Next Article in Journal
Heterojunctions Based on II-VI Compound Semiconductor One-Dimensional Nanostructures and Their Optoelectronic Applications
Next Article in Special Issue
Trinuclear Co(II) and Mononuclear Ni(II) Salamo-Type Bisoxime Coordination Compounds
Previous Article in Journal
Synchrotron Radiation Pair Distribution Function Analysis of Gels in Cements
Previous Article in Special Issue
What Can We Learn from the Crystal Structures of Metallacarboranes?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Binuclear Copper(I) Borohydride Complex Containing Bridging Bis(diphenylphosphino) Methane Ligands: Polymorphic Structures of [(µ2-dppm)2Cu22-BH4)2] Dichloromethane Solvate

1
A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences (INEOS RAS), 119991 Moscow, Russia
2
Inorganic Chemistry Department, Peoples’ Friendship University of Russia (RUDN University), 117198 Moscow, Russia
3
Institute of Problems of Chemical Physics, Russian Academy of Sciences (IPCP RAS), 142432 Moscow, Russia
4
Istituto di Chimica dei Composti Organometallici Consiglio Nazionale delle Ricerche (ICCOM CNR), 50019 Sesto Fiorentino, Italy
*
Author to whom correspondence should be addressed.
Crystals 2017, 7(10), 318; https://doi.org/10.3390/cryst7100318
Submission received: 18 September 2017 / Revised: 13 October 2017 / Accepted: 17 October 2017 / Published: 20 October 2017
(This article belongs to the Special Issue Crystal Structures of Boron Compounds)

Abstract

:
Bis(diphenylphosphino)methane copper(I) tetrahydroborate was synthesized by ligands exchange in bis(triphenylphosphine) copper(I) tetrahydroborate, and characterized by XRD, FTIR, NMR spectroscopy. According to XRD the title compound has dimeric structure, [(μ2-dppm)2Cu22-BH4)2], and crystallizes as CH2Cl2 solvate in two polymorphic forms (orthorhombic, 1, and monoclinic, 2) The details of molecular geometry and the crystal-packing pattern in polymorphs were studied. The rare Twisted Boat-Boat conformation of the core Cu2P4C2 cycle in 1 is found being more stable than Boat-Boat conformation in 2.

1. Introduction

The concept of cooperative catalytic effects [1] in multinuclear transition metal systems led to the broad development and extensive investigation of the chemistry of transition metal complexes, bearing “short-bite” ligands that are able to lock two or more metallocenters in close proximity [2,3,4,5,6,7]. Such compounds are of great interest due to their catalytic activity including the transformation of small molecules on metal centres [8,9], they can also be used as synthetic models of enzyme action [10,11,12].
Phosphines are ubiquitous ligands in transition metal chemistry. Among various types of diphosphine ligands, bis-(diphenylphosphino)methane (dppm) is one of the very efficient bridging ligands [2]. As other diphosphine ligands it is able to chelate metals, but rarely acts as a bidentate ligand (η2-dppm), forming a strained four-membered cycle (Scheme 1) [13,14,15,16,17]. Rather, it has a tendency to act as either a monodentate (η1-dppm) or bridging bidentate ligand (μ2-dppm) [18]. Many examples of binuclear complexes containing the eight-membered ring M(μ2-dppm)2M' are known with a variety of metals and stereochemistries [18].
The Cu(I)-dppm complexes are emerging class of polynuclear complexes, that are drawing considerable attention because of their photophysical properties [19,20,21,22] and prospective use as a catalyst [23,24,25] and a sensor for various organic bases [26] and anions [27]. Binuclear Cu(I) species possess an enhanced reactivity toward organic azides in copper-catalysed azide-alkyne cycloaddition compared to monomeric copper complexes [28,29,30,31,32,33,34,35]. Copper(I) tetrahydroborates with phosphine ligands featuring relative stability to air oxygen and moisture are used as selective reducing agents [36,37,38,39,40], catalysts of photosensitized isomerization of dienes [41,42,43] and hydrolytic dehydrogenation of ammonia borane [44]. Since metal tetrahydroborates have great potential in hydrogen storage technology [45,46,47,48,49,50], as catalysts [51,52,53,54,55] and selective reducing agents [56,57,58,59,60,61] their structural and dynamic properties have been actively investigated [52,62,63,64]. These studies revealed different modes of BH4 coordination to the metal atom, which can behave as mono-, bi-, or tridentate ligand [64].
Our studies of intermolecular interactions of BH4 [65,66] and several metal tetrahydroborates [67,68,69,70] with proton donors have shown the versatility of dihydrogen bonded (DHB) complexes formed and their crucial role in the reactivity of these compounds. In particular, we have shown that the formation of bifurcate DHB complexes involving both bridging and terminal hydride hydrogens of (Ph3P)2Cu(η2-BH4) (Scheme 2) is prerequisite for the subsequent proton transfer and dimerization to occur [67]. Continuing these studies, we attempted the synthesis of (η2-dppm)Cu(η2-BH4) following the published recipe [71]. However, in our hands, it gave, instead, a binuclear dimer bearing two bridging μ2-dppm ligands between the two {Cu(η2-BH4)} moieties. Herein we describe its spectroscopic characterization and analysis of polymorphic structures of its dichloromethane solvate.

2. Experimental Section

All manipulations were performed under a dry argon atmosphere using the standard Schlenk technique. Commercially-available argon (99.9%) was additionally purified from traces of oxygen and moisture by sequential passage through Ni/Cr catalyst column and 4 Å molecular sieves.
The HPLC grade solvents (Acros Organics, Morris Plains, NJ, United States) were used for sample preparation after additional purification by standard procedures. Dichloromethane (DCM) and toluene were dehydrated over CaH2 and Na/benzophenone, respectively. All solvents were freshly distilled under argon prior to use. Deuterated solvent (CD2Cl2) was dehydrated over CaH2 and was distilled and degassed by three freeze-pump-thaw cycles prior to use. Bis(diphenylphosphino)methane (dppm) from Sigma Aldrich (St. Louis, MO, USA) was used without additional purification. Bis(triphenylphosphine) copper(I) tetrahydroborate was prepared following the previously-described procedure [67].
IR spectra were recorded on Shimadzu IR Prestige21 (Shimadzu Corporation, Kyoto, Japan) and Nicolet 6700 FTIR (Thermo Fisher Scientific, Waltham, MA, USA) spectrometers in KBr pellets and Nujol mull in thin polyethylene film. NMR spectra were recorded on a Bruker Avance II 500 and 600 MHz spectrometers (Bruker Corporation, Billerica, MA, USA). 1H and 13C{1H} chemical shifts are reported in parts per million (ppm) downfield to tetramethylsilane (TMS) and were calibrated against the residual solvent resonance, while 31P{1H} spectra were referenced to 85% H3PO4 with a downfield shift taken as positive; 11B spectra were referenced to BF3⋅Et2O. The 13C{1H} NMR spectra were registered using the JMODECHO mode; the signals for the C atoms bearing odd and even numbers of protons have opposite polarities.

2.1. Preparation of µ2-Bis(Diphenylphosphino)Methane Copper(I) Tetrahydroborate [(μ2-dppm)2Cu2][η2-BH4]2

The complex was synthesized through a slight modification of a previously described procedure [71]. Bis(diphenylphosphino)methane (dppm) (0.5 g, 1.32 mmol) were added to a solution of bis(triphenylphosphine) copper(I) tetrahydroborate (0.8 g, 1.32 mmol) in 50 ml toluene. The reaction mixture was stirred for 3 h at 60 °C, then cooled to room temperatures and refrigerated (−15 °C) to afford the white powder precipitate (0.3 g) of pure µ2-bis(diphenylphosphino)methane copper(I) tetrahydroborate (yield: 48%). The monocrystals suitable for XRD analysis were obtained by slow solvent evaporation from CH2Cl2 (DCM) solution under an argon stream.
1H NMR (500 MHz, CD2Cl2, 298 K, ppm): 1.26 (br d, BH4), 2.88 (br q, CH2), 7.11 (t, meta Ph), 7.26 (t, para Ph), 7.33 (multiplet, ortho Ph). 31P{1H} NMR (202 MHz, CD2Cl2, 298 K, ppm): −14.6 ÷ −16.5 (s) depending on conc. 11B NMR (160 MHz, CD2Cl2, 298 K, ppm): −29.81 (broad multiplet). 13C NMR (126 MHz, CD2Cl2, 298 K, ppm) 25.74 (multiplet CH2), 129.93 (s) para Ph, 128.52 (s) meta Ph, 132.58 multiplet orto Ph, 132.99 multiplet ipso C Ph.
FTIR: 3075, 3049, 2382, 2360, 2294, 2249, 2019, 1967, 1934, 1484, 1433, 1384, 1368, 1331, 1312, 1278, 1187, 1158, 1133, 1095, 1025, 999, 918, 848, 777, 766, 741, 734, 719, 693, 516, 507, 477 cm−1 (KBr pellet); 521, 516, 507, 477, 430, 420, 412, 358 cm−1 (Nujol mull/polyethylene film).
Several attempts to obtain the pure complex by recrystallization from toluene gave samples containing the traces of this solvent. The satisfactory elemental analysis was obtained for the sample which contains according to 1H NMR approximately 0.7 molecules of toluene per one molecule of the copper dimer. Anal. calcd. for C50H52B2Cu2P4: C, 64.88; H, 5.66; B, 2.34; Cu, 13.73; P, 13.39. Found: C, 66.55; H, 5.83; B, 2.29; P, 12.71.

2.2. Computational Details

Full geometry optimization of 1 and 2 (with removed solvent molecules) was performed with the Gaussian09 (Revision D.01, Gaussian, Wallingford, CT, USA) [72] software package. The model was described by M06 [73], B3LYP [74,75,76], BP86 [77], and PBE0 [78] methods with spin-state-corrected s6-31G(d) [79] basis set for Cu atom and 6-311++G(d,p) for atoms of the BH4 and alcohol OH-groups [80,81]; 6-31G(d) for the phosphorus atoms [82]; and 6-31G for the carbon and hydrogen atoms of dppm ligand [80,83,84]. For B3LYP, BP86 and PBE0 functionals empirical dispersion correction suggested by Grimme (GD2 [85] and GD3BJ [86,87]) was applied. Frequency calculations were performed for all optimized complexes in the gas phase and are reported without the use of scaling factors. The nature of all the stationary points on the potential energy surfaces was confirmed by an absence of any imaginary frequencies in the vibrational analysis [88].
The inclusion of nonspecific solvent effects in the calculations was performed by using the SMD method [89]. The solute cavity was redefined with radii = UAHF, because this atomic cavity was found to be more suitable than the default atom cavity (radii = SMD-Coulomb) defined in the SMD model [70,90]. The interaction energies were calculated in CH2Cl2 (ε = 8.9) for the gas phase optimized geometries. Changes in Gibbs energies and enthalpies in the solvent were determined using corresponding corrections obtained for the gas phase [91]:
ΔHSolv. = ΔESolv. + ΔHcorrgas
ΔGSolv. = ΔESolv. + ΔGcorrgas

2.3. X-ray Crystallography

X-ray diffraction data were collected on an Bruker APEX II CCD diffractometer (Bruker Corporation, Billerica, MA, United States) using molybdenum radiation [λ(MoKα) = 0.71072 Å, ω-scans] for 1 and 2. The substantial redundancy in data allowed an empirical absorption correction to be applied with SADABS by multiple measurements of equivalent reflections. The structures were solved by direct methods and refined by the full-matrix least-squares technique against F2 in the anisotropic-isotropic approximation.
The positional and anisotropic displacement parameters of the disordered CH2Cl2 in 1 and 2 were refined with the constraints on the C–Cl bond length (DFIX) and anisotropic displacement parameters (EADP). C–H hydrogen atoms in all structures were placed in calculated positions and refined within the riding model. Hydrogen atoms of BH4 group in both structures were located from the Fourier density synthesis and refined in the riding model. All calculations were performed with the SHELXTL software package [92]. Crystal data and structure refinement parameters are listed in Table 1. Crystallographic data for the structures reported in this paper have been deposited to the Cambridge Crystallographic Data Centre as supplementary no.: CCDC-1572389 (for 1) and CCDC-1572388 (for 2). These data can be obtained free of charge from Cambridge Crystallographic Data via www.ccdc.cam.ac.uk/data_request/cif.

3. Results and Discussion

3.1. Experimental Characterization

The title copper(I) tetrahydroborate was synthesized by the ligand exchange in a manner similar to that given in [71], where of the reaction product was described as (η2-dppm)Cu(η2-BH4). In our hands the procedure described in the Experimental Section yielded, upon recrystallization from CH2Cl2, a bimetallic complex with two bridging dppm ligands [(μ2-dppm)2Cu2][η2-BH4]2, the possible formation of such compound was suggested earlier in [93].
The 1H NMR spectrum (Figure S1) of the complex obtained exhibits the signals of phenyl protons around δPh = 7.26 ppm, methylene protons δCH2 = 2.88 ppm, and broad multiplet at δBH4 = 1.26 ppm belonging to borohydride protons. The 11B NMR spectrum (Figure S2) consists of a broad multiplet at −29.81 ppm, which is similar to the resonance of the monometallic (Ph3P)2Cu(η2-BH4) (δB = −29.79 ppm) [67]. The 31P{1H} spectrum (Figure S3) shows only one singlet at −14.6 ÷ −16.5 (s) ppm (depending on concentration), which means the phosphorus atoms are magnetically equivalent. The signal is deshielded compared to the free dppm ligand (δP = −21.7); at the same time its position is significantly different from that reported for monometallic dppm compound [71] (−148 ppm relative to (MeO)3P as a reference which is 140 ppm relative to 85% H3PO4 used in this work). In the 13C{1H} NMR (see Figures S4–S7) hydrogen and carbon atoms of methylene bridge (−25.9 ppm) and carbon atoms in phenyl rings (ortho and ipso ones) give centrosymmetric multiplets (see Figures S5 and S6). For the sake of comparison the 31P{1H} and 1H NMR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 were also measured in CDCl3 (Figure S10).
FTIR spectra in the KBr pellets of 1 and 2 (Figures S8 and S9) show two BHterm at 2382 and 2360 cm−1 and two BHbr stretching vibrations at 2019 and 1967 cm−1, BH2 deformation at 1133 cm−1 (Table 2) and a band at 358 cm−1, which can be attributed to the vibrations of the four-membered CuHBH cycle (νCuB) [67,94]. The positions of these bands are within the range reported for bis-phosphine {Cu(η2-BH4)} complexes (Table 2). Moreover, they coincide with those reported previously for the analogue dppm compound formulated as (dppm)Cu(η2-BH4) [94].
The spectral criteria allow determining the coordination mode: the hapticity of the BH4 ligand. IR spectra of {M(η1-BH4)} complexes show only one BHbr stretching vibration instead of two BHbr stretching vibrations observed for {M(η2-BH4)} complexes. The latter also exhibits two resolved or one broad BHterm stretching vibrations at a higher frequency than {M(η1-BH4)} (Table 2). Additionally, stretching vibration of CuHBH cycle is a unique feature of {M(η2-BH4)} complexes [52,63]. Analysis of the data for copper(I) tetrahydroborate complexes shows that the 11B NMR chemical shift of the BH4 group is slightly different for different coordination types (−29.1 ÷ −30.2 ppm for {Cu(η2-BH4)} and −30.2 ÷ −40.0 ppm for {Cu(η1-BH4)}). Thus, the spectral analysis can serve as a base for the initial assignment of the BH4 coordination mode. The X-ray data on the Cu∙∙∙B distance (vide infra) should allow to unambiguously distinguish the type of BH4 coordination even if the position of hydrogens could not be accurately determined [63].
The XRD analysis of monocrystals obtained for this copper(I) tetrahydroborate compound revealed it is a binuclear complex bearing two dppm ligands bridging two {Cu(η2-BH4)} fragments. Previously it was found that the addition of an excess anion able to act as a capping ligand (e.g., halogen anions) can yield not only binuclear, but tri-, or even tetranuclear Cu(I)-dppm complexes. However in our case, BH4 does not act as a capping ligand, and the trinuclear structure was found previously only once for (μ2-PPh2NHPPh2)3Cu33-H)-(μ3-BH4) [105].
Two solvatomorphic structures were identified (Table 1, Figure 1 and Figure 2): one of orthorhombic space group P212121 with one DCM molecule (1), and the second one of monoclinic space group P21/c with ½ molecule of DCM per molecule of the copper complex (2) (Figure S11). In both structures the solvent molecules are disordered in a 1:1 ratio across a crystallographic inversion centre. The copper atoms have distorted tetrahedral geometry, being ligated with two phosphorus atoms of dppm ligands and two hydrogen atoms from tetrahydroborate; the selected bond distances and angles are presented in Table 3 (for additional details see Tables S2 and S4). Copper atoms and ligands form eight-member cycles Cu2P4C2 that have Twisted Boat-Boat conformation in 1 and Boat-Boat conformation in 2 (Figure S11). The Cu(1)–Cu(2) distance is 3.2035(4) Å for 1 and 3.392(1) Å for 2, which is above the sum of van der Waals radii for copper (2.8 Å), and is within the range (2.679–4.797 Å) determined for eight-member Cu2P4C2 cycles (Table S1).
The DFT calculations (see below) of 2 revealed its possible structural instability, during the optimization the conformation changes from Boat-Boat to Twist Boat-Boat.
The non-covalent interactions apparently play an important role in the stabilization of both structures. In both crystals the π-π stacking interaction between the pairs of phenyl rings of dppm ligands is suggested by short inter-ring distance (3.723 Å for 1 and 3.888 Å for 2). The analysis of molecular packing of 1 (Figure 2a) reveals four short contacts between the C–H of phenyl rings of dppm ligands and the chlorine atom of DCM (C–H···Cl) per unit cell, which can be considered as weak hydrogen bonds. The angles (C–H···Cl) for these interactions vary from 145.9 to 149.6° and H···Cl distances are in the range 2.659–2.679 Å that is less than the sum of van der Waals radii for these two atoms (2.95 Å). There is also a short B–H···Cl–C distance 2.755 Å between BH and DCM with angle C–Cl···H(B) = 158.2° that resembles a halogen bonding [106,107] and was referred to as a hydride-halogen bond [108,109,110,111,112,113]. This interaction has a donor-acceptor nature, where B–Hδ− acts as a donor of electron density and interacts with an electron deficient area (σ-hole) located on the halogen atom Hal–R.
The [(μ2-dppm)2Cu2][η2-BH4]2 molecules in 2 are connected with each other and with the DCM molecules via hydrogen bonds [r(H···Cl) = 3.031 Å, C–H···Cl = 149.2°] and dihydrogen bonds [r(H···H) = 2.246 Å, C–H···H(B) = 168.2°] leading to the formation of a two-dimensional network (Figure 2b).
As mentioned above, the borohydride ligand is coordinated to copper via two hydrogen atoms, the distances Cu∙∙∙B (2.190–2.198 Å) are typical for structures of {Cu(η2-BH4)} complexes (according to previously suggested structural criteria) [63] found in CCDC (Table S6), but are slightly shorter if compared with the value 2.212 Å found for (PPh3)2Cu(η2-BH4) [67]. The {Cu(η1-BH4)} complexes are characterized by longer Cu∙∙∙B distances of 2.441–2.499 Å.

3.2. CCDC Analysis

The CCDC search for the structures containing eight-membered [(μ2-dppm)2Cu2]2+ moieties gave 110 entries, but none of them bears a BH4 ligand. The crystal structures found are gathered in Table S1, subdivided according to their conformation type (Figure 3) and arranged in ascending order of Cu(1)···Cu(2) distances. Eight of the structures found are characterized by strong copper-copper interaction, the Cu(1)···Cu(2) distances being less than the sum of van der Waals radii for two Cu atoms 2.679–2.789 Å. For each conformation (Figure 3) we detected the most representative structure among the entries found and the boundary conditions. According to the boundary conditions for basic conformation types (Boat-Boat, Boat-Chair, Chair-Chair and Plain) the difference between two angles PCuP and between two dihedral angles χ(P,Cu,Cu,P) should be less or equal 5°. The Chair-Chair conformation is characterized by two straight dihedral angle χ(P,Cu,Cu,P). Twist-type conformations (Twisted Boat-Boat and Twisted Boat-Chair) should have one dihedral angle χ(P,Cu,Cu,P) close to 90° and another obtuse dihedral angle χ(P,Cu,Cu,P) (>117°). Remaining structures that have geometrical parameters between basic and twisted conformation types were named as distorted conformations (Distorted Boat-Boat and Distorted Boat-Chair). All conformational data are summarized in Table 4 and Table S1.
The analysis reveals that Boat-Boat and Distorted Boat-Boat conformations account for 48% of all found structures (Figure 4). Other conformation types are Chair-Chair–19%, Boat-Chair–14% and Twisted Boat-Boat, comprising 10% of conformations. The rest of the conformation types (Distorted Boat-Chair, Twisted Boat-Chair and Plain) account for 5% and less.
The majority of all three Boat-Boat structures (basic, distorted, and twisted) contains a bridging ligand (μ-R2S; μ-R2CO; RPy-O; μ-NO3; μ-RCOO) and is characterized by rather short Cu(1)···Cu(2) distances (2.679–3.852 Å). The Chair-Chair and Plain conformations feature the longest Cu(1)···Cu(2) distance (3.359–4.797 Å) because these complexes contain chelating or strongly-coordinating ligands, such as bipyridine, pyrazine, phenantroline, and their derivatives. Despite the difference in the solid state conformations, the 31P NMR chemical shifts of dppm ligand in eight-membered [(μ2-dppm)2Cu2]2+ moieties fall in the same, rather broad, range that does not allow discriminating of the conformation types on the basis of 31P NMR data in solution.

3.3. DFT Calculations

Since our attempt to synthesize the monomeric compound (η2-dppm)Cu(η2-BH4) (Figure 5a), previously described in [71] has failed, we attempted to optimize this structure by different DFT methods. However, these attempts were unsuccessful; instead of the proposed structure (Figure 5a), they gave the (η1-dppm)Cu(η2-BH4) complex stabilized by copper interaction with a phenyl ring (Figure 5b). The formation of the [(μ2-dppm)2Cu2][η2-BH4]2 dimer from two molecules of monomeric (η1-dppm)Cu(η2-BH4) is energetically favourable, ∆GDCM°form being −19.3 kcal/mol (M06) and −24.0 kcal/mol (B3LYP-D2) (Table S5). This is in agreement with only a few examples of Cu(I) complexes reported in which dppm acts as a chelate ligand (η2-dppm, Scheme 1) [114]; the dinuclear Cu(I) complexes with two bridging dppm ligands are by far more common [115].
The geometry optimizations by M06 and B3LYP-D2 methods (Table S7) reproduced quite well the X-ray determined geometry 1 of the binuclear copper complex; the difference between the calculated and experimentally determined Cu(1)∙∙∙Cu(2) distances (0.006 Å for M06 and 0.155 Å for B3LYP-D2) is the lowest among other DFT functionals used (Tables S2 and S3). The difference between the calculated and experimentally observed Cu–P bonds lengths is also rather small, 0.039–0.020 Å (M06) and less than 0.020 Å for B3LYP-D2. The difference in experimental and theoretical CuH bond length and Cu∙∙∙B distances is 0.008–0.101 Å (M06), 0.004–0.089 Å (B3LYP-D2), 0.001–0.005 Å (M06) and 0.011–0.023 Å (B3LYP-D2), respectively. The analogous performance for M06 and B3LYP-D2 methods was previously observed for calculations of the Cu(II)-silsesquioxane core [116]. When the optimization was attempted for the geometry of structure 2, it led to the conformation changes converting from Boat-Boat to Twisted Boat-Boat. No local minimum was found for the Boat-Boat conformation type.
The simulated IR spectra of 1 (Twisted Boat-Boat) optimized by M06 and B3LYP-D2 methods are in line with the experimental IR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 in KBr pellets (Figure 6, Table 5). The M06 and B3LYP-D2-optimized (η1-dppm)Cu(η2-BH4) monomer gives similar positions of IR-active stretching vibrations but has a significant difference in the relative IR intensity of BHterm stretching vibrations.

4. Summary and Conclusions

The XRD analysis of monocrystals revealed the first example of the bimetallic complex [(μ2-dppm)2Cu2][η2-BH4]2 bearing two dppm ligands bridging two {Cu(η2-BH4)} fragments. Two solvatomorphic structures were identified: one of orthorhombic space group P212121 with one DCM molecule 1 and the second one of monoclinic space group P21/c with ½ molecule of DCM per molecule of complex 2. The former structure possesses the twisted boat-boat conformation, which is rather rare for eight-membered [(μ2-dppm)2Cu2]2+ moieties. Analysis of the literature data revealed that, despite the difference in conformations, the 31P NMR chemical shift of dppm ligand in eight-membered [(μ2-dppm)2Cu2]2+ moieties does not enable identification of conformation type in solution. On the other hand, the 11B NMR and IR spectra could be used to discriminate between the η1 and η2-BH4 coordination modes. However, the final assignment should come from the XRD analysis.
The DFT calculations by M06 and B3LYP-D2 methods reproduced, quite well, the geometry of 1 and observed experimental IR spectra. Optimization of 2 revealed structural instability during the optimization conformation changes from Boat-Boat to Twisted Boat-Boat (Table S4). This finding is surprising because Boat-Boat and Distorted Boat-Boat conformations account, together, for 48% of the reported CCSD structures bearing eight-membered [(μ2-dppm)2Cu2]2+ fragments, whereas the Twisted Boat-Boat conformation is revealed only for 10% of compounds.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4352/7/10/318/s1. Figure S1. 1H NMR spectra (500 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2. Figure S2. 11B{1H} NMR spectra (160 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2. Figure S3. 31P{1H} NMR spectra (202 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2. Figure S4. 13C{1H} NMR spectra (126 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2. Figure S5. 13C{1H} NMR spectra (126 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2 (16,850–16,600 Hz). Figure S6. 13C{1H} NMR spectra (126 MHz, CD2Cl2, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2 (3290–3180 Hz). Figure S7. 13C{1H} NMR spectra (126 MHz, CD2Cl2, 298 K, ppm) in JMODECHO mode of [(μ2-dppm)2Cu2][η2-BH4]2. Figure S8. FTIR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 in KBr pellet. Figure S9. FTIR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 in Nujol mull/thin polyethylene film. Figure S10. 31P{1H} (202 MHz, 298 K, ppm) and 1H NMR spectra (500 MHz, 298 K, ppm) of [(μ2-dppm)2Cu2][η2-BH4]2 in CDCl3. Figure S11. General view of molecular structures of 1 and 2 conformations. The solvents molecules are omitted for clarity. Table S1. CCDC analysis of the structures, containing eight-membered [(μ2-dppm)2Cu2]2+ moieties. Table S2. Structural parameters of crystal structure 1 (Twisted Boat-Boat) and optimized structures. Table S3. The differences between structural parameters of crystal structure 1 (Twisted Boat-Boat) and optimized structures. Table S4. Structural parameters of crystal structure 2 (Boat-Boat) and optimized structures. Table S5. Energy of formation of DFT-optimized geometries binuclear complexes (1 and 2) computed relative monomer complexes. Table S6. CCDC analysis of the structures, containing {Cu(BH4)} moieties. Table S7. DFT-optimised geometries (Cartesian coordinates) and electronic energies.

Acknowledgments

Some parts of our work were financially supported by the Ministry of Education and Science of the Russian Federation (the Agreement number 02.a03.21.0008), the Russian-Italian bilateral project CNR-RFBR no. 15-53-78027 and RFBR project no. 16-03-00324. The X-ray diffraction and NMR spectroscopic data were obtained using the equipment of Centre for Molecule Composition Studies of INEOS RAS.

Author Contributions

Evgenii I. Gutsul and Andrea Rossin carried out the synthesis. Konstantin A. Lyssenko performed the XRD experiments. Alexander S. Peregudov measured NMR experiments. Evgenii I. Gutsul conducted the FTIR investigation. Viktor D. Makhaev contributed reagents. Lina M. Epstein analysed the literature data. Igor E. Golub and Oleg A. Filippov performed the DFT calculations. Natalia V. Belkova and Igor E. Golub analysed the data and wrote the paper. Maurizio Peruzzini and Elena S. Shubina conceived experiments and supervised the work.

Conflicts of Interest

The author declare no conflict of interest.

References

  1. Niedner-Schatteburg, G. Cooperative effects in clusters and oligonuclear complexes of transition metals in isolation. In Clusters—Contemporary Insight in Structure and Bonding; Dehnen, S., Ed.; Springer International Publishing: Cham, Switzerland, 2017; pp. 1–40. [Google Scholar]
  2. Chaudret, B.; Delavaux, B.; Poilblanc, R. Bisdiphenylphosphinomethane in dinuclear complexes. Coord. Chem. Rev. 1988, 86, 191–243. [Google Scholar] [CrossRef]
  3. Balakrishna, M.S.; Reddy, V.S.; Krishnamurthy, S.S.; Nixon, J.F.; Laurent, J.C.T.R.B.S. Coordination chemistry of diphosphinoamine and cyclodiphosphazane ligands. Coord. Chem. Rev. 1994, 129, 1–90. [Google Scholar] [CrossRef]
  4. Mague, J.T. “Short-bite” ligands in cluster synthesis. J. Cluster Sci. 1995, 6, 217–269. [Google Scholar] [CrossRef]
  5. Maggini, S. Classification of P,N-binucleating ligands for hetero- and homobimetallic complexes. Coord. Chem. Rev. 2009, 253, 1793–1832. [Google Scholar] [CrossRef]
  6. Naik, S.; Mague, J.T.; Balakrishna, M.S. Short-bite PNP ligand-supported rare tetranuclear [Cu4I4] clusters: Structural and photoluminescence studies. Inorg. Chem. 2014, 53, 3864–3873. [Google Scholar] [CrossRef] [PubMed]
  7. Fliedel, C.; Ghisolfi, A.; Braunstein, P. Functional short-bite ligands: Synthesis, coordination chemistry, and applications of N-functionalized bis(diaryl/dialkylphosphino)amine-type ligands. Chem. Rev. 2016, 116, 9237–9304. [Google Scholar] [CrossRef] [PubMed]
  8. Muetterties, E.L. Metal clusters in catalysis III.—Clusters as models for chemisorption processes and heterogeneous catalysis. Bull. Soc. Chim. Belg. 1975, 84, 959–986. [Google Scholar] [CrossRef]
  9. Muetterties, E.L.; Krause, M.J. Catalysis by molecular metal clusters. Angew. Int. Ed. 1983, 22, 135–148. [Google Scholar] [CrossRef]
  10. Adam, F.I.; Hogarth, G.; Richards, I. Models of the iron-only hydrogenase: Reactions of [Fe2(CO)6(μ-pdt)] with small bite-angle diphosphines yielding bridge and chelate diphosphine complexes [Fe2(CO)4(diphosphine)(μ-pdt)]. J. Organomet. Chem. 2007, 692, 3957–3968. [Google Scholar] [CrossRef]
  11. Adam, F.I.; Hogarth, G.; Kabir, S.E.; Richards, I. Models of the iron-only hydrogenase: Synthesis and protonation of bridge and chelate complexes [Fe2(CO)4{Ph2P(CH2)nPPh2}(μ-pdt)] (n = 2–4)—Evidence for a terminal hydride intermediate. C. R. Chim. 2008, 11, 890–905. [Google Scholar] [CrossRef]
  12. Yen, T.-H.; Chu, K.-T.; Chiu, W.-W.; Chien, Y.-C.; Lee, G.-H.; Chiang, M.-H. Synthesis and characterization of the diiron biomimics bearing phosphine borane for hydrogen formation. Polyhedron 2013, 64, 247–254. [Google Scholar] [CrossRef]
  13. Isaacs, E.E.; Graham, W.A.G. Phosphorus-31 nuclear magnetic resonance spectra of fac and mer isomers of tricarbonylbis[bis(diphenylphosphino)methane]molybdenum. Inorg. Chem. 1975, 14, 2560–2561. [Google Scholar] [CrossRef]
  14. Cotton, F.A.; Matusz, M. Synthesis, molecular structure and spectroscopic characterization of Mo(CO)2I22-dppm)(η1-dppm). Polyhedron 1987, 6, 261–267. [Google Scholar] [CrossRef]
  15. Gao, Y.; Holah, D.G.; Hughes, A.N.; Spivak, G.J.; Havighurst, M.D.; Magnuson, V.R.; Polyakov, V. Reactions of FeIII with LiAlH4 and LiBH4 in the presence of bis(diphenylphosphino)methane (dppm) and CO. The crystal and molecular structures of trans-[Fe(Cl)(CO)(η2-dppm)2][FeCl4], trans-Fe(H)22-dppm)2 and cis-[Fe(η2-S2CH(η2-dppm)2]BF4. Polyhedron 1997, 16, 2797–2807. [Google Scholar] [CrossRef]
  16. Gandhi, T.; Nethaji, M.; Jagirdar, B.R. Novel double insertion of carbon disulfide into two Ru−H bonds of [(dppm)2Ru(H)2] (dppm = Ph2PCH2PPh2): Synthesis and crystal structure of a methanedithiolate complex. Inorg. Chem. 2003, 42, 667–669. [Google Scholar] [CrossRef] [PubMed]
  17. Wilton-Ely, J.D.E.T.; Solanki, D.; Hogarth, G. Synthesis and reactivity of the Ruthenium(II) Dithiocarbonate complex [Ru(κ2-S2CO)(dppm)2] (dppm = bis(diphenylphosphino)methane). Inorg. Chem. 2006, 45, 5210–5214. [Google Scholar] [CrossRef] [PubMed]
  18. Puddephatt, R.J. Chemistry of bis(diphenylphosphino)methane. Chem. Soc. Rev. 1983, 12, 99–127. [Google Scholar] [CrossRef]
  19. Bera, J.K.; Nethaji, M.; Samuelson, A.G. Anion-controlled nuclearity and metal−metal distances in copper(I)−dppm complexes (dppm = Bis(diphenylphosphino)methane). Inorg. Chem. 1999, 38, 218–228. [Google Scholar] [CrossRef]
  20. Fan, W.-W.; Li, Z.-F.; Li, J.-B.; Yang, Y.-P.; Yuan, Y.; Tang, H.-Q.; Gao, L.-X.; Jin, Q.-H.; Zhang, Z.-W.; Zhang, C.-L. Synthesis, structure, terahertz spectroscopy and luminescent properties of copper (I) complexes with bis(diphenylphosphino)methane and N-donor ligands. J. Mol. Struct. 2015, 1099, 351–358. [Google Scholar] [CrossRef]
  21. El Sayed Moussa, M.; Evariste, S.; Wong, H.L.; Le Bras, L.; Roiland, C.; Le Polles, L.; Le Guennic, B.; Costuas, K.; Yam, V.W.W.; Lescop, C. A solid state highly emissive Cu(I) metallacycle: Promotion of cuprophilic interactions at the excited states. Chem. Commun. 2016, 52, 11370–11373. [Google Scholar] [CrossRef] [PubMed]
  22. He, L.-H.; Luo, Y.-S.; Di, B.-S.; Chen, J.-L.; Ho, C.-L.; Wen, H.-R.; Liu, S.-J.; Wang, J.-Y.; Wong, W.-Y. Luminescent three- and four-coordinate dinuclear copper(i) complexes triply bridged by bis(diphenylphosphino)methane and functionalized 3-(2-Pyridyl)-1,2,4-triazole ligands. Inorg. Chem. 2017, 56, 10311–10324. [Google Scholar] [CrossRef] [PubMed]
  23. Motokura, K.; Kashiwame, D.; Takahashi, N.; Miyaji, A.; Baba, T. Highly active and selective catalysis of copper diphosphine complexes for the transformation of carbon dioxide into silyl formate. Chem. Eur. J. 2013, 19, 10030–10037. [Google Scholar] [CrossRef] [PubMed]
  24. Nakamae, K.; Kure, B.; Nakajima, T.; Ura, Y.; Tanase, T. Facile insertion of carbon dioxide into Cu2(μ-H) dinuclear units supported by tetraphosphine ligands. Chem. Asian J. 2014, 9, 3106–3110. [Google Scholar] [CrossRef] [PubMed]
  25. Nakamae, K.; Tanaka, M.; Kure, B.; Nakajima, T.; Ura, Y.; Tanase, T. A fluxional Cu8H6 cluster supported by bis(diphenylphosphino)methane and its facile reaction with CO2. Chem. Eur. J. 2017, 23, 9457–9461. [Google Scholar] [CrossRef] [PubMed]
  26. Shiu, L.-C.; Liu, S.-A.; Wu, T.-S.; Shiu, K.-B. Synthesis and structures of d10–d10 M2(μ-dppm)2 complexes with sensitive metal–metal distances in response to the binding of sigma donors. J. Chin. Chem. Soc. 2016, 63, 1051–1055. [Google Scholar] [CrossRef]
  27. Bera, J.K.; Nethaji, M.; Samuelson, A.G. Synthesis and structures of oxyanion encapsulated copper(i)−dppm complexes (dppm = bis(diphenylphosphino)methane). Inorg. Chem. 1999, 38, 1725–1735. [Google Scholar] [CrossRef] [PubMed]
  28. Ahlquist, M.; Fokin, V.V. Enhanced reactivity of dinuclear copper(I) acetylides in dipolar cycloadditions. Organometallics 2007, 26, 4389–4391. [Google Scholar] [CrossRef]
  29. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide-alkyne cycloaddition (CuAAC) and beyond: New reactivity of copper(i) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef] [PubMed]
  30. Worrell, B.T.; Malik, J.A.; Fokin, V.V. Direct evidence of a dinuclear copper intermediate in Cu(I)-catalyzed azide-alkyne cycloadditions. Science 2013, 340, 457. [Google Scholar] [CrossRef] [PubMed]
  31. Jin, L.; Tolentino, D.R.; Melaimi, M.; Bertrand, G. Isolation of bis(copper) key intermediates in Cu-catalyzed azide-alkyne “click reaction”. Sci. Adv. 2015, 1, e1500304. [Google Scholar] [CrossRef] [PubMed]
  32. Singh, D.P.; Allam, B.K.; Singh, K.N.; Singh, V.P. Binuclear Cu(I) complex of (N1E,N2E)-N1,N2-bis(phenyl(pyridin-2-yl)methylene)oxalohydrazide: Synthesis, crystal structure and catalytic activity for the synthesis of 1,2,3-triazoles. J. Mol. Catal. A Chem. 2015, 398, 158–163. [Google Scholar] [CrossRef]
  33. Wang, C.; Ikhlef, D.; Kahlal, S.; Saillard, J.-Y.; Astruc, D. Metal-catalyzed azide-alkyne “click” reactions: Mechanistic overview and recent trends. Coord. Chem. Rev. 2016, 316, 1–20. [Google Scholar] [CrossRef]
  34. Zhu, L.; Brassard, C.J.; Zhang, X.; Guha, P.M.; Clark, R.J. On the mechanism of copper(i)-catalyzed azide–alkyne cycloaddition. Chem. Rec. 2016, 16, 1501–1517. [Google Scholar] [CrossRef] [PubMed]
  35. Ziegler, M.S.; Lakshmi, K.V.; Tilley, T.D. Dicopper Cu(I)Cu(I) and Cu(I)Cu(II) complexes in copper-catalyzed azide–alkyne cycloaddition. J. Am. Chem. Soc. 2017, 139, 5378–5386. [Google Scholar] [CrossRef] [PubMed]
  36. Fleet, G.W.J.; Harding, P.J.C. Convenient synthesis of bis(triphenylphosphine) copper(I) tetrahydroborate and reduction of acid chlorides to aldehydes. Tetrahedron Lett. 1979, 20, 975–978. [Google Scholar] [CrossRef]
  37. Fleet, G.W.J.; Harding, P.J.C.; Whitcombe, M.J. Bis(triphenylphosphine)copper (I) tetrahydroborate in the reduction of P-toluenesulphonylhydrazones and 2,4,6-triisopropylbenzenesulphonyl hydrazones (trisyl hdyrazones) to alkanes. Tetrahedron Lett. 1980, 21, 4031–4034. [Google Scholar] [CrossRef]
  38. Sorrell, T.N.; Pearlman, P.S. Preparation of aldehydes from acid chlorides using copper tetrahydroborate complexes. J. Org. Chem. 1980, 45, 3449–3451. [Google Scholar] [CrossRef]
  39. Fleet, G.W.J.; Harding, P.J.C. Selectivity in the acid catalysed reduction of carbonyl compounds to alcohols by bis(triphenylphosphine)copper(I) tetrahydroborate: Reduction of aldehydes in the presence of ketones. Tetrahedron Lett. 1981, 22, 675–678. [Google Scholar] [CrossRef]
  40. Bhanushali, M.J.; Nandurkar, N.S.; Bhor, M.D.; Bhanage, B.M. Direct reductive amination of carbonyl compounds using bis(triphenylphosphine) copper(I) tetrahydroborate. Tetrahedron Lett. 2007, 48, 1273–1276. [Google Scholar] [CrossRef]
  41. Grutsch, P.A.; Kutal, C. Use of copper(I) phosphine compounds to photosensitize the valence isomerization of norbornadiene. J. Am. Chem. Soc. 1977, 99, 6460–6463. [Google Scholar] [CrossRef]
  42. Grutsch, P.A.; Kutal, C. Photobehavior of copper(I) compounds. Role of copper(I)-phosphine compounds in the photosensitized valence isomerization of norbornadiene. J. Am. Chem. Soc. 1979, 101, 4228–4233. [Google Scholar] [CrossRef]
  43. Liaw, B.; Orchard, S.W.; Kutal, C. Photobehavior of copper(I) compounds. 4. Role of the triplet state of (arylphosphine)copper(I) complexes in the photosensitized isomerization of dienes. Inorg. Chem. 1988, 27, 1311–1316. [Google Scholar] [CrossRef]
  44. Hu, X.; Soleilhavoup, M.; Melaimi, M.; Chu, J.; Bertrand, G. Air-stable (CAAC)CuCl and (CAAC)CuBH4 complexes as catalysts for the hydrolytic dehydrogenation of BH3NH3. Angew. Chem. 2015, 127, 6106–6109. [Google Scholar] [CrossRef]
  45. Jepsen, L.H.; Ley, M.B.; Lee, Y.-S.; Cho, Y.W.; Dornheim, M.; Jensen, J.O.; Filinchuk, Y.; Jørgensen, J.E.; Besenbacher, F.; Jensen, T.R. Boron–nitrogen based hydrides and reactive composites for hydrogen storage. Mater. Today 2014, 17, 129–135. [Google Scholar] [CrossRef] [Green Version]
  46. Jepsen, L.H.; Ley, M.B.; Filinchuk, Y.; Besenbacher, F.; Jensen, T.R. Tailoring the properties of ammine metal borohydrides for solid-state hydrogen storage. ChemSusChem 2015, 8, 1452–1463. [Google Scholar] [CrossRef] [PubMed]
  47. Lai, Q.; Paskevicius, M.; Sheppard, D.A.; Buckley, C.E.; Thornton, A.W.; Hill, M.R.; Gu, Q.; Mao, J.; Huang, Z.; Liu, H.K.; et al. Hydrogen storage materials for mobile and stationary applications: Current state of the art. ChemSusChem 2015, 8, 2789–2825. [Google Scholar] [CrossRef] [PubMed]
  48. Dovgaliuk, I.; Filinchuk, Y. Aluminium complexes of B- and N-based hydrides: Synthesis, structures and hydrogen storage properties. Int. J. Hydrogen Energy 2016, 41, 15489–15504. [Google Scholar] [CrossRef]
  49. Mohtadi, R.; Orimo, S. The renaissance of hydrides as energy materials. Nat. Rev. Mater. 2016, 2, 16091. [Google Scholar] [CrossRef]
  50. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Cerny, R.; Ravnsbaek, D.B.; Filinchuk, Y.; Dornheim, M.; Besenbacher, F.; Jensen, T.R. Metal borohydrides and derivatives—Synthesis, structure and properties. Chem. Soc. Rev. 2017, 46, 1565–1634. [Google Scholar] [CrossRef] [PubMed]
  51. Mirviss, S.B.; Dougherty, H.W.; Looney, R.W. Catalyst of Transition Metal Borohydride Complexes. U.S. Patent 3,310,547, 21 March 1961. [Google Scholar]
  52. Marks, T.J.; Kolb, J.R. Covalent transition metal, lanthanide, and actinide tetrahydroborate complexes. Chem. Rev. 1977, 77, 263–293. [Google Scholar] [CrossRef]
  53. Langer, R.; Iron, M.A.; Konstantinovski, L.; Diskin-Posner, Y.; Leitus, G.; Ben-David, Y.; Milstein, D. Iron Borohydride pincer complexes for the efficient hydrogenation of ketones under mild, base-free conditions: Synthesis and mechanistic insight. Chem. Eur. J. 2012, 18, 7196–7209. [Google Scholar] [CrossRef] [PubMed]
  54. Chakraborty, S.; Lagaditis, P.O.; Förster, M.; Bielinski, E.A.; Hazari, N.; Holthausen, M.C.; Jones, W.D.; Schneider, S. Well-defined iron catalysts for the acceptorless reversible dehydrogenation-hydrogenation of alcohols and ketones. ACS Catal. 2014, 4, 3994–4003. [Google Scholar] [CrossRef]
  55. Zell, T.; Ben-David, Y.; Milstein, D. Highly efficient, general hydrogenation of aldehydes catalyzed by PNP iron pincer complexes. Catal. Sci. Technol. 2015, 5, 822–826. [Google Scholar] [CrossRef]
  56. Bhattacharyya, S.; Chatterjee, A.; Williamson, J.S. Reductive amination with zinc borohydride. efficient, safe route to fluorinated benzylamines. Synth. Commun. 1997, 27, 4265–4274. [Google Scholar] [CrossRef]
  57. Ravikumar, K.S.; Sinha, S.; Chandrasekaran, S. Diastereoselectivity in the reduction of acyclic carbonyl compounds with diisopropoxytitanium(iii) tetrahydroborate. J. Org. Chem. 1999, 64, 5841–5844. [Google Scholar] [CrossRef]
  58. Alinezhad, H.; Tajbakhsh, M.; Zamani, R. One-pot reductive amination of aldehydes and ketones using n-methyl-piperidine zinc borohydride (ZBNMPP) as a new reducing agent. Synlett 2006, 3, 431–0434. [Google Scholar] [CrossRef]
  59. Firouzabadi, H.; Iranpoor, N.; Alinezhad, H. Micellar media catalyzed highly efficient reductive amination of carbonyl compounds with bis(triphenylphosphine)(tetrahydroborato)zirconium(II), [Zr(BH4)2(Ph3P)2], as a new and a highly water tolerant tetrahydroborate reducing agent. J. Iran Chem. Soc. 2009, 6, 177–186. [Google Scholar] [CrossRef]
  60. Mohammadi, S.; Setamdideh, D. (Acridine)(tetrahydroborato) zinc complex [Zn(BH4)2(acr)]: A new stable and efficient reducing agent. Orient. J. Chem. 2015, 31, 2395–2399. [Google Scholar] [CrossRef]
  61. Abdollahpour, F.; Setamdideh, D. (Caffeine)(tetrahydroborato) zinc complex [Zn(BH4)2(caf)]: A new stable and efficient reducing agent. Orient. J. Chem. 2015, 31, 1787–1792. [Google Scholar] [CrossRef]
  62. Xu, Z.; Lin, Z. Transition metal tetrahydroborato complexes: An orbital interaction analysis of their structure and bonding. Coord. Chem. Rev. 1996, 156, 139–162. [Google Scholar] [CrossRef]
  63. Makhaev, V.D. Structural and dynamic properties of tetrahydroborate complexes. Russ. Chem. Rev. 2000, 69, 727. [Google Scholar] [CrossRef]
  64. Besora, M.; Lledós, A. Coordination modes and hydride exchange dynamics in transition metal tetrahydroborate. In Complexes Contemporary Metal Boron Chemistry I; Marder, T., Lin, Z., Eds.; Springer: Berlin, Germany, 2008; Volume 130, pp. 149–202. [Google Scholar]
  65. Filippov, O.A.; Filin, A.M.; Tsupreva, V.N.; Belkova, N.V.; Lledós, A.; Ujaque, G.; Epstein, L.M.; Shubina, E.S. Proton-transfer and H2-elimination reactions of main-group hydrides EH4 (E = B, Al, Ga) with alcohols. Inorg. Chem. 2006, 45, 3086–3096. [Google Scholar] [CrossRef] [PubMed]
  66. Golub, I.E.; Filippov, O.A.; Gulyaeva, E.S.; Gutsul, E.I.; Belkova, N.V. The interplay of proton accepting and hydride donor abilities in the mechanism of step-wise boron hydrides alcoholysis. Inorg. Chim. Acta 2017, 456, 113–119. [Google Scholar] [CrossRef]
  67. Golub, I.E.; Filippov, O.A.; Gutsul, E.I.; Belkova, N.V.; Epstein, L.M.; Rossin, A.; Peruzzini, M.; Shubina, E.S. Dimerization mechanism of bis(triphenylphosphine)copper(I) tetrahydroborate: Proton transfer via a dihydrogen bond. Inorg. Chem. 2012, 51, 6486–6497. [Google Scholar] [CrossRef] [PubMed]
  68. Belkova, N.V.; Bakhmutova-Albert, E.V.; Gutsul, E.I.; Bakhmutov, V.I.; Golub, I.E.; Filippov, O.A.; Epstein, L.M.; Peruzzini, M.; Rossin, A.; Zanobini, F.; et al. Dihydrogen bonding in complex (PP3)RuH(η1-BH4) featuring two proton-accepting hydride sites: Experimental and theoretical studies. Inorg. Chem. 2014, 53, 1080–1090. [Google Scholar] [CrossRef] [PubMed]
  69. Golub, I.E.; Filippov, O.A.; Belkova, N.V.; Epstein, L.M.; Rossin, A.; Peruzzini, M.; Shubina, E.S. Two pathways of proton transfer reaction to (triphos)Cu(η1-BH4) via a dihydrogen bond [triphos = 1,1,1-tris(diphenylphosphinomethyl)ethane]. Dalton Trans. 2016, 45, 9127–9135. [Google Scholar] [CrossRef] [PubMed]
  70. Golub, I.E.; Filippov, O.A.; Belkova, N.V.; Gutsul, E.I.; Epstein, L.M.; Rossin, A.; Peruzzini, M.; Shubina, E.S. Competition between the hydride ligands of two types in proton transfer to [{κ3-P-CH3C(CH2CH2PPh2)3}RuH(η2-BH4)]. Eur. J. Inorg. Chem. 2017. [Google Scholar] [CrossRef]
  71. Haddad, P.S.; Mauro, A.E.; Frem, R.C. Borohidreto complexos de cobre (I) contendo difosfinas. caracterizacao espectroscopica c comportamento termico. Quim. Nova 2001, 24, 786–789. [Google Scholar] [CrossRef]
  72. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.; et al. Gaussian 09; Revision D. 01; Gaussian: Wallingford, CT, USA, 2009. [Google Scholar]
  73. Zhao, Y.; Truhlar, D. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chim. Acta 2008, 120, 215–241. [Google Scholar]
  74. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  75. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  76. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  77. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  78. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  79. Swart, M.; Guell, M.; Luis, J.M.; Solà, M. Spin-state-corrected gaussian-type orbital basis sets. J. Phys. Chem. A 2010, 114, 7191–7197. [Google Scholar] [CrossRef] [PubMed]
  80. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self—Consistent molecular orbital methods. xii. further extensions of gaussian—Type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  81. Dill, J.D.; Pople, J.A. Self-consistent molecular orbital methods. XV. Extended gaussian-type basis sets for lithium, beryllium, and boron. J. Chem. Phys. 1975, 62, 2921–2923. [Google Scholar] [CrossRef]
  82. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  83. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  84. Pietro, W.J.; Francl, M.M.; Hehre, W.J.; DeFrees, D.J.; Pople, J.A.; Binkley, J.S. Self-consistent molecular orbital methods. 24. Supplemented small split-valence basis sets for second-row elements. J. Am. Chem. Soc. 1982, 104, 5039–5048. [Google Scholar] [CrossRef]
  85. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  86. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  87. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  88. Fritsch, J.; Zundel, G. Influence of the polarity of the environment on easily polarizable OH∙∙∙N = O-∙∙∙H+N hydrogen bonds. J. Phys. Chem. 1981, 85, 556–561. [Google Scholar] [CrossRef]
  89. 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] [PubMed]
  90. Dong, L.; Wen, J.; Li, W. A theoretical investigation of substituent effects on the stability and reactivity of N-heterocyclic olefin carboxylates. Org. Biomol. Chem. 2015, 13, 8533–8544. [Google Scholar] [CrossRef] [PubMed]
  91. Sumimoto, M.; Iwane, N.; Takahama, T.; Sakaki, S. Theoretical study of trans-metalation process in palladium-catalyzed borylation of iodobenzene with diboron. J. Am. Chem. Soc. 2004, 126, 10457–10471. [Google Scholar] [CrossRef] [PubMed]
  92. Sheldrick, G. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  93. Xi, Z.; Jin, D. Preliminary study on the syntheses and structures of copper complexes with bidentate phosphine ligands. Chinese J. Inorg. Chem 1993, 9, 379–383. [Google Scholar]
  94. Makhaev, V.; Borisov, A.; Baranetskaya, N.; Vil’chevskaya, V.; Gorelikova, Y.Y.; Krylova, A. Complexes of copper (I) tetrahydroborate with organometallic and polydentate phosphine ligands. Russ. J. Coord. Chem. 1992, 18, 416. [Google Scholar]
  95. Bommer, J.C.; Morse, K.W. Temperature-dependent phosphite complex equilibriums observable with NMR and IR. Inorg. Chem. 1981, 20, 1731–1734. [Google Scholar] [CrossRef]
  96. Makhaev, V.D.; Borisov, A.P.; Lobkovskii, É.B.; Polyakova, V.B.; Semenenko, K.N. Preparation of hydroborate complexes of copper and an investigation of the dependence of their structure on the nature of the organophosphorus ligands. Russ. Chem. Bull. 1985, 34, 1731–1736. [Google Scholar] [CrossRef]
  97. Khan, M.M.T.; Paul, P.; Venkatasubramanian, K. Synthesis and structural characterization of [α,α′-bis{bis(2-(diphenylphosphino)ethyl)amino}ethane] bis(tetrahydroborato)dicopper(I). Hydrogen bridge attachment of the tetrahydroborate group. Polyhedron 1991, 10, 1827–1829. [Google Scholar] [CrossRef]
  98. Makhaev, A.; Dolghusin, F.; Yanovskii, A.; Struchkov, Y.T. Crystal and molecular structure of the benzene solvate of [1,1′-bis(diphenylphosphino)ferrocene]copper(i)eta^2-tetrahydridoborate [Fe(C5H4PPh2)2] CuBH4·2C6H6. Russ. J. Coord. Chem. 1996, 22, 573–577. [Google Scholar]
  99. Lobkovskii, E.B.; Makhaev, V.D.; Borisov, A.P. Crystal and molecular structure of (tetrahydridoborato)bis(diphenylbutylphosphine)copper(I). J. Struct. Chem. 1984, 25, 496–498. [Google Scholar] [CrossRef]
  100. Lippard, S.J.; Ucko, D. Transition metal borohydride complexes. II. Th reaction of copper(I) compounds with boron hydride anions. Inorg. Chem. 1968, 7, 1051–1056. [Google Scholar] [CrossRef]
  101. Makhaev, V.; Borisov, A.; Antsyshkina, A.; Sadikov, G.; Poraj-Koshits, M.; Kudrova, N.; Mal’tseva, N.; Istomin, S.Y. Synthesis and crystal structure of (1, 10-phenanthroline)-(triethylphosphite) copper (1) tetrahydroborate. Russ. J. Coord. Chem. 1993, 19, 858–863. [Google Scholar]
  102. Green, B.E.; Kennard, C.H.L.; Smith, G.; Elcombe, M.M.; Moore, F.H.; James, B.D.; White, A.H. Crystal structures of α- and β-(1,10-phenanthroline)tetrahydroborato(triphenylphosphine)copper(I) and (2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline)tetrahydroboratocopper(I). Inorg. Chim. Acta 1984, 83, 177–189. [Google Scholar] [CrossRef]
  103. Bommer, J.C.; Morse, K.W. [(MePh2P)3CuBH4]: A single M-H bridged tetrahydroborate. J. Chem. Soc. Chem. Commun. 1977, 137–138. [Google Scholar] [CrossRef]
  104. Bianchini, C.; Ghilardi, C.A.; Meli, A.; Midollini, S.; Orlandini, A. Reactivity of copper(I) tetrahydroborates toward carbon dioxide and carbonyl sulfide. Structure of (triphos)Cu(η1-O2CH). Inorg. Chem. 1985, 24, 924–931. [Google Scholar] [CrossRef]
  105. Li, J.; White, J.M.; Mulder, R.J.; Reid, G.E.; Donnelly, P.S.; O’Hair, R.A.J. Synthesis, structural characterization, and gas-phase unimolecular reactivity of bis(diphenylphosphino)amino copper hydride nanoclusters [Cu3(X)(μ3-H)((PPh2)2NH)3](BF4), where X = μ2-Cl and μ3-BH4. Inorg. Chem. 2016, 55, 9858–9868. [Google Scholar] [CrossRef] [PubMed]
  106. Li, Q.-Z.; Dong, X.; Jing, B.; Li, W.-Z.; Cheng, J.-B.; Gong, B.-A.; Yu, Z.-W. A new unconventional halogen bond C–X···H–M between HCCX (X = Cl and Br) and HMH (M = Be and Mg): An Ab Initio study. J. Comput. Chem. 2010, 31, 1662–1669. [Google Scholar] [CrossRef] [PubMed]
  107. Smith, D.A.; Brammer, L.; Hunter, C.A.; Perutz, R.N. Metal hydrides form halogen bonds: Measurement of energetics of binding. J. Am. Chem. Soc. 2013, 136, 1288–1291. [Google Scholar] [CrossRef] [PubMed]
  108. Lipkowski, P.; Grabowski, S.J.; Leszczynski, J. Properties of the halogen−hydride interaction: An ab initio and “atoms in molecules” analysis. J. Phys. Chem. A 2006, 110, 10296–10302. [Google Scholar] [CrossRef] [PubMed]
  109. Mohajeri, A.; Alipour, M.; Mousaee, M. Halogen−hydride interaction between Z−X (Z = CN, NC; X = F, Cl, Br) and H−Mg−Y (Y = H, F, Cl, Br, CH3). J. Phys. Chem. A 2011, 115, 4457–4466. [Google Scholar] [CrossRef] [PubMed]
  110. Jabłoński, M.; Palusiak, M. Nature of a hydride–halogen bond. A SAPT-, QTAIM-, and NBO-based study. J. Phys. Chem. A 2012, 116, 2322–2332. [Google Scholar] [CrossRef] [PubMed]
  111. Esrafili, M.D.; Solimannejad, M. Revealing substitution effects on the strength and nature of halogen-hydride interactions: A theoretical study. J. Mol. Model. 2013, 19, 3767–3777. [Google Scholar] [CrossRef] [PubMed]
  112. Bai, X.; Li, Q.; Li, R.; Cheng, J.; Li, W. Is a MH (M = Be and Mg) radical a better electron donor in halogen-hydride interaction?: A theoretical comparison with HMH. Int. J. Quantum Chem. 2013, 113, 1293–1298. [Google Scholar] [CrossRef]
  113. Anufriev, S.A.; Sivaev, I.B.; Suponitsky, K.Y.; Bregadze, V.I. Practical synthesis of 9-methylthio-7,8-nido-carborane [9-MeS-7,8-C2B9H11]. Some evidences of BH···X hydride-halogen bonds in 9-XCH2(Me)S-7,8-C2B9H11 (X = Cl, Br, I). J. Organomet. Chem. 2017. [Google Scholar] [CrossRef]
  114. Moudam, O.; Kaeser, A.; Delavaux-Nicot, B.; Duhayon, C.; Holler, M.; Accorsi, G.; Armaroli, N.; Seguy, I.; Navarro, J.; Destruel, P.; et al. Electrophosphorescent homo- and heteroleptic copper(i) complexes prepared from various bis-phosphine ligands. Chem. Commun. 2007, 3077–3079. [Google Scholar] [CrossRef] [PubMed]
  115. Kaeser, A.; Mohankumar, M.; Mohanraj, J.; Monti, F.; Holler, M.; Cid, J.-J.; Moudam, O.; Nierengarten, I.; Karmazin-Brelot, L.; Duhayon, C.; et al. Heteroleptic copper(I) complexes prepared from phenanthroline and bis-phosphine ligands. Inorg. Chem. 2013, 52, 12140–12151. [Google Scholar] [CrossRef] [PubMed]
  116. Bilyachenko, A.N.; Kulakova, A.N.; Levitsky, M.M.; Petrov, A.A.; Korlyukov, A.A.; Shul’pina, L.S.; Khrustalev, V.N.; Dorovatovskii, P.V.; Vologzhanina, A.V.; Tsareva, U.S.; et al. Unusual tri-, hexa-, and nonanuclear Cu(II) cage methylsilsesquioxanes: Synthesis, structures, and catalytic activity in oxidations with peroxides. Inorg. Chem. 2017, 56, 4093–4103. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Possible coordination modes of dppm ligand in transition metal complexes.
Scheme 1. Possible coordination modes of dppm ligand in transition metal complexes.
Crystals 07 00318 sch001
Scheme 2. Possible structures of DHB complexes. Adapted with permission from ref [67]. Copyright 2012 American Chemical Society.
Scheme 2. Possible structures of DHB complexes. Adapted with permission from ref [67]. Copyright 2012 American Chemical Society.
Crystals 07 00318 sch002
Figure 1. General view of molecular structures 1 (a) and 2 (b). Thermal ellipsoids are drawn at the 50% probability level.
Figure 1. General view of molecular structures 1 (a) and 2 (b). Thermal ellipsoids are drawn at the 50% probability level.
Crystals 07 00318 g001
Figure 2. View of molecular packing of 1 (P212121) (a) and 2 (P21/c) (b) in a unit cell.
Figure 2. View of molecular packing of 1 (P212121) (a) and 2 (P21/c) (b) in a unit cell.
Crystals 07 00318 g002
Figure 3. The conformation types found from CCDC search for [(μ2-dppm)2Cu2]2+ moieties.
Figure 3. The conformation types found from CCDC search for [(μ2-dppm)2Cu2]2+ moieties.
Crystals 07 00318 g003
Figure 4. Conformational distribution of eight-membered [(μ2-dppm)2Cu2]2+ moieties.
Figure 4. Conformational distribution of eight-membered [(μ2-dppm)2Cu2]2+ moieties.
Crystals 07 00318 g004
Figure 5. The tentative structure of (η2-dppm)Cu(η2-BH4) monomer (a) and its M06-optimized geometry (b).
Figure 5. The tentative structure of (η2-dppm)Cu(η2-BH4) monomer (a) and its M06-optimized geometry (b).
Crystals 07 00318 g005
Figure 6. Experimental IR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 and simulated IR spectra of M06 and B3LYP-D2 optimized (η1-dppm)Cu(η2-BH4) monomer and 1. No scaling factors applied.
Figure 6. Experimental IR spectra of [(μ2-dppm)2Cu2][η2-BH4]2 and simulated IR spectra of M06 and B3LYP-D2 optimized (η1-dppm)Cu(η2-BH4) monomer and 1. No scaling factors applied.
Crystals 07 00318 g006
Table 1. Crystal data and structure refinement parameters for 1 and 2.
Table 1. Crystal data and structure refinement parameters for 1 and 2.
12
Brutto formulaC50H52B2Cu2 P4, CH2Cl2C50H52B2Cu2 P4, 0.5 CH2Cl2
Formula weight1010.42967.96
T, K120120
Space groupP212121P21/c
Z(Z’)4(1)4(1)
a/Å14.218(2)23.0884(18)
b/Å17.875(3)13.0448(10)
c/Å19.523(3)16.0830(13)
β/°90.0092.055(2)
Volume/Å34961.7(13)4840.8(7)
ρcalc, g/cm31.3531.328
μ/cm−111.2810.99
F(000)20882004
max, °5858
Reflections collected (Rint)50,044 (0.0480)56,740 (0.0429)
Independent reflections1314012851
Reflections with I > 2σ(I)118749781
Parameters547576
R1 [I > 2σ (I)]0.06040.0368
wR20.15690.0974
GOF1.0941.018
Residual electron density, e·Å−3 (ρmin/ρmax)−2.080/1.142−0.730/0.840
Table 2. Characteristic vibrations observed in IR spectra and 11B NMR chemical shifts of the BH4 group reported for copper(I) tetrahydroborate complexes.
Table 2. Characteristic vibrations observed in IR spectra and 11B NMR chemical shifts of the BH4 group reported for copper(I) tetrahydroborate complexes.
CompoundνBHtermνBHbrδBH2νCuBδBH4Ref.
[(EtO)3P]2Cu(η2-BH4)2380, 23501990, 19301135−29.1 c[95]
[(μ2-dppm)2Cu2][η2-BH4]22382, 23602019, 19671133358−29.5 bThis work
(PPh3)2Cu(η2-BH4)2403, 23941994,19371142374−29.7 b[67,96]
[{Ph2P(CH2)2}2NCH2]2Cu(η2-BH4)236520101120−30.2 c[97]
[(MeO)3P]2Cu(η2-BH4)2380, 23451990, 19351135−30.4 c[95]
(dppm)Cu(η2-BH4)2382, 23602018, 19651130358[94]
“[(μ2-dppm)2Cu2][η2-BH4]22391, 23451987, 19241144[93]
“(η2-dppm)Cu(η2-BH4)”2370, 22291984, 19491185378 a[71]
(dppe)Cu(η2-BH4)2384, 23411990, 19281141[93]
(dppe)Cu(η2-BH4)2380, 23602010, 19501140355[94]
(dppb)Cu(η2-BH4)2385, 23601985, 19501140[94]
(dpph)Cu(η2-BH4)2388, 23601982, 19401140356[94]
(FcPPh2)2Cu(η2-BH4)2398, 23602005, 19601140[94,98]
(Fc2PPh)22-BH4)2398, 23602005, 19501140368[94]
(dppf)Cu(η2-BH4)2397, 23542013, 19701130376[94]
(nBuPPh2)2Cu(η2-BH4)2404, 23941995, 19371139363[96,99]
[(EtO)3P]2Cu(η2-BH4)2397, 23601994, 19331137386[96]
[(iPrO)3P]2Cu(η2-BH4)2399, 23941999, 19321137384[96]
[(Me2N)3P]2Cu(η2-BH4)2392, 23662023, 19461137356[96]
[(p-MeOC6H4O)3P]2Cu(η2-BH4)2385, 23502005, 1961[100]
[(p-MeC6H4O)3P]2Cu(η2-BH4)2382, 23431990, 1930[100]
[(m-MeC6H4O)3P]2Cu(η2-BH4)2380, 23432018, 1944[100]
[(EtO)3P](phen)Cu(η2-BH4)2360, 23302080[101]
(PPh3)(phen)Cu(η2-BH4)2360, 23302070, 19101120[102]
(dmdp)Cu(η2-BH4)2385, 235019821128398[102]
(triphos)Cu(η1-BH4)2354, 23211988−32.8 b[69]
(MePPh2)3Cu(η2-BH4)2335, 231520501075, 1060−39.0 c[95,103]
[(MeO)3P]3Cu(η1-BH4)23402055−39.0 c[95]
[(EtO)3P]3Cu(η1-BH4)23352055−40.0 c[95]
(triphos)Cu(η1-BH4)2360, 23001980[104]
(NP3)Cu(η1-BH4)231020601130, 1060[104]
(EtP3)Cu(η1-BH4)237520001130[104]
For the ligands abbreviations see Supporting Information. a This stretching vibration was previously described as νCuP [71]; b CD2Cl2; c CDCl3.
Table 3. Selected structural parameters for 1 and 2.
Table 3. Selected structural parameters for 1 and 2.
Distances, Å1Distances, Å2
Cu(1)···Cu(2)3.392(1)Cu(1)–Cu(2)3.2035(4)
Cu(1)–P(2)2.238(2)Cu(1)–P(2)2.2234(7)
Cu(2)–P(1)2.253(2)Cu(2)–P(1)2.2608(7)
Cu(1)–P(3)2.254(2)Cu(1)–P(3)2.2288(6)
Cu(2)–P(4)2.257(2)Cu(2)–P(4)2.2542(6)
Cu(1)–B(1)2.194(9)Cu(1)–B(1)2.198(2)
Cu(2)–B(2)2.190(7)Cu(2)–B(2)2.192(3)
H(19)A···Cl(1’)2.722H(10)A···Cl(2)D3.031
H(13)A···Cl(1’)2.727
H(29)A···Cl(1’)2.816
H(28)A···Cl(1’)2.627
H(1)BD···Cl(2’)2.814H(26)A···H(1)BD2.246
Angles, °1Angles, °2
P(2)–Cu(1)–P(3)112.93(7)P(2)–Cu(1)–P(3)117.74(2)
P(1)–Cu(2)–P(4)111.33(6)P(1)–Cu(2)–P(4)117.29(2)
P(1)–C(1)–P(2)112.6(3)P(1)–C(1)–P(2)110.6(1)
P(3)–C(2)–P(4)109.9(4)P(3)–C(2)–P(4)111.5(1)
C(19)–H(19)A···Cl(1’)150.9C(10)–H(10)A···Cl(2)D149.2
C(13)–H(13)A···Cl(1’)140.8
C(29)–H(29)A···Cl(1’)136.0
C(28)–H(28)A···Cl(1’)150.4
B(1)–H(1)BD···Cl(2’)142.6C(26)–H(26)A···H(1)BD168.2
Dihedral Angles, °1Dihedral Angles, °2
χ1(P,Cu,Cu,P)−92.41(6)χ1(P,Cu,Cu,P)117.04(2)
χ2(P,Cu,Cu,P)133.69(6)χ2(P,Cu,Cu,P)−118.44(2)
Table 4. Summary of Cu(1)···Cu(2) distances, PCuP and dihedral χ(P,Cu,Cu,P) angles reported for eight-membered [(μ2-dppm)2Cu2]2+ moieties. N–number of CCSD structures of certain conformation.
Table 4. Summary of Cu(1)···Cu(2) distances, PCuP and dihedral χ(P,Cu,Cu,P) angles reported for eight-membered [(μ2-dppm)2Cu2]2+ moieties. N–number of CCSD structures of certain conformation.
ConformationNd[Cu(1)···Cu(2)], ÅPCuP’, °χ[P,Cu(1),Cu(2),P’], °PCHP’, °δ31P{1H}, ppm
Boat-Boat302.679–3.651113–136113–136110–117−7.8 ÷ −15.2
Distorted
Boat-Boat
212.931–3.85295–13387–115/117–139111–117−7.9 ÷ −25.7
Twist
Boat-Boat
112.743–3.757110–14089–103/134–164109–115+2.1 ÷ −14.6
Boat-Chair152.735–3.901117–133119–138111–116−6.3 ÷ −10.9
Distorted
Boat-Chair
52.712–4.644115–146113–171110–122−6.6 ÷ −18.7
Twist
Boat-Chair
32.925/3.133120–122/130–132105–108/145–148110–115−7.7/−8.2
Plain14.277148/150170117
Chair-Chair203.359–4.797130–145179–180110–147−5.6 ÷ −14.4
Table 5. Experimental and the calculated values of IR-active vibration bands for crystal structure 1 and optimized structures.
Table 5. Experimental and the calculated values of IR-active vibration bands for crystal structure 1 and optimized structures.
1Monomer
Vibration typeexptM06B3LYP-GD2M06B3LYP-GD2
νCHas(Ph)3075, 30493228, 32253230, 32203229, 32273228, 3218
νCHas(CH2)3109, 30243053, 304630583179
νBHtermas23822504250725612543
νBHterms23602459249324902497
νBHbr1as201920612127, 210119912035
νBHbr2as19672000205719682004
νCuH1433, 13841421, 14121417, 139614531444
δBH113311651187, 117811471168
νCuB358405392357362

Share and Cite

MDPI and ACS Style

Belkova, N.V.; Golub, I.E.; Gutsul, E.I.; Lyssenko, K.A.; Peregudov, A.S.; Makhaev, V.D.; Filippov, O.A.; Epstein, L.M.; Rossin, A.; Peruzzini, M.; et al. Binuclear Copper(I) Borohydride Complex Containing Bridging Bis(diphenylphosphino) Methane Ligands: Polymorphic Structures of [(µ2-dppm)2Cu22-BH4)2] Dichloromethane Solvate. Crystals 2017, 7, 318. https://doi.org/10.3390/cryst7100318

AMA Style

Belkova NV, Golub IE, Gutsul EI, Lyssenko KA, Peregudov AS, Makhaev VD, Filippov OA, Epstein LM, Rossin A, Peruzzini M, et al. Binuclear Copper(I) Borohydride Complex Containing Bridging Bis(diphenylphosphino) Methane Ligands: Polymorphic Structures of [(µ2-dppm)2Cu22-BH4)2] Dichloromethane Solvate. Crystals. 2017; 7(10):318. https://doi.org/10.3390/cryst7100318

Chicago/Turabian Style

Belkova, Natalia V., Igor E. Golub, Evgenii I. Gutsul, Konstantin A. Lyssenko, Alexander S. Peregudov, Viktor D. Makhaev, Oleg A. Filippov, Lina M. Epstein, Andrea Rossin, Maurizio Peruzzini, and et al. 2017. "Binuclear Copper(I) Borohydride Complex Containing Bridging Bis(diphenylphosphino) Methane Ligands: Polymorphic Structures of [(µ2-dppm)2Cu22-BH4)2] Dichloromethane Solvate" Crystals 7, no. 10: 318. https://doi.org/10.3390/cryst7100318

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