Next Article in Journal
Synthesis, Structure and Catechol Oxidase Activity of Mono Nuclear Cu(II) Complex with Phenol-Based Chelating Agent with N, N, O Donor Sites
Previous Article in Journal
Advanced Catalysts for the Water Gas Shift Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Rigidity of the (BH4)-Anion Dispersed in Halides AX, A = Na, K; X = Cl, Br, I, and in MBH4 with M = Na, K, Rb, Cs

by
Zeina Assi
1,
Alexander Gareth Schneider
2,
Anna Christina Ulpe
2,
Thomas Bredow
2 and
Claus Henning Rüscher
1,*
1
Institut für Mineralogie, Leibniz Universität Hannover, Callinstr. 3, D-30167 Hannover, Germany
2
Mulliken Center for Theoretical Chemistry, Rheinische Friedrich-Wilhelm-Universität Bonn, Bergstr. 4-6, D-53115 Bonn, Germany
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(4), 510; https://doi.org/10.3390/cryst12040510
Submission received: 9 March 2022 / Revised: 28 March 2022 / Accepted: 1 April 2022 / Published: 6 April 2022

Abstract

:
The B–H bond length of the borohydride anion (BH4) in alkali metal borohydrides MBH4 with M = Na, K, Rb, Cs, and diluted in different alkali halide matrices, was investigated experimentally by infrared spectroscopy (FTIR) and theoretically using first principles calculations. The peak positions in IR absorption spectra of NaBH4 pressed at 754 MPa in halides NaX and KX with X = Cl, Br, I show significant variations indicating ion exchange effects between the halide and NaBH4. For NaBH4 in NaBr, NaI, KBr and KI pellets, the peak positions indicate that BH4 could be highly diluted in the AX matrix, which renders an isolation of BH4 in AX (i-BH4). For NaBH4 in NaCl and KCl pellets, a solution of BH4 in AX occurred only after a further thermal treatment up to 450 °C. The observed peak positions are discussed with respect to the lattice parameter (a0), anion to cation ratio (R = rA/rX), standard enthalpy of formation (ΔfH) and ionic character (Ic) of the halides. A linear relation is obtained between ν3(i-BH4) and the short-range lattice energies of AX. Density functional theory (DFT) calculations at generalized gradient approximation (GGA) level were used to calculate the IR vibrational frequencies ν4, ν3 and ν2 + ν4 for series of compositions Na(BH4)0.25X0.75 with X = Cl, Br, I, and MBH4. The theoretical and experimental results show the same trends, indicating the rigidity of the B–H bond length and the failure of Badger’s rule.

1. Introduction

Borohydrides, conventionally called tetrahydroborates following the recommendation of the International Union of Pure and Applied Chemistry (IUPAC), have received attention due to their remarkable gravimetric and volumetric hydrogen contents [1]. Here, the anionic part is constituted of the tetrahedral BH4 group and the negative charge on the borohydride unit is generally compensated by alkali or alkaline earth metals. This has been known since the discovery of NaBH4 and moreover their potential use as reducing agents [2]. Alkaline solutions of NaBH4 were already used for producing electrical energy with fuel cell technology in Apollo missions. However, a breakthrough for a viable hydrogen economy is not achieved because of the irreversibility of the hydrogen release, although there were great efforts for better understanding the related problems and also receiving technical solutions for suitable applications of using NaBH4 as hydrogen storage [3].
Fundamental research shows that the thermal desorption of H2 occurs at temperatures above 450 °C and leads to a decomposition of the compound into the elements according to MBH4 => M + B + 2H2 [1,4]. Some novel approaches could reduce the hydrogen desorption temperature including catalysis [5,6], nano-engineering [7,8] and use of reactive additives [9,10,11]. Some work includes the enclosure and handling of the BH4 anion in sodalite cages [12,13], or NaBH4 in an aluminosilicate gel [13,14]. The investigations of the BH4 anion in the sodalite cages were mainly focused on the thermal control of steps of the effective reaction:
BH4 + 2H2O = > BO2 + 4H2,
Reaction (1) is supported by previous investigations showing that half of the released hydrogen is supplied by water [3]. In a combined theoretical and experimental approach [15], B(OH)3/(OH), B(OH)4 and BO(OH)2 could be identified as intermediates in the sodalite cages finally receiving BO2. In a more recent investigation [16], it was concluded that an initial reaction reveal (BH3) and (OH) in separated sodalite cages followed by a fast reaction:
BH3 + 3H2O = > B(OH)3 + 3H2,
Thus, the reaction sequence could be in close analogy to conclusions by Mesmer and Jolly [17] that hydrolysis of NaBH4 in aqueous solutions take place forming initially BH5 which separates into BH3 + H2. Kreevoy and Hutchins [18] attributed it to an acid catalyzed mechanism, where BH3 rather rapidly reacts with 3H2O to B(OH)3 plus 3H2, and B(OH)3 reacts successively with (OH) to B(OH)4.
Basic questions still concern details of the stability of the BH4 anion in various borohydrides. Nakamori et al. [19,20] reported a correlation between the first hydrogen desorption temperature (Td) of the metal borohydrides and the Pauling electronegativity (χP) of the metal cations, while Td decreases with increasing χP of the metal. An apparent linear relationship was obtained between the Td and χP for the borohydrides of Na, Li, Mg, Sc, Zr and Zn indicating that χP could be considered as an indicator for approximate estimations of the stability of borohydrides. This relation is applicable for other classes of hydrides as tetra-alanates (AlH4) [21] and hexa-alanates (AlH63) [22], too. Hence, one approach to affect the thermal stability of metal borohydrides is to introduce a metal with higher electronegativity through lattice cation substitution. For example, for the cation substitution of Na (χP = 0.93), (The χP values are dimensionless and are given according to Pauling’s scale for the elements of the periodic table; χP is defined as the power of an atom in a molecule to attract electrons [23]) and Li (χP = 0.98) with Zn (χP = 1.65) in NaBH4 and LiBH4 salts, respectively, by mechanically ball milling the borohydride salts with ZnCl2 produces mixed-metal of borohydrides NaZn(BH4)3, NaZn2(BH4)5 and LiZn2(BH4)5, which are thermally less stable than their unsubstituted patterns [24].
The anion substitution in the alkali metal borohydride is also used to tune the thermodynamics of the borohydrides. The mixed hydride-fluoride Na3AlH(6-x)Fx prepared from NaF, Al and H2 is destabilized relative to the pure hydride Na3AlH6 [25]. Likewise, first-principles calculations of the decomposition reaction of LiBH4 with and without fluorine F suggested that doping LiBH4 with F results in a partial substitution of H by F in the hydride lattice yielding favorable thermodynamics modifications [26]. In a more recent study on ball milled samples of NaBH4 and NaBF4, the authors reported a fluorine substitution in the temperature range 200–215 °C yielding NaBH2F2. The NaBH4-NaBF4 composite decomposes at temperatures by at least 100 °C lower in comparison with NaBH4 [10].
The anion substitution may involve the whole borohydride unit as the substitution of BH4 by Cl in NaBH4 [27,28], and with Br [29] and I [30] in LiBH4, respectively. Differential Scanning Calorimetry (DSC) measurements of the solid solutions reveal the endothermic peaks of the melt as an indicator of the thermal stability at higher temperatures compared to the unsubstituted borohydride salts, suggesting that hydrogen release is not improved by these substitutions.
The stabilization/destabilization effect upon the lattice substitutions could be attributed to the different types of interactions altered during the ionic exchange. The ionic bonding is effective between the cation element and the BH4 group, whereas the bonding between the boron B and the hydrogen H atoms within the BH4 group shows a covalent character [31]. The present work is focused on a better understanding of the interactions between the BH4 anion and its environment. Room temperature, Raman and infrared spectroscopic studies of alkali borohydrides MBH4, M = Na, K, Rb and Cs, could be fully assigned by Renaudin et al. [32] for both hydrides and deuterides, including the overtones with Fermi resonance and combination bands of the BH4 anion, which was in agreement with previous assignments [33,34,35,36]. The data were found to be in perfect agreement with Badger′s rule possessing a linear relationship between vibrational frequencies and the B–D bond length. We show here that for the BH4 anion highly diluted in halide matrices AX, A = Na, K; X = Cl, Br, I, a simple space argument does not consistently predict extension or contraction of B–H bond distances. Instead, details of the effective potential function are found to be more important. First principles calculations support the conclusion that there are only rather small changes in B–H bond distances although there are large variations in lattice parameter dependent on composition.

2. Materials and Methods

2.1. Experimental Study of BH4 Anion Isolated in AX Halides

Commercial samples of sodium borohydride NaBH4 (Merck, ≥98%), potassium borohydride (53.94 g·mol−1, Aldrich, 99.9%), sodium chloride NaCl (Merck, ≥9.5%), sodium bromide NaBr (Fluka, ≥99.0%), sodium iodide NaI (Fluka, ≥99.0%), potassium chloride KCl (Merck, ≥99.5%), potassium bromide KBr (Roth) and potassium iodide KI (Riedel-de Haën, 99–100.5%) were used in this study.
Attenuated total reflectance FTIR-ATR spectra of as-received NaBH4 and KBH4 samples were recorded on the Bruker ifs 66 v/S using an ATR accessory with a diamond crystal. The vibrational spectra of borohydrides can be divided into two regions: the bond bending region (with ν2 and ν4, the symmetric and asymmetric bending modes, respectively) from 1050 to 1300 cm−1 and the bond stretching region (with ν1 and ν3, the symmetric and asymmetric stretching modes, respectively) from 2100 to 2500 cm−1 [32]. The asymmetric modes ν4 and ν3 are IR active. The B–H bond bending vibrations occur at about half the bond stretching vibrations implying overtone (2ν4) and combination (ν2 + ν4) with Fermi resonance [35,36].
MBH4/AX pellets (M = Na and K) of 13 mm diameter were produced by exerting a force of 100 kN (equivalent to a pressure of 754 MPa) on a mixture consisting approximately of 0.5 wt% MBH4 and 99.5 wt% AX for one minute. Fourier transform infrared (FTIR) spectra of NaBH4 diluted in different halide matrices were taken using the Bruker Vertex 80 v FTIR spectrometer under vacuum. NaBH4/NaCl and NaBH4/KCl pellets were further heated under vacuum to 450 °C. IR spectra after the thermal treatment were recorded on the Bruker ifs 66 v/S FTIR spectrometer.

2.2. First Principles Calculations

All calculations were performed with the crystalline orbital program CRYSTAL14 [37,38]. All parameters and functionals were tested in a previous work [15]. We used the PWGGA [39] functional, CRYSTAL standard basis sets for sodium, potassium, boron, hydrogen and chlorine, a TZVP basis set for bromine with an additional f-function and pseudopotential basis sets for rubidium, cesium and iodine (Table 1).
The truncation criteria in all calculations for bielectronic integrals were set to 10−9 (overlap and penetration threshold for Coulomb-integrals, overlap threshold for exchange-integrals, pseudo-overlap) and 10−18 (pseudo-overlap), respectively. The SCF (self-consistent-field) convergence threshold on total energy was set to 10−7 a.u. for geometry optimizations and to 10−10 a.u. for frequency calculations, respectively. A Monkhorst-Pack-net of 8 × 8 × 8 was employed and a maximum trust radius of 0.3 Bohr was used in the optimizations. The Anderson method for accelerating convergence was used.
We used the experimental structures [32,47] as a starting point and fixed the lattice constants in the geometry optimizations, but not the atomic positions. Due to the computational afford, we calculated no supercells for dilution of BH4 in metal halides but unit cells with one BH4 and three halides, which is a ratio of 25% of BH4. The frequencies were calculated with fully optimized atomic positions. The antisymmetric stretching mode was enharmonically corrected [15]. The calculated modes were visualized with the program Jmol [48] and could thereby be assigned to the different mode types.

3. Results and Discussion

3.1. Spectra of NaBH4/KBH4 in ATR and Pressed in Various Halides

The ATR spectra of as-received NaBH4 and KBH4 powders are shown in Figure 1. The positions of the peak maxima are in good agreement with values reported by Renaudin et al. [32] as collected in Table 2 together with the assignment given. It is well known that the bands in the ATR spectra are generally shifted to lower frequencies compared to the transmission spectra due to the dispersion of the refractive index. This is one reason that the IR peak positions obtained using the standard pressed pellet technique, e.g., in KBr halide shows deviations from peak positions determined using ATR.
Another reason for deviations in the peak positions determined in the pressed pellet method is related to a rather uncontrolled anion or cation exchange effect between the probe and the matrix. This is shown in Figure 2 for NaBH4 in KBr (spectrum b, c). Rather broad and less well-determined peak positions are observed for short mixing and pressing time (30 s, spectrum b). At variance, good mixing and 90 s pressing at the same force (100 kN/13 mm diameter sample = 754 MPa) reveal much sharper and shifted peak positions (spectrum c). The positions of the peak maxima are also given in Table 2 for comparison. The better mixing and longer pressing time obviously leads to better resolved peaks, e.g., for the isotope effect 11B/10B depicted by the shoulder at higher wavenumbers seen for ν4 and 2ν4. This better resolution is explained, however, by a higher degree of solution of the BH4 anion into the KBr matrix. In this respect, the broader peaks seem to be more representative for NaBH4. This is supported by inspection of the spectrum obtained for NaBH4 diluted and pressed into NaCl also shown in Figure 2 (spectrum a). The peaks well coincide with those of NaBH4 in KBr also taken from the typically broader one. Spectra given for NaBH4 in NaBr (d) and KI (e) also show rather sharp peaks but at significant different peak positions. This indicates the effect of a sufficient dilution and incorporation of the BH4 anion in the various metal halide type lattices.
It has been reported that the BH4 anions are highly diluted in AX by heating discs constituted of AX and traces of borohydride (NaBH4 or KBH4) to temperatures between 500 and 600 °C [49]. For our purpose, this effect is demonstrated in Figure 3. Shown are spectra for NaBH4 (1 mg) diluted into 200 mg of NaCl and KCl (dashed curves) compared to those after heating to 450 °C (red solid lines). For NaBH4 in NaCl after the thermal treatment, the IR peaks appeared to be sharpened and shifted to higher wavenumbers compared to spectra taken before the heating. For NaBH4 diluted and pressed with KCl, the spectrum shows peak splitting indicating the formation of two types of solid solution with a higher and smaller BH4 anion concentration in KCl. However, after thermal treatment, a more homogeneous dilution of BH4 anions in KCl crystals is obtained.

3.2. BH4-Frequency Variations Dependences on Halide Parameters

The pellet preparation procedure using a pressure of approximately 754 MPa was sufficient to solve NaBH4 (a0 = 6.16 Å) in NaI (a0 = 6.48 Å), NaBr (a0 = 5.97 Å), KBr (a0 = 6.60 Å) and KI (a0 = 7.06 Å), whereas for NaCl (a0= 5.64 Å) and KCl (a0 = 6.29 Å), a subsequent annealing to 450 °C was required to achieve the total solution of BH4 in the halide lattice. This can be related to the smaller effective ionic radius of Cl (1.80 Å [28]) compared to BH4 (2.03 Å [29]). On the other hand, the ionic radii of Br (1.96 Å [29]) and I (2.20 Å [30]) are close or significantly larger, respectively. Therefore, a much better exchange is already achieved by the pressure effect. The larger cation radius of K+ compared to Na+ may enhance the solution effect as could be suggested from the two type of compositions indicated in the NaBH4 spectrum in KCl compared to NaBH4 in NaCl (Figure 3).
The obtained IR frequencies of BH4 in different halide matrices in presumably the high dilution limit, denominated as i-BH4 (i for isolated) in the following, are summarized in Table 3 and compared to previous works [49,50]. The IR values for BH4 in RbX matrices are taken from ref. [50]. The cell parameter (a0) [47] and the radius ratio of the effective ionic radius (NaCl structure, coordination number 6) of the cation to the ionic radius of the anion (R = rA/rX) [51] are also given in Table 3.
It is observed that ν3(i-BH4) is similar in NaI and RbBr, 2281 and 2279 cm−1, respectively, while the corresponding a0 (6.479 Å and 6.877 Å, respectively) and ion radius ratio R (0.4636 and 0.7755, respectively) of the halide matrices are very different. Therefore, the frequencies of ν3(i-BH4) cannot be interpreted in terms of only the lattice parameter a0 or R determining the available space for the substituting BH4 in AX. At variance, the IR frequencies of i-BH4 are dependent on the nature of the host matrix, as suggested in previous works for many polyatomic ions isolated in halide matrices [30,50,]. To show this effect more clearly, graphical representations of ν3(i-BH4) as a function of the structural parameters, a0 and R of the corresponding AX, are given in Figure 4a,b, respectively.
Approximately, linear relationships are observed between ν3(i-BH4) and a0(AX) for the homologous series of halides with same cationic element A (A-series, solid red lines) or with the same anionic element X (X-series, dashed blue lines). The linearity between ν3(i-BH4) and a0(AX) applies within each of the defined series implying a decrease of ν3(i-BH4) with increasing a0(AX) (Figure 4a). Furthermore, according to Figure 4b, an increase of ν3(i-BH4) is accompanied with an increase of R for the A-series (for a given A-series rA is constant while rX is decreasing) but with a decrease of R for the X-series (for a given X-series rA is increasing while rX is maintained constant). This indicates that ν3(i-BH4) varies in an opposite direction to the ionic radii of the cations (rA) and the anions (rX). Smaller cations and anions tend to increase ν3(i-BH4). In other words, cations and anions of higher Pauling electronegativity (χP) within the A- or X-series exhibit higher ν3(i-BH4) since χP varies in an opposite direction to that observed for the ionic radius in the periodic table (χP increases by going from left to right across a period, and decreases by moving down a group). This still could indicate that increasing the available space for the substituting BH4 anion along a given A-series with increasing anion size (for e.g., NaCl, NaBr and NaI) decreases ν3 by increasing the B–H distances. A similar effect occurs with increasing the size of the cation along a given X-series (for e.g., NaI, KI and RbI), an elongation of the B–H distances could be responsible of the decrease of ν3(i-BH4). This explanation fails, however, between different series as alluded above for the case of NaI and RbBr which exhibit similar ν3(i-BH4) but completely different available anion spaces. The same holds for NaBr and KCl matrices. Therefore, the main factor responsible of frequency variations is not based on the B–H bond distance considerations in these cases.
The linear relationship established between ν3(i-BH4) and a0(AX) can be presented empirically according to:
ν3(i-BH4) = α.a0(AX) + β,
α is the gradient and β a constant obtained from the linear regressions and given in Table 4. The R2 values (>0.98) are satisfactory and confirm the effectiveness of the correlations.
The absolute value of α of the A-series decreases when descending group 1 of the periodic table, i.e., in the order Na-series > K-series > Rb-series. Similarly, the absolute values of α of the X-series decrease in the order Cl-series > Br-series > I-series. Assuming that the gradient α of the linear regression gives insight into the rate at which changes are taking place, i.e., how much the variations of a0(AX) would affect the values of ν3(i-BH4), obtained results suggest that the dependency of ν3(i-BH4) on a0(AX) diminishes across group 1 and group 17 of the periodic table, i.e., when increasing rA of A-series and rX of X-series, respectively. Furthermore, the absolute values of α of A-series are greater than those of the X-series. Since in a given A-series, the anionic element is varying and for a given X-series, the cationic element is changing; the obtained results suggest that the anionic element X of AX has more impact on ν3(i-BH4) than the cationic element A.
The interrelations established between ν3(i-BH4), a0(AX) and R(AX) can be extended and discussed in term of other parameters of AX as the negative value of the standard enthalpy of formation (ΔfH, or exothermicity = −ΔfH), the fractional ionic character of the bond for AX (Ic) and the electronegativity χP of A and X, respectively. It has been reported that -ΔfH increase with increase of Ic of AX. Moreover, both parameters increase when R increases within a given A-series (rX is changing) or X-series (rA is changing) with the increase being deeper for A-series [51]. By combining these correlations with those deduced from Figure 4, it can be deduced that for the same cationic A-series, ν3(i-BH4) is positively correlated to -ΔfH and Ic of AX, i.e., an increase of ν3(i-BH4) is accompanied with an increase of -ΔfH and Ic of AX, respectively (Figure 5). For the homologous series with the same anionic element, an inverse correlation lies between ν3(i-BH4) and -ΔfH and Ic of AX, i.e., an increase of ν3(i-BH4) is accompanied with a decrease of -ΔfH and Ic of AX, respectively.
BH4 anion isolated in AX replaces the anion X at individual sites in the alkali halide lattice. Several complex empirical functions have been developed to carry out a quantitative evaluation of the frequencies shift of an isolated ion in different surroundings. In a general way, the potential function (V′) governing the vibration of a molecular ion substituted in NaCl-type cube can be regarded as the sum of the potential function of the free ion (VF) and the interaction energy (V) arising from the presence of the environment expressed in terms of the interaction energy (U) (Equation (4)) [52].
V’ = VF + V,
The perturbing potential function (V, expressed in terms of the interaction energy) has two distinct effects on the considered ion: (i) the second derivatives of the perturbing potential with respect to appropriate symmetry coordinates which represent additional contributions to the effective force constants of the impurity ion and (ii) the first derivatives of the perturbing potential represent forces which modify the equilibrium internuclear spacing within the ion [52]. Accordingly, the frequency shift of the i th normal mode of the solute is given by:
Δ ν i ν i   =   1 2 k i 2 V S i 2     3 a i i i a i i V S i     r a i i r a r r V S r ,
where k i is the force constant of the i th mode, S i are the normal co-ordinates and a are the potential constants [53]. For a diatomic ion described by a Morse potential of the form V F   =   D 1     e     β ξ 2 the equation is reduced to:
Δ ν ν   =   1 2 k 2 V ξ 2   +   3 β V ξ ,
where ξ   =   r     r e is the change in the internuclear spacing [53,54]. The perturbing potential can be explicitly expressed based on the classical theory of ionic crystals given by Born and Mayer as the sum of a polarization term (Vp) arising from static or dynamic polarization of the isolated ion or the surrounding lattice, a columbic term (Vc) describing the charge–charge interaction between the solute ion and the supporting alkali halide lattice and a repulsive short-range term (Vs) due to the overlap of electron density of the isolated ion with its surrounding ions of the lattice [52,54]. In most cases where the terms of Equation (4) or its equivalent have been evaluated, it was found that the first derivative term (force multiplied by anharmonicity) dominated the frequency shift calculations [53]. For example, for the cyanide ion CN isolated in various alkali halides [54], the authors observed that the frequency shifts calculations are sensitive to very small changes in the position of the nearest neighbor ions and are dominated by the repulsive term shortening of the CN bond.
An approximately linear relation is obtained between ν3(i-BH4) and the short-range lattice energies of AX (Figure 6). The short lattice energy of AX can be considered as the contribution of the repulsive, dipole–dipole and dipole–quadrupole interaction energies given by the Born–Mayer equation AX. These values are tabulated in Table 5 and make clear evidence that the repulsive forces constitute the major part. Similar linear correlation was reported for the nitrate anion NO3 embedded in different halide matrices [55]. We also obtained the same linear dependency for CN in different halides basing on the IR frequencies of reference [54].
The correlations between ν3(i-BH4) and χP(A) and χP(X) already discussed for the different series, can provide information about the stability of the borohydrides. Cations and anions of higher χP within the A- or X-series exhibit higher ν3(i-BH4). A more pronounced charge transfer to BH4 anion results in stronger B–H bonds stabilizing the borohydride [1]. A more electronegative cation destabilizes the borohydride by attracting the electron density and reducing the electron density in the B–H bonds [19]. For MBH4 salts, the thermal stability increases in the order NaBH4 < KBH4 < RbBH4 < CsBH4 [57], whereas the χP of the cation increases in the opposite way. Orimo et al. [57] found that for lithium, sodium and potassium borohydrides, the Raman active stretching ν1 and bending ν2 modes decrease in the order (ν2 (LiBH4) = 1295 cm−1 and ν2′ (LiBH4) = 1305 cm−1) > (ν2 (NaBH4) = 1280 cm−1) > (ν2 (KBH4) = 1240 cm−1) and (ν1 (NaBH4) = 2325 cm−1) > (ν1 (KBH4) = 2305 cm−1), whereas the melting temperature (Tm) varies in the opposite direction, Tm (LiBH4) < Tm (NaBH4) < Tm (KBH4). The value of ν1 of LiBH4 was reported at 2293 cm−1 and did not follow the relation due to the difference of the crystal system. LiBH4 adopts an orthorhombic structure (space group Pcmn) at room temperature, whereas the other alkali metal (Na, K, Rb and Cs) borohydrides crystallize in a rock salt, face-centered-cubic structure. According to the ATR spectra of Renaudin et al. [32], ν3 is found to decrease in the order NaBH4 > KBH4 > RbBH4 > CsBH4. This indicates an inverse correlation between Td of the borohydride and ν3 of BH4 anion. Based on these results, it can be said that in a given X-group where A is changing, an increase of ν3(i-BH4) would be accompanied with a decrease of Td. At the same time, a0 (AX), R (AX), Ic (AX) and -ΔH (AX) decrease.
DSC data reveal stabilization with increasing chloride-substitution of NaBH4 [28]. Sieverts measurements of the hexagonal Li(BH4)0.5Br0.5 indicate similar H2 release as for the unsubstituted pattern LiBH4 [29]. This indicates that the substitution, with a more electronegative anion, localizes the negative charge on the B-H bond, stabilizing, thus, the borohydride. Therefore, it could be deduced that for the A-series where the anionic element X is varying, an increase of ν3(i-BH4) is accompanied with an increase of Td. At the same time, a0 (AX) decreases, R (AX), Ic (AX) and -ΔfH (AX) are increased. Further experiments are still needed to demonstrate the role of the electronegativity of the anion and the thermal stability of the borohydride.

3.3. BH4-Frequency Variations Dependences on Fundamental Parameter Sets in DFT

The experimental IR frequencies of i-BH4 in NaX (Cl, Br, I) are shown in Figure 7a as a function of the lattice parameter a0 of the halides: ν4 is represented by 2*ν4 and obtained by simply doubling the values of ν4 (* denotes the multiplication), together with the three other peaks, namely 2ν4F− and ν3F+ (the subscript F− and F+ are used to denote the Fermi resonance effect) and the combination mode ν2 + ν4. Values given by Renaudin et al. [32] for compositions MBH4, M = N, K, Rb, Cs are also included for comparison (the values are taken as obtained by ATR-method for 2ν4F− and ν3F+ and ν2 + ν4; for 2*ν4 we used the Raman values of v4). All frequencies increase with decreasing lattice parameter as discussed above. It is observed that the frequency-dependence on the lattice parameter is weaker for MBH4 compared to i-BH4 in NaX. The overtone frequency 2*ν4 comes in all cases well between 2ν4F− and ν3F+. This visualizes the Fermi resonance effect characteristic of the BH4 anion which occurs between ν3 and the overtone and combination modes. The Fermi resonance observed in the Raman and IR spectra of the tetrahedral BH4 occurs between the stretching and overtone and combination modes and redistributes the intensities and changes the frequencies [35,36]. As a result, the overtone 2ν4 and combination ν2 + ν4 modes appear with much greater intensity than would have been expected and the IR frequencies are shifted from the values at which they would otherwise have appeared (2ν4F− < 2*ν4 < ν3F+).
The same trends are obtained from DFT calculations at generalized gradient approximation (GGA) level (Figure 7b). It has to be noted that the calculations could not be completed for i-BH4 directly, i.e., in the highly diluted limit. Instead the calculations were carried out for compositions Na(BH4)0.25X0.75 with X = Cl, Br, I. The Fermi resonance effect cannot be considered since CRYSTAL does not allow for incorporation of anharmonic effects. Thus, simple approximations of overtones and combination bands had to be made. Therefore, the theoretical values for 2 × ν4 and ν3, as well as the combination mode ν2 + ν4 (by simple addition of the harmonic values) are shown related on the lattice parameter values as obtained with DFT. A quasi-linear relationship between lattice parameter and composition Na(BH4)1-xXx with x = 0, 0.25, 0.5, 1 and X = Cl, Br or I, was obtained, where the parameter values are in excellent agreement with experiment for the endmembers. Moreover, the frequency variation dependent on lattice parameter is more pronounced for NaX, X = Cl, Br, I, compared to the series MBH4 with M = Na, K, Rb, Cs. The absolute values of calculated and measured frequencies and lattice parameters are close. Different from the expectations, the theoretically optimized structures show only slight variations in B–H bond distances. The changes are about one order of magnitude smaller compared to the data given by Renaudin et al. [32]. Rather small variation were also obtained by others using ab initio methods [58]. We therefore conclude that the B–H bond variation is not responsible for the observed frequency changes, but rather the long-distance tail of the B–H potential which is affected by the nearest and next-nearest neighbors.

4. Conclusions

The nearest-neighbor coordination has significant influence on the B–H bonding strength of the BH4 anion predicting the variation of the IR frequencies in different halides. The BH4 anion isolated in AX crystal (i-BH4/AX) yields very sharp IR absorption bands. The IR frequencies of i-BH4 are found to depend on the nature of the AX matrix. Systematic variations are observed between the asymmetrical stretching ν3(i-BH4) and different structural parameters of AX. The correlations established are applicable to homologous series of AX possessing the same cationic A element (A-series: Na-, K-, Rb-series) or the same anionic X element (X-series: Cl-, Br and I-series). For example, when the cationic element A of AX is kept constant for the A-series of AX (A-series: Na-, K- and Rb-series), ν3(i-BH4) decreases linearly with increasing cell parameter a0 of AX. A similar quasi-linear trend is observed for the respective anionic X-series. The correlation between ν3(i-BH4) and a0(AX) for the homologous series is almost linear and can be written as ν3(i-BH4) = α.a0 (AX) + β. The absolute value of the slope decreases when increasing the ionic radii of the cation A or the anion X of A-groups and X-groups, respectively. Moreover, higher values are obtained for the A-groups compared to the X-groups. This indicates that smaller cations and anions produces deeper changes in ν3(i-BH4) and that ν3(i-BH4) is strongly affected by the nature of X rather than A. The correlation study on the ν3(i-BH4) is extended in terms of the ratio of the effective ionic radius of the cation to the ionic radius of the anion (R = rA/rX), the Pauling electronegativity (χP) of A and X, respectively, the ionic character (Ic) and the standard enthalpy of formation of AX (ΔfH). An approximately linear relationship is observed between ν3(i-BH4) and the contribution of the short-range interactions energies of AX pointing the strong impact of the repulsive forces between BH4 and its nearest neighbors on ν3(i-BH4). DFT calculations obtain a reasonable agreement with experimental values, without showing significant variations of the B–H bond length.

Author Contributions

Investigation, Z.A., A.G.S., A.C.U., T.B. and C.H.R.; Methodology, Z.A., A.G.S., A.C.U., T.B. and C.H.R.; Writing—Review & Editing, Z.A., A.G.S., A.C.U., T.B. and C.H.R. All authors have read and agreed to the published version of the manuscript.

Funding

ZA was funded by DAAD Ref 441: A/10/97515. CHR used LUH-internal support and RU764/6-1, AGS and ACU were funded within the DFG project of BR1768/8-1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data obtained as described.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Züttel, A.; Borgschulte, A.; Orimo, S.I. Tetrahydroborates as new hydrogen storage materials. Scr. Mater. 2007, 56, 823–828. [Google Scholar] [CrossRef]
  2. Schlesinger, H.I.; Brown, H.C.; Finholt, A.E.; Gilbreath, J.R.; Hoekstra, H.R.; Hyde, E.K. Sodium Borohydride, Its Hydrolysis and its Use as a Reducing Agent and in the Generation of Hydrogen. J. Am. Chem. Soc. 1953, 75, 215–219. [Google Scholar] [CrossRef]
  3. Demirci, U.B.; Akdim, O.; Andrieux, J.; Hannauer, J.; Chamoun, R.; Miele, P. Sodium borohydride hydrolysis as hydrogen generator: Issues, state of the art and applicability upstream froma fuel cell. Fuel Cells 2010, 10, 335–350. [Google Scholar] [CrossRef] [Green Version]
  4. Urgnani, J.; Torres, F.J.; Palumbo, M.; Baricco, M. Hydrogen release from solid state NaBH4. Int. J. Hydrogen Energy 2008, 33, 3111–3115. [Google Scholar] [CrossRef]
  5. Mao, J.; Guo, Z.; Nevirkovets, I.P.; Liu, H.K.; Dou, S.X. Hydrogen De-/absorption improvement of NaBH4 catalyzed by titanium-based additives. J. Phys. Chem. C 2012, 116, 1596–1604. [Google Scholar] [CrossRef] [Green Version]
  6. Humphries, T.D.; Kalantzopoulos, G.N.; Llamas-Jansa, I.; Olsen, J.E.; Hauback, B.C. Reversible hydrogenation studies of NaBH4 milled with Ni-containing additives. J. Phys. Chem. C 2013, 117, 6060–6065. [Google Scholar] [CrossRef] [Green Version]
  7. Ngene, P.; Van Den Berg, R.; Verkuijlen, M.H.W.; De Jong, K.P.; De Jongh, P.E. Reversibility of the hydrogen desorption from NaBH4 by confinement in nanoporous carbon. Energy Environ. Sci. 2011, 4, 4108–4115. [Google Scholar] [CrossRef] [Green Version]
  8. Peru, F.; Garroni, S.; Campesi, R.; Milanese, C.; Marini, A.; Pellicer, E.; Baro, M.D.; Mulas, G. Ammonia-free infiltration of NaBH4 into highly-ordered mesoporous silica and carbon matrices for hydrogen storage. J. Alloys Compd. 2013, 580, S309–S312. [Google Scholar] [CrossRef]
  9. Zhang, L.; Xiao, X.; Fan, X.; Li, S.; Ge, H.; Wang, Q.; Chen, L. Fast hydrogen release under moderate conditions from NaBH4 destabilized by fluorographite. RSC Adv. 2014, 4, 2550–2556. [Google Scholar] [CrossRef]
  10. Rude, L.H.; Filsø, U.; D’Anna, V.; Spyratou, A.; Richter, B.; Hino, S.; Jensen, T.R. Hydrogen-fluorine exchange in NaBH4-NaBF4. Phys. Chem. Chem. Phys. 2013, 15, 18185–18194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Zhang, Z.G.; Wang, H.; Zhu, M. Hydrogen release from sodium borohydrides at low temperature by the addition of zinc fluoride. Int. J. Hydrogen Energy 2011, 36, 8203–8208. [Google Scholar] [CrossRef]
  12. Buhl, J.C.; Gesing, T.M.; Rüscher, C.H. Synthesis, crystal structure and thermal stability of tetrahydroborate sodalite Na8[AlSiO4]6(BH4)2. Microporous Mesoporous Mater. 2005, 80, 57–63. [Google Scholar] [CrossRef]
  13. Buhl, J.C.; Schomborg, L.; Rüscher, C.H. Enclosure of Sodium Tetrahydroborate (NaBH4) in Solidified Aluminosilicate Gels and Microporous Crystalline Solids for Fuel Processing. In Hydrogen Storage; InTechOpen: London, UK, 2012. [Google Scholar]
  14. Rüscher, C.H. Boronhydride-geopolymer composites. J. Ceram. Sci. Technol. 2017, 8, 399–410. [Google Scholar] [CrossRef]
  15. Schneider, A.G.; Bredow, T.; Schomborg, L.; Rscher, C.H. Structure and IR vibrational spectra of Na8[AlSiO4]6(BH4)2: Comparison of theory and experiment. J. Phys. Chem. A 2014, 118, 7066–7073. [Google Scholar] [CrossRef]
  16. Rüscher, C.H.; Schomborg, L.; Bredow, T. Experiments on the thermal activation of hydrogen release of NaBH4-sodalites characterized by IR- and MAS-NMR spectroscopy. Int. J. Hydrogen Energy 2022. under review. [Google Scholar]
  17. Mesmer, R.E.; Jolly, W.L. The Hydrolysis of Aqueous Hydroborate. Inorg. Chem. 1962, 1, 608–612. [Google Scholar] [CrossRef]
  18. Kreevoy, M.M.; Hutchins, J.E.C. H2BH3 as an Intermediate in Tetrahydridoborate Hydrolysis. J. Am. Chem. Soc. 1972, 94, 6371–6376. [Google Scholar] [CrossRef]
  19. Nakamori, Y.; Miwa, K.; Ninomiya, A.; Li, H.; Ohba, N.; Towata, S.I.; Orimo, S.I. Correlation between thermodynamical stabilities of metal borohydrides and cation electronegativites: First-principles calculations and experiments. Phys. Rev. B—Condens. Matter Mater. Phys. 2006, 74, 1–9. [Google Scholar] [CrossRef] [Green Version]
  20. Nakamori, Y.; Li, H.W.; Kikuchi, K.; Aoki, M.; Miwa, K.; Towata, S.I.; Orimo, S.I. Thermodynamical stabilities of metal-borohydrides. J. Alloys Compd. 2007, 446–447, 296–300. [Google Scholar] [CrossRef]
  21. Matsunaga, T.; Buchter, F.; Miwa, K.; Towata, S.; Orimo, S.; Züttel, A. Magnesium borohydride: A new hydrogen storage material. Renew. Energy 2008, 33, 193–196. [Google Scholar] [CrossRef]
  22. Graetz, J. New approaches to hydrogen storage. Chem. Soc. Rev. 2009, 38, 73–82. [Google Scholar] [CrossRef] [PubMed]
  23. Hinze, J. The concept of electronegativity of atoms in molecules. Theor. Comput. Chem. 1999, 6, 189–212. [Google Scholar] [CrossRef]
  24. Ravnsbæk, D.; Filinchuk, Y.; Cerenius, Y.; Jakobsen, H.J.; Besenbacher, F.; Skibsted, J.; Jensen, T.R. A Series of Mixed-Metal Borohydrides. Angew. Chem. 2009, 121, 6787–6791. [Google Scholar] [CrossRef]
  25. Brinks, H.W.; Fossdal, A.; Hauback, B.C. Adjustment of the stability of complex hydrides by anion substitution. J. Phys. Chem. C 2008, 112, 5658–5661. [Google Scholar] [CrossRef]
  26. Yin, L.; Wang, P.; Fang, Z.; Cheng, H. Thermodynamically tuning LiBH4 by fluorine anion doping for hydrogen storage: A density functional study. Chem. Phys. Lett. 2008, 450, 318–321. [Google Scholar] [CrossRef]
  27. 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]
  28. Olsen, J.E.; Sørby, M.H.; Hauback, B.C. Chloride-substitution in sodium borohydride. J. Alloys Compd. 2011, 509, L228–L231. [Google Scholar] [CrossRef]
  29. Rude, L.H.; Zavorotynska, O.; Arnbjerg, L.M.; Ravnsbæk, D.B.; Malmkjær, R.A.; Grove, H.; Jensen, T.R. Bromide substitution in lithium borohydride, LiBH4-LiBr. Int. J. Hydrogen Energy 2011, 36, 15664–15672. [Google Scholar] [CrossRef] [Green Version]
  30. Rude, L.H.; Groppo, E.; Arnbjerg, L.M.; Ravnsbaek, D.B.; Malmkjaer, R.A.; Filinchuk, Y.; Jensen, T.R. Iodide substitution in lithium borohydride, LiBH4-LiI. J. Alloys Compd. 2011, 509, 8299–8305. [Google Scholar] [CrossRef] [Green Version]
  31. Vajeeston, P.; Ravindran, P.; Kjekshus, A.; Fjellvåg, H. Structural stability of alkali boron tetrahydrides ABH4 (A=Li, Na, K, Rb, Cs) from first principle calculation. J. Alloys Compd. 2005, 387, 97–104. [Google Scholar] [CrossRef]
  32. Renaudin, G.; Gomes, S.; Hagemann, H.; Keller, L.; Yvon, K. Structural and spectroscopic studies on the alkali borohydrides MBH4 (M = Na, K, Rb, Cs). J. Alloys Compd. 2004, 375, 98–106. [Google Scholar] [CrossRef] [Green Version]
  33. Harvey, K.B.; McQuaker, N.R. Infrared and Raman Spectra of Potassium and Sodium Borohydride. Can. J. Chem. 1971, 49, 3272–3281. [Google Scholar] [CrossRef]
  34. Heyns, A.M.; Schutte, C.J.H.; Scheuermann, W. Low-temperature infrared studies. IX. The infrared and Raman spectra of sodium d4-borohydride NaBD4. J. Mol. Struct. 1971, 9, 271–281. [Google Scholar] [CrossRef]
  35. Memon, M.I.; Wilkinson, G.R.; Sherman, W.F. Vibrational studies of BH4- and BD4- isolated in alkali halides. J. Mol. Struct. 1982, 80, 113–116. [Google Scholar] [CrossRef]
  36. Memon, M.I.; Sherman, W.F.; Wilkinson, G.R. Fermi resonances in the Raman spectra of alkali halide/BH4-. J. Mol. Struct. 1984, 115, 213–216. [Google Scholar] [CrossRef]
  37. Dovesi, R.; Orlando, R.; Civalleri, B.; Roetti, C.; Saunders, V.R.; Zicovich-Wilson, C.M. CRYSTAL: A computational tool for the ab initio study of the electronic properties of crystals. Z. Für. Krist.–Cryst. Mater. 2005, 220, 571–573. [Google Scholar] [CrossRef]
  38. Dovesi, R.; Roberto, O.; Alessandro, E.; Claudio, M.Z.-W.; Bartolomeo, C.; Silvia, C.; Lorenzo, M.; Matteo, F.; Marco, D.L.P.; Philippe, D.; et al. CRYSTAL14 User’s Manual; University of Turin: Torino, Italy, 2014; p. 382. [Google Scholar]
  39. Perdew, J.P.; Yue, W. Accurate and simple density functional for the electronic exchange energy: Generalized gradient approximation. Phys. Rev. B Condens. Matter 1986, 33, 8800–8802. [Google Scholar] [CrossRef]
  40. Dovesi, R.; Roetti, C.; Freyria-Fava, C.; Prencipe, M.; Saunders, V.R. On the elastic properties of lithium, sodium and potassium oxide. An ab initio study. Chem. Phys. 1991, 156, 11–19. [Google Scholar] [CrossRef]
  41. Leininger, T.; Nicklass, A.; Küchle, W.; Stoll, H.; Dolg, M.; Bergner, A. The accuracy of the pseudopotential approximation: Non-frozen-core effects for spectroscopic constants of alkali fluorides XF (X = K, Rb, Cs). Chem. Phys. Lett. 1996, 255, 274–280. [Google Scholar] [CrossRef]
  42. Dovesi, R.; Ermondi, C.; Ferrero, E.; Pisani, C.; Roetti, C. Hartree-Fock study of lithium hydride with the use of a polarizable basis set. Phys. Rev. B 1984, 29, 3591–3600. [Google Scholar] [CrossRef]
  43. Orlando, R.; Dovesi, R.; Roetti, C.; Saunders, V.R. Ab initio Hartree-Fock calculations for periodic compounds: Application to semiconductors. J. Phys. Condens. Matter 1990, 2, 7769–7789. [Google Scholar] [CrossRef]
  44. Apra, E.; Causa, M.; Prencipe, M.; Dovesi, R.; Saunders, V.R. On the structural properties of {NaCl}: An ab initio study of the B1-B2 phase transition. J. Phys. Condens. Matter 1993, 5, 2969–2976. [Google Scholar] [CrossRef]
  45. Peintinger, M.F.; Oliveira, D.V.; Bredow, T. Consistent Gaussian basis sets of triple-zeta valence with polarization quality for solid-state calculations. J. Comput. Chem. 2013, 34, 451–459. [Google Scholar] [CrossRef] [PubMed]
  46. Stoll, H.; Metz, B.; Dolg, M. Relativistic energy-consistent pseudopotentials--recent developments. J. Comput. Chem. 2002, 23, 767–778. [Google Scholar] [CrossRef] [PubMed]
  47. The Inorganic Crystal Structure Database (ICSD). Available online: https://icsd.products.fiz-karlsruhe.de/ (accessed on 9 March 2022).
  48. Hanson, R.M. Jmol—A paradigm shift in crystallographic visualization. J. Appl. Crystallogr. 2010, 43, 1250–1260. [Google Scholar] [CrossRef] [Green Version]
  49. Hisatsune, I.C.; Suarez, N.H. Infrared Spectra of Metaborate Monomer and Trimer Ions. Inorg. Chem. 1964, 3, 168–174. [Google Scholar] [CrossRef]
  50. Ketelaar, J.A.A.; Schutte, C.J.H. The borohydride ion (BH4-) in a face-centred cubic alkali-halide lattice. Spectrochim. Acta 1961, 17, 1240–1243. [Google Scholar] [CrossRef]
  51. Nasar, A. Correlation between standard enthalpy of formation, structural parameters and ionicity for alkali halides. J. Serb. Chem. Soc. 2013, 78, 241–253. [Google Scholar] [CrossRef] [Green Version]
  52. Bryant, J.I.; Turrell, G.C. Infrared Spectra of the Azide Ion in Alkali-Halide Lattices. J. Chem. Phys. 1962, 37, 1069–1077. [Google Scholar] [CrossRef]
  53. Sherman, W.F.; Wilkinson, G.R. Infrared and raman studies on the vibrational spectra of impurities in ionic and covalent crystals. In Vibrational Spectroscopy of Trapped Species; Wiley: London, UK, 1973; p. 291. [Google Scholar]
  54. Field, G.R.; Sherman, W.F. Cyanide ion—Environmental perturbation of its vibrational and rotational motion when isolated in alkali halides. J. Chem. Phys. 1967, 47, 2378–2389. [Google Scholar] [CrossRef]
  55. Strasheim, A.; Buijs, K. Infrared absorption of nitrate ions dissolved in solid alkali halides. J. Chem. Phys. 1961, 34, 691–692. [Google Scholar] [CrossRef]
  56. Seitz, F. The Modern Theory of Solids; McGraw-Hill Book Company, Incorporated: New York, NY, USA, 1940. [Google Scholar]
  57. Orimo, S.; Nakamori, Y.; Züttel, A. Material properties of MBH4 (M = Li, Na, and K). Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2004, 108, 51–53. [Google Scholar] [CrossRef]
  58. Zhang, X.D.; Hou, Z.F.; Jiang, Z.Y.; Hou, Y.Q. Elastic properties of MBH4 (M = Na, K, Rb, Cs). Phys. B Condens. Matter 2011, 406, 2196–2199. [Google Scholar] [CrossRef]
Figure 1. FTIR−ATR spectra of the as-received NaBH4 and KBH4 samples.
Figure 1. FTIR−ATR spectra of the as-received NaBH4 and KBH4 samples.
Crystals 12 00510 g001
Figure 2. Room temperature transmission IR spectra of as−received NaBH4 in NaCl (a), in KBr short pressed (b) and long pressed (c), in NaBr (d) and in KI (e). Vertical dashed lines are used to show the coincidence of peak positions of ν4 and ν3 of (a,b).
Figure 2. Room temperature transmission IR spectra of as−received NaBH4 in NaCl (a), in KBr short pressed (b) and long pressed (c), in NaBr (d) and in KI (e). Vertical dashed lines are used to show the coincidence of peak positions of ν4 and ν3 of (a,b).
Crystals 12 00510 g002
Figure 3. IR spectra of NaBH4 diluted in (a) NaCl and (b) KCl pellets before (black dashed lines) and after thermal treatment to 450 °C (red solid lines).
Figure 3. IR spectra of NaBH4 diluted in (a) NaCl and (b) KCl pellets before (black dashed lines) and after thermal treatment to 450 °C (red solid lines).
Crystals 12 00510 g003
Figure 4. Graphical representation of ν3(i−BH4) as a function of (a) the cell parameter a0 and (b) the radius ratio (R = rA/rX) of the corresponding AX halide. Red solid lines represent common cations (A-series) and blue dashed lines are for the common anion X-series.
Figure 4. Graphical representation of ν3(i−BH4) as a function of (a) the cell parameter a0 and (b) the radius ratio (R = rA/rX) of the corresponding AX halide. Red solid lines represent common cations (A-series) and blue dashed lines are for the common anion X-series.
Crystals 12 00510 g004
Figure 5. Graphical representation of ν3(i−BH4) in function of (a) the exothermicity (–ΔfH) of AX and (b) the ionic character (Ic) of the bond for AX halides. Red solid lines are for A−series and blue dashed lines are for X−series.
Figure 5. Graphical representation of ν3(i−BH4) in function of (a) the exothermicity (–ΔfH) of AX and (b) the ionic character (Ic) of the bond for AX halides. Red solid lines are for A−series and blue dashed lines are for X−series.
Crystals 12 00510 g005
Figure 6. Graphical representation of ν3(i−BH4) as a function of the contribution of the short-range interaction energies in the corresponding AX. Dashed line is the linear regression.
Figure 6. Graphical representation of ν3(i−BH4) as a function of the contribution of the short-range interaction energies in the corresponding AX. Dashed line is the linear regression.
Crystals 12 00510 g006
Figure 7. (a) Experimental peak positions of v4 expressed in the term of 2*v4, 2v4F−, v3F+ (F−, F+ denotes the Fermi−resonance effect) and v2 + v4 of i-BH4 in NaX, X = Cl, Br, I and MBH4 (taken from Renaudin et al. [32]) as a function of the cell parameter a0. (b) Calculated (first principles) peak positions of BH4 anion in Na(BH4)0.25X0.75 with X = Cl, Br, I and MBH4, M = Na, K, Rb, Cs as a function of calculated unit cell parameters.
Figure 7. (a) Experimental peak positions of v4 expressed in the term of 2*v4, 2v4F−, v3F+ (F−, F+ denotes the Fermi−resonance effect) and v2 + v4 of i-BH4 in NaX, X = Cl, Br, I and MBH4 (taken from Renaudin et al. [32]) as a function of the cell parameter a0. (b) Calculated (first principles) peak positions of BH4 anion in Na(BH4)0.25X0.75 with X = Cl, Br, I and MBH4, M = Na, K, Rb, Cs as a function of calculated unit cell parameters.
Crystals 12 00510 g007
Table 1. Atomic basis sets used in the CRYSTAL calculations.
Table 1. Atomic basis sets used in the CRYSTAL calculations.
AtomBasis Set
Na8-511G [40]
K86-511G [40]
RbECP28MWB [41]
CsECP46MWB [41]
H5-11G * [42]
B6-21G * [43]
Cl86-311G [44]
BrPob-TZVP [45]
IECP [46]
* related to the nomenclature of the basis set.
Table 2. ATR frequencies of NaBH4 and KBH4 compared to ATR values of Renaudin et al. [32] (values in brackets) in addition to the transmission IR frequencies of NaBH4 in KBr and NaCl pellets.
Table 2. ATR frequencies of NaBH4 and KBH4 compared to ATR values of Renaudin et al. [32] (values in brackets) in addition to the transmission IR frequencies of NaBH4 in KBr and NaCl pellets.
ν44ν3ν2 + ν4
ATRNaBH41110 (1110)2217 (2217)2283 (2284)2397 (2404)
KBH41111 (1112)2205 (2208)2268 (2270)2375 (2376)
Transmission
(pellets)
NaBH4/NaCl1121222923022403
NaBH4/KBr
(<30 s pressed)
1120222022992398
NaBH4/KBr
(90 s pressed)
1126222422912387
Table 3. IR frequencies of i-BH4 in different AX. Structural parameters of AX are also given.
Table 3. IR frequencies of i-BH4 in different AX. Structural parameters of AX are also given.
AXIR Frequencies of i-BH4 in AXStructural Parameters of AX
ν44ν3ν2 + ν4a0R = rA/rX
NaCl11652303237524825.64020.5635
1166 *2307 *2373 *2486 *
1167 **2311 **2382 **2493 **
NaBr11382252232624265.97380.5204
1136 *2254 *2328 *2427 *
1135 **2257 **2331 **2430 **
NaI11122208228123646.4790.4636
1111 *2202 *2278 *
KCl11442257232324226.2900.7624
1144 *2257 *2324 *2422 *
1146 **2262 **2328 **2428 **
KBr11272226229323906.5980.7041
1128 *2226 *2293 *2387 *
1128 **2229 **2295 **2391 **
KI11072191225823507.0640.6273
1108 *2191 *2258 *2341 *
1108 **2192 **2258 **2350 **
RbCl1132 **2236 **2303 **2395 **6.5820.8398
RbBr1118 **2212 **2279 **2369 **6.87680.7755
RbI1104 **2181 **2250 **2335 **7.32910.6909
* IR values according to ref. [49]. ** IR values according to ref. [50].
Table 4. Empirical constants (α and β) of the linear relationship ν3(i-BH4) = α.a0(AX) + β for common cation A- series and common anion X-series.
Table 4. Empirical constants (α and β) of the linear relationship ν3(i-BH4) = α.a0(AX) + β for common cation A- series and common anion X-series.
A-Groupsα (cm−1 Å−1)β (cm−1)R2
Na-group−119.92753054.09620.9848
K-group−88.84812884.80240.9928
Rb-group−69.17652757.64900.9956
X-groups
Cl-group−83.14492850.94770.9999
Br-group−56.62202668.95450.9999
I-group−35.71902511.72330.9957
Table 5. Contribution of the short-range forces of the alkali halides as given by the Born–Mayer equation (in kg.cal.mol−1 [56]).
Table 5. Contribution of the short-range forces of the alkali halides as given by the Born–Mayer equation (in kg.cal.mol−1 [56]).
AXRepulsiveDipole-DipoleDipole-QuadrupoleShort Range
NaCl−23.55.20.1−18.2
NaBr−20.65.50.1−15
NaI−17.16.30.1−10.7
KCl−21.57.10.1−14.3
KBr−18.66.90.1−11.6
KI−15.97.10.1−8.7
RbCl−19.97.90.1−11.9
RbBr−17.67.90.1−9.6
RbI−15.47.90.1−7.4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Assi, Z.; Schneider, A.G.; Ulpe, A.C.; Bredow, T.; Rüscher, C.H. The Rigidity of the (BH4)-Anion Dispersed in Halides AX, A = Na, K; X = Cl, Br, I, and in MBH4 with M = Na, K, Rb, Cs. Crystals 2022, 12, 510. https://doi.org/10.3390/cryst12040510

AMA Style

Assi Z, Schneider AG, Ulpe AC, Bredow T, Rüscher CH. The Rigidity of the (BH4)-Anion Dispersed in Halides AX, A = Na, K; X = Cl, Br, I, and in MBH4 with M = Na, K, Rb, Cs. Crystals. 2022; 12(4):510. https://doi.org/10.3390/cryst12040510

Chicago/Turabian Style

Assi, Zeina, Alexander Gareth Schneider, Anna Christina Ulpe, Thomas Bredow, and Claus Henning Rüscher. 2022. "The Rigidity of the (BH4)-Anion Dispersed in Halides AX, A = Na, K; X = Cl, Br, I, and in MBH4 with M = Na, K, Rb, Cs" Crystals 12, no. 4: 510. https://doi.org/10.3390/cryst12040510

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