Next Article in Journal
Imaging-Guided Delivery of a Hydrophilic Drug to Eukaryotic Cells Based on Its Hydrophobic Ion Pairing with Poly(hexamethylene guanidine) in a Maleated Chitosan Carrier
Next Article in Special Issue
Hydrolysis of the Borohydride Anion BH4: A 11B NMR Study Showing the Formation of Short-Living Reaction Intermediates including BH3OH
Previous Article in Journal
Current Progress in the Development of Hepatitis B Virus Capsid Assembly Modulators: Chemical Structure, Mode-of-Action and Efficacy
Previous Article in Special Issue
Synthesis of Zwitter-Ionic Conjugate of Nido-Carborane with Cholesterol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Boron Hydrogen Compounds: Hydrogen Storage and Battery Applications

Département de Chimie Physique, Université de Genève, 30, Quai E. Ansermet, CH1211 Geneva 4, Switzerland
Molecules 2021, 26(24), 7425; https://doi.org/10.3390/molecules26247425
Submission received: 15 October 2021 / Revised: 29 November 2021 / Accepted: 2 December 2021 / Published: 7 December 2021

Abstract

:
About 25 years ago, Bogdanovic and Schwickardi (B. Bogdanovic, M. Schwickardi: J. Alloys Compd. 1–9, 253 (1997) discovered the catalyzed release of hydrogen from NaAlH4. This discovery stimulated a vast research effort on light hydrides as hydrogen storage materials, in particular boron hydrogen compounds. Mg(BH4)2, with a hydrogen content of 14.9 wt %, has been extensively studied, and recent results shed new light on intermediate species formed during dehydrogenation. The chemistry of B3H8, which is an important intermediate between BH4 and B12H122−, is presented in detail. The discovery of high ionic conductivity in the high-temperature phases of LiBH4 and Na2B12H12 opened a new research direction. The high chemical and electrochemical stability of closo-hydroborates has stimulated new research for their applications in batteries. Very recently, an all-solid-state 4 V Na battery prototype using a Na4(CB11H12)2(B12H12) solid electrolyte has been demonstrated. In this review, we present the current knowledge of possible reaction pathways involved in the successive hydrogen release reactions from BH4 to B12H122−, and a discussion of relevant necessary properties for high-ionic-conduction materials.

1. Introduction

Boron hydrogen compounds have been intensively studied for almost a century since the pioneering studies of A. Stock [1]. Boron hydrogen compounds are also energetic materials and were considered as rocket or jet fuels [2]; however, the toxicity of boranes has prevented their extended use. Currently, nontoxic compounds such as ammonia-borane are also studied as hypergolic propellants [3,4]. Recently, many different applications of boron hydrogen compounds have emerged [5]. In particular, compounds derived from closo-hydroborates such as B12H122− have found many new applications, including new all-solid-state batteries, medical applications, and as catalysts [6,7,8,9,10,11]. Since the discovery of catalyzed hydrogen release in NaAlH4 by Bogdanovic and Schwickardi [12], light boron and aluminum hydrides were intensively studied and reviewed as potential hydrogen storage materials [13,14,15,16,17,18,19,20,21,22]. The dehydrogenation reactions of metal borohydrides ultimately lead to hydrogen, metal and boron, or metal borides. In this reaction process, intermediate species are formed, particularly compounds with closo-hydroborate anion B12H122− [23,24]. B12H122− is particularly stable and can thus also act as a detrimental thermodynamic sink for further dehydrogenation reactions. The properties of closo-hydroborates and related anions were addressed in several recent publications [6,25,26,27,28]. New research on the thermal properties of closo-hydroborate salts revealed a high-temperature phase transition in Na2B12H12 leading to a superionic phase [29]. Thus, the controlled dehydrogenation of a borohydride salt can be used to safely prepare new closo- and nido- hydroborate salts for potential battery applications [30] without using toxic boranes such as B10H14, which were used for the synthesis of this large boron species [31].
In this review, we first describe experimental results on hydrogen storage in Mg(BH4)2, which has a large hydrogen content of 14.9 wt %. Hydrogen storage in other borohydrides, such as LiBH4, was recently reviewed [32]. Recent results on potential dehydrogenation intermediates for Mg(BH4)2 provide new insights on the potential reaction intermediates and are reported here. In this context, we then present recent results based on DFT calculations to explore possible reaction paths for successive dehydrogenation reactions starting from BH4. These paths are described in more detail in the following section, which discusses the formation and reactions of B3H8, as this ion is considered to be one of the reaction intermediates during the dehydrogenation of borohydride compounds. The high-temperature dehydrogenation of B3H8 leads to the formation of closo-hydroborate anions B10H102− and B12H122−, which form excellent solid ionic conductors for new all-solid-state batteries [30]. The properties of these ionic conductors are presented in the last section.

2. Magnesium Borohydride

Among the many compounds considered for hydrogen storage, Mg(BH4)2 is particularly interesting and has been studied by many authors. The earlier studies on Mg(BH4)2 were reviewed in detail in 2016 [22]. Mg(BH4)2 has a hydrogen content of 14.9 mass % [22,33]. This compound can be prepared in different crystalline modifications, and high pressure-phase transitions were also observed [33]. Porous γ-Mg(BH4)2 can also adsorb 0.8 H2 at low temperatures and 100 bar to achieve a total hydrogen mass content of 17.4% [33]. High-pressure phase δ-Mg(BH4)2 has a very high volumetric hydrogen content of 147 g H2/L. Mg(BH4)2 can also form amorphous solids. Overall dehydrogenation reaction
Mg(BH4)2 ➔ MgB2 + 4 H2
is, in fact, a multistep reaction (see Figure 1) with various reaction intermediates, such as Mg(B3H8)2, MgH2, and MgB12H12, which were proposed both experimentally and theoretically [22,34,35,36]. MgB2 is the decomposition product obtained after heating to 500 °C [37]. Boron-rich MgB7 films are obtained by heating volatile Mg(B3H8)2 solvates with dimethyl ether and diethyl ether [38].
MgB2 can be rehydrogenated, although under drastic conditions (950 bar H2 at 400 °C) [40]. The rehydrogenation of MgB2 can be accelerated with THF, MgH2, and Mg [41]. Mechanically milled mixtures of MgB2, THF, and 40 mol % Mg could thus absorb 6 wt % of H2 at 300 °C under 700 bar of H2, which is less drastic than that without THF. Recently, rehydrogenation at room temperature with mechanical activation by ball milling was reported [42]. These rehydrogenation reactions of MgB2 demonstrate the principle that hydrogen storage in Mg(BH4)2 is indeed reversible. A recent combined experimental and theoretical study concluded that the initial stages of rehydrogenation are associated with the formation of σ bonds of hydrogen with boron on the reactive edges of the MgB2 solid [43]. The rehydrogenation of intermediate compounds was also studied. MgB3H8.THF can be rehydrogenated under milder conditions than those of dry MgB3H8 (50 bar H2 and 200 °C for 5 h vs. 120 bar H2 and 250 °C for 48 h) [44]. MgH2 is formed in intermediate reaction steps, such as
6 Mg(BH4)2 ➔ MgB12H12 + 5 MgH2 + 13 H2
Magnesium hydride dissociates into Mg and H2 at high temperatures and low H2 pressures. The different reaction products observed under various conditions (see Figure 1) show that the reaction kinetics can be influenced by various parameters, which also include the initial crystalline modification of Mg(BH4)2.
The overall enthalpy of reaction for the dehydrogenation of Mg(BH4)2fH° = −208 kJ/mol) to form MgB2fH° = −91.96 kJ/mol) and hydrogen can be calculated [45,46,47] to be equal to +116 kJ/mol, i.e., less than 30 kJ/mol per hydrogen molecule released, which is, in principle, in the correct range for a hydrogen storage material [13]. The first step of a dehydrogenation reaction of BH4 is likely to be the breaking of a B–H bond. Isotope exchange reactions of Mg(BH4)2 with D2 allow for producing a complete exchange to form Mg(BD4)2, and the corresponding activation energy was estimated to be about 51 kJ/mol [48]. For Ca(BH4)2, the corresponding activation energy was found to be 82 and 98.5 kJ/mol for the reverse reaction, confirming that breaking a bond with hydrogen or deuterium is the rate-limiting step [49]. Theoretical calculations of potential defects in Mg(BH4)2 suggest that, in the initial phase of the dehydrogenation, a H ion is formed that can diffuse in the lattice [50]. On the other hand, gas diffusion in the solid is also a contribution to exchange kinetics, as was shown by isotope exchange reactions with the highly porous modification of γ-Mg(BH4)2 with a high surface area compared to a ball-milled sample with a strongly reduced surface aera [51].
The reaction kinetics of hydrogen release in Mg(BH4)2 can be significantly enhanced by various additives, such as TiCl3 [52] or NbF5 and TiO2 [53]. Lewis bases in the form of solvates of Mg(BH4)2 can also accelerate the hydrogen release [54]. As shown in Figure 1, the THF solvate releases H2 gas below 200 °C to form Mg(B10H10). The formation of B3H8 and B12H122− was also observed, but with THF and dimethyl ether, B12H122− remained a minor reaction product. The physical properties of Mg(BH4)2.3THF were recently investigated in detail [55]. In this compound, Mg2+ is coordinated to 2 BH4 ions and 3 THF molecules. The orientational mobilities of the BH4 ions are not particularly sensitive to the presence of THF. The authors concluded that “the presence of THF also disrupts the stability of the crystalline phase leading to enhanced kinetics for the dehydrogenations”. Recently, Tran et al. [56] reported that the presence of different glymes with Mg(BH4)2 results in various ratios of MgB10H10 and MgB12H12 upon thermolysis at 160–200 °C, and allows for selectively obtaining MgB10H10 with one equivalent of monoglyme. Mixtures of Mg(BH4)2 with (CH3)4NBH4 (5:1 molar) reveal reversible melting around 180–195 °C [57] with enhanced stability compared to melts of pure Mg(BH4)2 and (CH3)4NBH4. [Ph4P]2[Mg(BH4)4] gradually loses mass over 225–230 °C, but heating to 500 °C does not lead to the mass loss expected for the formation of MgB2. A similar behavior was observed for [Me4N]2[Mg(BH4)4] [58]. These findings suggest that derivates of Mg(BH4)2 with organic cations are rather stabilized.
Solvent-free Mg(B3H8)2 can be prepared by ball milling MgBr2 with NaB3H8 [38,59]. Kim et al. [38] reported the formation of boron-rich MgB7 films upon heating under vacuum above 425 °C due to some evaporation of Mg under these conditions. Thermogravimetry (TG) experiments [59] revealed a 30 wt % mass loss setting in above ca 80 °C corresponding to the evolution of B2H6, B5H9 and H2. The residual solid after heating to 200 °C was a mixture of mainly Mg(BH4)2, Mg(B10H10), and Mg(B12H12), and the combined evolution of H2, B2H6, and B5H9 was confirmed by mass spectrometry [60]. The addition of activated (ball-milled) MgH2 to Mg(B3H8)2 results in a strong reduction in borane evolution and up to 88% conversion back to Mg(BH4)2 at 100 °C. The presence of activated MgH2 thus substantially decreases the formation of (closo-hydro)borates and provides the necessary hydrogen for the conversion of B3H8 back into BH4.
These experiments suggest that, while Lewis acids may favor the dehydrogenation reactions of Mg(BH4)2, they do not necessarily catalyze the rehydrogenation reactions, as transition metal halides do not appear to affect the rehydrogenation of MgB2 [40,61]. THF and other Lewis bases appear to accelerate both the dehydrogenation and rehydrogenation reactions of Mg(BH4)2, and encourage more studies to even further improve the kinetics.

3. DFT Calculations

The results presented above for Mg(BH4)2 suggest the formation of various intermediate species such as B2H62−, B3H8, B4H102−, B5H92− and the closo-borates BnHn2− (n = 8–12). For hydrogen storage applications, the only gaseous species resulting from dehydrogenation reactions should be hydrogen; thus, neutral boranes are a priori not involved in the reaction mechanisms. Many other anionic boron hydrides have been reported in the literature and could be involved in one reaction step or another. In 1999, some reactions between neutral and anionic boron hydrides related to the formation of B3H8, B5 anions, and some other species were reviewed [62].
In order to assess the driving forces for different reactions, thermodynamic information can be very useful, but experimental data are very scarce. For alkali borohydrides, thermodynamical data are available [47], but only few other experimental data are available. Using the experimental values of the formation enthalpy of Mg(BH4)2 [45] and La(BH4)3 [63], the formation enthalpy of other M(BH4)2 and M(BH4)3 compounds were estimated, assuming that the lattice enthalpy of bromides and borohydrides with the same metal ion were identical within about 15 kJ/mol [46]. The experimental formation enthalpy of NH4B3H8 (−530 ± 33 kJ/mol) [64], (NH4)2B10H10 (−359.2 ± 10 kJ/mol) [65], and of guanidinium and other nitrogen-based closo-borates was reported [66]. Recently, new heat capacity measurements for Na, K, Rb, Cs, Mg, Ca borohydrides were reported [67]. The knowledge of all thermodynamic properties in principle allows for quantitatively describing the phase diagram of a system, which was performed using available data for the Mg–B–H system [68].
In the absence of experimental data, theoretical data are obtained. It is quite challenging to obtain accurate results of formation enthalpies using DFT. Nguyen et al. [69] calculated for the formation enthalpy of (NH4)2B10H10 with the G3 method the value of −184 kJ/mol, which is quite different from the experimental value of −359.2 kJ/mol. For α-Mg(BH4)2, formation enthalpy values ranging from −67 to −277 kJ/mol were reported in the literature [68], while the experimental value was −208 kJ/mol [45]. Zhang et al. [23] computed relative formation energies of potential solid intermediates formed during the dehydrogenation of Mg(BH4)2, in combination with a Monte Carlo-based structure prediction method. They predicted a potential Mg3(B3H6)2 intermediate with a B3H63− ion, while Mg(B3H8)2 was found to be very high in relative energy and thereby unlikely to be formed.
The principal difficulty for estimating the formation enthalpy of crystalline solids is the evaluation of lattice energy, as different approaches (volume-based, Kaputinski equation etc.) lead to different values. Further, lattice energies can only be computed for crystalline materials, preferentially on the basis of experimental structure data, but experiments showed that a significant fraction of the reaction intermediates remain amorphous, complicating things even further.
DFT calculations in the gas phase are quite reliable, and allow for obtaining good structural data and vibrational frequencies, in particular when anharmonicity is included. Several studies report the formation enthalpy of borohydride ions in the gas phase [69,70,71,72] Anharmonic DFT calculations allow for obtaining improved agreement with experimental vibrational spectra, from which heat capacity data were calculated [73]. Figure 2 compares experimental [74] and DFT calculated [69,70,71,72] formation enthalpy data for neutral and anionic boron hydrogen species. Figure 2 shows that the calculated formation enthalpy for a given species (e.g., B3H8) can differ by about 100 kJ/mole for different sources. These values are derived, for instance, from isodesmic reactions with known formation heat [69], thus generating a potential propagation of errors if the initial formation enthalpy values are different. We outline all reported values to highlight the limitations of the accuracy of these data.
Figure 2 shows that the experimental formation enthalpies of neutral species are all positive [74], with values ranging from 36 kJ/mol (for B2H6) to 210 kJ/mol for (B2H4). Gas phase reaction
2 B2H6 ➔ B4H10 + H2
has an enthalpy change of 66.1 − 2 × 36.4 = −6.7 kJ/mol, and shows that increasing the number of boron atoms in the cluster can be thermodynamically favorable for neutral species. Other reactions towards larger hydroboranes may become favorable at higher temperatures from the liberation of hydrogen. The first theoretical studies of enthalpy changes for reactions of neutral boranes were reported by M.L. McKee in 1990 [70], who showed that a sequence of BH3 additions followed by H2 elimination from B2H6 to B6H10 is overall exothermic, but with two less stable reaction intermediates (B3H9 and B4H8) that can act as barrier steps for the kinetics. Figure 2 shows that anionic species with 9–12 boron atoms are the most stable, which indicates that there is a thermodynamic driving force towards these anions. The most stable species in this figure is the closo B12H122− ion, and its stability is related to its 3-dimensional aromaticity [6]. The formation enthalpy of B12H122− in the gas phase was estimated to be between −325.5 and −428.6 kJ/mol according to different theoretical studies [72,75,76]. One key intermediate in the overall dehydrogenation reactions of BH4 appears to be ion B3H8, which is discussed in the next section.

4. Formation and Reactions of B3H8

As mentioned above, the formation of Mg(B3H8)2 was observed during the decomposition of Mg(BH4)2 under dynamic vacuum [54,77], and Y(B3H8)3 was obtained after heating Y(BH4)3 under hydrogen pressure of 1–10 bar [78]. There are several reports in the literature on the synthesis of B3H8 that highlight that various routes can lead to this ion. Starting from diborane under strongly reducing conditions, dianion B2H62− was reported to form [62,79]
2 B2H6 + 2 C8H10 ➔ [BH32−] + BH3 + 2 C8H10 ➔ [B2H62−] + 2 C8H10
BH32− and B2H62− intermediates were identified by NMR. The reaction of B2H62− with additional diborane yields B3H8 + BH4, and no further intermediate was observed:
B2H6 + B2H62− ➔ B3H8 + BH4
Another reaction observed was the reaction of potassium metal with THF.BH3 [80].
2 K + 4 THF.BH3 ➔ 2 K+ + B3H8 + BH4
Beall and Gaines [62] argue that also in this case, B2H62− is the reaction intermediate, which can then react with THF–BH3 to form either B2H5 + BH4 with the addition of the 4th THF.BH3 B3H8 or first with THF–BH3 the ion B3H92−, which then reacts with THF.BH3 to yield B3H8 + BH4. B3H8 can also be formed from the reaction of BH4 with diborane [81]:
BH4 + B2H6 ➔ B3H8 + H2
BH4 + B2H6 ➔ BH3 + B2H7 ➔ BH3 + B2H5 + H2 ➔ B3H8 + H2
BH4 + B2H6 ➔ BH3 + B2H7 ➔ B3H10 ➔ B3H8 + H2
This reaction can proceed either via B2H7 (hydride transfer) and B3H10 (BH3 addition) followed by H2 detachment or via B2H7, which first loses H2 to form B2H5, which then adds BH3. The efficient synthesis of alkali metal octahydrotriborates (M = Na, K, Rb, Cs) from the reaction of MBH4 with 2 equivalents of dimethyl sulfide borane was reported [82]. The formation of ion B2H7 was observed by NMR for the reaction of LiBH4 with THF.BH3 in THF [83], and during the solvothermal reaction of BH4 with CH2Cl2 at 70 °C [84]. The reaction of BD4 requires higher temperatures (90 °C) [84], which suggests that the rate-determining reaction step is associated with the breaking of a boron–hydrogen (deuterium) bond, which could be the formation of a reactive Lewis adduct of BH3 from BH4, which then reacts with other BH4 to form B2H7 etc., as outlined above.
Once formed, B3H8 can further react to yield B9 to B12 hydroborate anions. Using the DFT calculation formation enthalpies of B9H14, B3H8 and BH4 [71], for the gas phase reactions, one obtains
4 B3H8 ➔ B9H14 + 3 BH4 + 3 H2
4 B3H8 ➔ B10H102− + 2 BH4 + 9 H2
exothermic reaction enthalpy values of −413 and −49.8 kJ/mol, respectively, and a strong entropy increase that even further favors the reaction at higher temperatures. These spontaneous overall reaction enthalpies also explain why potential reaction intermediates with 6 to 8 boron atoms are practically not observed. The simultaneous production of BH4 in these reactions adds a thermodynamic driving force (as the formation enthalpy of BH4 is negative) for these reactions.
In the presence of hydrides, Grinderslev et al. [85] observed the following decomposition reaction at 150 and 200 °C of KB3H8 under 380 bar of H2:
KB3H8 + 2KH ➔ KBH4 + K2B12H12 + K2B10H10 + K2B9H9
As shown above, heating solvent-free Mg(B3H8)2 + 4 MgH2 either with or without H2 gas results in up to 88% back conversion to Mg(BH4)2 with some MgB12H12 [60]. These results show that B3H8 can react in many different ways to either form larger boron hydride clusters or regenerate BH4. This can be exploited, for instance, to achieve the direct synthesis of B10H102− and B12H122− to prepare solid ionic conductors such as Na4(B10H10)(B12H12), as demonstrated by Gigante et al. [86]. This synthesis starts with the conversion of NaBH4 into (Et4N)BH4, which reacts solvothermally with CH2Cl2 to form (Et4N)B3H8. (Et4N)B3H8 is then heated in toluene to 185 °C to form a mixture of (Et4N)2B10H10 and (Et4N)2B12H12, which can then either be separated by fractional crystallization or directly converted with sodium tetraphenylborate into ionic conductor Na4(B10H10)(B12H12).

5. Closoborates and Related Species as Solid Ionic Conductors

Solid ionic conductors for lithium or sodium batteries allow for avoiding the use of a flammable organic electrolyte and are thus expected to considerably improve the safety of batteries. A good solid electrolyte must fulfill several empirical criteria, according to [87]:
-
“open structure” with a low coordination number of the mobile ion;
-
The presence of structural phase transitions at low pressure. In the case of AgI, the ambient pressure wurtzite structure (space group P63mc) transforms at 3 kbar and 315 K into a NaCl structure (space group Fm-3m), thus going from a rather covalent network with coordination number 4 to a rather ionic structure with coordination number 6. The associated charge fluctuations between ions can potentially be coupled to vibrational motions and thus dynamically favor ionic conduction.
For practical applications, the conductivity of the material should be higher than 1 mS/cm. Further, the material should have high chemical and thermal stability, and a high electrochemical stability window. Additionally, it must be electronically insulating to avoid battery self-discharge or shortage. Further, the electrolyte should be deformable in order to accommodate the volume changes of anode and cathode materials upon lithium or sodium insertion and removal. This can thus limit the formation of fractures that reduce the performance of the battery. Lastly, the material should not be toxic and be cheap enough for the considered applications.
The discovery of superionic conductivity in the high-temperature phases of LiBH4 [88] and Na2B12H12 [29] has stimulated new research for similar compounds with high ionic conductivity at lower temperatures. These compounds include closo-hydroborates, nido-hydroborates (B11H14), and closo-hydrocarborates (CB9H10, CB11H12). Ions B10H102− and B12H122− are not very toxic. Mutterties et al. [89] reported LD50 values for Na2B10H10 and Na2B12H12 administered orally to rats to be around or higher than 7.5 g/kg of body weight for both compounds.
The crystal chemistry of inorganic hydroborates except BH4 was recently presented in detail [90], while the crystal chemistry of salts with BH4 was addressed in an earlier review [18]: “All nonmolecular hydroborate crystal structures can be derived by simple deformation of the close-packed anionic lattices, i.e., cubic close packing (ccp) and hexagonal close packing (hcp), or bodycentered cubic (bcc), by filling tetrahedral or octahedral sites” [90]. This observation can be illustrated considering group–subgroup relationships of encountered crystal structures, as illustrated in Figure 3 for some relevant compounds [90,91,92,93,94,95,96,97,98,99,100,101,102]. Crystal packing is governed by large anions, leaving in some space groups empty cationic sites, which, of course, favor ionic conduction. For instance, β-Na2B12H12 crystallizes in the Pm-3n space group with a statistical population of 6 sites occupied by 4 Na+ ions.
Perturbations of the anionic sublattice further allow for stabilizing the conductive phase at lower temperatures. This was first demonstrated for solid solutions of LiBH4 with LiBr and LiI [103]. Phase stability and ionic conductivity in mixed LiBH4–LiX (X = Cl, Br) was recently studied in detail [104]. Perturbations of the structure by ball milling or partial substitution was demonstrated for Na2B12H12 with a partial introduction of iodine ions in the closo-hydroborate [105]. In a further step, solid solutions of closo-hydroborate and closo-carbahydroborates, and solid solutions of nido-hydroborates with closo-hydroborates were studied [106,107,108,109,110,111,112]. Representative examples of mixed borate ionic conductors are shown in Table 1.
The mechanism of ionic conduction in these compounds is related to the dynamical properties of the borohydride or carbohydride ions. These properties can be addressed using NMR [113] and neutron techniques [114], in conjunction with temperature-dependent conductivity and X-ray diffraction, and are supported by theoretical calculations [76,77,88]. A detailed study of ionic conductor Na4(B12H12)(B10H10) [115] with all these techniques revealed 3 different regimes with increasing temperature. Below −50 °C, conductivity remains very low. Above this temperature, an apparent activation energy of 0.6 eV was found, related to significant couplings of anionic and cationic motions. Above 70 °C, activation energy decreases to 0.37 eV, as thermal energy leads to noncorrelated ionic motions.
One important aspect of solid ionic conductors is their electrochemical stability, which is a critical limit for a reversible battery application. Asakura et al. [116] developed a linear sweep voltammetry method to reliably measure the electrochemical stability of borohydride-based solid electrolytes. The measured oxidative stability of LiBH4 of 2.0 V vs. Li+/Li was significantly smaller than that in initial reports claiming a stability of up to 5 V [117]. For Na4(B12H12)(B10H10), two oxidation onsets at 3.02 and 3.22 V vs. Na+/Na were tentatively assigned to the onset of decomposition of the less stable [B10H10]2− and more stable [B12H12]2− ions, respectively [116]. Closo-carborane ions are even more stable, as for Na4(CB11H12)2(B12H12), where a large anodic current was observed above 4 V vs. Na+/Na, together with a small onset at 2.93 V. For Li2(CB9H10)(CB11H12), the onset of decomposition was observed at 2.86 V vs. Li+/Li [116]. Nido-borates are electrochemically less stable. The oxidative stability limit for Na5(B11H14)(B12H12)2 was 2.6 V vs. Na+/Na, and for LiB11H14, 2.6 V vs. Li+/Li [107].
These developments have also led to several all-solid-state battery prototypes based on these mixed borate ionic conductors. Duchêne et al. [118] presented a 3 V sodium battery using Na4(B12H12)(B10H10), and Murgia et al. [119] showed Na stripping/plating over >500 h in a Na cell with Na4(CB11H12)2(B12H12). Recently, Asakura et al. [120] demonstrated a 4 V sodium battery with the same solid-state conductor, Na4(CB11H12)2(B12H12). These results show that closo-hydroborates and their derivatives are very promising materials for chemically and electrochemically stable all-solid-state ionic conductors.

6. Conclusions

In the last 20 years, many studies on borohydride species have considerably increased our knowledge on the properties of these materials. For hydrogen storage applications, the kinetics and reversibility of the dehydrogenation reactions remain a major challenge for practical applications. The chemistry of borohydrides from BH4 to B12H122− in the gas phase and in solution has been theoretically and experimentally addressed; however, in solids, these studies are very challenging, as structural data of potential reaction intermediates such as Mg(B3H8)2 are elusive, and not all intermediates can be observed. If the reaction intermediates are amorphous, X-ray diffraction cannot be used, and theoretical approaches can lead to many different potential structures. The presence of additional hydrides or of Lewis bases such as THF, as shown for the reactions of KB3H8 and Mg(B3H8)2, strongly modifies the reaction products upon heating. We are thus still very far from a full microscopic understanding of these hydrogenation–dehydrogenation reactions and in the search for optimal catalysts for these processes.
For hydrogen storage, B3H8 is an interesting species that can be rehydrogenated back to BH4. Even though only 25% of the hydrogen is available for this reversible hydrogen storage, the temperatures (less than 200 °C) and kinetics of these reactions approach practical conditions.
The closo-hydroborate ions that are formed and identified as intermediates of dehydrogenation reactions have found new and very promising applications as solid-state ionic conductors, as they present many very favorable properties for this use. The recent demonstration of a 4 V all-solid-state battery using solid sodium electrolyte Na4(CB11H12)2(B12H12) [120] highlights this potential. Whether compounds such as Mg(B10H10), which can be obtained starting from Mg(BH4)2.2THF, are applicable for new Mg-based batteries remains to be demonstrated. In the preparation of these closo-hydroborates and their derivatives, starting from BH4 instead of neutral boranes, has the great advantage to reduce the toxicity of the reactants. B2H6, B5H9 and B10H14 are highly toxic and thereby not really suitable for industrial production processes of closo-hydroborates at a higher scale. Thus, boron–hydrogen compounds have a future for new green energy applications.

Funding

This research was funded by the Swiss National Science Foundation, grant number 200020_182494.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Stock, A. The Hydrides of Boron and Silicon; Cornell University Press: New York, NY, USA, 1933. [Google Scholar]
  2. Martin, D.R. The Development of Borane Fuels. J. Chem. Educ. 1959, 36, 208–214. [Google Scholar] [CrossRef]
  3. Baier, M.J.; Veeraraghavan Ramachandran, P.; Son, S.F. Characterization of the Hypergolic Ignition Delay of Ammonia Borane. J. Propuls. Power 2019, 35, 182–189. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Zhao, Z.; Wang, B.; Zhang, J. Boron based hypergolic ionic liquids: A review. Green Energy Environ. 2020, 6, 794–822. [Google Scholar] [CrossRef]
  5. Huang, Z.; Wang, S.; Dewhurst, R.D.; Ignat’ev, N.V.; Finze, M.; Braunschweig, H. Boron: Its Role in Energy-Related Processes and Applications. Angew. Chem. Int. Ed. 2020, 59, 8800–8816. [Google Scholar] [CrossRef]
  6. Zhao, X.; Yang, Z.; Chen, H.; Wang, Z.; Zhou, X.; Zhang, H. Progress in three-dimensional aromatic-like closo-dodecaborate. Coord. Chem. Rev. 2021, 444, 214042. [Google Scholar] [CrossRef]
  7. Stauber, J.M.; Schwan, J.; Zhang, X.; Axtell, J.C.; Jung, D.; McNicholas, B.J.; Oyala, P.H.; Martinolich, A.J.; Winkler, J.R.; See, K.A.; et al. A Super-Oxidized Radical Cationic Icosahedral Boron Cluster. J. Am. Chem. Soc. 2020, 142, 12948–12953. [Google Scholar] [CrossRef]
  8. Tu, D.; Yan, H.; Poater, J.; Solà, M. The nido-Cage···π Bond: A Non-covalent Interaction between Boron Clusters and Aromatic Rings and Its Applications. Angew. Chemie Int. Ed. 2020, 59, 9018–9025. [Google Scholar] [CrossRef] [PubMed]
  9. Alamón, C.; Dávila, B.; García, M.F.; Sánchez, C.; Kovacs, M.; Trias, E.; Barbeito, L.; Gabay, M.; Zeineh, N.; Gavish, M.; et al. Sunitinib-Containing Carborane Pharmacophore with the Ability to Inhibit Tyrosine Kinases Receptors FLT3, KIT and PDGFR-β, Exhibits Powerful In Vivo Anti-Glioblastoma Activity. Cancers 2020, 12, 3423. [Google Scholar] [CrossRef] [PubMed]
  10. Ali, F.; S Hosmane, N.; Zhu, Y. Boron Chemistry for Medical Applications. Molecules 2020, 25, 828. [Google Scholar] [CrossRef] [Green Version]
  11. Stockmann, P.; Gozzi, M.; Kuhnert, R.; Sárosi, M.B.; Hey-Hawkins, E. New keys for old locks: Carborane-containing drugs as platforms for mechanism-based therapies. Chem. Soc. Rev. 2019, 48, 3497–3512. [Google Scholar] [CrossRef] [Green Version]
  12. Bogdanovic, B.; Schwickardi, M. Ti-doped alkali metal aluminium hydrides as potential novel reversible hydrogen storage materials. J. Alloys Compd. 1997, 253–254, 1–9. [Google Scholar] [CrossRef]
  13. Yang, J.; Sudik, A.; Wolverton, C.; Siegel, D.J. High capacity hydrogen storage materials: Attributes for automotive applications and techniques for materials discovery. Chem. Soc. Rev. 2010, 39, 656–675. [Google Scholar] [CrossRef] [Green Version]
  14. Bellosta von Colbe, J.; Ares, J.-R.; Barale, J.; Baricco, M.; Buckley, C.; Capurso, G.; Gallandat, N.; Grant, D.M.; Guzik, M.N.; Jacob, I.; et al. Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. Int. J. Hydrogen Energy 2019, 44, 7780–7808. [Google Scholar] [CrossRef]
  15. Schneemann, A.; White, J.L.; Kang, S.; 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]
  16. Ohno, S.; Banik, A.; Dewald, G.F.; Kraft, M.A.; Krauskopf, T.; Minafra, N.; Till, P.; Weiss, M.; Zeier, W.G. Materials design of ionic conductors for solid state batteries. Prog. Energy 2020, 2, 022001. [Google Scholar] [CrossRef]
  17. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Černý, R.; Ravnsbæk, D.B.; Filinchuk, Y.; Dornheim, M.; Besenbacher, F.; Jensen, T.R. Metal borohydrides and derivatives–synthesis, structure and properties. Chem. Soc. Rev. 2017, 46, 1565–1634. [Google Scholar] [CrossRef]
  18. Černý, R.; Schouwink, P. The crystal chemistry of inorganic metal borohydrides and their relation to metal oxides. Acta Cryst. 2015, B71, 619–640. [Google Scholar] [CrossRef]
  19. Moussa, G.; Moury, R.; Demirci, U.B.; Sener, T.; Miele, P. Boron-based Hydrides for Chemical Hydrogen Storage. Int. J. Energy Res. 2013, 37, 825–842. [Google Scholar] [CrossRef]
  20. Suárez-Alcántara, K.; Tena-Garcia, J.R.; Ricardo Guerrero-Ortiz, R. Alanates, a Comprehensive Review. Materials 2019, 12, 2724. [Google Scholar] [CrossRef] [Green Version]
  21. 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]
  22. Zavorotynska, O.; El-Kharbachi, A.; Deledda, S.; Hauback, B.C. Recent progress in magnesium borohydride Mg(BH4)2: Fundamentals and applications for energy storage. Int. J. Hydrogen Energy 2016, 41, 14387–14403. [Google Scholar] [CrossRef] [Green Version]
  23. Zhang, Y.; Majzoub, E.; Ozoliņš, V.; Wolverton, C. Theoretical Prediction of Metastable Intermediates in the Decomposition of Mg(BH4)2. J. Phys. Chem. C 2012, 116, 10522–10528. [Google Scholar] [CrossRef]
  24. Hwang, S.-J.; Bowman, R.C.; Reiter, J.W.; Rijssenbeek, 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]
  25. Golub, I.E.; Filippov, O.A.; Belkova, N.V.; Epstein, L.M.; Shubina, E.S. The Reaction of Hydrogen Halides with Tetrahydroborate Anion and Hexahydro-closo-hexaborate Dianion. Molecules 2021, 26, 3754. [Google Scholar] [CrossRef]
  26. Voinova, V.V.; Selivanov, N.A.; Plyushchenko, I.V.; Vokuev, M.F.; Bykov, A.Y.; Klyukin, I.N.; Novikov, A.S.; Zhdanov, A.P.; Grigoriev, M.S.; Rodin, I.A.; et al. Fused 1,2-Diboraoxazoles Based on closo-Decaborate Anion–Novel Members of Diboroheterocycle Class. Molecules 2021, 26, 248. [Google Scholar] [CrossRef]
  27. Andreichuk, E.P.; Anufriev, S.A.; Suponitsky, K.Y.; Sivaev, I.B. The First Nickelacarborane with closo-nido Structure. Molecules 2020, 25, 6009. [Google Scholar] [CrossRef]
  28. Klyukin, I.N.; Vlasova, Y.S.; Novikov, A.S.; Zhdanov, A.P.; Hagemann, H.R.; Zhizhin, K.Y.; Kuznetsov, N.T. B-F bonding and reactivity analysis of mono- and perfluoro-substituted derivatives of closo-borate anions (6, 10, 12): A computational study. Polyhedron 2022, 211, 115559. [Google Scholar] [CrossRef]
  29. Udovic, T.J.; Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Rush, J.J.; Takamura, H.; Orimo, S. Sodium Superionic Conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750–3752. [Google Scholar] [CrossRef]
  30. Duchêne, L.; Remhof, A.; Hagemann, H.; Battaglia, C. Status and prospects of hydroborate electrolytes for all-solid-state batteries. Energy Storage Mat. 2020, 26, 543–549. [Google Scholar] [CrossRef]
  31. He, L.; Li, H.-W.; Hwang, S.-J.; Akiba, E. Facile Solvent-Free Synthesis of anhydrous alkali metal dodecaborate M2B12H12 (M = Li, Na, K). J. Phys. Chem. C 2014, 118, 6084–6089. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, W.; Zhang, X.; Huang, Z.; Li, H.-W.; Gao, M.; Pan, H.; Liu, Y. Recent Development of Lithium Borohydride-Based Materials for Hydrogen Storage. Adv. Energy Sustain. Res. 2021, 2, 2100073. [Google Scholar] [CrossRef]
  33. 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]
  34. Yang, J.; Zhang, X.; Zheng, J.; Song, P.; Li, X. Decomposition pathway of Mg(BH4)2 under pressure: Metastable phases and thermodynamic parameters. Scr. Mater. 2011, 64, 225–228. [Google Scholar] [CrossRef]
  35. Vitillo, J.G.; Bordiga, S.; Baricco, M. Spectroscopic and Structural Characterization of Thermal Decomposition of γ-Mg(BH4)2: Dynamic Vacuum versus H2 Atmosphere. J. Phys. Chem. C 2015, 119, 25340–25351. [Google Scholar] [CrossRef]
  36. Wang, X.; Xiao, X.; Zheng, J.; Yao, Z.; Zhang, M.; Huang, X.; Chen, L. Insights into magnesium borohydride dehydrogenation mechanism from its partial reversibility under moderate conditions. Mater. Today Energy 2020, 18, 100552. [Google Scholar] [CrossRef]
  37. Crociani, L.; Rossetto, G.; Kaciulis, S.; Mezzi, A.; El-Habra, N.; Palmieri, V. Study of Magnesium Boride Films Obtained From Mg(BH4)2 by CVD. Chem. Vap. Depos. 2007, 13, 414–419. [Google Scholar] [CrossRef]
  38. Kim, D.Y.; Yang, Y.; Abelson, J.R.; Girolami, G.S. Volatile magnesium octahydrotriborate complexes as potential CVD Precursors to MgB2. Synthesis and Characterization of Mg(B3H8)2 and its etherates. Inorg. Chem. 2007, 46, 9060–9066. [Google Scholar] [CrossRef]
  39. Pistidda, C.; Garroni, S.; Dolci, F.; Gil Bardají, E.; Khandelwal, A.; Nolis, P.; Dornheim, M.; Gosalawit, R.; Jensen, T.; Cere-nius, Y.; et al. Synthesis of amorphous Mg(BH4)2 from MgB2 and H2 at room temperature. J. Alloys Comp. 2010, 508, 212–215. [Google Scholar] [CrossRef]
  40. Severa, G.; Rönnebro, E.; Jensen, C.M. Direct hydrogenation of magnesium boride to magnesium borohydride: Demonstration of >11 weight percent reversible hydrogen storage. Chem. Commun. 2010, 46, 421–423. [Google Scholar] [CrossRef]
  41. Sugai, C.; Kim, S.; Severa, G.; White, J.L.; Leick, N.; Martinez, M.B.; Gennett, T.; Stavila, V.; Jensen, C. Kinetic Enhancement of Direct Hydrogenation of MgB2 to Mg(BH4)2 upon Mechanical Milling with THF, MgH2, and/or Mg. ChemPhysChem 2019, 20, 1301–1304. [Google Scholar] [CrossRef]
  42. Pistidda, C.; Santhosh, A.; Jerabek, P.; Shang, 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]
  43. Ray, K.G.; Klebanoff, L.E.; Lee, J.R.I.; Stavila, V.; Wook Heo, T.; Shea, P.; Baker, A.A.; Kang, S.; Bagge-Hansen, M.; Liu, Y.-S.; et al. Elucidating the mechanism of MgB2 initial hydrogenation via a combined experimental–theoretical study. Phys. Chem. Chem. Phys. 2017, 19, 22646–22658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Chong, M.; Matsuo, M.; Orimo, S.; Autrey, T.; Jensen, C.M. Selective Reversible Hydrogenation of Mg(B3H8)2/MgH2 to Mg(BH4)2: Pathway to Reversible Borane-Based Hydrogen Storage? Inorg. Chem. 2015, 54, 4120–4125. [Google Scholar] [CrossRef]
  45. Yue, Y.; Chen, L.; Peng, J. Thermal Behaviors and Their Correlations of Mg(BH4)2-Contained Explosives. J. Energetic Mater. 2018, 36, 82–92. [Google Scholar] [CrossRef]
  46. Hagemann, H. Estimation of Thermodynamic Properties of Metal Hydroborates. ChemistrySelect 2019, 4, 8989–8992. [Google Scholar] [CrossRef]
  47. Chase, M.W. NIST-JANAF Thermochemical Tables. J. Phys. Chem. Ref. Data Monogr. 1998, 2. [Google Scholar] [CrossRef]
  48. Hagemann, H.; D’Anna, V.; Rapin, J.-P.; Yvon, K. Deuterium-Hydrogen Exchange in Solid Mg(BH4)2. J. Phys. Chem. C 2010, 114, 10045–10047. [Google Scholar] [CrossRef]
  49. Sharma, M.; Sethio, D.; D’Anna, V.; Fallas, J.C.; Schouwink, P.; Cerný, R.; Hagemann, H. Isotope Exchange Reactions in Ca(BH4)2. J. Phys. Chem. C 2015, 119, 29–32. [Google Scholar] [CrossRef]
  50. Huang, Z.; Wang, Y.; Wang, D.; Yang, F.; Wu, Z.; Wua, L.; Zhang, Z. Role of native defects and the effects of metal additives on the kinetics of magnesium borohydride. Phys. Chem. Chem. Phys. 2019, 21, 11226–11233. [Google Scholar] [CrossRef]
  51. Heere, M.; Zavorotynska, O.; Deledda, S.; Sørby, M.H.; Book, D.; Steriotis, T.; Hauback, B.C. Effect of additives, ball milling and isotopic exchange in porous magnesium borohydride. RSC Adv. 2018, 8, 27645–27653. [Google Scholar] [CrossRef] [Green Version]
  52. 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]
  53. Saldan, I.; Frommen, C.; Llamas-Jansa, I.; Kalantzopoulos, G.N.; Hino, S.; Arstad, B.; Heyn, R.H.; Zavorotynska, O.; Deledda, S.; Sørby, M.H.; et al. Hydrogen storage properties of gamma-Mg(BH4)2 modified by MoO3 and TiO2. Int. J. Hydrogen Energy 2015, 40, 12286–12293. [Google Scholar] [CrossRef]
  54. Chong, M.; Autrey, T.; Jensen, C.M. Lewis Base Complexes of Magnesium Borohydride: Enhanced Kinetics and Product Selectivity upon Hydrogen Release. Inorganics 2017, 5, 89. [Google Scholar] [CrossRef] [Green Version]
  55. Dimitrievska, M.; Chong, M.; Bowden, M.E.; Wu, H.; Zhou, W.; Nayyar, I.; Ginovska, B.; Gennett, T.; Autrey, T.; Jensen, C.M.; et al. Structural and reorientational dynamics of tetrahydroborate (BH4) and tetrahydrofuran (THF) in a Mg(BH4)2·3THF adduct: Neutron-scattering characterization. Phys. Chem. Chem. Phys. 2020, 22, 368–378. [Google Scholar] [CrossRef]
  56. 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]
  57. Bell, R.T.; Strange, N.A.; Leick, N.; Stavila, V.; Bowden, M.E.; Autrey, T.S.; Gennett, T. Mg(BH4)2-Based Hybrid Metal–Organic Borohydride System Exhibiting Enhanced Chemical Stability in Melt. ACS Appl. Energy Mater. 2021, 4, 1704–1713. [Google Scholar] [CrossRef]
  58. Wegner, W.; Jaroń, T.; Dobrowolski, M.A.; Dobrzycki, Ł.; Cyrańskic, M.K.; Grochala, W. Organic derivatives of Mg(BH4)2 as precursors towards MgB2 and novel inorganic mixed-cation borohydrides. Dalton Trans. 2016, 45, 14370–14377. [Google Scholar] [CrossRef]
  59. Moury, R.; Gigante, A.; Remhof, A.; Roedern, E.; Hagemann, H. Experimental investigation of Mg(B3H8)2 dimensionality, materials for energy storage applications. Dalton Trans. 2020, 49, 12168–12173. [Google Scholar] [CrossRef]
  60. Gigante, A.; Leick, N.; Lipton, A.S.; Tran, B.; Strange, N.A.; Bowden, M.; Martinez, M.B.; Moury, R.; Gennett, T.; Hagemann, H.; et al. Thermal Conversion of Unsolvated Mg(B3H8)2 to BH4 in the Presence of MgH2. ACS Appl. Energy Mater. 2021, 4, 3737–3747. [Google Scholar] [CrossRef]
  61. Newhouse, R.J.; Stavila, V.; Hwang, S.-J.; Klebanoff, L.E.; Zhang, J.Z. Reversibility and Improved Hydrogen Release of Magnesium Borohydride. J. Phys. Chem. C 2010, 114, 5224–5232. [Google Scholar] [CrossRef] [Green Version]
  62. Beall, H.; Gaines, D.F. Mechanistic aspects of boron hydride reactions. Inorg. Chim. Acta 1999, 289, 1–10. [Google Scholar] [CrossRef]
  63. Kurbonbekov, A.; Alikhanova, T.K.; Badalov, A.; Murufi, V.K.; Mirsaidov, U. Solubility in the lanthanum borohydride -potassium borohydride -tetrahydrofuran system at 25 °C and some thermodynamic characteristics of lanthanum borohydride. Dokl. Akad. Nauk Tadzh. SSR 1990, 33, 393–395. [Google Scholar]
  64. Huang, Z.; Chen, X.; Yisgedu, T.; Meyers, E.A.; Shore, S.G.; Zhao, J.-C. Ammonium Octahydrotriborate (NH4B3H8): New Synthesis, Structure, and Hydrolytic Hydrogen Release. Inorg. Chem. 2011, 50, 3738–3742. [Google Scholar] [CrossRef] [Green Version]
  65. Good, W.D.; Mansson, M. The Thermochemistry of Boron and Some of Its Compounds. The Enthalpies of Formation of Orthoboric Acid, Trimethylamineborane, and Diammoniumdecaborane. J. Phys. Chem. 1966, 70, 97–104. [Google Scholar] [CrossRef]
  66. Hanumantha Rao, M.; Muralidharan, K. closo-Dodecaborate (B12H12)2− salts with nitrogen based cations and their energetic properties. Polyhedron 2016, 115, 105–110. [Google Scholar]
  67. Dematteis, E.M.; Jensen, S.R.; Jensen, T.R.; Baricco, M. Heat capacity and thermodynamic properties of alkali and alkali-earth borohydrides. J. Chem. Thermodyn. 2020, 143, 106055. [Google Scholar] [CrossRef]
  68. Pinatel, E.R.; Albanese, E.; Civalleri, B.; Baricco, M. Thermodynamic modelling of Mg(BH4)2. J. Alloys Comp. 2015, 645, S64–S68. [Google Scholar] [CrossRef] [Green Version]
  69. Nguyen, M.T.; Matus, M.H.; Dixon, D.A. Heats of Formation of Boron Hydride Anions and Dianions and Their Ammonium Salts [BnHmy−][NH4+]y with y = 1−2. Inorg. Chem. 2007, 46, 7561–7570. [Google Scholar] [CrossRef]
  70. McKee, M.L. Estimation of Heats of Formation of Boron Hydrides from ab Initio Energies. J. Phys. Chem. 1990, 94, 435–440. [Google Scholar] [CrossRef]
  71. Kelley, S.P.; McCrary, P.D.; Flores, L.; Garner, E.B.; Dixon, D.A.; Rogers, R.D. Structural and Theoretical Study of Salts of the [B9H14] Ion: Isolation of Multiple Isomers and Implications for Energy Storage. ChemPlusChem 2016, 81, 922–925. [Google Scholar] [CrossRef]
  72. Sethio, D.; Lawson Daku, L.M.; Hagemann, H.; Kraka, E. Quantitative Assessment of B–B–B, B–Hb–B, and B–Ht Bonds: From BH3 to B12H122−. ChemPhysChem 2019, 20, 1967–1977. [Google Scholar] [CrossRef]
  73. Maillard, R.; Sethio, D.; Hagemann, H.; Lawson Daku, L.M. Accurate Computational Thermodynamics Using Anharmonic Density Functional Theory Calculations: The Case Study of B−H Species. ACS Omega 2019, 4, 8786–8794. [Google Scholar] [CrossRef] [Green Version]
  74. Yu, C.L.; Bauer, S.H. Thermochemistry of the boranes. J. Phys. Chem. Ref. Data 1998, 27, 807–835. [Google Scholar] [CrossRef]
  75. Lee, T.B.; Mc Kee, M.L. Redox Energetics of Hypercloso Boron Hydrides BnHn (n = 6–13) and B12X12 (X = F, Cl, OH, and CH3). Inorg. Chem. 2012, 51, 4205–4214. [Google Scholar] [CrossRef]
  76. Lee, T.B.; Mc Kee, M.L. Dissolution Thermochemistry of Alkali Metal Dianion Salts (M2X1, M = Li+, Na+, and K+ with X = CO32−, SO42−, C8H82−, and B12H122−). Inorg. Chem. 2011, 50, 11412–11422. [Google Scholar] [CrossRef]
  77. Yan, Y.; Remhof, A.; Rentsch, D.; Züttel, A. The role of MgB12H12 in the hydrogen desorption process of Mg(BH4)2. Chem. Commun. 2015, 51, 700–702. [Google Scholar]
  78. Yan, Y.; Remhof, A.; Rentsch, D.; Lee, Y.S.; Cho, Y.W.; Züttel, A. Is Y2(B12H12)3 the main intermediate in the decomposition process of Y(BH4)3? Chem. Commun. 2013, 49, 5234–5236. [Google Scholar] [CrossRef] [Green Version]
  79. Godfroid, R.A.; Hill, T.G.; Onak, T.P.; Shore, S.G. Formation of [BH3]2− and [B2H6]2− From the Homogeneous Reduction of B2H6. J. Am. Chem. Soc. 1994, 116, 12107–12108. [Google Scholar] [CrossRef]
  80. Hill, T.G.; Godfroid, R.A.; White III, J.P.; Shore, S.G. Reduction of borane THF by alkali metal (potassium, rubidium, cesium) and ytterbium mercury amalgams to form salts of octahydrotriborate(1-); a simple procedure for the synthesis of tetraborane(10). Inorg. Chem. 1991, 30, 2952–2954. [Google Scholar] [CrossRef]
  81. Gaines, D.F.; Schaeffer, R.; Tebbe, F. Convenient Preparations of Solutions Containing the Triborohydride Ion. Inorg. Chem. 1963, 2, 526–528. [Google Scholar] [CrossRef]
  82. Chen, X.; Liu, X.-R.; Wang, X.; Chen, X.-M.; Jing, Y.; Wie, D. A Safe and Efficient Synthetic Method of the Alkali Metal Octahydrotriborate, Unravelling a General Mechanism of Constructing the Delta B3 Unit of Polyhedral Boranes. Dalton Trans. 2021, 50, 13676–13679. [Google Scholar] [CrossRef]
  83. Chen, X.-M.; Ma, N.; Zhang, Q.-F.; Wang, J.; Feng, X.; Wei, C.; Wang, L.-S.; Zhang, J.; Chen, X. Elucidation of the Formation Mechanisms of the Octahydrotriborate Anion (B3H8) through the Nucleophilicity of the B–H Bond. J. Amer. Chem. Soc. 2018, 140, 6718–6726. [Google Scholar] [CrossRef]
  84. Moury, R.; Gigante, A.; Hagemann, H. An alternative approach to the synthesis of NaB3H8 and Na2B12H12 for solid electrolyte applications. Int. J. Hydrogen Energy 2017, 42, 22417–22421. [Google Scholar] [CrossRef]
  85. 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]
  86. Gigante, A.; Duchêne, L.; Moury, R.; Pupier, M.; Remhof, A.; Hagemann, H. Direct solution–based synthesis of the Na4(B12H12)(B10H10) solid electrolyte. ChemSusChem 2019, 12, 4832–4837. [Google Scholar] [CrossRef]
  87. Aniya, M. A chemical approach for the microscopic mechanism of fast ion transport in solids. Solid State Ion. 1992, 50, 125–129. [Google Scholar] [CrossRef]
  88. 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]
  89. Muetterties, E.L.; Balthis, J.H.; Chia, Y.T.; Knoth, W.H.; Miller, H.C. Chemistry of Boranes. VIII. Salts and Acids of B10H10−2 and B12H12−2. Inorg. Chem. 1964, 3, 444–451. [Google Scholar] [CrossRef]
  90. Černý, R.; Brighi, M.; Murgia, F. The Crystal Chemistry of Inorganic Hydroborates. Chemistry 2020, 2, 805–826. [Google Scholar] [CrossRef]
  91. Wu, H.; Tang, W.S.; Stavila, V.; Zhou, W.; Rush, J.J.; Udovic, T.J. Structural Behavior of Li2B10H10. J. Phys. Chem. C 2015, 119, 6481–6487. [Google Scholar] [CrossRef]
  92. Her, J.-H.; Yousufuddin, M.; Zhou, W.; Jalisatgi, S.S.; Kulleck, J.G.; Zan, J.A.; Hwang, S.-J.; Bowman, R.C.; 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] [Green Version]
  93. Paskevicius, M.; Pitt, M.P.; Brown, D.H.; Sheppard, D.A.; Chumphongphan, S.; Buckley, C.E. First-order phase transition in the Li2B12H12 system. Phys. Chem. Chem. Phys. 2013, 15, 15825–15828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Wu, H.; Tang, W.S.; Zhou, W.; Stavila, V.; Rush, J.J.; Udovic, T.J. The structure of monoclinic Na2B10H10: A combined diffraction, spectroscopy, and theoretical approach. Cryst. Eng. Comm. 2015, 17, 3533–3540. [Google Scholar] [CrossRef]
  95. Verdal, N.; Her, J.-H.; Stavila, V.; Soloninin, A.V.; Babanova, O.A.; Skripov, A.V.; Udovic, T.J.; Rush, J.J. Complex high-temperature phase transitions in Li2B12H12 and Na2B12H12. J. Solid State Chem. 2014, 212, 81–91. [Google Scholar] [CrossRef]
  96. Wunderlich, J.A.; Lipscomb, W.N. Structure of B12H12−2 ion. J. Am. Chem. Soc. 1960, 82, 4427–4428. [Google Scholar] [CrossRef]
  97. Hofmann, K.; Albert, B. Crystal structures of M2[B10H10] (M = Na, K, Rb) via real space simulated annealing powder techniques. Z. Kristall. 2005, 220, 142–146. [Google Scholar] [CrossRef]
  98. Bukovsky, E.V.; Peryshkov, D.V.; Wu, H.; Zhou, W.; Tang, W.S.; Jones, W.M.; Stavila, V.; Udovic, T.J.; Strauss, S.H. Comparison of the Coordination of B12F122−, B12Cl122−, and B12H122− to Na+ in the Solid State: Crystal Structures and Thermal Behavior of Na2B12F12,Na2(H2O)4B12F12, Na2B12Cl12, and Na2(H2O)6B12Cl12. Inorg. Chem. 2017, 56, 4369–4379. [Google Scholar] [CrossRef] [PubMed]
  99. Tiritiris, I.; Schleid, T. The Dodecahydro-closo-Dodecaborates M2[B12H12] of the Heavy Alkali Metals (M = K+, Rb+, NH4+, Cs+) and their Formal Iodide Adducts M3I [B12H12] (= MI·M2[B12H12]). Z. Anorg. Allg. Chem. 2003, 629, 1390–1402. [Google Scholar] [CrossRef]
  100. Guggenberger, L.J. Chemistry of boranes. XXXIII. The crystal structure of Rb2B9H9. Inorg. Chem. 1968, 7, 2260–2264. [Google Scholar] [CrossRef]
  101. Verdal, N.; Wu, H.; Udovic, T.J.; Stavila, V.; Zhou, W.; Rush, J.J. Evidence of a transition to reorientational disorder in the cubic alkali-metal dodecahydro-closo-dodecaborates. J. Solid State Chem. 2011, 184, 3110–3116. [Google Scholar] [CrossRef]
  102. Moury, R.; Lodziana, Z.; Remhof, A.; Duchêne, L.; Roedern, E.; Gigante, A.; Hagemann, H. Pressure-induced phase transitions in Na2B12H12, structural investigation on a candidate for solid-state electrolyte. Acta Cryst. B 2019, B75, 406–413. [Google Scholar] [CrossRef] [Green Version]
  103. Maekawa, H.; Matsuo, M.; Takamura, H.; Ando, M.; Noda, Y.; Karahashi, T.; Orimo, S. Halide-stabilized LiBH4, a room-temperature lithium fast-ion conductor. J. Am. Chem. Soc. 2009, 894–895. [Google Scholar] [CrossRef]
  104. Gulino, V.; Brighi, M.; Dematteis, E.M.; Murgia, F.; Nervi, C.; Cerny, R.; Baricco, M. Phase Stability and Fast Ion Conductivity in the Hexagonal LiBH4-LiBr-LiCl Solid Solution. Chem. Mater. 2019, 31, 5133–5144. [Google Scholar] [CrossRef] [Green Version]
  105. Sadikin, Y.; Schouwink, P.; Brighi, M.; Łodziana, Z.; Cerný, R. Modified anion packing of Na2B12H12 in close to room temperature superionic conductors. Inorg. Chem. 2017, 56, 5006–5016. [Google Scholar] [CrossRef]
  106. Kim, S.; Oguchi, H.; Toyama, N.; Sato, T.; Takagi, S.; Otomo, T.; Arunkumar, D.; Kuwata, N.; Kawamura, J.; Orimo, S. A complex hydride lithium superionic conductor for high-energy-density all-solid-state lithium metal batteries. Nat. Comm. 2019, 10, 1081. [Google Scholar] [CrossRef] [Green Version]
  107. Payandeh, S.H.; Rentsch, D.; Łodziana, Z.; Asakura, R.; Bigler, L.; Černý, R.; Battaglia, C.; Remhof, A. Nido-Hydroborate-Based Electrolytes for All-Solid-State Lithium Batteries. Adv. Funct. Mater. 2021, 31, 2010046. [Google Scholar] [CrossRef]
  108. Brighi, M.; Murgia, F.; Łodziana, Z.; Schouwink, P.; Wołczyk, A.; Černý, R. A mixed anion hydroborate/carba-hydroborate as a room temperature Na ion solid electrolyte. J. Power Sources 2019, 404, 7–12. [Google Scholar] [CrossRef]
  109. Duchêne, L.; Kühnel, R.-S.; Rentsch, D.; Remhof, A.; Hagemann, H.; Battaglia, C. A highly stable sodium solid-state electrolyte based on a dodeca/deca-borate equimolar mixture. Chem. Commun. 2017, 53, 4195–4198. [Google Scholar] [CrossRef] [Green Version]
  110. Yoshida, K.; Sato, T.; Unemoto, A.; Matsuo, M.; Ikeshoji, T.; Udovic, T.J.; Orimo, S.I. Fast sodium ionic conduction in Na2B10H10-Na2B12H12 pseudo-binary complex hydride and application to a bulk-type all-solid-state battery. Appl. Phys. Lett. 2017, 110, 103901. [Google Scholar] [CrossRef] [Green Version]
  111. Payandeh, S.H.; Asakura, R.; Avramidou, P.; Rentsch, D.; Łodziana, Z.; Černý, R.; Remhof, A.; Battaglia, C. Nido-Borate/Closo-borate mixed-anion electrolytes for all-solid-state batteries. Chem. Mater. 2020, 32, 1101–1110. [Google Scholar] [CrossRef]
  112. Brighi, M.; Murgia, F.; Cerny, R. Closo-Hydroborate Sodium Salts as an Emerging Class of Room-Temperature Solid Electrolytes. Cell Rep. Phys. Sci. 2020, 1, 100217. [Google Scholar] [CrossRef]
  113. 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]
  114. Lohstroh, W.; Heere, M. Structure and Dynamics of Borohydrides Studied by Neutron Scattering Techniques: A Review. J. Phys. Soc. Japan. 2020, 89, 1–12. [Google Scholar] [CrossRef]
  115. Duchêne, L.; Lunghammer, S.; Burankova, T.; Liao, W.-C.; Embs, J.P.; Coperet, C.; Wilkening, H.M.R.; Remhof, A.; Hagemann, H.; Battaglia, C. Ionic conduction mechanism in the Na2(B12H12)0.5(B10H10)0.5 closo-borate solid-state electrolyte: Interplay of disorder and ion–ion interactions. Chem. Mater. 2019, 31, 3449–3460. [Google Scholar] [CrossRef]
  116. Asakura, R.; Duchêne, L.; Kühnel, R.-S.; Remhof, A.; Hagemann, H.; Battaglia, C. Electrochemical Oxidative Stability of Hydroborate-Based Solid-State Electrolytes. ACS Appl. Energy Mater. 2019, 2, 6924–6930. [Google Scholar] [CrossRef]
  117. Matsuo, M.; Orimo, S. Lithium Fast-Ionic Conduction in Complex Hydrides: Review and Prospects. Adv. Energy Mater. 2011, 1, 161–172. [Google Scholar] [CrossRef]
  118. Duchêne, L.; Kühnel, R.-S.; Stilp, E.; Reyes, E.C.; Remhof, A.; Hagemann, H.; Battaglia, C. A Stable 3 V all-solid-state sodium-ion battery based on a closo -borate electrolyte. Energy Environ. Sci. 2017, 10, 2609–2615. [Google Scholar] [CrossRef]
  119. Murgia, F.; Brighi, M.; Cerny, R. Room-temperature-operating Na-ion solid state-battery with complex hydride as electrolyte. Electrochem. Comm. 2019, 106, 106534. [Google Scholar] [CrossRef]
  120. Asakura, R.; Reber, D.; Duchêne, L.; Payandeh, S.; Remhof, A.; Hagemann, H.; Battaglia, C. 4 V room-temperature all-solid-state sodium battery enabled by a passivating cathode/hydroborate solid electrolyte interface. Energy Environ. Sci. 2020, 13, 5048–5058. [Google Scholar] [CrossRef]
Figure 1. Illustration of Mg(BH4)2 dehydrogenation reactions (blue arrows) and rehydrogenation reactions (red arrows) reported in the literature [22,34,35,36,37,38,39,40,41,42,43,44]. Upon further heating, these intermediate species, which are associated with (amorphous) MgH2, form MgB2.
Figure 1. Illustration of Mg(BH4)2 dehydrogenation reactions (blue arrows) and rehydrogenation reactions (red arrows) reported in the literature [22,34,35,36,37,38,39,40,41,42,43,44]. Upon further heating, these intermediate species, which are associated with (amorphous) MgH2, form MgB2.
Molecules 26 07425 g001
Figure 2. Experimental (bold) and theoretical formation enthalpy values for neutral (red) monoanionic (black) and dianionic (blue) species. Closo species, circles; nido, #; arachno, crossed squares. Data from [69,70,71,72,74,75,76]. For closo ions BnHn2−, data (blue circles) from 2 different studies [69,72] reveal systematic differences. All monoanionic species (in black) have negative formation enthalpies, while all neutral boranes (in red) have positive formation enthalpy.
Figure 2. Experimental (bold) and theoretical formation enthalpy values for neutral (red) monoanionic (black) and dianionic (blue) species. Closo species, circles; nido, #; arachno, crossed squares. Data from [69,70,71,72,74,75,76]. For closo ions BnHn2−, data (blue circles) from 2 different studies [69,72] reveal systematic differences. All monoanionic species (in black) have negative formation enthalpies, while all neutral boranes (in red) have positive formation enthalpy.
Molecules 26 07425 g002
Figure 3. Group–subgroup relationships between space groups (in Herrmann-Mauguin notation) of closo-hydro borates and some closo-halogeno borates. t, “translationengleich” subgroups; k, “klassengleich” subgroups.
Figure 3. Group–subgroup relationships between space groups (in Herrmann-Mauguin notation) of closo-hydro borates and some closo-halogeno borates. t, “translationengleich” subgroups; k, “klassengleich” subgroups.
Molecules 26 07425 g003
Table 1. Examples of ionic conductivity in mixed borate salts.
Table 1. Examples of ionic conductivity in mixed borate salts.
CompoundTemperatureConductivityReference
0.7 Li(CB9H10)–0.3 Li(CB11H12)298 K6.7 mS/cm[106]
Li2(B11H14)(CB11H12)298 K0.11 mS/cm[107]
Li3(B11H14)(CB11H12)2298 K1.1 mS/cm[107]
Na3(CB11H12)(B12H12)298 K2 mS/cm[108]
Na4(CB11H12)2(B12H12)298 K2 mS/cm[108]
Na4(B10H10)(B12H12)298 K0.9 mS/cm[109]
Na2(B10H10)−3 Na2(B12H12)298 K0.34 mS/cm[110]
Nax+2y(B11H14)x(B12H12)y298 K3–4 mS/cm[111]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hagemann, H. Boron Hydrogen Compounds: Hydrogen Storage and Battery Applications. Molecules 2021, 26, 7425. https://doi.org/10.3390/molecules26247425

AMA Style

Hagemann H. Boron Hydrogen Compounds: Hydrogen Storage and Battery Applications. Molecules. 2021; 26(24):7425. https://doi.org/10.3390/molecules26247425

Chicago/Turabian Style

Hagemann, Hans. 2021. "Boron Hydrogen Compounds: Hydrogen Storage and Battery Applications" Molecules 26, no. 24: 7425. https://doi.org/10.3390/molecules26247425

Article Metrics

Back to TopTop