Next Article in Journal
The Oxidation Behaviour and Notch Wear Formation of TiAlN Coated Tools Using Different Oxidation Techniques
Next Article in Special Issue
Hydrogen Generation by Hydrolysis of MgH2-LiH Composite
Previous Article in Journal
Evaluation of the Hydration Characteristics and Anti-Washout Resistance of Non-Dispersible Underwater Concrete with Nano-SiO2 and MgO
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Polymorphism and Thermal Decomposition Process of (n-C4H9)4NRE(BH4)4 for RE = Ho, Tm and Yb

1
College of Inter-Faculty Individual Studies in Mathematics and Natural Sciences, University of Warsaw, Banacha 2c, 02-097 Warsaw, Poland
2
Centre of New Technologies, University of Warsaw, Banacha 2c, 02-097 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Submission received: 27 January 2021 / Revised: 26 February 2021 / Accepted: 4 March 2021 / Published: 10 March 2021
(This article belongs to the Special Issue Hydrogen Storage in Metal Hydrides and Related Materials)

Abstract

:
In total, three novel organic derivatives of lanthanide borohydrides, n-But4NRE(BH4)4 (TBAREB), RE = Ho, Tm, Yb, have been prepared utilizing mechanochemical synthesis and purified via solvent extraction. Studies by single crystal and powder X-ray diffraction (SC-XRD and PXRD) revealed that they crystalize in two polymorphic forms, α- and β-TBAREB, adopting monoclinic (P21/c) and orthorhombic (Pnna) unit cells, previously found in TBAYB and TBAScB, respectively. Thermal decomposition of these compounds has been investigated using thermogravimetric analysis and differential scanning calorimetry (TGA/DSC) measurements, along with the analysis of the gaseous products with mass spectrometry (MS) and with analysis of the solid decomposition products with PXRD. TBAHoB and TBAYbB melt around 75 °C, which renders them new ionic liquids with relatively low melting points among borohydrides.

1. Introduction

Metal borohydrides constitute a continuously growing family of chemical compounds. Although the dominant applicability of this group of compounds remains linked with their reducing properties (exploited mostly for NaBH4 and for LiBH4 [1]), the recent expansion of their chemistry is related to other potential applications. In the last two decades, due to extremely high hydrogen content of a BH4- anion (>27 wt.%), such interest was predominantly stimulated by their prospective use as chemical hydrogen storage materials for supply of the fuel cells installed in vehicles [2,3,4,5,6,7,8,9]. Although most of studies focus on examining of the process of hydrogen release, several reports indicate at least some degree of reversibility of such systems, which is crucial for their application, e.g., [10]. Various borohydrides can also serve as precursors of metal borides [11,12,13,14,15] or boron nitride [16], and some of them have been explored as prospective solid-state electrolytes in Li+ batteries [17,18,19,20,21] or luminescent and magnetic materials [22,23], while the borohydride complexes of rare-earth elements were also tested for organic catalysis [24,25].
Several methods have been proposed to modify the physico-chemical properties of borohydride-based materials, among which substitution of the cations and formation of mixed-cation or mixed-anion compounds appeared most common [26,27,28]. Consequently, novel synthetic approaches have been developed to broaden the scope of available borohydride-based systems showing potentially interesting properties [22,29,30,31,32,33,34]. Among the methods available for preparation of the mixed-cation borohydrides in a solvent-free form either those based on mechanochemically-driven addition or metathesis, or on a solvent-mediated metathesis reaction reveal general applicability to diverse systems. While purification of the products of the mechanochemical synthesis may often be problematic, the approach utilizing weakly coordinating anions [35] and the solvents of low basicity [36] leads to mixed-cation borohydrides of rather high purity [31,32,37], e.g.,
[Cat]Y(BH4)4 + Cs[An] → CsY(BH4)4↓ + [Cat][An]
The symbols [Cat] and [An] in Equation (1) denote a bulky organic cation, like [(n‑C4H9)4N]+ or [(C6H5)4P]+, and a weakly coordinating anion, like [Al(OC(CF3)3)4], respectively. The borohydride product (in this case CsY(BH4)4) is precipitated from the solvent like dichloromethane, leaving most contaminants in the remaining solution. The above synthetic method has been utilized for a number of products, containing various metal cations, including those belonging to the s- and d- blocks of periodic table, while among the f-elements only the borohydrides containing EuIII have been attempted [37].
In this contribution we discuss synthesis, crystal structures, and thermal decomposition of a series of borohydrides of trivalent rare-earth (RE) elements: (n‑C4H9)4NRE(BH4)4, for RE = Ho, Tm and Yb. These compounds can be conveniently prepared and are thermally stable, therefore they may serve as precursors towards new mixed-cation borohydrides of the RE mentioned above, which would be performed according to Equation (1). Notably, only a limited number of inorganic, solvent-free borohydrides of RE = Ho, Tm, and Yb is currently known, including KHo(BH4)4 [38], RbTm(BH4)4 [23] and MYb(BH4)4, where M = Li [11], Na, K [39]. It is worth to mention that although a few (n‑C4H9)NRE(BH4)4 compounds have been reported before [40], only two of them, RE = Y [41] and Sc [42], were fully identified on a basis of their crystal structures.

2. Materials and Methods

2.1. Reagents Handling

All reagents were handled in a glovebox under argon atmosphere (H2O < 1 ppm, O2 < 1 ppm). All chemicals were anhydrous, fine quality reagents of high purity (RECl3 > 99.9%; LiBH4 > 95%; (n‑C4H9)4N(BH4) (TBAB) >98%; n-hexane > 99%; dichloromethane (DCM) > 99.8%) purchased from Sigma Aldrich. Anhydrous DCM was additionally distilled with P2O5 and to some portions molecular sieves were added as well.

2.2. Synthesis

Mechanochemical reactions were carried out utilizing LMW–S vibrational mill (Testchem, 1400 rpm) and milling vessel made of silicon carbide. All millings were conducted in 5 min periods, with overall time of 60 min. Between these periods the vessel was cooled down using liquid nitrogen to maintain room temperature (RT), thereby avoiding thermal decomposition of the products. The vessels were sealed inside the glovebox filled with argon. The samples were synthesized using 1 mmol of RECl3, and other reagents, according to the following reaction scheme:
[(n-C4H9)4N]BH4 + RECl3 + 3 LiBH4 → [(n-C4H9)4N]RE(BH4)4 + 3 LiCl

2.3. Powder X-ray Diffraction (PXRD)

PXRD measurements were performed using two diffractometers: Bruker D8 Discover and Empyrean series 2, both with parallel beam and the CuKα1 and CuKα2 radiation (intensity ratio of 2:1). Powder samples were sealed under argon atmosphere inside quartz capillaries (diameter of 0.3–0.5 mm).

2.4. Rietveld Refinement

Jana2006 program [43] was used for Rietveld refinement. [(n-C4H9)4N]Y(BH4)4 [41] and [(n-C4H9)4N]Sc(BH4)4 [42] were utilized as preliminary structural models for α-[(n-C4H9)4N]RE(BH4)4 and β-[(n-C4H9)4N]RE(BH4)4, respectively. Pseudo–Voight function was used for peak shape description, and Berar–Baldinozzi function or correction by divergence to describe their asymmetry. The background was described by 36 Legendre polynomials. A set of restrains was applied to keep suitable geometries. For BH4 groups the B–H distances were fixed at 1.15 Å (with standard uncertainty, s.u. (standard uncertainty) equal to 0.001 Å) and H–B–H angles were fixed at 109.47° (s.u. = 0.01°). For [RE(BH4)4] a tridentate coordination scheme was set, where RE–H distances for the three H atoms from each borohydride group were fixed to be equal (s.u. = 0.01 Å). The Atomic Displacement Parameters (ADP) of hydrogen atoms were fixed as 1.2 ADP of boron atoms. If needed, inter H...H contact between TBA groups was restrain to at least 2 Å. For [n‑But4N] two approaches have been applied. For α-[(n‑C4H9)4N]RE(BH4)4 ADP for all B atoms were kept the same, as well as ADP for all C and N atoms. The C–H distances were restrained to 1 Å (s.u. = 0.001 Å), C–C and C–N distances to 1.5 Å (s.u. = 0.01 Å). Suitable C–C–C, C–N–C, N–C–C, N–C–H, C–C–H, and H–C–H angles were restrained to 109.47° (s.u. = 1°) to keep TBA+ geometry. ADPs of hydrogen atoms were fixed as 1.2 ADP of bonded C-atom. For β-[(n‑C4H9)4N]RE(BH4)4 (higher symmetry) suitable atoms were kept tetrahedral and ADP of hydrogen contained in organic cation as 1.2 or 1.5 ADP of the neighboring carbon for the middle and terminal carbon atom, respectively (riding model).

2.5. Recrystallization of the Samples

Samples were recrystallized using DCM in which TBAREB is very well soluble, while the insoluble LiCl can be removed easily using filtration or decantation. A simple, slow evaporation of the solvent at room temperature led to poor quality, highly mosaic, and twinned crystals. Much better results were obtained when the process of crystallization has been carried out via diffusion of hexane vapor to the DCM solution of TBAREB.

2.6. Single-Crystal X-ray Diffraction (SC-XRD)

The crystals were measured on Agilent Supernova X-ray diffractometer with CuKα micro-source. Data collection and reduction were performed with CrysAlisPro software (version 39.46, by Rigaku Oxford Diffraction) [44], while SHELXT [45] and Olex2 [46,47] programs were applied for structure solution and refinement, respectively.

2.7. The Cambridge Crystallographic Data Centre (CCDC)

α‑[(n‑C4H9)4N]RE(BH4)4, RE = Ho (PXRD: 2058633), Tm (PXRD: 2058634), and β‑[(n‑C4H9)4N]RE(BH4)4, RE = Ho (SC-XRD: 2058684 for 100 K, 2058683 for 200K and 2058678 for 300 K), Tm (PXRD: 2058972; SC-XRD: 2058681 for 100 K, 2058680 for 200 K, and 2058682 for 300 K), Yb (PXRD: 2058635; SC-XRD: 2058679 for 100 K), contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html or from the Cambridge Crystallographic Data Centre. Additionally, preliminary structure for α‑[(n‑C4H9)4N]Ho(BH4)4 from SC-XRD (100 K) is placed in S4 part of ESI, due to poor quality of obtained crystals.
The crystal structures were visualized using VESTA [48].

2.8. Thermal Decomposition: TGA/DSC/EGA

Samples were placed in Al2O3 crucibles with a small hole in the lid, allowing the gaseous thermal decomposition products to be analyzed. Powder samples were analyzed using Netzsch STA 409 PG instrument, utilizing combined thermogravimetry (TGA) and differential scanning calorimetry (DSC) simultaneous measurements. Measurements were performed in the range from room temperature (RT) to 650 °C at 5 °C/min heating rate, at a constant 100 mL/min flow of high-purity argon. The evolved gaseous products were transported through quartz capillary preheated to 200 °C to the quadrupole mass spectrometer (Netzsch / Pfeiffer Vacuum, model QMS 403C). After the measurements the samples were cooled down to RT in argon atmosphere. Then solid products of thermal decomposition were collected, put inside the glovebox and analyzed using PXRD. A short contact with air was not fully prevented during collection of solid decomposition products.

2.9. Spectral Analysis: FTIR

Spectra were collected in 4000–400 cm−1 range using Vertex 80v FTIR spectrometer from Bruker and anhydrous KBr (from Sigma Aldrich, 200 mg per pellet of 12 mm diameter) as pellets material.

3. Results and Discussion

3.1. Synthesis Outcome and Product Identification

A series of [(n‑C4H9)4N]RE(BH4)4 compounds, RE = Ho, Tm, Yb has been prepared according to Equation (2). The products obtained were identified by PXRD, and are summarized in Table 1. The prepared [(n‑C4H9)4N]RE(BH4)4 (abbreviated TBAREB) crystalize in two different structure types observed previously for similar compounds: α- of TBAYB crystal structure (P21/c) [41], and β-, isostructural to TBAScB (Pnna) [42], cf. Figure 1. The polymorphs have been named in analogy to those of the single-cation rare earth borohydrides for which α- denotes a lower symmetry form [11,38,49,50,51,52,53]. For better structural description, the samples were recrystallized and single crystals were investigated by SC-XRD (see: Materials and Methods); the crystal structures have been discussed in the Section 3.2. The plots showing Rietveld analysis of the as-milled samples are placed in Supplementary Materials, Figures S1–S3.
The identity of polymorph dominating in the raw, post-reaction mixture (α- or β-) strongly correlates with RE3+ ionic radius. In the case of the largest Ho3+, revealing similar radius to Y3+, only α-TBAHoB phase (along with LiCl) can be detected, which reminds the behavior of TBAYB [41]. On the other hand, the smallest Yb3+ forms β-TBAYbB isostructural to β‑TBAScB [42]. In this case, the same polymorph is preferred despite the effective size of Yb3+ is ca. 15% larger than that of Sc3+, while there is only <4% difference between Yb3+ and Ho3+ [54]. For Tm3+ of intermediate size, both polymorphs, α‑TBATmB and β‑TBATmB, are present in a comparable amount, with the molar ratio of 1:0.96, respectively, according to Rietveld refinement (Figure S2).
Recrystallization of the as prepared TBAHoB and TBATmB via slow evaporation of DCM led to formation of α‑TBAHoB and β‑TBATmB, respectively, with traces of unidentified crystalline phase according to PXRD (Figures S4 and S5, respectively). However, the unidentified impurities contribute only to minor diffraction intensity, especially for RE = Tm. Unfortunately, it appeared that the crystals obtained this way revealed rather poor quality for RE = Ho and Tm (showing high mosaicity and severe twinning), while the crystals of sufficient quality for SC-diffraction measurements were obtained only for β‑TBAYbB. For RE = Tm and Ho much better single crystals were obtained via diffusion of hexane vapor to the DCM solution of TBAREB, surprisingly resulting only in β‑TBAREB in both cases. Apparently, the choice of solvent and the conditions of crystallization influence the polymorphs obtained, indicating their proximity in formation enthalpy.

3.2. Crystal Structures and Polymorphism of [(n‑C4H9)4N]RE(BH4)4

Crystal structures obtained from the diffraction measurements performed at RT, ca. 293 K (PXRD) and at controlled temperature of 100, 200, and 300 K (SC-XRD) for TBAREB, RE = Y, Ho, Tm, Yb, Sc, are summarized and compared in Table 2 and Table S1 (in the Supplementary Materials).
α‑[(n‑C4H9)4N]RE(BH4)4 adopt monoclinic unit cell crystalizing in P21/c centrosymmetric space group, presented at Figure 1a. This structure was previously described for α‑TBAYB analogue [41]. A comparison of the unit cells for RE = Ho, Y, Tm is presented in Table 2. It is worth noting, that β angle is similar for all those compounds (around 129°). β‑[(n‑C4H9)4N]RE(BH4)4 has a higher symmetry and adopt orthorhombic unit cell, space group Pnna, seen at Figure 1b. It strongly resembles the crystal structure of TBAFeCl4 [55], where the chloride ions are replaced by the borohydride groups. This structure was previously described for β‑TBAScB analogue [42].
Both α‑TBAREB and β‑TBAREB polymorphs of these ionic compounds are composed of [(n‑C4H9)4N]+ cations (abbreviated TBA+) and [RE(BH4)4] anions, with Z’ = 1 and 0.5, respectively, Figure 1. In the first of them each TBA+ is surrounded by four [RE(BH4)4], forming a distorted tetrahedron, while [RE(BH4)4] is surrounded in a deformed trigonal pyramid pattern by the four nitrogen centers of TBA+ cations (Figure S6). In the case of β‑TBAREB polymorph both TBA+ and [RE(BH4)4] have the same coordination spheres, with five counterions around them constituting deformed trigonal bipyramids. The latter form a three-dimensional network with N and RE atoms placed in a honeycomb-like structure, taking alternating positions in the corners of each hexagon (Figure S7).
In the case of β‑TBAREB, RE = Sc [42], and described here by powder method for RE = Tm, Yb a better Rietveld fit for Yb has been acquired modeling slight disorder of the aliphatic chains in the TBA+ cation. However, such disordered model was not necessary for a reliable refinement of SC-XRD data for RE = Ho, Tm, Yb. The aliphatic chains present in both structures of TBAREB are rather close to the idealized geometry. This was not the case for the parent TBAB showing significantly deformed and severely disordered TBA+ moieties already at 100 K [41].
In both cases RE3+ is coordinated by the four borohydride groups with boron centers forming a tetrahedron around RE. Although the positions of hydrogen atoms are significantly biased by their low scattering factor, the FTIR spectra (Figure S20 in ESI) show the features characteristic for BH4 anions acting as tridentate ligands [56], which results in 12-fold coordination. The observed absorption bands are also similar to those reported for the related derivatives of yttrium borohydride [57], where the DFT calculations for [(CH3)4N][Y(BH4)4] confirmed 12-fold coordination of Y3+ in the [Y(BH4)4] complex. [RE(BH4)4] complex anion in such geometry is found in numerous RE3+ borohydrides, also in all the known mixed-cation derivatives of RE(BH4)3, RE = Ho, Tm, Yb, namely K[Ho(BH4)4] [38], Rb[Tm(BH4)4] [23], [Ph4P][Tm(BH4)4] [58], and M[Yb(BH4)4], M = Li [11], Na, K [39], as well as in the related compounds of Sc and Y [34,57,59,60].
Under specified temperature conditions, the molecular volume (Vm) of β-TBAREB is related to RE3+ ionic radius, Table 2 and Table S1. At 100 K the α‑TBAREB polymorph of TBAHoB reveals slightly smaller molecular volume (by ca. 2%, from SC-XRD data), while this trend inverts at RT—the molecular volume of α‑TBATmB is ca. 1% larger than that of β‑TBATmB, as judged from the PXRD data, Figure 2a. On the other hand, the minimal distances between the ion centers, as probed by the RE…N distances, are significantly shorter for α‑TBAREB polymorph (>4%), regardless of the temperature, Figure 2b. This difference is even more distinct for TBAHoB, showing slightly different geometry of the TBA+ cation resulting in smaller ion separation, Figure S8.

3.3. Thermal Decomposition Process and Evolved Gas Analysis

TGA curves for all the samples reveal major drop of mass between ca. 190 and 300 °C, indicating the main stages of thermal decomposition, Figure 3 and Figure S10–S16 in Supplementary Materials. For Ho and Yb samples (Table 1), thermal decomposition is preceded by melting around 75 and 80 °C, respectively, as judging from DSC extrema. In the case of Yb melting is preceded probably by a polymorphic transition revealing minor endothermic effect around 54 °C. This renders TBAHoB and TBAYbB ionic liquids, similar to their analogues described previously: TBAYB melting around 78 °C [41] and TBAScB melting at 88 °C [42]. In the case of Tm sample an exothermic event can be observed close to this temperature which is associated with a minor mass loss (<1 wt.%) due to emission of volatile products (Figures S12 and S15). The signals observed in the time-resolved mass spectra suggest emission of B2H6 during this decomposition stage, which could indicate partial reduction to Tm2+. This sample shows no signs of melting, as judged from lack of clear endothermic DSC signal and the physical appearance of the post-decomposition residue, which remains powder-like, contrary to the other two samples (see Section 3.4).
Around 175 °C, an intense evolution of H2 and B2H6 is observed on MS spectrum of Yb sample, along with an exothermic effect seen at DSC curve, and a minor mass loss (ca. 1.2 wt.%). This should be associated with Yb3+ → Yb2+ reduction which takes place in the case of other borohydrides containing ytterbium [11,50]. In this respect, TBAYbB is slightly more stable thermally in comparison to its inorganic analogues; a similar effect has been observed for the borohydrides of Eu3+ and Sm3+ [11,37,51,61].
According to temperature-resolved MS spectra, in the case of Ho and Yb samples volatile organic decomposition products are evolved within the range of 175–375 °C, along with the fragments of BH4 groups. The signals from hydrogen are present in the range of 250 to 325 °C and 225 to 650 °C for Ho and Yb sample, respectively. On the other hand, during thermal decomposition of Tm sample the MS signals from the volatile organic decomposition products are detected in the range of 135 to 500 °C, with an additional upcast of signals just before 600 °C. In this case, strong signals from hydrogen are observed starting from 175 °C till the end of the measurement.
Above 600 °C the melting of LiCl is observed at DSC curves of all the samples, with much smaller effect seen for Yb. Such weakening of the thermal effect may be caused by partial consumption of LiCl due to formation of a halide-substituted derivative, as it has been observed during heating for inorganic borohydrides of ytterbium and scandium [50,62] and in as-milled samples of various others lanthanides borohydrides [51,63,64,65]. The total mass drop at TGA curves for Ho, Tm and Yb samples is 40.2, 36.6, and 39.5%, respectively. As the LiCl does not contribute to the drop of mass, it can be associated only with thermal decomposition of TBAREB, which leads to the corrected mass drops of 51.2, 46.5, and 50.1%, respectively. Thermal decomposition processes discussed here for TBAREB, RE = Ho, Tm, Yb, are similar to those observed for RE = Y [41], Sc [42]. TBAYB and TBAScB decompose endothermally, above 230 and 175 °C, respectively. Evolution of volatile organic decomposition products in the case of yttrium and scandium analogs leads to ca. 60% (corrected for pure compound) and 64% mass drop, respectively. These complex borohydrides are usually more stable thermally in comparison to organic (n‑C4H9)4NBH4, which melts around 130 °C and its exothermic decomposition starts above 160 °C with a mass drop around 87.5% [41].

3.4. Analysis of Solid Residues from Thermal Decomposition

The process of thermal decomposition of borohydrides may be utilized for synthetic purposes. The recent studies show that lanthanide borohydrides can serve as precursors of crystalline lanthanide borides [11,51], which can be prepared via annealing of the former compounds to 650 °C. What is more, thermal decomposition of (NH4)3Mg(BH4)5 was reported to led to formation of amorphous BN, which can be obtained in the pure form after rinsing the product with water [16]. These findings encouraged us to analyze the chemical composition of the residue left after heating of TBAREB to 650 °C. The crystalline phases detected by PXRD are summarized in Table 1.
In the case of Ho and Yb samples the residue is embedded in the crucible, and also on the top of the lid, which suggests melting and boiling of the samples and blending into the crucible pores during the process in which the residue is formed. Contrary, the post-decomposition residue of the Tm sample remains in the form of a pellet, and it is not spread on the top of the lid and it is not blended into the crucible, resembling the form of residue found after thermal decomposition of simple lanthanide borohydrides [11]. It appeared that the respective oxides and borates are the only RE-containing crystalline phases detected in the decomposed samples instead of the lanthanide borides (which should also be stable during short contact with air). Similar crystalline oxide-containing phases were detected for several samples of decomposed inorganic borohydrides of RE(BH4)3, which indicated formation of the active REHx, instantly reacting with atmospheric O2/H2O [11]. Note that the contact with air was possible during the transportation of the crucibles (containing residue) to the glovebox. In all the samples the respective REBO3 have been observed in crystalline form, along with remaining LiCl. Additionally, in the case of RE = Yb, Yb2O3 was observed, and in the case of RE = Ho additional signals were seen at the PXRD pattern, which can be assigned to Li3B14 and to a phase isostructural to Li2Yb5O4(BO3)3 [66], probably of a formula of Li2Ho5O4(BO3)3 (Figures S17–S19, structures from [66,67,68,69,70]). Unfortunately, the quality of PXRD pattern obtained does not allow for a reliable refinement of the latter structure. In the case of RE = Tm only TmBO3 and LiCl are present as crystalline phases according to the PXRD results.

4. Conclusions

Successful synthesis of three novel compounds, namely n‑But4NRE(BH4)4, RE = Ho, Tm, Yb, is reported, alongside with their crystal structures, polymorphism, and thermal properties. TBAREB are synthesized using mechanochemical milling and are purified (from LiCl byproduct) via extraction with DCM. SC-XRD and PXRD studies show that they crystalize in a monoclinic (P21/c, α-TBAREB) or an orthorhombic (Pnna, β-TBAREB) crystal system. The former was found in the as-milled TBAREB samples, RE = Ho, Tm, and in the TBAHoB samples recrystallized in DCM, while the latter in the as-milled and crystallized in DCM TBAREB samples for RE = Yb, Tm and in the TBAREB samples, RE = Ho, Tm, when a mixture of DCM and hexane has been utilized for crystallization.
Unit cells of both polymorphs have similar volumes, β-TBAREB form being less than 1% smaller at RT according to PXRD data, while the trend is being inverted at 100 K, where α-TBAREB unit cell has smaller molecular volume by around 2%, as judged from SC‑XRD data. Size of the unit cell in the case of all RE increases with rising RE3+ ionic radius for β-TBAREB polymorphs under specified temperature conditions, and also grows with rising temperature for both polymorphs. We found out that the minimal distances between the ion centers, probed by the RE…N distances, are significantly shorter for α-TBAREB polymorph (>4%), while the difference is even more distinct for TBAHoB.
TBAHoB and TBAYbB melt around 75 °C, which renders them new ionic liquids. Such molten salts can be used in synthesis of other compounds in a reductive environment, extending the current utility of the ionic liquids with incorporated rare earth elements [71]. n‑But4NRE(BH4)4, RE = Ho, Tm, Yb, decompose with evolution of organic compounds, diborane, and hydrogen. Additionally, a reduction in Yb3+ → Yb2+ is observed with simultaneous intense evolution of diborane and hydrogen. Solid decomposition products after heating at 650 °C and short contact with air at RT consist of the corresponding RE boranes, oxides and lithium borides.
As the compounds obtained are soluble in the weakly basic solvents, they can be used as precursors of various inorganic mixed-cation lanthanide borohydrides, which synthesis in a pure form would be possible employing metathetic synthesis protocol of mixed-metal borohydrides [14,31,32,37].

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1996-1944/14/6/1329/s1, Figure S1. Rietveld refinement for as-milled sample Ho, Figure S2. Rietveld refinement for as-milled sample Tm, Figure S3. Rietveld refinement for as-milled sample Yb, Figure S4. PXRD patterns with subtracted backgrounds for as-milled Ho sample (containing LiCl), purified Ho (with unknown impurities) and simulated pattern of α TBAHoB, Figure S5. PXRD patterns with subtracted backgrounds for as-milled Tm sample (containing LiCl), purified Tm (with unknown impurities) and simulated patterns of α‑TBATmB and β‑TBATmB, Table S1, supplementary to Table 2. Unit cell dimensions for TBAREB, RE = Y, Ho, Tm, Yb, Sc. 100/200/300 K obtained from SC-XRD, RT from PXRD, Figure S6. Nitrogen centers of TBA+ cations and RE3+ in α‑TBAREB, Figure S7. Honeycomb-like structure of nitrogen centers of TBA+ cations and RE3+ in β‑TBAREB, Figure S8. RE…N distances for α‑TBAHoB and α‑TBATmB, Figure S9. Evolution of a, b, and c lattice parameters of β‑TBAHoB and β‑TBATmB in the function of temperature. SC-XRD data, Figure S10. TGA/DSC curves of the samples (a) Ho, (b) Yb and (c) Tm, up to 650 °C, Figure S11. MS spectrum for Ho sample, Figure S12. MS spectrum for Tm sample, Figure S13. MS spectrum for Yb sample, Figure S14. TGA/DSC curves for Ho sample, Figure S15. TGA/DSC curves for Tm sample, Figure S16. TGA/DSC curves for Yb sample, Figure S17. PXRD pattern (with subtracted background) of Ho sample heated to 650 °C, and simulated patterns of identified crystalline pyrolysis products, Figure S18. PXRD pattern (with subtracted background) of Tm sample heated to 650 °C, and simulated patterns of identified crystalline pyrolysis products, Figure S19. PXRD pattern (with subtracted background) of Yb sample heated to 650 °C, and simulated patterns of identified crystalline pyrolysis products, Figure S20. FTIR spectra of Ho, Yb, and Tm samples, and preliminary CIF of α-TBAHoB from SC-XRD data (100 K) in part S4.

Author Contributions

Conceptualization, W.W.; methodology, W.W. and T.J.; validation, W.W. and T.J.; formal analysis, W.W. and T.J.; investigation, W.W. and T.J.; resources, W.W.; data curation, W.W. and T.J.; writing—original draft preparation, W.W.; writing—review and editing, T.J., W.W; project administration, W.W.; funding acquisition, W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre, Poland, grant number 2017/27/N/ST5/01406.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Crystallographic data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html or from the Cambridge Crystallographic Data Centre. For details see Section 2.7.

Acknowledgments

The authors thank the Biopolymers Laboratory, Faculty of Physics, University of Warsaw, for the access to Agilent Supernova X-ray single-crystal diffractometer, co-financed by the European Union within the ERDF Project POIG.02.01.00-14-122/09.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Nora de Souza, M.V.; Alves Vasconcelos, T.R. Recent methodologies mediated by sodium borohydride in the reduction of different classes of compounds. Appl. Organomet. Chem. 2006, 20, 798–810. [Google Scholar] [CrossRef]
  2. Grochala, W.; Edwards, P.P. Thermal Decomposition of the Non-Interstitial Hydrides for the Storage and Production of Hydrogen. Chem. Rev. 2004, 104, 1283–1316. [Google Scholar] [CrossRef] [PubMed]
  3. Bannenberg, L.J.; Heere, M.; Benzidi, H.; Montero, J.; Dematteis, E.M.; Suwarno, S.; Jaroń, T.; Winny, M.; Orłowski, P.A.; Wegner, W.; et al. Metal (boro-) hydrides for high energy density storage and relevant emerging technologies. Int. J. Hydrogen Energy 2020, 45, 33687–33730. [Google Scholar] [CrossRef]
  4. Rivard, E.; Trudeau, M.; Zaghib, K. Hydrogen Storage for Mobility: A Review. Materials (Basel) 2019, 12, 1973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rude, L.H.; Nielsen, T.K.; Ravnsbæk, D.B.; Bösenberg, U.; Ley, M.B.; Richter, B.; Arnbjerg, L.M.; Dornheim, M.; Filinchuk, Y.; Besenbacher, F.; et al. Tailoring properties of borohydrides for hydrogen storage: A review. Phys. Status Solidi 2011, 208, 1754–1773. [Google Scholar] [CrossRef]
  6. Nakamori, Y.; Orimo, S. Borohydrides as hydrogen storage materials. In Solid-State Hydrogen Storage; Elsevier: Amsterdam, The Netherlands, 2008; pp. 420–449. [Google Scholar]
  7. Li, H.-W.; Yan, Y.; Orimo, S.; Züttel, A.; Jensen, C.M. Recent Progress in Metal Borohydrides for Hydrogen Storage. Energies 2011, 4, 185–214. [Google Scholar] [CrossRef]
  8. Orimo, S.; Nakamori, Y.; Eliseo, J.R.; Züttel, A.; Jensen, C.M. Complex Hydrides for Hydrogen Storage. Chem. Rev. 2007, 107, 4111–4132. [Google Scholar] [CrossRef]
  9. Churchard, A.J.; Banach, E.; Borgschulte, A.; Caputo, R.; Chen, J.C.; Clary, D.; Fijalkowski, K.J.; Geerlings, H.; Genova, R.V.; Grochala, W.; et al. A multifaceted approach to hydrogen storage. Phys. Chem. Chem. Phys. 2011, 13, 16955–16972. [Google Scholar] [CrossRef]
  10. Zhang, X.; Zhang, L.; Zhang, W.; Ren, Z.; Huang, Z.; Hu, J.; Gao, M.; Pan, H.; Liu, Y. Nano-synergy enables highly reversible storage of 9.2 wt% hydrogen at mild conditions with lithium borohydride. Nano Energy 2021, 83, 105839. [Google Scholar] [CrossRef]
  11. Wegner, W.; Jaroń, T.; Grochala, W. Preparation of a series of lanthanide borohydrides and their thermal decomposition to refractory lanthanide borides. J. Alloys Compd. 2018, 744, 57–63. [Google Scholar] [CrossRef]
  12. Jensen, J.A.; Gozum, J.E.; Pollina, D.M.; Girolami, G.S. Titanium, zirconium, and hafnium tetrahydroborates as “tailored” CVD precursors for metal diboride thin films. J. Am. Chem. Soc. 1988, 110, 1643–1644. [Google Scholar] [CrossRef]
  13. 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]
  14. Wegner, W.; Jaroń, T.; Dobrowolski, M.A.; Dobrzycki, Ł.; Cyrański, 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]
  15. Yan, Y.; Remhof, A.; Rentsch, D.; Züttel, A.; Giri, S.; Jena, P. A novel strategy for reversible hydrogen storage in Ca(BH4)2. Chem. Commun. 2015, 51, 11008–11011. [Google Scholar] [CrossRef] [PubMed]
  16. Wegner, W.; Fijalkowski, K.J.; Grochala, W. A low temperature pyrolytic route to amorphous quasi-hexagonal boron nitride from hydrogen rich (NH4)3Mg(BH4)5. Dalton Trans. 2020, 49, 336–342. [Google Scholar] [CrossRef]
  17. Cuan, J.; Zhou, Y.; Zhou, T.; Ling, S.; Rui, K.; Guo, Z.; Liu, H.; Yu, X. Borohydride-Scaffolded Li/Na/Mg Fast Ionic Conductors for Promising Solid-State Electrolytes. Adv. Mater. 2019, 31, 1803533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Gulino, V.; Brighi, M.; Dematteis, E.M.; Murgia, F.; Nervi, C.; Černý, 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]
  19. Payandeh GharibDoust, S.; Brighi, M.; Sadikin, Y.; Ravnsbæk, D.B.; Černý, R.; Skibsted, J.; Jensen, T.R. Synthesis, Structure, and Li-Ion Conductivity of LiLa(BH4)3 X, X = Cl, Br, I. J. Phys. Chem. C 2017, 121, 19010–19021. [Google Scholar] [CrossRef]
  20. Zhang, T.; Wang, Y.; Song, T.; Miyaoka, H.; Shinzato, K.; Miyaoka, H.; Ichikawa, T.; Shi, S.; Zhang, X.; Isobe, S.; et al. Ammonia, a Switch for Controlling High Ionic Conductivity in Lithium Borohydride Ammoniates. Joule 2018, 2, 1522–1533. [Google Scholar] [CrossRef] [Green Version]
  21. Yan, Y.; Grinderslev, J.B.; Lee, Y.S.; Jørgensen, M.; Cho, Y.W.; Černý, R.; Jensen, T.R. Ammonia-assisted fast Li-ion conductivity in a new hemiammine lithium borohydride, LiBH4·1/2NH3. Chem. Commun. 2020, 56, 3971–3974. [Google Scholar] [CrossRef]
  22. Schouwink, P.; Ley, M.B.; Tissot, A.; Hagemann, H.; Jensen, T.R.; Smrčok, Ľ.; Černý, R. Structure and properties of complex hydride perovskite materials. Nat. Commun. 2014, 5, 5706. [Google Scholar] [CrossRef]
  23. Wegner, W.; van Leusen, J.; Majewski, J.; Grochala, W.; Kögerler, P. Borohydride as Magnetic Superexchange Pathway in Late Lanthanide Borohydrides. Eur. J. Inorg. Chem. 2019, 2019, 1776–1783. [Google Scholar] [CrossRef]
  24. Fadlallah, S.; Jothieswaran, J.; Rosal, I.D.; Maron, L.; Bonnet, F.; Visseaux, M. Rationalizing the reactivity of mixed allyl rare-earth borohydride complexes with DFT studies. Catalysts 2020, 10, 820. [Google Scholar] [CrossRef]
  25. Visseaux, M.; Bonnet, F. Borohydride complexes of rare earths, and their applications in various organic transformations. Coord. Chem. Rev. 2011, 255, 374–420. [Google Scholar] [CrossRef]
  26. Nickels, E.A.; Jones, M.O.; David, W.I.F.; Johnson, S.R.; Lowton, R.L.; Sommariva, M.; Edwards, P.P. Tuning the Decomposition Temperature in Complex Hydrides: Synthesis of a Mixed Alkali Metal Borohydride. Angew. Chem. Int. Ed. 2008, 47, 2817–2819. [Google Scholar] [CrossRef]
  27. Ravnsbæk, D.B.; Nickels, E.A.; Černý, R.; Olesen, C.H.; David, W.I.F.; Edwards, P.P.; Filinchuk, Y.; Jensen, T.R. Novel alkali earth borohydride Sr(BH4)2 and borohydride-chloride Sr(BH4)Cl. Inorg. Chem. 2013, 52, 10877–10885. [Google Scholar] [CrossRef] [PubMed]
  28. Jaroń, T.; Wegner, W.; Grochala, W. M[Y(BH4)4] and M2Li[Y(BH4)6−xClx] (M = Rb, Cs): New borohydride derivatives of yttrium and their hydrogen storage properties. Dalton Trans. 2013, 42, 6886. [Google Scholar] [CrossRef]
  29. Ley, M.B.; Paskevicius, M.; Schouwink, P.; Richter, B.; Sheppard, D.A.; Buckley, C.E.; Jensen, T.R. Novel solvates M(BH4)3S(CH3)2 and properties of halide-free M(BH4)3 (M = Y or Gd). Dalton Trans. 2014, 43, 13333–13342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Richter, B.; Grinderslev, J.B.; Møller, K.T.; Paskevicius, M.; Jensen, T.R. From Metal Hydrides to Metal Borohydrides. Inorg. Chem. 2018, 57, 10768–10780. [Google Scholar] [CrossRef] [PubMed]
  31. Jaroń, T.; Orłowski, P.A.; Wegner, W.; Fijałkowski, K.J.; Leszczyński, P.J.; Grochala, W. Hydrogen Storage Materials: Room-Temperature Wet-Chemistry Approach toward Mixed-Metal Borohydrides. Angew. Chem. Int. Ed. 2015, 54, 1236–1239. [Google Scholar] [CrossRef]
  32. Jaroń, T.; Wegner, W.; Fijałkowski, K.J.; Leszczyński, P.J.; Grochala, W. Facile Formation of Thermodynamically Unstable Novel Borohydride Materials by a Wet Chemistry Route. Chem. A Eur. J. 2015, 21, 5689–5692. [Google Scholar] [CrossRef]
  33. Hagemann, H.; Černý, R. Synthetic approaches to inorganic borohydrides. Dalton Trans. 2010, 39, 6006. [Google Scholar] [CrossRef]
  34. Starobrat, A.; Jaroń, T.; Grochala, W. Two new derivatives of scandium borohydride, MSc(BH4)4, M = Rb, Cs, prepared via a one-pot solvent-mediated method. Dalton Trans. 2019, 48, 11829–11837. [Google Scholar] [CrossRef]
  35. Malinowski, P.J.; Jaroń, T.; Domańska, M.; Slattery, J.M.; Schmitt, M.; Krossing, I. Building blocks for the chemistry of perfluorinated alkoxyaluminates [Al{OC(CF3)3}4]: Simplified preparation and characterization of Li+-Cs+, Ag+, NH4+, N2H5+ and N2H7+ salts. Dalton Trans. 2020, 49, 7766–7773. [Google Scholar] [CrossRef] [PubMed]
  36. Malinowski, P.J.; Zhuravlev, V.; Jaroń, T.; Santiso-Quinones, G.; Krossing, I. Extending the chemistry of weakly basic ligands: Solvates of Ag+ and Cu+ stabilized by [Al{OC(CF3)3}4] anion as model examples in the screening for useful weakly-interacting solvents. Dalton Trans. 2021, 50, 2050–2056. [Google Scholar] [CrossRef]
  37. Starobrat, A.; Tyszkiewicz, M.J.; Wegner, W.; Pancerz, D.; Orłowski, P.A.; Leszczyński, P.J.; Fijalkowski, K.J.; Jaroń, T.; Grochala, W. Salts of highly fluorinated weakly coordinating anions as versatile precursors towards hydrogen storage materials. Dalton Trans. 2015, 44, 19469–19477. [Google Scholar] [CrossRef] [PubMed]
  38. Wegner, W.; Jaroń, T.; Grochala, W. Polymorphism and hydrogen discharge from holmium borohydride, Ho(BH4)3, and KHo(BH4)4. Int. J. Hydrogen Energy 2014, 39, 20024–20030. [Google Scholar] [CrossRef]
  39. Wegner, W.; Jaroń, T.; Grochala, W. MYb(BH4)4 (M = K, Na) from laboratory X-ray powder data. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2013, 69, 1289–1291. [Google Scholar] [CrossRef]
  40. Makhaev, V.D.; Borisov, A.P.; Tarasov, B.P. Complexes of Tetrahydroborates of Rare-Earth Elements with Tetrabutylammonium Tetrahydroborate (Bu4N)[Ln(BH4)4 · nDME] (n = 0,1). Russ. J. Inorg. Chem. 2000, 45, 40–45. [Google Scholar]
  41. Jaroń, T.; Wegner, W.; Cyrański, M.K.; Dobrzycki, G.W. Tetrabutylammonium cation in a homoleptic environment of borohydride ligands: [(n-Bu)4N][BH4] and [(n-Bu)4N][Y(BH4)4]. J. Solid State Chem. 2012, 191, 279–282. [Google Scholar] [CrossRef]
  42. Starobrat, A.; Jaroń, T.; Grochala, W. Synthesis and characterization of a series of mixed-cation borohydrides of scandium: [Cat][Sc(BH4)4], [Cat] = [Me4N], [n-Bu4N], and [Ph4P]. Inorg. Chim. Acta 2015, 437, 70–73. [Google Scholar] [CrossRef]
  43. Petricek, V.; Dusek, M.; Palatinus, L. Crystallographic computing system JANA2006: General features. Z. Krist. 2014, 229, 345–352. [Google Scholar]
  44. Agilent. CrysAlis PRO; Agilent Technologies Ltd.: Yarnton, Oxfordshire, UK, 2014. [Google Scholar]
  45. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  46. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  47. Bourhis, L.J.; Dolomanov, O.V.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. The anatomy of a comprehensive constrained, restrained refinement program for the modern computing environment—Olex2 dissected. Acta Crystallogr. Sect. A 2015, 71, 59–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  49. Jaroń, T.; Grochala, W. Y(BH4)3—An old-new ternary hydrogen store aka learning from a multitude of failures. Dalton Trans. 2010, 39, 160–166. [Google Scholar] [CrossRef] [PubMed]
  50. Olsen, J.E.; Frommen, C.; Sørby, M.H.; Hauback, B.C. Crystal structures and properties of solvent-free LiYb(BH4)4−xClx, Yb(BH4)3 and Yb(BH4)2−xClx. RSC Adv. 2013, 3, 10764. [Google Scholar] [CrossRef]
  51. Olsen, J.E.; Frommen, C.; Jensen, T.R.; Riktor, M.D.; Sørby, M.H.; Hauback, B.C. Structure and thermal properties of composites with RE-borohydrides (RE = La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, Er, Yb or Lu) and LiBH4. RSC Adv. 2014, 4, 1570–1582. [Google Scholar] [CrossRef]
  52. Frommen, C.; Aliouane, N.; Deledda, S.; Fonneløp, J.E.; Grove, H.; Lieutenant, K.; Llamas-Jansa, I.; Sartori, S.; Sørby, M.H.; Hauback, B.C. Crystal structure, polymorphism, and thermal properties of yttrium borohydride Y(BH4)3. J. Alloys Compd. 2010, 496, 710–716. [Google Scholar] [CrossRef]
  53. Sato, T.; Miwa, K.; Nakamori, Y.; Ohoyama, K.; Li, H.-W.; Noritake, T.; Aoki, M.; Towata, S.; Orimo, S. Experimental and computational studies on solvent-free rare-earth metal borohydrides R(BH4)3 (R = Y, Dy, and Gd). Phys. Rev. B 2008, 77, 104114. [Google Scholar] [CrossRef]
  54. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  55. Hay, M.T.; Geib, S.J. Tetrabutylammonium tetrachloroferrate(III). Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, m190–m191. [Google Scholar] [CrossRef]
  56. Marks, T.J.; Kolb, J.R. Covalent Transition Metal, Lanthanide, and Actinide Tetrahydroborate Complexes. Chem. Rev. 1977, 77, 263–293. [Google Scholar] [CrossRef]
  57. Jaroń, T.; Grochala, W. Probing Lewis acidity of Y(BH4)3 via its reactions with MBH4 (M = Li, Na, K, NMe4). Dalton Trans. 2011, 40, 12808–12817. [Google Scholar] [CrossRef]
  58. Antsyshkina, A.S.; Sadikov, G.G.; Borisov, P.; Makhaev, V.D. Complexes of Yttrium, Thulium, and Lutetium Tetrahydridoborates with Tetraphenylphosphonium Tetrahydridoborate (Ph4P)[M(BH4)4] (M = Y, Tm, Lu): Crystal Structure of (Ph4P)[Tm(BH4)4]. Russ. J. Inorg. Chem. 2001, 46, 1141–1146. [Google Scholar]
  59. Hagemann, H.; Longhini, M.; Kaminski, J.W.; Wesolowski, T.A.; Černý, R.; Penin, N.; Sørby, M.H.; Hauback, B.C.; Severa, G.; Jensen, C.M. LiSc(BH4)4: A novel salt of Li and discrete Sc(BH4)4- Complex anions. J. Phys. Chem. A 2008, 112, 7551–7555. [Google Scholar] [CrossRef] [Green Version]
  60. Černý, R.; Schouwink, P. The crystal chemistry of inorganic metal boro-hydrides and their relation to metal oxides. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2015, 71, 619–640. [Google Scholar] [CrossRef] [PubMed]
  61. Humphries, T.D.; Ley, M.B.; Frommen, C.; Munroe, K.T.; Jensen, T.R.; Hauback, B.C. Crystal structure and in situ decomposition of Eu(BH4)2 and Sm(BH4)2. J. Mater. Chem. A 2015, 3, 691–698. [Google Scholar] [CrossRef] [Green Version]
  62. Starobrat, A.; Derzsi, M.; Jaroń, T.; Malinowski, P.; Grochala, W. The Ground State and Polymorphism of LiSc(BH4)4 Finally Understood by Density Functional Theory Modelling. arXiv 2020. preprint. Available online: https://arxiv.org/abs/2009.00046 (accessed on 4 March 2021).
  63. Ley, M.B.; Boulineau, S.; Janot, R.; Filinchuk, Y.; Jensen, T.R. New li ion conductors and solid state hydrogen storage materials: LiM(BH4)3Cl, M = La, Gd. J. Phys. Chem. C 2012, 116, 21267–21276. [Google Scholar] [CrossRef]
  64. Frommen, C.; Sørby, M.H.; Ravindran, P.; Vajeeston, P.; Fjellvag, H.; Hauback, B.C. Synthesis, crystal structure, and thermal properties of the first mixed-metal and anion-substituted rare earth borohydride LiCe(BH4)3Cl. J. Phys. Chem. C 2011, 115, 23591–23602. [Google Scholar] [CrossRef] [Green Version]
  65. Ley, M.B.; Ravnsbæk, D.B.; Filinchuk, Y.; Lee, Y.-S.; Janot, R.; Cho, Y.W.; Skibsted, J.; Jensen, T.R. LiCe(BH4)3Cl, a New Lithium-Ion Conductor and Hydrogen Storage Material with Isolated Tetranuclear Anionic Clusters. Chem. Mater. 2012, 24, 1654–1663. [Google Scholar] [CrossRef]
  66. Jubera, V.; Gravereau, P.; Chaminade, J.P.; Fouassier, C. A new oxyborate in the ternary phase diagrams Li2O-Ln2O3-B2O3:Li2Ln5O4(BO3)3 (Ln=Yb,Lu): Crystal structure of the ytterbium phase. J. Solid State Chem. 2001, 156, 161–167. [Google Scholar] [CrossRef]
  67. Ievin’sh, A.F.; Straumanis, M.; Karlsons, K. Praezisionsbestimmung von Gitterkonstanten hygroskopischer Verbindungen (Li Cl, Na Br). Z. Phys. Chem. Abt. B Chem. Elem. Aufbau Mater. 1938, 40, 146–150. [Google Scholar]
  68. Mair, G.; Nesper, R.; von Schnering, H.G. Trilithium tetradecaboride Li3B14: Synthesis, structure, and properties. J. Solid State Chem. 1988, 75, 30–40. [Google Scholar] [CrossRef]
  69. Newnham, R.E.; Redman, M.J.; Santoro, R.P. Crystal Structure of Yttrium and Other Rare-Earth Borates. J. Am. Ceram. Soc. 1963, 46, 253–256. [Google Scholar] [CrossRef]
  70. Saiki, A.; Ishizawa, N.; Mizutani, N.; Kato, M. Structural Change of C-Rare Earth Sesquioxides Yb2O3 and Er2O3 as a Function of Temperature. J. Ceram. Assoc. Jpn. 1985, 93, 649–654. [Google Scholar] [CrossRef] [Green Version]
  71. Binnemans, K. Lanthanides and Actinides in Ionic Liquids. Chem. Rev. 2007, 107, 2592–2614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Crystal unit cells of: (a) α‑[(n‑C4H9)4N]RE(BH4)4 (here RE = Ho); (b) β‑[(n‑C4H9)4N]RE(BH4)4 (here RE = Ho). Blue color—RE atoms; green—B atoms; brown—C atoms; grey—N atoms; pink—H atoms.
Figure 1. Crystal unit cells of: (a) α‑[(n‑C4H9)4N]RE(BH4)4 (here RE = Ho); (b) β‑[(n‑C4H9)4N]RE(BH4)4 (here RE = Ho). Blue color—RE atoms; green—B atoms; brown—C atoms; grey—N atoms; pink—H atoms.
Materials 14 01329 g001
Figure 2. (a) Molecular volumes (Vm), and (b) minimal RE…N distances in the function of temperature for a series of TBAREB compounds in both α– and β– forms. SC-XRD data—filled symbols, PXRD data—open symbols.
Figure 2. (a) Molecular volumes (Vm), and (b) minimal RE…N distances in the function of temperature for a series of TBAREB compounds in both α– and β– forms. SC-XRD data—filled symbols, PXRD data—open symbols.
Materials 14 01329 g002
Figure 3. TGA/DSC curves of the crude, not purified samples (a) Ho, (b) Yb, and (c) Tm, prepared as in Table 1. *—artifact, coming from the instrument.
Figure 3. TGA/DSC curves of the crude, not purified samples (a) Ho, (b) Yb, and (c) Tm, prepared as in Table 1. *—artifact, coming from the instrument.
Materials 14 01329 g003
Table 1. Summary of the reactions of type (2) performed in this work, their conditions and crystalline products after the synthesis and after heating them to 650 °C (and short contact with air1 in the last case).
Table 1. Summary of the reactions of type (2) performed in this work, their conditions and crystalline products after the synthesis and after heating them to 650 °C (and short contact with air1 in the last case).
Sample NameSubstratesMilling TimeCrystalline Products (Synthesis)Crystalline Products (Heating to 650 °C)
HoHoCl3 + 3 LiBH4 + TBAB12 × 5 minα‑TBAHo(BH4)4, LiClHoBO3, LiCl, Li3B14, Li2Ho5O4(BO3)3
TmTmCl3 + 3 LiBH4 + TBAB12 × 5 minα‑TBATm(BH4)4, β‑TBATm(BH4)4, LiClTmBO3, LiCl
YbYbCl3 + 3 LiBH4 + TBAB12 × 5 minβ‑TBAYb(BH4)4, LiClYbBO3, LiCl, Yb2O3
1 lid of a crucible has a little hole in the middle, allowing brief contact with air during transportation from TGA/DSC instrument to glovebox.
Table 2. Unit cell dimensions for TBAREB, RE = Y, Ho, Tm, Yb, Sc. Data for 300 K were obtained from SC-XRD, data for RT (ca. 293 K) were obtained from PXRD.
Table 2. Unit cell dimensions for TBAREB, RE = Y, Ho, Tm, Yb, Sc. Data for 300 K were obtained from SC-XRD, data for RT (ca. 293 K) were obtained from PXRD.
Compoundα-TBAHoBα-TBAYB [41]α-TBATmBβ-TBAHoBβ-TBATmBβ-TBAYbBβ-TBAScB [42]
RE3+ r 1 [Å]0.9010.9000.8800.9010.8800.8680.745
spc. groupP21/cP21/cP21/cPnnaPnnaPnnaPnna
T [K]RTRTRT300RTRTRT
a [Å]11.4218(9)11.4181(10)11.4063(18)19.3238(12)19.287(3)19.2235(10)19.1399(10)
b [Å]20.553(2)20.510(3)20.545(4)12.0529(5)12.0317(17)11.9943(6)11.8849(6)
c [Å]15.3049(17)15.2811(19)15.319(3)11.8820(6)11.8591(19)11.8244(6)11.7325(6)
β [°]129.433(7)129.464(8)129.423(12)90909090
V [Å3]2775.0(5)2762.772773.1(10)2767.4(3)2751.9(7)2726.4(2)2668.9(2)
Z4444444
Vm3]693.8690.7693.3691.9688.0682.3667.2
1 Effective ionic radius (6-coordinate, octahedral environment) from [54].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wegner, W.; Jaroń, T. Synthesis, Polymorphism and Thermal Decomposition Process of (n-C4H9)4NRE(BH4)4 for RE = Ho, Tm and Yb. Materials 2021, 14, 1329. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14061329

AMA Style

Wegner W, Jaroń T. Synthesis, Polymorphism and Thermal Decomposition Process of (n-C4H9)4NRE(BH4)4 for RE = Ho, Tm and Yb. Materials. 2021; 14(6):1329. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14061329

Chicago/Turabian Style

Wegner, Wojciech, and Tomasz Jaroń. 2021. "Synthesis, Polymorphism and Thermal Decomposition Process of (n-C4H9)4NRE(BH4)4 for RE = Ho, Tm and Yb" Materials 14, no. 6: 1329. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14061329

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