ACS Publications. Most Trusted. Most Cited. Most Read
Monodentate σ-Accepting Boron-Based Ligands Bearing Square-Planar Ni(0) Centers
My Activity

Figure 1Loading Img
  • Open Access
Article

Monodentate σ-Accepting Boron-Based Ligands Bearing Square-Planar Ni(0) Centers
Click to copy article linkArticle link copied!

  • Yutaka Mondori
    Yutaka Mondori
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
  • Yasuhiro Yamauchi
    Yasuhiro Yamauchi
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
  • Takahiro Kawakita
    Takahiro Kawakita
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
  • Sensuke Ogoshi
    Sensuke Ogoshi
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
  • Yuta Uetake*
    Yuta Uetake
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    Innovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, Suita, Osaka 565-0871, Japan
    *Email: [email protected]
    More by Yuta Uetake
  • Yasuo Takeichi
    Yasuo Takeichi
    Department of Applied Physics, Graduate School of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
  • Hidehiro Sakurai
    Hidehiro Sakurai
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    Innovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, Suita, Osaka 565-0871, Japan
  • Yoichi Hoshimoto*
    Yoichi Hoshimoto
    Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    Center for Future Innovation (CFi), Division of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    *Email: [email protected]
Open PDFSupporting Information (3)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2025, 147, 10, 8326–8335
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.4c15892
Published February 28, 2025

Copyright © 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Transition metals are known to work as electron donors toward electron-accepting heavier-group-13 elements (Al, Ga, and In), called Z-type ligands. However, complexes with boron-based Z-type ligands are stable only in the presence of additional coordination units (the so-called “supported-ligand” strategy). Here, we report the synthesis and characterization of square-planar Ni(0) complexes that bear tris(perfluoroaryl)boranes as monodentate Z-type ligands, even though such coordination geometry has been traditionally associated with Ni(II) species based on the well-established ligand-field theory. A combined theoretical and experimental approach revealed a mixed covalent/dative character for the Ni–B bonds. This strategy uses frustrated L/Z-ligand pairs that combine sterically encumbered electron-donating (L-type) and electron-accepting ligands to form noncovalent interactions over L–M–Z units to achieve unprecedented low-valent transition metal species with monodentate Z-type ligands.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2025 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

The concept of electron acceptors (i.e., Lewis acids; LAs) and donors (i.e., Lewis bases; LBs) has played a pivotal role in understanding chemical bonds between molecular fragments in various fields of chemistry (Figure 1a, top). (1) As a representative example, transition metal complexes, which are employed in the synthesis of pharmaceuticals, polymers, and commodities, (2,3) are constructed via the sharing of electrons between transition metals (TM) as LAs and ligands as LBs. Such TM ← LB interactions involve the transfer of electron(s) from an electron-occupied orbital (e.g., s, p, and their hybridized orbitals) in the LB to an empty d-orbital of TM (Figure 1a, middle); the arrow indicates the direction of the donor-to-acceptor-transfer of electrons. TM can simultaneously act as an electron donor via so-called back-donation when adequate orbital(s) in the ligands, such as antibonding σ* and π* orbitals, are available to accept electrons from the filled d-orbital in TM, leading to the formation of TM ⇄ LB bonds. (4) Additionally, electron donation from the filled d-orbitals in TM to empty p orbitals of main-group LAs (i.e., TM → LA) has also been known since the 1960s, although examples remain limited (Figure 1a, bottom). (5−10) These Lewis-acidic σ-accepting ligands are often referred to as Z-type ligands (formal two-electron acceptors) based on the number of electrons donated/accepted to/from TM, as proposed by Green (11) and Parkin, (9) while formal one- and two-electron donors are known as X- and L-types, respectively.

Figure 1

Figure 1. Concept of electron acceptors and donors. (a) A simplified representation of complexation between Lewis acids (LAs) and Lewis bases (LBs) and the related orbitals. (b) Coordination of transition metals (TM) to trivalent group-13 species. (c) Pioneering examples of Ni complexes that bear LnZ-type borane ligands. (d) A strategy for monodentate Z-type borane ligands based on frustrated L/Z-ligand pairs (this work).

Trivalent group-13 molecules, i.e., EX3 (E = B, Al, Ga, In), are typical LAs, as the empty p orbitals on the central E atom can accept electrons from various LBs, including main-group (12,13) and transition metals (Figure 1b). Since Burlitch, Hughes and co-workers crystallographically identified the presence of the Fe → Al bonding interaction in [Et4N][Fe(AlPh3)(cyclopentadienyl)(CO)2], (14) EX3 ligands have been employed as monodentate, unsupported Z-type ligands to accept electrons from a variety of transition metals. (5,7,10,15) However, the formation of the unsupported TM → BX3 interaction remains elusive due to the absence of crystallographical evidence, even though such a bonding situation has often been proposed based on experimental results using NMR and/or IR spectroscopy techniques. (5,7,16,17) In this context, Burlitch, Hughes, and co-workers proposed, based on IR and NMR analyses, the formation of Fe → BPh3 interactions after the reaction between [NEt4][(η5-cyclopentadienyl)Fe(CO)2] and BPh3; (17) however, even though the formation of a four-coordinated boron species was suggested, such spectroscopic analyses still remain questionable when discussing the existence of TM → BX3 interactions, as we demonstrate in this manuscript (see our results on e.g., 2a, 6a, and 7g). Remarkably, Campos and co-workers have recently reported that a Rh(I) complex bearing HB(C6F5)2 includes a bonding interaction between the Rh center and the B–H bond, and proposed that this complex may be potentially described as a transition metal complex bearing an unsupported borane ligand. (18) Based on our experience in custom-tailoring the Lewis acidity of triarylboranes (X = aryl) (19−21) and the report of Frenking (22) and co-workers, we are convinced that the striking differences in the reactivity of BX3 and its heavier analogues can be explained by the energetic penalty for the geometric deformation (Edef) from the trigonal-planar to the tetrahedral conformation via the formation of TM–EX3 bonds (Figure 1b). In fact, the Edef values for BX3 (X = halogen, alkyl, aryl) have been predicted to be at least twice those of the corresponding Al, Ga, and In analogues when triethylphosphine oxide was used as a model LB. (23) One strategy to compensate for this high Edef and to stabilize the tetrahedral TM–BX3 motif is the use of boranes that are equipped with additional L-type coordination units such as phosphines. In 1999, Hill and co-workers reported the synthesis and identification of several transition metal complexes that bear B(mt)3 (mt = 2-sulfanyl-1-methylimidazole) as a multidentate L3Z-type ligand. (24,25) However, these authors also pointed out the significant contribution of an imidazole-coordinated σ-boryl canonical form (an L2X2-type) of the B(mt)3 ligand; thus, a careful discussion of the nature of TM–borane bonds is required. (26,27) Later, Bourissou and co-workers successfully demonstrated that triarylboranes that bear di- or triphosphanyl moieties can serve as an adequate platform to construct supported TM–borane complexes (6) such as Ni/B-1 (28) (Figure 1c). Moreover, Ni/B-2 and Ni/B-3 have been synthesized by Emslie (29) and Peters, (30) respectively, and both have been proposed to include an η3-BCC coordination of an aryl borane unit to the nickel center. These pioneering results demonstrated that BX3 has sufficient electrophilicity to accept electrons donated from TM if the energetic penalty in the geometry deformation is effectively compensated. Nevertheless, an approach to successfully realize unsupported, monodentate Z-type borane ligands has remained elusive so far.
Herein, we report the synthesis of square-planar nickel(0) complexes that bear boranes as monodentate Z-type ligands via L-to-Z ligand substitution on the nickel(0) carbonyl complexes (Figure 1d). We found that the combination of sterically encumbered N-heterocyclic carbenes (NHCs) as L-type ligands and tris(perfluoroaryl)boranes (Bn) as Z-type ligands (i.e., frustrated L/Z-ligand pairs, named in analogy to the concept of frustrated Lewis pairs, FLPs) (31,32) is critical to generate these elusive species. The participation of multiple noncovalent interactions (NCI) over the NHC–TM–Bn units is discussed based on a systematic theoretical analysis.

Results and Discussion

Click to copy section linkSection link copied!

We have reported the formation of heterobimetallic Ni/Al complexes that bear N-phosphine-oxide-substituted imidazolinylidene L1 (Figure 2) as a bridging ligand via the Al(C6F5)3-induced rotation of the N-phosphinoyl moieties in (κ-C,O-L1)Ni(CO)2 (1a). (33) In the present study, we first attempted to extend this strategy, i.e., the Lewis acid-mediated reversible modulation of the local environment around the metal center, by exploring the reaction between 1a and B(C6F5)3 (B1), i.e., the boron analogue of Al(C6F5)3, with the expectation of forming the corresponding Ni/B bimetallic complex. However, upon dissociation of CO, we unexpectedly obtained (κ-C,O-L1)Ni(CO)(B1) (2a), which was isolated in 73% yield (Figure 2). In the 11B NMR spectrum of 2a in α,α,α-trifluorotoluene (TFT), the resonance of the B1 unit is observed at δB –10.3, which represents a significant upfield shift compared to that of free B1B 59.9) and suggests the formation of a four-coordinated boron species. For comparison, complexes bearing phosphine-boranes such as Ni/B-1 display downfield-shifted resonances (δB 15–30), (28−30) highlighting the difference between the present unsupported Z-type systems and the hitherto reported supported systems. We have also isolated (κ-C,O-L2)Ni(CO)(B1) (2b) and (κ-C,O-L3)Ni(CO)(B1) (2c) in 71% and 48% yield, respectively, via L-to-Z ligand substitution. Under an inert-gas atmosphere at −30 °C, crystalline samples of 2a–2c are stable for at least several months. These complexes are nearly insoluble in nonpolar hydrocarbons such as n-pentane and n-hexane, slightly soluble in aromatic hydrocarbons such as benzene and toluene, and fully soluble in TFT. These nickel–borane complexes slowly decompose into unidentified products once dissolved in TFT (e.g., 2a: t1/2 = 615 min at 23 °C).

Figure 2

Figure 2. Reaction between Ni-carbonyl complexes and triarylboranes. The isolated yield is shown with the NMR yield in parentheses; N.R.: no reaction; Dipp: 2,6-diisopropylphenyl; Mes: mesityl.

To investigate which properties are essential to the ability of triarylboranes to induce the Ni → BAr3 interaction, we treated 1a with B(2,3,5,6-F4C6H)3 (B2) and B(2,6-F2C6H3)3 (B3); the Lewis acidity of these boranes decreases in the order B1 > B2 > B3 (for details of the Lewis acidity of B1–B5 toward Et3P═O, see Table S1). (34) Interestingly, (κ-C,O-L1)Ni(CO)(B2) (3a) was obtained in 71% yield; however, no reaction proceeded when B3 was employed, representing a plausible threshold of Lewis acidity below which the reactions between 1a and boranes cannot proceed. This hypothesis is consistent with the fact that no reaction occurred using B(C6Cl5)3 (B4), which exhibits lower Lewis acidity but higher electrophilicity than B1. (35) We then explored the effect of the ortho-F atoms in the B-aryl groups by using B(3,5-(CF3)2C6H3)3 (B5). B5 has been reported to exhibit higher Lewis acidity than B1 due to the electron-withdrawing nature of the meta-CF3 groups as well as the reduced intramolecular repulsion between ortho-H atoms in the tetrahedral four-coordinated boron unit. (36) Although a change in the NMR spectra was virtually negligible when 1a and B5 were mixed under ambient conditions, nickel complex 6a was isolated in 86% yield when the reaction mixture was stirred under reduced pressure to promote the removal of CO (Figure 2). Notably, in stark contrast to 2a–c and 3a, no resonance was observed in the 11B NMR spectrum of 6a. We therefore concluded that the corresponding Ni → B5 interaction does not manifest in this complex. Based on the results of the single-crystal X-ray diffraction (SC-XRD) and density functional theory (DFT) calculation analyses (vide infra), we concluded that in 6a, interactions between the Ni–CO bond and the boron atom, as well as between the Ni atom and one of the ipso-C atoms in the aryl group occur. Thus, B5 does not act as a typical Z-type ligand under the applied reaction conditions, although the observed bonding situations are nearly identical to those found in Ni/B-2 and Ni/B-3 (Figure S20). We subsequently used 6a as a precursor for preparing Ni–borane complexes via the borane–exchange reaction. The quantitative formation of 2a was confirmed when 6a was treated with B1. These results demonstrate the role of the ortho-F atoms of the Z-type tris(perfluoroaryl)borane ligands in the formation of the Ni → Bn interactions in 2a–2c and 3a.
Complexes 2a–c, 3a, and 6a were unambiguously characterized using SC-XRD analysis. Selected geometric parameters are summarized and compared with the reported parameters in Ni/B-1, Ni/B-2, and Ni/B-3 in Table S2. The molecular structures of 2a and 6a are shown in Figure 3a,b, respectively, together with a topological analysis of the electron density based on the atoms-in-molecules (AIM) theory. The Ni–B bond length in 2a (2.245(1) Å) falls within the expected bond-length range for supported TM–borane complexes (7) (ca. 2.05–2.90 Å) but is clearly longer than those in Ni/B-1 (2.168(2) Å) and Ni/B-3 (2.1543(9) Å). Notably, the interatomic distance between the Ni and B atoms in 6a (2.267(4) Å) is nearly identical to that in 2a, despite the fact that the NMR analyses of 6a do not support the formation of a four-coordinated boron species (vide supra). Furthermore, in the case of 2a, a bond critical point (BCP) was confirmed between the Ni and B atoms (electron density ρ = 0.07 [e × rBohr–3]; Laplacian of electron density ∇2ρ = 0.00 [e × rBohr–5]) (Figure 3a, bottom) in AIM analysis. These results suggest that the Ni–B bond in 2a possesses mixed covalent/dative bond properties. On the other hand, a corresponding BCP is not observed in 6a, again suggesting the absence of a significant Ni → B interaction (Figure 3b, bottom). These results suggest that the existence of TM → B bonding interactions should not be discussed based on the interatomic distance between the TM and B atoms. A natural bond orbital (NBO) analysis also highlights characteristic interactions involved in 2a (Figure S23). Donor–acceptor interactions between the Ni and B atoms (E(2) = +26.9 kcal mol–1) and carbonyl C2 and B atoms (E(2) = +57.6 kcal mol–1) were confirmed in 2a based on second-order-perturbation-theory analysis, suggesting a three-center four-electron bond among the Ni, B, and C2 atoms. It should also be noted here that the Ni → BX3 interaction can be interpreted in terms of a reversed bonding situation compared to the TM ← BLnX3–n (n = 1–2) interaction in borylene-coordinated transition metal complexes, wherein TMs and borylenes play the roles of LAs and LBs, respectively. (37,38)

Figure 3

Figure 3. Molecular structures and electron densities for (a) 2a and (b) 6a. Each structure was obtained from the corresponding SC-XRD analysis. The quantum theory of AIM bond paths (white lines) and bond critical points (BCPs, green dots) are also shown with overlaid contour plots of ∇2ρ (e × rBohr–5) through the plane defined by the C2, Ni, and B atoms (∇2ρ < 0 shown in blue; ∇2ρ > 0 shown in red). The AIM methods used the SCF electron density calculated at the PBE0-D3BJ/Def2-TZVPD//M06L/Def2-SVPD(Ni,O,F), Def2-SVP(others) level. Electron densities (ρ in e × rBohr–3) at selected BCPs are given. Values of ∇2ρ are also given in parentheses. Selected bond lengths (Å) for 2a: Ni–C1 1.976(1), Ni–C2 1.697(1), Ni–B 2.245(1), Ni–O1 1.9916(9), C2–O2 1.154(1), Ni···F 2.7450(9); 6a: Ni–C1 1.946(2), Ni–C2 1.735(3), Ni–B 2.267(4), Ni–C3 2.110(2), Ni···C4 2.238(2), Ni–O1 2.320(2), C2–B 2.373(4), C2–O2 1.146(3).

Although the AIM analysis does not support the existence of a BCP between the C2 and B atoms in 2a, the unusual topology of ∇2ρ between the C2 and B atoms (Figure 3a, bottom) is consistent with a partial contribution of the C2 → B electron transfer. Similarly, the corresponding C2 → B interaction (E(2) = +49.0 kcal mol–1) in 6a was also confirmed by the NBO analysis, while the Ni → B interaction (E(2) = +14.5 kcal mol–1) was found to be less significant (Figure S25). Based on these results, we conclude that the coordination of the Ni–C2 bond to the B atom is a key feature for understanding the connection between the Ni(Lm)(CO) and Bn (n = 1 and 5, m = 1–3) units, as seen in 6a. When ortho-F atoms are involved in Bn, the Ni → B interaction tends to become predominant, affording complexes 2a–c with monodentate Z-type borane ligands. Indeed, the AIM analysis confirmed the participation of an NCI between the Ni center and one of the ortho-F atoms in the B1 unit (interatomic Ni···F distance: 2.7450(9) Å; determined by SC-XRD). The corresponding Ni···F NCIs were also observed in 2c (2.688(49) Å) and 3a (2.642(62) Å), whereas 2b, which bears N-mesityl-substituted L2, can not be expected to include the interaction between Ni and F atoms (the distance between Ni and the nearest F atoms is 2.948(1) Å). Instead, one of the ipso-carbon atoms (Cip) in the B-C6F5 groups of 2b closely approaches the Ni center at 2.529(1) Å, which is, however, still longer than those found in Ni/B-2 (2.018(3) Å) (29) and Ni/B-3 (2.0750(7) Å). (30) A BCP is found between these two atoms in 2a (ρ = 0.02 [e × rBohr–3]; ∇2ρ = 0.09 [e × rBohr–5]; δ(Ni|F) = 0.10), suggesting that this Ni···F interaction is significantly weaker than e.g., Ni–O1 (ρ = 0.06 [e × rBohr–3]; ∇2ρ = 0.32 [e × rBohr–5]; δ(Ni|O1) = 0.34) and Ni–C1 (ρ = 0.11 [e × rBohr–3]; ∇2ρ = 0.34 [e × rBohr–5]; δ(Ni|C1) = 0.71) interactions, wherein δ(X|Y) values show the average number of electrons shared at a BCP between X and Y atoms. In addition, all ortho-F atoms in 2a are observed at nearly identical shifts in the 19F NMR analysis (δF –137 ∼ −122 at −28 °C), suggesting a labile and exchangeable nature of the Ni···F NCIs in solution. Moreover, multiple NCIs are simultaneously observed between the L1 and B1 ligands over the Ni atoms in 2a. We think that these NCIs are not individually essential for the stability of the Ni–borane complexes but rather work cooperatively with other weak interactions.
The geometry around the Ni centers in 2a–c and 3a was evaluated based on the value of τMM = {360 – (α + β)}/141 × β/α, where α and β are the largest and second-largest L–M–L angles obtained from their SC-XRD analyses]. (39) A τM value approaching 0 indicates that M adopts an ideal square-planar geometry, while a value approaching 1 corresponds to an ideal tetrahedral geometry. The Ni centers in 2a–c and 3a were found to adopt a distorted square-planar geometry with τNi values of 0.10–0.16, as has often been found for low-spin Ni(II) complexes; however, we eventually concluded that the electronic state of the Ni center in 2a would be close to Ni(0) (vide infra) (Figure 6). For the geometry around the boron center, as expected, distorted tetrahedral structures (τB = 0.7–0.8) were confirmed in 2a–c and 3a, while the B atom in 6a retains a rather planar geometry (sum of C–B–C angles = 353.8°, see also results of the tetrahedral characters [THC(B)] in Table S2).
To further confirm the prerequisite factors for the monodentate Z-type coordination of tris(perfluoroaryl)boranes, we carried out control experiments using B1 and nickel carbonyl complexes 1d–1f, which bear L4–L6 (Figure 4a); however, we did not observe the formation of the corresponding Ni–B1 complex under the applied conditions. In fact, the reaction between N-phosphanyl-substituted 1d and B1 resulted in the formation of (κ-C-L4)Ni(CO)3 in 19% yield through B1-induced disproportionation of the CO ligands in 1d, while other nickel carbonyl species were not identified by 31P and 19F NMR analyses. Thus, the N-phosphinoyl oxygen atom in L1 plays a role in stabilizing 2a as an isolable species. No reaction took place when 1e and 1f, which bear L5 and L6, respectively, were employed. We then turned our attention to 1g, which bears 1,3-di-tert-butylimidazol-2-ylidene L7, as it can generate an FLP species with B1 (Figure 4b). (40,41) We confirmed the generation of an equilibrium mixture including 1g (56%) and 7g (41%) at −30 °C after a period of 5 h. The characterization of 7g was accomplished based on NMR and SC-XRD analyses, which clearly confirmed the presence of the Ni(μ-η12-CO)B unit. The resonance for the corresponding four-coordinated boron moiety was observed at δB –14.2 in the 11B NMR spectrum, a shift that is comparable to that reported for the four-coordinated boron complex [Cp2Ti(THF-d8)(μ-η12-CO)B1] (42) (Cp = cyclopentadienyl) (δB –11). Complex 7g gradually decomposed even at −10 °C and eventually gave an unidentifiable mixture, probably via the dissociation of the end-on CO ligand. A crystal suitable for the SC-XRD analysis was prepared from this resultant mixture by recrystallization at −30 °C, and we confirmed the presence of binuclear nickel carbonyl complex 8g (Figure 4b). (43−46) Inspired by these results, we explored a possible mechanistic scenario for the formation of 7g from 1g and B1 based on DFT calculations at the PBE0-D3BJ/Def2-TZVPD//M06L/Def2-SVPD(Ni,F,O), Def2-SVP(others) level. The relative Gibbs free energies with respect to [1g + B1] (0.0 kcal mol–1) for selected molecules are shown in Figure 4b (for details, see Figure S29). We found that square-planar Ni complex 2g (+8.0 kcal mol–1), which bears B1 as a monodentate Z-type ligand, is formed via TS1′ (+11.9 kcal mol–1). While the Ni–B bond (2.43 Å) is longer than those in 2a–2c (ca. 2.2 Å), the BCP found between these atoms by the AIM analysis supports the presence of the Ni → B interaction in 2g (Figure S21). Moreover, as was seen for 2a, the participation of multiple NCIs between the L7 and B1 units in 2g was confirmed. Subsequently, the Ni-to-C2 migration of B1 proceeds via TS2′ (+23.9 kcal mol–1) to yield 7g (+1.6 kcal mol–1). These results imply that the use of multidentate carbene ligands (e.g., L1–L3) that can generate FLP species with tris(perfluoroaryl)boranes represents a strategy for generating transition metal complexes that bear monodentate Z-type borane ligands.

Figure 4

Figure 4. Effect of ligands. (a) Reaction with 1d–1f. (b) Reaction with 1g. The conversion of 1g through the transformation of 7g is shown in square brackets. The gas-phase optimized structure of 2g and the SC-XRD structure of 7g (except hydrogen atoms; thermal ellipsoids at 30% probability) are also shown. Selected bond lengths (Å) for 2g: Ni–C1 2.00, Ni–C2 1.82, Ni–B 2.43, C2–O 1.15; 7g: Ni–C1 1.976(4), Ni–O 1.966(3), Ni–C2 1.843(4), C2–O 1.207(5), C2–B 1.639(5).

Based on the results obtained up to this point, we considered possible reaction mechanisms to afford 2a from 1a and B1. Figure 5 depicts two selected paths with key intermediates and transition states (for further details, see Figure S28). Following the generation of association complex Int1-v1 (+5.1 kcal mol–1) from 1a and B1, the formation of the C2–B bond proceeds via TSadd1 (+13.6 kcal mol–1) to afford Ni(μ-η1-CO)B species Int2-v1 (+7.1 kcal mol–1; C2–B 1.80 Å) (blue colored path). Subsequently, another CO ligand smoothly dissociates to yield Int3-v1 (+12.1 kcal mol–1) without a significant energy barrier, which eventually forms Int4-v1 (+5.1 kcal mol–1) via the complete removal of CO. Coordination of the Cip atom in the C6F5 unit stabilizes Int4-v1, and the B atom effectively approaches the Ni center (Ni···B 2.31 Å). The C2-to-Ni migration of the B atom subsequently occurs via TSmig1 (+7.0 kcal mol–1), resulting in the construction of the Ni–B bond (2.22 Å) together with the elongation of the C2···B distance to 2.41 Å from 1.78 Å in Int4-v1. This mechanistic scenario predicts the facile formation of 2a after mixing 1a and B1, as the reaction rate would be predominantly determined by the formation of the C2–B bond via TSadd1, i.e., with an energy barrier of only +13.6 kcal mol–1. In contrast, for B3 and B4, the corresponding paths to form the C2–B bond and dissociate another CO ligand should become energetically unfavorable due to their decreased Lewis acidity, which would explain why no reaction was observed with 1a (Figure 2). On the other hand, we also found that the formation of Ni(μ-η12-CO)B species Int2-v4 (−1.6 kcal mol–1) from Int1-v1 via TSism1 (+14.4 kcal mol–1) can compete with the above pathway via TSadd1 (gray colored path). The geometric parameters around the Ni atom in Int2-v4 (e.g., Ni–C2 1.85 Å; Ni–O2 2.01 Å; C2–B 1.62 Å) are nearly identical to those in 7g (Figure 4b), which bears L4, although the phosphinoyl O1 weakly coordinates to Ni in Int2-v4. This weak coordination of the phosphinoyl moiety should prevent the formation of a Ni–borane complex that bears two CO ligands and the carbene in L1, as seen in the interconversion between 2g and 7g. It would be possible to yield 2a from Int2-v4 via the regeneration of Int1-v1 and then TSmig1 as a result of the change in the coordination mode of CO from a μ-η12 to a μ-η1 fashion; the overall energy barrier in this case is +16.0 kcal mol–1. Conversely, the path directly connecting Int2-v4 and 2a is unfavorable due to the higher energy barrier to overcome TSmig2.

Figure 5

Figure 5. Plausible mechanisms for the formation of 2a. Relative Gibbs free energies (kcal mol–1) with respect to [1a + B1] (0.0 kcal mol–1) are shown, calculated at the PBE0-D3BJ/Def2-TZVPD//M06L/Def2-SVPD(Ni,F,O), Def2-SVP(others) level. Parts of the molecular structures for the selected compounds are also shown with selected bond lengths (Å).

As mentioned earlier, the Ni center in 2a–c and 3a adopts a distorted square-planar geometry, which represents a fingerprint of the Ni(II) configuration in classical ligand-field theory. However, several transition metal complexes that bear Z-type ligands have been found to adopt unusual electronic configurations different from those expected based on their geometry. (47−51) Thus, we employed Ni K- and L2,3-edge X-ray absorption spectroscopy (XAS) to gain insight into the electronic configuration of 2a. Given that transition metal K-edge XAS is sensitive to the ligand-field geometry rather than the electronic state, we synthesized (κ-C,O-L1)Ni(C6F5)2 (9), a low-spin square-planar nickel complex with the formal oxidation state +II, as a reference. The Ni K-edge X-ray absorption near edge structure (XANES) data are shown in Figure 6a. The absorption edge of 2a appeared at 8333 eV with the pre-edge peak at 8327 eV; these values were almost identical to those of 1a. In addition, the absorption edge of 9 appeared at 8336 eV, i.e., 3 eV higher than that of 2a. Considering the identical ligand-field geometries in 2a and 9, this edge-shift suggests that 2a adopts the formal electric state Ni(0). The Ni L2,3-edge XAS analysis afforded deeper insight, as the hole number in Ni 3d orbitals is proportional to the sum of the Ni L3- and Ni L2-edge peak area based on the sum rule. (52,53) To this end, Ni d-electron counts are inversely proportional to the sum of the Ni L3- and Ni L2-edge peak area, which allows transition metal L2,3-edge XAS to be used as a direct probe for d-electron occupancy. The Ni L2,3-edge XAS experiments were conducted using the total electron yield (TEY) method under vacuum conditions, and the obtained XAS data of the (κ-C,O-L1)Ni complexes 1a, 2a, 9, and 10 are shown in Figure 6b, together with those of Ni(COD)(DQ) (COD: 1,5-cyclooctadiene, DQ: duroquinone). (54) For the formal Ni(II) complexes 9 (low-spin) and 10 (high-spin), sharp and intense L3-edge peaks were observed at 856.6 and 854.8 eV, respectively. In contrast, Ni(COD)(DQ) exhibited a broadened first peak at 856.2 eV. A prominent second peak was detected at higher energy, which was ascribed to the Ni 2p → ligand transitions. (55) In the cases of 1a and 2a, the XAS signals are split into several peaks with low intensity, which was attributed to the influence of the electron-accepting ligands, such as carbonyl and borane. The Ni 3d electron counts were then calculated from the sum of the normalized peak area using the linear function established from the XAS data of the d-count calibrants [Et4N]2[NiCl4] and [Na(benzo-15-crown-5)]2[Ni(mnt)2] (mnt: 1,2-dicyano-ethene-1,2-dithiolate) (Figure 6c and Table S6). (56,57) The Ni 3d count of 2a (9.1) is comparable to those of the other formal Ni(0) complexes, such as 1a (9.0) and Ni(COD)(DQ) (9.0). These results are also consistent with the estimated electron configuration of (3d)9.10(4s)0.32 calculated by the NBO7 program.

Figure 6

Figure 6. Experimental and theoretical evaluation of the electronic state of the Ni complexes examined in this study. (a) Ni K-edge XAS spectra. (b) Ni L2,3-edge XAS spectra. (c) Sum-rule analysis. (d) A simplified Frontier-molecular-orbital energy diagram of 2a, focusing on the interaction between the Ni and B centers, calculated at PBE0-D3BJ/Def2-TZVPD level. (e) Kohn–Sham HOMO and LUMO+1 in 2a and their orbital composition.

Finally, we disclose the role of tris(perfluoroaryl)boranes that contain ortho-F atoms, such as B1 and B2, in constructing the unusual square-planar geometry at the Ni(0) center with the occupied 3d electron configuration (Figure 6d). Based on ligand-field theory, Ni(0) d10 complexes usually adopt a tetrahedral geometry; this tendency is clearly observed in 6a bearing ortho-protonated B5, which coordinates to the Ni center via the Cip (labeled as C3 in Figure 3b). On the other hand, in the case of the B1/B2 ligands, the repulsive effects generated between the Ni and ortho-F atoms would prohibit such coordination of the Cip atoms, and lead to orbital rehybridization to stabilize the high-lying occupied Ni 3dx2-y2 orbital with the assistance of σ-accepting B1/B2 ligands. As a result, 2a–c and 3a adopt sterically acceptable but electronically unfavorable square-planar geometries, i.e., the square-planar Ni d9.1 state, where the electron density on the Ni center is modulated through interactions with the supporting ligands. In fact, DFT calculations showed that the HOMO and LUMO+1 in 2a are predominantly located on the Ni–B bonds with significant contributions of the Ni 3d and B 2p atomic orbitals (HOMO: Ni 29.2% and B 15.5%; LUMO+1: Ni 15.1%, B 8.7%) (Figure 6e), supporting the conclusion that the Ni and B atoms form a typical Lewis base → acid interaction. Furthermore, an energy decomposition analysis (EDA) indicated a significant contribution of the electrostatic interaction (ΔEelstat) in the stabilization of the square-planar Ni centers bearing B1 and B2 (Table S4).

Conclusions

Click to copy section linkSection link copied!

In summary, we have isolated and fully characterized square-planar nickel carbonyl (Ni–CO) complexes that bear tris(perfluoroaryl)boranes as monodentate σ-accepting (Z-type) ligands. In these complexes, the Ni center exhibits a d9.1 configuration with a high-lying occupied 3dx2y2 orbital, which is effectively stabilized by the σ-accepting ability of tris(perfluoroaryl)boranes with ortho-F atoms. The Ni–B bonds in these complexes exhibit mixed covalent/dative character. In contrast, B(3,5-(CF3)2C6H3)3 afforded a tetrahedral nickel complex via Ni–CO bond coordination to the boron atom, highlighting the critical role of the ortho-F atoms in constructing square-planar nickel–borane complexes. Under the explored experimental and theoretical conditions, such nickel–borane complexes seem to be generated through the combination of sterically encumbered N-heterocyclic carbenes (L-type ligands) and tris(perfluoroaryl)boranes (Z-type ligands). In addition, we observed the participation of multiple noncovalent interactions over the L–Ni–Z units, which should support the deformed tetrahedral four-coordinated boron units. This work thus demonstrates a strategy, namely, the use of frustrated L/Z-ligand pairs, for the formation of a hitherto elusive class of low-valent transition metal complexes.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c15892.

  • Full details pertaining to the experimental methods, identification of the compounds, and DFT calculations (PDF)

  • AIM analysis (XLSX)

  • DFT coordinate (XLSX)

Accession Codes

Deposition Numbers 23847522384760 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via the joint Cambridge Crystallographic Data Centre (CCDC) and Fachinformationszentrum Karlsruhe Access Structures service.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Yuta Uetake - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, JapanInnovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0002-4742-8085 Email: [email protected]
    • Yoichi Hoshimoto - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, JapanCenter for Future Innovation (CFi), Division of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0003-0882-6109 Email: [email protected]
  • Authors
    • Yutaka Mondori - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    • Yasuhiro Yamauchi - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    • Takahiro Kawakita - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, Japan
    • Sensuke Ogoshi - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0003-4188-8555
    • Yasuo Takeichi - Department of Applied Physics, Graduate School of Engineering, Osaka University, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0003-3334-0274
    • Hidehiro Sakurai - Department of Applied Chemistry, Faculty of Engineering, Osaka University, Suita, Osaka 565-0871, JapanInnovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0001-5783-4151
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We are grateful to Daisuke Hashizume (RIKEN) for insightful discussions on the AIM analysis. Ni K-edge XAS measurements were performed at the BL14B2 beamline of SPring-8 under the approval of the Japan Synchrotron Radiation Research Institute (proposal numbers 2020A1871, 2021A1630, 2022A1767, and 2022A1784). Ni L2,3-edge XAS measurements were performed at the BL19B beamline of KEK under the approval of the Photon Factory Program Advisory Committee (proposal numbers 2022P013 and 2024G031), and at the BL4B beamline of the UVSOR Synchrotron Facility with the approval of the Institute for Molecular Science (proposal number 21-697). Parts of the theoretical calculations were performed using resources from the Research Center for Computational Science, Okazaki, Japan (24-IMS-C089). This project was supported by Grants-in-Aid for Transformative Research Area (A) Digitalization-driven Transformative Organic Synthesis (22H05363 to Y.H.); Green Catalysis Science for Renovating Transformation of Carbon-Based Resources (24H01851 to Y.U.); JSPS KAKENHI grant Scientific Research (C) (22K05095 to Y.U.); the JST FOREST Program (JPMJFR2222 to Y.H.); a JSPS Research Fellowship (to Y.Y.); and JST SPRING (to Y.M.).

References

Click to copy section linkSection link copied!

This article references 57 other publications.

  1. 1
    Jensen, W. B. The Lewis Acid-Base Definitions: A Status Report. Chem. Rev. 1978, 78, 131,  DOI: 10.1021/cr60311a002
  2. 2
    Crawley, M. L.; Trost, B. M. Applications of Transition Metal Catalysis in Drug Discovery and Development: An Industrial Perspective; John Wiley & Sons, 2012; . DOI: 10.1002/9781118309872 .
  3. 3
    Striegler, S.; Rieger, B.; Kacker, S.; Baugh, L. S. Late Transition Metal Polymerization Catalysis; Wiley VCH, 2003.
  4. 4
    Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books, 2010.
  5. 5
    Braunschweig, H.; Dewhurst, R. D.; Schneider, A. Electron-Precise Coordination Modes of Boron-Centered Ligands. Chem. Rev. 2010, 110, 39243957,  DOI: 10.1021/cr900333n
  6. 6
    Amgoune, A.; Bourissou, D. σ-Acceptor, Z-type ligands for transition Metals. Chem. Commun. 2011, 47, 859871,  DOI: 10.1039/C0CC04109B
  7. 7
    Braunschweig, H.; Dewhurst, R. D. Transition metals as Lewis bases: “Z-type” boron ligands and metal-to-boron dative bonding. Dalton Trans. 2011, 40, 549558,  DOI: 10.1039/C0DT01181A
  8. 8
    Bauer, J.; Braunschweig, H.; Dewhurst, R. D. Metal-Only Lewis Pairs with Transition Metal Lewis Bases. Chem. Rev. 2012, 112, 43294346,  DOI: 10.1021/cr3000048
  9. 9
    Parkin, G. Impact of the coordination of multiple Lewis acid functions on the electronic structure and vn configuration of a metal center. Dalton Trans. 2022, 51, 411427,  DOI: 10.1039/D1DT02921E
  10. 10
    Komuro, T.; Nakajima, Y.; Takaya, J.; Hashimoto, H. Recent progress in transition metal complexes supported by multidentate ligands featuring group 13 and 14 elements as coordinating atoms. Coord. Chem. Rev. 2022, 473, 214837214864,  DOI: 10.1016/j.ccr.2022.214837
  11. 11
    Green, M. L. H. A new approach to the formal classification of covalent compounds of the elements. J. Organomet. Chem. 1995, 500, 127148,  DOI: 10.1016/0022-328X(95)00508-N
  12. 12
    Davydova, E. I.; Sevastianova, T. N.; Timoshkin, A. Y. Molecular complexes of group 13 element trihalides, pentafluorophenyl derivatives and Lewis superacids. Coord. Chem. Rev. 2015, 297, 91126,  DOI: 10.1016/j.ccr.2015.02.019
  13. 13
    Cowley, A. H. From group 13–group 13 donor–acceptor bonds to triple-decker cations. Chem. Commun. 2004, 23692375,  DOI: 10.1039/B409497M
  14. 14
    Burlitch, J. M.; Leonowicz, M. E.; Petersen, R. B.; Hughes, R. E. Coordination of Metal Carbonyl Anions to Triphenylaluminum, -gallium, and -indium and the Crystal Structure of Tetraethylammonium Triphenyl((η5-Cyclopentadienyl)dicarbonyliron)aluminate(Fe–Al). Inorg. Chem. 1979, 18, 10971105,  DOI: 10.1021/ic50194a040
  15. 15
    Kameo, H.; Nakazawa, H. Saturated Heavier Group 14 Compounds as σ-Electron-Acceptor (Z-type) Ligands. Chem. Rec. 2017, 17, 268286,  DOI: 10.1002/tcr.201600061
  16. 16
    For examples including Pt-complexes, see:Bauer, J.; Braunschweig, H.; Dewhurst, R. D.; Radacki, K. Reactivity of Lewis Basic Platinum Complexes Towards Fluoroboranes. Chem.─Eur. J. 2013, 19, 87978805,  DOI: 10.1002/chem.201301056
  17. 17
    For examples including Fe-complexes, see:Burlitch, J. M.; Burk, J. H.; Leonowicz, M. E.; Hughes, R. E. Migration of triphenylboron from iron to a geminal η5-cylopentadienyl ligand. Inorg. Chem. 1979, 18, 17021709,  DOI: 10.1021/ic50196a061
  18. 18
    For examples including Rh-complexes, see:Alférez, M. G.; Moreno, J. J.; Gaona, M. A.; Maya, C.; Campos, J. Ligand Postsynthetic Functionalization with Fluorinated Boranes and Implications in Hydrogenation Catalysis. ACS Catal. 2023, 13, 1605516066,  DOI: 10.1021/acscatal.3c02764
  19. 19
    Hashimoto, T.; Asada, T.; Ogoshi, S.; Hoshimoto, Y. Main group catalysis for H2 purification based on liquid organic hydrogen carriers. Sci. Adv. 2022, 8, eade0189  DOI: 10.1126/sciadv.ade0189
  20. 20
    Sakuraba, M.; Morishita, T.; Hashimoto, T.; Ogoshi, S.; Hoshimoto, Y. Remote Back Strain: A Strategy for Modulating the Reactivity of Triarylboranes. Synlett 2023, 34, 21872192,  DOI: 10.1055/a-2110-5359
  21. 21
    Hisata, Y.; Washio, T.; Takizawa, S.; Ogoshi, S.; Hoshimoto, Y. In-silico-assisted derivatization of triarylboranes for the ccatalytic reductive functionalization of aniline-derived amino acids and peptides with H2. Nat. Commun. 2024, 15, 3708,  DOI: 10.1038/s41467-024-47984-0
  22. 22
    Goedecke, C.; Hillebrecht, P.; Uhlemann, T.; Haunschild, R.; Frenking, G. The Dewar-Chatt-Duncanson model reversed: Bonding analysis of group-10 complexes [(PMe3)2M–EX3] (M = Ni, Pd, Pt; E = B, Al, Ga, In, Tl; X = H, F, Cl, Br, I). Can. J. Chem. 2009, 87, 14701479,  DOI: 10.1139/V09-099
  23. 23
    Erdmann, P.; Greb, L. What Distinguishes the Strength and the Effect of a Lewis Acid: Analysis of the Gutmann–Beckett Method. Angew. Chem., Int. Ed. 2022, 61, e202114550  DOI: 10.1002/anie.202114550
  24. 24
    For examples including Ru-complexes, see:Hill, A. F.; Owen, G. R.; White, A. J. P.; Williams, D. J. The Sting of the Scorpion: A Metallaboratrane. Angew. Chem., Int. Ed. 1999, 38, 27592761,  DOI: 10.1002/(SICI)1521-3773(19990917)38:18<2759::AID-ANIE2759>3.0.CO;2-P
  25. 25
    Hill, A. F. An Unambiguous Electron-Counting Notation for Metallaboratranes. Organometallics 2006, 25, 47414743,  DOI: 10.1021/om0602512
  26. 26
    For examples including Rh-complexes, see:Crossley, I. R.; Hill, A. F.; Willis, A. C. Metallaboratranes: Tris(methimazolyl)borane Complexes of Rhodium(I). Organometallics 2006, 25, 289299,  DOI: 10.1021/om050772+
  27. 27
    Parkin, G. A. Simple Description of the Bonding in Transition-Metal Borane Complexes. Organometallics 2006, 25, 47444747,  DOI: 10.1021/om060580u
  28. 28
    For examples including Ni-, Cu-, Pd-, Ag-, Pt- and Au-complexes, see:Sircoglou, M.; Bontemps, S.; Bouhadir, G.; Saffon, N.; Miqueu, K.; Gu, W.; Mercy, M.; Chen, C.-H.; Foxman, B. M.; Maron, L.; Ozerov, O. V.; Bourissou, D. Group 10 and 11 Metal Boratranes (Ni, Pd, Pt, CuCl, AgCl, AuCl, and Au+) Derived from a Triphosphine–Borane. J. Am. Chem. Soc. 2008, 130, 1672916738,  DOI: 10.1021/ja8070072
  29. 29
    For examples including Ni- and Pd-complexes, see:Emslie, D. J. H.; Harrington, L. E.; Jenkins, H. A.; Robertson, C. M.; Britten, J. F. Group 10 Transition-Metal Complexes of an Ambiphilic PSB-Ligand: Investigations into η3(BCC)-Triarylborane Coordination. Organometallics 2008, 27, 53175325,  DOI: 10.1021/om800670e
  30. 30
    For examples including Ni-complexes, see:Harman, W. H.; Peters, J. C. Reversible H2 Addition across a Nickel-Borane Unit as a Promising Strategy for Catalysis. J. Am. Chem. Soc. 2012, 134, 50805082,  DOI: 10.1021/ja211419t
  31. 31
    Stephan, D. W.; Erker, G. Frustrated Lewis Pair Chemistry: Development and Perspectives. Angew. Chem., Int. Ed. 2015, 54, 64006441,  DOI: 10.1002/anie.201409800
  32. 32
    Jupp, A. R.; Stephan, D. W. New Directions for Frustrated Lewis Pair Chemistry. Trends Chem. 2019, 1, 3548,  DOI: 10.1016/j.trechm.2019.01.006
  33. 33
    Yamauchi, Y.; Mondori, Y.; Uetake, Y.; Takeichi, Y.; Kawakita, T.; Sakurai, H.; Ogoshi, S.; Hoshimoto, Y. Reversible Modulation of the Electronic and Spatial Environment around Ni(0) Centers Bearing Multifunctional Carbene Ligands with Triarylaluminum. J. Am. Chem. Soc. 2023, 145, 1693816947,  DOI: 10.1021/jacs.3c06267
  34. 34
    Carden, J. L.; Dasgupta, A.; Melen, R. L. Halogenated triarylboranes: synthesis, properties and applications in catalysis. Chem. Soc. Rev. 2020, 49, 17061725,  DOI: 10.1039/C9CS00769E
  35. 35
    Ashley, A. E.; Herrington, T. J.; Wildgoose, G. G.; Zaher, H.; Thompson, A. L.; Rees, N. H.; Krämer, T.; O’hare, D. Separating Electrophilicity and Lewis Acidity: The Synthesis, Characterization, and Electrochemistry of the Electron Deficient Tris(aryl)boranes B(C6F5)3-n(C6Cl5)n (n = 1–3). J. Am. Chem. Soc. 2011, 133, 1472714740,  DOI: 10.1021/ja205037t
  36. 36
    Sakuraba, M.; Hoshimoto, Y. Recent Trends in Triarylborane Chemistry: Diversification of Structures and Reactivity via meta-Substitution of the Aryl Groups. Synthesis 2024, 56, 3421,  DOI: 10.1055/s-0043-1775394
  37. 37
    Braunschweig, H.; Dewhurst, R. D.; Gessner, V. H. Transition metal borylene complexes. Chem. Soc. Rev. 2013, 42, 31973208,  DOI: 10.1039/c3cs35510a
  38. 38
    Kong, L.; Lu, W.; Yongxin, Y.; Ganguly, R.; Kinjo, R. Formation of Boron–Main-Group Element Bonds by Reactions with a Tricoordinate Organoboron L2PhB: (L = Oxazol-2-ylidene). Inorg. Chem. 2017, 56, 55865593,  DOI: 10.1021/acs.inorgchem.6b02993
  39. 39
    Reineke, M. H.; Sampson, M. D.; Rheingold, A. L.; Kubiak, C. P. Synthesis and Structural Studies of Nickel(0) Tetracarbene Complexes with the Introduction of a New Four-Coordinate Geometric Index, τδ. Inorg. Chem. 2015, 54, 32113217,  DOI: 10.1021/ic502792q
  40. 40
    Holschumacher, D.; Bannenberg, T.; Hrib, C. G.; Jones, P. G.; Tamm, M. Heterolytic Dihydrogen Activation by a Frustrated Carbene-Borane Lewis Pair. Angew. Chem., Int. Ed. 2008, 47, 74287432,  DOI: 10.1002/anie.200802705
  41. 41
    Chase, P. A.; Stephan, D. W. Hydrogen and Amine Activation by a Frustrated Lewis Pair of a Bulky N-Heterocyclic Carbene and B(C6F5)3. Angew. Chem., Int. Ed. 2008, 47, 74337437,  DOI: 10.1002/anie.200802596
  42. 42
    Choukroun, R.; Lorber, C.; Lepetit, C.; Donnadieu, B. Reactivity of [Cp2Ti(CO)2] and B(C6F5)3: Formation of the Acylborane Complexes [Cp2Ti(CO)(η2-OCB(C6F5)3)] and [Cp2Ti(THF)(η2-OCB(C6F5)3)]. Organometallics 2003, 22, 19951997,  DOI: 10.1021/om030185t
  43. 43
    Lyngdoh, R. H. D.; Schaefer, H. F.; King, R. B. Metal-Metal (MM) Bond Distances and Bond Orders in Binuclear Metal Complexes of the First Row Transition Metals Titanium Through Zinc. Chem. Rev. 2018, 118, 1162611706,  DOI: 10.1021/acs.chemrev.8b00297
  44. 44
    Shriver, D. F. BASCITY AND REACTIVITY OF METAL CARBONYLS. J. Organomet. Chem. 1975, 94, 259271,  DOI: 10.1016/S0022-328X(00)88721-5
  45. 45
    Gong, J.-K.; Kubiak, C. P. A New Route to the Ni(0) ‘Cradle’ Complex Ni2(μ-CO)(CO)2(PPh2CH2PPh2)2: μ-CO Ligand and Metal-centered Reactivity. Inorg. Chim. Acta 1989, 162, 1921,  DOI: 10.1016/S0020-1693(00)83112-6
  46. 46
    Bernhardt, E.; Finze, M.; Willner, H.; Lehmann, C. W.; Aubke, F. Salts of the Cobalt(I) Complexes [Co(CO)5]+ and [Co(CO)2(NO)2]+ and the Lewis Acid-Base Adduct [Co2(CO)7CO-B(CF3)3]. Chem.─Eur. J. 2006, 12, 82768283,  DOI: 10.1002/chem.200600210
  47. 47
    Wade, C. R.; Lin, T.-P.; Nelson, R. C.; Mader, E. A.; Miller, J. T.; Gabbaï, F. P. Synthesis, Structure, and Properties of a T-shaped 14-Electron Stiboranyl-Gold Complex. J. Am. Chem. Soc. 2011, 133, 89488955,  DOI: 10.1021/ja201092g
  48. 48
    Ansmann, N.; Münch, J.; Schorpp, M.; Greb, L. Neutral and Anionic Square Planar Palladium(0) Complexes Stabilized by a Silicon Z-Type Ligand. Angew. Chem., Int. Ed. 2023, 62, e202313636  DOI: 10.1002/anie.202313636
  49. 49
    For examples including Pt-complexes, see:Kameo, H.; Tanaka, Y.; Shimoyama, Y.; Izumi, D.; Matsuzaka, H.; Nakajima, Y.; Lavedan, P.; Le Gac, A.; Bourissou, D. Square-Planar Anionic Pt0 Complexes. Angew. Chem., Int. Ed. 2023, 62, e202301509  DOI: 10.1002/anie.202301509
  50. 50
    Wächtler, E.; Gericke, R.; Block, T.; Pöttgen, R.; Wagler, J. Trivalent Antimony as L-, X-, and Z-Type Ligand: The Full Set of Possible Coordination Modes in Pt-Sb Bonds. Inorg. Chem. 2020, 59, 1554115552,  DOI: 10.1021/acs.inorgchem.0c02615
  51. 51
    For examples including Au-complexes, see:Sircoglou, M.; Bontemps, S.; Mercy, M.; Saffon, N.; Takahashi, M.; Bouhadir, G.; Maron, L.; Bourissou, D. Transition-Metal Complexes Featuring Z-Type Ligands: Agreement or Discrepancy between Geometry and dn Configuration?. Angew. Chem., Int. Ed. 2007, 46, 85838586,  DOI: 10.1002/anie.200703518
  52. 52
    Dimucci, I. M.; Lukens, J. T.; Chatterjee, S.; Carsch, K. M.; Titus, C. J.; Lee, S. J.; Nordlund, D.; Betley, T. A.; MacMillan, S. N.; Lancaster, K. M. The Myth of d8 Copper(III). J. Am. Chem. Soc. 2019, 141, 1850818520,  DOI: 10.1021/jacs.9b09016
  53. 53
    DiMucci, I. M.; Titus, C. J.; Nordlund, D.; Bour, J. R.; Chong, E.; Grigas, D. P.; Hu, C. H.; Kosobokov, M. D.; Martin, C. D.; Mirica, L. M.; Nebra, N.; Vicic, D. A.; Yorks, L. L.; Yruegas, S.; MacMillan, S. N.; Shearer, J.; Lancaster, K. M. Scrutinizing Formally NiIV centers through the lenses of core spectroscopy, molecular orbital theory, and valence bond theory. Chem. Sci. 2023, 14, 69156929,  DOI: 10.1039/D3SC02001K
  54. 54
    Tran, V. T.; Li, Z. Q.; Apolinar, O.; Derosa, J.; Joannou, M. V.; Wisniewski, S. R.; Eastgate, M. D.; Engle, K. M. Ni(COD)(DQ): An Air-Stable 18-Electron Nickel(0)-Olefin Precatalyst. Angew. Chem., Int. Ed. 2020, 59, 74097413,  DOI: 10.1002/anie.202000124
  55. 55
    Hatsui, T.; Kosugi, N. Metal-to-ligand charge transfer in polarized metal L-edge X-ray absorption of Ni and Cu Complexes. J. Electron Spectrosc. Relat. Phenom. 2004, 136, 6775,  DOI: 10.1016/j.elspec.2004.02.133
  56. 56
    Solomon, E. I.; Hedman, B.; Hodgson, K. O.; Dey, A.; Szilagyi, R. K. Ligand K-Edge X-ray aabsorption spectroscopy: covalency of ligand-metal bonds. Coord. Chem. Rev. 2005, 249, 97129,  DOI: 10.1016/j.ccr.2004.03.020
  57. 57
    Sarangi, R.; George, S. D. B.; Rudd, D. J.; Szilagyi, R. K.; Ribas, X.; Rovira, C.; Almeida, M.; Hodgson, K. O.; Hedman, B.; Solomon, E. I. Sulfur K-Edge X-ray Absorption Spectroscopy as a Probe of Ligand-Metal Bond Covalency: Metal vs Ligand Oxidation in Copper and Nickel Dithiolene Complexes. J. Am. Chem. Soc. 2007, 129, 23162326,  DOI: 10.1021/ja0665949

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2025, 147, 10, 8326–8335
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.4c15892
Published February 28, 2025

Copyright © 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

4816

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Concept of electron acceptors and donors. (a) A simplified representation of complexation between Lewis acids (LAs) and Lewis bases (LBs) and the related orbitals. (b) Coordination of transition metals (TM) to trivalent group-13 species. (c) Pioneering examples of Ni complexes that bear LnZ-type borane ligands. (d) A strategy for monodentate Z-type borane ligands based on frustrated L/Z-ligand pairs (this work).

    Figure 2

    Figure 2. Reaction between Ni-carbonyl complexes and triarylboranes. The isolated yield is shown with the NMR yield in parentheses; N.R.: no reaction; Dipp: 2,6-diisopropylphenyl; Mes: mesityl.

    Figure 3

    Figure 3. Molecular structures and electron densities for (a) 2a and (b) 6a. Each structure was obtained from the corresponding SC-XRD analysis. The quantum theory of AIM bond paths (white lines) and bond critical points (BCPs, green dots) are also shown with overlaid contour plots of ∇2ρ (e × rBohr–5) through the plane defined by the C2, Ni, and B atoms (∇2ρ < 0 shown in blue; ∇2ρ > 0 shown in red). The AIM methods used the SCF electron density calculated at the PBE0-D3BJ/Def2-TZVPD//M06L/Def2-SVPD(Ni,O,F), Def2-SVP(others) level. Electron densities (ρ in e × rBohr–3) at selected BCPs are given. Values of ∇2ρ are also given in parentheses. Selected bond lengths (Å) for 2a: Ni–C1 1.976(1), Ni–C2 1.697(1), Ni–B 2.245(1), Ni–O1 1.9916(9), C2–O2 1.154(1), Ni···F 2.7450(9); 6a: Ni–C1 1.946(2), Ni–C2 1.735(3), Ni–B 2.267(4), Ni–C3 2.110(2), Ni···C4 2.238(2), Ni–O1 2.320(2), C2–B 2.373(4), C2–O2 1.146(3).

    Figure 4

    Figure 4. Effect of ligands. (a) Reaction with 1d–1f. (b) Reaction with 1g. The conversion of 1g through the transformation of 7g is shown in square brackets. The gas-phase optimized structure of 2g and the SC-XRD structure of 7g (except hydrogen atoms; thermal ellipsoids at 30% probability) are also shown. Selected bond lengths (Å) for 2g: Ni–C1 2.00, Ni–C2 1.82, Ni–B 2.43, C2–O 1.15; 7g: Ni–C1 1.976(4), Ni–O 1.966(3), Ni–C2 1.843(4), C2–O 1.207(5), C2–B 1.639(5).

    Figure 5

    Figure 5. Plausible mechanisms for the formation of 2a. Relative Gibbs free energies (kcal mol–1) with respect to [1a + B1] (0.0 kcal mol–1) are shown, calculated at the PBE0-D3BJ/Def2-TZVPD//M06L/Def2-SVPD(Ni,F,O), Def2-SVP(others) level. Parts of the molecular structures for the selected compounds are also shown with selected bond lengths (Å).

    Figure 6

    Figure 6. Experimental and theoretical evaluation of the electronic state of the Ni complexes examined in this study. (a) Ni K-edge XAS spectra. (b) Ni L2,3-edge XAS spectra. (c) Sum-rule analysis. (d) A simplified Frontier-molecular-orbital energy diagram of 2a, focusing on the interaction between the Ni and B centers, calculated at PBE0-D3BJ/Def2-TZVPD level. (e) Kohn–Sham HOMO and LUMO+1 in 2a and their orbital composition.

  • References


    This article references 57 other publications.

    1. 1
      Jensen, W. B. The Lewis Acid-Base Definitions: A Status Report. Chem. Rev. 1978, 78, 131,  DOI: 10.1021/cr60311a002
    2. 2
      Crawley, M. L.; Trost, B. M. Applications of Transition Metal Catalysis in Drug Discovery and Development: An Industrial Perspective; John Wiley & Sons, 2012; . DOI: 10.1002/9781118309872 .
    3. 3
      Striegler, S.; Rieger, B.; Kacker, S.; Baugh, L. S. Late Transition Metal Polymerization Catalysis; Wiley VCH, 2003.
    4. 4
      Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books, 2010.
    5. 5
      Braunschweig, H.; Dewhurst, R. D.; Schneider, A. Electron-Precise Coordination Modes of Boron-Centered Ligands. Chem. Rev. 2010, 110, 39243957,  DOI: 10.1021/cr900333n
    6. 6
      Amgoune, A.; Bourissou, D. σ-Acceptor, Z-type ligands for transition Metals. Chem. Commun. 2011, 47, 859871,  DOI: 10.1039/C0CC04109B
    7. 7
      Braunschweig, H.; Dewhurst, R. D. Transition metals as Lewis bases: “Z-type” boron ligands and metal-to-boron dative bonding. Dalton Trans. 2011, 40, 549558,  DOI: 10.1039/C0DT01181A
    8. 8
      Bauer, J.; Braunschweig, H.; Dewhurst, R. D. Metal-Only Lewis Pairs with Transition Metal Lewis Bases. Chem. Rev. 2012, 112, 43294346,  DOI: 10.1021/cr3000048
    9. 9
      Parkin, G. Impact of the coordination of multiple Lewis acid functions on the electronic structure and vn configuration of a metal center. Dalton Trans. 2022, 51, 411427,  DOI: 10.1039/D1DT02921E
    10. 10
      Komuro, T.; Nakajima, Y.; Takaya, J.; Hashimoto, H. Recent progress in transition metal complexes supported by multidentate ligands featuring group 13 and 14 elements as coordinating atoms. Coord. Chem. Rev. 2022, 473, 214837214864,  DOI: 10.1016/j.ccr.2022.214837
    11. 11
      Green, M. L. H. A new approach to the formal classification of covalent compounds of the elements. J. Organomet. Chem. 1995, 500, 127148,  DOI: 10.1016/0022-328X(95)00508-N
    12. 12
      Davydova, E. I.; Sevastianova, T. N.; Timoshkin, A. Y. Molecular complexes of group 13 element trihalides, pentafluorophenyl derivatives and Lewis superacids. Coord. Chem. Rev. 2015, 297, 91126,  DOI: 10.1016/j.ccr.2015.02.019
    13. 13
      Cowley, A. H. From group 13–group 13 donor–acceptor bonds to triple-decker cations. Chem. Commun. 2004, 23692375,  DOI: 10.1039/B409497M
    14. 14
      Burlitch, J. M.; Leonowicz, M. E.; Petersen, R. B.; Hughes, R. E. Coordination of Metal Carbonyl Anions to Triphenylaluminum, -gallium, and -indium and the Crystal Structure of Tetraethylammonium Triphenyl((η5-Cyclopentadienyl)dicarbonyliron)aluminate(Fe–Al). Inorg. Chem. 1979, 18, 10971105,  DOI: 10.1021/ic50194a040
    15. 15
      Kameo, H.; Nakazawa, H. Saturated Heavier Group 14 Compounds as σ-Electron-Acceptor (Z-type) Ligands. Chem. Rec. 2017, 17, 268286,  DOI: 10.1002/tcr.201600061
    16. 16
      For examples including Pt-complexes, see:Bauer, J.; Braunschweig, H.; Dewhurst, R. D.; Radacki, K. Reactivity of Lewis Basic Platinum Complexes Towards Fluoroboranes. Chem.─Eur. J. 2013, 19, 87978805,  DOI: 10.1002/chem.201301056
    17. 17
      For examples including Fe-complexes, see:Burlitch, J. M.; Burk, J. H.; Leonowicz, M. E.; Hughes, R. E. Migration of triphenylboron from iron to a geminal η5-cylopentadienyl ligand. Inorg. Chem. 1979, 18, 17021709,  DOI: 10.1021/ic50196a061
    18. 18
      For examples including Rh-complexes, see:Alférez, M. G.; Moreno, J. J.; Gaona, M. A.; Maya, C.; Campos, J. Ligand Postsynthetic Functionalization with Fluorinated Boranes and Implications in Hydrogenation Catalysis. ACS Catal. 2023, 13, 1605516066,  DOI: 10.1021/acscatal.3c02764
    19. 19
      Hashimoto, T.; Asada, T.; Ogoshi, S.; Hoshimoto, Y. Main group catalysis for H2 purification based on liquid organic hydrogen carriers. Sci. Adv. 2022, 8, eade0189  DOI: 10.1126/sciadv.ade0189
    20. 20
      Sakuraba, M.; Morishita, T.; Hashimoto, T.; Ogoshi, S.; Hoshimoto, Y. Remote Back Strain: A Strategy for Modulating the Reactivity of Triarylboranes. Synlett 2023, 34, 21872192,  DOI: 10.1055/a-2110-5359
    21. 21
      Hisata, Y.; Washio, T.; Takizawa, S.; Ogoshi, S.; Hoshimoto, Y. In-silico-assisted derivatization of triarylboranes for the ccatalytic reductive functionalization of aniline-derived amino acids and peptides with H2. Nat. Commun. 2024, 15, 3708,  DOI: 10.1038/s41467-024-47984-0
    22. 22
      Goedecke, C.; Hillebrecht, P.; Uhlemann, T.; Haunschild, R.; Frenking, G. The Dewar-Chatt-Duncanson model reversed: Bonding analysis of group-10 complexes [(PMe3)2M–EX3] (M = Ni, Pd, Pt; E = B, Al, Ga, In, Tl; X = H, F, Cl, Br, I). Can. J. Chem. 2009, 87, 14701479,  DOI: 10.1139/V09-099
    23. 23
      Erdmann, P.; Greb, L. What Distinguishes the Strength and the Effect of a Lewis Acid: Analysis of the Gutmann–Beckett Method. Angew. Chem., Int. Ed. 2022, 61, e202114550  DOI: 10.1002/anie.202114550
    24. 24
      For examples including Ru-complexes, see:Hill, A. F.; Owen, G. R.; White, A. J. P.; Williams, D. J. The Sting of the Scorpion: A Metallaboratrane. Angew. Chem., Int. Ed. 1999, 38, 27592761,  DOI: 10.1002/(SICI)1521-3773(19990917)38:18<2759::AID-ANIE2759>3.0.CO;2-P
    25. 25
      Hill, A. F. An Unambiguous Electron-Counting Notation for Metallaboratranes. Organometallics 2006, 25, 47414743,  DOI: 10.1021/om0602512
    26. 26
      For examples including Rh-complexes, see:Crossley, I. R.; Hill, A. F.; Willis, A. C. Metallaboratranes: Tris(methimazolyl)borane Complexes of Rhodium(I). Organometallics 2006, 25, 289299,  DOI: 10.1021/om050772+
    27. 27
      Parkin, G. A. Simple Description of the Bonding in Transition-Metal Borane Complexes. Organometallics 2006, 25, 47444747,  DOI: 10.1021/om060580u
    28. 28
      For examples including Ni-, Cu-, Pd-, Ag-, Pt- and Au-complexes, see:Sircoglou, M.; Bontemps, S.; Bouhadir, G.; Saffon, N.; Miqueu, K.; Gu, W.; Mercy, M.; Chen, C.-H.; Foxman, B. M.; Maron, L.; Ozerov, O. V.; Bourissou, D. Group 10 and 11 Metal Boratranes (Ni, Pd, Pt, CuCl, AgCl, AuCl, and Au+) Derived from a Triphosphine–Borane. J. Am. Chem. Soc. 2008, 130, 1672916738,  DOI: 10.1021/ja8070072
    29. 29
      For examples including Ni- and Pd-complexes, see:Emslie, D. J. H.; Harrington, L. E.; Jenkins, H. A.; Robertson, C. M.; Britten, J. F. Group 10 Transition-Metal Complexes of an Ambiphilic PSB-Ligand: Investigations into η3(BCC)-Triarylborane Coordination. Organometallics 2008, 27, 53175325,  DOI: 10.1021/om800670e
    30. 30
      For examples including Ni-complexes, see:Harman, W. H.; Peters, J. C. Reversible H2 Addition across a Nickel-Borane Unit as a Promising Strategy for Catalysis. J. Am. Chem. Soc. 2012, 134, 50805082,  DOI: 10.1021/ja211419t
    31. 31
      Stephan, D. W.; Erker, G. Frustrated Lewis Pair Chemistry: Development and Perspectives. Angew. Chem., Int. Ed. 2015, 54, 64006441,  DOI: 10.1002/anie.201409800
    32. 32
      Jupp, A. R.; Stephan, D. W. New Directions for Frustrated Lewis Pair Chemistry. Trends Chem. 2019, 1, 3548,  DOI: 10.1016/j.trechm.2019.01.006
    33. 33
      Yamauchi, Y.; Mondori, Y.; Uetake, Y.; Takeichi, Y.; Kawakita, T.; Sakurai, H.; Ogoshi, S.; Hoshimoto, Y. Reversible Modulation of the Electronic and Spatial Environment around Ni(0) Centers Bearing Multifunctional Carbene Ligands with Triarylaluminum. J. Am. Chem. Soc. 2023, 145, 1693816947,  DOI: 10.1021/jacs.3c06267
    34. 34
      Carden, J. L.; Dasgupta, A.; Melen, R. L. Halogenated triarylboranes: synthesis, properties and applications in catalysis. Chem. Soc. Rev. 2020, 49, 17061725,  DOI: 10.1039/C9CS00769E
    35. 35
      Ashley, A. E.; Herrington, T. J.; Wildgoose, G. G.; Zaher, H.; Thompson, A. L.; Rees, N. H.; Krämer, T.; O’hare, D. Separating Electrophilicity and Lewis Acidity: The Synthesis, Characterization, and Electrochemistry of the Electron Deficient Tris(aryl)boranes B(C6F5)3-n(C6Cl5)n (n = 1–3). J. Am. Chem. Soc. 2011, 133, 1472714740,  DOI: 10.1021/ja205037t
    36. 36
      Sakuraba, M.; Hoshimoto, Y. Recent Trends in Triarylborane Chemistry: Diversification of Structures and Reactivity via meta-Substitution of the Aryl Groups. Synthesis 2024, 56, 3421,  DOI: 10.1055/s-0043-1775394
    37. 37
      Braunschweig, H.; Dewhurst, R. D.; Gessner, V. H. Transition metal borylene complexes. Chem. Soc. Rev. 2013, 42, 31973208,  DOI: 10.1039/c3cs35510a
    38. 38
      Kong, L.; Lu, W.; Yongxin, Y.; Ganguly, R.; Kinjo, R. Formation of Boron–Main-Group Element Bonds by Reactions with a Tricoordinate Organoboron L2PhB: (L = Oxazol-2-ylidene). Inorg. Chem. 2017, 56, 55865593,  DOI: 10.1021/acs.inorgchem.6b02993
    39. 39
      Reineke, M. H.; Sampson, M. D.; Rheingold, A. L.; Kubiak, C. P. Synthesis and Structural Studies of Nickel(0) Tetracarbene Complexes with the Introduction of a New Four-Coordinate Geometric Index, τδ. Inorg. Chem. 2015, 54, 32113217,  DOI: 10.1021/ic502792q
    40. 40
      Holschumacher, D.; Bannenberg, T.; Hrib, C. G.; Jones, P. G.; Tamm, M. Heterolytic Dihydrogen Activation by a Frustrated Carbene-Borane Lewis Pair. Angew. Chem., Int. Ed. 2008, 47, 74287432,  DOI: 10.1002/anie.200802705
    41. 41
      Chase, P. A.; Stephan, D. W. Hydrogen and Amine Activation by a Frustrated Lewis Pair of a Bulky N-Heterocyclic Carbene and B(C6F5)3. Angew. Chem., Int. Ed. 2008, 47, 74337437,  DOI: 10.1002/anie.200802596
    42. 42
      Choukroun, R.; Lorber, C.; Lepetit, C.; Donnadieu, B. Reactivity of [Cp2Ti(CO)2] and B(C6F5)3: Formation of the Acylborane Complexes [Cp2Ti(CO)(η2-OCB(C6F5)3)] and [Cp2Ti(THF)(η2-OCB(C6F5)3)]. Organometallics 2003, 22, 19951997,  DOI: 10.1021/om030185t
    43. 43
      Lyngdoh, R. H. D.; Schaefer, H. F.; King, R. B. Metal-Metal (MM) Bond Distances and Bond Orders in Binuclear Metal Complexes of the First Row Transition Metals Titanium Through Zinc. Chem. Rev. 2018, 118, 1162611706,  DOI: 10.1021/acs.chemrev.8b00297
    44. 44
      Shriver, D. F. BASCITY AND REACTIVITY OF METAL CARBONYLS. J. Organomet. Chem. 1975, 94, 259271,  DOI: 10.1016/S0022-328X(00)88721-5
    45. 45
      Gong, J.-K.; Kubiak, C. P. A New Route to the Ni(0) ‘Cradle’ Complex Ni2(μ-CO)(CO)2(PPh2CH2PPh2)2: μ-CO Ligand and Metal-centered Reactivity. Inorg. Chim. Acta 1989, 162, 1921,  DOI: 10.1016/S0020-1693(00)83112-6
    46. 46
      Bernhardt, E.; Finze, M.; Willner, H.; Lehmann, C. W.; Aubke, F. Salts of the Cobalt(I) Complexes [Co(CO)5]+ and [Co(CO)2(NO)2]+ and the Lewis Acid-Base Adduct [Co2(CO)7CO-B(CF3)3]. Chem.─Eur. J. 2006, 12, 82768283,  DOI: 10.1002/chem.200600210
    47. 47
      Wade, C. R.; Lin, T.-P.; Nelson, R. C.; Mader, E. A.; Miller, J. T.; Gabbaï, F. P. Synthesis, Structure, and Properties of a T-shaped 14-Electron Stiboranyl-Gold Complex. J. Am. Chem. Soc. 2011, 133, 89488955,  DOI: 10.1021/ja201092g
    48. 48
      Ansmann, N.; Münch, J.; Schorpp, M.; Greb, L. Neutral and Anionic Square Planar Palladium(0) Complexes Stabilized by a Silicon Z-Type Ligand. Angew. Chem., Int. Ed. 2023, 62, e202313636  DOI: 10.1002/anie.202313636
    49. 49
      For examples including Pt-complexes, see:Kameo, H.; Tanaka, Y.; Shimoyama, Y.; Izumi, D.; Matsuzaka, H.; Nakajima, Y.; Lavedan, P.; Le Gac, A.; Bourissou, D. Square-Planar Anionic Pt0 Complexes. Angew. Chem., Int. Ed. 2023, 62, e202301509  DOI: 10.1002/anie.202301509
    50. 50
      Wächtler, E.; Gericke, R.; Block, T.; Pöttgen, R.; Wagler, J. Trivalent Antimony as L-, X-, and Z-Type Ligand: The Full Set of Possible Coordination Modes in Pt-Sb Bonds. Inorg. Chem. 2020, 59, 1554115552,  DOI: 10.1021/acs.inorgchem.0c02615
    51. 51
      For examples including Au-complexes, see:Sircoglou, M.; Bontemps, S.; Mercy, M.; Saffon, N.; Takahashi, M.; Bouhadir, G.; Maron, L.; Bourissou, D. Transition-Metal Complexes Featuring Z-Type Ligands: Agreement or Discrepancy between Geometry and dn Configuration?. Angew. Chem., Int. Ed. 2007, 46, 85838586,  DOI: 10.1002/anie.200703518
    52. 52
      Dimucci, I. M.; Lukens, J. T.; Chatterjee, S.; Carsch, K. M.; Titus, C. J.; Lee, S. J.; Nordlund, D.; Betley, T. A.; MacMillan, S. N.; Lancaster, K. M. The Myth of d8 Copper(III). J. Am. Chem. Soc. 2019, 141, 1850818520,  DOI: 10.1021/jacs.9b09016
    53. 53
      DiMucci, I. M.; Titus, C. J.; Nordlund, D.; Bour, J. R.; Chong, E.; Grigas, D. P.; Hu, C. H.; Kosobokov, M. D.; Martin, C. D.; Mirica, L. M.; Nebra, N.; Vicic, D. A.; Yorks, L. L.; Yruegas, S.; MacMillan, S. N.; Shearer, J.; Lancaster, K. M. Scrutinizing Formally NiIV centers through the lenses of core spectroscopy, molecular orbital theory, and valence bond theory. Chem. Sci. 2023, 14, 69156929,  DOI: 10.1039/D3SC02001K
    54. 54
      Tran, V. T.; Li, Z. Q.; Apolinar, O.; Derosa, J.; Joannou, M. V.; Wisniewski, S. R.; Eastgate, M. D.; Engle, K. M. Ni(COD)(DQ): An Air-Stable 18-Electron Nickel(0)-Olefin Precatalyst. Angew. Chem., Int. Ed. 2020, 59, 74097413,  DOI: 10.1002/anie.202000124
    55. 55
      Hatsui, T.; Kosugi, N. Metal-to-ligand charge transfer in polarized metal L-edge X-ray absorption of Ni and Cu Complexes. J. Electron Spectrosc. Relat. Phenom. 2004, 136, 6775,  DOI: 10.1016/j.elspec.2004.02.133
    56. 56
      Solomon, E. I.; Hedman, B.; Hodgson, K. O.; Dey, A.; Szilagyi, R. K. Ligand K-Edge X-ray aabsorption spectroscopy: covalency of ligand-metal bonds. Coord. Chem. Rev. 2005, 249, 97129,  DOI: 10.1016/j.ccr.2004.03.020
    57. 57
      Sarangi, R.; George, S. D. B.; Rudd, D. J.; Szilagyi, R. K.; Ribas, X.; Rovira, C.; Almeida, M.; Hodgson, K. O.; Hedman, B.; Solomon, E. I. Sulfur K-Edge X-ray Absorption Spectroscopy as a Probe of Ligand-Metal Bond Covalency: Metal vs Ligand Oxidation in Copper and Nickel Dithiolene Complexes. J. Am. Chem. Soc. 2007, 129, 23162326,  DOI: 10.1021/ja0665949
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c15892.

    • Full details pertaining to the experimental methods, identification of the compounds, and DFT calculations (PDF)

    • AIM analysis (XLSX)

    • DFT coordinate (XLSX)

    Accession Codes

    Deposition Numbers 23847522384760 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via the joint Cambridge Crystallographic Data Centre (CCDC) and Fachinformationszentrum Karlsruhe Access Structures service.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.