Next Article in Journal
Unveiling the Dynamic Evolution of Catalytic Sites in N-Doped Leaf-like Carbon Frames Embedded with Co Particles for Rechargeable Zn–Air Batteries
Next Article in Special Issue
Gold/DNA-Cu2+ Complex Nanozyme-Based Aptamer Lateral Flow Assay for Highly Sensitive Detection of Kanamycin
Previous Article in Journal
Ni–Doped Pr0.7Ba0.3MnO3−δ Cathodes for Enhancing Electrolysis of CO2 in Solid Oxide Electrolytic Cells
Previous Article in Special Issue
Hybrid Hydroxyapatite–Metal Complex Materials Derived from Amino Acids and Nucleobases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards Nickel–NHC Fluoro Complexes—Synthesis of Imidazolium Fluorides and Their Reactions with Nickelocene

by
Siobhan S. Wills
1,
Corinne Bailly
2 and
Michael J. Chetcuti
3,*
1
School of Chemistry, University of New South Wales (UNSW), Sydney, NSW 2052, Australia
2
Fédération de Chimie Le Bel UAR2042, BP20296, 1 rue Blaise Pascal, 67008 Strasbourg, France
3
Organometallic Research Group (OMCAT), LIMA CNRS UMR 7042, ECPM, University of Strasbourg, 25 rue Becquerel, 67087 Strasbourg, France
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(18), 4493; https://doi.org/10.3390/molecules29184493
Submission received: 22 July 2024 / Revised: 16 September 2024 / Accepted: 19 September 2024 / Published: 21 September 2024
(This article belongs to the Special Issue Advances in Coordination Chemistry 2.0)

Abstract

:
While hundreds of complexes of the general formula [Ni(η5-C5H5)(NHC)(X)] exist (NHC = a N-heterocyclic carbene, X = Cl, Br, I), none is yet known with X = F. We attempted to prepare such a species by reacting nickelocene with imidazolium fluorides. Three imidazolium fluorides (ImH)+ F [Im = (N,N′-bis-(R)-imidazolium: 1a, IMe, R = Me; 1b, IMes, R = 2,4,6-trimethylphenyl; 1c, IPr, R = 2,6-diisopropylphenyl)] were prepared and characterized by spectroscopic methods. In addition, the salts 1b [(IMesH)+ F] and 1c [(IPrH)+ F] were subjected single-crystal X-ray diffraction experiments. The reactions of these imidazolium fluorides with nickelocene did not lead to [Ni(η5-C5H5)(NHC)(F)] species. Instead, the reaction of 1a [(IMeH)+ F] and 1b [(IMesH)+ F] with nickelocene led to the salt 2 [Ni(η5-C5H5)(IMe)2]+ F and to the square planar complex 3a trans-[NiF2(IMes)2] respectively. Both complexes were characterized spectroscopically and by single crystal X-ray diffraction. All four X-ray diffraction studies reveal hydrogen bonding and hydrogen interactions with the F atom or anion, and in some cases with solvent molecules of crystallization, and these phenomena are all discussed. Complex 2, in particular, exhibited a wide range of interesting H-bonded interactions in the solid state. Complexes 2 and 3a were tested as catalysts for Suzuki–Miyaura coupling but were not promising: complex 2 was inactive, and while 3a did indeed catalyze the reaction, it gave widely diverging results owing to its instability in solution.

Graphical Abstract

1. Introduction

Organofluorine compounds are key molecules in the pharmaceutical and agrochemical industries. In 2020, it was estimated that ≈20% of commercial pharmaceuticals are fluoro-pharmaceuticals, [1] including best-sellers such as the anti-cholesterol drug Lipitor, the anti-depressive Prozac (fluoxetine) and the antibiotic Cipro (ciprofloxacin), and the proportion has increased since then. Nevertheless, not many transition-metal catalysts are known that can introduce fluorine into organic molecules. Fluorine is the most electronegative element and (after He) the smallest. Established synthetic techniques for the introduction of halogens into organometallic compounds are not generally applicable to fluorine [2], as carbon forms its strongest single bond with fluorine [3], and this inert C–F bond makes oxidative addition of RF compounds to organometallic reagents more difficult. Fluoride-for-halide exchange reactions are thus more difficult; furthermore, the ease of reaction of F ions with glass [4] potentially complicates its reactions. For these reasons, fluorine organometallics are less common, and this area is “still in its early stages” [5]. Such species are more common for the early transition metals [6,7], though this is slowly changing [8,9,10,11,12].
Our group has been investigating the syntheses and the catalytic potential of nickel complexes with N-heterocyclic carbene (NHC) ligands over the past two decades and we have prepared and extensively investigated nickel NHC complexes of the type [Ni(Cp)(NHC)(X)] (Cp = η5-C5H5 or η5-C5Me5, X = Cl, Br, I) [13,14]. Most complexes are easy to prepare, as treatment of nickelocene with the appropriate imidazolium salt in a solvent, under reflux or via microwave synthesis, leads directly to [Ni(Cp)(NHC)(X)] species in satisfactory to excellent yields [13,14,15]. These complexes are catalytically active for a wide range of reactions that have been investigated by us and others, including Suzuki–Miyaura coupling reactions [16,17,18,19,20,21,22], α-ketone arylation [23,24,25], hydrosilylation [14,26,27,28] and hydroboration [29,30,31,32]. We were interested to see if [Ni(Cp)(NHC)(F)] complexes were accessible by using the lesser-known imidazolium fluorides, in an extension of the just discussed method of synthesis.
Nickel NHC species with fluoride ligands may be implicated in cross coupling reactions between aryl Grignard reagents and aryl halides, as NiF2 carries out the cross coupling in the presence of imidazolium chloride or tetrafluoroborate salts [33]. In general, however, there are not many isolated nickel NHC complexes with fluoride ligands [9,10,34], so this remains an area that has not been thoroughly investigated.
Imidazolium salts of Cl, Br and I are well known and are readily accessible via many synthetic routes. One involves direct alkylation of imidazole using the appropriate alkyl halides in the presence of a base in a classic nucleophilic substitution reaction. For bulkier N-substituted imidazolium salts, an alternative two-step process is used, starting from glyoxal and a primary amine, followed by ring closure of the di-imine produced with formaldehyde and a Lewis acid source. This method often leads to imidazolium chlorides, and both methods have been well investigated [35,36,37]. The Cl can be replaced with Br or I anions by adding KBr and KI [38]. Yields are good to excellent and do not require expensive reagents or anerobic synthetic conditions. However, these synthetic methods do not work for the synthesis of imidazolium fluorides, presumably because of the significant strength of the C–F bonds in the alkyl fluoride.
The first synthetic report of a N,N′-bis-alkyl-imidazolium fluoride appeared in 2012. It involved halide exchange between a N,N′-bis-alkyl imidazolium iodide and silver fluoride [39]. The reaction is driven by the formation of insoluble AgI and was only reported for R = Me and R’ = Me, Et or Pr. Incidentally, similar reactions have been used to prepare metal-fluoro complexes; for example, Pilon and Grushin reported the synthesis of a series of [Pd(PPh3)2(R)F] complexes that were synthesized this way [40].
Other reports of imidazolium fluorides have since appeared in the literature. While these were sometimes prepared for their properties and use as ionic liquids, over the last ten years, imidazolium fluorides have found increasing applications in organic chemistry towards the nucleophilic fluorination of various substrates [41,42]. Indeed, it is noted, in an article that briefly reviews the field and applies imidazolium fluorides towards nucleophilic fluorination [43], that “the reagent PhenoFluor [44] and its air-stable successor AlkylFluor [45] … seem to be the reagents of choice for deoxyfluorination of phenols and aliphatic alcohols, respectively”. Nevertheless, only a handful of single-crystal X-ray diffraction studies of such fluorides have been reported [39,46]. In this work, we describe the isolation and characterization of three imidazolium fluorides, the structural determination of two of them, and their subsequent reactions with nickelocene, which in fact do not appear to yield [Ni(Cp)(NHC)(F)] species.

2. Results and Discussion

2.1. Synthesis of the Imidazolium Fluoride Salts 1a1c

The salt N,N′-bis-methylimidazolium fluoride, [(IMeH)+ F] 1a, which, incidentally, has been the subject of a theoretical study on its stability [47], was prepared following the procedure reported by Xiao [39] and was obtained in good yield. We attempted to prepare other imidazolium fluorides by an adaptation of this method, using the exchange reaction with PbF2 rather than with AgF. The lead salt is not light sensitive and is significantly cheaper. However, this did not work well, as the halide exchange was not quantitative, so we resorted to AgF as the source of fluoride in the exchange reactions for the syntheses of two other imidazolium, fluorides, [(IMesH)+ F], 1b, and [(IPrH)+ F], 1c (IMesH = N,N′-bis-(2,4,6-trimethylphenyl)imidazolium); IPrH = N,N′-bis-(2,6-diisopropylphenyl)imida-zolium). 1c was prepared by reacting the free carbene N,N′-bis-(2,6-diisopropyl-pheny)imidazol-2-ylidene with KHF2 or HF [48], but we preferred avoiding possible HF contact. Our reactions were conducted in glassware wrapped in aluminum foil to minimize light exposure to the light-sensitive silver compounds. The syntheses of these aryl imidazolium fluorides were not clean, as colloidal AgCl tended to form when [(IMesH)+ Cl] was treated with AgF. However, careful control of reaction conditions, including the slow addition of a solution of AgF led to relatively pure [(IMesH)+ F]. The other salts were similarly prepared. These reactions are outlined in Scheme 1.
The cations in 1b and 1c are the same as in the chlorides [(IMesH)+ Cl] and [(IPrH)+ Cl], but there are significant spectral differences between the F and Cl salts due to H- bonding of the C(1)–H proton to the fluoride in the fluorides. Indeed, the C(1)–H proton, usually observed at δ = 9–11 ppm in other imidazolium halides, is usually not seen in 1H NMR spectra of 1b and 1c though occasionally a very weak broad peak is observed in the 11.5 ppm range. This phenomenon has previously been reported for 1c [46]. Significant signal broadening could arise from H F hydrogen bonding or through H–D exchange due to the presence of traces of water in solution, which is perhaps more likely.
The absence of this signal could conceivably arise from the formation of silver NHC complexes of stoichiometry F–Ag–NHC, but this was refuted by the 19F NMR spectra of both species, which exhibit peaks at −131.0 and 131.5 ppm for 1b and 1c, not far from the signal at −120.3 ppm observed for free fluoride ions. In addition, there is no evidence of 107Ag–F or 109Ag–F coupling, and the C–H–N microanalysis negates the presence of silver in these species.
The 13C NMR of the imidazolium fluoride salts does not exhibit a signal for the C(1)–H carbon atom, possibly due to its coupling to the two flanking quadrupolar nitrogen atoms. All other carbon atoms show their 13C NMR peaks in their expected chemical shift range, and there is no evidence of any signals coupling to silver. Furthermore, IR spectra of 1b and 1c exhibit strong ν(H F) peaks at 3640 and 3650 cm1 respectively.

2.2. X-ray Diffraction Studies on the Imidazolium Fluoride Salts 1b and 1c

The H-bonding was confirmed by X-ray diffraction studies on single crystals of 1b and 1c, obtained from concentrated CH2Cl2 solutions. Their structures are shown in Figure 1 and Figure 2. In both cases, there are disordered solvent molecules (H2O in 1b, CH2Cl2 molecules in 1c). The structure of 1c has been previously reported [46], but with acetonitrile in the lattice (as opposed to CH2Cl2 in ours), and the previously reported cell parameters are (unsurprisingly) different from those we found. Neither structure is of high quality; there is significant disorder in both, that could not be properly modeled. Nevertheless, we believe that they are useful and should appear in the literature, as very few imidazolium fluoride structures are known. Key data for both salts are collected in Table 1 with relevant data for 1a [39] included as well. Figure 1 and Figure 2 show the molecular structures of the salts. In 1b, the fluoride atom is disordered symmetrically about two positions.
A feature of interest in both structures are the short C(1)Imid–H F distances of 1.85 and 1.84 Å in 1b and 1c respectively, indicative of strong hydrogen bonding. The previously reported structure of 1c [46] showed a shorter corresponding distance of 1.72 Å, but in all cases, these distances have relatively high estimated standard deviations (esds). The structure of 1a [39] also shows short C(1)Imid–H F distances of 1.92 and 2.04 Å. There are many other longer-range H-interactions in both 1b and 1c; those of 1b are shown in Figure 3.
In 1b, the meta-H-atom of the mesityl ring, one of the mesityl CH3 hydrogen atoms and the two other C–H hydrogen atoms of the imidazolium ring have longer interactions in the 2.59 to 2.78 Å range with F ions. There are disordered water molecules, and their O-atoms have shorter interactions of 2.38, 2.28 and 2.53 Å with (respectively) hydrogen atoms of two ortho-CH3 groups and one of the imidazolium –CH=CH– atoms that probably help stabilize the structure. The short F O distances of 2.474 and 2.537 Å suggest that there are OHF interactions as well, but the water molecules’ hydrogen atoms could not be located or modeled.
In compound 1c, the situation is interesting, as, in addition to the short imidazolium H(2) F interaction, mentioned earlier, each para-C–H bond of the two aromatic rings on each imidazolium moiety is linked in a symmetric fashion with the fluoride ion (C–Hpara  F = 2.41 Å) as shown in Figure 4. The cations are thus arranged in long chains, effectively linked together via these long H–bonds. The fluoride ion here is in a distorted, but close to trigonal, planar geometry with the C–Hpara  F C–Hpara angle being 139°: the two other C–Hpara  F C(2)–H angles are equivalent and equal to 109° so as to have the fluoride anions surrounded in a very close to trigonal geometry (the sum of the three angles equals 356°). Some of the chlorine atoms of the CH2Cl2 solvent molecules also have long interactions with some hydrogen atoms: they range from 2.94 to 3.15 Å, though these are probably of marginal significance. There are short FH2CCl2 interactions of 2.39 Å with one of the dichloromethane hydrogen atoms with the fluoride ion. While occasionally fluoride anions can exhibit short interactions with other halogens [11,49], no F Cl2CH2 short contacts were observed here.

2.3. Reactions of Imidazolium Fluoride 1a with Nickelocene and an X-ray Diffraction Study of [Ni(Cp)(IMe)2]+ F, 2

The reaction of a wide spectrum of imidazolium halide salts (halide = Cl, Br or I) with nickelocene has been shown by us and others to lead to the half-sandwich (or two-legged piano-stool) molecules of formula [Ni(Cp)(NHC)(X)], and many of these well-known species have been structurally characterized by us [13,14,16,17] and others [50,51,52,53,54,55]. However, no fluoro-complexes of formula [Ni(Cp)(NHC)(F)] have been reported, and we were interested in exploring the reaction of nickelocene with imidazolium fluorides 1a1c to try to prepare such a species (Scheme 2).
The reaction was first attempted with 1a. A relatively rapid color change was quickly observed, with the deep forest green of the nickelocene solution giving way to a cherry red, then purple color within 10 min. More slowly, a dark, almost black, color developed during the total 41 h reaction time. This contrasts with the reaction of the N,N′-bis-methylimidazolium iodide with nickelocene, which is much slower.
After workup (see experimental section), red crystals of 2 were obtained and they exhibited no ν(H F) in their IR spectrum. 1H NMR data indicated the presence of a Cp and two NHC ligands and suggested that 2 was probably the bis-NHC complex [Ni(Cp)(IMe)2]+ F. The 13C NMR spectrum corroborated the 1H NMR data and, in addition, exhibited a signal for the carbene carbon atoms at characteristic downfield peak at 162.6 ppm, in line with the signals observed for NHC carbene carbon atoms in other similar bis-NHC complexes of nickel, which we have previously observed in the 158–168 ppm range [56]. The 19F NMR showed a single peak at −162 ppm, not very close to the signal for free F in chloroform (which appears at −120 ppm). The F anion is thus likely to be closely associated with the complex in solution, as some sort of ion pair.
The structure of [Ni(Cp)(IMe)2]+ F, 2 (Figure 5) was confirmed by single-crystal X-ray diffraction on a crystal grown by slow diffusion of Et2O into a CH2Cl2 solution of 2. Key structural parameters are collected in Table 2. The cation forms a 2-legged piano stool, whose structure is similar to those of other [NiCp(NHC)X]+ (X = Cl, Br, I, NHC’) species described by us [13,14,16,17,56] and others [15,50,51,52,53,54,55]. The Ccarb1–Ni–Ccarb2 angle is 94.5°, which is in the range of what is seen in other complexes of this geometry (96.9° is observed for this angle in [Ni(Cp)(IMes)(IMe)]+) [56]. As is also commonly observed in [Ni(Cp)(NHC)2]+ salts [15,56], if one takes the centroid of the Cp ligand and the two carbene carbon atoms, the nickel is in the center of a very close to trigonal planar geometry [the sum of the three angles subtended by the nickel atom to these three centers is 359.9°].
There are many significant supramolecular H-interactions in the structure that lead to a three-dimensional network in the crystal. While such phenomena were also observed in the structures of the two imidazolium fluorides described earlier in this manuscript, the hydrogen interactions are much more extensive here. The salt crystallizes in the high symmetry trigonal space-group R3, and an imposed three-fold rotational axis relates each cation to the other two that are ordered around it, via this axis. There are also three disordered water molecules incorporated in the structure. Short H F distances of 2.28 and 2.49 Å are seen for a fluoride ion that is effectively bridging two C(H)=C(H) hydrogen atoms from two different carbene ligands on two different cations, via this H-bond interaction. One of the two C(H)=C(H) atoms of the other carbene ligand that is not involved in a H F interaction instead shows two short interactions of 2.45 and 2.58 Å with two oxygen atoms of the water molecules present. There are also short interactions between some of the relatively acidic Cp hydrogen atoms with the fluoride ions (2.41 Å) and with the oxygen atoms of the water (2.58 Å). Even some of the sp3-carbon bonded hydrogen atoms of the dimethylated imidazolyidene methyl groups are involved with these electronegative elements, and HMe  F distances of 2.68 Å are seen for HF distances and 2.72, 2.97 and 2.65 Å for HO distances. A water molecule’s hydrogen atom is also strongly interacting with the fluoride ion (FH–O–H = 1.72Å). These interactions are collected in Figure 6. Finally, and remarkably, six hydrogen atoms, from different N–CH3 groups, are pseudo-octahedrally coordinated to a (water) oxygen atom with H O distances all equivalent to 2.81 Å in a hexagonal channel (Figure 7).
The base-assisted deprotonation of the imidazolium salt to the free carbene was initially hypothesized to be the first step in the synthesis of 2. We have previously reported the protonation of a η5-C5H5 group with HCl, and the subsequent loss of free cyclopentadiene in an unusual reaction of an 18-electron nickel complex [57], so such a reaction mechanism is not unprecedented. In addition, nickelocene is formally a 20-electron complex which makes the Cp ligand much more reactive and labile. As 60% yields of 2 were obtained using a NiCp2: 1a ratio of 1:2, Cp is unlikely to be the base, as a maximum yield of 50% would be expected. Furthermore, with a NiCp2: 1a ratio of 1:4, 96% yield could be obtained. The HF produced could protonate one of the nickel cyclopentadienyl ligands to give free cyclopentadiene, and the process would then be repeated for the coordination of the second carbene.

2.4. Reactions of the Imidazolium Fluorides 1b and 1c with Nickelocene and an X-ray Diffraction Study of trans-[NiF2(IMes)2], 3a

Cationic nickel complexes of stoichiometry [Ni(Cp)(NHC1)(NHC2)]+ were reported by us and by others as previously mentioned [15,56]. The two NHC ligands can be different from each other, but they all share one characteristic feature: the two NHC ligands are relatively small. Indeed, our failed attempts to prepare the complex [Ni(Cp)(IMes)2]+ led us to discover an interesting base-promoted C–H activation of acetonitrile [58].
The reaction of 1b, [(IMesH]+ F] with nickelocene thus was unlikely to lead to [Ni(Cp)(IMes)2]+ cations and this, we believed, increased the odds of preparing a complex of formula [Ni(Cp)(NHC)(F)]. The reaction of two equivalents of 1b with nickelocene was rapid: within 10 min of mixing, the deep green color of nickelocene gave way to a red color. Complex 3a was isolated as orange-red powder in relatively poor yield (17%) after workup. However, no cyclopentadienyl signals were observed in the 1H NMR spectrum. There were resonances consistent with one (or more than one equivalent) IMes carbene ligand that was presumably coordinated to the nickel. The spectrum was not clean and appeared to show at least two products as well as free 1b; the quantity of the second minor product increased with time. Reactions of 1b and 1c with nickelocene are shown in Scheme 2.
Clearly, 3a is not stable in solution, as decomposition products rapidly formed within minutes. Despite repeated attempts, we were unable to characterize any of these other products. A few crystals of 3a could be obtained and a 1H NMR spectrum of these dissolved crystals was obtained at low temperature
The 1H NMR spectrum showed signals that corresponded to the major product in the crude reaction mixture, but it had broad peaks, possibly suggesting a paramagnetic impurity (see Supplementary Materials); the sample degraded at room temperature affording 1b and uncharacterizable decomposition products. It was not possible to obtain a clean (and non-broadened) spectrum of 3a, even at low temperatures. The 13C NMR spectrum was also messy and no carbene carbon signal could be seen. A 19F NMR spectrum exhibited a signal at −157 ppm and a small unidentified peak at −113 ppm. As pointed out by a thorough reviewer, fluorides often react with borosilicate glass with its traces of adsorbed water to give [SiF5] and perhaps [SiF6]2− and [BF4] [4]. This could account for the instability of 3a in solution and the unidentified peak as a decomposition product.
The first reported 19F chemical shift of [SiF5] was given as +136.7 ppm [59] and later at +136.0 [60]. Discrepancies in 19F NMR chemical shift values have been the subject of a recent publication [61]; furthermore, older publications may have a sign convention problem, so the quoted values may in fact be at negative ppm values. The 19F chemical shifts of [SiF5(H2O)] and [SiF6]2− are reported at −130.5 and −129.5 ppm, respectively [62], and that of [BF4] at −139.8 ppm) [63], so the unidentified peak cannot be attributed to any of them. However, it cannot be ruled out that the instability of the compound results from reactions with the borosilicate glass surface.
A single crystal X-ray diffraction study on an isolated crystal established that 3a is the square planar Ni(II) complex trans-[NiF2(IMes)2]. Complexes of generic formula [NiX2(NHC)2] (X = Cl, Br, I) are well known [64], but to our knowledge this is the first example in which X = F. The structure is shown in Figure 8; key structural parameters of 3a and of a few other square-planar Ni-NHC complexes that contain a single fluoride as a ligand are collected in Table 3.
Only a handful of Ni(NHC) complexes bearing fluoride ligands are known [9,10,25], and to our knowledge, this is the first example of a Ni-NHC complex with two fluoride groups coordinated to nickel. In complex 3a, the two halves of the molecule are equivalent via the crystallographically imposed symmetry from the high symmetry trigonal space group P 3 ¯ c1. The nickel coordination geometry is similar to those of square planar trans-[NiX2(NHC)2] complexes that have been characterized with other halide ligands [65], and to [trans-[Ni(NHC)2(aryl)(F)] complexes (shown in Table 1) [9,10]. The F–Ni–F angle is 178.3 (18)°, while the F–Ni–Ccarb angles are 90.79 (13) and 89.23 (13)°. These minor distortions are likely due to the steric pressure exerted by the relatively large IMes ligands. Furthermore, as is seen with all other structures presented in this paper, each fluorine atom has close contacts (six in all) with neighboring methyl hydrogen atoms that range from 2.466 to 2.659 Å (mean distance = 2.553 Å). The Ni–F distance of 1.816 (2) Å is significantly shorter than that observed for other complexes shown in Table 3, probably because of the less significant steric strain exerted by the smaller IMes ligand as compared to IPr. The mesitylene ligands on each carbene are not in the plane of the NHC ligand but are twisted at angles of 76 and 84° with respect to it.
As is seen with all other structures in this paper, the fluorine atoms have close contacts (six in all) with neighboring hydrogen atoms. These distances range from 2.46 to 2.65 Å (mean = 2.58 Å), and four of these (per fluorine) are intramolecular, coming from four different hydrogen atoms from four different ortho-methyl hydrogens on each mesityl group. The remaining two are intermolecular, one from an ortho-methyl hydrogen atom (at a distance of 2.47 Å) and the other from the meta-CH aromatic hydrogen atom (2.49 Å). The intramolecular H-interactions are shown in Figure 9.
Some evidence suggests that the analogous reaction of the imidazolium salt 1c, [(IPrH)+ F], with nickelocene affords the analogous square planar complex 3b, trans-[NiF2(IPr)2]. However, this species was not fully characterized. Despite repeated attempts, it could not be obtained pure and free of the imidazolium salt 1c. However, the 1H NMR spectrum of the impure product was obtained and is included in the Supplementary Materials together with its IR spectrum. Perhaps the increased steric footprint of the large IPr ligands contributes to this instability in solution.

2.5. Catalytic Tests

Preliminary catalytic studies were carried out on complexes 2 and 3a. The coupling reaction between phenylboronic acid and 4-bromoacetophenone was implemented at 90 °C in toluene with the addition of 3 mol% of either 2 or 3a as the catalyst and the addition of K3PO4 as a base (Scheme 3). After 1 h, an aliquot of the reaction mixture was taken and analyzed by 1H NMR using methods previously described [16,17,18].
Complex 2 did not lead to any coupling product, as only starting organic materials could be observed in the 1H NMR spectrum. This is not entirely surprising, as previous work in our labs has shown that [Ni(Cp)(NHC)2]+ X complexes have been ineffective in exhibiting catalytic activity towards Suzuki–Miyaura coupling [56]. This is probably due to the high stability of having two strongly σ-bonded NHC ligands on the nickel atom (an electronic effect) as well the overall steric protection of the nickel center, which is surrounded by two NHCs and a η5-C5H5 ligand. This prevents the coordination of starting materials and any possible ring-slippage of the η5-C5H5 group to provide a vacant coordination site, or reduction to catalytically active Ni(0) species.
The square planar complex 3a, trans-[NiF2(IMes)2] was catalytically active, but widely varying yields of coupling products were observed in three catalytic runs with conversion percentages of 29, 58 and 78%. Furthermore, as has already been noted, 3a decomposes in solution, and color changes from pale orange to light brown were observed in each catalytic run. No further catalytic experiments were performed.
The high chemical stability of 2 was shown by no observable degradation of the complex in its attempted protonation using 0.01 M HCl(aq) or 0.01 M HNO3(aq) when stirred at room temperature for 1 h with these acids. In both cases, no color changes were observed and only signals for 2 were observed after the acid exposure. Experimental details of the catalytic studies and on the inertness of 2 towards mineral acids are given in the “Materials and Methods” section below.

3. Materials and Methods

Reactions were carried out using standard Schlenk techniques, anaerobically under argon unless otherwise stated. All chemicals were purchased from the Sigma-Aldrich chemical company (Saint Quentin, France) or Fisher (Illkirch, France). Commercial compounds were used as received. Solvents were distilled over sodium/benzophenone (diethyl ether, thf, 1-2-dimethoxyethane), sodium (toluene) or CaH2 (pentane, dichloromethane. NMR spectra were recorded at ambient temperature on a Bruker FT 300 or 400 spectrometer (Karlsruhe, Germany) operating at 300.13 or 400 MHz (1H), 75.47 or 100.62 MHz (13C{1H}) and 282.38 MHz for 19F NMR spectra in CDCl3 unless otherwise stated. Chemical shifts were references to residual solvent peaks for 1H and 13C NMR spectra and to trifluorotoluene for the 19F NMR (set at −63.3 ppm). Chemical shifts (∂) and coupling constants (J) are in ppm and Hz, respectively. IR spectra were collected on a Nicolet 380 FT IR (Thermofisher, Schwerte, Germany) equipped with a diamond SMART ORBIT ATR accessory. Elemental analyses were performed by the Service Analysis Center at the University of Strasbourg. Nickelocene [66] and the salts N,N′-bis-methyl-imidazolium fluoride [39], N,N′-bis-methylimidazolium iodide [17,66] and both N,N′-bis-2,4,6-trimethylphenyl)imidazolium and N,N′-bis-2,6-diisopropylphenyl)imidazolium chlorides were prepared following their literature methods [36,37].
X-ray diffraction data for 1b, 1c, 2 and 3a were all collected on a Bruker APEX II DUO Kappa-CCD diffractometer (Karlsruhe, Germany) equipped with an Oxford Cryosystem liquid N2 device, using Mo-Kα radiation (λ = 0.71073 Å) at the University of Strasbourg structural facility. All structures were solved using SHELXS-97 [67,68]. Single crystals of 1b, 1c and 2 were obtained by crystallization at 4 °C from dichloromethane (1b, 1c) or by slow diffusion of Et2O into a dichloromethane solution of 2. Crystals of 3a were grown from a thf/pentane solution at room temperature for 12 h and then at 4 °C for 1 day. The crystal–detector distance was 38 mm. The cell parameters were determined (APEX2 software) [69] from reflections taken from three sets of six frames, each at 10 s exposure. The refinement and all further calculations were carried out using SHELXL-2019 [70]. The H-atoms were included in calculated positions and treated as riding atoms using SHELXL default parameters. The non-H atoms were refined anisotropically, using weighted full-matrix least-squares on F2. A semi-empirical absorption correction was applied using SADABS in APEX2 [55]; transmission factors: Tmin/Tmax = 0.8198/0.9265 for 3a, Tmin/Tmax = 0.6805/0.7456 for 1b, Tmin/Tmax = 0.6383/0.7456 for 1c, Tmin/Tmax = 0.6700/0.7456 for 2. For compounds 2 and 3b, the SQUEEZE instruction in PLATON [71] was applied. Residual electron density was assigned to one molecule of water present in the crystal for compound 2.

3.1. Syntheses and Spectroscopic Data

All imidazolium fluoride salts were prepared similarly, by an adaptation of Xiao’s method [39]. As the silver salts are photosensitive, glassware was protected with aluminum foil to prevent their photodecomposition.

3.2. N,N′-bis-(2,4,6-Trimethylphenyl)imidazolium Fluoride, [IMesH+ F], 1b

AgF (272 mg, 2.14 mmol) was dissolved in H2O (20 mL) and [IMesH+ Cl] (733 mg, 2.15 mmol) was added slowly. The mixture was stirred for 1 h and then refluxed for 1 h. The solution was then cooled in an ice-bath, filtered, and the water was removed under reduced pressure. The resulting solid was dried in vacuo overnight to give 1b (662 mg, 2.04 mmol, 95%) as a cream solid, mpt. 184 °C. 1H NMR (CDCl3): 11.30 (? CHF), 7.57 (2H, im, H 4, 5), 6.95 (4H, meta-H), 2.28 (6H, para-CH3), 2.08 (12H, ortho-CH3). 13C{1H}: 141.0 (ipso-CAr), 134.2 (ortho-CAr), 130.9 (para-CAr), 129.7 (meta-CAr), 124.3 (im C4, 5), 21.1 (para-CH3), 17.4 (ortho-CH3). 19F{1H} −131.0. IR: 3642 cm1 ν(HF). Anal: calcd. (expt.) for C21H25N2F.2H2O: C, 69.97 (69.48); H, 8.11 (7.80); N, 7.77 (7.43).

3.3. N,N′-bis-(2,6-Diisopropylphenyl)imidazolium Fluoride, [IPrH+ F], 1c

AgF (254 mg, 2.00 mmol) was dissolved in H2O (20 mL) and [IPrH+ Cl] (850 mg, 2.00 mmol) was added slowly. The mixture was stirred for 2 h and then filtered. The filtrate turned cloudy and was filtered a second time after heating. A milky suspension was still obtained, so it was heated for 2 h and then allowed to sit for 2 d and then filtered a third time. The now clear solution was evaporated under vacuum and deposited a cream-white solid 1c (497 mg, 1.21 mmol, 61%), mpt. 170–173°. 1H NMR (CDCl3): 7.94 (2H, im, H 4, 5), 7.53 (t, 2H, para-H, 3J = 8.0), 7.30 (d, 4H, meta-H, 3J = 8.1), 2.39 (septet, 4H, Me2CH, 3J = 6.9), 1.22 (d, 12H, Me, 3J = 6.3), 1.16 (d, 12H, Me, 3J = 6.9). 13C{1H}:145.3 (ortho-CAr), 131.9 (para-CAr), 130.9 (im C4, 5), 130.4 (ipso-CAr) (124.7 (meta-CAr), 29.1 (Me2CH), 24.6 (CH3), 23.8 (CH3). 19F{1H} −131.5. IR: 3651 cm1 ν(HF). Anal: calcd. (expt.) for C27H37N2F.CH2Cl2: C, 68.14 (72.69); H, 7.97 (8.87); N, 5.68 (6.06).

3.4. Synthesis of [Ni(Cp)(IMe)2]+ F, 2

A solution of nickelocene (384 mg, 2.03 mmol) in 1,2-dimethoxyethane (20 mL) was added to excess N,N′-bis-methylimidazolium fluoride (897 mg, 7.72 mmol) and the mixture was refluxed for 41h. The solution changed rapidly from dark green to deep red within 5 min and then darkened to purple and then to black as the rection proceeded. The solvent was then removed under vacuum and the deep red residue was dissolved in a minimum of toluene and filtered through a 3 × 3 cm Celite pad until the filtrate ran clear. The toluene was then evaporated, and the residue extracted with dichloromethane, again filtered through Celite, and the solvent was removed again under vacuum. A mixture of green and red microcrystals was obtained. Recrystallization from a dichloromethane/diethyl–ether mixture deposited red crystals of 2 (66 mg 0.215 mmol, 11% based on nickelocene). Mpt. (dec.) 130 °C. 1H NMR: 7.13 (4H, NCH), 5.41 (5H, C5H5), 3.94 (12H, Me). 13C NMR: 162.6 (NCN), 124.8 (NCH), 91.2 (C5H5), 38.9 (Me). 19F NMR: −131.50. Anal: calcd. (expt.) for C15H21N4NiF.2H2O: C, 48.55 (48.86); H, 6.32 (6.79); N, 15.10 (14.88).

3.5. Synthesis of 3a, trans-[NiF2(IMes)2]

A solution of nickelocene (277 mg, 1.47 mmol) in thf (20 mL) was added to 1b, [(IMesH+ F], 920 mg, 3.06 mmol), and the deep green color of the nickelocene solution gave way to a red-brown color within 10 min. The mixture was refluxed for 2 h and the now deep-red reaction mixture was cooled, filtered through a 3 × 3 cm Celite pad, and washed with thf until the filtrate was clear. The solution was then concentrated and placed in a 4° C refrigerator. Small needles of 3a (178 mg, 0.252 mmol, 17%) formed on standing overnight. Mpt. (dec.) > 200 °C. 1H NMR: 6.99 (8H, meta-CH), 6.64 (d, 4H, im, H 4,5), 2.58 (d, 6H, para-CH), 1.86 (d, 12H, ortho-CH). 13C NMR: 137.4 (ipso-CAr), 136.7 (meta-CAr), 136.6 (para-CAr), (129.0 ortho-CAr), 121.9 (NCH), 21.6, 17.8 (Me). 19F NMR: −156.7. Anal: calcd. (expt.) for C42H48N4F2Ni: C, 71.50 (70.36); H, 6.86 (7.02); N, 7.94 (7.39).

3.6. Catalytic Studies on the Reaction of 4-Bromoacetophenone with Phenylboronic Acid

A Schlenk tube was charged with 3.0 mol % of the catalyst (2, 10 mg or 3a, 21.2 mg) and degassed. 4-bromoacetophenone (1.0 mmol), phenylboronic acid (1.3 mmol) and K3PO4 (2.6 mmol) were added. Toluene (3 mL) was injected, and the tube was sealed with an unpierced septum and the mixture was immediately placed in a pre-heated oil bath at 90 °C. After 1 h, the reaction was quenched by exposing the reaction medium to air and cooled. 1H NMR yields were determined by removing a sample of the supernatant solution with a pipette, filtering the extract into an NMR tube and comparing the methyl signal integrations from the cross-coupling product and the 4-bromoacetophenone. We found that 2 gave 0% conversion, while in three separate experiments, 3a gave 29, 58 and 78% conversion.

3.7. Ligand Lability Studies

Complex 2 was mixed with 1 equivalent of KPF6 in acetonitrile (10 mL) and 1 equivalent of 0.01 M HNO3 (aq) [or 0.01 M HCl (aq) in a separate experiment]. The mixture was stirred at room temperature for 1h. No color changes were observed. The solvent was then removed under reduced pressure and the crude residue was extracted with CDCl3, filtered and then a 1H NMR spectrum was obtained. In each case, only signals for 2 were observed.

4. Conclusions

These results provide more evidence of the extreme tendency of fluoride or fluoro-complexes of metals to engage in hydrogen bonding in the solid state. In all four structures, there is extensive hydrogen bonding in the crystal. This is not entirely surprising, as fluorine is the most electronegative element, but nevertheless the number of such interactions was unexpected. The reactions of the imidazolium fluorides with nickelocene also show that the fluoride salts do not replicate the chemistry that is observed in these reactions of nickelocene with other imidazolium halides (X = Cl, Br, I). We have been unable to prepare complexes of the type [Ni(Cp)(NHC)(F)] even though a plethora of such complexes are known with many different NHC ligands and X = Cl, Br or I [13,14,15,16,17,18,20,21,22,23]. Furthermore, it appears that trans-square-planar complexes of the type trans-[NiF2(NHC)2] are unstable in solution, for reasons that are not clear to us but that may be due to “vessel effects”, i.e., reaction with the borosilicate glass and adsorbed water [4].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29184493/s1. Spectroscopic and X-ray data, including the CheckCif files for 1b, 1c, 2 and 3. Crystallographic data for compounds 1b, 1c, 2 and 3a have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication nos. CCDC 2383952, 2383953, 2383954 and 2383955. Copies of the data can be obtained free of charge from the Director, CCDC, 12 Union Road, Cambridge CB2 1EZ, UK (fax: +44-1223-336-033; e-mail: [email protected]; http://www.ccdc.cam.ac.uk).

Author Contributions

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

Funding

This research was funded by the CNRS (France) and the University of Strasbourg. S.S.W. thanks the University of Western Australia for a Hackett Foundation Alumni Honours Scholarship and a Raoul Robellaz Kahan Honours Scholarship in Chemistry.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials.

Acknowledgments

The authors thank E. Wasielewski and M. Chessé, who are responsible for the research platforms and facilities (Strasbourg NMR platform and Strasbourg chromatography facilities) of LIMA (UMR 7042 CNRS-Unistra-UHA) and who contributed, by their valuable technical and scientific support, to the achievement of this research project. We would also like to especially thank the crystallographers Lydia Karmazin and Corinne Bailly for their excellent work on poor crystals.

Conflicts of Interest

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

References

  1. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of Organofluorine Compounds to Pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef] [PubMed]
  2. Dolbier, W.R., Jr. Fluorine chemistry at the millennium. J. Fluorine Chem. 2005, 126, 157–163. [Google Scholar] [CrossRef]
  3. Zhou, L. Recent Advances in C–F Bond Cleavage Enabled by Visible Ligh Photoredox Catalysis. Molecules 2021, 26, 7051. [Google Scholar] [CrossRef] [PubMed]
  4. Nielsen, M.M.; Pedersen, C.M. Vessel effects in organic chemical reactions; a century old, overlooked phenomenon. Chem. Sci. 2022, 13, 6181–6196. [Google Scholar] [CrossRef] [PubMed]
  5. Grushkin, V.V. The Organometallic Fluorine Chemistry of Palladium and Rhodium: Studies toward Aromatic Fluorination. Acc. Chem. Res. 2010, 43, 160–171. [Google Scholar] [CrossRef]
  6. Murphy, E.F.; Ramaswarmy, M.; Roesky, H.W. Organometallic Fluorides: Compounds Containing Carbon-Metal–Fluorine Fragments of d-Block Metals. Chem. Rev. 1997, 97, 3425–3468. [Google Scholar] [CrossRef]
  7. Roesky, H.W. Playing the Keyboard of Fluorine Chemistry. Inorg. Chem. 1999, 38, 5934–5943. [Google Scholar] [CrossRef]
  8. Jasim, N.A.; Perutz, R.N.; Whitwood, A.C.; Braun, T.; Izundu, J.; Neumann, B.; Rothfeld, S.; Stammler, H.-G. Complexes Contrasting Reactivity of Fluoropyridines at Palladium and Platinum:  C−F Oxidative Addition at Palladium, P−C and C−F Activation at Platinum. Organometallics 2004, 26, 6140–6149. [Google Scholar] [CrossRef]
  9. Schaub, T.; Fischer, P.; Steffen, A.; Braun, T.; Radius, U.; Mix, A. C-F Activation of Fluorinated Arenes using NHC-Stabilized Nickel(0) Complexes: Selectivity and Mechanistic Investigations. J. Am. Chem. Soc. 2008, 130, 9304–9317. [Google Scholar] [CrossRef]
  10. Schaub, T.; Backes, M.; Radius, U. Catalytic C−C Bond Formation Accomplished by Selective C−F Activation of Perfluorinated Arenes. J. Am. Chem. Soc. 2006, 128, 15964–15965. [Google Scholar] [CrossRef]
  11. Rachor, S.G.; Müller, R.; Kaupp, M.; Braun, T. Hydrogen and Halogen Bonding to Au(I) Fluorido Complexes. Eur. J. Inorg. Chem. 2023, 26, e202200668. [Google Scholar] [CrossRef]
  12. Gauthier, R.; Nikolaos, V.; Tzouras, N.V.; Zhang, Z.; Bédard, S.; Saab, M.; Falivene, L.; Van Hecke, K.; Luigi Cavallo, L.; Nolan, S.P.; et al. Gold N-Heterocyclic Carbene Catalysts for the Hydrofluorination of Alkynes Using Hydrofluoric Acid: Reaction Scope, Mechanistic Studies and the Tracking of Elusive Intermediates. Chem. Eur. J. 2022, 22, e202103886. [Google Scholar] [CrossRef] [PubMed]
  13. Ritleng, V.; Barth, C.; Brenner, E.; Milosevic, S.; Chetcuti, M.J. Synthesis, structure, and solution dynamics of pentamethyl-cyclopentadienyl nickel complexes bearing N-heterocyclic carbene ligands. Organometallics 2008, 27, 4223–4228. [Google Scholar] [CrossRef]
  14. Ulm, F.; Poblador-Bahamonde, A.I.; Choppin, S.; Bellemin-Laponnaz, S.; Chetcuti, M.J.; Achard, T.; Ritleng, V. Synthesis, characterization, and catalytic application in aldehyde hydrosilylation of half- sandwich nickel complexes bearing (κ1-C)- and hemilabile (κ2-C,S)-thioether functionalized NHC ligands. Dalton Trans. 2018, 47, 17134–17145. [Google Scholar] [CrossRef] [PubMed]
  15. Abernethy, C.D.; Clyburne, J.A.C.; Cowley, A.H.; Jones, R.A. Reactions of Transition-Metal Metallocenes with Stable Carbenes. J. Am. Chem. Soc. 1999, 121, 2329–2330. [Google Scholar] [CrossRef]
  16. Ritleng, V.; Oertel, A.M.; Chetcuti, M.J. Half-sandwich NHC-nickel(II) complexes as pre-catalysts for the fast Suzuki coupling of aryl halides: A comparative study. Dalton Trans. 2010, 39, 8153–8160. [Google Scholar] [CrossRef]
  17. Oertel, A.M.; Ritleng, V.; Chetcuti, M.J. Synthesis and Catalytic Activity in Suzuki Coupling of Nickel Complexes Bearing n-Butyl- and Triethoxysilylpropyl-Substituted NHC Ligands: Toward the Heterogenization of Molecular Catalysts. Organometallics 2012, 7, 2829–2840. [Google Scholar] [CrossRef]
  18. Aloui, L.; Abidi, R.; Chetcuti, M.J. Synthesis and characterization of nickel-N-heterocyclic carbenes linked to a calix[6]arene platform and their applications in Suzuki-Miyaura cross-coupling catalysis. Inorg. Chim. Acta 2020, 505, 119494. [Google Scholar] [CrossRef]
  19. Zell, T.; Fischer, P.; Schmidt, D.; Radius, U. C–Br Activation of Aryl Bromides at Ni0(NHC)2: Stoichiometric Reactions, Catalytic Application in Suzuki–Miyaura Cross-Coupling, and Catalyst Degradation. Organometallics 2012, 31, 5065–5073. [Google Scholar] [CrossRef]
  20. Gu, S.; Du, J.; Huang, J.; Guo, Y.; Yang, L.; Xu, W.; Chen, W. Unsymmetrical NCN-pincer mononuclear and dinuclear nickel(II) complexes of N-heterocyclic carbene (NHC): Synthesis, structure and catalysis for Suzuki–Miyaura cross-coupling. Dalton Trans. 2017, 46, 586–594. [Google Scholar] [CrossRef]
  21. Rahman, M.M.; Zhao, Q.; Meng, G.; Szostak, R.; Szostak, M. [Ni(Np#)(η5-Cp)Cl]: Flexible, Sterically Bulky, Well-Defined, Highly Reactive Complex for Nickel-Catalyzed Cross-Coupling. Organometallics 2022, 41, 2597–2604. [Google Scholar] [CrossRef] [PubMed]
  22. Liao, C.-Y.; Chan, K.-T.; Chang, Y.-C.; Chen, C.-Y.; Tu, C.-Y.; Hu, C.-H.; Lee, H.M. Unexpected Solvent-Induced Cis/Trans Isomerization and Catalytic Application of a Bis-bidentate Nickel(II) Complex with N-Heterocyclic Carbene and Amido Functionalities. Organometallics 2007, 26, 5826–5833. [Google Scholar] [CrossRef]
  23. Henrion, M.; Ritleng, V.; Chetcuti, M.J. From acetone metalation to the catalytic α-arylation of acyclic ketones with NHC–nickel(II) complexes. Chem. Commun. 2014, 50, 4624–4627. [Google Scholar] [CrossRef] [PubMed]
  24. Cai, Y.; Ruan, L.-X.; Rahman, A.; Shi, S.-L. Fast Enantio- and Chemoselective Arylation of Ketones with Organoboronic Esters Enabled by Nickel/N-Heterocyclic Carbene Catalysis. Ang. Chem. Int. Ed. 2021, 60, 5262–5267. [Google Scholar] [CrossRef] [PubMed]
  25. Fernandez-Salas, J.A.; Marelli, E.; Cordes, D.B.; Slawin, A.M.Z.; Nolan, S.P. General and Mild Ni0-Catalyzed α-Arylation of Ketones Using Aryl Chlorides. Chem. Eur. J. 2015, 21, 3906–3909. [Google Scholar] [CrossRef]
  26. Ulm, F.; Shahane, S.; Truong-Phuoc, L.; Romero, T.; Papaefthimiou, V.; Chessé, M.; Chetcuti, M.J.; Pham-Huu, C.; Michon, C.; Ritleng, V. Half-Sandwich Nickel(II) NHC-Picolyl Complexes as Catalysts for the Hydrosilylation of Carbonyl Compounds: Evidence for NHC-Nickel Nanoparticles under Harsh Reaction Conditions. Eur. J. Inorg. Chem. 2021, 30, 3074–3082. [Google Scholar] [CrossRef]
  27. Chang, A.S.; Kawamura, K.E.; Henness, H.S.; Salpino, V.M.; Greene, J.C.; Zakharov, L.N.; Cook, A.K. (NHC)Ni(0)-Catalyzed Branched-Selective Alkene Hydrosilylation with Secondary and Tertiary Silanes. ACS Catal. 2022, 12, 11002–11014. [Google Scholar] [CrossRef]
  28. Junquera, L.B.; Puerta, M.C.; Valerga, P. R-Allyl Nickel(II) Complexes with Chelating N-Heterocyclic Carbenes: Synthesis, Structural Characterization, and Catalytic Activity. Organometallics 2012, 31, 2175–2183. [Google Scholar] [CrossRef]
  29. Ulm, F.; Djukic, J.-P.; Chetcuti, M.J.; Ritleng, V. Hydroboration of alkenes catalyzed by a nickel N-heterocyclic carbene complex: Reaction and mechanistic aspects. Chem. Eur. J. 2020, 26, 8916–8925. [Google Scholar] [CrossRef]
  30. Afandiyeva, M.; Kadam, A.A.; Wu, X.; Brennessel, W.W.; Kennedy, C.R. Synthesis, Structure, and Hydroboration Reactivity of Anionic Nickel(0) Complexes Supported by Bidentate NHC-Pyridone Ligands. Organometallics 2022, 41, 3014–3023. [Google Scholar] [CrossRef]
  31. Touney, E.E.; Van Hoveln, R.; Buttke, C.T.; Freidberg, M.D.; Guzei, I.A.; Schomaker, J.M. Heteroleptic Nickel Complexes for the Markovnikov-Selective Hydroboration of Styrenes. Organometallics 2016, 35, 3436–3439. [Google Scholar] [CrossRef]
  32. Lee, B.C.; Liu, C.-F.; Lin, L.Q.H.; Yap, K.Z.; Song, N.; Ko, C.H.M.; Chan, P.H.; Koh, M.J. N-Heterocyclic carbenes as privileged ligands for nickel-catalysed alkene functionalisation. Chem. Soc. Rev. 2023, 52, 2946–2991. [Google Scholar] [CrossRef] [PubMed]
  33. Hatakeyama, T.; Hashimoto, S.; Ishizuka, K.; Nakamura, M. Highly Selective Biaryl Cross-Coupling Reactions between Aryl Halides and Aryl Grignard Reagents: A New Catalyst Combination of N-Heterocyclic Carbenes and Iron, Cobalt, and Nickel Fluorides. J. Am. Chem. Soc. 2009, 131, 11949–11963. [Google Scholar] [CrossRef] [PubMed]
  34. Schaub, T.; Backes, M.; Radius, U. Square-Planar (Pentafluorophenyl)nickel(II) Complexes by Derivatization of a C–F Activation Product. Eur. J. Inorg. Chem. 2008, 17, 2680–2690. [Google Scholar] [CrossRef]
  35. Jahnke, M.C.; Hahn, F.E. Chemistry of N-heterocyclic Carbene Ligands. Top. Organomet. Chem. 2010, 30, 95–129. [Google Scholar] [CrossRef]
  36. Arduengo, A.J., III; Krafczyk, R.; Schmutzler, R.; Craig, H.A.; Goerlich, J.R.; Marshall, W.J.; Unverzagt, M. Imidazolylidenes, imidazolinylidenes and imidazolidines. Tetrahedron 1999, 55, 14523–14534. [Google Scholar] [CrossRef]
  37. Hintermann, L. Expedient syntheses of the N-heterocyclic carbene precursor imidazolium salts Ipr·HCl, IMes·HCl and IXy·HCl. Belstein J. Org. Chem. 2007, 3, 22. [Google Scholar] [CrossRef]
  38. Oertel, A.M.; Ritleng, V.; Chetcuti, M.J. N′-Activation of N-Arylimidazoles: Facile Syntheses of N-Alkyl-N′-arylimidazolium Iodides from Less Expensive Chloro Substrates. Synthesis 2009, 10, 1647–1650. [Google Scholar] [CrossRef]
  39. Zhu, Z.-Q.; Jiang, M.-Y.; Zheng, C.-G.; Xiao, J.-C. Efficient synthesis of 1-alkyl-3-methylimidazolium fluorides and possibility of the existence of hydrogen bonding between fluoride anion and C(sp3)–H. J. Fluorine Chem. 2012, 133, 160–162. [Google Scholar] [CrossRef]
  40. Pilon, M.C.; Grushin, V. Synthesis and Characterization of Organopalladium Complexes Containing a Fluoro Ligand. Organometallics 1998, 17, 1774–1781. [Google Scholar] [CrossRef]
  41. Bouvet, S.; Pégot, B.; Marrot, J.; Magnier, E. Solvent free nucleophilic introduction of fluorine with [bmim][F]. Tetrahedron Lett. 2014, 55, 826–829. [Google Scholar] [CrossRef]
  42. Bouvet, S.; Pégot, B.; Diter, P.; Marrot, J.; Magnier, E. Oxidative desulfurization–fluorination reaction promoted by [bdmim][F] for the synthesis of difluorinated methyl ethers. Tetrahedron Lett. 2015, 56, 1682–1686. [Google Scholar] [CrossRef]
  43. Alič, B.; Petrovčic, J.; Jelen, J.; Tavčar, G.; Iskra, J. Renewable Reagent for Nucleophilic Fluorination. J. Org. Chem. 2022, 87, 5987–5993. [Google Scholar] [CrossRef] [PubMed]
  44. Tang, P.; Wang, W.; Ritter, T. Deoxyfluorination of Phenols. J. Am. Chem. Soc. 2011, 133, 11482–11484. [Google Scholar] [CrossRef] [PubMed]
  45. Goldberg, N.W.; Shen, X.; Li, J.; Ritter, T. AlkylFluor: Deoxyfluorination of Alcohols. Org. Lett. 2016, 18, 6102–6104. [Google Scholar] [CrossRef]
  46. Alič, B.; Tavčar, G. Reaction of N-heterocyclic carbene (NHC) with different HF sources and ratios—A free fluoride reagent based on imidazolium fluoride. J. Fluorine Chem. 2016, 192, 141–146. [Google Scholar] [CrossRef]
  47. Huang, H.; Zhong, J.; Tan, X.; Guo, X.; Yuan, B.; Lin, Y.; Francisco, J.S.; Zeng, X.G. New Insights into the Stability of Anhydrous 2H-Imidazolium Fluoride and its High Dissolution Capability toward a Strongly Hydrogen-Bonded Compound. J. Am. Chem. Soc. 2020, 142, 10314–10318. [Google Scholar] [CrossRef]
  48. Alič, B.; Tramšek, M.; Kokalj, A.; Tavcar, G. Discrete GeF5 Anion Structurally Characterized with a Readily Synthesized Imidazolium Based Naked Fluoride Reagent. Inorg. Chem. 2017, 56, 10070–10077. [Google Scholar] [CrossRef]
  49. Libri, S.; Jasim, N.A.; Perutzm, R.N.; Brammer, L. Metal Fluorides Form Strong Hydrogen Bonds and Halogen Bonds: Measuring Interaction Enthalpies and Entropies in Solution. J. Am. Chem. Soc. 2008, 130, 7842–7844. [Google Scholar] [CrossRef]
  50. Kelly, R.A.; Scott, N.M.; Díez-González, S.; Stevens, E.D.; Nolan, S.P. Simple Synthesis of CpNi(NHC)Cl Complexes (Cp = Cyclopentadienyl; NHC = N-Heterocyclic Carbene). Organometallics 2005, 24, 3442–3447. [Google Scholar] [CrossRef]
  51. Abernethy, C.D.; Cowley, A.H.; Jones, R.A. Reactions of Nickelocene with 1,3-dimesitylimidazolium chloride. J. Organomet. Chem. 2000, 596, 3–5. [Google Scholar] [CrossRef]
  52. Malan, F.P.; Singleton, E.; van Rooyen, P.H.; Conradie, J.; Landman, M. Synthesis, structure and DFT conformation analysis of CpNiX(NHC) and NiX2(NHC)2 (X = SPh or Br) complexes. J. Mol. Struct. 2017, 1147, 235–243. [Google Scholar] [CrossRef]
  53. Banach, L.; Gunka, P.A.; Zachara, J.; Buchowicz, W. Half-sandwich Ni(II) complexes [Ni(Cp)(X)(NHC)]: From an underestimated discovery to a new chapter in organonickel chemistry. Coord. Chem. Rev. 2019, 389, 19–58. [Google Scholar] [CrossRef]
  54. Bielinski, E.A.; Dai, W.; Guard, L.M.; Hazari, N.; Takase, M.K. Synthesis, properties, and reactivity of palladium and nickel NHC complexes supported by combinations of allyl, cyclopentadienyl, and indenyl ligands. Organometallics 2013, 32, 4025–4037. [Google Scholar] [CrossRef]
  55. Pankov, R.O.; Tarabrin, I.R.; Son, A.G.; Minyaev, M.E.; Prima, D.O.; Ananikov, V.P. Synthesis and comparative study of (NHCF)PdCl2Py and (NHCF)Ni(Cp)Cl complexes: Investigation of the electronic properties of NHC ligands and complex characteristics. Dalton Trans. 2024, 53, 12503–12518. [Google Scholar] [CrossRef]
  56. Oertel, A.M.; Ritleng, V.; Burr, L.; Chetcuti, M.J. Synthesis and Structural Characterization of Half-Sandwich Nickel Complexes Bearing Two Different N-Heterocyclic Carbene Ligands. Organometallics 2011, 30, 6685–6691. [Google Scholar] [CrossRef]
  57. Henrion, M.; Oertel, A.M.; Ritleng, V.; Chetcuti, M.J. Facile displacement of η5-cyclopentadienyl ligands from half-sandwich alkyl,NHC–nickel complexes: An original route to robust cis-C,C-nickel square planar complexes. Chem. Commun. 2013, 49, 6424–6426. [Google Scholar] [CrossRef]
  58. Oertel, A.M.; Ritleng, V.; Chetcuti, M.J.; Veiros, L.F. C-H Activation of Acetonitrile at Nickel: Ligand Flip and Conversion of N-Bound Acetonitrile into a C-Bound Cyanomethyl Ligand. J. Am. Chem. Soc. 2010, 132, 13588–13589. [Google Scholar] [CrossRef]
  59. Gibson, J.A.; Ibbott, D.G.; Janzen, A.F. Fluorine NMR spectra of sulfur, phosphorus, and silicon fluorides. Hydrolysis and intra-intermolecular mechanism of fluorine exchange. Can. J. Chem. 1973, 51, 3203–3210. [Google Scholar] [CrossRef]
  60. Klanberg, F.; Muetterties, E.L. Nuclear Magnetic Resonance Studies on Pentacoordinate Silicon Fluorides. Inorg. Chem. 1968, 7, 155–160. [Google Scholar] [CrossRef]
  61. Rosenau, C.P.; Jelier, B.J.; Gossert, A.D.; Togni, A. Exposing the Origins of Irreproducibility in Fluorine NMR Spectroscopy. Ang. Chem. 2018, 57, 9528–9533. [Google Scholar] [CrossRef]
  62. Finney, W.F.; Wilson, E.; Callender, A.; Morris, M.D.; Beck, L.W. Reexamination of Hexafluorosilicate Hydrolysis by 19F NMR and pH Measurement. Environ. Sci. Technol. 2006, 40, 2572–2577. [Google Scholar] [CrossRef] [PubMed]
  63. Brownstein, S. Complex fluoroanions in solution. VII. Replacement of fluoride by trifluoroacetate. Can. J. Chem. 1978, 56, 343–347. [Google Scholar] [CrossRef]
  64. Zhang, D.; Zhou, S.; Li, Z.; Wang, Q.; Weng, L. Direct synthesis of cis-dihalido-bis(NHC) complex of nickel(II) and catalytic application in olefin addition polymerization: Effect of halogen co-ligands and density functional theory study. Dalton Trans. 2013, 42, 12020–12030. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, Z.-H.; Xu, Y.-C.; Xie, L.-Z.; Sun, H.-M.; Shen, Q.; Zhang, Y. Controlled synthesis of nickel(II) dihalides bearing two different or identical N-heterocyclic carbene ligands and the influence of carbene ligands on their structures and catalysis. Dalton Trans. 2011, 40, 4697. [Google Scholar] [CrossRef]
  66. Ritleng, V.; Brenner, E.; Chetcuti, M.J. Preparation of a N-heterocyclic carbene nickel(II) complex. Synthetic experiments in current organic and organometallic chemistry. J. Chem. Educ. 2008, 85, 1646–1648. [Google Scholar] [CrossRef]
  67. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  68. Sheldrick, G.M. SHELXS-97 Program for Crystal Structure Determination. Acta Crystallogr. 1990, 46, 467–473. [Google Scholar] [CrossRef]
  69. M86-E01078 APEX2 User Manual; Bruker AXS Inc.: Madison, WI, USA, 2006.
  70. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  71. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Cryst. 2003, 36, 7–13. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of imidazolium fluorides 1a1c based on Xiao’s method [39].
Scheme 1. Synthesis of imidazolium fluorides 1a1c based on Xiao’s method [39].
Molecules 29 04493 sch001
Figure 1. Structure of 1b, [(IMesH)+ F]. C, N and F atoms are shown as black, blue and yellow ellipsoids at the 50% probability level; both positions of the symmetrically disordered F ions are shown. Only the imidazolium H-atom is shown as a sphere of arbitrary radius, for clarity, showing its close interactions with the F ions (H  F = 1.85 Å).
Figure 1. Structure of 1b, [(IMesH)+ F]. C, N and F atoms are shown as black, blue and yellow ellipsoids at the 50% probability level; both positions of the symmetrically disordered F ions are shown. Only the imidazolium H-atom is shown as a sphere of arbitrary radius, for clarity, showing its close interactions with the F ions (H  F = 1.85 Å).
Molecules 29 04493 g001
Figure 2. Structure of the salt [(IPrH)+ F], 1c. Ellipsoids are shown at the 50% probability level. C, N and F atoms are color coded as shown. Only the imidazolium H(2) atom is shown, with a short H  F interaction of 1.86 Å, for clarity.
Figure 2. Structure of the salt [(IPrH)+ F], 1c. Ellipsoids are shown at the 50% probability level. C, N and F atoms are color coded as shown. Only the imidazolium H(2) atom is shown, with a short H  F interaction of 1.86 Å, for clarity.
Molecules 29 04493 g002
Figure 3. HF and HO interactions < 3.0 Å in the crystal of 1b, viewed approximately down the y-axis of the unit cell. C, N, O and F atoms are color coded.as shown. HimiO = 2.28, 2.38, 2.53 Å; C(2)HImid F = 1.85; HMe F = 2.78, 2.71 Å; Hmeta F = 2.70 Å.
Figure 3. HF and HO interactions < 3.0 Å in the crystal of 1b, viewed approximately down the y-axis of the unit cell. C, N, O and F atoms are color coded.as shown. HimiO = 2.28, 2.38, 2.53 Å; C(2)HImid F = 1.85; HMe F = 2.78, 2.71 Å; Hmeta F = 2.70 Å.
Molecules 29 04493 g003
Figure 4. H  F and interactions < 3.0 Å in 1c, viewed down the z-axis of the unit cell. C, N, Cl and F atoms are color coded as shown. Many C and H atoms of the i-Pr-group and the CH2Cl2 H-atoms are hidden, for clarity. Himi  F = 1.84 Å; Hpara  F = 2.41 Å.
Figure 4. H  F and interactions < 3.0 Å in 1c, viewed down the z-axis of the unit cell. C, N, Cl and F atoms are color coded as shown. Many C and H atoms of the i-Pr-group and the CH2Cl2 H-atoms are hidden, for clarity. Himi  F = 1.84 Å; Hpara  F = 2.41 Å.
Molecules 29 04493 g004
Scheme 2. Reactions of the imidazolium fluorides 1a1c with nickelocene.
Scheme 2. Reactions of the imidazolium fluorides 1a1c with nickelocene.
Molecules 29 04493 sch002
Figure 5. Structure of 2, [Ni(Cp)(IMe)2]+ F. Ellipsoids are shown at the 50% probability level and the C, N, Ni and F atoms are colored as shown. Two relatively short H  F interactions are shown (HImid  F = 2.28 Å, HCp  F = 2.41.
Figure 5. Structure of 2, [Ni(Cp)(IMe)2]+ F. Ellipsoids are shown at the 50% probability level and the C, N, Ni and F atoms are colored as shown. Two relatively short H  F interactions are shown (HImid  F = 2.28 Å, HCp  F = 2.41.
Molecules 29 04493 g005
Figure 6. Hydrogen short contacts and H-bonds (in Å) for (a) the η5-C5H5 and (b) the Ni(IMe)2 portions of complex 2 with the F anion and the O atom of H2O molecules present in the lattice. C, N, Ni, F and O atoms are colored as shown.
Figure 6. Hydrogen short contacts and H-bonds (in Å) for (a) the η5-C5H5 and (b) the Ni(IMe)2 portions of complex 2 with the F anion and the O atom of H2O molecules present in the lattice. C, N, Ni, F and O atoms are colored as shown.
Molecules 29 04493 g006
Figure 7. View down the crystallographically imposed 3-fold rotational z-axis in 2 that shows the pseudo-octahedral geometry (with a significant rhombic distortion) of an oxygen atom of water, that interacts weakly with 6 methyl group hydrogen atoms of 6 different IMe groups. All H  O = 2.81 Å; HOH angles are 74, 105 or 180°. C, N, O atoms are colored as shown with H atoms shown as small grey spheres.
Figure 7. View down the crystallographically imposed 3-fold rotational z-axis in 2 that shows the pseudo-octahedral geometry (with a significant rhombic distortion) of an oxygen atom of water, that interacts weakly with 6 methyl group hydrogen atoms of 6 different IMe groups. All H  O = 2.81 Å; HOH angles are 74, 105 or 180°. C, N, O atoms are colored as shown with H atoms shown as small grey spheres.
Molecules 29 04493 g007
Figure 8. Structure of complex 3a, trans-[NiF2(IMes)2]. Ellipsoids are shown at the 50% probability level. C, N, F and Ni atoms are colored as shown. Hydrogen atoms are not shown for clarity.
Figure 8. Structure of complex 3a, trans-[NiF2(IMes)2]. Ellipsoids are shown at the 50% probability level. C, N, F and Ni atoms are colored as shown. Hydrogen atoms are not shown for clarity.
Molecules 29 04493 g008
Figure 9. Intramolecular H  F interactions in 3a, trans-[NiF2(IMes)2]. C, N, F and Ni atoms are color coded as displayed. Only H atoms (of arbitrary radius) on the ortho-Me groups are depicted, for clarity. Intramolecular close contacts are shown (distances in Å). Intermolecular H  F interactions of 2.49 (with a meta-hydrogen) and with a CH3 group hydrogen atom of 2.47 are not shown.
Figure 9. Intramolecular H  F interactions in 3a, trans-[NiF2(IMes)2]. C, N, F and Ni atoms are color coded as displayed. Only H atoms (of arbitrary radius) on the ortho-Me groups are depicted, for clarity. Intramolecular close contacts are shown (distances in Å). Intermolecular H  F interactions of 2.49 (with a meta-hydrogen) and with a CH3 group hydrogen atom of 2.47 are not shown.
Molecules 29 04493 g009
Scheme 3. Catalytic tests of 2 and 3a for Suzuki–Miyaura coupling.
Scheme 3. Catalytic tests of 2 and 3a for Suzuki–Miyaura coupling.
Molecules 29 04493 sch003
Table 1. Key parameters of 1b, 1c. a From Ref. [39]; average values of two independent molecules are given. b Angle between the normal to the least square planes of the aromatic C6 ring atoms and the imidazolium C3N2 atoms.
Table 1. Key parameters of 1b, 1c. a From Ref. [39]; average values of two independent molecules are given. b Angle between the normal to the least square planes of the aromatic C6 ring atoms and the imidazolium C3N2 atoms.
Parameter (Å or °)1a, [(IMe)+ F] a1b, [(IMesH)+ F]1c, [(IPrH)+ F]
C–Himid  F1.92 and 2.041.851.84
C=Cimid1.331 (3)1.341 (6)1.360 (17)
C(1)imid–N1.318 (3)1.332 (3)1.328 (9)
C(2)imid–N1.362 (3)1.377 (3)1.355 (11)
b Ar–imid. least sq. planes angle70.282.0
Table 2. Structural parameters of complex 2. a The average of chemically equivalent bonds is given. b Angle between the normals of the 2 NHC C3N2 least square planes.
Table 2. Structural parameters of complex 2. a The average of chemically equivalent bonds is given. b Angle between the normals of the 2 NHC C3N2 least square planes.
Parameter (Å, Bond-Distances; Angles,°)2, [Ni(Cp)(IMe)2+] F.
a Ni–Ccarb1.892 (6)
a C=Cimid1.320 (11)
a C(1)imid–N1.350 (8)
a C(2)imid–N and C(2′)imid–N1.388 (9)
a N–CMe1.463 (9)
a Ni–Cp2.124 (6)
Ni–Cpcentroid1.752
Ccarb–Ni–Ccarb94.5 (2)
Ccarb–Ni–Cpcentroid132.3, 133.1
b IMe(1)–IMe(2) angle94.5
Table 3. Key bond lengths (Å) and angles (°) with estimated standard deviations in parenthesis for trans-[NiF2(IMes)2], 3a and a few other trans-square planar Ni(NHC)2 complexes with Ni–F bonds. 1 F–NiC(1) angle. A = trans-[Ni(N,N′-(i-Pr2Im)2(C6F4CF3-4)(F)] [10]; B = trans-[Ni(N,N′-(i-Pr2Im)2(C6F5)(F) [9]; C = trans-[Ni(N,N′-(i-Pr2Im)2(C6F4SiMe3-4)(F)] [9].
Table 3. Key bond lengths (Å) and angles (°) with estimated standard deviations in parenthesis for trans-[NiF2(IMes)2], 3a and a few other trans-square planar Ni(NHC)2 complexes with Ni–F bonds. 1 F–NiC(1) angle. A = trans-[Ni(N,N′-(i-Pr2Im)2(C6F4CF3-4)(F)] [10]; B = trans-[Ni(N,N′-(i-Pr2Im)2(C6F5)(F) [9]; C = trans-[Ni(N,N′-(i-Pr2Im)2(C6F4SiMe3-4)(F)] [9].
Parameter3aABC
Ni–Ccarb1.905 (4)1.932 (8), 1.911 (8)1.924 (2), 1.933 (2)1.911 (6), 1.905 6)
Ni–F1.816 (2)1.856 (4)1.891 (1)1.856 (3)
F–Ni–F178.3 (18)178.6 (3)175.5 (1) 1177.7 (2) 1
F–Ni–Ccarb89.23 (13), 90.79 (13)89.1 (2), 88.0 (3)90.2 (1), 88.6 (1)89.4 (2), 89.0 (2)
Ccarb–Ni–Ccarb178.3 (18)175.0 (3)-175.0 (2)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wills, S.S.; Bailly, C.; Chetcuti, M.J. Towards Nickel–NHC Fluoro Complexes—Synthesis of Imidazolium Fluorides and Their Reactions with Nickelocene. Molecules 2024, 29, 4493. https://doi.org/10.3390/molecules29184493

AMA Style

Wills SS, Bailly C, Chetcuti MJ. Towards Nickel–NHC Fluoro Complexes—Synthesis of Imidazolium Fluorides and Their Reactions with Nickelocene. Molecules. 2024; 29(18):4493. https://doi.org/10.3390/molecules29184493

Chicago/Turabian Style

Wills, Siobhan S., Corinne Bailly, and Michael J. Chetcuti. 2024. "Towards Nickel–NHC Fluoro Complexes—Synthesis of Imidazolium Fluorides and Their Reactions with Nickelocene" Molecules 29, no. 18: 4493. https://doi.org/10.3390/molecules29184493

Article Metrics

Back to TopTop