Next Article in Journal
Effect of Naringenin and Its Derivatives on the Probing Behavior of Myzus persicae (Sulz.)
Next Article in Special Issue
Mercury(II) Complexes of Anionic N-Heterocyclic Carbene Ligands: Steric Effects of the Backbone Substituent
Previous Article in Journal
Versatile Polypeptide-Functionalized Plasmonic Paper as Synergistic Biocompatible and Antimicrobial Nanoplatform
Previous Article in Special Issue
N-Heterocyclic Carbene Platinum(IV) as Metallodrug Candidates: Synthesis and 195Pt NMR Chemical Shift Trend
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syntheses and Reactivity of New Zwitterionic Imidazolium Trihydridoborate and Triphenylborate Species

1
School of Science and Technology, Chemistry Division, University of Camerino, via S. Agostino 1, 62032 Camerino, Macerata, Italy
2
Department of Chemistry and Biochemistry, The University of Texas at Arlington, Arlington, TX 76019-0065, USA
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(14), 3184; https://doi.org/10.3390/molecules25143184
Submission received: 18 June 2020 / Revised: 7 July 2020 / Accepted: 10 July 2020 / Published: 13 July 2020
(This article belongs to the Special Issue Carbon Ligands: From Fundamental Aspects to Applications)

Abstract

:
In this study, four new N-(alkyl/aryl)imidazolium-borates were prepared, and their deprotonation reactions were investigated. Addition of BH3•THF to N-benzylimidazoles and N-mesitylimidazoles leads to imidazolium-trihydridoborate adducts. Ammonium tetraphenylborate reacts with benzyl- or mesityl-imidazoles with the loss of one of the phenyl groups yielding the corresponding imidazolium-triphenylborates. Their authenticity was confirmed by CHN analysis, 1H-NMR, 13C-NMR, 11B-NMR, FT-IR spectroscopy, and electrospray ionization mass spectrometry (ESI-MS). 3-Benzyl-imidazolium-1-yl)trihydridoborate, (HImBn)BH3, and (3-mesityl-imidazolium-1-yl)trihydridoborate, (HImMes)BH3, were also characterized by X-ray crystallography. The reactivity of these new compounds as carbene precursors in an effort to obtain borate-NHC complexes was investigated and a new carbene-borate adduct (which dimerizes) was obtained via a microwave-assisted procedure.

Graphical Abstract

1. Introduction

N-heterocyclic carbenes (NHCs) [1,2,3] are an extremely useful and versatile class of ligands [4,5,6,7,8,9,10] with donor properties similar to phosphanes [11,12,13,14]. By tuning the steric and electronic properties around the carbene center, several carbenes featuring various σ-donating and π-accepting properties have been developed to date [15,16,17]. Their chemical versatility not only implies a wide variety of structural diversity and coordination modes, but also a capability to form stable complexes with a large number of transition metals with different oxidation states [6,7,18,19,20] and labile ligands [21,22,23,24,25,26]. Metal-NHCs complexes gained considerable interest in recent years because of their application in material chemistry [27], in catalysis [19,28,29,30,31,32,33,34,35,36,37], in carbene transfer reactions [38,39], and in medicinal inorganic chemistry [40,41,42,43,44,45,46,47,48,49].
In the last thirty years, several carbenes based on the imidazol-2-ylidene (Scheme 1a) as well as the imidazolin-2-ylidene [50] and the chain-like carbene compound have been reported [51]. Such NHCs compounds have in common the presence of only organic substituents attached to the nitrogen atoms, whereas carbenes with other main-group elements as substituents (Scheme 1b) are scarce [52,53,54,55,56]. Substitution of one of the groups attached to nitrogen by a borane would result in the generation of carbene-borate anions NHC-BR3 (Scheme 1c), as anionic analogs of the neutral imidazol-2-ylidenes. To the best of our knowledge, only few examples of monoanionic carbenes such as the on in Scheme 1c have been published as yet [53,57,58,59,60,61,62,63].
In 1998 and 2002, Siebert and co-workers reported that deprotonation of imidazole-borane complexes or imidazolium-borate species (Scheme 2a) with BuLi leads to the formation of the carbene-borate anions NHC-BH3 [59,64]. These kinds of nucleophilic carbenes allowed the formation of neutral manganese complexes and anionic iron compounds by reactions with BrMn(CO)5 and Fe(CO)5, respectively [59]. The analogous reaction with [(C7H11)Fe(CO)2Br], Cp2TiCl, VCl3, and ScCl3 yielded the corresponding metal complexes [64]. Bis(imidazolyl) compounds with BH3 or BEt3 (Scheme 2b) and their behavior towards treatment with butyllithium to give dianionic chelating dicarbene-diborate ligands have also been reported [53]. Among them, the dianionic bis(imidazol-2-ylidene) species obtained from b2 (Scheme 2) reacted with Cp2TiCl2 and Cp2ZrCl2 allowing the formation of the corresponding carbene-borate complexes [53]. Isomerization to the 2-borate imidazole forms by 1,2-BR3 migration [65], intramolecular addition/elimination or dimerization reactions may or may not occur on deprotonation [57,58,59,60,61,66]. For example, deprotonation of the triethylborane adduct (Scheme 2c) produced the isomerized N-heterocyclic carbene-borate species (Scheme 2d) [59]. Attempts to synthesize the carbene-borate anions by deprotonation of the parent imidazole (Scheme 2(e1,e2)) and benzimidazole (Scheme 2(e3)) adducts, have invariably resulted in the formation of isomers (Scheme 2f) [57,67] by ring-closure due to a rapid intramolecular nucleophilic aromatic substitution. On the other hand, Contreras et al. [60,66] reported the imidazaboles (Scheme 2(g1,g2)), by elimination of H2 from the (N-alkylimidazolium)borate species with iodine at 270 °C. Okada et al. [61] reported the synthesis of analogous imidazaboles (Scheme 2(g3,g4)), from reaction of the parent (N-alkylimidazolium)borates with organolithium reagents. Recently, Chiu and coworkers [65] reported that dimerization of 2-borylimidazoles through B−N coordination yielded the head-to-tail dimers g5 and g6 (Scheme 2). Compound g7 is the only isolable product of the reaction of [Ph2B(ImtBu)2Br] and [Ca{N(SiMe3)2}2(THF)2] [68].
Functionalized imidazole-based NHCs have attracted special interest because they can be utilized to tune the environment and properties at the coordinated metal [4,69]. Whereas there are many studies describing the coordination of chelate and pincer N-heterocyclic carbene ligands, the use of anionic NHC-borates is still scarce [52,63]. Recently, significant research efforts have been devoted to the development of ionic liquids based on (N-alkylimidazolium)borate as new potential hypergolic fuels owing to their excellent physiochemical properties including and unique hypergolic reactivity [70]. The first chelating tricarbene ligand with the topology of Trofimenko’s tris(pyrazolyl)borates [71,72], tris(3-methylimidazolin-2-ylidene-1-yl)borate, in which the carbene units are connected via a BH group, was introduced in 1995 by Fehlhammer and co-workers [54] together with its hexacarbene iron(III) and cobalt(III) complexes [73,74]. The synthesis of monoanionic chelating dicarbene bis(imidazol-2-ylidene-1-yl)borates and their use as ligands in various homoleptic and heteroleptic metal complexes has been described [75,76] and recently reviewed [15,52].
In the last years, we developed several classes of coinage metal NHCs complexes obtained from the chelating precursors [HB(RImH)3]Br2 (R = Benzyl, Mesityl and t-Butyl) [77], [H2B(HTzBn)2]Br [78], H2C(HTzR)2, and H2C(HImR)2 (HTz = 1,2,4-triazole; HIm = imidazole; R = (CH2)3SO3 or (CH2)2COO) [79]. Recently, we have focused the research work on the development of new group 11 metal-NHCs complexes obtained from the water-soluble precursors HIm1R,3RCl (R = COOCH3, COOCH2CH3, or CON(CH2CH3)2) [80,81] or the zwitterionic water-soluble precursor NaHIm1R,3R (R = (CH2)3SO3) [82].
Despite the impressive chemistry based on parent poly(azolyl)borate, the analogous mono(azolyl)borate have received very little attention in recent years [83,84]. Recently, we prepared trihydro(pyrazolyl)borates such as Na[H3B(5-(CF3)pz)] and Na[H3B(3-(NO2)pz)] and related copper(I) and silver(I) phosphane complexes [85,86].
Here, we present the synthesis of (N-(alkyl/aryl)imidazolium)borate-based systems (Scheme 3) and their reactivity as carbene precursors in the effort to obtain borate-NHCs silver(I) complexes.

2. Result and Discussion

Synthesis and Characterization

The N-(alkyl/aril)imidazolium-borate adducts 14 were synthesized in one step by two different routes (Scheme 4).
The addition of one equivalent of BH3•THF to a solution of N-benzylimidazole or N-mesitylimidazole at room temperature yields the colorless imidazolium-borate adducts 1 or 2, respectively, in nearly quantitative yields (Scheme 4a). By dissolving the crude ligands 1 and 2 in CHCl3 and CHCl3/THF solution, respectively, single crystals suitable for X-ray diffraction analysis were obtained.
Compounds 3 and 4 were prepared by addition of NH4BPh4 to an acetonitrile solution of methylimidazole or benzylimidazole under reflux conditions. It is known that under acidic conditions the tetraphenylborate anion has limited stability producing triphenylboranes [87], and when heated with alkylammonium salts can lose a phenyl ring to form a B–N bond with the ammonium compound [88]. This kind of displacement was observed in our studies: the loss of a phenyl ring and the formation of imidazolium-triphenylborate species occurred in good yields, volatile benzene and ammonia being also produced. Compound 3 was previously obtained as a crystalline byproduct of the reaction mixture of [ReO2(1-MeIm)4]+ complex and NaBPh4 in acidic conditions [89].
Derivatives 1 and 2 are white and brownish solids, respectively, both soluble in CH3OH, CHCl3, CH2Cl2, THF, DMSO, and acetone. Derivatives 3 and 4 are white solids, both soluble in THF, CH2Cl2, CHCl3, CH3CN, DMSO, and acetone.
The authenticity of compounds 14 was confirmed by CHN analysis, 1H-NMR, 13C-NMR, 11B-NMR, FT-IR spectroscopy, and electrospray ionization mass spectrometry (ESI-MS). Compounds 1 and 2 were also characterized by X-ray crystallography.
The (HImBn)BH3 (1) crystallizes in the Orthorhombic P212121 space group. The molecular structure is illustrated in Figure 1. It is monomeric in the solid state and C1-N1 distance is slightly longer than the C1-N2 distance.
The molecular structure of (HImMes)BH3 (2) is shown in Figure 2. It crystallizes in the Monoclinic P21/n space group with two chemically similar but crystallographically different molecules in the asymmetric unit. Structural features of 2 are similar to those observed for 1.
The FT-IR spectra of Compounds 14 showed weak absorptions in the range 3010–3177 cm−1, due to the azolyl ring C-H stretching and the presence of the BH3 moiety in Compounds 1 and 2 was detected by intense absorptions at 2255–2374 cm−1.
The 1H- and 13C-NMR spectra of 1 and 2 were recorded in CDCl3 and CD3OD, while the spectra of 3 and 4 were recorded in DMSO solution. Compounds 14 showed a single set of resonances for the imidazolium rings. The 1H NMR spectra of Compounds 1 and 2 at the 2-CH position does not show any reduced intensity after two days in CD3OD solution at room temperature, suggesting the absence of fast H-D exchange and therefore lack of deuteration at this position.
The 11B-NMR spectra showed a quartet at δ −19.38 and −19.21 ppm for Compounds 1 and 2, respectively, in CDCl3 solution, indicating a coordination of the imidazole rings at the BH3 group [62,90]. The single broad 11B resonances observed at δ −6.52 ppm for Compound 3 and at δ −6.37 ppm for Compound 4, in (CD3)2CO and CDCl3 solutions, respectively, are indicative of a four-coordinate boron center; they are in the range observed for analogously triphenylborate species [91], being considerably shifted in comparison with the triphenylborane one, which is observed at δ −60.2 ppm [92].
In the ESI(+)-MS spectra of 1 and 2 we observed peaks at m/z 195 and 223, due to the molecular specie [(HImBn)BH3 + Na]+ and [(HImMes)BH3 + Na]+, respectively. In addition the ESI(+)MS spectra displayed peaks due to the fragmentation species [HImR + H]+ and to the aggregates [(HImR)2BH2]+ (R = Bn or Mes). Analogously, the ESI(+)-MS spectra of Compounds 3 and 4 were dominated by the peaks at m/z 83 and 159 due to the [HImCH3 + H]+ and [HImBn + H]+, respectively, along with a fragment at m/z 247 ([(HImCH3)BPh2]+, 25%) and an aggregate at 481 ([(ImBn)2BPh2]+, 45%), in 3 and 4 respectively.
Our aim was to synthesize new N-(alkyl/aryl)imidazolium-borates and study their reactivity and investigate their reactivity as carbene precursors in an effort to obtain borate-NHCs silver(I) complexes. However, treatment of Compounds 14 with nBuLi to yield the imidazol-2-ylidenes always led to decomposition species. Further direct reactions of 14 with Ag2O, in different reaction conditions (r.t. or reflux; reaction times = 5, 24, 48, and 120 h; solvent = THF, CH2Cl2, CH3OH, and CH3CN), or with silver acetate (in CH3OH or CH3CN) to give silver carbene complexes were unsuccessful: only mixtures of unreacted or decomposition products were detected. The only partially isolable product of the reaction of 4 and Ag2O in CH3CN was the imidazabole species 5. After these efforts, we found that the direct synthesis of imidazaboles [60,61,66] could be achieved by using microwave-assisted procedure [93], following a pre-set heating ramp of 1 h up to 80 °C, in technical-grade CH3CN and in the presence of Ag2O (Scheme 5). Unfortunately, this methodology was only successful for Compound 5, and mixtures of products were obtained by using microwave-assisted procedure employing Compounds 13 as starting materials.
Compound 5 is an oil soluble in CH3OH, CH3CN and DMSO. Its formation can be explained by the abstraction of proton at 2-position of the imidazolium-triphenylborate and the successive bimolecular condensation of the produced anions with elimination of two benzene molecules (Scheme 5) [65]. Compound 5 has a framework of 1,4-diazonia-2,5-diboratacyclohexa-3,6-diene, which can also be regarded as an intramolecular carbene-borate adduct [59,60].
NMR spectra showed significant changes going from Compound 4 to the corresponding imidazabole species 5. In particular, the 1H-NMR spectrum of 5 recorded in deuterated DMSO showed the disappearance of the diagnostic 2-CH imidazolium signal of 4 at 8.37 ppm upon cyclization. Analogously, in the 13C-NMR spectrum, the 2-CH imidazolium signal of 4 at 136.33 ppm was no longer observed in the spectrum of 5 that instead showed a new, albeit poorly intense, 2-C signal at 159.18 ppm indicative of the carbene-borate formation [94]. The remaining 13C-NMR data are very similar to those of 4. The 11B-NMR spectrum contains a singlet at δ 1.43. The decreased 11B-NMR nuclear shielding in 5 as compared to 411B −6.37) points towards lower delocalization of the positive charge in the imidazabole system [95].
Isomerization to the 2-borate imidazole forms by 1,2-BR3 migration [65], intramolecular addition/elimination or dimerization reactions may occur on deprotonation [57,58,59,60,61,66], presumably involving intermediates such as in Scheme 6A,B.
DFT studies by Vagedes et al. [57] suggested that direct interconversion of such anions by 1,2-migration is very unlikely. The borate substituent thermodynamically prefers to be bound to C-2 of the anionic heterocyclic moiety. Presumably, the Lewis acidic borane compensates the negative charge much more efficiently when bound to the carbon atom than when bound to the nitrogen atom, but their interconversion was precluded by a very high barrier of the respective 1,2-BR3 shift [57]. In particular, for Compound 5, the probably initially generated “anionic Arduengo carbene” product A is proved unstable under the reaction conditions and it must be assumed that the rearrangement, experimentally observed to yield species C, is likely to have proceeded intermolecularly by two successive nucleophilic substitutions or by radical pathway as recently proposed by Chiu et al. [65].
As demonstrated in the BR3-functionalized NHC, the incorporation of anionic borate functionality enhances the donating ability of NHC [96,97]. However, we must conclude that the N-borato carbene anion A could exhibit its characteristic NHC chemistry when prepared or generated under conditions precluding intermolecular rearrangement pathways to their thermodynamically favored C(2)-borated imidazole isomers or head-to-tail imidazabole dimers.

3. Experimental Section

3.1. Materials and General Methods

All syntheses and handling were carried out under an atmosphere of dry oxygen-free dinitrogen, using standard Schlenk techniques or a glove box. Glassware was dried with a heat-gun under high vacuum. Solvents were purchased from commercial sources and purified by conventional methods prior to use. Elemental analyses (C, H, N, and S) were performed with a Fisons Instruments EA-1108 CHNS-O Elemental Analyzer (Thermo Fisher Scientific Inc., Waltham, MA, USA). Melting points were taken on an SMP3 Stuart Scientific Instrument (Bibby Sterilin Ltd., London, UK). IR spectra were recorded from 4000 to 400 cm−1 on a PerkinElmer Frontier FT-IR instrument (Perkin Elmer Inc., Waltham, MA, USA), equipped with single reflection universal diamond ATR top-plate. IR annotations used were as follows: br = broad, m = medium, mbr = medium broad, s = strong, sbr = strong broad, sh = shoulder, vs = very strong, w = weak, wbr = weak broad. 1H-, 13C-, and 11B-NMR spectra were recorded with an Oxford AS400 Varian spectrometer (400.4 MHz for 1H, 100.1 MHz for 13C, and 128.4 MHz for 11B) (Oxford Instruments, MA, USA) or with a 500 Bruker Ascend (500.1 MHz for 1H, 125 MHz for 13C, and 160.5 MHz for 11B) (Bruker BioSpin Corporation, 15 Fortune Drive, Billerica, MA, USA). Referencing was relative to tetramethylsilane (TMS) (1H and 13C) and BF3.Et2O (11B). NMR annotations used were as follows: br = broad; d = doublet, m = multiplet, s = singlet. Syntheses under microwave irradiation were performed by means of a Flexible Microwave Platform FlexSynth Milestone apparatus (Milestone Srl, Via Fatebenefratelli, Sorisole (BG), Italy). The reactions were performed in a 100-mL PTFE vessel, sealed using a Teflon crimp top. Electrospray mass spectra (ESI-MS) were obtained in positive—(ESI(+)MS) or negative-ion (ESI(−)MS) mode on an Agilent Technologies Series 1100 LC/MSD Mass Spectrometer (Agilent Technologies Inc, Santa Clara, CA, USA), using a methanol or acetonitrile mobile phase. The compounds were added to reagent grade methanol to give approximately 0.1 mM solutions, injected (1 µL) into the spectrometer via a Hewlett Packard 1090 Series II UV-Visible HPLC system (Agilent Technologies Inc, Santa Clara, CA, USA) fitted with an autosampler. The pump delivered the solutions to the mass spectrometer source at a flow rate of 300 mL min−1, and nitrogen was employed both as a drying and nebulizing gas. Capillary voltages were typically 4000 and 3500 V for the ESI(+)MS and ESI(−)MS modes, respectively. Confirmation of all major species in this ESI-MS study was supported by comparison of the observed and predicted isotope distribution patterns, the latter calculated using the IsoPro 3.1 computer program (T-Tech Inc., Norcross, GA, USA). 1-Benzylimidazole, 1-methylimidazole, BH3•THF complex, ammonium tetraphenylborate, and silver oxide were purchased from Sigma-Aldrich (Merck Life Science S.r.l., Via Monte Rosa, Milano, Italy). The 1-mesitylimidazole was synthesized in accordance with the literature method [98].
Caution! The materials used and synthesized in this study are energetic. They should be handled in quantities not exceeding the millimolar scale. Manipulations should be carried out behind blast shields and with adequate personal safety gear.

3.1.1. Synthesis of (HImBn)BH3 (1)

1-Benzylimidazole (1.840 g, 11.631 mmol) was dissolved in dry THF (50 mL) under N2 atmosphere and BH3•THF complex (12.0 mL, 1M) was added drop by drop. The reaction mixture was stirred at room temperature for 24 h. Then, the volatiles were removed under reduced pressure to give a colorless oil. It was re-crystallized by CHCl3/diethyl ether/n-hexane (1/3/3) solution to obtain a white precipitate; it was filtered, washed with diethyl ether, and dried under reduced pressure to give 1 in 80% yield (1.601 g). Single crystals of 1 suitable for X-ray analysis were obtained by slow evaporation of a CHCl3 solution of 1. Melting point: 92–94 °C. IR (cm−1): 3159w, 3135m, 3061w, 3038w (C-H); 2352m, 2297m, 2255m (B-H); 1540m, 1533m (C=C/C=N). 1H-NMR (CDCl3, 293 K): δ 2.2 (br, 3H, BH3), 5.13 (s, 2H, CH2Ph), 6.91 (s, 1H, 4-CH or 5-CH), 7.14 (s, 1H, 4-CH or 5-CH), 7.23–7.44 (m, 5H, C6H5), 7.79 (s, 1H, 2-CH). 1H-NMR (CD3OD, 293 K): δ 2.2 (qbr, 3H, BH3), 5.24 (s, 2H, CH2Ph), 7,03 (s, 1H, 4-CH or 5-CH), 7.19 (s, 1H, 4-CH or 5-CH), 7.26–7.43 (m, 5H, C6H5), 8.13 (s, 1H, 2-CH). 13C{1H}-NMR (CDCl3, 293 K): δ 52.35 (CH2Ph), 119.94, 127.98, 128.21, 129.33, 129.47, 133.46 (CH), 136.33 (2-CH). 11B{1H}-NMR (CDCl3, 293 K): δ −19.38 (s, BH3). 11B-NMR (CDCl3, 293 K): δ −19.38 (q, BH3, JB-H = 96 Hz). ESI-MS (major positive-ions, CH3OH), m/z (%): 159 (40) [HImBn + H]+, 181 (40) [HImBn + Na]+, 195 (90) [(HImBn)BH3 + Na]+, 329 (100) [(HImBn)2BH2]+. Anal. Calcd. for C10H13BN2: C 69.82, H 7.62, N 16.28%. Found: C 69.52, H 7.30, N 15.91%.

3.1.2. Synthesis of (HImMes)BH3 (2)

1-mesityl-imidazole (0.930 g, 5.000 mmol) was dissolved in dry THF (30 mL) under N2 atmosphere and BH3•THF complex (5.2 mL, 1M) was added drop by drop. The reaction mixture was stirred at room temperature for 24 h. Then, the volatiles were removed under reduced pressure to give a brown oil. It was re-crystallized by CHCl3/diethyl ether/n-hexane (1/3/3) solution to obtain a brown precipitate; it was filtered, washed with diethyl ether, and dried under reduced pressure to give 1 in 68% yield (0.680 g). Single crystals of 2 suitable for X-ray analysis were obtained by slow evaporation of a CHCl3/THF solution of 2. Melting point: 109–111 °C. IR (cm−1): 3177w, 3155w, 3132w, 3061w, 3028w (C-H); 2374m, 2338m, 2323m, 2300m, 2259m (B-H); 1526s (C=C/C=N). 1H-NMR (CDCl3, 293 K): δ 2.03 (s, 6H, CH3Mes), 2.3 (br, 3H, BH3), 2.37 (s, 3H, CH3Mes), 6.90 (s, 1H, 4-CH or 5-CH), 7.02 (s, 2H, CHMes), 7.31 (s, 1H, 4-CH or 5-CH), 7.75 (s, 1H, 2-CH). 1H-NMR (CD3OD, 293 K): δ 2.04 (s, 6H, CH3Mes), 2.1 (br, 3H, BH3), 2.35 (s, 3H, CH3Mes), 7.08 (s, 2H, CHMes), 7.23 (s, 1H, 4-CH or 5-CH), 7.25 (s, 1H, 4-CH or 5-CH), 8.14 (s, 1H, 2-CH). 13C{1H}-NMR (CDCl3, 293 K): δ 17.33, 21.06 (CH3Mes), 121.25, 128.12, 129.49, 131.71, 134.86, 136.85 (CH), 140.34 (2-CH). 11B-NMR (CDCl3, 293 K): δ −19.21 (dbr, BH3). ESI-MS (major positive-ions, CH3OH), m/z (%): 187 (15) [HImMes + H]+, 223 (55) [(HImMes)BH3 + Na]+, 385 (100) [(HImMes)2BH2]+. Anal. Calcd. for C12H17BN2: C 72.03, H 8.56, N 14.00%. Found: C 71.81, H 8.25, N 13.60%.

3.1.3. Synthesis of (HImCH3)BPh3 (3)

A large excess of 1-methylimidazole (0.603 g, 7.344 mmol) was dissolved in acetonitrile (CH3CN, 60 mL). Then, ammonium tetraphenylborate (NH4BPh4, 1.770 g, 5.248 mmol) was added to the solution. A white precipitate was formed, but the solution became limpid after 1 h. The reaction proceeded for 70 h at reflux under magnetic stirring. At the end, the solution was dried at reduced pressure, obtaining a white solid. Et2O was added to the round-bottom flask to purify the residue from the starting materials that did not react. The resulting suspension was filtered, dried under reduced pressure, and furthe purified with CHCl3 to precipitate the excess of NH4BPh4. The mixture was filtered and the mother liquors were dried at reduced pressure to give the white ligand (HImCH3)BPh3 (3) in 76% yield (1.293 g). Melting point: 209–212 °C. IR (cm−1): 3158w, 3133m, 3085w, 3064m, 3054mbr, 3010mbr (C-H); 1546m, 1531m, 1483mbr (C=C/C=N). 1H-NMR (DMSO-d6, 293 K): δ 3.79 (s, 3H, NCH3), 6.90 (d, 1H, 4-CH or 5-CH), 7.03–7.15 (m, 15H, CH), 7.44 (d, 1H, 4-CH or 5-CH), 8.09 (s, 1H, 2-CH). 13C{1H}-NMR (DMSO-d6, 293 K): δ 35.13 (NCH3), 122.34, 124.85, 126.42, 127.02, 134.49, 138.58 (CH). 11B-NMR (Acetone-d6, 293 K): δ −6.52 (s, BPh3). ESI-MS (major positive ions, CH3CN), m/z (%): 83 (100) [HImCH3 + H]+, 247 (25) [(HImCH3)BPh2]+. Anal. Calcd. for C22H21BN2: C 81.50, H 6.53, N 8.64. Found: C 81.14, H 6.56, N 8.38.

3.1.4. Synthesis of (HImBn)BPh3 (4)

A large excess of 1-benzylimidazole (0.633 g, 4.000 mmol) was dissolved in CH3CN (60 mL). Then, NH4BPh4 (0.961 g, 2.850 mmol) was added to the solution. A white precipitate was formed, but the solution became limpid after 1 h. The reaction proceeded for 70 h at reflux under magnetic stirring. At the end, the solution was dried at reduced pressure, obtaining a white solid. EtOH was added to the round-bottom flask to purify the residue from the starting materials that did not react. The resulting suspension was filtered and dried at reduced pressure to give the white ligand (HImBn)BPh3 (4) in 50% yield (0.570 g). Melting point: 175–178 °C. IR (cm−1): 3163m, 3140m, 3125m, 3064mbr, 3023m (C-H); 1531mbr, 1506m, 1489mbr (C=C/C=N). 1H-NMR (DMSO-d6, 293 K): δ 5.39 (s, 2H, CH2Ph), 6.91 (s, 1H, 4-CH or 5-CH), 7.04–7.43 (m, 20H, C6H5), 7.49 (s, 1H, 4-CH or 5-CH), 8.37 (s, 1H, 2-CH). 13C{1H}-NMR (DMSO-d6, 293 K): δ 51.24 (CH2Ph), 121.25, 124.90, 127.23, 128.21, 129.33, 129.47, 133.46 (CH), 136.33 (2-CH). 11B-NMR (CDCl3, 293 K): δ −6.37 (s, BPh3). ESI-MS (major positive ions, CH3CN), m/z (%): 91 (80) [C7H7]+, 159 (100) [HImBn + H]+, 242 (50) [BPh3 + H]+, 481 (45) [(ImBn)2BPh2]+. Anal. Calcd. for C28H25BN2: C 84.01, H 6.29, N 7.00. Found: C 83.72, H 6.03, N 7.06.

3.1.5. Synthesis of (ImBnBPh2)2 (5)

In a 100-mL PTFE vessel equipped with a magnetic stir bar, Compound 4 (0.360 g, 0.900 mmol), silver oxide (Ag2O, 0.104 g, 0.450 mmol), and CH3CN (25 mL) were added. The reaction mixture was heated in the microwave reactor following a pre-set heating ramp, up to 80 °C. Once the temperature was reached, the reaction proceeded for 1 h and then it was cooled following a pre-set cooling ramp, to room temperature. All the steps were performed always under magnetic stirring. At the end, the mixture was filtered and the obtained mother liquors were dried at reduced pressure to give the oily brownish residue (ImBnBPh2)2 (5) in 54% yield (0.157 g). IR (cm−1): 3161m, 3143m, 3113sh, 3087m, 3064m, 3038m, 3024m, 3010m, 2999m, 2972m, 2938wbr (C-H); 1600m, 1587m, 1571m, 1534s, 1509s, 1496sbr (C=C/C=N). 1H-NMR (DMSO-d6, 293 K): δ 5.18 (s, 2H, CH2), 6.91 (s, 1H, 4-CH or 5-CH), 7.18–7.36 (m, 15H, ArH), 7.77 (s, 1H, 4-CH or 5-CH). 1H-NMR (CDCl3, 293 K): δ 5.13 (s, 2H, CH2), 6.92 (s, 1H, 4-CH or 5-CH), 7.11–7.44 (m, 15H, ArH), 7.67 (s, 1H, 4-CH or 5-CH). 13C{1H}-NMR (DMSO-d6, 293 K): δ 50.04 (CH2Ph), 120.14, 127.98, 128.28, 128.79, 128.94, 129.17, 130.56, 134.47 (CH), 159.18 (2-C). 11B-NMR (DMSO-d6, 293 K): δ 1.43 (s). ESI-MS (major positive ions, CH3CN), m/z (%): 91 (95) [C7H7]+, 159 (100) [HImBn + H]+. Elemental analysis for C29H27AgBN2 (%): calculated: H 5.94, C 82.01, N 8.69; found: H 6.04, C 81.27, N 8.89.

3.2. Crystallographic Data Collection and Refinement

A suitable crystal covered with a layer of hydrocarbon/Paratone-N oil was selected and mounted on a Cryo-loop and immediately placed in the low temperature nitrogen stream. X-ray intensity data were measured at 100(2) K on a Bruker SMART APEX II CCD area detector system equipped with an Oxford Cryosystems 700 series cooler, a graphite monochromator, and a Mo Kα fine-focus sealed tube (λ = 0.71073 Å). Intensity data were processed using the Bruker ApexII program suite. Absorption corrections were applied by using SADABS. Initial atomic positions were located by direct methods using XS, and the structures of the compounds were refined by the least-squares method using SHELXL [99]. All the non-hydrogen atoms were refined anisotropically. The hydrogen atoms attached to boron (B-H) were located in difference Fourier maps, included and refined freely with isotropic displacement parameters. All the other hydrogen atoms were placed at calculated positions and refined using a riding model. X-ray structural figures were generated using Olex2 [100]. The CCDC 2010217–2010218 contain the supplementary crystallographic data. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html or from the Cambridge Crystallographic Data Centre (CCDC), 12 Union Road, Cambridge, CB2 1EZ, UK).

4. Conclusions

Two imidazolium-trihydridoborate adducts were obtained by addition of BH3•THF to N-benzyl- and N-mesitylimidazoles. In addition, two imidazolium-triphenylborates were obtained by displacement of one phenyl group of ammonium tetraphenylborate reacting with methyl- or benzyl-imidazoles. 3-Benzyl-imidazolium-1-yl)trihydridoborate and (3-mesityl-imidazolium-1-yl)trihydridoborate were also characterized by X-ray crystallography. The reactivity of these new compounds as carbene precursors was investigated and a new dimeric carbene-borate adduct was obtained via a microwave-assisted procedure. The intermolecular rearrangement pathway to the head-to-tail imidazabole dimer prevented the isolation of this type of compounds and the development of their characteristic NHC chemistry.

Author Contributions

Conceptualization, C.S.; Data curation, R.V. and H.V.R.D.; Formal analysis, L.B. and H.V.R.D.; Investigation, M.P.; Methodology, M.P.; Supervision, M.P. and C.S.; Writing – original draft, M.P., H.V.R.D. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the University of Camerino (FAR 2019). H.V.R.D. is thankful for the financial support by the Robert A. Welch Foundation (Grant Y-1289).

Acknowledgments

We are grateful to CIRCMSB (Consorzio Interuniversitario di Ricerca in Chimica dei Metalli nei Sistemi Biologici).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Arduengo, A.J.; Dias, H.V.R.; Harlow, R.L.; Kline, M. Electronic stabilization of nucleophilic carbenes. J. Am. Chem. Soc. 1992, 114, 5530–5534. [Google Scholar] [CrossRef]
  2. Igau, A.; Grutzmacher, H.; Baceiredo, A.; Bertrand, G. Analogous.alpha.,.alpha.’-bis-carbenoid, triply bonded species: Synthesis of a stable.lambda.3-phosphino carbene-.lambda.5-phosphaacetylene. J. Am. Chem. Soc. 1988, 110, 6463–6466. [Google Scholar] [CrossRef]
  3. Arduengo, A.J.; Harlow, R.L.; Kline, M. A stable crystalline carbene. J. Am. Chem. Soc. 1991, 113, 361–363. [Google Scholar] [CrossRef]
  4. Hopkinson, M.N.; Richter, C.; Schedler, M.; Glorius, F. An overview of N-heterocyclic carbenes. Nature 2014, 510, 485–496. [Google Scholar] [CrossRef]
  5. Martin, C.D.; Soleilhavoup, M.; Bertrand, G. Carbene-Stabilized Main Group Radicals and Radical Ions. Chem. Sci. 2013, 4, 3020–3030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hahn, F.E.; Jahnke, M.C. Heterocyclic Carbenes: Synthesis and Coordination Chemistry. Angew. Chem. Int. Ed. 2008, 47, 3122–3172. [Google Scholar] [CrossRef] [PubMed]
  7. Nolan, S.P. N-Heterocyclic Carbenes in Synthesis; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2006. [Google Scholar]
  8. Dröge, T.; Glorius, F. The Measure of All Rings-N-Heterocyclic Carbenes. Angew. Chem. Int. Ed. 2010, 49, 6940–6952. [Google Scholar] [CrossRef]
  9. Hahn, F.E. Heterocyclic Carbenes. Angew. Chem. Int. Ed. 2006, 45, 1348–1352. [Google Scholar] [CrossRef]
  10. Díez-González, S. N-Heterocyclic Carbenes: From Laboratory Curiosities to Efficient Synthetic Tools: Edition 2; Royal Society of Chemistry: Cambridge, UK, 2017. [Google Scholar]
  11. Öfele, K.; Herrmann, W.A.; Mihalios, D.; Elison, M.; Herdtweck, E.; Scherer, W.; Mink, J. Multiple bonds between Main-Group elements and transition metals. CXXVI. Heterocyclene-carbenes as phosphine-analog ligands in metal complexes. J. Organomet. Chem. 1993, 459, 177–184. [Google Scholar] [CrossRef]
  12. Kühl, O. Functionalised N-Heterocyclic Carbene Complexes; John Wiley & Sons Ltd.: Chichester, UK, 2010. [Google Scholar]
  13. Zeng, X.; Soleilhavoup, M.; Bertrand, G. Gold-Catalyzed Intermolecular Markovnikov Hydroamination of Allenes with Secondary Amines. Org. Lett. 2009, 11, 3166–3169. [Google Scholar] [CrossRef] [Green Version]
  14. Crabtree, R.H. NHC ligands versus cyclopentadienyls and phosphines as spectator ligands in organometallic catalysis. J. Organomet. Chem. 2005, 690, 5451–5457. [Google Scholar] [CrossRef]
  15. Nesterov, V.; Reiter, D.; Bag, P.; Frisch, P.; Holzner, R.; Porzelt, A.; Inoue, S. NHCs in Main Group Chemistry. Chem. Rev. 2018, 118, 9678–9842. [Google Scholar] [CrossRef] [PubMed]
  16. Arduengo, A.J.; Bertrand, G. Carbenes Introduction. Chem. Rev. 2009, 109, 3209–3210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Doddi, A.; Peters, M.; Tamm, M. N-Heterocyclic Carbene Adducts of Main Group Elements and Their Use as Ligands in Transition Metal Chemistry. Chem. Rev. 2019, 119, 6994–7112. [Google Scholar] [CrossRef]
  18. Arnold, P.L.; Casely, I.J. F-Block N-Heterocyclic Carbene Complexes. Chem. Rev. 2009, 109, 3599–3611. [Google Scholar] [CrossRef]
  19. Cazin, C.S.J. N-Heterocyclic Carbenes in Transition Metal Catalysis and Organocatalysis; Springer Science & Business Media: Dordrecht, The Netherlands, 2011; Volume 32. [Google Scholar]
  20. Lin, J.C.Y.; Huang, R.T.W.; Lee, C.S.; Bhattacharyya, A.; Hwang, W.S.; Lin, I.J.B. Coinage Metal−N-Heterocyclic Carbene Complexes. Chem. Rev. 2009, 109, 3561–3598. [Google Scholar] [CrossRef]
  21. Dash, C.; Kroll, P.; Yousufuddin, M.; Dias, H.V.R. Isolable, gold carbonyl complexes supported by N-heterocyclic carbenes. Chem. Commun. 2011, 47, 4478. [Google Scholar] [CrossRef]
  22. Celik, M.A.; Dash, C.; Adiraju, V.A.; Das, A.; Yousufuddin, M.; Frenking, G.; Dias, H.V.R. End-On and Side-On π-Acid Ligand Adducts of Gold(I): Carbonyl, Cyanide, Isocyanide, and Cyclooctyne Gold(I) Complexes Supported by N-Heterocyclic Carbenes and Phosphines. Inorg. Chem. 2012, 52, 729–742. [Google Scholar] [CrossRef]
  23. Dash, C.; Das, A.; Yousufuddin, M.; Dias, H.V.R. Isolable, Copper(I) Dicarbonyl Complexes Supported by N-Heterocyclic Carbenes. Inorg. Chem. 2013, 52, 1584–1590. [Google Scholar] [CrossRef]
  24. Dash, C.; Yousufuddin, M.; Cundari, T.R.; Dias, H.V.R.; Dias, H.V.R. Gold-Mediated Expulsion of Dinitrogen from Organic Azides. J. Am. Chem. Soc. 2013, 135, 15479–15488. [Google Scholar] [CrossRef]
  25. Wang, G.; Ponduru, T.T.; Wang, Q.; Zhao, L.; Frenking, G.; Dias, H.V.R. Heterobimetallic Complexes Featuring Fe(CO)5 as a Ligand on Gold. Chem. - A Eur. J. 2017, 23, 17222–17226. [Google Scholar] [CrossRef] [PubMed]
  26. Dash, C.; Wang, G.; Muñoz-Castro, A.R.; Ponduru, T.T.; Zacharias, A.O.; Yousufuddin, M.; Dias, H.V.R. Organic Azide and Auxiliary-Ligand-Free Complexes of Coinage Metals Supported by N-Heterocyclic Carbenes. Inorg. Chem. 2019, 59, 2188–2199. [Google Scholar] [CrossRef] [PubMed]
  27. Smith, C.A.; Narouz, M.R.; Lummis, P.A.; Singh, I.; Nazemi, A.; Li, C.H.; Crudden, C.M. N-Heterocyclic Carbenes in Materials Chemistry. Chem. Rev. 2019, 119, 4986–5056. [Google Scholar] [CrossRef] [PubMed]
  28. Rovis, T.; Nolan, S.P. Stable Carbenes: From ‘Laboratory Curiosities’ to Catalysis Mainstays. Synlett 2013, 24, 1188–1189. [Google Scholar] [CrossRef] [Green Version]
  29. Schaper, L.-A.; Hock, S.J.; Herrmann, W.A.; Kuehn, F.E. Synthesis and Application of Water-Soluble NHC Transition-Metal Complexes. Angew. Chem. int. Ed. 2013, 44, 270–289. [Google Scholar] [CrossRef]
  30. He, Y.; Lv, M.-F.; Cai, C. A simple procedure for polymer-supported N-heterocyclic carbene silver complex via click chemistry: An efficient and recyclable catalyst for the one-pot synthesis of propargylamines. Dalton Trans. 2012, 41, 12428–12433. [Google Scholar] [CrossRef]
  31. Li, Y.; Chen, X.; Song, Y.; Fang, L.; Zou, G. Well-defined N-heterocyclic carbene silver halides of 1-cyclohexyl-3-arylmethylimidazolylidenes: Synthesis, structure and catalysis in A3-reaction of aldehydes, amines and alkynes. Dalton Trans. 2011, 40, 2046. [Google Scholar] [CrossRef]
  32. Herrmann, W.A. N-heterocyclic carbenes: A new concept in organometallic catalysis. Angew. Chem. Int. Ed. 2002, 41, 1290–1309. [Google Scholar] [CrossRef]
  33. Glorius, F.; Glorius, F. N-Heterocyclic Carbenes in Transition Metal Catalysis; Springer-Verlag Berlin Heidelberg: Heidelberg, Germany, 2007. [Google Scholar]
  34. Marion, N.; Nolan, S.P.; Díez-González, S. N-Heterocyclic Carbenes as Organocatalysts. Angew. Chem. Int. Ed. 2007, 46, 2988–3000. [Google Scholar] [CrossRef]
  35. Izquierdo, J.; Hutson, G.E.; Cohen, D.T.; Scheidt, K.A. A continuum of progress: Applications of N-hetereocyclic carbene catalysis in total synthesis. Angew. Chem. Int. Ed. 2012, 51, 11686–11698. [Google Scholar] [CrossRef] [Green Version]
  36. Bugaut, X.; Glorius, F. Organocatalytic umpolung: N-heterocyclic carbenes and beyond. Chem. Soc. Rev. 2012, 41, 3511. [Google Scholar] [CrossRef] [PubMed]
  37. Díez-González, S.; Marion, N.; Nolan, S.P. N-Heterocyclic Carbenes in Late Transition Metal Catalysis. Chem. Rev. 2009, 109, 3612–3676. [Google Scholar] [CrossRef] [PubMed]
  38. Kantchev, E.A.B.; O’Brien, C.J.; Organ, M.G. Palladium Complexes of N-Heterocyclic Carbenes as Catalysts for Cross-Coupling Reactions — A Synthetic Chemist′s Perspective. Angew. Chem. Int. Ed. 2007, 38, 2768–2813. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, H.M.J.; Lin, I.J.B. Facile Synthesis of Silver(I)−Carbene Complexes. Useful Carbene Transfer Agents. Organometallics 1998, 17, 972–975. [Google Scholar] [CrossRef]
  40. Mjos, K.D.; Orvig, C. Metallodrugs in Medicinal Inorganic Chemistry. Chem. Rev. 2014, 114, 4540–4563. [Google Scholar] [CrossRef] [PubMed]
  41. Aher, S.B.; Muskawar, P.N.; Thenmozhi, K.; Bhagat, P.R. Recent developments of metal N-heterocyclic carbenes as anticancer agents. Eur. J. Med. Chem. 2014, 81, 408–419. [Google Scholar] [CrossRef] [PubMed]
  42. Ceresa, C.; Bravin, A.; Cavaletti, G.; Pellei, M.; Santini, C. The combined therapeutical effect of metal-based drugs and radiation therapy: The present status of research. Curr. Med. Chem. 2014, 21, 2237–2265. [Google Scholar] [CrossRef]
  43. Budagumpi, S.; Haque, R.A.; Endud, S.; Rehman, G.U.; Salman, A.W. Biologically Relevant Silver(I)-N-Heterocyclic Carbene Complexes: Synthesis, Structure, Intramolecular Interactions, and Applications. Eur. J. Inorg. Chem. 2013, 2013, 4367–4388. [Google Scholar] [CrossRef]
  44. Liu, W.; Gust, R. Metal N-heterocyclic carbene complexes as potential antitumor metallodrugs. Chem. Soc. Rev. 2013, 42, 755–773. [Google Scholar] [CrossRef]
  45. Monteiro, D.C.F.; Phillips, R.M.; Crossley, B.D.; Fielden, J.; Willans, C.E. Enhanced cytotoxicity of silver complexes bearing bidentate N-heterocyclic carbeneligands. Dalton Trans. 2012, 41, 3720. [Google Scholar] [CrossRef] [Green Version]
  46. Hindi, K.M.; Panzner, M.J.; Tessier, C.A.; Cannon, C.L.; Youngs, W.J. The Medicinal Applications of Imidazolium Carbene−Metal Complexes. Chem. Rev. 2009, 109, 3859–3884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Teyssot, M.-L.; Jarrousse, A.-S.; Manin, M.; Chevry, A.; Roche, S.; Norre, F.; Beaudoin, C.; Morel, L.; Boyer, D.; Mahiou, R.; et al. Metal-NHC complexes: A survey of anti-cancer properties. Dalton Trans. 2009, 35, 6894. [Google Scholar] [CrossRef] [PubMed]
  48. Hartinger, C.G.; Dyson, P.J. Bioorganometallic chemistry—from teaching paradigms to medicinal applications. Chem. Soc. Rev. 2009, 38, 391–401. [Google Scholar] [CrossRef] [PubMed]
  49. Porchia, M.; Pellei, M.; Marinelli, M.; Tisato, F.; Del Bello, F.; Santini, C. New insights in Au-NHCs complexes as anticancer agents. Eur. J. Med. Chem. 2018, 146, 709–746. [Google Scholar] [CrossRef] [PubMed]
  50. Arduengo, A.J.; Goerlich, J.R.; Marshall, W.J. A stable diaminocarbene. J. Am. Chem. Soc. 1995, 117, 11027–11028. [Google Scholar] [CrossRef]
  51. Alder, R.W.; Allen, P.R.; Murray, M.; Orpen, A.G. Bis(diisopropylamino)carbene. Angew. Chem. Int. Ed. 1996, 35, 1121–1123. [Google Scholar] [CrossRef]
  52. Santini, C.; Marinelli, M.; Pellei, M. Boron-Centered Scorpionate-Type NHC-Based Ligands and Their Metal Complexes. Eur. J. Inorg. Chem. 2016, 2016, 2312–2331. [Google Scholar] [CrossRef]
  53. Weiss, A.; Pritzkow, H.; Siebert, W. Synthesis, Structures and Reactivity ofN-Borane-Protected 1,1′-Bisimidazoles with Different Bridging Functions. Eur. J. Inorg. Chem. 2002, 2002, 1607–1614. [Google Scholar] [CrossRef]
  54. Kernbach, U.; Ramm, M.; Luger, P.; Fehlhammer, W.P. A Chelating Triscarbene Ligand and Its Hexacarbene Iron Complex. Angew. Chem. Int. Ed. 1996, 35, 310–312. [Google Scholar] [CrossRef]
  55. Lapointe, R.E.; Roof, G.R.; Abboud, K.A.; Klosin, J. New Family of Weakly Coordinating Anions. J. Am. Chem. Soc. 2000, 122, 9560–9561. [Google Scholar] [CrossRef]
  56. Asada, T.; Hoshimoto, Y.; Ogoshi, S. Rotation-Triggered Transmetalation on a Heterobimetallic Cu/Al N-Phosphine-Oxide-Substituted Imidazolylidene Complex. J. Am. Chem. Soc. 2020, 142, 9772–9784. [Google Scholar] [CrossRef]
  57. Vagedes, D.; Kehr, G.; König, D.; Wedeking, K.; Fröhlich, R.; Erker, G.; Mück-Lichtenfeld, C.; Grimme, S. Formation of Isomeric BAr3 Adducts of 2-Lithio-N-methylimidazole. Eur. J. Inorg. Chem. 2002, 2002, 2015–2021. [Google Scholar] [CrossRef]
  58. Wacker, A.; Pritzkow, H.; Siebert, W. Nucleophilic Substitution Reactions with the 3-Borane-1,4,5-trimethylimidazol-2-ylidene Anion. – Unexpected Formation of an Imidazabole Isomer. Eur. J. Inorg. Chem. 1999, 5, 789–793. [Google Scholar] [CrossRef]
  59. Wacker, A.; Pritzkow, H.; Siebert, W. Borane-substituted imidazol-2-ylidenes. Syntheses, structures, and reactivity. Eur. J. Inorg. Chem. 1998, 6, 843–849. [Google Scholar] [CrossRef]
  60. Padilla-Martínez, I.I.; Martínez-Martínez, F.J.; López-Sandoval, A.; Girón-Castillo, K.I.; Brito, M.A.; Contreras, R. New imidazabole derivatives: Dimers of carbene-borane adducts. Eur. J. Inorg. Chem. 1998, 10, 1547–1553. [Google Scholar]
  61. Okada, K.; Suzuki, R.; Oda, M. Novel boron–nitrogen containing compounds from the reaction of organolithiums with complexes between dimesitylfluoroborane and six- or five-membered aza aromatic compounds. J. Chem. Soc., Chem. Commun. 1995, 20, 2069–2070. [Google Scholar] [CrossRef]
  62. Padilla-Martínez, I.I.; Ariza-Castolo, A.; Contreras, R. NMR Study of isolobal N-CH3+, N-BH3 and N-BF3 imidazole derivatives. Magn. Reson. Chem. 1993, 31, 189–193. [Google Scholar] [CrossRef]
  63. Nasr, A.; Winkler, A.; Tamm, M. Anionic N-heterocyclic carbenes: Synthesis, coordination chemistry and applications in homogeneous catalysis. Coord. Chem. Rev. 2016, 316, 68–124. [Google Scholar] [CrossRef]
  64. Wacker, A.; Yan, C.G.; Kaltenpoth, G.; Ginsberg, A.; Arif, A.M.; Ernst, R.; Pritzkow, H.; Siebert, W. Metal complexes of anionic 3-borane-1-alkylimidazol-2-ylidene derivatives. J. Organomet. Chem. 2002, 641, 195–202. [Google Scholar] [CrossRef]
  65. Liu, W.-C.; Liu, Y.-H.; Lin, T.-S.; Peng, S.-M.; Chiu, C.-W. 1,2-Migration of N-Diarylboryl Imidazol-2-ylidene through Intermolecular Radical Process. Inorg. Chem. 2017, 56, 10543–10548. [Google Scholar] [CrossRef]
  66. Padilla-Martínez, I.I.; Rosalez-Hoz, M.D.J.; Contreras, R.; Kerschl, S.; Wrackmeyer, B. From Azole—Borane Adducts to Azaboles—Molecular Structure of an Imidazabole. Eur. J. Inorg. Chem. 1994, 127, 343–346. [Google Scholar] [CrossRef]
  67. Vagedes, D.; Erker, G.; Kehr, G.; Bergander, K.; Kataeva, O.; Fröhlich, R.; Grimme, S.; Mück-Lichtenfeld, C. Tris(pentafluorophenyl)borane adducts of substituted imidazoles: Conformational features and chemical behavior upon deprotonation. Dalton Trans. 2003, 7, 1337–1344. [Google Scholar] [CrossRef]
  68. Arrowsmith, M.; Heath, A.; Hill, M.S.; Hitchcock, P.B.; Kociok-Köhn, G. Tris(imidazolin-2-ylidene-1-yl)borate Complexes of the Heavier Alkaline Earths: Synthesis and Structural Studies. Organometallics 2009, 28, 4550–4559. [Google Scholar] [CrossRef]
  69. Crudden, C.M.; Allen, D.P. Stability and reactivity of N-heterocyclic carbene complexes. Co-ord. Chem. Rev. 2004, 248, 2247–2273. [Google Scholar] [CrossRef]
  70. Huang, S.; Zhang, W.; Liu, T.; Wang, K.; Qi, X.; Zhang, J.; Zhang, Q. TowardsN-Alkylimidazole Borane-based Hypergolic Fuels. Chem. – Asian J. 2016, 11, 3528–3533. [Google Scholar] [CrossRef]
  71. Santini, C.; Pellei, M.; Gioia Lobbia, G.; Papini, G. Synthesis and properties of poly(pyrazolyl)borate and related boron-centered scorpionate ligands. Part A: Pyrazole-based systems. Mini-Rev. Org. Chem. 2010, 7, 84–124. [Google Scholar] [CrossRef]
  72. Pellei, M.; Lobbia, G.G.; Papini, G.; Santini, C. Synthesis and Properties of Poly(pyrazolyl)borate and Related Boron-Centered Scorpionate Ligands. Part B: Imidazole-, Triazole- and Other Heterocycle-Based Systems. Mini-Reviews Org. Chem. 2010, 7, 173–203. [Google Scholar] [CrossRef]
  73. Fränkel, R.; Birg, C.; Kernbach, U.; Habereder, T.; Noth, H.; Fehlhammer, W.P. A Homoleptic Carbene–Lithium Complex. Angew. Chem., Int. Ed. 2001, 40, 1907–1910. [Google Scholar] [CrossRef]
  74. Frankel, R.; Kernbach, U.; Bakola-Christianopoulou, M.; Plaia, U.; Suter, M.; Ponikwar, W.; Noth, H.; Moinet, C.; Fehlhammer, W.P. Homoleptic carbene complexes. Part VIII. Hexacarbene complexes. J. Organomet. Chem. 2001, 530–545. [Google Scholar] [CrossRef]
  75. Fränkel, R.; Kniczek, J.; Ponikwar, W.; Noth, H.; Polborn, K.; Fehlhammer, W.P. Homoleptic carbene complexes: Part IX. Bis(imidazolin-2-ylidene-1-yl)borate complexes of palladium(II), platinum(II) and gold(I). Inorg. Chim. Acta 2001, 312, 23–39. [Google Scholar]
  76. Nieto, I.; Bontchev, R.P.; Smith, J.M. Synthesis of a Bulky Bis(carbene)borate Ligand – Contrasting Structures of Homoleptic Nickel(II) Bis(pyrazolyl)borate and Bis(carbene)borate Complexes. Eur. J. Inorg. Chem. 2008, 2008, 2476–2480. [Google Scholar] [CrossRef]
  77. Biffis, A.; Lobbia, G.G.; Papini, G.; Pellei, M.; Santini, C.; Scattolin, E.; Tubaro, C. Novel scorpionate-type triscarbene ligands and their silver and gold complexes. J. Organomet. Chem. 2008, 693, 3760–3766. [Google Scholar] [CrossRef]
  78. Papini, G.; Bandoli, G.; Dolmella, A.; Lobbia, G.G.; Pellei, M.; Santini, C. New homoleptic carbene transfer ligands and related coinage metal complexes. Inorg. Chem. Commun. 2008, 11, 1103–1106. [Google Scholar] [CrossRef]
  79. Papini, G.; Pellei, M.; Lobbia, G.G.; Burini, A.; Santini, C. Sulfonate- or carboxylate-functionalized N-heterocyclic bis-carbene ligands and related water soluble silver complexes. Dalton Trans. 2009, 6985. [Google Scholar] [CrossRef]
  80. Giorgetti, M.; Aquilanti, G.; Pellei, M.; Gandin, V. The coordination core of Ag( i ) N-heterocyclic carbene (NHC) complexes with anticancer properties as revealed by synchrotron radiation X-ray absorption spectroscopy. J. Anal. At. Spectrom. 2014, 29, 491–497. [Google Scholar] [CrossRef]
  81. Pellei, M.; Gandin, V.; Marinelli, M.; Marzano, C.; Yousufuddin, M.; Dias, H.V.R.; Santini, C. Synthesis and Biological Activity of Ester- and Amide-Functionalized Imidazolium Salts and Related Water-Soluble Coinage Metal N-Heterocyclic Carbene Complexes. Inorg. Chem. 2012, 51, 9873–9882. [Google Scholar] [CrossRef]
  82. Gandin, V.; Pellei, M.; Marinelli, M.; Marzano, C.; Dolmella, A.; Giorgetti, M.; Santini, C. Synthesis and in vitro antitumor activity of water soluble sulfonate- and ester-functionalized silver(I) N-heterocyclic carbene complexes. J. Inorg. Biochem. 2013, 129, 135–144. [Google Scholar] [CrossRef]
  83. Maria, L.; Paulo, A.; Santos, I.C.; Santos, I.; Kurz, P.; Spingler, B.; Alberto, R. Very Small and Soft Scorpionates: Water Stable Technetium Tricarbonyl Complexes Combining a Bis-agostic (k3-H, H, S) Binding Motif with Pendant and Integrated Bioactive Molecules. J. Am. Chem. Soc. 2006, 128, 14590–14598. [Google Scholar] [CrossRef]
  84. Lu, D.; Tang, H. Theoretical survey of the ligand tunability of poly(azolyl)borates. Phys. Chem. Chem. Phys. 2015, 17, 17027–17033. [Google Scholar] [CrossRef]
  85. Pellei, M.; Papini, G.; Lobbia, G.G.; Ricci, S.; Yousufuddin, M.; Dias, H.V.R.; Santini, C.; Dias, H.V.R. Scorpionates bearing nitro substituents: Mono-, bis- and tris-(3-nitro-pyrazol-1-yl)borate ligands and their copper(i) complexes. Dalton Trans. 2010, 39, 8937. [Google Scholar] [CrossRef]
  86. Dias, H.V.R.; Alidori, S.; Lobbia, G.G.; Papini, G.; Pellei, M.; Santini, C.; Dias, H.V.R. Small Scorpionate Ligands: Silver(I)-Organophosphane Complexes of 5-CF3-Substituted Scorpionate Ligand Combining a B−H··Ag Coordination Motif. Inorg. Chem. 2007, 46, 9708–9714. [Google Scholar] [CrossRef] [PubMed]
  87. Meisters, M.; VandeBerg, J.T.; Cassaretto, F.P.; Posvic, H.; Moore, C.E. Studies in the tetraarylborates: Part V. The influence of substituents on the stability of tetraarylborates. Anal. Chim. Acta 1970, 49, 481–485. [Google Scholar] [CrossRef]
  88. Bakshi, P.K.; Linden, A.; Vincent, B.R.; Roe, S.P.; Adhikesavalu, D.; Cameron, T.S.; Knop, O. Crystal chemistry of tetraradial species. Part 4. Hydrogen bonding to aromatic π systems: Crystal structures of fifteen tetraphenylborates with organic ammonium cations. Can. J. Chem. 1994, 72, 1273–1293. [Google Scholar] [CrossRef] [Green Version]
  89. Belanger, S.; Beauchamp, A.L. (1-Methylimidazole- N 3 )triphenylboron. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1998, 54, IUC9800057. [Google Scholar] [CrossRef]
  90. Bélanger-Chabot, G.; Kaplan, S.M.; Deokar, P.; Szimhardt, N.; Haiges, R.; Christe, K.O. Synthesis and Characterization of Nitro-, Trinitromethyl-, and Fluorodinitromethyl-Substituted Triazolyl- and Tetrazolyl-trihydridoborate Anions. Chem. - A Eur. J. 2017, 23, 13087–13099. [Google Scholar] [CrossRef]
  91. Ridlen, S.G.; Kulkarni, N.; Dias, H.V.R. Monoanionic, Bis(pyrazolyl)methylborate [(Ph3B)CH(3,5-(CH3)2Pz)2)]−as a Supporting Ligand for Copper(I)-ethylene, cis-2-Butene, and Carbonyl Complexes. Inorg. Chem. 2017, 56, 7237–7246. [Google Scholar] [CrossRef] [PubMed]
  92. Good, C.D.; Ritter, D.M. Alkenylboranes. II. Improved Preparative Methods and New Observations on Methylvinylboranes. J. Am. Chem. Soc. 1962, 84, 1162–1166. [Google Scholar] [CrossRef]
  93. Winkelmann, O.H.; Navarro, O. Microwave-Assisted Synthesis of N-Heterocyclic Carbene- Palladium(II) Complexes. Adv. Synth. Catal. 2010, 352, 212–214. [Google Scholar] [CrossRef]
  94. Curran, D.P.; Solovyev, A.; Brahmi, M.M.; Fensterbank, L.; Malacria, M.; Lacôte, E. Synthesis and Reactions of N-Heterocyclic Carbene Boranes. Angew. Chem. Int. Ed. 2011, 50, 10294–10317. [Google Scholar] [CrossRef]
  95. Wrackmeyer, B. Nuclear Magnetic Resonance Spectroscopy of Boron Compounds Containing Two-, Three- and Four-Coordinate Boron. Annual Reports on NMR Spectroscopy 1988, 20, 61–203. [Google Scholar] [CrossRef]
  96. Kronig, S.; Theuergarten, E.; Daniliuc, C.; Jones, P.G.; Tamm, M. Anionic N-Heterocyclic Carbenes That Contain a Weakly Coordinating Borate Moiety. Angew. Chem. Int. Ed. 2012, 51, 3240–3244. [Google Scholar] [CrossRef] [PubMed]
  97. Kolychev, E.L.; Kronig, S.; Brandhorst, K.; Freytag, M.; Jones, P.G.; Tamm, M. Iridium(I) Complexes with Anionic N-Heterocyclic Carbene Ligands as Catalysts for the Hydrogenation of Alkenes in Nonpolar Media. J. Am. Chem. Soc. 2013, 135, 12448–12459. [Google Scholar] [CrossRef]
  98. Liu, J.; Chen, J.; Zhao, J.; Zhao, Y.; Li, L.; Zhang, H. A Modified Procedure for the Synthesis of 1-Arylimidazoles. Synthesis 2003, 17, 2661–2666. [Google Scholar] [CrossRef]
  99. Sheldrick, G.M. A short history ofSHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2007, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  100. Dolomanov, O.; Bourhis, L.J.; Gildea, R.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Scheme 1. Structure of carbenes based on the imidazol-2-ylidene moieties: (a) with alkyl and aryl substituents; (b) with a main-group element substituent; (c) with a borate moiety substituent.
Scheme 1. Structure of carbenes based on the imidazol-2-ylidene moieties: (a) with alkyl and aryl substituents; (b) with a main-group element substituent; (c) with a borate moiety substituent.
Molecules 25 03184 sch001
Scheme 2. Structure of: (a) imidazolium trihydridoborate species; (b) bis(imidazolium)borate species; (c) imidazolium triethylborate species; (d) 2-substituted imidazolylborate species; (e) imidazolium and benzimidazolium triarylborate species; (f) ring-closed imidazolium and benzimidazolium triarylborate species; (g) imidazaboles.
Scheme 2. Structure of: (a) imidazolium trihydridoborate species; (b) bis(imidazolium)borate species; (c) imidazolium triethylborate species; (d) 2-substituted imidazolylborate species; (e) imidazolium and benzimidazolium triarylborate species; (f) ring-closed imidazolium and benzimidazolium triarylborate species; (g) imidazaboles.
Molecules 25 03184 sch002
Scheme 3. Chemical structures of (N-(alkyl/aryl)imidazolium)borates 14.
Scheme 3. Chemical structures of (N-(alkyl/aryl)imidazolium)borates 14.
Molecules 25 03184 sch003
Scheme 4. Synthesis of: (a) imidazolium trihydridoborates 1 and 2; (b) imidazolium triphenyborates 3 and 4.
Scheme 4. Synthesis of: (a) imidazolium trihydridoborates 1 and 2; (b) imidazolium triphenyborates 3 and 4.
Molecules 25 03184 sch004
Figure 1. Molecular structure of (HImBn)BH3 (1). Selected bond distances (Å) and angles (˚): N1-C1 1.343(2), N2-C1 1.323(2), N2-B1 1.587(2), N1-C4 1.474(2), N1-C1-N2 110.29(14), C1-N2-B1 126.58(14).
Figure 1. Molecular structure of (HImBn)BH3 (1). Selected bond distances (Å) and angles (˚): N1-C1 1.343(2), N2-C1 1.323(2), N2-B1 1.587(2), N1-C4 1.474(2), N1-C1-N2 110.29(14), C1-N2-B1 126.58(14).
Molecules 25 03184 g001
Figure 2. Molecular structure of (HImMes)BH3 (2). There are two chemical similar but crystallographically different molecules of (HImMes)BH3 in the asymmetric unit but only one is shown here. Selected bond distances (Å) and angles (˚): N1-C1 1.3442(14), N2-C1 1.3207(14), N2-B1 1.5836(16), N1-C4 1.4465(14), N1-C1-N2 110.36(10), C1-N2-B1 127.91(10).
Figure 2. Molecular structure of (HImMes)BH3 (2). There are two chemical similar but crystallographically different molecules of (HImMes)BH3 in the asymmetric unit but only one is shown here. Selected bond distances (Å) and angles (˚): N1-C1 1.3442(14), N2-C1 1.3207(14), N2-B1 1.5836(16), N1-C4 1.4465(14), N1-C1-N2 110.36(10), C1-N2-B1 127.91(10).
Molecules 25 03184 g002
Scheme 5. Microwave-assisted synthesis of Compound 5.
Scheme 5. Microwave-assisted synthesis of Compound 5.
Molecules 25 03184 sch005
Scheme 6. Rearrangement species (B,C) by isomerization or dimerization of the NHC-borate form (A).
Scheme 6. Rearrangement species (B,C) by isomerization or dimerization of the NHC-borate form (A).
Molecules 25 03184 sch006

Share and Cite

MDPI and ACS Style

Pellei, M.; Vallesi, R.; Bagnarelli, L.; Dias, H.V.R.; Santini, C. Syntheses and Reactivity of New Zwitterionic Imidazolium Trihydridoborate and Triphenylborate Species. Molecules 2020, 25, 3184. https://doi.org/10.3390/molecules25143184

AMA Style

Pellei M, Vallesi R, Bagnarelli L, Dias HVR, Santini C. Syntheses and Reactivity of New Zwitterionic Imidazolium Trihydridoborate and Triphenylborate Species. Molecules. 2020; 25(14):3184. https://doi.org/10.3390/molecules25143184

Chicago/Turabian Style

Pellei, Maura, Riccardo Vallesi, Luca Bagnarelli, H. V. Rasika Dias, and Carlo Santini. 2020. "Syntheses and Reactivity of New Zwitterionic Imidazolium Trihydridoborate and Triphenylborate Species" Molecules 25, no. 14: 3184. https://doi.org/10.3390/molecules25143184

Article Metrics

Back to TopTop