Next Article in Journal
The Effect of Bulk Nucleation Parameters on the Solidification Structure of Large Slabs During Electroslag Remelting and Optimization of Production Process Parameters
Previous Article in Journal
Multi-State Photoluminescence of Donor–π–Acceptor Tetrafluorinated Tolane Mesogenic Dimers in Solution, Crystal, and Liquid-Crystalline Phases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Comparative Study of Four Hexachloroplatinate, Tetrachloroaurate, Tetrachlorocuprate, and Tetrabromocuprate Benzyltrimethylammonium Salts: Synthesis, Single-Crystal X-Ray Structures, Non-Classical Synthon Preference, Hirshfeld Surface Analysis, and Quantum Chemical Study

1
Chemistry Department, Institute of Ecological and Inorganic Chemistry, Technical University of Lodz, 116 Zeromskiego, 90-924 Lodz, Poland
2
Department of Physical Chemistry, Slovak Technical University, Radlinskeho 9, 81237 Bratislava, Slovakia
3
Faculty of Chemistry, Warsaw University of Technology, 3 Noakowskiego, 00-664 Warsaw, Poland
4
Biological and Chemical Research Centre, University of Warsaw, Żwirki i Wigury 101, 02-089 Warsaw, Poland
5
Institute of Biochemistry and Biophysics, Polish Academy of Sciences, Pawinskiego 5a, 02-106 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(12), 1051; https://doi.org/10.3390/cryst15121051
Submission received: 29 October 2025 / Revised: 3 December 2025 / Accepted: 8 December 2025 / Published: 11 December 2025

Abstract

Four benzyltrimethylammonium (BTMA) salts were successfully prepared: bis(benzyltrimethylammonium) hexachloroplatinate (1), benzyltrimethylammonium tetrachloroaurate (2), bis(benzyltrimethylammonium) tetrachlorocuprate (3), and bis(benzyltrimethylammonium) tetrabromocuprate (4) from benzyltrimethylammonium hydroxide (Triton B). Their crystal structures were determined by single-crystal X-ray diffraction, and the supramolecular architectures were characterized hierarchically. Extended Hirshfeld surface analysis, including enrichment ratio calculations, was performed to evaluate intermolecular interactions. Nonclassical hydrogen bonds, such as C–HCl(Br), involving the anions, contribute to the formation of self-assembled architectures. Additional stabilization arises from ππ and Cu–Brπ interactions, particularly in crystals 2 and 4, respectively. Hirshfeld surface analysis showed that HH and CH/HC interactions are the dominant contributors in all crystals. According to enrichment ratio calculations, CH/HC interactions in 1, 3, and 4; ClH/HCl in 1 and 3; CuH/HCu in 3 and 4; and BrH/HBr and BrC/CBr in 4 are statistically favored in the crystal packing. Halogen bonding ClCl was observed in 1 but does not significantly influence packing. Energy framework calculations indicated that dispersive interactions are favorable in the analyzed crystals. A library of H-bonding supramolecular patterns, including interchangeable synthons, is provided and may guide the rational design of new derivatives with controllable features. Finally, the topology of intermolecular connections and the electronic structure of the benzyltrimethylammonium cation, investigated by quantum-chemical calculations, provide insights into its reactivity.

1. Introduction

Hybrid organic–inorganic halometallates, possessing organic, quaternary ammonium cations and inorganic metal-containing anions, could find significant importance in material sciences, chemistry, and crystal engineering. These compounds, such as 1-ethyl-3-methylimidazolium tetrachloroaurate, could be used as low-temperature ionic liquids, exhibiting ionic liquid behavior from 58 to 220 °C with possible applications in catalysis and electrochemical technologies [1]. The formation of tetrachloroaurate salts from gold(III) chloride and tetraalkylammonium chlorides facilitates the extraction of gold from water solutions to the organic phase [2]. Appropriate salts obtained from hexadecyltrimethylammonium hydroxide and hexachloroplatinic or tetrachloroauric acid were used for the synthesis of nanoparticles and nanofibers on HOPG (highly oriented pyrolytic graphite) [3]. Three hexachloroplatinate complexes [(C2H5)2NH2]2[PtCl6], [(C2H5)4N]2[PtCl6], and [(CH3)3NH]2[PtCl6] were synthetized and tested for their antibacterial activities against Escherichia coli strain M-17. A reliable antimicrobial effect was observed for the latter compound [4]. Organic tetrachlorocuprate salts exhibited termochromic properties [5,6] and possess promising antibacterial activities [7,8]. Another example of such compounds is trimethylchloromethylammonium tetrachloroferrate, a plastic crystal capable of magnetic-optic-electric triple switching and thermal energy storage [9]. Finally, a magnetic ionic liquid—1-butyl-3-methylimidazolium tetrachloroferrate—was obtained and its magnetic behavior and susceptibility was investigated [10].
Crystal engineering focuses on the employment of intermolecular interactions, including hydrogen/halogen bonding, π-based interactions, and electrostatic, hydrophobic, and van der Waals interactions, to self-assemble crystalline materials with controllable properties [11,12]. Hydrogen bonds are, without a doubt, the most popular interactions driving self-assembled structures (supramolecules). Halogen bonding and other non-classical interactions, which play a significant role in supramolecular structures, have occasionally been reported in the scientific literature.
In a continuation of our supramolecular explorations of organic salts [13,14,15], in the present manuscript, we report the synthesis and thorough supramolecular investigation of four salts, namely, bis(benzyltrimethylammonium) hexachloroplatinate (1), benzyltrimethylammonium tetrachloroaurate (2), bis(benzyltrimethylammonium) tetrachlorocuprate (3), and bis(benzyltrimethylammonium) tetrabromocuprate (4), as shown in Scheme 1. It is interesting to note that the supramolecular architectures of these structures are formed only by non-classical interactions. The nature and topology of these interactions are studied by an extended Hirshfeld surface analysis and energy frameworks. The supramolecular preferences of synthetized salts were investigated, and a library of H-bonding supramolecular synthons was developed. Furthermore, quantum chemical studies of compounds containing benzyltrimethylammonium cations, BTMA+, are of two types. Luminescence [16] and magnetic properties [17] were investigated using periodic DFT calculations of the entire crystal (including anions). On the other hand, the vibrational spectra of a crystal were assigned by DFT calculations of the sole BTMA+ [18]. The same holds for DFT studies of BTMA+ degradation by hydroxides’ [19,20,21] adsorption on electrodes [22,23] and catalysis [24]. None of the above studies deals with a detailed study of the electronic structure of BTMA+, which should explain some of its physical and chemical properties. Thus, we will address this omission.
The crystal structures of bis(benzyltrimethylammonium) tetrachlorocuprate (3) and bis(benzyltrimethylammonium) tetrabromocuprate (4) have been reported previously [25,26]; however, they were determined at room temperature, and those studies do not provide a comprehensive analysis of their supramolecular features, including the intermolecular interactions that govern the crystal packing. Moreover, no quantum–mechanical calculations were carried out (in the case of 3 this is understandable, because its crystal structure was described in the 1960s). In view of these limitations, the aim of the present work was to deliver a more complete structural and theoretical description of these systems, with particular emphasis on their supramolecular organization and the nature of the interactions stabilizing the crystal structures. Additionally, a polymorphic form (P21/c) of 3 was also described in the literature [27].

2. Materials and Methods

2.1. Preparation of Salts 14

Bis(benzyltrimethylammonium) hexachloroplatinate (1): To 1.37 g of a 30% solution of hexachloroplatinic acid H2PtCl6 in water (410 mg of H2PtCl6, 1 mmol, 1 equiv.), diluted with 2 mL of distilled water, 824 mg of a 40% solution of benzyltrimethylammonium hydroxide (Triton B, 2 mmol, 2 equiv.) in methanol, diluted with 2 mL of distilled water, was added dropwise and vigorously stirred at room temperature for 5 min. The slow evaporation of the obtained solution in room temperature led to the formation of 1 crystals.
Benzyltrimethylammonium tetrachloroaurate (2): To 1.13 g of a 30% solution of tetrachloroauric acid HAuCl4 in water (340 mg of HAuCl4, 1 mmol, 1 equiv.), diluted with 2 mL of distilled water, 418 mg of a 40% solution of benzyltrimethylammonium hydroxide (1 mmol, 1 equiv.) in methanol, diluted with 2 mL of distilled water, was added dropwise and vigorously stirred in room temperature for 5 min. The slow evaporation of the obtained solution at room temperature led to the formation of 2 crystals.
Bis(benzyltrimethylammonium) tetrachlorocuprate (3): To 135 mg of copper(II) chloride CuCl2 in 2 mL of distilled water, 200 μL of a 37% solution of hydrochloric acid in water (73 mg of HCl, 2 mmol, 2 equiv.) was added, followed by 824 mg of a 40% solution of benzyltrimethylammonium hydroxide (2 mmol, 2 equiv.) diluted with 2 mL of distilled water, and the obtained solution was vigorously stirred at room temperature for 5 min. The slow evaporation at room temperature led to the formation of 3 crystals.
Bis(benzyltrimethylammonium) tetrabromocuprate (4): To 223 mg of copper(II) bromide CuBr2 in 2 mL of distilled water, 404 μL of a 40% solution of hydrobromic acid in water (162 mg of HBr, 2 mmol, 2 equiv.) was added, followed by 824 mg of a 40% solution of benzyltrimethylammonium hydroxide (2 mmol, 2 equiv.) diluted with 2 mL of distilled water, and the obtained solution was vigorously stirred at room temperature for 5 min. The slow evaporation at room temperature led to the formation of 4 crystals.

2.2. Crystallography

Good quality single crystals of the investigated compounds were selected for the X-ray diffraction experiments at T = 100(2) K. Diffraction data were collected on the Agilent Technologies (Santa Clara, CA, USA) SuperNova Double Source diffractometer with CuKα (λ = 1.54184 Å) radiation (1, 3, and 4) and the Agilent Technologies (Santa Clara, CA, USA) SuperNova Single Source diffractometer with MoKα (λ = 0.71073 Å) (2), using the CrysAlis Pro software version 1.171.40.67a [28]. In all cases, the analytical numeric absorption correction using a multifaceted crystal model based on expressions derived by R.C. Clark and J.S. Reid was used [29]. The structural determination procedure was carried out using the SHELX package [30,31]. The structure was solved with the intrinsic phasing method, and a successive least-square refinement was then carried out based on the full-matrix least-squares method on F2 using the SHELXL-2018/3 program [30,31]. All H-atoms were positioned geometrically, with the C–H equal to 0.93, 0.96, and 0.97 Å for the aromatic, methylene, and methyl H-atoms, respectively, and refined as fixed.

2.3. Computational Study

2.3.1. Quantum Chemical Studies

Starting from the X-ray structure of (BTMA)2[PtCl6], the geometry of the BTMA+ cation in the singlet ground spin state was optimized in vacuo and in aqueous solutions at the MP2 [32,33] level of theory using the cc-pVTZ basis set from the Gaussian library [34]. Solvent effects were accounted for using the SMD (solvent model based on density) [35] approximation. Stable structures were tested for the absence of imaginary vibrations using vibrational analysis. All quantum chemical calculations were performed using the Gaussian09 software, version 9.0 (Gaussian Inc.: Wallingford, CT, USA, 2011) [34]. Frontier molecular orbitals were drawn using the MOLEKEL software, version 5.4.0.8 (Swiss National Supercomputing Centre: Manno, Switzerland, 2009) [36] from the Gaussian09 output. The map of the electrostatic potential of BTMA+ at the Hartree–Fock level of theory was drawn using the Avogadro software, version 1.98 [37].
The electronic structure in terms of the QTAIM (quantum theory of atoms-in-molecule) topological analysis of electron density is described as follows [38]. Atomic charges and volumes were obtained by integration of electron density and space over atomic basins (up to 0.001 a.u. isosurface), respectively. Bond critical points (BCPs) are saddle points of electron density on the bond paths between individual atom pairs. A set of bond paths is denoted as a molecular graph. The BCP electron density, ρBCP, reflects the bond strength. Its Laplacian, ∇2ρBCP,
2ρBCP = λ1 + λ2 + λ3
where λ1 < λ2 < 0 < λ3 are the eigenvalues of the Hessian of the BCP electron density, which depicts the relative electron density contribution of the bonded atoms (negative values correspond to covalent bonds). The BCP bond ellipticity, εBCP,
εBCP = λ12 − 1
describes the bond deviation from cylindrical symmetry (such as in ideal single or triple bonds) due to its double-bond character, mechanical strain, or other perturbations.
The electron delocalization index, DI, is the average number of electrons delocalized (shared) between atom pairs. It corresponds to a bond index if both atoms are connected by a bond path [39]. The QTAIM analysis was performed using the AIMAll software [39] from *.wfn files produced by the Gaussian09 software, version 9.0 (Gaussian Inc.: Wallingford, CT, USA, 2011). Molecular graphs were drawn using the AIM2000 software, version 1.0 [40]. The superposition of the optimized structures was optimized using the PyMol software, ver. 1.8 [41,42]. The root mean square deviations (RMSDs) of the corresponding atomic positions of individual pairs of structures were evaluated using the script (www.github.com).

2.3.2. Hirshfeld Surface Analysis

The 3D Hirshfeld surfaces (HSs) and corresponding 2D fingerprint calculations were carried out in CrystalExplorer [43] based on the methods reported in the literature [44]. The maps of Hirshfeld’s surfaces were prepared using dnorm, shape index, curvedness, and fragment patch propensity. Fingerprint plots show a qualitative description of close contacts in the crystals as a function of the di and de values [44].

2.3.3. Enrichment Ratio

The enrichment ratios (ERs) of the intermolecular interactions observed in the crystal structures 14 were calculated based on the HS methodology [45] as the ratio between actual and random inter-contacts. In this context, privileged (with ER larger than unity) and disfavored (with ER smaller than unity) inter-contacts were described in the analyzed crystals.

2.3.4. Energy Frameworks

The pairwise interaction energies between the moieties in the crystals were calculated using the CrystalExplorer 21.5 program [46] using the wave function calculated at the B3LYP/6-31G(d,p) functional basis set that is widely accepted in computational chemistry. More specifically, the electrostatic (Eele), polarization (Epol), dispersion (Edisp), and exchange repulsion (Erep) energy components, according to the Equation presented below, were calculated [47].
Etot = Eele + Epol + Edisp + Erep
A molecular cluster with a radius of 3.8 Å was generated around a selected single moiety of the reference compound.
The total energy was calculated using scaling factors such as Kele = 1.057, Kpol = 0.740, Kdis = 0.871, and Krep = 0.618 [47,48].

2.3.5. CSD-Materials Module

The Cambridge Crystallographic Data Centre Mercury 4.0 program (version 2024.3.1) [49], along with selected functionalities from the CSD-Materials module, was used in this study. Before any calculations, the molecular structures were standardized according to the Cambridge Structural Database conventions for bond types, and default settings were applied for all subsequent calculations. To assess the preferred interaction behavior, full interaction maps (FIMs) were generated for the studied crystals based on CSD interaction data [50]. This approach allowed for the visualization of interaction landscapes using 3D coordinates obtained from X-ray crystallographic data. Calculations were performed using the Aromatic Analyser, a neural network-based tool that quantitatively evaluates aromatic ring interactions within crystal structures. Only interactions classified as strong (scores of 10–6) or moderate (scores of 4–6) on a 1–10 scale were considered. The topology of intermolecular connections in the analyzed series was assessed by representing each cation as a single point located at its center of gravity. Connections were established based on identified weak interactions. Anions were represented by the position of their heavy atom, and the charge-assisted C–H···Cl hydrogen bonds were depicted as linkers connecting the cations.

3. Results and Discussion

3.1. Crystallographic Study

Crystals of 1–4 were successfully obtained, and their 3D structures were determined by the single-crystal X-ray diffraction method with high precision at low temperatures. The compounds 13 crystallize in the monoclinic centrosymmetric space groups (1 in P21/c, 2 in C2/c, and 3 in P21/n), whereas the compound 4 crystallizes in the orthorhombic noncentrosymmetric space group P212121. The asymmetric unit of the crystal lattice of 1 contains one benzyltrimethylammonium cation and half of the inorganic anion, while in 2, one cation and two individual halves of an anion are present. The metal atoms of all the above anions lie on the symmetry elements. On the other hand, in the analyzed crystals of 3 and 4, we are dealing with a situation where two benzyltrimethylammonium cations assist one inorganic anion, and all atoms of the latter have a general position.
As mentioned, all four analyzed salts contain benzyltrimethylammonium cations; however, differences in their intrinsic geometries are rather small. As can be seen in the figure showing the overlay of their structural skeletons, differences manifest as a slight rotation of the trimethylammonium moiety (Figure 1). The substantial distinctions among the individual salts manifest in their overall supramolecular architecture. The geometry around the Pt atom in the hexachloroplatinate anion (in 1) is octahedral, while around the Cu atom in the tetrachlorocuprate (in 3) and tetrabromocuprate (in 4) anions it is slightly distorted and tetrahedral. Interestingly, in 1, the hexachloroplatinate anion lies on the inversion center and two-fold axis.
The molecular structures of the analyzed salts are depicted in Figure 1. Complete crystal data and refinement details of the analyzed salts are collected in Table 1. The datasets have been deposited in the Cambridge Crystallographic Data Centre under CCDC numbers 2464968–2464971. Selected bond lengths and angles of 14 are summarized in Tables S1–S12 in the Supplementary Materials Section. The values are similar to those found in the literature.

3.2. Supramolecular Features of 14

The full interaction maps (FIMs) for 14 that were calculated based on statistical information on interactions derived from the CSD [50] are presented in Figure 2. They helped to check the preferred interaction behavior and understand the effect of different anions on the formation of supramolecular synthons. Overall, the dark zones indicate a high tendency for synthon formation. The predicted most-likely positions of functional groups with a crystal packing indicated that corresponding crystals fulfill the expected H-bonding and π-based interactions. First of all, a thorough comparative analysis indicated a lack of classical interactions in all the analyzed structures since blue and red zones for the H-bonding donor and acceptor probabilities are absent. Moreover, nearly identical beige–brown landscapes of aromatic–hydrophobic interactions were obtained. Notably, subtle but essential differences are visible. The color intensity of the regions correlates with the likelihood of the corresponding interactions. The functional groups have slightly different probabilities of being involved in aromatic–hydrophobic (beige–brown areas) and non-classical H-bonding (claret areas) interactions. The generated synthons by the latter interactions are discussed in the section below. Moving forward, the specificity of the studied salts resulted from the restricted functional groups and low availability of H-atom acceptors; hence, the crystal packing in all supramolecular systems is driven only by non-classical intermolecular interactions, such as C–HCl and C–HBr (in 4). The donor-to-acceptor distances vary from 3.563 Å for C–HCl in 1 to 3.838 Å for C–HBr in 4. Thus, the supramolecular networks of the investigated compounds are governed via the presence of the anions containing the chlorine and bromine atoms as H-acceptors of hydrogen bonds. On the other hand, significant ππ interconnects are observed in 2 and, additionally, Cu–Brπ interconnects are observed in 4. The geometric parameters associated with H-bonding interactions in compounds 14 are summarized in Table 2, while π-based interconnects are listed in Table S13. A detailed analysis using the Aromatic Analyser module is presented below.
The results of the calculations performed using the Aromatic Analyser module, integrated into the CSD Mercury 2024.3 software (CCDC, Cambridge, UK), are presented in Table S14. For compound 1, only moderate C–Hπ interactions (score 5.8) were observed. In contrast, compound 2 exhibited strong ππ interactions (score 8.9), characterized by a parallel orientation of the rings with a centroid distance of 3.932(2) Å. Additionally, moderate off-set stacking interactions (score 5.3) were detected. In structure 3, both cations are only weakly connected. The highest score of 4.3 corresponds to C–Hπ interactions between crystallographically different cations. Similar interactions were found in compound 4, with a maximum score of 4.5.
The crystal packing of the studied structures is illustrated in Figure 3, Figure 4, Figure 5 and Figure 6, showing, at first glance, different supramolecular architectures. The crystals of the salts differ by networks of non-covalent (supramolecular) interactions and stacking modes. Nevertheless, a thorough examination revealed a high similarity of the topology of packing in 3 and 4 crystals and a lower similarity between 1 and 2 crystals. In 1 and 2, benzyltrimethylammonium cations are arranged in an antiparallel fashion bonded by the hexachloroplatinate and tetrachloroaurate anions, respectively, giving rise to a layer-like structure with the alternation of two sub-layers of apolar cations and polar anions prolongated along the crystallographic axis b. Nevertheless, in 2, the different positions of anions adopt different twisted (about 90°) arrangements relative to each other. Notably, in 1, the octahedral anion participates in bi- and tetrafurcated hydrogen bonds. The crystal packing of 3 and 4 is very similar. Benzyltrimethylammonium cations are arranged in a head-to-tail fashion and connected via tetrachlorocuprate and tetrabromocuprate anions, respectively. Consequently, twisted cationic sub-layers are observed. Both in 3 and 4, distorted tetrahedral anions participate in multi-furcated hydrogen bonds.
In all the analyzed crystals, the ions are linked together by C–HCl (in 13) and C–HBr (in 4), respectively, enclosing diverse dimers (D) at the first level of the graph set theory, and chains (C) and rings (R) at the second level of the graph set theory, according to Etter’s rules [51,52]. Additionally, in 24, D motifs are also constructed at the second level. Thus, only two types of non-classical interactions resulted in a rich portfolio of supramolecular H-bonding motifs (Table S15). It can be observed that the benzyltrimethylammonium cation may be considered as a supramolecular tecton engaging in the building of synthonic recurring cyclic motifs in all the analyzed crystals. Only heterosynthons are generated, by different functional groups, in all supramolecular systems. All cyclic intermolecular synthons (R) are formed only between cations and anions.
More importantly, supramolecular interchangeable/equivalent synthons are observed (Figure 7 and Figure S1). They may be helpful in the design of similar compounds with controllable properties. Among them, bifurcated synthons are dominant. It should be noted that some of the robust synthons are stabilized by ππ (in 2) and Cu–Brπ (in 4) (Figure 7) interactions. It is worth mentioning that the R12(6) bifurcated synthon is present in all four crystal structures.
Taking into account the aromatic interactions between cations and the C–HCl contacts between differently charged entities, the topology of the resulting supramolecular networks was analyzed. In Figure 8, the cations are reduced to a node (black balls), representing the center of gravity of each moiety. In compounds 3 and 4, two crystallographically independent cations are shown in white and black, respectively. In 1, the cations are linked via two C–Hπ interactions between aromatic rings, with one of the CH3 groups acting as a weak donor. These two interactions operate between the same pair of cations, resulting in the formation of a chain of cations propagating along the shortest unit cell axis c. Two detected cation–anion hydrogen bonds of the C–HCl type link these chains into an undulated layer perpendicular to the [100] direction. The topology descriptor of the layer can be given as 4(3·2.6) + 4(3·2.6) and its symmetry is in accordance with the p21/b layer symmetry group.
The main interaction occurring between cations in 2 is ππ stacking, which topologically leads to the formation of a dimer. The two independent anions bind weakly to the cations via C–HCl interactions, resulting directly in the formation of a three-dimensional (3D) structure. It should be noted that the topological environment of these two anions is different (LK = 2 and LK = 4), which further complicates the description of this network. However, it can be seen that one of the anions (LK = 4) and the cation form infinite chains along the [10-1] direction with a 4(3·2) + 4(3·2) topology and the p2/c rod symmetry group. Through the second anion (vertex with LK = 2), they combine into a 3D structure. In 3, there are two crystallographically independent cations, denoted by white and black balls, respectively. Weak aromatic interactions join two different cations, forming a dimeric, finite motif, whereas relatively short CH3π contacts are observed between cations labeled as B (black balls). The connections visible in Figure 8 are denoted with red and orange lines, topologically representing a decorated chain along the [010] direction. Together with the observed C–HCl hydrogen bonds (see Table S15, shown as green shaded lines in Figure 8), the resulting topology forms a layer perpendicular to the longest unit cell vector c. The topological descriptor can be given as 4(4·4.64) + 5(4·5) + 2(4) and the symmetry of the layer group is p21/b. Similarly to salt 3, in 4, two independent cations are connected via weak aromatic interactions, forming a decorated chain propagating along the [100] direction. However, more numerous C–HBr hydrogen bonds, as well as detected Brπ interactions, link the supramolecular cationic chains directly into a three-dimensional (3D) structure. In this case, the topology can be described using the same layer descriptors as in structure 3, and only the linkers between the metal vertex and the cation are taken into account, marked as a black sphere.
Interestingly, the crystal structure similarity calculations (performed using Mercury) revealed that, in the crystals of 3 and 4, 13 out of 15 molecules are identically oriented (see Figure S2; the mismatched molecule is denoted with a circle).

3.3. Hirshfeld Surface Analysis

A Hirshfeld surface (HS) extended analysis was performed to obtain a deeper insight into the hierarchical preferences of close non-covalent interactions in the supramolecular assemblies of 14. HS analysis is a valuable advancement that enables crystal engineers, including supramolecular chemists, to understand crystal packing behavior and design desired hierarchically ordered structures, thereby modulating their properties. HS maps were generated using not only a standard surface solution of dnorm surfaces but also shape index, curvedness, and fragment patch surfaces (Figure 9). The Au-containing structure (2) was not included in the Hirshfeld surface and energy framework analyses owing to limitations of CrystalExplorer in treating such heavy atoms.
The surfaces were calculated for the cation, which is identical in all crystals, to identify differences in different environments. It should be noted that the red spots on the domain surfaces signify strong interactions, white indicates moderate interactions, while blue indicates negligible interactions. A thorough analysis of the maps revealed that the red regions are predominantly localized around the methyl groups (and also the phenyl groups) of the cations, signifying that this is a primary hydrogen-bonding site involved in intermolecular interactions. The surfaces are indicative of non-classical hydrogen-bonding contacts, representing interactions with symmetry-related anions. More specifically, the small red spots on the maps of the dnorm surfaces characterize closer distances than the van der Waals radii; here, H-bonding interactions such as C–HCl(Br) can be observed. To further investigate the similarities and differences in the weak interactions among the analyzed crystals, color maps of additional Hirshfeld surface properties were generated. Alternating red and blue areas indicate different weak interactions connecting cations with anions. Poorly visible adjacent red and blue triangles on the shape index surfaces, as well as poorly outlined flat green regions on the curvedness maps, indicate a lack of significant π-based interactions in the analyzed crystals. However, yellow–orange deformations in 4 may reveal Cu–Brπ interconnects. The colored fragment patches may indicate the number of nearest surrounding moieties that interact with the cation.
According to the commonly accepted criteria in Hirshfeld surface analysis, only contacts contributing more than 0.9% to the Hirshfeld surface are considered statistically meaningful for interpretation. In the studied structures, the CuH/HCu and ClH/HCl contacts exceed this threshold (1.5–1.6%), which allows their reliable quantitative evaluation. Their enrichment ratios (E ≥ 1) further indicate that these interactions are statistically favored, although not dominant, within the crystal packing.
Fingerprint plots summarizing the intricate information are unique for each crystal. Nevertheless, the calculated fingerprint diagrams for 14 are very similar. They are characteristic of strong non-classical interactions due to a lack of inter alia long spikes (specific for OH/HO hydrogen bonds) on the histograms. The HH contacts are observed in the middle area between the short spikes of ClH/HCl (in 1 and 3) and BrH/HBr (in 4), while the CH/HC interactions are visualized as ‘wings’ in the upper side of fingerprint plots of all crystals (Figure 10). HH interactions are the largest contributors to the total crystal packing (from 55.2% in 1 to 61.7% in 3), while CH/HC interactions are the next most significant contributors, representing from 9.5% in 3 to ~19% in 1. ClH/HCl (in 1 and 3) and BrH/HBr (in 4) are additional contributors, at the 25.5% level. Furthermore, CuH/HCu (~1.6%) in 3 and 4, and CBr/BrC in 4 (1.5%) are also present. Notably, CC interactions represent only a small fraction in 3 and 4 (below 1%). The percentage contribution of close contacts in these crystal structures is depicted in Figure 10. When considering the entire salts, halogen bonds ClCl and ClPt/PtCl in 1 (~5%) are additionally observed (Figure S3). The enrichment ratio calculations based on Hirshfeld surface analysis indicate the propensity for specific intercontacts. Only contacts contributing more than 0.9% to the Hirshfeld surface are considered statistically meaningful [44]. In our structures, CuH/HCu and ClH/HCl contacts exceed this threshold (1.5–1.6%), allowing a reliable evaluation of their enrichment ratios. The calculated ratios (E ≥ 1) show that these interactions are statistically favored, although not dominant, in the crystal packing. Similarly, BrH/HBr and BrC/CBr interactions in 4 are also favored. CH/HC interactions contribute prominently to all analyzed crystals. Notably, HH contacts are always disfavored in the crystal packing despite their high contribution to the HSs in all structures (Tables S16–S20). In the context of halogen bonds such as ClCl, these interactions do not play a significant role in the crystal packing.

3.4. Energy Frameworks

The energy framework concept is helpful in either interpreting or visualizing the pairwise interaction energy between the neighboring moieties in the crystal. The 3D energy frameworks for the analyzed crystals, emphasizing the neighboring moieties in a radius of 3.8 Å from the central moiety, were calculated for the investigated salts. Energy types such as coulombic (red) and dispersion (green) are illustrated in Figure 11, Figure 12, Figure 13, Figure 14 and Figure 15. It should be mentioned that the strength of interactions correlates with the size of the cylinders. The details, such as the crystallographic symmetry operations and their corresponding molecular energy values, which are crucial for calculating lattice energy [53] are summarized in Figures S4–S6. Salts 1 and 4 may be considered to be the most stable thermodynamically. It is noteworthy that the dispersion term is the significantly dominant component in all the analyzed crystals (Table S21), meaning that van der Waals forces have relevance.

3.5. Quantum Chemical Study

The relevant structural parameters of the X-ray structures under study and of the MP2-optimized structure of BTMA+ in aqueous solutions are presented in Table 3. As expected, the largest differences are observed for the torsion angles. The X-ray BTMA+ structures are affected by environmental influences, and so they lost their mirror plane. The symmetry group Cs is preserved in the MP2-optimized structures. Another problem with the X-ray structures is the location of the H-atoms. The C–H bond lengths are fixed at 0.92–0.96 Å, whereas the MP2 optimizations produce the value of 1.08–1.09 Å, which should be closer to reality. In general, the differences between the X-ray structures of BTMA+ in Table 3 are comparable with the differences between them and the MP2-optimized ones.
The overlay of the X-ray structure of BTMA+ in (BTMA)2[PtCl6] and the MP2-optimized ones in vacuum and in aqueous solutions is shown in Figure 16. The differences between them can be explained as above. A quantitative measure of similarity between two or more structures is the root mean square deviation (RMSD) of their corresponding atomic positions (i.e., the average distance between the corresponding atoms of superimposed molecules). The data in Table 4 confirm that the similarity between the X-ray and both MP2-optimized structures is significantly lower than that between the MP2-optimized structures in vacuum and in aqueous solution.
The energies and locations of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) are crucial for the chemical reactivity of compounds. According to the Fukui rules [54,55], the locations where the HOMO electron density is largest are most readily attacked by electrophilic or oxidizing reagents. Analogously, the locations of the maximal LUMO density are most probably attacked by nucleophilic or reducing reagents. An additional condition is the suitable orbital energy of the attacking reagent. The energy gap between HOMO and LUMO in BTMA+ is approximately 12 eV, indicating its high stability (Table 5). An aqueous solvent increases the BTMA+ orbital energies and stability. Both HOMO and LUMO (Figure 17 and Figure 18) are located mainly at the phenyl ring. It can be expected that in the case of basic BTMA+ degradation [19,20,21,56], the attacking OH- anion is subsequently transferred to the C7 site and the tetramethylammonium group is split.
Additionally, active sites of BTMA+ can be deduced on the basis of its electrostatic potential map (Figure 19). As expected, negative potential values are observed at the aromatic phenyl π ring. The positive potential at hydrogen atoms explains the less-preferred formation of H2O molecules under OH- attack [19].
The molecular graph of BTMA+ consisting of atoms, the bond paths between them, and critical points is presented in Figure 20. The corresponding atomic charges and volumes are collected in Table 6. The charges of the phenyl carbons are slightly negative, unlike the positive charges of the aliphatic carbons. Hydrogen charges are slightly positive. The volumes of atoms with the same bonding pattern should be larger for more negatively charged (less positively charged) atoms. This is in contradiction with the lower volumes of the phenyl C atoms in ortho-positions, which indicates their higher reactivity in comparison with the C atoms in the meta- and para-positions. An analogous situation also holds for the phenyl hydrogens.
The selected properties of the BTMA+ bonds are presented in Table 7. Bond strengths can be compared according to electron densities at their bond critical points, ρBCP, or, alternatively, according to the average number of electrons that are shared between bonded atoms (delocalization indices, DI). Both treatments indicate maximal strengths of the phenyl aromatic bonds, whereas the C7–N and N–Cmet bonds are the weakest ones (even weaker than the C–H bonds). This finding is in agreement with the experimentally observed degradation in basic solutions [19,20,21,56] connected with the splitting of these weakest bonds, as follows.
C6H5CH2N+(CH3)3 + OH→C6H5CH2OH + N(CH3)3
C6H5CH2N+(CH3)3 + OH→C6H5CH2N(CH3)2 + CH3OH
All bonds in BTMA+ have negative values of BCP Laplacians of electron density, ∇2ρBCP, which are typical for covalent bonds. Their values reflect the various polarities of individual bond types.
Bond ellipticities, εBCP, reflect the double-bond π character of individual bonds. We can see that the ellipticities of the phenyl aromatic bonds are approximately one order higher than their σ bond counterparts (single C1–C7, C7-N, N-Cmet, and all C-H bonds).

4. Conclusions

In summary, we successfully obtained four salts of benzyltrimethylammonium, namely, bis(benzyltrimethylammonium) hexachloroplatinate (1), benzyltrimethylammonium tetrachloroaurate (2), bis(benzyltrimethylammonium) tetrachlorocuprate (3), and bis(benzyltrimethylammonium) tetrabromocuprate (4). Interestingly, in 1, the anion is completed by inversion symmetry. Overall, the synthetized salts demonstrate varied supramolecular features arising from different anionic species. However, high similarity is observed in the 3 and 4 crystals. The supramolecular systems of the investigated salts, driven by only non-classical interactions, were thoroughly explored using diverse modern computational approaches.
Hirshfeld surface analysis and enrichment ratio calculations indicate that CH/HC interactions contribute prominently to the stabilization of all supramolecular assemblies. In addition, ClH/HCl and CuH/HCu interactions have relevance in 13, and in 34, respectively. BrH/HBr, BrC/CBr interactions are favorable in 4. The hierarchy of synthons was examined using extended Hirshfeld surface analysis, 2D fingerprint plots, enrichment ratio, and energy framework calculations. As a result, a library of hydrogen-bonding motifs was constructed. Equivalent synthons were identified that may serve as interchangeable motifs in the design of derivatives with controllable features. A bifurcated synthon R12(6) between cation and anion is present in all four crystals.
Our MP2 calculations pointed out differences between the structures in solutions and the solid-state structures affected by intermolecular interactions. The Cs symmetry point group is preserved in solutions, which is reflected in the symmetry of their electronic structural characteristics. Frontier molecular orbitals (HOMO and LUMO), the electrostatic potential map, and electron structure knowledge enable the explanation of BTMA+ reactivity and its active sites.
The long-distance ordering in a crystal resembles the local ordering in a liquid [57]. In liquid solutions, a short-range order of solvent molecules persists, and the solvent molecules move in the potential field of the solute ions or molecules. The liquid solvent contains regions with ordered bonds of a regular lattice as well as regions with non-ordered molecules in a random array. Its dynamic behavior with short rotational and translational correlation times indicates high bond-exchange rates. Solvent effects can affect chemical reactions (reaction rate, mechanism, and selectivity), chemical equilibria (relative stability of ions and transition states), and physical properties (solubility, crystal shape, and shifts in spectral lines).
This article has been completed while Sepideh Jafari, the fifth author, was the Doctoral Candidate in the Interdisciplinary Doctoral School at Lodz University of Technology, Poland.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/cryst15121051/s1. Figure S1: Equivalent/interchangeable H-bonding synthons observed in the analyzed crystals 14 (X1, X2 = Cl, Br). Figure S2: Crystal packing similarity between 3 (green) and 4 (orange). Figure S3: Upper: Percentage contributions of inter-contacts in crystals 14 (above 1%)—calculations for entire salt; Below: fingerprint plots for overall interactions for 14. Figure S4: Energy framework calculations for cations in 1. Figure S5: Energy framework calculations for cations in 3. Figure S6: Energy framework calculations for cations in 4. Table S1: Bond lengths for 1. Table S2: Valence angles for 1. Table S3: Torsion angles for 1. Table S4: Bond lengths for 2. Table S5: Valence angles for 2. Table S6: Torsion angles for 2. Table S7: Bond lengths for 3. Table S8: Valence angles for 3. Table S9: Torsion angles for 3. Table S10: Bond lengths for 4. Table S11: Valence angles for 4. Table S12: Torsion angles for 4. Table S13: Π-based inter-contacts in 14 (<4 Å). Table S14: Aromatic Analyser results. Table S15: Library of H-bonding synthons (<20-membered) in 1–4. Table S16: Enrichment ratios for cations in 1 (>0.9). Table S17: Enrichment ratio for cation no. 1 in 3 (>0.9). Table S18: Enrichment ratio for cation no. 2 in 3 (>0.9). Table S19: Enrichment ratio for cation no. 1 in 4 (>0.9). Table S20: Enrichment ratio for cation no. 2 in 4 (>0.9). Table S21: The total interaction energies for 14 via energy framework calculations.

Author Contributions

Conceptualization, A.M. and J.B.; methodology, M.B., J.B., A.M. and D.T.; validation, D.T. and J.B.; formal analysis, J.B., D.T. and M.B.; investigation, D.T., J.B., A.M., M.B., I.D.M., K.W., S.J. and I.J.; data curation, D.T., K.W. and J.B.; writing—original draft preparation, J.B., D.T., M.B. and A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been supported by the Slovak Scientific Grant Agency VEGA (project no. 1/0175/23).

Data Availability Statement

CCDC 2464968–2464971 contain the supplementary crystallographic data for this paper, accessed on 17 June 2025. These data can be obtained free-of-charge via https://www.ccdc.cam.ac.uk/structures.

Acknowledgments

The authors thank the HPC center at the Slovak University of Technology in Bratislava, which is a part of the Slovak Infrastructure of High-Performance Computing (SIVVP project ITMS 26230120002, funded by European Region Development Funds) for the computational time and resources made available.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hasan, M.; Kozhevnikov, I.V.; Siddiqui, M.R.H.; Steiner, A.; Winterton, N. Gold compounds as ionic liquids. Synthesis, structures, and thermal properties of N, N ‘-Dialkylimidazolium tetrachloroaurate salts. Inorg. Chem. 1999, 38, 5637–5641. [Google Scholar] [CrossRef]
  2. Yamamoto, K.; Inada, S. Liquid-liquid distribution of ion associates of tetrahalogenoaurate (III) with quaternary ammonium counter ions. Anal. Sci. 1995, 11, 643–649. [Google Scholar] [CrossRef]
  3. Kawasaki, H.; Uota, M.; Yoshimura, T.; Fujikawa, D.; Sakai, G.; Arakawa, R.; Kijima, T. Self-Organization of Surfactant−Metal-Ion Complex Nanofibers on Graphite Surfaces and Their Application to Fibrously Concentrated Platinum Nanoparticle Formation. Langmuir 2007, 23, 11540–11545. [Google Scholar] [CrossRef]
  4. Tkacheva, A.; Sharutin, V.; Sharutina, O.; Shlepotina, N.; Kolesnikov, O.; Shishkova, Y.S.; Peshikova, M. Tetravalent platinum complexes: Synthesis, structure, and antimicrobial activity. Russ. J. Gen. Chem. 2020, 90, 655–659. [Google Scholar] [CrossRef]
  5. Xie, D.; Xu, J.; Cheng, H.; Wang, N.; Zhou, Q. The role played by amine and ethyl group in the reversible thermochromic process of [(C2H5)2NH2]2CuCl4 probing by FTIR and 2D-COS analysis. J. Mol. Struct. 2018, 1161, 267–272. [Google Scholar] [CrossRef]
  6. Mande, H.M.; Ghalsasi, P.S.; Navamoney, A. Synthesis, structural and spectroscopic characterization of the thermochromic compounds A2CuCl4:[(Naphthyl ethylammonium)2CuCl4]. Polyhedron 2015, 91, 141–149. [Google Scholar] [CrossRef]
  7. Gilmore, B.F.; Andrews, G.P.; Borberly, G.; Earle, M.J.; Gilea, M.A.; Gorman, S.P.; Lowry, A.F.; McLaughlin, M.; Seddon, K.R. Enhanced antimicrobial activities of 1-alkyl-3-methyl imidazolium ionic liquids based on silver or copper containing anions. New J. Chem. 2013, 37, 873–876. [Google Scholar] [CrossRef]
  8. Kaur, G.; Kumar, S.; Dilbaghi, N.; Bhanjana, G.; Guru, S.K.; Bhushan, S.; Jaglan, S.; Hassan, P.; Aswal, V. Hybrid surfactants decorated with copper ions: Aggregation behavior, antimicrobial activity and anti-proliferative effect. Phys. Chem. Chem. Phys. 2016, 18, 23961–23970. [Google Scholar] [CrossRef]
  9. Li, D.; Zhao, X.-M.; Zhao, H.-X.; Long, L.-S.; Zheng, L.-S. Coexistence of magnetic-optic-electric triple switching and thermal energy storage in a multifunctional plastic crystal of trimethylchloromethyl ammonium tetrachloroferrate (III). Inorg. Chem. 2018, 58, 655–662. [Google Scholar] [CrossRef]
  10. Hayashi, S.; Saha, S.; Hamaguchi, H. A new class of magnetic fluids: Bmim [fecl/sub 4/] and nbmim [fecl/sub 4/] ionic liquids. IEEE Trans. Magn. 2006, 42, 12–14. [Google Scholar] [CrossRef]
  11. Desiraju, G.R. Crystal engineering: A holistic view. Angew. Chem. Int. Ed. 2007, 46, 8342–8356. [Google Scholar] [CrossRef]
  12. Aakeröy, C.B.; Champness, N.R.; Janiak, C. Recent advances in crystal engineering. CrystEngComm 2010, 12, 22–43. [Google Scholar] [CrossRef]
  13. Bojarska, J.; Łyczko, K.; Breza, M.; Mieczkowski, A. Recurrent Supramolecular Patterns in a Series of Salts of Heterocyclic Polyamines and Heterocyclic Dicarboxylic Acids: Synthesis, Single-Crystal X-ray Structure, Hirshfeld Surface Analysis, Energy Framework, and Quantum Chemical Calculations. Crystals 2024, 14, 733. [Google Scholar] [CrossRef]
  14. Bojarska, J.; Łyczko, K.; Mieczkowski, A. Synthesis, Crystal Structure and Supramolecular Features of Novel 2, 4-Diaminopyrimidine Salts. Crystals 2024, 14, 133. [Google Scholar] [CrossRef]
  15. Bojarska, J.; Łyczko, K.; Mieczkowski, A. Novel Salts of Heterocyclic Polyamines and 5-Sulfosalicylic Acid: Synthesis, Crystal Structure, and Hierarchical Supramolecular Interactions. Crystals 2024, 14, 497. [Google Scholar] [CrossRef]
  16. Lian, L.; Zhang, P.; Gao, J.; Zhang, D.; Zhang, J. Self-assemblies of isomeric copper iodide trimers with geometry-dependent photophysical properties. Chem. Mater. 2023, 35, 9339–9345. [Google Scholar] [CrossRef]
  17. Mączka, M.; Gągor, A.; Stroppa, A.; Gonçalves, J.N.; Zaręba, J.K.; Stefańska, D.; Pikul, A.; Drozd, M.; Sieradzki, A. Two-dimensional metal dicyanamide frameworks of BeTriMe [M(dca)3(H2O)] (BeTriMe = benzyltrimethylammonium; dca = dicyanamide; M = Mn2+, Co2+, Ni2+): Coexistence of polar and magnetic orders and nonlinear optical threshold temperature sensing. J. Mater. Chem. C 2020, 8, 11735–11747. [Google Scholar] [CrossRef]
  18. Maczka, M.; Ptak, M.; Trzebiatowska, M.; Kucharska, E.; Hanuza, J.; Palka, N.; Czerwinska, E. THz, Raman, IR and DFT studies of noncentrosymmetric metal dicyanamide frameworks comprising benzyltrimethylammonium cations. Spectrochim. Acta Part A-Mol. Biomol. Spectrosc. 2021, 251, 119416. [Google Scholar] [CrossRef]
  19. Chempath, S.; Boncella, J.M.; Pratt, L.R.; Henson, N.; Pivovar, B.S. Density functional theory study of degradation of tetraalkylammonium hydroxides. J. Phys. Chem. C 2010, 114, 11977–11983. [Google Scholar] [CrossRef]
  20. Sturgeon, M.R.; Macomber, C.S.; Engtrakul, C.; Long, H.; Pivovar, B.S. Hydroxide based benzyltrimethylammonium degradation: Quantification of rates and degradation technique development. J. Electrochem. Soc. 2015, 162, F366. [Google Scholar] [CrossRef]
  21. Karibayev, M.; Myrzakhmetov, B.; Kalybekkyzy, S.; Wang, Y.; Mentbayeva, A. Binding and degradation reaction of hydroxide ions with several quaternary ammonium head groups of anion exchange membranes investigated by the DFT method. Molecules 2022, 27, 2686. [Google Scholar] [CrossRef]
  22. McCrum, I.; Hickner, M.; Janik, M. Quaternary ammonium cation specific adsorption on platinum electrodes: A combined experimental and density functional theory study. J. Electrochem. Soc. 2018, 165, F114. [Google Scholar] [CrossRef]
  23. Guan, K.; Tao, L.; Yang, R.; Zhang, H.; Wang, N.; Wan, H.; Cui, J.; Zhang, J.; Wang, H.; Wang, H. Anti-corrosion for reversible zinc anode via a hydrophobic interface in aqueous zinc batteries. Adv. Energy Mater. 2022, 12, 2103557. [Google Scholar] [CrossRef]
  24. Wang, T.; Zheng, D.; Zhang, Z.; Wang, L.; Zhang, J. Exploration of catalytic species for highly efficient preparation of quinazoline-2,4(1H, 3H)-diones by succinimide-based ionic liquids under atmospheric pressure: Combination of experimental and theoretical study. Fuel 2022, 319, 123628. [Google Scholar] [CrossRef]
  25. Bonamico, M.; Dessy, G.; Vaciago, A. The crystal structure of bis-trimethylbenzylammoniumtetrachlorocuprate (II). Theor. Chim. Acta 1967, 7, 367–374. [Google Scholar] [CrossRef]
  26. Jin, L.; Liu, N.; Li, Y.-J.; Wu, D.-H. Bis (benzyltrimethylammonium) tetrabromidocuprate (II). Struct. Rep. 2011, 67, m1325. [Google Scholar] [CrossRef]
  27. Jin, Y.; Yu, C.-H.; Zhang, W. Structural diversity of a series of chlorocadmate (II) and chlorocuprate (II) complexes based on benzylamine and its N-methylated derivatives. J. Coord. Chem. 2014, 67, 1156–1173. [Google Scholar] [CrossRef]
  28. CrysAlis, C. CrysAlis Red. Xcalibur PX Software, version 1.171.40.67a; Oxford Diffraction Ltd.: Abingdon, UK, 2008.
  29. Clark, R.; Reid, J. The analytical calculation of absorption in multifaceted crystals. Found. Crystallogr. 1995, 51, 887–897. [Google Scholar] [CrossRef]
  30. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71 Pt 1, 3–8. [Google Scholar] [CrossRef]
  31. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  32. Head-Gordon, M.; Pople, J.A.; Frisch, M.J. MP2 energy evaluation by direct methods. Chem. Phys. Lett. 1988, 153, 503–506. [Google Scholar] [CrossRef]
  33. Head-Gordon, M.; Head-Gordon, T. Analytic MP2 frequencies without fifth-order storage. Theory and application to bifurcated hydrogen bonds in the water hexamer. Chem. Phys. Lett. 1994, 220, 122–128. [Google Scholar] [CrossRef]
  34. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Petersson, G.; Nakatsuji, H. GAUSSIAN09, version 9.0; Gaussian Inc.: Wallingford, CT, USA, 2011.
  35. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  36. Varetto, U. Molekel, version 5.4. 0.8; Swiss National Supercomputing Centre: Manno, Switzerland, 2009.
  37. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef]
  38. Bader Richard, F. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, UK, 1994. [Google Scholar]
  39. Keith, T.A. AIMAll, version 17.11. 14; TK Gristmill Software: Overland Park, KS, USA, 2017.
  40. Biegler-Konig, F.; Schonbohm, J.; Bayles, D. Software News and Updates-AIM2000-A Program to Analyze and Visualize Atoms in Molecules; John Wiley & Sons Inc.: New York, NY, USA, 2001; pp. 545–559. [Google Scholar]
  41. Schrödinger, L. The AxPyMOL Molecular Graphics Plugin for Microsoft PowerPoint, version 1.8; Schroedinger, LLC: New York, NY, USA, 2015.
  42. Kabsch, W. A solution for the best rotation to relate two sets of vectors. Found. Crystallogr. 1976, 32, 922–923. [Google Scholar] [CrossRef]
  43. Turner, M.; McKinnon, J.; Wolff, S.; Grimwood, D.; Spackman, P.; Jayatilaka, D.; Spackman, M. CrystalExplorer, version 17.5; University of Western Australia: Crawley, Australia, 2017.
  44. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  45. Jelsch, C.; Ejsmont, K.; Huder, L. The enrichment ratio of atomic contacts in crystals, an indicator derived from the Hirshfeld surface analysis. IUCrJ 2014, 1, 119–128. [Google Scholar] [CrossRef]
  46. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  47. Mackenzie, C.F.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer model energies and energy frameworks: Extension to metal coordination compounds, organic salts, solvates and open-shell systems. IUCrJ 2017, 4, 575–587. [Google Scholar] [CrossRef]
  48. Grimme, S. Semiempirical GGA—type density functional constructed with a long—range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  49. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M. Mercury 4.0: From visualization to analysis, design and prediction. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef]
  50. Wood, P.A.; Olsson, T.S.; Cole, J.C.; Cottrell, S.J.; Feeder, N.; Galek, P.T.; Groom, C.R.; Pidcock, E. Evaluation of molecular crystal structures using Full Interaction Maps. CrystEngComm 2013, 15, 65–72. [Google Scholar] [CrossRef]
  51. Bernstein, J.; Davis, R.E.; Shimoni, L.; Chang, N.L. Patterns in hydrogen bonding: Functionality and graph set analysis in crystals. Angew. Chem. Int. Ed. Engl. 1995, 34, 1555–1573. [Google Scholar] [CrossRef]
  52. Etter, M.C. Encoding and decoding hydrogen-bond patterns of organic compounds. Acc. Chem. Res. 1990, 23, 120–126. [Google Scholar] [CrossRef]
  53. Tan, S.L.; Jotani, M.M.; Tiekink, E.R. Utilizing Hirshfeld surface calculations, non-covalent interaction (NCI) plots and the calculation of interaction energies in the analysis of molecular packing. Struct. Rep. 2019, 75, 308–318. [Google Scholar] [CrossRef]
  54. Fukui, K.; Yonezawa, T.; Nagata, C.; Shingu, H. Molecular orbital theory of orientation in aromatic, heteroaromatic, and other conjugated molecules. J. Chem. Phys. 1954, 22, 1433–1442. [Google Scholar] [CrossRef]
  55. Fukui, K. Role of frontier orbitals in chemical reactions. Science 1982, 218, 747–754. [Google Scholar] [CrossRef]
  56. Long, H.; Pivovar, B.S. Hydroxide degradation pathways for substituted benzyltrimethyl ammonium: A DFT study. ECS Electrochem. Lett. 2015, 4, F13. [Google Scholar] [CrossRef]
  57. Reichardt, C.; Welton, T. Solvents and Solvent Effects in Organic Chemistry; Wiley Online Library: Hoboken, NJ, USA, 2011. [Google Scholar]
Scheme 1. The synthesis of the analyzed salts: bis(benzyltrimethylammonium) hexachloroplatinate (1), benzyltrimethylammonium tetrachloroaurate (2), bis(benzyltrimethylammonium) tetrachlorocuprate (3), and bis(benzyltrimethylammonium) tetrabromocuprate (4).
Scheme 1. The synthesis of the analyzed salts: bis(benzyltrimethylammonium) hexachloroplatinate (1), benzyltrimethylammonium tetrachloroaurate (2), bis(benzyltrimethylammonium) tetrachlorocuprate (3), and bis(benzyltrimethylammonium) tetrabromocuprate (4).
Crystals 15 01051 sch001
Figure 1. (a) Molecular structures of 14 showing the atom numbering scheme. (b) Overlay of the structural skeletons of the benzyltrimethylammonium moieties in the crystals of 14.
Figure 1. (a) Molecular structures of 14 showing the atom numbering scheme. (b) Overlay of the structural skeletons of the benzyltrimethylammonium moieties in the crystals of 14.
Crystals 15 01051 g001
Figure 2. The 3D full interaction maps of 14, with the possible areas of weak acceptors in dark red.
Figure 2. The 3D full interaction maps of 14, with the possible areas of weak acceptors in dark red.
Crystals 15 01051 g002
Figure 3. Crystal packing of 1 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Figure 3. Crystal packing of 1 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Crystals 15 01051 g003
Figure 4. Crystal packing of 2 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Figure 4. Crystal packing of 2 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Crystals 15 01051 g004
Figure 5. Crystal packing of 3 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Figure 5. Crystal packing of 3 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Crystals 15 01051 g005
Figure 6. Crystal packing of 4 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Figure 6. Crystal packing of 4 along the (a) (on the left), (b) (middle), and (c) (on the right) crystallographic axes.
Crystals 15 01051 g006
Figure 7. Fragments of the supramolecular architectures of 14 showing the intermolecular interactions between neighboring moieties and recurring H-bonding motifs that are stabilized by ππ interactions in 2 and Cu–Brπ interactions in 4 (pink circles).
Figure 7. Fragments of the supramolecular architectures of 14 showing the intermolecular interactions between neighboring moieties and recurring H-bonding motifs that are stabilized by ππ interactions in 2 and Cu–Brπ interactions in 4 (pink circles).
Crystals 15 01051 g007
Figure 8. Topological representation of intermolecular interactions in the 14 crystals. The black and white balls represent the centers of gravity of the cations. The anion is reduced to the position of the heavy metal only. The red and orange lines represent C–H...π and ππ interactions between cations. Links are shown in green, and cyan shades are in line with the C–H...Cl(Br) hydrogen bonds. In 4, the brownish line represents Br...π interactions.
Figure 8. Topological representation of intermolecular interactions in the 14 crystals. The black and white balls represent the centers of gravity of the cations. The anion is reduced to the position of the heavy metal only. The red and orange lines represent C–H...π and ππ interactions between cations. Links are shown in green, and cyan shades are in line with the C–H...Cl(Br) hydrogen bonds. In 4, the brownish line represents Br...π interactions.
Crystals 15 01051 g008
Figure 9. Comparison of Hirshfeld surface maps for cations in 1,3 and 4: (a) dnorm, (b) shape index, (c) curvedness, (d) fragment patches, (e) di, and (f) de.
Figure 9. Comparison of Hirshfeld surface maps for cations in 1,3 and 4: (a) dnorm, (b) shape index, (c) curvedness, (d) fragment patches, (e) di, and (f) de.
Crystals 15 01051 g009
Figure 10. (a) Fingerprint plots for the overall interactions for 1,3 and 4 (b) Percentage contributions of inter-contacts in crystals 1,3 and 4 (above 1%). The enrichment ratio values are signified in the diagram by the same colors as their corresponding interactions.
Figure 10. (a) Fingerprint plots for the overall interactions for 1,3 and 4 (b) Percentage contributions of inter-contacts in crystals 1,3 and 4 (above 1%). The enrichment ratio values are signified in the diagram by the same colors as their corresponding interactions.
Crystals 15 01051 g010
Figure 11. Energy framework plots for cations in 1, showing the electrostatic and dispersion terms along a, b, and, c crystallographic axes, as well as the total energy interactions (the tube size is set to 100).
Figure 11. Energy framework plots for cations in 1, showing the electrostatic and dispersion terms along a, b, and, c crystallographic axes, as well as the total energy interactions (the tube size is set to 100).
Crystals 15 01051 g011
Figure 12. Energy framework plots for cation no. 1 in 3, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Figure 12. Energy framework plots for cation no. 1 in 3, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Crystals 15 01051 g012
Figure 13. Energy framework plots for cation no. 2 in 3, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Figure 13. Energy framework plots for cation no. 2 in 3, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Crystals 15 01051 g013
Figure 14. Energy framework plots for cation no. 1 in 4, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Figure 14. Energy framework plots for cation no. 1 in 4, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Crystals 15 01051 g014
Figure 15. Energy framework plots for cation no. 2 in 4, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Figure 15. Energy framework plots for cation no. 2 in 4, showing the electrostatic and dispersion terms as well as the total energy interactions (the tube size is set to 100).
Crystals 15 01051 g015
Figure 16. Overlay of the X-ray structure of the benzyltrimethylammonium cation in (C10H16N)2[PtCl6] (C—gray, N—blue) and of its DFT-optimized structures in vacuum (green) and in aqueous solution (magenta). H-atoms have been omitted for the sake of clarity.
Figure 16. Overlay of the X-ray structure of the benzyltrimethylammonium cation in (C10H16N)2[PtCl6] (C—gray, N—blue) and of its DFT-optimized structures in vacuum (green) and in aqueous solution (magenta). H-atoms have been omitted for the sake of clarity.
Crystals 15 01051 g016
Figure 17. HOMO of BTMA+ in aqueous solution (drawn at 0.07 a.u. isosurface).
Figure 17. HOMO of BTMA+ in aqueous solution (drawn at 0.07 a.u. isosurface).
Crystals 15 01051 g017
Figure 18. LUMO of BTMA+ in aqueous solution (drawn at 0.07 a.u. isosurface).
Figure 18. LUMO of BTMA+ in aqueous solution (drawn at 0.07 a.u. isosurface).
Crystals 15 01051 g018
Figure 19. Electrostatic potential map of BTMA+ in aqueous solutions (isovalue 0.1 a.u., blue and red colors denote positive and negative values, respectively).
Figure 19. Electrostatic potential map of BTMA+ in aqueous solutions (isovalue 0.1 a.u., blue and red colors denote positive and negative values, respectively).
Crystals 15 01051 g019
Figure 20. Molecular graph of BTMA+ in aqueous solution (C—black, N—blue, H—gray, bond critical points—red, and ring critical point—yellow).
Figure 20. Molecular graph of BTMA+ in aqueous solution (C—black, N—blue, H—gray, bond critical points—red, and ring critical point—yellow).
Crystals 15 01051 g020
Table 1. Crystal data and refinement details of 14.
Table 1. Crystal data and refinement details of 14.
Compound1 234
Empirical formula C20H32Cl6N2Pt C10H16AuCl4N C20H32Cl4CuN2 C20H32Br4CuN2
Formula weight 708.26 489.00 505.81 683.65
Temperature/K 100(2) 100(2) 100(2) 100(2)
Crystal system monoclinic monoclinic monoclinic orthorhombic
Space group P21/cC2/cP21/nP212121
a10.9703(2) 16.7055(3) 9.49010(16) 9.1205(2)
b11.9807(3) 14.3876(2) 9.02423(19) 9.6086(3)
c9.6767(2) 13.9779(2) 27.9964(4) 28.8970(7)
α90 90 90 90
β91.7727(19) 114.613(2) 92.1082(14) 90
γ90 90 90 90
Volume/Å3 1271.22(5) 3054.37(9) 2396.01(8) 2532.38(11)
Z2 8 4 4
ρcalcg/cm31.850 2.127 1.402 1.793
μ/mm−116.206 10.308 5.442 8.661
F(000) 692.0 1840.0 1052.0 1340.0
Crystal size/mm3 0.39 × 0.06 × 0.05 0.18 × 0.09 × 0.04 0.24 × 0.17 × 0.08 0.36 × 0.28 × 0.06
Radiation CuKα (λ = 1.54184) MoKα (λ = 0.71073) CuKα (λ = 1.54184) CuKα (λ = 1.54184)
2Θ range for data collection/° 8.064 to 136.446 4.28 to 52.744 6.318 to 134.158 6.118 to 134.138
Index ranges −13 ≤ h ≤ 13,
−14 ≤ k ≤ 13,
−11 ≤ l ≤ 11
−20 ≤ h ≤ 20,
−17 ≤ k ≤ 17,
−17 ≤ l ≤ 17
−9 ≤ h ≤ 11,
−10 ≤ k ≤ 10,
−33 ≤ l ≤ 32
−10 ≤ h ≤ 8,
−11 ≤ k ≤ 11,
−30 ≤ l ≤ 34
Reflections collected 18,563 30,718 24,060 8980
Independent reflections 2327 [Rint = 0.0414,
Rsigma = 0.0185]
3126 [Rint = 0.0305,
Rsigma = 0.0149]
4294 [Rint = 0.0305,
Rsigma = 0.0194]
4524 [Rint = 0.0271,
Rsigma = 0.0313]
Data/restraints/parameters 2327/0/136 3126/0/150 4294/0/250 4524/0/250
Goodness-of-fit on F2 1.072 1.069 1.202 1.049
Final R indexes [I ≥ 2σ (I)] R1 = 0.0203,
wR2 = 0.0466
R1 = 0.0120,
wR2 = 0.0250
R1 = 0.0361,
wR2 = 0.0873
R1 = 0.0226,
wR2 = 0.0562
Final R indexes [all data] R1 = 0.0223,
wR2 = 0.0481
R1 = 0.0145,
wR2 = 0.0257
R1 = 0.0388,
wR2 = 0.0884
R1 = 0.0240,
wR2 = 0.0567
Largest diff. peak/hole/e Å−3 1.65/−0.92 0.61/−0.35 0.56/−0.45 0.31/−0.51
Table 2. Geometric parameters of H-bonds for 14.
Table 2. Geometric parameters of H-bonds for 14.
D-HAD-H [Å]HA [Å]DA [Å]HA [o]
1
C7-H7ACl15 i0.972.633.563(4)162
C10-H10ACl130.962.763.655(4)155
Symmetry code: (i) x, 3/2-y, ½ + z
2
C7-H7ACl16 i0.972.823.701(2)152
(i) 3/2-x, −1/2 + y, 3/2-z
3
C4A-H4ACl15 i0.932.773.615(3)152
C7B-H7BBCl13 ii0.972.833.724(3)154
C9B-H9BBCl16 iii0.962.823.688(3)151
C11B-H11DCl16 iv0.962.753.646(3)155
C11B-H11ECl16 iii0.962.773.654(3)153
Symmetry codes: (i) x, −1 + y, z; (ii) 1 + x, −1 + y, z; (iii) 1 + x, y, z; (iv) 3/2-x, −1/2 + y, ½-z
4
C9A-H9AABr13 i0.962.923.802(5)153
C4B-H4BBr15 ii0.932.863.693(5)149
C11A-H11ABr13 iii0.962.893.786(5)155
C11B-H11DBr14 iv0.962.913.838(5)164
Symmetry codes: (i) −1 + x, y, z; (ii) ½ + x, ½-y, 1-z; (iii) −1/2 + x, ½-y, 1-z; (iv) ½-x, 1-y, ½ + z
Table 3. Relevant bond lengths, bond angles, and dihedral angles in the compounds under study (see Figure 1 for atom notation, the subscript met is related to methyl groups).
Table 3. Relevant bond lengths, bond angles, and dihedral angles in the compounds under study (see Figure 1 for atom notation, the subscript met is related to methyl groups).
Compound(BTMA)2[PtCl6](BTMA)[AuCl4](BTMA)2[CuCl4](BTMA)2[CuBr4]BTMA+ in H2O
MethodX-rayX-rayX-rayX-rayMP2
Bond lengths [Å]
C1–C21.393(5)1.392(3)1.396(4)/1.392(4)1.391(6)/1.393(7)1.399
C2–C31.392(5)1.389(3)1.386(5)/1.389(4)1.388(7)/1.370(7)1.393
C3–C41.383(6)1.379(4)1.382(5)/1.381(5)1.379(8)/1.379(7)1.395
C4–C51.383(6)1.380(4)1.384(4)/1.386(6)1.388(7)/1.379(7)1.395
C5–C61.385(5)1.379(3)1.388(4)/1.390(4)1.377(7)/1.402(6)1.393
C6–C11.395(5)1.388(3)1.394(4)/1.384(4)1.391(7)/1.383(6)1.399
C1–C71.505(5)1.500(3)1.503(4)/1.502(4)1.497(6)/1.511(6)1.495
C7–N1.521(4)1.532(2)1.527(4)/1.530(4)1.528(5)/1.532(5)1.516
N–Cmet1.488(4)1.508(3)1.494(4)/1.499(4)1.499(5)/1.493(6)1.488
1.518(4)1.499(3)1.502(4)/1.498(4)1.507(6)/1.507(6)1.490
1.492(4)1.497(3)1.500(4)/1.500(4)1.504(5)/1.495(6)1.488
Bond angles [deg]
C1–C2–C3120.6(3)119.9(2)120.1(3)/120.3(3)120.5(5)/120.2(4)120.3
C2–C3–C4120.1(4)120.4(2)120.1(3)/120.2(3)120.2(4)/120.7(5)120.0
C3–C4–C5119.6(3)119.8(2)120.2(3)/119.9(3)119.5(5)/119.5(4)119.9
C4–C5–C6120.8(4)120.1(2)120.1(3)/119.8(3)120.4(5)/120.7(4)120.0
C5–C6–C1120.3(3)120.6(2)120.0(3)/120.7(3)120.7(4)/119.0(4)120.3
C6–C1–C2118.7(3)119.1(2)119.4(3)/119.1(3)118.6(4)/119.9(4)119.3
C2–C1–C7120.6(3)119.8(2)119.9(3)/120.5(3)120.7(4)/119.5(4)120.3
C6–C1–C7120.4(3)121.0(2)120.7(2)/120.4(3)120.6(5)/120.6(4)120.3
C1–C7–N115.3(2)114.7(2)114.0(2)/113.3(2)113.9(3)/113.9(3)113.9
C7–N–Cmet111.1(3)110.6(2)111.2(2)/109.2(2)111.5(3)/111.3(3)111.1
106.8(2)107.6(2)107.2(2)/107.6(2)107.7(3)/107.4(3)107.4
111.9(2)112.1(2)111.2(2)/111.0(2)110.7(3)/111.2(3)111.1
Dihedral angles [deg]
C1–C2–C3–C4−1.1(5)−1.0(4)−0.5(5)/0.5(5)0.8(7)/0.6(7)0.0
C2–C3–C4–C51.0(6)0.1(4)0.0(5)/−1.6(5)−1.9(7)/−1.5(7)0.0
C3–C4–C5–C6−0.3(5)0.2(4)−0.2(5)/0.7(5)0.4(7)/0.9(6)0.0
C4–C5–C6–C1−0.3(5)0.4(4)1.0(5)/1.3(5)2.3(7)/0.5(6)0.0
C6–C1–C2–C30.5(5)1.6(3)1.3(5)/1.5(5)1.9(7)/0.8(6)0.0
C3–C2–C1–C7174.8(3)178.6(2)179.0(3)/−177.1(3)−178.3(4)/179.8(4)179.7
C2–C1–C7–N90.7(4)89.8(3)92.0(3)/92.2(3)94.2(5)/91.6(5)90.7
C1–C7–N–Cmet71.6(3)63.0(2)62.5(3)/63.7(3)65.6(4)/68.9(5)61.0
−167.5(3)−177.7(2)−178.4(3)/−177.4(2)−175.6(3)/−172.3(4)180.0
−51.8(4)−57.8(2)−59.7(3)/−58.4(3)−56.2(5)/−53.8(5)−61.0
Table 4. RMSDs of the corresponding BTMA+ atomic positions in (C10H16N)2[PtCl6] and MP2 calculations [Å].
Table 4. RMSDs of the corresponding BTMA+ atomic positions in (C10H16N)2[PtCl6] and MP2 calculations [Å].
MP2 optimized in vacuum vs. X-ray structure0.131
MP2 optimized in solvent vs. X-ray structure0.130
MP2 optimized in vacuum vs. in solution0.006
Table 5. MP2 molecular orbital energies of HOMO, εHOMO, and LUMO, εLUMO, and their difference, Δε, in BTMA+ in various environments.
Table 5. MP2 molecular orbital energies of HOMO, εHOMO, and LUMO, εLUMO, and their difference, Δε, in BTMA+ in various environments.
EnvironmentVacuumAqueous Solution
εHOMO [eV]−12.66−9.28
εLUMO [eV]−0.703.01
Δε [eV]11.9612.28
Table 6. MP2 atomic charges, q, and volumes, V, of BTMA+ in aqueous solutions (see Figure 1 for atom notation, the subscript met is related to methyl groups).
Table 6. MP2 atomic charges, q, and volumes, V, of BTMA+ in aqueous solutions (see Figure 1 for atom notation, the subscript met is related to methyl groups).
qV [bohr3]
C1−0.0465
C2−0.0580
C3−0.0585
C4−0.0585
C5−0.0585
C6−0.0580
C70.2452
N−0.9849
Cmet0.25 (3×)62 (3×)
H20.0746
H30.0747
H40.0747
H50.0747
H60.0746
H70.09 (2×)43 (2×)
Hmet0.08 (9×)44 (9×)
Table 7. Electron densities, ρBCP, their Laplacians, ∇2ρBCP, bond ellipticities, εBCP, at the BCPs and bond delocalization indices, DI, of BTMA+ in aqueous solution (see Figure 1 for atom notation, the subscript met is related to methyl groups).
Table 7. Electron densities, ρBCP, their Laplacians, ∇2ρBCP, bond ellipticities, εBCP, at the BCPs and bond delocalization indices, DI, of BTMA+ in aqueous solution (see Figure 1 for atom notation, the subscript met is related to methyl groups).
ρBCP [e/bohr3]2ρBCP [e/bohr5]εBCPDI
C1–C20.3163−1.02450.1941.134
C2–C30.3196−1.05430.1921.177
C3–C40.3188−1.05070.1871.173
C4–C50.3188−1.05070.1871.173
C5–C60.3196−1.05430.1921.177
C6–C10.3163−1.02450.1941.134
C1–C70.2660−0.73860.0190.867
C7–N0.2350−0.58420.0340.734
N–Cmet0.2472 (2×)−0.6567 (2×)0.010 (2×)0.759 (2×)
0.2464−0.65420.0030.756
C2–H20.2930−1.18070.0120.830
C3–H30.2940−1.19010.0120.843
C4–H40.2940−1.19020.0120.844
C5–H50.2940−1.19010.0120.843
C6–H60.2930−1.18070.0120.830
C7–H70.2945 (2×)−1.1864 (2×)0.019 (2×)0.790 (2×)
Cmet–Hmet0.295 (2×)−1.19 (9×)0.024 (9×)0.80 (2×)
0.294 (7×) 0.81 (7×)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bojarska, J.; Breza, M.; Jelemenska, I.; Madura, I.D.; Jafari, S.; Trzybiński, D.; Woźniak, K.; Mieczkowski, A. The Comparative Study of Four Hexachloroplatinate, Tetrachloroaurate, Tetrachlorocuprate, and Tetrabromocuprate Benzyltrimethylammonium Salts: Synthesis, Single-Crystal X-Ray Structures, Non-Classical Synthon Preference, Hirshfeld Surface Analysis, and Quantum Chemical Study. Crystals 2025, 15, 1051. https://doi.org/10.3390/cryst15121051

AMA Style

Bojarska J, Breza M, Jelemenska I, Madura ID, Jafari S, Trzybiński D, Woźniak K, Mieczkowski A. The Comparative Study of Four Hexachloroplatinate, Tetrachloroaurate, Tetrachlorocuprate, and Tetrabromocuprate Benzyltrimethylammonium Salts: Synthesis, Single-Crystal X-Ray Structures, Non-Classical Synthon Preference, Hirshfeld Surface Analysis, and Quantum Chemical Study. Crystals. 2025; 15(12):1051. https://doi.org/10.3390/cryst15121051

Chicago/Turabian Style

Bojarska, Joanna, Martin Breza, Ingrid Jelemenska, Izabela D. Madura, Sepideh Jafari, Damian Trzybiński, Krzysztof Woźniak, and Adam Mieczkowski. 2025. "The Comparative Study of Four Hexachloroplatinate, Tetrachloroaurate, Tetrachlorocuprate, and Tetrabromocuprate Benzyltrimethylammonium Salts: Synthesis, Single-Crystal X-Ray Structures, Non-Classical Synthon Preference, Hirshfeld Surface Analysis, and Quantum Chemical Study" Crystals 15, no. 12: 1051. https://doi.org/10.3390/cryst15121051

APA Style

Bojarska, J., Breza, M., Jelemenska, I., Madura, I. D., Jafari, S., Trzybiński, D., Woźniak, K., & Mieczkowski, A. (2025). The Comparative Study of Four Hexachloroplatinate, Tetrachloroaurate, Tetrachlorocuprate, and Tetrabromocuprate Benzyltrimethylammonium Salts: Synthesis, Single-Crystal X-Ray Structures, Non-Classical Synthon Preference, Hirshfeld Surface Analysis, and Quantum Chemical Study. Crystals, 15(12), 1051. https://doi.org/10.3390/cryst15121051

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop