohydride, hexamminechromium(III) borohydride and ammonium borhydride. J. Am. Chem. Soc. 1958, 80, 1–3. [Google Scholar] [CrossRef]
  400. Crabtree, R.H. A New Type of Hydrogen Bond. Science 1998, 282, 2000–2001. [Google Scholar] [CrossRef]
  401. Sit, V.; Geanangel, R.A.; Wendlandt, W.W. The thermal dissociation of NH3BH3. Thermochim. Acta 1987, 113, 379–382. [Google Scholar] [CrossRef]
  402. Li, L.; Yao, X.; Sun, C.; Du, A.; Cheng, L.; Zhu, Z.; Yu, C.; Zou, J.; Smith, S.C.; Wang, P.; et al. Lithium-catalyzed dehydrogenation of ammonia borane within mesoporous carbon framework for chemical hydrogen storage. Adv. Funct. Mater. 2009, 19, 265–271. [Google Scholar] [CrossRef] [Green Version]
  403. Sepehri, S.; Feaver, A.; Shaw, W.J.; Howard, C.J.; Zhang, Q.; Autrey, T.; Cao, G. Spectroscopic studies of dehydrogenation of ammonia borane in carbon cryogel. J. Phys. Chem. B 2007, 111, 14285–14289. [Google Scholar] [CrossRef] [PubMed]
  404. Kim, H.; Karkamkar, A.; Autrey, T.; Chupas, P.; Proffen, T. Determination of structure and phase transition of light element nanocomposites in mesoporous silica: Case study of NH3BH3 in MCM-41. J. Am. Chem. Soc. 2009, 131, 13749–13755. [Google Scholar] [CrossRef] [PubMed]
  405. Paolone, A.; Palumbo, O.; Rispoli, P.; Cantelli, R.; Autrey, T.; Karkamkar, A. Absence of the structural phase transition in ammonia borane dispersed in mesoporous silica: Evidence of novel thermodynamic properties. J. Phys. Chem. C 2009, 113, 10319–10321. [Google Scholar] [CrossRef]
  406. Wolf, G.; Baumann, J.; Baitalow, F.; Hoffmann, F.P. Calorimetric process monitoring of thermal decomposition of B–N–H compounds. Thermochim. Acta 2000, 343, 19–25. [Google Scholar] [CrossRef]
  407. Lai, S.-W.; Lin, H.-L.; Yu, T.L.; Lee, L.-P.; Weng, B.-J. Hydrogen release from ammonia borane embedded in mesoporous silica scaffolds: SBA-15 and MCM-41. Int. J. Hydrogen Energy 2012, 37, 14393–14404. [Google Scholar] [CrossRef]
  408. Feaver, A.; Sepehri, S.; Shamberger, P.; Stowe, A.; Autrey, T.; Cao, G. Coherent carbon cryogel−ammonia borane nanocomposites for H2 storage. J. Phys. Chem. B 2007, 111, 7469–7472. [Google Scholar] [CrossRef] [PubMed]
  409. Moussa, G.; Bernard, S.; Demirci, U.B.; Chiriac, R.; Miele, P. Room-temperature hydrogen release from activated carbon-confined ammonia borane. Int. J. Hydrogen Energy 2012, 37, 13437–13445. [Google Scholar] [CrossRef]
  410. Srinivas, G.; Travis, W.; Ford, J.; Wu, H.; Guo, Z.-X.; Yildirim, T. Nanoconfined ammonia borane in a flexible metal–organic framework Fe–MIL-53: Clean hydrogen release with fast kinetics. J. Mater. Chem. A 2013, 1, 4167–4172. [Google Scholar] [CrossRef]
  411. Chung, J.-Y.; Liao, C.-W.; Chang, Y.-W.; Chang, B.K.; Wang, H.; Li, J.; Wang, C.-Y. Influence of metal–organic framework porosity on hydrogen generation from nanoconfined ammonia borane. J. Phys. Chem. C 2017, 121, 27369–27378. [Google Scholar] [CrossRef]
  412. Ramachandran, P.V.; Raju, B.C.; Gagare, P.D. One-pot synthesis of ammonia-borane and trialkylamine-boranes from trimethyl borate. Org. Lett. 2012, 14, 6119–6121. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Convergence of hydrogen storage methods towards a sustainable hydrogen economy.
Figure 1. Convergence of hydrogen storage methods towards a sustainable hydrogen economy.
Materials 15 02286 g001
Figure 2. Theoretical (maximum) hydrogen storage capacity of metal borohydrides. Metal borohydrides are arranged by cation valence (I, II, III, and IV). Highlighted in blue are the borohydrides that exceed the DOE’s hydrogen storage capacity set for 2020 (4.5 wt %). Ln = La, Ce, Pr, Nd, Sm(II, III), Eu(II, III), Gd, Tb, Dy, Ho, Er, Tm, Yb(II, III), Lu; Ac = Th, Pa, U, Np, Pu; lt-isolable at low temperatures; in bold are shown borohydrides reported by several groups (higher confidence).
Figure 2. Theoretical (maximum) hydrogen storage capacity of metal borohydrides. Metal borohydrides are arranged by cation valence (I, II, III, and IV). Highlighted in blue are the borohydrides that exceed the DOE’s hydrogen storage capacity set for 2020 (4.5 wt %). Ln = La, Ce, Pr, Nd, Sm(II, III), Eu(II, III), Gd, Tb, Dy, Ho, Er, Tm, Yb(II, III), Lu; Ac = Th, Pa, U, Np, Pu; lt-isolable at low temperatures; in bold are shown borohydrides reported by several groups (higher confidence).
Materials 15 02286 g002
Figure 3. X-ray structure of lithium borohydride, LiBH4. Projections along a-axis (left), b-axis (middle), and c-axis (right).
Figure 3. X-ray structure of lithium borohydride, LiBH4. Projections along a-axis (left), b-axis (middle), and c-axis (right).
Materials 15 02286 g003
Figure 4. Physical vs. chemical hydrogen storage methods.
Figure 4. Physical vs. chemical hydrogen storage methods.
Materials 15 02286 g004
Figure 5. Complexity of ball-milling process, exemplified for synthesis of Zn(BH4)2 from ZnCl2 and MBH4 (M = Na, K) under various ratios of starting materials (1:1, 1:1.5, 1:2, and 1:2.5); THF-tetrahydrofuran, RT-room temperature, b.m - ball milling.
Figure 5. Complexity of ball-milling process, exemplified for synthesis of Zn(BH4)2 from ZnCl2 and MBH4 (M = Na, K) under various ratios of starting materials (1:1, 1:1.5, 1:2, and 1:2.5); THF-tetrahydrofuran, RT-room temperature, b.m - ball milling.
Materials 15 02286 g005
Figure 6. Soft nanocasting showing infiltration of MgnBu2 solution into SBA-15 mesoporous silica template to yield finally porous MgH2 nanoconfined inside SBA-15 pores (Dp~7.5 nm).
Figure 6. Soft nanocasting showing infiltration of MgnBu2 solution into SBA-15 mesoporous silica template to yield finally porous MgH2 nanoconfined inside SBA-15 pores (Dp~7.5 nm).
Materials 15 02286 g006
Figure 7. Two formulations (I [102] and II [159]) of possible solvate structures for ligand-coordinated metal alanates of select alkali metals (M: Li, Na).
Figure 7. Two formulations (I [102] and II [159]) of possible solvate structures for ligand-coordinated metal alanates of select alkali metals (M: Li, Na).
Materials 15 02286 g007
Figure 8. Structural diversity exhibited by known and characterized main group (alkali and alkali-earth) metal borohydrides. Crystal system for various borohydrides depicted as projections along “a” crystallographic axis. 1—LiBH4 (α-, LT polymorph, a-axis view); 2—LiBH4 (α-, HT polymorph, a-axis view); 3—KBH4; 4—RbBH4; 5—Be(BH4)2 (b-axis view); 6—Be(BH4)2 (c-axis view); 7α-Ca(BH4)2 (b-axis view); 8α-Ca(BH4)2 (c-axis view); 9α-Mg(BH4)2 (a-axis view); 10β-Mg(BH4)2 (a-axis view); 11γ-Mg(BH4)2 (c-axis view); 12—BaB12H12.
Figure 8. Structural diversity exhibited by known and characterized main group (alkali and alkali-earth) metal borohydrides. Crystal system for various borohydrides depicted as projections along “a” crystallographic axis. 1—LiBH4 (α-, LT polymorph, a-axis view); 2—LiBH4 (α-, HT polymorph, a-axis view); 3—KBH4; 4—RbBH4; 5—Be(BH4)2 (b-axis view); 6—Be(BH4)2 (c-axis view); 7α-Ca(BH4)2 (b-axis view); 8α-Ca(BH4)2 (c-axis view); 9α-Mg(BH4)2 (a-axis view); 10β-Mg(BH4)2 (a-axis view); 11γ-Mg(BH4)2 (c-axis view); 12—BaB12H12.
Materials 15 02286 g008
Figure 9. Selected examples of trivalent-, TM- (transition metal) and mixed-cation borohydrides. 1—Al(BH4)3 (a-axis view); 2—Y(BH4)3 (c-axis view); 3—Mn(BH4)2 (a-axis view); 4—Mn(BH4)2 (c-axis view); 5—LiSc(BH4)4 (a-axis view); 6—LiSc(BH4)4 (c-axis view); 7—NaZn2(BH4)5 (a-axis view); 8—NaZn2(BH4)5 (c-axis view). Color code: Sc—orange; Zn—blue.
Figure 9. Selected examples of trivalent-, TM- (transition metal) and mixed-cation borohydrides. 1—Al(BH4)3 (a-axis view); 2—Y(BH4)3 (c-axis view); 3—Mn(BH4)2 (a-axis view); 4—Mn(BH4)2 (c-axis view); 5—LiSc(BH4)4 (a-axis view); 6—LiSc(BH4)4 (c-axis view); 7—NaZn2(BH4)5 (a-axis view); 8—NaZn2(BH4)5 (c-axis view). Color code: Sc—orange; Zn—blue.
Materials 15 02286 g009
Figure 10. Overview of dehydrogenation pathways in Ca(BH4)2·4/3N2H4 (10.8 wt.% H2); the system is not yet reversible [252].
Figure 10. Overview of dehydrogenation pathways in Ca(BH4)2·4/3N2H4 (10.8 wt.% H2); the system is not yet reversible [252].
Materials 15 02286 g010
Figure 11. Boron–hydrogen speciation and influence on ionic conductivity of corresponding alkali and alkali-earth salts.
Figure 11. Boron–hydrogen speciation and influence on ionic conductivity of corresponding alkali and alkali-earth salts.
Materials 15 02286 g011
Figure 12. Decomposition temperature of M(BH4)x dependent on cation (*Pauling) electronegativity following a van’t Hoff equation plot. In green are highlighted elements forming borohydrides with predicted decomposition temperature 50–150 °C, suitable for mobile application. Very toxic elements are in red [295,296,297,298,299].
Figure 12. Decomposition temperature of M(BH4)x dependent on cation (*Pauling) electronegativity following a van’t Hoff equation plot. In green are highlighted elements forming borohydrides with predicted decomposition temperature 50–150 °C, suitable for mobile application. Very toxic elements are in red [295,296,297,298,299].
Materials 15 02286 g012
Figure 13. Formation enthalpies ΔH0f (ΔHboro, kJ/mol BH4) for metal borohydrides that are either reported or believed to exist (Figure 2).
Figure 13. Formation enthalpies ΔH0f (ΔHboro, kJ/mol BH4) for metal borohydrides that are either reported or believed to exist (Figure 2).
Materials 15 02286 g013
Figure 14. Dihydrogen bonding represented between borohydride [BH4] and amidic [NH2] moieties.
Figure 14. Dihydrogen bonding represented between borohydride [BH4] and amidic [NH2] moieties.
Materials 15 02286 g014
Figure 15. XRD data showing the evolution of phases in a Mo-catalyzed LiBH4-Mo:MSU-H mesoporous silica nanocomposite: commercial LiBH4 (A), rehydrogenated LiBH4 - Mo:MSU-H (B) and dehydrogenated LiBH4 Mo:MSU-H (C). Reprinted with permission from [335].
Figure 15. XRD data showing the evolution of phases in a Mo-catalyzed LiBH4-Mo:MSU-H mesoporous silica nanocomposite: commercial LiBH4 (A), rehydrogenated LiBH4 - Mo:MSU-H (B) and dehydrogenated LiBH4 Mo:MSU-H (C). Reprinted with permission from [335].
Materials 15 02286 g015
Figure 16. XRD diffraction spectra of MTBE-recrystallized LiBH4 and of three nanocomposites LiBH4-C-MSU- H (8 wt.%, 20 wt.%, and 40 wt.% LiBH4). Reprinted with permission from [334].
Figure 16. XRD diffraction spectra of MTBE-recrystallized LiBH4 and of three nanocomposites LiBH4-C-MSU- H (8 wt.%, 20 wt.%, and 40 wt.% LiBH4). Reprinted with permission from [334].
Materials 15 02286 g016
Figure 17. Hydrogen desorption kinetics of nanoconfined LiBH4@C-MSU-H with a heating ramp of 2 °C/min (initial nanocomposites, left; rehydrogenated samples, right). A: 8 wt.%, B: 20 wt.%, and C: 40 wt.% loading of LiBH4. Reprinted with permission from [334].
Figure 17. Hydrogen desorption kinetics of nanoconfined LiBH4@C-MSU-H with a heating ramp of 2 °C/min (initial nanocomposites, left; rehydrogenated samples, right). A: 8 wt.%, B: 20 wt.%, and C: 40 wt.% loading of LiBH4. Reprinted with permission from [334].
Materials 15 02286 g017
Figure 18. Direct re-hydrogenation of MgB2 under extreme conditions (950-bar H2, 450 °C) to regenerate Mg(BH4)2 material with ~11 wt.% reversible hydrogen capacity. Figure compiled from reaction data presented in [386].
Figure 18. Direct re-hydrogenation of MgB2 under extreme conditions (950-bar H2, 450 °C) to regenerate Mg(BH4)2 material with ~11 wt.% reversible hydrogen capacity. Figure compiled from reaction data presented in [386].
Materials 15 02286 g018
Figure 19. TPD curves of Ca(BH4)2@MC-a nanocomposites with a heating rate of 2.5 °C/min. Reprinted with permission from [143].
Figure 19. TPD curves of Ca(BH4)2@MC-a nanocomposites with a heating rate of 2.5 °C/min. Reprinted with permission from [143].
Materials 15 02286 g019
Figure 20. TPA curves of nanocomposite Ca.(BH4)/2 = MC 650-a for successive hydrogenation cycles at pressures in the range 20–45 atm. Reprinted with permission from [143].
Figure 20. TPA curves of nanocomposite Ca.(BH4)/2 = MC 650-a for successive hydrogenation cycles at pressures in the range 20–45 atm. Reprinted with permission from [143].
Materials 15 02286 g020
Table 1. Selected metal borohydrides with corresponding space group, decomposition temperature (and rehydrogenation temperature, where available) (°C), and crystal system.
Table 1. Selected metal borohydrides with corresponding space group, decomposition temperature (and rehydrogenation temperature, where available) (°C), and crystal system.
Metal BorohydrideDecomposition
Temperature (°C)
Rehydrogenation
Temperature (°C)
Space GroupCrystal
System
Reference
o-LiBH4(mp = 277); 600PnmaOrthorhombic[30,31,32,33,34,35,36,37,38,164]
h-LiBH4(107)600P63mcHexagonal[30,31,32,33,34,35,36,37,38]
α-NaBH4535270–400 (2NaH/
MgH2)
F m 3 ¯ m Cubic[165,166,167,168,169]
Be(BH4)2(mp = 91.3)
td = 123
Explosive
decomp. in air/moisture
I41/cdTetragonal;
helical
polymeric chains BeH2BH2BeH2BH2 and terminal bidentate BH4 groups
[170]
α-Mg(BH4)2
(6 known
polymorphs)
~300(390 °C, 90-bar H2, 72 h)
(400 °C, 80-bar D2)
RT stable
P6122Hexagonal[106,171]
β-Mg(BH4)2~300HT polymorph;
RT metastable
FdddOrthorhombic[172,173]
γ-Mg(BH4)2~300RT metastableIa3dCubic; 3D network of interpenetrated channels (1st borohydride with permanent porosity, S = 1505 m2/g)[158]
α-Ca(BH4)2
(4 known
polymorphs)
347–387 °C, 397–497 °C (2-step process)RT stableF2ddOrthorhombic[146,147,148,149,150,151]
β-Ca(BH4)2 α Ca ( BH 4 ) 2   167   ° C β-Ca(BH4)2RT, metastable, HT polymorph P 4 ¯ Tetragonal[146,147,148,149,150,151,157]
α-Al(BH4)3<100 −123C2/cMonoclinic[83,84,85,86,87,88]
β-Al(BH4)3<100−78Pna21Orthorhombic[83,84,85,86,87,88]
α-Y(BH4)3
(β-Y(BH4)3
is the HT polymorph)
190, 270 180 °C (α-to-β phase
transition)
P a 3 ¯ Cubic[67,68,69,70,71,72,73]
α-Mn(BH4)2
(4 known
polymorphs)
130RT stableP3112Hexagonal[174,175,176,177,178,179,180]
LiSc(BH4)4400(400), irreversible hydrogen
storage in Li-Sc-B-H system
P 4 ¯ 2cTetragonal[181,182,183]
NaZn(BH4)3>85 °C (at 110 °C NaBH4 reacts with Na2ZnCl4 to form Zn and NaCl)stable RTP21cMonoclinic[41,184]
NaZn2(BH4)5unstable, converts at low temp. to NaZn(BH4)3unstable at RT
or −32 °C
P21cMonoclinic[41,184]
Table 2. Ion conductivity σ M n +   ( S c m ) of representative alkali and alkali-earth metal borohydrides.
Table 2. Ion conductivity σ M n +   ( S c m ) of representative alkali and alkali-earth metal borohydrides.
Compound Ion   Conductivity   σ M n +   ( S c m ) Experimental Conditions/ObservationsReference
LiBH42 ∙ 10−3107 °C[261]
LiBH4·H2O 4.89 ∙ 10−4t = 45 °C[264]
LiBH4·x NH32.21 ∙ 10−3 0   <   x   <   2 ;   σ = σ ( x ) ; σ m a x @ 40   ° C [265,266]
LiBH4·x NH3@Li2O5.4 ∙ 10−40.67 < x < 0.8; 20 °C; max for 78 wt.% Li2O[267]
LiBH4·x NH3BH31.47 ∙ 10−5–4.04 ∙ 10−4½ < x < 1; highest for x = 1[268]
LiBH4·NH3BH310−1t = 55 °C; record value for Li+ [263]
Li2B12H1210−1110 °C[257]
Li2B10H103 ∙ 10−281 °C[258]
Na2B10H103 ∙ 10−2RT[258]
Na2B12H1210−1256 °C[259]
LiBH4-nano10−3RT; nanoconfined electrolyte[260]
Mg(BH4)2<10−12t = 30 °C[269]
Mg(BH4)(NH2)10−6t = 150 °C[270]
Mg3(BH4)4(NH2)24.1 ∙ 10−5100 °C [271]
Mg(BH4)2·en
en = NH2CH2CH2NH2
6 ∙ 10−570 °C, crucial role of [BH4], Mg(en)2X2 showed
σ very low
[268,272]
Mg(BH4)2·1/2dy
dy = diglyme
2 ∙ 10−580 °C; chelating of flexible dy ligand[273]
Mg(BH4)2·x NH33.3 ∙ 10−4x = 1,2,3 6; t = 80 °C; pas-de-deux mechanism proposed[274]
Mg(BH4)2·1.6NH32.2 ∙ 10−3t = 55 °C[275]
Table 3. Thermodynamic parameters characteristic to metal borohydrides: ΔHboro, Td, and td. In brown are shown elements with χ P 1.57 (ΔH > 0, unstable); decomposition temperature is given both as Td(K) and td(°C), for easier evaluation.
Table 3. Thermodynamic parameters characteristic to metal borohydrides: ΔHboro, Td, and td. In brown are shown elements with χ P 1.57 (ΔH > 0, unstable); decomposition temperature is given both as Td(K) and td(°C), for easier evaluation.
Metal (M)M(BH4)x χ P T d = 175.914 ( χ P 3.11 ) 2   [ K ] td [°C] Δ H b o r o = 248.7   χ P 390.8 [ kJ mol   BH 4 ] Reference
LiLiBH40.98798.1524.95−147.074[30,31,32,33,34,35,36,37,38,196,199]
NaNaBH40.93836.01562.86−159.509[165,168,169]
KKBH40.82922.51649.36−186.866[165,166,168]
CuCuBH41.9257.56−15.5981.73[113] forms at 253 K, dec.at 261–273 K
RbRbBH40.82 922.51649.36−186.866[167]
AgAgBH41.93244.94−28.2189.191[114]; forms at 193 K, dec. at 243 K
CsCsBH40.79946.84673.69−194.327[165]
AuIAuBH42.5457.15−216240.898[115] * unstable
AuIIIAu(BH4)32.5457.15−216240.898[116]
BeBe(BH4)21.57417.2144.05−0.341[170]
MgMg(BH4)21.31569.96296.81−65.003[106,158,171,172,173,189,190,191,192,193,194,195]
CaCa(BH4)21783.19510.04−142.1[146,147,148,149,150,151,157]
CrCr(BH4)21.66369.8696.7122.042[28,29]
MnMn(BH4)21.55428.1154.95−5.315[175,189,190,191,224]
FeFe(BH4)21.83288.2215.0764.321[192,193]
NiNi(BH4)21.91253.32−19.8384.217[195] * as heteroleptic complex
CoCo(BH4)21.88266.14−7.0176.756[194] * presumed, but unstable
ZnZn(BH4)21.65374.98101.8319.555[119,120,121,122]
SrSr(BH4)20.95820.74547.59−154.535[147,148]
BaBa(BH4)20.89866.97593.82−169.457[148]
CdCd(BH4)21.69354.7181.5629.503[149]
HgHg(BH4)22216.74−56.41106.6[151] * failed attempts
AlAl(BH4)31.61395.81122.669.607[83,84,85,86,87,88]
ScSc(BH4)31.36538.74265.59−52.568[182,183,295,296,297,298,299]
TiIIITi(BH4)31.54433.61160.46−7.802[176,177,178,179,206]
* Ti(IV) is reduced in situ to Ti(III)
VV(BH4)31.63385.32112.1714.581[66]
GaGa(BH4)31.81297.2924.1459.347[84,85] * as GdH(BH4)2 (volatile, 203K) and GdH2(BH4) (unstable)
YY(BH4)31.22628.38355.23−87.386[65,66,68,69,70,71,72,73]
NbNb(BH4)31.6401.1127.957.12[180] * no homoleptic form; as complex
InIn(BH4)31.78311.1738.0251.886[86]
LaLa(BH4)31.1710.71437.56−117.23[212]
CeCe(BH4)31.12696.64423.49−112.256[212]
NdNd(BH4)31.14682.7409.55−107.282[69]
SmIISm(BH4)21.13689.65416.5−109.769[213,214]
SmIIISm(BH4)31.13689.65416.5−109.769[214]
Pr Pr(BH4)31.13689.65416.5−109.769[64]
EuIIEu(BH4)21.2641.75368.6−92.36[213]
EuIIIEu(BH4)31.2641.75368.6−92.36[70,148]
GdGd(BH4)31.2641.75368.6−92.36[68,71,72,214]
TbTb(BH4)31.22628.38355.23−87.386[73,214]
DyDy(BH4)31.23621.75348.6−84.899[68]
HoHo(BH4)41.24615.15342−82.412[69]
ErEr(BH4)31.24615.15342−82.412[214]
TmTm(BH4)31.25608.59335.44−79.925[73]
YbIIYb(BH4)21.1710.71437.56−117.23[215]
YbIIIYb(BH4)31.1710.71437.56−117.23[215]
LuLu(BH4)31.27595.57322.42−74.951[73]
ThTh(BH4)41.3576.31303.16−67.49[78,206,207]
PaPa(BH4)31.5455.99182.84−17.75[78]
UU(BH4)41.38526.49253.34−47.594[78,80,208,209]
NpNp(BH4)41.36538.74265.59−52.568[79,80,210]
PuIIIPu(BH4)31.28589.12315.97−72.464[78,80]
PuIVPu(BH4)4 [79]
TlITlBH41.62390.55117.412.094[87]
TlIIITl(BH4)31.62390.55117.412.094[88] * failed attempts; as TlCl(BH4)2 (dec. at 178 K)
GeGe(BH4)42.01212.86−60.29109.087no conclusive reports
ZrZr(BH4)41.33557.37284.22−60.029[203,204]
SnSn(BH4)41.96232.65−40.596.652no conclusive reports
HfHf(BH4)41.3576.31303.16−67.49[205]
Table 5. Kinetic parameters for dehydrogenation of binary destabilized systems Ca(BH4)2-MII(NH2)2 (M = Mg, Ca).
Table 5. Kinetic parameters for dehydrogenation of binary destabilized systems Ca(BH4)2-MII(NH2)2 (M = Mg, Ca).
Binary SystemEa (kJ/mol)A × 10−9 (min−1)k × 102 (min−1)Reference
Ca(BH4)2-2Mg(NH2)2132.7151.2[309]
Ca(BH4)2-2Ca(NH2)2119.33.24.3[309]
Table 6. DSC analysis data for RF-CA-nanoconfined 2LiBH4-MgH2 vs. 2LiBH4-MgH2-bulk [346].
Table 6. DSC analysis data for RF-CA-nanoconfined 2LiBH4-MgH2 vs. 2LiBH4-MgH2-bulk [346].
CompoundDSC Peak 1; Polymorphic Transformation
o-LiBH4h-LiBH4
DSC Peak 2; Melting PointDSC Peak 3; DehyDrogenation
(MgH2)
DSC Peak 4; DehyDrogenation
(LiBH4)
2LiBH4-MgH2@RF-CA21nm113 °C267 °C332 °C351 °C
2LiBH4-MgH2-bulk117 °C290 °C364 °C462 °C
Table 7. Nanoconfinement of ammonia borane in nanoporous silica (MCM-41, SBA-15) or carbon (CMK-3 OMC, RF-CC).
Table 7. Nanoconfinement of ammonia borane in nanoporous silica (MCM-41, SBA-15) or carbon (CMK-3 OMC, RF-CC).
SubstanceScaffoldScaffold PoreLoading (wt.%)td (°C)Ea (kJ/mol)ΔH
(kJ/mol)
Ref.
NH3BH3--100114 (110), 155184−19.6 to −21[145,406]
NH3BH3MCM-414 nm33, 50, 75 (outside pores)100, 182N/A34.09 (2 steps: 30.33 and 3.76)[404,407]
NH3BH3SBA-157.5 nm5050–100, 13067−1[145]
NH3BH3OMC4.5 nm (50-)95 −2.1[402]
NH3BH3RF-CC (carbon cryogel)2–20 (mean 5) nm2485–150; 80–90 for CC(exo- peak)N/A−120 (CC)[403,408]
NH3BH3AC (activated carbon)3 nm17.6 wt.% (computed from pore volume decrease from 0.36 cm3/g to 0.14 cm3/g for AC@ABTHF)3–4; RT(25)N/A−1.6 (AC@ABTHF), −3.0 (AC@ABMET) and −5.9 (AC@ABmilled)[409]
NH3BH3MOF (Fe–MIL-53)11–13 nmAB:Fe = 6.5% (0.5:1 molar); 13% (1:1 molar)80–110130 +/− 7;
135 +/−3
N/A[410]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Comanescu, C. Complex Metal Borohydrides: From Laboratory Oddities to Prime Candidates in Energy Storage Applications. Materials 2022, 15, 2286. https://doi.org/10.3390/ma15062286

AMA Style

Comanescu C. Complex Metal Borohydrides: From Laboratory Oddities to Prime Candidates in Energy Storage Applications. Materials. 2022; 15(6):2286. https://doi.org/10.3390/ma15062286

Chicago/Turabian Style

Comanescu, Cezar. 2022. "Complex Metal Borohydrides: From Laboratory Oddities to Prime Candidates in Energy Storage Applications" Materials 15, no. 6: 2286. https://doi.org/10.3390/ma15062286

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