Abstract
A diverse range of heteromultinuclear NiI/[MCO] clusters (MCO = CpFe(CO)2, CpRu(CO)2, Cp*W(CO)3) supported by a N-heterocyclic carbene ligand have been synthesized by reacting the NiI precursor, [IPrNi(μ-Cl)]2, with [MCO]− reagents under various conditions. Clusters with Ni2Fe2, NiFe2, Ni2Ru, Ni2Ru2, NiRu2, and Ni2W, and NiW cores were all characterized using NMR and IR spectroscopies and X-ray crystallography. The NiI-containing paramagnetic heterobinuclear species, IPrNi-Wp* (7), was further characterized by EPR spectroscopy and DFT calculations. Notably, unlike previously studied (NHC)CuI-[MCO] derivatives, complex 7 was found to coordinate Lewis bases like 3-chloropyridine to produce (IPr)(3-Clpy)NiWp* (9). Complex 9 further underwent thermolytic C-Cl activation, proposed to involve NHC-free [(3-Clpy)Ni(μ-Wp*)]2 (10), to provide the C-arylated N-heterocyclic carbene product, [IPr(py-3-yl)]+[Cp*WCl2(CO)2]− (11).
Graphical Abstract

INTRODUCTION
The development of new platforms for cooperative bond activation is a frontier research area in homogeneous catalysis.1 One strategy for achieving cooperative bond activation is heterobinuclear cooperativity, which involves use of polar metal-metal interactions to activate inert small molecules for incorporation into organic substrates.1b,1c,1f,1i,2 In recent years, reported examples of organic transformations catalyzed by heterobinuclear metal complexes include C-H functionalization,3 carbonylation,4 hydrogenation,5 and hydrofunctionalization.6
Within this context, our group has been specifically interested in heterobimetallic systems consisting of a copper N-heterocyclic carbene (Cu-NHC) Lewis acid and a metal carbonyl anion ([MCO]− = e.g. FeCp(CO)2−, Mn(CO)5−, MoCp(CO)3−, etc.) Lewis base.7 These complexes feature dative [MCO]→Cu(NHC) bonds8 that are capable of C-X,9 H-H,10 H-B,11 and H-Sn6a activation. We have applied this platform to develop several classes of catalytic transformations,3b,4d,5b,6,12 and the modular nature of the catalyst design should allow for more applications to emerge as we expand our library of available heterobimetallic catalysts.
The catalytic mechanisms for the previously explored reactions catalyzed by (NHC)Cu-[MCO] complexes involve redox cycling at the metal carbonyl unit, while the Cu(NHC) group remains redox neutral. New reactivity paradigms might be accessible if the d10 CuI center were replaced with a redox-active metal ion with an open valence shell. Generally, heterobimetallic complexes featuring [MCO]-M′ bonds are limited to closed-shell systems where M′ is a d0 metal ion13 or a d10 metal ion.7 Exceptions include d8 NiII in square planar environments that have a strong preference for populating diamagnetic ground states and do not engage in Ni-based redox cycling.14 A recent survey of paramagnetic, metal-metal bonded, heterometallic complexes did not include any [MCO]-M′ derivatives,15 and to our knowledge the only open shell examples are a series of ArM-FeCp(CO)2 complexes reported by Power (M = Cr, Mn, Fe)16 and a (Ph3P)2Ni-WCp(CO)3 complex,17 none of which have available reactivity data in the literature.
This paper reports our efforts at synthesizing (NHC)Ni-[MCO] complexes in which CuI has been replaced by NiI, a redox-active d9 metal ion. Due to the possibility for redox chemistry, the synthesis of these heterobimetallic NiI complexes is not as straightforward as the CuI derivatives, and several heteromultinuclear Ni/Fe and Ni/Ru complexes were discovered serendipitously while targeting (NHC)Ni-MCp(CO)2 molecules (M = Fe, Ru). In one case, a paramagnetic (NHC)Ni-WCp*(CO)3 complex was successfully obtained. Reported here are this complex’s characterization by NMR and EPR spectroscopies, X-ray crystallography, DFT modeling, and preliminary reactivity data. While historically NiI has been considered a rare oxidation state,18 recently the diverse chemistry of this metal ion has emerged in many stoichiometric and catalytic transformations.19 We anticipate discovery of novel heterobinuclear catalysis based on the new heterobimetallic NiI complexes reported here.
RESULTS AND DISCUSSION
Synthesis and characterization of heteromultinuclear Ni/Fe and Ni/Ru clusters.
The most common method for synthesizing (NHC)Cu-[MCO] complexes is to react (NHC)CuCl and [MCO]− precursors.20 These salt metathesis reactions are typically quite clean and high-yielding, in part because there is little chance of competing electron transfer from [MCO]− to the d10 CuI center. Thus, we began our studies using [IPrNi(μ-Cl)]2 as a NiI precursor21 in combination with NaFp and KRp (Fp = FeCp(CO)2, Rp = RuCp(CO)2). The products were fully characterized by 1H NMR spectroscopy, IR spectroscopy, and X-ray crystallography. Although the targeted IPrNi-[MCO] complexes were expected to be paramagnetic due to retention of d9 NiI (S = ½), these reactions with Fp− and Rp− produced exclusively diamagnetic products (Scheme 1), indicative of electron transfer between the [MCO]− anions and the NiI center.
Scheme 1.

Synthesis of heteromultinuclear (a) Ni/Fe and (b) Ni/Ru clusters (Fp = FeCp(CO)2, Rp = RuCp(CO)2)
The reaction between [IPrNi(μ-Cl)]2 and NaFp in THF at room temperature produced a color change from yellow to dark brown immediately upon mixing, and 1H NMR spectroscopy was used to monitor the reaction progress. In the first 5 minutes, two new resonances in the Cp region were evident at 3.88 and 4.28 ppm in a ratio of 2:1, while the 4.28-ppm resonance dominated after 6 h. These resonances were found to arise from the heterotetranuclear cluster, [IPrNi(μ-Fp)]2 (1), and the cyclotrinuclear cluster, IPrNiFp2 (2), respectively. The mixtures of 1 and 2 (ratio 2: 1) were found to quantitatively convert to 2 in C6D6 over 12 h at 60°C; presumably a IPrNi0 byproduct is lost during this transformation.
The Ni/Ru chemistry, while analogous to the Ni/Fe chemistry, was more complex. The reaction between [IPrNi(μ-Cl)]2 and KRp in THF at room temperature gave mixtures of cyclotrinuclear IPrNiRp2 (3) as the major product with two minor products, [IPrNi(μ-Rp)]2 (4) and (IPrNi)2(μ-Rp)(μ-CO)3(μ3-CO) (5). All of the assignments were confirmed by X-ray crystallography, as discussed below. Running the same reaction at 40°C instead of room temperature provided complex 3 exclusively. Reacting [IPrNi(μ-Cl)]2 and KRp in toluene at −35°C gave yet another complex, (IPrNi)2(μ-Rp)(μ-Cl) (6). Interestingly, complex 4 could be obtained by adding NaOtBu to 6 in toluene at 50 °C for 1 h. Adding additional KRp to 6 produced 3. When isolated complex 4 was heated to 80°C in C6D6 for 24 h, it quantitatively converted to 3 with loss of “IPrNi0” as with the Fe case. Thus, synthetic methods were identified for isolating complexes 3, 4, and 6. Although we accidently obtained a crystal of 5, we failed to identify conditions that allowed us to isolate 5, which we hypothesize is a decomposition product generated by loss of “CpRu” from 4.
Solid-state structures of complexes 1-6 are shown in Figure 1. The solid-state structure of 2 displays an isosceles NiFe2 triangle with bent Ni-Fe-Fe (60.329(14) and 58.768(14)°) and Fe-Ni-Fe (60.903(14)°) angles and significantly bent CNHC-Ni-Fe angles (139.33(8)° and 159.63(8)°). The Ni-Fe bond distances (2.4652(5) and 2.5050(5) Å) are in the range of Ni-Fe single bonds (typically 2.41–2.64 Å), and the Fe-Fe bond distance (2.5192(5) Å) is in the range of a Fe-Fe single bond (2.46–2.78 Å) based on literature precedents.22 The analogous NiRu2 complex 3 is isostructural, with bent Ni-Ru-Ru (57.916(11) and 57.304(12)°) and Ru-Ni-Ru (64.780(12)°) angles along with significantly bent CNHC-Ni-Ru angles (142.56(8) and 151.81(8)°). The Ni-Ru distances (2.5165(5) and 2.5337(4) Å are close to the Ni-Ru single bond distance in Tp#Ni-RuCp(CO)2 (2.512(1) Å; Tp# = hydrotris(3,5-dimeth-yl-4-bromopyrazolyl)borato))23, and the Ru-Ru bond distance (2.7053(5) Å) is in the range of a Ru-Ru single bond (2.59–2.75 Å) in some Ru cluster cases.24 Previously, similar structures of related MoNi2 clusters supported by NHCs have been reported by Chetcuti.25 Calculation of asymmetry parameters26 indicate that all CO ligands in 2 and 3 are in the semi-bridging regime. In each case, one CO unit can be considered as a semi-bridging μ3-CO between Ni and two M (M = Fe or Ru), while the others are assigned as μ2-CO ligands between Ni and M or M and M. The solid-state IR spectra of 2 and 3 have features at 1656 and 1665 cm−1, respectively, assigned to the more activated semi-bridging μ3-CO ligands, while the μ2-CO ligands are assigned to features at 1841 and 1749 cm−1, respectively.
Figure 1.

Solid-state structures of complexes 1-6 determined by X-ray crystallography. For clarity, hydrogen atoms and crystallized solvent molecules are omitted. NHC ligands and Cp groups are shown as wireframes. All other atoms are shown as 50% probability ellipsoids.
The solid-state structure of 1 displays a tetranuclear Ni-Fe-Fe-Ni core, featuring bent Ni-Fe-Fe (82.75(6)°) and CNHC-Ni-Fe (158.60(19)°) angles. The Ni-Fe bond distance in 1 (2.3974(14) Å) is shorter than in 2, and the Fe-Fe bond distance is 2.525(2) Å, which is shorter than the Fe-Fe distance in Fp2 (2.5389(3) Å)27. The analogous Ni2Ru2 complex 4 is isostructural to 1, with bent Ni-Ru-Ru (81.454(4)°) and CNHC-Ni-Ru (156.8(2)°) angles. The Ni-Ru bond distance in 4 (2.4785(11) Å) is shorter than in 3, and the Ru-Ru bond distance is 2.7334(15) Å, which is similar to that in Rp2 (2.7412(4) Å)28. According to the solid-state structures of 1 and 4, in each case two CO units can be considered semibridging μ3-CO ligands between Ni and two M (M = Fe or Ru), while the others are assigned as semibridging μ2-CO ligands between Ni and M. The elongated C-O bond distances in 1 (1.218(8) Å) and in 4 (1.220(9) Å) indicate the significant CO activation. The solid-state IR spectra of 1 and 4 show features at 1552 and 1554 cm−1, respectively, assigned to semi-bridging μ3-CO ligands, consistent with the longer C-O bond distances. The μ2-CO ligands are assigned to features at 1757 and 1753 cm−1, respectively.
Complex 6 has Ni-Ru bond distances of 2.5702(6) and 2.5580(6) Å and a Ni-Ni distance of 2.2974(7) Å. Interestingly, 6 displays an unusually short Ni-Ni distance compared to [IPrNi(μ-Cl)]2 (2.5194(5) Å)2, and slightly shorter than the Ni-Ni distances measured for [(IPr)Ni]2(μ-Cl)(μ:η2-4-CH3-C6H4) (2.3954(5) Å)19e and [(IPr)Ni]2(μ-CO)(μ-η2,η2-CO2) (2.3374(4) Å)29. Complex 6 has bent Ni-Ni-Ru (63.657(19) and 63.113(19)°) and Ni-Ru-Ni (53.231(17)°) angles and shows significantly bent CNHC-Ni-Ni angles (171.69(10) and 172.72(11)°). The CO units can be considered as semi-bridging μ2-CO ligands between Ni and Ru. The solid-state IR spectrum of the highly symmetric 6 displays a single feature at 1749 cm−1.
Synthesis and characterization of a heterobinuclear Ni/W complex.
Compared to Fp− (E° = −1.352 V vs. ferrocene) and Rp− (E° = −1.057 V vs. ferrocene), Wp− derivatives are known to be weaker reductants (Wp = WCp(CO)3, E° = −0.491 V vs. ferrocene).30 Furthermore, a (Ph3P)2Ni-Wp derivative was studied in the 1980s,17 indicating that NiI is compatible with Wp−. Thus, we hypothesized that electron transfer from [MCO]− to NiI could be avoided by use of tungsten carbonyl anions under appropriate reaction conditions. Accordingly, we were successful in preparing IPrNiWp* (7) from a 1:4 mixture of [IPrNi(μ-Cl)]2 and LiWp* in THF at 40°C (Scheme 2, Wp* = WCp*(CO)3). As expected, 7 displays a 1H NMR spectrum with broadened and paramagnetically shifted resonances between 0 and 15 ppm. In addition, we successfully identified and isolated the intermediate, (IPrNi)2(μ-Wp*)(μ-Cl) (8), from a 1:2 mixture of [IPrNi(μ-Cl)]2 and LiWp* that had been allowed to react in toluene at room temperature for 2 h. Isolated complex 8 was converted to 7 with additional LiWp*. During investigation of the thermal stability of 7, we found that heating to 80°C for 24 h in C6D6 formed free IPr plus a decomposition product we are tentatively calling [Ni(μ-Wp*)]n. In two cases, decomposition mixtures yielded X-ray quality crystals in which free IPr co-crystallized with either [Ni(μ-Wp*)]3 or [Ni(μ-Wp*)]4 that featured planar NinWn metallacycles with complex μ-CO networks. However, we have been unable to model these data sufficiently for publication.
Scheme 2.

Synthesis of a heteromultinuclear Ni/W complexes
The solid-state structures of 7 and 8 were determined by X-ray crystallography (Figure 2). Intermediate 8 has a structure closely related to that of 6, with Ni-W bond distances of 2.5910(7) and 2.6138(7) Å, a Ni-Ni distance of 2.4335(7) Å, and bent Ni-Ni-W (62.601(19) and 61.649(19)°), Ni-W-Ni (55.749(18)°), and CNHC-Ni-Ni (166.03(6) and 161.08(7)°) angles. The three CO units can be considered as semibridging μ2-CO ligands between Ni and W. The solid-state IR spectrum of 8 has features at 1763 and 1722 cm−1 assigned to these semi-bridging μ2-CO ligands.
Figure 2.

Solid-state structures of complexes 7-9 determined by X-ray crystallography. For clarity, hydrogen atoms and crystallized solvent molecules are omitted. NHC ligands and Cp groups are shown as wireframes. All other atoms are shown as 50% probability ellipsoids.
Heterobinuclear complex 7 features a Ni-W bond distance of 2.4977(10) Å, and a bent CNHC-Ni-W angle (156.8(2)°). It has a significantly shorter Ni-W bond than the Ni-W single bond in (Ph3P)2Ni-WCp(CO)3 (2.574(3) Å)12 and is in the range of the Ni=W double bond (~2.466 Å) reported for the complex of Cp*Ni-W(η-C5H4Me)(CO)3.31 According to the solid-state structure, two of the CO units are assigned as semi-bridging ligands according to their asymmetry parameters.6 The solid-state IR spectrum of 7 has two features at 1733 and 1760 cm−1 assigned to the two semi-bridging CO ligands, and a feature at 1891 cm−1 assigned to the terminal CO ligand. The closest Cu-W analogue to 7 is IPrCuWp (Wp = WCp(CO)3),1a which has been structurally characterized in two forms, one with two semi-bridging CO ligands and one with three. Selected metrics for the form with two semi-bridging CO ligands are presented in Table 1 for comparison to complex 7. The metal-metal bond distance in 7 is comparatively short despite the larger covalent radius of Ni compared to Cu. This could be due to Ni-W bond with more covalent character, or due to an enhanced W→Ni dative from the more electron rich [Wp*]− compared to Wp−. Otherwise, the two complexes are roughly isostructural.
Table 1.
Comparison of Ni-W and Cu-W heterobinuclear complexes.
| (NHC)Ni-WCpR(CO)3 | (NHC)Cu-WCp(CO)3c | |
|---|---|---|
| d(M-W) (Å)a | 2.4977(10) | 2.5599(6) |
| d(M···CO) (Å)a | 2.021(9), 2.072(9), 3.364(9) | 2.294(7), 2.280(5), 3.858(6) |
| ∠(CNHC-M-W) (°)a | 156.8(2) | 165.7(1) |
| νCO (cm−1)a | 1733, 1760, 1891 | 1784, 1818, 1920 |
| q(M)b | 0.42 | 0.40 |
| q(W)b | −0.89 | −0.95 |
| q(IMeM)a | 0.57 | 0.69 |
| q(WCpR(CO)3)b | −0.57 | −0.69 |
| M-W Wiberg bond indexb | 0.31 | 0.29 |
Solid-state data for NHC = IPr and CpR = η5-C5Me5 (M = Ni) or CpR = η5-C5H5 (M = Cu).
Charges (q) and bond indices from NBO calculations for NHC = IMe and CpR = η5-C5H5.
From ref. 1a.
The X-band EPR spectrum of 7 in 2-methyltetrahydrofuran frozen solution is shown in Figure 3. The spectrum is rhombic;32 however, the g value pattern33 is characteristic of a d9 metal ion with a singly-occupied molecular orbital (SOMO) of dx2-y2 character, as typically seen for CuII, and of particular relevance here for NiI.34 Particularly notable are two weak, satellite features associated only with gmid. We assign these to a hyperfine-split doublet arising from coupling of the spin primarily on NiI to 183W (I = 1/2, 14% abundant). This hyperfine coupling is very anisotropic. EPR simulations allow a very rough estimate as to the coupling at all g values: A(183W = [30, 390, 120] MHz. The resulting A(183W) tensor gives an isotropic coupling that corresponds to the SOMO having a maximum of ~3% W 6s character.35,36 For comparison, no tungsten hyperfine was resolved in the EPR spectrum of (Ph3P)2Ni-Wp.12 In general, previous examples of low-valent tungsten complexes characterized by EPR spectroscopy are somewhat rare and involve mononuclear complexes. The closest analogue to 7 is Bullock’s CpW(CO)2(IMes) (IMes = N,N’-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene).37 Interestingly, despite Bullock’s complex being assigned as a W(I) metalloradical, no tungsten hyperfine features were resolved in the EPR spectra. In contast, Connelly characterized Tp′W(CO)2(η2-RC≡CR) (Tp′ = hydridorotris(3,5-dimethylpyrazolyl)borate) complexes by EPR spectroscopy and measured an Aiso(183W) value of 128 MHz (for R = Me; EPR was insensitive to R and gave resolved coupling at all g values: A(183W) = –[37,38,53] MHz).38,39 Additionally, Connelly’s complexes gave a different g value pattern: rhombic, with one value below ge (2.00), g = [2.102, 2.015, 1.955]. In contrast, complex 7 gives all g values above ge, and the maximum deviation from ge is much larger. Assuming a limiting W(I) description for both systems, this difference could qualitatively be due to the monotungsten(I) having a SOMO of dyz parentage but 7 having a SOMO of dx2-y2 parentage. Thus, the comparison between Connelly’s monotungsten(I) complexes and heterobinuclear 7 is of limited utility but still indicates some measurable W(I) character of 7.
Figure 3.

X-band EPR spectrum of IPrNiWp* (7) in frozen Me-THF solution (black trace), with simulation (red trace). Experimental conditions: temperature, 77 K; microwave frequency, 9.255 GHz; microwave power, 6.3 mW (15 dB); field modulation amplitude, 0.1 mT; time constant, 1 s; scan time, 240 s. Simulation parameters: S = 1/2, g = [2.038, 2.125, 2.427], W = 25, 40, 60 MHz (gaussian, hwhm); a spectrum with a normalized intensity of 85.7 is added to a spectrum with a normalized intensity of 14.3 that includes hyperfine coupling to 183W (I = 1/2) with A = [30, 390, 120] MHz (assumed collinear with g).
A truncated model of 7, IMeNiWp (7′, IMe = N,N’-dimethylimidazol-2-ylidene), was studied using density functional theory (DFT) calculations. The optimized structure of 7′ includes the presence of two semi-bridging CO ligands and a bent CNHC-Ni-W angle of 164°, although the Ni-W distance is overestimated in the calculation as 2.57 Å. The calculated CO stretching frequencies are 1745, 1773, and 1870 cm−1. In accordance with EPR analysis of 7, the calculated spin density for 7′ (Figure 4) resides mainly on Ni, with Mulliken spin densities of 0.92 on Ni, 0.10 on W, and −0.02 on CNHC (sum = 1.0). Model complex 7′ was also analyzed with natural bond orbital (NBO) calculations, which allowed us to calculate natural charges and bond indices. These values are presented in Table 1 for comparison to previously calculated IMeCuWp.1a Based on partial charges, the Ni-W bond is less polar than the corresponding Cu-W bond; for example, charge difference between [IMeM]+ and Wp− fragments is 1.14e− (M = Ni) vs. 1.38e− (M = Cu). Similarly, the Ni-W bond is calculated to have slightly more covalent character than the corresponding Cu-W bond according to its Wiberg bond index. Collectively, the data are indicative of a dative W0→NiI interaction with slightly more covalent character than the corresponding W0→CuI derivative, thus giving rise to the minor WI-Ni0 contribution indicated by EPR spectroscopy.
Figure 4.

Spin density plot (isovalue = 0001 au) for IMeNiWp (7′) determined by DFT (B3LYP/LANL2TZ(f) for Ni,W/6–31+G(d) for all other atoms).
Reactivity studies.
Recently, we reported the (IPr)CuFp-catalyzed 1,4-hydroboration of pyridines.6b To expand our knowledge on NiI chemistry, we thus conducted preliminary reactivity studies between 7 and various substituted pyridines (Rpy) to form complexes. These reactions were found to give products of the type (IPr)(Rpy)Ni-WCp*(CO)3. For example, addition of 3-chloropyridine to 7 immediately produced the mixtures of 3-chloropyridine adducts, with majority of (IPr)(3-Clpy)NiWp* (9) along with minor [(3-Clpy)Ni(μ-Wp*)]2 (10). We propose that there is an equilibrium between 9 and 10 + IPr (Scheme 3), analogous to the IPr dissociation equilibrium proposed for 7. This behavior is notable, as all the (NHC)Cu-[MCO] derivatives we studied previously were inert towards binding Lewis bases, presumably due to the preference of (NHC)CuX complex to adopt linear coordination geometries. The solid-state structure of 9 (Figure 2) shows a longer Ni-W bond distance (2.5793(8) Å) than in 7. This is a consequence of the increase in coordination number at Ni, and this Ni-W distance comparable to that in (Ph3P)2Ni-Wp (2.574(3) Å).10 The CNHC-Ni-W angle is also more bent in 9 (136.02(16) °) than in 7. The identity of minor component 10 was also determined by X-ray crystallography.
Scheme 3.

Reactivity studies of 7 with 3-chloropyridine.
Heating complex 9 to 150°C for 0.5 h or to 100°C for 12 h resulted in ionic complex 11, in which the IPr ligand has been arylated at CNHC upon C-Cl bond activation of the 3-chloropyridine substrate (Scheme 3). A trace amount of 3,3’-bipyridine was also detected in the product mixture. Monitoring formation of 11 by 1H NMR in toluene-d8 indicated the presence of an intermediate forming on the way to 11, which was determined to be [(3-Clpy)Ni(μ-Wp*)]2 (10). Given that Ni-catalyzed arylation of the CNHC position of free N-heterocyclic carbenes has already been reported,40 it is likely that 10 mediates the arylation of free IPr with concurrent release of [Cp*WCl2(CO)2]−. Collectively, the ease with which IPr dissociates from 7 and 9 should be noted for any future reactivity studies.
CONCLUSIONS
In conclusion, a new series of heteromutinuclear, NiI-containing complexes supported by a N-heterocyclic carbene ligand has been synthesized and characterized crystallographically. The transformations of complexes 1–8 indicate the versatile reactivity of redox-active d9 NiI compared to redox-inert d10 CuI metal sites. Moreover, the paramagnetic heterobimetallic complex 7 was successfully synthesized and further investigated by EPR spectroscopy and DFT calculations, indicating its NiI metalloradical character. While the Ni-W bond in 7 is slightly more covalent than its Cu-W analogue, stoichiometric reactivity data of 7 are consistent with a more labile Ni-NHC bond compared to the inert Cu-NHC bond. Additionally, unlike previously studied heterobinuclear CuI complexes,7,8 complex 7 was found to coordinate pyridines at the NiI metal site, leading to heterobimetallic C-Cl activation of 3-chloropyridine. Further studies towards catalytic applications of 7 and its more detailed spectroscopic characterization are ongoing in our laboratory and will be informed by the results reported here.
EXPERIMENTAL SECTION
General considerations.
All reactions and manipulations were conducted under purified N2 using standard Schlenk line techniques or in a glovebox. Reaction solvents (THF, toluene, diethyl ether, and pentane) were purified using a Glass Contour Solvent System built by Pure Process Technology, LLC. Deuterated solvents (C6D6, CD2Cl2, CDCl3, and toluene-d8) were purchased from Sigma-Aldrich and stored over activated 3-Å molecular sieves prior to use. 1H and 13C NMR spectra were recorded using Bruker Avance 400-MHz or 500-MHz NMR spectrometers. NMR spectra were recorded at room temperature unless otherwise indicated, and chemical shifts were referenced to residual solvent peaks. FT-IR spectra were recorded in a glovebox on powder samples using a Bruker ALPHA spectrometer fitted with a diamond-ATR detection unit. Elemental analyses were performed by Midwest Microlab, LLC, in Indianapolis, IN. Literature methods were used to synthesize NaFp,20, 41 KRp,42 LiWp*,43 and [IPrNi(μ-Cl)]2.21 Although satisfactory combustion analysis data were not obtained after multiple attempts for complexes 4, 6, 7, 8, 9, and 11, NMR and IR data are included in the Supporting Information as indications of purity.
X-ray crystallography.
For each complex, a crystal was selected for data collection, encased in silicone oil, transferred to a nylon loop, and cooled to 100 K. Data collection was carried out using a Bruker D8 QUEST ECO diffractometer under its default manufacturer settings. Standard solution and refinement methods44 within the SHELX package45 were applied as indicated in the supporting CIF files.
EPR spectroscopy.
Frozen solution X-band EPR spectra were recorded on a modified Varian E-9 spectrometer using a liquid nitrogen finger dewar. Room temperature EPR spectra were attempted, but gave only very weak signals, so that no aiso(183W) value (i.e., the true isotropic value, as opposed to the average of the A matrix components in frozen solution) could be determined, in contrast to Connelly and co-workers who were able to record fluid solution (toluene) spectra.38 EPR spectra were simulated using the program QPOW, originally by Belford and co-workers,46 as modified by J. Telser.
Computational methods.
DFT calculations were performed with Gaussian09 Revision B.0147 using the built-in hybrid functional, B3LYP.48 The LANL2DZ(f) basis set was used for transition metals,49 and the Gaussian09 internal 6–31+G(d) basis set was used for all other atoms. Input geometries were obtained from the crystallographic coordinates of 7. To reduce the computational cost, the 2,6-di-iso-propylphenyl groups in the structure were replaced by methyl groups. Vibrational frequencies were calculated to verify identification of an energy minimum (zero imaginary frequencies), and the optimized coordinates are provided as Supporting Information. Vibrational frequencies were corrected using standard scaling factors.50
Preparation of [IPrNi(μ-Fp)]2 (1).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (20.8 mg, 0.0215 mmol) and NaFp (9.0 mg, 0.046 mmol) were dissolved in tetrahydrofuran (THF, 4 mL) in a 20-mL scintillation vial and reacted at room temperature for 5 mins. After the reaction completed, the solution was dried in vacuo, resulting in a dark-brown solid residue. The solid was extracted with diethyl ether (Et2O, 2 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in Et2O (1 mL) with few drops of pentane at −35 °C to give dark brown crystals. 1H NMR yield with tetraethoxysilane as internal standard: 41%. 1H NMR (500 MHz, C6D6): δ 7.06 (s, 12H, Ar-H), 6.56 (s, 4H, NCH), 3.88 (s, 10H, Cp), 2.96 (sept., J = 7.0 Hz, 8H, CH(CH3)2), 1.50 (d, J = 7.0 Hz, 24H, CH(CH3)2), 1.06 (d, J = 7.0 Hz, 24H, CH(CH3)2). 13C{1H} NMR (400 MHz, C6D6): δ 194.4 (NCNi), 146.7, 137.1, 130.1, 130.0, 124.3, 123.2, 85.3(Cp*), 28.8, 25.7, 23.4 (Note: no peak corresponding to bound CO was located). IR (solid, cm−1): 1757 (νCO), 1552 (νCO). This mixture contained complexes 1 and 2 in the ratio of 2:1; spectral features from 2 are omitted in the peak lists.
Preparation of IPrNiFp2 (2).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (20.5 mg, 0.0212 mmol) and NaFp (12.8 mg, 0.0640 mmol) were dissolved in THF (5 mL) in a 20-mL scintillation vial and reacted at room temperature for 6 h. The reaction turned dark-brown and then the solution was dried in vacuo, resulting in a dark-brown residue, which was extracted with toluene (2 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of pentane at −35 °C to give brown crystals. Yield: 22.4 mg, 0.0280 mmol, 66%. 1H NMR (500 MHz, C6D6): δ 7.31 (t, J = 8.0 Hz, 2H, p-H), 7.20 (d, J = 8.0 Hz, 4H, m-H), 6.54 (s, 2H, NCH), 4.28 (s, 10H, Cp), 2.86 (sept., J = 6.5 Hz, 4H, CH(CH3)2), 1.44 (d, J = 7.0 Hz, 12H, CH(CH3)2), 1.03 (d, J = 7.0 Hz, 12H, CH(CH3)2). 13C{1H} NMR (500 MHz, C6D6): δ 189.6 (NCNi), 146.3, 136.2, 130.0, 124.2, 124.1, 89.9 (Cp*), 28.8, 25.9, 22.9 (Note: no peak corresponding to bound CO was located). IR (solid, cm−1): 1837 (νCO), 1797 (νCO), 1749 (νCO), 1656 (νCO). Anal. Calcd for C41H46Fe2N2NiO4: C, 61.46; H, 5.79; N, 3.50. Found: C, 60.96; H, 6.35; N, 3.78.
Preparation of IPrNiRp2 (3).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (16 mg, 0.0166 mmol) and KRp (3.0 equiv.) were dissolved in THF (4 mL) in a 20-mL scintillation vial and reacted at 40 °C for 4 h. After the reaction completed, the solution was dried in vacuo, resulting in a dark-brown solid residue. The solid was extracted with toluene (2 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of THF at −35 °C to give brown crystals. Yield: 21.3 mg, 0.0239 mmol, 72%. 1H NMR (500 MHz, C6D6): δ 7.28 (t, J = 8.5 Hz, 2H, p-H), 7.20 (d, J = 7.5 Hz, 4H, m-H), 6.63 (s, 2H, NCH), 4.81 (s, 10H, Cp), 3.04 (sept., J = 7.0 Hz, 4H, CH(CH3)2), 1.49 (d, J = 7.0 Hz, 12H, CH(CH3)2), 1.06 (d, J = 7.0 Hz, 12H, CH(CH3)2).13C{1H} NMR (500 MHz, C6D6): δ 235.6 (CO), 195.3(NCNi), 146.5, 136.2, 130.0, 124.2, 124.1, 90.2 (Cp*), 28.6, 26.3, 23.0. IR (solid, cm−1): 1841 (νCO), 1801 (νCO), 1754 (νCO), 1665 (νCO). Anal. Calcd for C41H46Ru2N2NiO4: C, 55.23; H, 5.20; N, 3.14. Found: C, 55.42; H, 5.94; N, 3.58.
Preparation of [IPrNi(μ-Rp)]2 (4).
In a nitrogen filled glovebox, [(IPrNi)2(μ-Rp)(μ-Cl)] (34.0 mg, 0.0295 mmol) and NaOtBu (4.8 mg, 0.0499 mmol) were dissolved in THF (4 mL) in a 20-mL scintillation vial and reacted at 50 °C for 1 h. After the reaction completed, the solution was dried in vacuo, resulting in a red-brown solid residue that was washed with cold pentane (1 mL). The solid was extracted with toluene (2 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of pentane at −35 °C to give reddish brown crystals. Yield: 34.0 mg, 0.0252 mmol, 85%. 1H NMR (500 MHz, C6D6): δ 7.14–7.07 (m, 12H), 6.54 (s, 4H, NCH), 4.35 (s, 10H, Cp), 2.98 (sept., J = 7.0 Hz, 8H, CH(CH3)2), 1.46 (d, J = 7.0 Hz, 24H, CH(CH3)2), 1.05 (d, J = 7.0 Hz, 24H, CH(CH3)2).13C{1H} NMR (500 MHz, C6D6): δ 195.2 (NCNi), 146.8, 137.1, 130.1, 124.3, 123.1, 87.6 (Cp*), 28.7, 25.7, 23.5 (Note: no peak corresponding to bound CO was located). IR (solid, cm−1): 1753 (νCO), 1554 (νCO).
Preparation of [(IPrNi)2(μ-Rp)(μ-Cl)] (6).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (55.8 mg, 0.0578 mmol) and KRp (32.4 mg, 0.124 mmol) were dissolved at −35 °C in toluene (4 mL) in a 20-mL scintillation vial and reacted allowed to warm from −35 °C to room temperature for 1 h. After the reaction completed, the solution was pipette-filtered through Celite and dried in vacuo, resulting in a brown solid residue that was then washed with pentane (2 × 2 mL). Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of pentane at −35 °C to give brown crystals. Yield: 61.0 mg, 0.0529 mmol, 91%. 1H NMR (500 MHz, C6D6): δ 7.33 (t, J = 7.5 Hz, 4H, p-H), 7.25 (d, J = 7.5 Hz, 8H, m-H), 6.30 (s, 4H, NCH), 4.51 (s, 5H, Cp), 2.98 (sept., J = 7.0 Hz, 8H, CH(CH3)2), 1.52 (d, J = 7.0 Hz, 24H, CH(CH3)2), 1.00 (d, J = 7.0 Hz, 24H, CH(CH3)2).13C{1H} NMR (500 MHz, C6D6): δ 240.9 (CO), 181.8 (NCNi), 146.0, 137.4, 129.7, 124.1, 124.0, 80.5(Cp*), 28.8, 25.4, 23.4. IR (solid, cm−1): 1749 (νCO).
Preparation of IPrNiWp* (7).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (95.0 mg, 0.098 mmol) and LiWp* (162.1 mg, 0.394 mmol) were dissolved in THF (4 mL) in a 20-mL scintillation vial and reacted at 40 °C for 4 h. After the reaction completed, the solution was dried in vacuo, resulting in a dark greenish solid residue that was washed with pentane (2 × 2 mL). The solid was extracted with toluene (4 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of pentane at −35 °C to give green crystals. Yield: 141.2 mg, 0.166 mmol, 84%. 1H NMR (500 MHz, C6D6): δ 14.58 (br,s), 7.99 (br,s), 7.58 (br,s), 6.61 (br,s), 3.71 (br,s,), 2.36 (br,s), 1.37 (br,s). IR (solid, cm−1): 1891 (νCO), 1760 (νCO), 1733 (νCO).
Preparation of [(IPrNi)2(μ-Wp*)(μ-Cl)] (8).
In a nitrogen filled glovebox, [IPrNi(μ-Cl)]2 (20.7 mg, 0.0214 mmol) and LiWp* (18.0 mg, 0.043 mmol) were dissolved in toluene (4 mL) in a 20 ml scintillation vial and reacted at room temperature for 2 h. After the reaction completed, the solution was dried in vacuo, resulting in a dark brown-greenish solid residue that was washed with pentane (2 × 2 mL). The solid was extracted with toluene (2 mL) and pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (1 mL) with few drops of pentane at −35 °C to give brown crystals. Yield: 24.5 mg, 0.0184 mmol, 86%. 1H NMR (500 MHz, C6D6): δ 7.38 (t, J = 7.5 Hz, 4H, p-H), 7.32 (d, J = 7.5 Hz, 8H, m-H), 6.55 (s, 4H, NCH), 3.22 (sept., J = 6.5 Hz, 8H, CH(CH3)2), 1.97 (s, 15H, Cp*), 1.54 (d, J = 7.0 Hz, 24H, CH(CH3)2), 0.98 (d, J = 7.0 Hz, 24H, CH(CH3)2).13C{1H} NMR (500 MHz, C6D6): δ 248.4 (CO), 146.0, 137.8, 129.4, 127.1, 124.1, 99.7(Cp*), 28.7, 26.2, 23.4, 11.6 (Note: no peak corresponding to NCNi was located). IR (solid, cm−1): 1763 (νCO), 1722 (νCO).
Preparation of mixture containing (IPr)(3-Clpy)NiWp* (9) and [(3-Clpy)Ni(μ-Wp*)]2 (10).
In a nitrogen filled glovebox, IPrNiWp* (31.0 mg, 0.0364 mmol) and 3-chloropyridine (4.3 mg, 0.0375 mmol) were dissolved in pentane (4 mL) and toluene (0.5 mL) in a 20-mL scintillation vial, resulting into brown solution immediately. After 5 min, the solution was pipette-filtered through Celite. Crystallization was accomplished by concentrating the filtrate (1.5 mL) and allowing to sit at −35 °C to give brown crystals. Complexes 9 and 10 were separated by sequential recrystallization: crystals of 10 were obtained in the 1st recrystallization (but still contained some impurities); then, the mother liquor was filtered by Celite pipette and provided crystals of 9 in a 2nd recrystallization. Yield of major complex 9: 17.6 mg, 0.0182 mmol, 50%. Complex 9, 1H NMR (500 MHz, C6D6): δ 13.58 (br,s), 7.97 (br,s), 7.75 (br,s), 7.32 (br,s, overlap with d-benzene), 2.73 (br,s), 1.03 (br,s). IR (solid, cm−1): 1852 (νCO), 1749 (νCO), 1716 (νCO).
Yield of minor complex 10: 3.3 mg, 0.0028 mmol, 14%. Complex 10, 1H NMR (500 MHz, C6D6): δ 9.89 (m, 2H, 3-Cl-py), 8.04 (m, 2H, 3-Cl-py), 7.34 (m, 4H, 3-Cl-py), 2.36 (s, 30H, Cp*). IR (solid, cm−1): 1872 (νCO), 1739 (νCO).
Preparation of [IPr(py-3-yl)]+[Cp*WCl2(CO)2]− (11).
In a nitrogen filled glovebox, IPrNiWp* (29.8 mg, 0.035) and 3-chloropyridine (16.1 mg, 0.140 mmol) were dissolved in toluene (1 mL) in a 20-mL scintillation vial, resulting into brown solution that was immediately transferred to a J-Young NMR tube. The reaction was performed at 150 °C for 0.5 h or 100 °C for 12 h, after which the solution was pipette-filtered through Celite. Crystallization was accomplished by leaving a concentrated solution in toluene (0.5 mL) with few drops of pentane at −35 °C. Yield: 16.4 mg, 0.018 mmol, 51%. 1H NMR (500 MHz, CDCl3): δ 7.62 (br), 7.55 (br), 2.23 (br), 2.17 (br, s), 1.97 (br), 1.42 (br), 1.26 (br). IR (solid, cm−1): 2009 (νCO), 1903 (νCO). Note: trace amount of 3,3’-bipyridine was found in the crude mixture, with 1H NMR peaks matching reported values in the literature.51
Supplementary Material
ACKNOWLEDGMENT
This material is based upon work supported by the U.S. Department of Energy (DOE), Office of Science, Office of Basic Energy Sciences (BES), under Award Number DE-SC0021055. Instrumentation for X-ray crystallography was funded, in part, by the National Institutes of Health under grant R01 GM116820. We thank Prof. Brian M. Hoffman (Northwestern University) for use of the EPR spectrometer, which is supported by the US DOE, Office of Science, BES, under Award Number DE-SC0019342. Computational resources and services were provided by the Advanced Cyberinfrastructure for Education and Research (ACER) group at UIC.
Footnotes
The authors declare no competing financial interests.
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website.
Spectral data and computational output (PDF)
Crystallographic data (CIF)
REFERENCES
- 1.(a) Buchwalter P; Rose J; Braunstein P Multimetallic Catalysis Based on Heterometallic Complexes and Clusters. Chem. Rev. 2015, 115, 28–126. [DOI] [PubMed] [Google Scholar]; (b) Campos J Bimetallic Cooperation across the Periodic Table. Nat. Rev. Chem. 2020, 4, 696–702. [DOI] [PubMed] [Google Scholar]; (c) Pye D, R. D; Mankad NP Bimetallic catalysis for C–C and C–X coupling reactions. Chem. Sci. 2017, 8, 1705–1718. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) . Kim, Bin U; Jung DJ; Jeon HJ; Rathwell K; Lee S Synergistic Dual Transition Metal Catalysis. Chem. Rev. 2020, 120, 13382–13433. [DOI] [PubMed] [Google Scholar]; (e) Gunanathan C; Milstein D Metal–Ligand Cooperation by Aromatization–Dearomatization: A New Paradigm in Bond Activation and “Green” Catalysis. Acc. Chem. Res. 2011, 44, 588–602. [DOI] [PubMed] [Google Scholar]; (f) Powers IG; Uyeda C Metal−Metal Bonds in Catalysis. ACS Catal. 2017, 7, 936–958. [Google Scholar]; (g) Allen AE; MacMillan DWC Synergistic Catalysis: A Powerful Synthetic Strategy for New Reaction Development. Chem Sci 2012, 3, 633– 658. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Lee JM; Na Y; Han H; Chang S Cooperative Multi-Catalyst Systems for One-Pot Organic Transformations. Chem. Soc. Rev. 2004, 33, 302–311. [DOI] [PubMed] [Google Scholar]; (i) Karunananda MK; Mankad NP Cooperative Strategies for Catalytic Hydrogenation of Unsaturated Hydrocarbons. ACS Catal. 2017, 7, 6110–6119. [Google Scholar]
- 2.Yu H-C; Mankad NP Catalytic Reactions by Heterobimetallic Carbonyl Complexes with Polar Metal-Metal Interactions. Synthesis 2021, 53, 1409–1422. [Google Scholar]
- 3.(a) Reed SA; White MC Catalytic Intermolecular Linear Allylic C-H Amination via Heterobimetallic Catalysis. J. Am. Chem. Soc. 2008, 130, 3316–3318. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Mazzacano TJ; Mankad NP Base Metal Catalysts For Photochemical C–H Borylation That Utilize Metal-Metal Cooperativity. J. Am. Chem. Soc. 2013, 135, 17258–17261. [DOI] [PubMed] [Google Scholar]
- 4.(a) Getzler YDYL; Mahadevan V; Lobkovsky EB; Coates GW Synthesis of Beta-Lactones: a Highly Active and Selective Catalyst for Epoxide Carbonylation. J. Am. Chem. Soc. 2002, 124, 1174–1175. [DOI] [PubMed] [Google Scholar]; (b) Mulzer M; Whiting BT; Coates GW Regioselective Carbonylation of trans-Disubstituted Epoxides to β-Lactones: A Viable Entry into syn-Aldol-Type Products. J. Am. Chem. Soc. 2013, 135, 10930–10933. [DOI] [PubMed] [Google Scholar]; (c) Hubbell AK; LaPointe AM; Lamb JR; Coates GW Regioselective Carbonylation of 2,2-Disubstituted Epoxides: An Alternative Route to Ketone-Based Aldol Products. J. Am. Chem. Soc. 2019, 141, 2474–2480. [DOI] [PubMed] [Google Scholar]; (d) Pye DR; Cheng L-J; Mankad NP Cu/Mn Bimetallic Catalysis Enables Carbonylative Suzuki–Miyaura Coupling with Unactivated Alkyl Electrophiles. Chem. Sci. 2017, 8, 4750–4755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.(a) Yao C; Dahmen T; Gansauer A; Norton J Anti-Markovnikov Alcohols via Epoxide Hydrogenation through Cooperative Catalysis. Science 2019, 364, 764–767. [DOI] [PubMed] [Google Scholar]; (b) Karunananda MK; Mankad NP E-Selective Semi-Hydrogenation of Alkynes by Heterobimetallic Catalysis. J. Am. Chem. Soc. 2015, 137, 14598–14601. [DOI] [PubMed] [Google Scholar]; (c) Cammarota RC; Clouston LJ; Lu CC Leveraging Molecular Metal–Support Interactions for H2 and N2 Activation. Coord. Chem. Rev. 2017, 334, 100–111. [Google Scholar]
- 6.(a) Cheng L-J; Mankad NP Heterobimetallic Control of Regioselectivity in Alkyne Hydrostannylation: Divergent Syntheses of α- and (E)–β–Vinylstannanes via Cooperative Sn–H Bond Activation. J. Am. Chem. Soc. 2019, 141, 3710–3716. [DOI] [PubMed] [Google Scholar]; (b) Yu H-C; Islam SM; Mankad NP Cooperative Heterobimetallic Substrate Activation Enhances Catalytic Activity and Amplifies Regioselectivity in 1,4-Hydroboration of Pyridines. ACS Catal. 2020, 10, 3670–3675. [Google Scholar]
- 7.Mankad NP Diverse Bimetallic Mechanisms Emerging From Transition Metal Lewis Acid/Base Pairs: Development Of Co-Catalysis With Metal Carbenes And Metal Carbonyl Anions. Chem. Commun. 2018, 54, 1291–1302. [DOI] [PubMed] [Google Scholar]
- 8.(a) Jayarathne U; Mazzacano TJ; Bagherzadeh S; Mankad NP Heterobimetallic Complexes with Polar, Unsupported Cu–Fe and Zn–Fe Bonds Stabilized by N-Heterocyclic Carbenes. Organometallics 2013, 32, 3986–3992. [Google Scholar]; (b) Karunananda MK; Vázquez FX; Alp EE; Bi W; Chattopadhyay S; Shibata T; Mankad NP Experimental Determination of Redox Cooperativity and Electronic Structures in Catalytically Active Cu-Fe and Zn-Fe Heterobimetallic Complexes. Dalt. Trans. 2014, 43, 13661–13671. [DOI] [PubMed] [Google Scholar]
- 9.Karunananda M; Parmelee S; Waldhart G; Mankad N Experimental and Computational Characterization of the Transition State for C–X Bimetallic Oxidative Addition at a Cu–Fe Reaction Center. Organometallics 2015, 34, 3857–3864. [Google Scholar]
- 10.(a) Karunananda M; Mankad N Heterobimetallic H2 Addition and Alkene/Alkane Elimination Reactions Related to the Mechanism of E-Selective Alkyne Semihydrogenation. Organometallics 2017, 36, 220–227. [Google Scholar]; (b) Zhang Y; Karunananda MK; Yu H-CC; Clark KJ; Williams W; Mankad NP; Ess DH Dynamically Bifurcating Hydride Transfer Mechanism and Origin of Inverse Isotope Effect for Heterodinuclear AgRu-Catalyzed Alkyne Semihydrogenation. ACS Catal. 2019, 9, 2657–2663. [Google Scholar]
- 11.Parmelee SR; Mazzacano TJ; Zhu Y; Mankad NP; Keith JA A Heterobimetallic Mechanism for C–H Borylation Elucidated from Experimental and Computational Data. ACS Catal. 2015, 5, 3689–3699. [Google Scholar]
- 12.Bagherzadeh S; Mankad NP Catalyst Control of Selectivity in CO2 Reduction Using a Tunable Heterobimetallic Effect. J. Am. Chem. Soc. 2015, 137, 10898–10901. [DOI] [PubMed] [Google Scholar]
- 13.(a) Gade LH Highly Polar Metal–Metal Bonds in “Early–Late” Heterodimetallic Complexes. Angew. Chem. Int. Ed. 2000, 39, 2658–2678. [DOI] [PubMed] [Google Scholar]; (b) King RB Perspectives in the Syntheses of Novel Organometallic Compounds Using Metal Carbonyl Anions. J. Organomet. Chem. 1975, 100, 111–125. [Google Scholar]
- 14.(a) Mazzacano TJ; Leon NJ; Waldhart GW; Mankad NP Fundamental Organometallic Chemistry under Bimetallic Influence: Driving β-Hydride Elimination and Diverting Migratory Insertion at Cu and Ni. Dalton Trans. 2017, 46, 5518–5521. [DOI] [PubMed] [Google Scholar]; (b) Levina VA; Rossin A; Belkova NV; Chierotti MR; Epstein LM; Filippov OA; Gobetto R; Gonsalvi L; Lledós A; Shubina ES; Zanobini F; Peruzzini M Acid-Base Interaction between Transition-Metal Hydrides: Dihydrogen Bonding and Dihydrogen Evolution. Angew. Chemie Int. Ed. 2010, 50, 1367–1370. [DOI] [PubMed] [Google Scholar]
- 15.Chipman JA; Berry JF Paramagnetic Metal-Metal Bonded Heterometallic Complexes. Chem. Rev. 2020, 120, 2409–2447. [DOI] [PubMed] [Google Scholar]
- 16.Lei H; Guo J-D; Fettinger JC; Nagase S; Power PP Two-Coordinate First Row Transition Metal Complexes with Short Unsupported Metal−Metal Bonds. J. Am. Chem. Soc. 2010, 132, 17399–17401. [DOI] [PubMed] [Google Scholar]
- 17.(a) Carlton L; Llndsell WE; McCullough KJ; Preston PN Preparation, Structure, and Properties of Paramagnetic, Heteroblnuclear Complexes Containing Nickel and Molybdenum or Tungsten. X-ray Crystal Structure of [NiW(CO)3(PPh3)2(η-C5H5)] Organometallics 1985, 4, 1138–1140. [Google Scholar]; (b) Carlton L; Lindsell WE; McCullough KJ; Preston PN Heteronuclear Complexes Containing Group 6 Transition Metals; Chemical, Spectroscopic, and Theoretical Studies on some Binuclear Complexes and the X-Ray Crystal Structure Determination of [NiW(CO)3(PPh3)2(η5-C5H5)]. J. Chem. Soc., Dalton Trans, 1987, 2741–2749. [Google Scholar]
- 18.(a) Lin C-Y; Power PP Complexes of Ni(I): A “Rare” Oxidation State of Growing Importance. Chem. Soc. Rev. 2017, 46, 5347–5399. [DOI] [PubMed] [Google Scholar]; (b) Zimmermann P; Limberg C Activation of Small Molecules at Nickel(I) Moieties. J. Am. Chem. Soc. 2017, 139, 4233–4242. [DOI] [PubMed] [Google Scholar]
- 19.(a) Zhang K; Conda-Sheridan M; Cooke SR; Janis Louie J N-Heterocyclic Carbene Bound Nickel(I) Complexes and Their Roles in Catalysis. Organometallics 2011, 30, 2546–2552. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Tasker SZ; Standley EA; Jamison TF Recent Advances in Homogeneous Nickel Catalysis. Nature 2014, 509, 299−309. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Matsubara K; Fukahori Y; Inatomi T; Tazaki S; Yamada Y; Koga Y; Kanegawa S; Nakamura T Monomeric Three-Coordinate N-Heterocyclic Carbene Nickel(I) Complexes: Synthesis, Structures, and Catalytic Applications in Cross-Coupling Reactions. Organometallics 2016, 35, 3281−3287. [Google Scholar]; (d) Dgrr AB; Fisher HC; Kalvet I; Truong K-N; Schoenebeck F Divergent Reactivity of a Dinuclear (NHC)Nickel(I) Catalyst versus Nickel(0) Enables Chemoselective Trifluoromethylselenolation. Angew. Chem. Int. Ed. 2017, 56, 13431 –13435. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Matsubara K; Yamamoto H; Miyazaki S; Inatomi T; Nonaka K; Koga Y; Yamada Y; Veiros LF; Kirchner K Dinuclear Systems in the Efficient Nickel-Catalyzed Kumada−Tamao−Corriu Cross-Coupling of Aryl Halides. Organometallics 2017, 36, 255−265. [Google Scholar]; (f) Inatomi T; Fukahori Y; Yamada Y; Ishikawa R; Kanegawa S; Koga Y; Matsubara K Ni(I)-Ni(III) Cycle in Buchwald-Hartwig Amination of Aryl Bromide Mediated by NHC-Ligated Ni(I) Complexes. Catal. Sci. Technol. 2019, 9, 1784–1793. [Google Scholar]; (g) Horn B; Limberg C; Herwig C; Braun B Nickel(I)-Mediated Transformations of Carbon Dioxide in Closed Synthetic Cycles: Reductive Cleavage and Coupling of CO2 Generating NiICO, NiIICO3 and NiIIC2O4NiII Entities. Chem. Commun. 2013, 49, 10923–10925. [DOI] [PubMed] [Google Scholar]; (h) Yin H; Fu GC Mechanistic Investigation of Enantioconvergent Kumada Reactions of Racemic α-Bromoketones Catalyzed by a Nickel/Bis(Oxazoline) Complex. J. Am. Chem. Soc. 2019, 141, 15433–15440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.(a) Banerjee S; Karunananda MK; Bagherzadeh S; Jayarathne U; Parmelee SR; Waldhart GW; Mankad NP Synthesis and Characterization of Heterobimetallic Complexes with Direct Cu–M Bonds (M = Cr, Mn, Co, Mo, Ru, W) Supported by N-Heterocyclic Carbene Ligands: A Toolkit for Catalytic Reaction Discovery. Inorg. Chem. 2014, 53, 11307–11315. [DOI] [PubMed] [Google Scholar]; (b) Jayarathne U; Mazzacano TJ; Bagherzadeh S; Mankad NP Heterobimetallic Complexes with Polar, Unsupported Cu–Fe and Zn–Fe Bonds Stabilized by N-Heterocyclic Carbenes. Organometallics 2013, 32, 3986–3992. [Google Scholar]; (c) Jayarathne U; Parmelee S; Mankad N Small Molecule Activation Chemistry of Cu–Fe Heterobimetallic Complexes Toward CS2 and N2O. Inorg. Chem. 2014, 53 (14), 7730–7737. [DOI] [PubMed] [Google Scholar]
- 21.Dibble BR; Sigman MS; Arif AM Oxygen-Induced Ligand Dehydrogenation of a Planar Bis-μ-Chloronickel(I) Dimer Featuring an NHC Ligand. Inorg. Chem. 2005, 44, 3774–3776. [DOI] [PubMed] [Google Scholar]
- 22.(a) Akita M; Terada M; Tanaka M; Moro-oka Y Interaction of Iron Ethynyl and Ethynediyl Complexes with Cp2Ni2(CO)2: Formation of Tri- and Tetranuclear Adducts with a Multiply Bridging C2H ligand. Organometallics 1992, 11, 3468–3472. [Google Scholar]; (b) Colbran SB; Robinson BH; Simpson J Synthesis, Structure, and Redox Properties of [(η-C5H5)Fe{η-σ:η4-NiC4R4(η-C5H5)}]. A Ferrocene Analogue with a Nickelapentadiene Ring. Organometallics 1985, 4, 1594–1601. [Google Scholar]; (c) Yempally V; Zhu L; Captain B A Bimetallic Iron–Nickel Cluster from the Reaction of Nickelocene with a Pentairon Carbonyl Carbide Cluster Complex. J. Clust. Sci. 2009, 20, 695–705. [Google Scholar]; (d) Pergolla RD; Fumagalli A; Garlaschelli L; Manassero C; Manassero M; Sansoni M; Sironi A Iron–Nickel Mixed Metal Nitrido Clusters: Synthesis and Solid State Structure of the Anions [HFe5NiN(CO)14]2− and [HFe4Ni2N(CO)13]2−. Inorg. Chim. Acta. 2008, 361, 1763–1769. [Google Scholar]; (e) Sappa E Synthesis and Crystal Structure of (η5-C5H5)NiFe3(CO)7-(μ-PPh2)(μ4,η2-HC≡CPri): a Butterfly Cluster with an Unprecedented Metal Core Stoichiometry. J. Organomet. Chem.1989, 359, 419–428. [Google Scholar]; (f) Pauling L Metal-Metal Bond Lengths in Complexes of Transition Metals. Proc. Nati. Acad. Sci. USA 1976, 73, 4290–4293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Uehara K; Hikichi S; Inagaki A; Akita M Xenophilic Complexes Bearing a TpR Ligand, [TpRM–M′Ln] [TpR=TpiPr2,Tp# (TpMe2,4-Br); M = Ni, Co, Fe, Mn; M′Ln = Co(CO)4, Co(CO)3(PPh3), RuCp(CO)2]: The Two Metal Centers are Held Together not by Covalent Interaction but by Electrostatic Attraction. Chem. Eur. J. 2005, 11, 2788–2809. [DOI] [PubMed] [Google Scholar]
- 24.(a) Jeynes TP; Cifuentes MP; Humphrey MG Ruthenium Carbonyl Cluster Complexes with Oxygen Ligands. Reactions between Ru3(CO)12 and 4-Methoxyphenol or 2-Naphthol. Crystal Structure of Ru4(μ3-OC6H4OMe-4)2(μ-Cl)(μ-OC6H4OMe-4)(CO)10, an Unusual Mixed-Valence Cluster Complex. J. Organomet. Chem.1994, 476, 133–143. [Google Scholar]; (b) Patra SK; Majumdar M; Bera JK Ligand Assisted Homolytic Cleavage of the Ru–Ru Single Bond in [Ru2(CO)4]2+ Core and the Chemical Consequence. J. Organomet. Chem.2006, 691, 4779–4787. [Google Scholar]
- 25.Milosevic S; Brenner E; Ritleng V;Chetcuti MJ Unsaturated Dinickel–molybdenum Clusters with N-heterocyclic Carbene Ligands. Dalton Trans. 2008, 1973–1975. [DOI] [PubMed] [Google Scholar]
- 26.(a) Curtis MD; Han KR; Butler WM Metal-Metal Multiple Bonds. 5. Molecular Structure and Fluxional Behavior of Tetraethylammonium μ-cyano-bis(cyclopentadienyldicarbonylmolybdate)(Mo-Mo) and the Question of Semibridging Carbonyls. Inorg. Chem. 1980, 19, 2096–2101. [Google Scholar]; (b) Parmelee SR; Mankad NP A Data-Intensive Re-evaluation of Semibridging Carbonyl Ligands. Dalton Trans. 2015, 44, 17007–17014. [DOI] [PubMed] [Google Scholar]
- 27.Mitschler A; Rees B; Lehmann MS Electron Density in Bis(Dicarbonyl-π-Cyclopentadienyliron) at Liquid Nitrogen Temperature by X-Ray and Neutron Diffraction. J. Am. Chem. Soc. 1978, 100, 3390–3397. [Google Scholar]
- 28.Straub T; Haukka M; Pakkanen TA Unbridged Homo and Hetero Dinuclear Complexes of Group 6 and 8 Metals: Synthesis, Characterization and Comparison of X-Ray Crystallographic Data. J. Organomet. Chem. 2000, 612, 106–116. [Google Scholar]
- 29.Lee CH; Laitar DS; Mueller P; Sadighi JP Generation of a Doubly Bridging CO2 Ligand and Deoxygenation of CO2 by an (NHC)Ni(0) Complex. J. Am. Chem. Soc. 2007, 129, 13802–13803. [DOI] [PubMed] [Google Scholar]
- 30.Tilset M; Parker VD Solution Homolytic Bond Dissociation Energies of Organotransition-Metal Hydrides. J. Am. Chem. Soc. 1989, 111, 6711–6717. [Google Scholar]
- 31.Chetcuti MJ; Grant BE Unsaturated Mixed-Metal Complexes: Syntheses of (η-C5Me5)N¡-M(CO)3(η-C5H5) (M = Mo, W) and X-ray Diffraction Study of (η-C5Me5)Ni-W(CO)3(η-C5H4Me). Organometallics 1990, 9, 1343–1345. [Google Scholar]
- 32. where 0.5 would be perfectly rhombic; gavg = (gmax + gmid + gmin)/3 = 2.197
- 33.g|| >> g⊥ > 2.0;
- 34.Telser J, Overview of Ligand versus Metal Centered Redox Reactions in Retraaza Macrocyclic Complexes of Nickel with a Focus on Electron Paramagnetic Resonance Studies. J. Braz. Chem. Soc. 2010, 21, 1139–1157. [Google Scholar]
- 35.Morton JR; Preston KF, Atomic Parameters for Paramagnetic Resonance Data. J. Magn. Reson. 1978, 30, 577–582. [Google Scholar]
- 36.A = [30, 390, 120] MHz, aiso = (30 + 390 + 120)/3 = 180 MHz; aiso(183W 6s1) = 5777 MHz. This rough calculation assumes that the A components all have the same sign, the signs not being obtainable from EPR alone. In ref 37, Connelly and co-workers were able to impute a negative sign for these based on a ligand-field analysis that is inappropriate here. The minimum aiso = 80 MHz would occur if A = [−30, 390, −120] MHz, and would correspond to only ~1.4% of a spin.
- 37.Roberts JAS; Franz JA; Van Der Eide EF; Walter ED; Petersen JL; Dubois DL; Bullock RM Comproportionation of Cationic and Anionic Tungsten Complexes Having an N-Heterocyclic Carbene Ligand to Give the Isolable 17-Electron Tungsten Radical CpW(CO)2(IMes)•. J. Am. Chem. Soc. 2011, 133, 14593–14603. [DOI] [PubMed] [Google Scholar]
- 38.Adams CJ; Bartlett IM; Boonyuen S; Connelly NG; Harding DJ; Hayward OD; McInnes EJL; Orpen AG; Quayle MJ; Rieger PH Structural Consequences of the One-Electron Reduction of d4 [Mo(CO)2(η-PhC≡CPh)Tp′]+ and the Electronic Structure of the d5 Radicals [M(CO)L(η-MeC≡CMe)Tp′] {L = CO and P(OCH2)3 CEt}. Dalton Trans. 2006, 2, 3466–3477. [DOI] [PubMed] [Google Scholar]
- 39.It is noteworthy that Connelly’s report (ref 37) included [Tp′W(CO){P(OCH2)3CEt}(η-MeC≡CMe)] that gave similar EPR parameters as the dicarbonyl, g = [2.126, 2.002, 1.951], A(183W)=−[120, 108, 177] MHz, but also with hyperfine coupling from the phosphite: A(31P) = 33 MHz (isotropic).
- 40.Ho NKT; Neumann B; Stammler H-G; Silva VHM; Watanabe DG; Braga AAC; Ghadwal RS Nickel-Catalysed Direct C2-Arylation of N-Heterocyclic Carbenes. Dalton Trans. 2017, 46, 12027–12031. [DOI] [PubMed] [Google Scholar]
- 41.Ohishi T; Shiotani Y; Yamashita M A Convenient One-Flask Preparation of Pure Potassium Cyclopentadienyldicarbonylferrate, K[(η5-C5H5)Fe(CO)2]. J. Org. Chem. 1994, 59, 250. [Google Scholar]
- 42.Kalz KF; Kindermann N; Xiang S-Q; Kronz A; Lange A; Meyer F Revisiting the Synthesis and Elucidating the Structure of Potassium Cyclopentadienyldicarbonylruthenate, K[CpRu(CO)2]. Organometallics 2014, 33, 1475–1479. [Google Scholar]
- 43.Einstein FWB; Jones RH; Dryden NH; Legzdins P Synthesis and Characterization of (η5-C5Me5)W(NO)I2, a Monomeric, Formally 16-Electron Complex. Can. J. Chem. 1988, 66, 2100–2103. [Google Scholar]
- 44.Müller P Practical Suggestions for Better Crystal Structures. Crystallogr. Rev. 2009, 15, 57–83. [Google Scholar]
- 45.Sheldrick GM A Short History of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [DOI] [PubMed] [Google Scholar]
- 46.(a) Belford RL; Belford GG, Eigenfield Expansion Technique for Efficient Computation of Field-Swept Fixed-Frequency Spectra from Relaxation Master Equations. J. Chem. Phys. 1973, 59, 853–854. [Google Scholar]; (b) Belford RL; Nilges MJ In Computer Simulations of Powder Spectra, EPR Symposium, 21st Rocky Mountain Conference, Denver, Colorado, Denver, Colorado, August,1979. [Google Scholar]
- 47.Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Montgomery JA Jr., Peralta JE, Ogliaro F, Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN, Keith T, Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas O, Foresman JB, Ortiz JV, Cioslowski J, and Fox DJ, Gaussian 09, revision B. 01. Wallingford CT: (2009). [Google Scholar]
- 48.(a) Becke AD Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98 5648–5652. [Google Scholar]; (b) Lee C; Yang W; Parr RG Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- 49.(a) Hay PJ; Wadt WR Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 213–299. [Google Scholar]; (b) Wadt WR; Hay PJ Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for Main Group Elements Na to Bi. J. Chem. Phys. 1985, 82, 216–284. [Google Scholar]; (c) Hay PJ; Wadt WR Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270. [Google Scholar]
- 50.Alecu IM; Zheng J; Zhao Y; Truhlar DG Computational Thermochemistry: Scale Factor Databases and Scale Factors for Vibrational Frequencies Obtained from Electronic Model Chemistries, J. Chem. Theory Comput. 2010, 6, 2872–2887. [DOI] [PubMed] [Google Scholar]
- 51.Gao G-L; Xia W; Jain P; Yu J-Q Pd(II)-Catalyzed C3-Selective Arylation of Pyridine with (Hetero)arenes. Org. Lett. 2016, 18, 744–747. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
