Abstract
NiII(IB) dihalide [IB = (3aR,3a’R,8aS,8a’S)-2,2’-(cyclopropane-1,1-diyl)bis(3a,8a-dihydro-8H-indeno[1,2-d]-oxazole)] complexes are representative of a growing class of first-row transition metal catalysts for the enantioselective reductive cross-coupling of C(sp2) and C(sp3) electrophiles. Recent mechanistic studies highlight the complexity of these ground state cross-couplings, but also illuminate new reactivity pathways stemming from one-electron redox and their significant sensitivities to reaction conditions. For the first time, a diverse array of spectroscopic methods coupled to electrochemistry has been applied to NiII-based pre-catalysts to evaluate specific ligand field effects governing key Ni-based redox potentials. We also experimentally demonstrate DMA solvent coordination to catalytically-relevant Ni complexes. Coordination is shown to favorably influence key redox-based reaction steps and prevent other deleterious Ni-based equilibria. Combined with electronic structure calculations, we further provide a direct correlation between reaction intermediate frontier molecular orbital energies and cross-coupling yields. Considerations developed herein demonstrate the use of synergic spectroscopic and electrochemical methods to provide concepts for catalyst ligand design and rationalization of reaction condition optimization.
Graphical Abstract
1. Introduction
The number of accessible oxidation states and the possibility of both one- and two-electron redox reactions make nickel a well-suited alternative to precious metals such as iridium and platinum for cross-coupling catalysis. Since the first report of reductive coupling of aryl halides to biaryl products by bis(1,5-cyclooctadiene)nickel0 in 1971,1 numerous reports of coupling reactions, including enantioselective cross-coupling, have appeared.2 Early studies used electrochemistry to render these reactions catalytic in nickel,3 and in 2007 Durandetti et al. demonstrated that elemental manganese could be used as the terminal reductant for the Ni(bipyridine)Br2-catalyzed reductive cross-coupling of iodobenzene with α-chloroesters in up to 87% yield.4 The first highly selective enantioconvergent reductive cross-coupling was reported by one of our groups in 2013.5 Using bis(4-phenyloxazoline) (PhBOX), elemental manganese, and a 30% DMA/THF solvent mixture, high yields and enantioselectivity were obtained for cross-coupling of acyl chlorides and benzyl chlorides. Following this, a variety of chiral ligand frameworks were shown to be effective for the reductive cross-coupling of different electrophile pairs.2 For example, bis(indanyloxazoline) (Scheme 1, bottom, IB) Ni complexes catalyze a variety of reductive alkenylation reactions, including formation of enantioenriched allylic silanes and alkenes with aryl-substituted tertiary stereogenic centers (Scheme 1, bottom).6,7 More recently, one of our groups has also demonstrated electrocatalytic competency of these NiII complexes for cross-coupling reactions.8
Scheme 1.
Outline of key findings from this comprehensive spectroscopic, electrochemical, and computational study of NiII precatalysts for enantioselective reductive alkenylation. FMO = frontier molecular orbital.
A fundamental and detailed description of both the ground- and excited-state electronic properties of transition metal complexes can help infer catalytic operativity and competency. For example, one of our groups has recently provided detailed electronic structural and mechanistic studies of the ground- and excited-state properties of low-spin NiII bipyridine aryl halide complexes and their NiI photogenerated intermediates. These studies identified key structure-function relationships relevant for excited state bond homolysis and oxidative addition reactivity, which allowed for the mechanistic analysis of NiI-mediated activation of strong C(sp2)–Cl bonds.9,10 Our groups also recently participated in collaborative work that utilized cyclic voltammetry under catalytic conditions and UV-vis-NIR spectroelectrochemistry to better understand the interplay of Ni and Cr oxidation states for Nozaki-Hiyama-Kishi coupling.11 Another study by Neidig et al. utilized a combination of Mössbauer spectroscopy, magnetic circular dichroism, and computations to demonstrate that the kinetic competency of four-coordinate organoiron toward halogen abstraction is dependent on the accessibility and relative energy of iron-based orbitals.12
Despite the rapidly growing number of Ni-catalyzed reductive cross-coupling reactions being developed, there have been relatively few detailed mechanistic studies, particularly of bis(oxazoline) Ni complexes. Additionally, although solvent typically has a profound influence on the yield and selectivity of these reactions,13–18 the influence of solvent coordination to the active catalyst has not been thoroughly studied. A previous study has crystallographically characterized a five-coordinate, DMSO-bound NiII phenanthroline complex, also suggesting the potential influence of solvent in the catalytic cycle.19 A recent study by Diao et al. has suggested a NiI/III redox cycle and radical formation and capture are pertinent for catalysis by (biOx)NiIIArX catalysts, which revised previous suggestions of involvement of Ni0/II oxidation states.20 In general, uncertainty still exists regarding how the specific ligand, as well as the potential role of solvent coordination, influence both low-spin and high-spin NiII cross-coupling catalysis.
To this end, we provide the first comprehensive spectroscopic and electrochemical investigation of two reductive cross-coupling catalysts, NiII(IB)Cl2 and NiII(IB)Br2 [IB = (3aR,3a’R,8aS,8a’S)-2,2’-(cyclopropane-1,1-diyl)bis(3a,8a-dihydro-8H-indeno[1,2-d]-oxazole)], to provide broad insight into the complex conditions and mechanism (Scheme 1). Although studies related to these complexes by one of our groups were primarily in the context of developing reductive alkenylation, we note that these and similar complexes have also been reported for a number of other reactions, including reductive arylation, 1,2-alkynylboration, and 1,2-vinylboration, further motivating their detailed experimental and computational investigation.21–31 Here we use variable-temperature (VT) UV-vis-NIR absorption, circular dichroism (CD), vibrational CD (VCD), and magnetic CD (MCD) spectroscopies, coupled with cyclic voltammetry, spectroelectrochemistry, and DFT/TDDFT and multi-reference (CASSCF/CASPT2) calculations to elucidate specific electronic structure contributions to reactivity, as well as the influence of solvent on the efficacy of catalytic transformations (Scheme 1). Quantifying halide-dependent spectral features directly connects ligand field strength and redox potentials. Similarly, studying three solvents with different donor numbers (DNs) provides a route toward understanding previously reported empirical solvent optimization studies. We demonstrate that DMA solvent can coordinate to catalytically-relevant Ni species and evaluate the potential role of solvent coordination in their reactivity. Based on previously reported yields and findings reported here, we suggest solvent coordination and low temperature can favorably influence driving forces and kinetic barriers of key reaction steps leading to cross-coupled product. These studies provide new insights relevant to the catalytic reactivity of chiral bis(oxazoline) Ni catalysts that have been recently popularized for a variety of asymmetric Ni-catalyzed cross-coupling reactions (Scheme 1).
2. Results and Analysis
2.1. Room-Temperature UV-Vis-NIR, CD, and MCD Characterization of NiII(IB)Cl2 and NiII(IB)Br2 in a Non-Coordinating Solvent
Due to the absence of low-energy charge transfer bands in the UV-vis-NIR, electronic absorption spectra of NiII(IB)Cl2 and NiII(IB)Br2 in DCM (donor number, DN = 2.4)32,33 (Figures 1A–C and 1D–F, respectively) provide rich insights into the NiII ligand field through observation of numerous spin-allowed and spin-forbidden ligand field transitions (Table 1). These transitions arise from the orbital triplet ground state (i.e., 3T1(F) in an idealized Td geometry) of the S = 1 d8 NiII complexes and exhibit correspondingly weak molar absorptivities at 294 K. CD and MCD spectroscopies provide complementary, signed, spectroscopic methods to resolve overlapping ligand field transitions. Due to the presence of the chiral IB ligand, the NiII-based ligand field transitions exhibit CD intensity (Figures 1B and 1E), and both complexes exhibit room temperature MCD signals at 1.4 T (Figures 1C and 1F, respectively). A detailed discussion of band assignments can be found in Supporting Information Section S3.
Figure 1.
Spectroscopic characterization of the Ni complexes. Room temperature (A, D) UV-vis-NIR absorption, (B, E) CD, and (C, F) MCD spectra of NiII(IB)Cl2 and NiII(IB)Br2 in DCM with Gaussian resolutions of individual transitions (orange). Overall fit to each spectrum given in dashed blue.
Table 1.
Band maxima and full widths at half maxima (FWHM) for NiII(IB)Cl2 and NiII(IB)Br2 in DCM obtained from simultaneous Gaussian resolution of UV-vis-NIR, CD, MCD, and vibrational CD spectra. Numbered states correspond to spin-allowed ligand field transitions.
NiII(IB)Cl2 | |||||||
---|---|---|---|---|---|---|---|
Absorption (cm−1) | CD (cm−1) | MCD (cm−1) | Assignment | ||||
# | Band max. | FWHM | Band max. (Sign) |
FWHM | Band max. (Sign) |
FWHM | Td: C2v Geometry |
1 | --- | --- | 2170 (+) | 730 | --- | --- | 3T1(F): 3B1(F) → 3A2(F) |
2 | 6640 | 2160 | --- | --- | --- | --- | 3T2: 3B1(F) → 3B1/3B2 |
3 | 8240 | 1680 | 8820 (−) | 1480 | 8010 (−) | 2000 | 3T2: 3B1(F) → 3A1 |
4 | 10 270 | 2110 | 10 750 (+) | 1500 | 10 030 (−) | 2110 | 3A2(F): 3B1(F) → 3A2(F) |
i | 12 100 | 400 | 12 020 (+) | 330 | 12 030 (−) | 250 | 3B1(F) → 1T2/1E |
--- | --- | 12 120 (−) | 180 | 12 100 (+) | 170 | 3B1(F) → 1T2/1E | |
ii | 12 400 | 1630 | 12 200 (+) | 1150 | --- | --- | Spin-forbidden LF |
5 | 14 940 | 1430 | --- | --- | 14 610 (+) | 1000 | 3T1(P): 3B1(F) → 3B1(P) |
iii | 15 890 | 1910 | 15 500 (−) | 1950 | 15 530 (+) | 1570 | --- |
6 | 18 210 | 1220 | 18 000 (−) | 1250 | 18 320 (+) | 1280 | 3T1(P): 3B1(F) → 3B2(P) |
iv | --- | --- | 19 050 (−) | 1250 | --- | --- | --- |
7 | 20 130 | 2800 | 20 130 (+) | 2300 | 20 270 (−) | 2330 | 3T1(P): 3B1(F) → 3A2(P) |
v | --- | --- | 23 740 (−) | 1600 | --- | --- | LF of Trimer |
NiII(IB)Br2 | |||||||
Absorption (cm−1) | CD (cm−1) | MCD (cm−1) | Assignment | ||||
# | Band max. | FWHM | Band max. (Sign) | FWHM | Band max. (Sign) | FWHM | Td: C2v Geometry |
1 | --- | --- | 2210 (+) | 970 | --- | --- | 3T1(F): 3B1(F) → 3A2(F) |
2 | 6150 | 2000 | --- | --- | --- | --- | 3T2: 3B1(F) → 3B1/3B2 |
3 | 7980 | 1670 | 8670 (−) | 1450 | 7980 (−) | 2150 | 3T2: 3B1(F) → 3A1 |
4 | 10 150 | 2080 | 10 460 (+) | 1280 | 10 060 (−) | 1930 | 3A2(F): 3B1(F) → 3A2(F) |
--- | --- | --- | --- | 11 500 (−) | 180 | 3B1(F) → 1T2/1E | |
i | 11 670 | 500 | 11 660 (+) | 500 | 11 750 (+) | 500 | 3B1(F) → 1T2/1E |
ii | 12 380 | 1750 | 12 000 (+) | 1830 | --- | --- | Spin-forbidden LF |
5 | 14 460 | 1450 | 14 900 (−) | 1530 | 14 210 (+) | 1280 | 3T1(P): 3B1(F) → 3B1(P) |
iii | 15 600 | 1330 | --- | --- | 15 290 (+) | 970 | --- |
6 | 17 500 | 2000 | 17 560 (−) | 1650 | 17 920 (+) | 1500 | 3T1(P): 3B1(F) → 3B2(P) |
iv | --- | --- | 18 500 (−) | 1030 | --- | --- | --- |
7 | 19 410 | 2660 | 19 200 (+) | 1670 | 19 420 (−) | 2500 | 3T1(P): 3B1(F) → 3A2(P) |
v | --- | --- | 21 130 (+) | 1030 | --- | --- | 3B1(F) → 1T1g, 1T2g, 1A1g, 1Eg |
vi | --- | --- | 25 100 (+) | 2400 | --- | --- | LMCT |
The ligand field transitions of NiII(IB)Cl2 relative to NiII(IB)Br2 are blueshifted by ~120 – 710 cm−1 in DCM (overlaps in Figure S17), consistent with stronger donation from chloride relative to bromide. To estimate the relative ligand field strengths, we average the assigned spin-allowed transitions (bands 1 – 7). Doing so provides relative ligand field strengths of ~12 365 cm−1 and ~11 975 cm−1 (Δ = ~390 cm−1) for NiII(IB)Cl2 and NiII(IB)Br2, respectively, in accord with the greater ligand field strength in the chloride complex relative to bromide. As will be shown below, the ligand field bands and ligand field strengths can be directly related to the energies of the NiII-based redox active molecular orbitals (RAMOs) and, thus, complex redox potentials (vide infra, Section 3.3). To the best of our knowledge, this provides the first experimental connection between ligand field spectroscopy and the electrochemical potentials of NiII-based enantioselective cross-coupling catalysts.
2.2. Vibrational CD Spectroscopy of NiII(IB)Cl2 and NiII(IB)Br2
Pseudo-Td Ni(II)/Co(II) and pseudo-Oh V(III) complexes can exhibit large splittings of their orbital triplet ground states. These splittings can be observed using techniques such as electronic Raman or VCD. The latter is the infrared analogue of electronic CD spectroscopy and provides a means to determine absolute stereochemical configurations or probe low-energy electronic transitions in chiral transition metal complexes.34,35 VCD signals are observed for both NiII(IB)Cl2 and NiII(IB)Br2 complexes dissolved in d2-DCM (Figure 2). Note high sample concentrations are necessary for VCD measurements. Despite substantial trimerization of NiII(IB)Cl2 at these concentrations (vide infra, Section 2.4), the observed VCD transition is still assigned to the four-coordinate species, as evidenced by calculations (Figure S121) and spectral consistency with the NiII(IB)Br2 analogue, which does not undergo trimerization at high concentrations in DCM. Furthermore, to confirm the observed VCD spectral intensity for NiII(IB)X2 complexes corresponds to a ligand field transition, the d10 complex, ZnII(IB)Cl2, was synthesized in a manner analogous to the NiII complexes. Note that an analogous complex, CuII(IB)Cl2, has been previously reported.36 The VCD spectrum of ZnII(IB)Cl2 in d2-DCM does not exhibit electronic absorption in the region of 1800 – 3200 cm−1 (Figure 2). NiII(IB)Cl2 exhibits a moderately sharp (FWHM = ~730 cm−1) ligand field transition at ~2170 cm−1, while NiII(IB)Br2 exhibits a transition (FWHM = ~970 cm−1) at ~2210 cm−1. For both complexes, this band (band 1, 3B1(F) → 3A2 transition in C2v) arises from a transition within the orbital triplet ground state (3T1(F) in Td), which is split by low symmetry distortions and spin-orbit coupling. Note the positive sign of the VCD band is consistent with the other 3A2 states observed with CD (Figure 1B and 1E). While the relative ground state splittings are quite similar between the two complexes, the small increase in splitting for NiII(IB)Br2 is likely due to the greater spin-orbit coupling constant for Br relative to Cl.
Figure 2.
Background-corrected vibrational CD spectra of 160.0 mM NiII(IB)Cl2, 139.0 mM NiII(IB)Br2, and 115.2 mM ZnII(IB)Cl2 in d2-DCM.
This application of VCD determines electronic excited state energies of the low-symmetry split orbital triplet ground state of transition metal-based enantioselective cross-coupling catalysts for the first time. These data, combined with the UV-vis-NIR, electronic CD, and MCD data, have allowed for the experimental determination of a complete ligand field energy level diagram for the NiII(IB)X2 complexes. All transitions to individual excited states in Td and C2v symmetry are assigned in the correlation diagram in Figure 3.
Figure 3.
Descent in symmetry state diagram for NiII(IB)Cl2. Starred states are derived from CASSCF/CASPT2 calculations and shifted by the average ratio of the seven experimental spin-allowed transitions vs. seven calculated ligand field energies. Numbered states given in Table 1.
2.3. Room Temperature UV-Vis-NIR, CD, and MCD Characterization of NiII(IB)Cl2 and NiII(IB)Br2 in Coordinating Solvents
To complement data obtained in DCM (poor σ and π donor) and to assess the solvent-dependent behavior of NiII(IB)X2 complexes, room temperature UV-vis-NIR absorption, CD, and 1.4 T MCD spectra were also acquired in MeCN (moderate σ donor and moderate π acceptor) and DMA (very strong σ donor and moderate π acceptor). These solvents were selected for their relative differences in dielectric constant and Lewis basicity/acidity to assess the effects of solvent environment on geometric and electronic structure, which can subsequently be correlated with catalytic activity and selectivity.
UV-vis-NIR spectra of NiII(IB)Cl2 and NiII(IB)Br2 exhibit significant solvent dependence (Figures S9 and S10). In addition to decreased overall intensities compared to transitions observed in DCM, new spectral intensity grows in at ~23 000 cm−1, with weak intensity in MeCN and greatest intensity in DMA. These spectral changes are also manifested in the solvent-dependent CD and MCD spectra (Figures S44–S47), with the new spectral intensity at ~23 000 cm−1 corresponding to a new negative band in CD and MCD. As demonstrated in Section 2.4 below, this new band reflects an equilibrium between the four- and five-coordinate, solvent coordinated species. Thus, electronic spectroscopies provide a direct handle on DMA coordination to NiII dihalide complexes relevant to catalysis.
2.4. Variable-Concentration and Variable-Temperature UV-Vis-NIR Spectroscopy of NiII(IB)Cl2 and NiII(IB)Br2
Solvent and additive evaluation are necessary steps in the optimization of transition metal-catalyzed organic reactions. However, these steps can be somewhat arbitrary and rely on a large empirical screening matrix. Previous studies of NiII(IB)X2 demonstrated catalytic yields are maximized in DMA and when the reaction is cooled to 0–5 °C.6–8 By obtaining VT UV-vis-NIR spectra for both NiII(IB)Cl2 and NiII(IB)Br2 in DCM and DMA, we aimed to provide insight into precatalyst speciation and, in turn, to provide experimental thermodynamic data for rational catalyst and condition design.
As the concentration of NiII(IB)Cl2 is increased in DCM, a noticeable color change from pink to orange is observed. Correspondingly, new electronic absorption bands are observed at 13 100 cm−1 and 22 840 cm−1 with increasing concentration (Figure S16). These additional absorption bands are ascribed to the formation of a [NiII(IB)Cl2]3 μ-Cl trimer. A previously obtained crystal structure of this species shows it possesses both five- and six-coordinate formal NiII centers.37
The UV-vis-NIR spectra of NiII(IB)Cl2 in DCM (4.3 mM) and DMA (3.6 mM) also depend on temperature (Figure 4A and 4B). Because of the number of overlapping transitions present in each spectrum, the VT spectra were resolved using global modeling of the temperature dependence through nonlinear regression at multiple wavelengths and bootstrapping (Figures S13–S15).38,39 In DMA, the VT UV-vis-NIR spectra reflect an equilibrium between the four- and five-coordinate, DMA coordinated species for NiII(IB)Cl2 (five-coordinate absorption maximum at 22 840 cm−1) and NiII(IB)Br2 (five-coordinate absorption maximum at 23 230 cm−1). In DCM, the VT UV-vis-NIR spectra reflect an equilibrium between a monomeric and trimeric form (trimer absorption maxima at 22 840 cm−1 and 20 160 cm−1). We note the excellent agreement between the resolved spectra of the four-coordinate species and the spectra of isolated NiII(IB)Cl2 and NiII(IB)Br2 obtained in DCM (Figure 4 and S12). In contrast to NiII(IB)Cl2, NiII(IB)Br2 exhibited no trimerization up to 152.1 mM or in VT studies down to –85 °C (Figures S18–S19).
Figure 4.
VT studies of NiII precatalysts. VT UV-vis-NIR spectra of (A) 4.3 mM NiII(IB)Cl2 in DCM and (B) 3.6 mM NiII(IB)Cl2 in DMA, as well as corresponding resolved spectra from global nonlinear regression for (C) NiII(IB)Cl2 monomer and trimer and (D) four- and five-coordinate NiII(IB)Cl2 monomer. Note analogous spectra for NiII(IB)Br2 are presented in Figures S11–12.
In addition to resolving the spectra for four-coordinate, five-coordinate, and trimeric species, these fits provide thermodynamic parameters based on the two equilibrium expressions,
(1) |
values for DMA coordination to NiII(IB)Cl2 and NiII(IB)Br2 are similar (0.58 and 0.97, respectively, Table 2). Furthermore, is calculated at 294 K for formation of the [NiII(IB)Cl2]3 μ-Cl trimer in DCM. Despite the exergonicity of this process (Table 2), the extent of reaction is only 0.09% at 4.3 mM, which is why essentially no trimer is observed at room temperature at low NiII concentrations. As expected for associative reactions, all entropy values for solvent coordination and trimerization are negative. Note that the extent of temperature-dependent reaction is primarily sensitive to the ratio of the enthalpy to entropy, and the global fitting can extract this value with <1% uncertainty. However, global fitting can additionally extract the absolute standard enthalpies and entropies, albeit with higher uncertainty. From values, we estimate concentrations of ~37% and ~49% for DMA-coordinated NiII(IB)Cl2 (3.6 mM) and NiII(IB)Br2 (3.5 mM), respectively, at 294 K. Upon cooling to 273 K, these values increase significantly to ~53% and ~69%, respectively.
Table 2.
Thermodynamic reaction parameters for transition metal oligomers and solvent adducts of NiII(IB)X2 species. 95% confidence intervals of global fitting given in brackets.
Complex | Standard Enthalpy [Range] | Standard Entropy [Range] | Free Energy (294 K) |
Keq (294 K) |
---|---|---|---|---|
Ni II (IB)(DMA)Cl 2 | −5.0 kcal mol−1 [−4.3, −7.1] | −18.1 cal mol−1 K−1 [−15.4, −25.1] | 0.3 kcal mol−1 | 0.58 |
Ni II (IB)(DMA)Br 2 | −6.1 kcal mol−1 [−4.2, −8.3] | −20.8 cal mol−1 K−1 [−13.4, −28.9] | 0.02 kcal mol−1 | 0.97 |
[Ni II (IB)Cl 2 ] 3 | −8.2 kcal mol−1 [−4.1, −12.5] | −22.3 cal mol−1 K−1 [−1.5, −43.7] | −1.6 kcal mol−1 | 16.7 |
As depicted in Figure 5, the preceding analysis provides the first detailed view of the effects of dielectric constant, concentration, and solvent donicity on the speciation equilibria of NiII precatalyst solutions. With high solvent donicity, the equilibrium shifts toward a five-coordinate, solvent coordinated species for both NiII(IB)Cl2 and NiII(IB)Br2. Increasing concentration and low dielectric constant shifts the equilibrium toward a trimeric species for NiII(IB)Cl2. Thus, solvent coordination, temperature, and catalyst concentration are important effects that contribute to catalyst speciation and activity under reaction conditions (vide infra, Discussion). VT spectroscopies provide a direct handle on the speciation and the associated thermodynamics.
Figure 5.
Equilibrium behavior of NiII(IB)X2 based on solvent conditions and concentration. Note while trimerization is not observed for NiII(IB)Br2, it may form at higher concentrations or lower temperatures accessed herein.
2.5. Cyclic Voltammetry of NiII(IB)Cl2 and NiII(IB)Br2 in DCM, MeCN, and DMA
To assess the effects of solvent donicity and dielectric constant on the electrochemical properties of NiII(IB)X2 complexes, scan rate-dependent cyclic voltammetry data were acquired in DCM, MeCN, and DMA (Figure 6 and Figures S48–S59). Table 3 provides peak and formal potentials for initial redox events, while Table S1 provides peak and formal potentials for unique re-oxidation and re-reduction events that result from chemical reactions following initial electron transfers. Additional details and discussions are provided in Supporting Information Section S5.
Figure 6.
Solvent-dependent, diffusion- and concentration-normalized voltammetry of NiII(IB)Br2 (details of normalization are provided in Section S5 in the Supporting Information). All electrochemistry acquired in 0.1 M TBAPF6 solution at a scan rate of 100 mV s−1.
Table 3.
Electrochemical parameters for NiII(IB)Cl2 and NiII(IB)Br2 in 0.1 M TBAPF6 electrolyte solution using a glassy carbon working electrode, 0.01 M Ag+/0 non-aqueous reference electrode, and platinum wire counter electrode. All peak and formal potentials are given in volts, obtained using a 100 mV s−1 scan rate (unless otherwise stated), and referenced to Fc+/0.
Complex | Solvent | E p,a,1 | E p,a,2 | E p,c | E p/2, c | a | a | D0 (cm2 s−1)b |
---|---|---|---|---|---|---|---|---|
Ni II (IB)Cl 2 | DCM | 1.51d | --- | −1.58 | −1.44 | −1.47 | 1.38d | 9.25 × 10−6 |
Ni II (IB)Br 2 | DCM | 0.94 | --- | −1.46 | −1.26 | −1.26 | 0.85 | 9.25 × 10−6 |
Ni II (IB)Cl 2 | MeCN | 0.96 | --- | −1.46 | −1.25 | −1.32 | 0.81 | 1.05 × 10−5 |
Ni II (IB)Br 2 | MeCN | 0.36 | 0.75 | −1.21 | −1.07 | −1.05 | 0.30/0.67 | 9.55 × 10−6 |
Ni II (IB)Cl 2 | DMA | 0.67e | --- | −1.54 | −1.42 | −1.47 | 0.48 | 2.81 × 10−6 |
Ni II (IB)Br 2 | DMA | 0.22 | --- | −1.37 | −1.23 | −1.23 | 0.15 | 3.06 × 10−6 |
Ni II ( diBn biOx)Br 2 c | 1,2-DFB | --- | --- | −1.90 | −1.55 | --- | --- | --- |
Ni II ( iPr biOx)Br 2 c | 1,2-DFB | --- | --- | −1.64 | −1.46 | --- | --- | --- |
Ni II ( diMe biOx)(Dipp)Br c | THF | --- | --- | −2.39 | −2.12 | --- | --- | --- |
From the inflection potential of the redox process at 100 mV s−1, which approximates the formal potential.
Derived from mass transport-controlled current at a disk microelectrode.
See reference 40.
Estimated from the local minimum of dj/dV at 25 mV s−1.
Estimated from the local minimum of dj/dV at 100 mV s−1.
Ep/2, c is the potential at half of the peak current.
Previous studies have provided formal potentials for both aromatic and non-aromatic NiII diimine systems, with many reports providing kinetic analyses with substrate present.41,20 However, to our knowledge, this is the first example of detailed solvent-dependent electroanalytical chemistry for non-aromatic NiII cross-coupling catalysts. In general, precatalyst electrochemical responses are remarkably solvent dependent. In all three solvents, NiII(IB)Cl2 and NiII(IB)Br2 both exhibit single, electrochemically irreversible reduction events with significantly shifted oxidative waves (Figures 6, and S50, S54, S58). The general irreversibility required use of peak potentials (Ep,a or Ep,c), potentials at half of the peak current value (Ep/2), and inflection potentials (accurate estimate of formal potential, E0’) for analysis.42 Our measured Fc formal potentials in DMA and MeCN are 85 and 91 mV vs 0.01 M Ag+/0, respectively, indicating accurate conclusions can be drawn regarding solvent effects on measured formal potentials of the Ni complexes in these solvents. In DCM, the measured Fc formal potential is 215 mV. Therefore, measured formal potentials in DCM will appear negatively shifted relative to values in DMA and MeCN.
Based on shifts in peak potential as a function of scan rate and scan rate normalized voltammetry (current function) in all three solvents (Figures S48–S58), as well as differential pulse voltammetry and variable temperature voltammetry in DMA for NiII(IB)Cl2 (Figure S58), we can draw some insightful conclusions regarding the reduction mechanism, as the current function and shift in peak potential are dictated by the particular chemical and electrochemical mechanism. These conclusions also apply to NiII(IB)Br2. We ascribe the reduction of both complexes to a concerted EqCi (in DMA and MeCN, solvent coordination and/or halide loss occur in concert) or step-wise EiCi mechanism (DCM), where slow electron transfer is followed by rapid halide loss. In DCM, three-coordinate NiI(IB)X will be generated upon reduction, with no subsequent solvent coordination. The lack of return current, shift in peak potential as a function of the logarithm of the scan rate near 29.6 mV, with ~33 mV observed here, and decrease in the current function toward a limiting value as the scan rate is increased supports a kinetically-controlled, stepwise reduction followed by rapid halide loss (Figures S48 and S50). Activation of DCM by other nickel complexes supported by naphthyridine-diimine ligands has been observed previously.43 However, spectroelectrochemical data obtained in DCM do not support reactivity of the NiI(IB)X with solvent (Figures S73–S75). Overall, this analysis featuring electron transfer coupled to rapid halide loss is consistent with halide dissociation observed previously using extended X-ray absorption fine structure (EXAFS) for a low-spin NiII biOx aryl halide complex upon reduction with potassium graphite.20
Experimental formal potentials for chemically-coupled reduction of NiII(IB)Cl2 and NiII(IB)Br2 to NiI(IB)Cl and NiI(IB)Br in DCM are −1.47 V and −1.26 V vs. Fc+/0, respectively (Table 3). It is therefore ~0.21 V (~1695 cm−1) harder to reduce NiII(IB)Cl2 relative to NiII(IB)Br2. This observation is consistent with the energetic shifts in the spin-allowed ligand field bands in DCM in experiment (vide supra, Section 2.1) and calculations (vide infra, Section 3.2 and 3.3). In MeCN and DMA, NiII(IB)Cl2 (Figures S54 and S58) and NiII(IB)Br2 (Figure 6, left) both exhibit superficially quasi-reversible voltammetry for the reduction. Experimental formal potentials for chemically-coupled reduction in MeCN/DMA of NiII(IB)Cl2 and NiII(IB)Br2 are −1.32/−1.47 V and −1.05/−1.23 V vs. Fc+/0, respectively (Δ = ~0.27/~0.24 V (~2180/1935 cm−1)). Thus, for all solvents used here, it is harder to reduce NiII(IB)Cl2 relative to NiII(IB)Br2.
Based on the VT UV-vis-NIR data in DMA (vide supra, Section 2.4), both the NiII four-coordinate and five-coordinate solvent adducts exist in equilibrium, and this can potentially influence the electrochemistry measured in this solvent. One possibility for the reduction mechanism for these species is reduction followed by halide loss and, for the four-coordinate portion of the complex, coordination of DMA to the NiI center, which could occur in a concerted or stepwise fashion. For a concerted mechanism, the anticipated shift in peak potential as a function of log(v) is 29.6/α mV, where α is the transfer coefficient for electron transfer.42 Based on the observation of only one differential pulsed voltammetry wave on the forward scan and the shift in peak potential with log(v) (~77–104 mV), we propose that the reduction and chemical follow up reaction in both MeCN and DMA (i.e., solvent coordination at NiI) is a concerted process. The two return waves observed scanning oxidatively suggest generation of a halide-dissociated species that is re-oxidized at more positive potentials. This conclusion is supported by VT differential pulse voltammetry (Figure S58), where the differential current at the more positive wave decreases as temperature is decreased, while the differential current at the wave ascribed to re-oxidation of five-coordinate NiI increases. Based on behavior previously observed for these systems and our computed formal potentials,20 the more positive re-oxidation could arise from re-oxidation of a NiI/NiI dimer that forms after the initial reduction. However, we favor the interpretation featuring re-oxidation of the halide-dissociated species based on computed formal potentials (vide infra, Section 3.3) and lack of return oxidation near the reduction event in DCM, where NiI is anticipated to dimerize rapidly. Further supporting our hypothesis, an additional wave near where three-coordinate NiI is predicted to oxidize is present in MeCN, but not in DMA (Figures S52 and S54), supporting the weaker coordination affinity of MeCN and our assignment of the species generated upon reduction.
Potentials for chemically-coupled reductions in DMA are more negative relative to MeCN by ~200 mV for both complexes. As discussed further in Section 3.3, this difference is ascribed to DMA being a higher donicity solvent and coordinating to the NiII center. Note that solvent coordination is not observed in DCM and only weakly so in MeCN. It is further interesting to note that the reduction potential for NiII(IB)Cl2 in both DCM and DMA is −1.47 V vs. Fc+/0, respectively; for NiII(IB)Br2, these are −1.26 V and −1.23 V vs. Fc+/0, respectively. The similarity in reduction potentials in DCM and DMA is ascribed to the relative Fc formal potentials in DCM vs. DMA and the role of solvent in facilitating the Ni–X bond rupture upon one-electron reduction, with the anionic halide loss more facile in DMA relative to DCM. Because of these considerations and the electronic structure calculations presented in Section 3.3, the more quantitative comparison of potentials for the reduction with and without coordinated solvent is that between MeCN and DMA. Furthermore, the temperature-dependent cyclic voltammetry demonstrates a negatively shifted reduction potential as the temperature is lowered, which may be due to increasing the relative amount of five-coordinate species. Thus, overall, solvent coordination results in a harder to reduce NiII center. By extension, this can be further translated to a more reducing NiI species, which, under catalytic conditions, can facilitate oxidative addition (vide infra, Discussion).44
Ligand field and bonding effects on NiII-based redox potentials can be further elucidated using electronic structure calculations (vide infra, Section 3.3) and through correlations to electronic spectroscopy, as transitions to the RAMO are also observed experimentally. Differences in measured redox potentials correlate directly with specific structural influences on the energy of the RAMO.
Finally, based on the measured formal potentials, proposed electrochemical mechanisms, and additional electronic structure calculations of redox potentials (vide infra, Section 3.3), we do not believe Ni0(IB)X2 (or Ni0 in any form) is thermodynamically accessible in the electrochemical window of common electrochemistry solvents, which supports a NiI/III catalytic cycle for reductive alkenylation and potentially related reactions involving bis(oxazoline)–Ni complexes.20 No additional reduction beyond NiI is required for oxidative addition of substrates for which this catalyst has been previously demonstrated to be competent, and this finding has important mechanistic implications for bis(oxazoline)–Ni-catalyzed reactions more generally.
2.6. Spectroelectrochemistry of NiII(IB)Cl2 and NiII(IB)Br2 in DCM and DMA
To rationalize the noticeable difference in electrochemical response of NiII(IB)X2 in DCM vs. DMA and to further understand the solvent-dependent catalytic activity, time-based spectroelectrochemical measurements were performed for both negative and positive polarizations. These measurements are the first comprehensive solvent-dependent spectroelectrochemistry for NiII cross-coupling catalysts. All experimental plots are shown in Figures S61–S87. In conjunction with calculations (vide infra, Figures S133–S140), we can assign transient spectra to possible species generated under polarized conditions. In DCM, polarization negative of the first reduction generates a spectrum with a slightly blue-shifted, higher-energy ligand field transition, consistent with calculated spectra for NiI(IB)X (Figures S133 and S135).
Reduction of both NiII(IB)Br2 and NiII(IB)Cl2 in DMA generates spectra consistent with calculated spectra corresponding to a solvent-coordinated NiI species, NiI(IB)(DMA)X (Figures S134 and S136). All ligand field bands decay in intensity upon oxidation, which is ascribed to oxidative degradation or oligomerization. Spectra obtained after positive polarization of NiII(IB)Br2 are consistent with bromide speciation.45
3. Computational Results
In this section, we sought to gain further insights into the electronic structures of the precatalysts by comparing experimental spectra and redox potentials with computed ground- and excited-state properties obtained from a combination of DFT, time-dependent DFT (TDDFT), and multireference CASSCF/CASPT2 calculations. TDDFT and multireference methods predict electronic transitions below ~25 000 cm−1 are ligand field excitations, and differences between computed electronic spectra of NiII(IB)X2 complexes arise from differences in ligand field strength. Generally, increased ligand field strength destabilizes the β lowest unoccupied molecular orbitals (β-LUMOs), which negatively shifts the reduction potential of NiII(IB)Cl2 relative to NiII(IB)Br2. DFT calculations also corroborate that, depending on reaction conditions (such as the choice of coordinating/non-coordinating solvent or NiII concentration), precatalysts can exist in complex equilibria featuring four-coordinate NiII(IB)X2, five-coordinate NiII(IB)(solv.)X2, or trimeric [NiII(IB)X2]3 μ-X species. Finally, we demonstrate the role of ligand–metal covalency in tuning the relative reactivity of the catalyst resting state; the bonding of this species further manifests in a correlation between reaction yield and its oxidation potential relative to those of C(sp3) radicals.
3.1. DFT and TDDFT Calculated Thermodynamics and Spectra of NiII(IB)Cl2 and NiII(IB)Br2 Precatalysts
Using a TPSSh functional with a conductor-like polarizable continuum model (CPCM) (see Computational Details in Section S8 of the Supporting Information), the ground state wave functions of NiII(IB)Cl2 and NiII(IB)Br2 are high-spin (S = 1) NiII. TDDFT was used to calculate the electronic transition energies (Table 4). The intensities of the lower-energy ligand field bands are underestimated relative to experiment and do not contribute significantly to the overall predicted spectra (Figures S92–S93). Additionally, two-electron excitations and spin-flip transitions are inaccessible through conventional TDDFT and, thus, are not observed in the spectral predictions (e.g., 3A2(F), band 4 and 1T2/1E, band i excited states in Td). For d8 NiII, there are only six spin-allowed ligand field transitions accessible using TDDFT (Table 4). Therefore, full assignment of all absorption bands from Section 2.1 cannot be achieved using this approach. A more detailed analysis must be obtained from the multi-reference CASSCF/CASPT2 calculations (vide infra, Section 3.2). Nonetheless, qualitative correlations can be made by comparing to assignments in the parent Td point group (Table 4). While the absolute calculated energies of the ligand field transitions are not well-reproduced with TDDFT, the average energies agree well with the experimentally determined relative ligand field strengths (vide supra, Section 2.1). For example, experimental values of ~12 365 cm−1 and ~11 975 cm−1 (Δ = ~390 cm−1) were determined for NiII(IB)Cl2 and NiII(IB)Br2, respectively, and are computationally estimated to be ~13 780 cm−1 and ~13 370 cm−1 (Δ = ~410 cm−1).
Table 4.
Experimental (UV-vis-NIR absorption) and predicted electronic transition energies for NiII(IB)Cl2 and NiII(IB)Br2.
Predicted transition energies (cm−1) | Assignment, (#)b | |||||
---|---|---|---|---|---|---|
Ni(IB)Cl2 | Ni(IB)Br2 | |||||
Experiment | DFT | CASPT2 | Experiment | DFT | CASPT2 | |
2170a | 7629 | 2009 | 2210a | 7242 | 1824 | 3T1(F): 3B1(F) → A2(F), (1) |
--- | 10 056 | 3318 | --- | 9730 | 2974 | 3T2: 3B1(F) → 3B1/3B2 |
6640 | 14 393 | 7443 | 6150 | 13 924 | 7229 | 3T2: 3B1(F) → 3B1/3B2, (2) |
8240 | 14 991 | 9786 | 7980 | 14 972 | 9322 | 3T2: 3B1(F) → 3A1, (3) |
--- | --- | 10 810 | --- | --- | 10 244 | 3T2: 3B1(F) → 3B1/3B2 |
10 270 | --- | 12 709 | 10 150 | --- | 12 382 | 3A2(F): 3B1(F) → 3A2(F), (4) |
14 940 | 18 462 | 16 354 | 14 460 | 17 529 | 16 112 | 3T1(P): 3B1(F) → 3B1(P), (5) |
18 210 | --- | 20 607 | 17 500 | --- | 19 514 | 3T1(P): 3B1(F) → 3B2(P), (6) |
20 130 | 17 166 | 21 586 | 19 410 | 16 834 | 19 561 | 3T1(P): 3B1(F) → 3A2(P), (7) |
Experimental energies for the 3B1(F) → A2(F) band were taken from the VCD spectroscopy.
Numbered states correspond to spin-allowed ligand field transitions observed in experiments.
In addition to correlating TDDFT calculations to experimental precatalyst spectra, they can be further utilized to understand the equilibria discussed in Section 2.4 (see Supporting Information Section S8.1). The calculations corroborate that precatalysts can exist in complex equilibria featuring four-coordinate NiII(IB)X2, five-coordinate NiII(IB)(solv.)X2, or trimeric [NiII(IB)X2]3 μ-X species. Consistent with experiment, the calculated spectra for five-coordinate species and trimeric species exhibit a significant blue shift for the most intense calculated ligand field band (band 7; Figures S130–S132).
3.2. Ab initio Multireference Calculations of the NiII(IB)Cl2 and NiII(IB)Br2 Precatalysts
Due to the inherent complications with TDDFT described above, we have also used ab initio multireference calculations to compute and assign the experimentally observed ligand field excitations. First, we have systematically probed the effects of active space variation in CASSCF calculations (see Computational Details in Section S8 of the Supporting Information) on the qualitative agreement of NiII(IB)Cl2 and NiII(IB)Br2 with experimental spectra. Regardless of active space size (Tables S24–S39), the ground state is exclusively high spin. Notably, the largest active space used in this work, 22e,12o (five Ni 3d orbitals, six halide 2p/3p orbitals, and the Ni(IB) σ bonding orbital), results in a single-reference ground-state solution, with the highest weight of a single configuration in the CI vector of ~95% for both NiII(IB)Cl2 and NiII(IB)Br2 (Tables S29 and S35). This configuration corresponds to an S = 1 triplet ground state with unpaired electrons in the d(x2-y2) and d(xz) orbitals (i.e., the same configuration as obtained from DFT calculations; cf. Figure 7). We note that inclusion of the occupied halide 2p/3p orbitals and the Ni(IB) σ bonding orbital in the active space was essential to reproduce the experimental spectra. With this optimized active space, UV-vis-NIR absorption, CD, and MCD spectra of NiII(IB)Cl2 and NiII(IB)Br2 were calculated (Figure 8 and S141–S142). These calculations generally support assignments of experimental data given in Section 2.1 (Table 4). Individual states can be assigned based on the configuration state function with the largest weight in the CI vector, in conjunction with the location of the 3d holes (see Supporting Information Section S8.2).
Figure 7.
Comparison between 3d β-orbital manifolds of NiII(IB)X2 precatalysts demonstrating the destabilization of the β-LUMOs due to differences in Cl vs. Br ligand field strength.
Figure 8.
Comparison between calculated (CASPT2, 22e,12o; blue) and experimental (black) spectra. (A) UV-vis-NIR absorption spectra of NiII(IB)Cl2 (top) and NiII(IB)Br2 (bottom). Group theory assignments are given along with bands numbered as in experiment. The state ordering is the same for NiII(IB)Br2. (B) UV-vis-NIR absorption, CD, and MCD spectra of NiII(IB)Cl2. Note absorption and CD spectra are not normalized, while the computed MCD spectrum has normalized intensities to match experiment. Corresponding data for NiII(IB)Br2 are presented in Figure S142 in the Supporting Information.
Calculated signs for CD and MCD transitions are also consistent with experiment. All 3A2 excited states exhibit experimental and calculated positive CD bands and negative MCD bands. The 3A1 excited state (band 3 in experiment) exhibits negative bands for both CD and MCD. As supported by theory, the 3B1 excited state exhibits a negative CD band and a positive MCD band. In contrast, the predicted sign for CD does not match that observed experimentally for the higher-energy 3B2 excited state, though this is likely due to this transition being formally forbidden and gaining intensity through either spin-orbit or vibronic coupling, which can give either positive or negative differential intensity in CD.46 The overall success of the CASSCF/CASPT2 approach in calculating UV-vis-NIR, and in particular CD/MCD spectra, is especially encouraging for future analyses of experimental spectra for other chiral first row transition metal cross-coupling catalysts.
3.3. Computed Electrochemical Properties
Since the CASSCF CI vector indicates that the ground-state solutions are single-referent, we can use DFT to interpret the effects of different halide ligands on the NiIII/II, NiII/I, and NiI/0 reduction potentials of the NiII(IB)X2 complexes and connect them to thermodynamically accessible redox pathways of Ni-based reductive cross-coupling catalysis. Table 5 provides computed potentials for various electrochemical processes (vide supra, Section 2.5). Experimentally, the reduction of NiII(IB)X2 complexes is chemically irreversible due to halide dissociation. DFT calculations predict slightly positive halide dissociation energies of ΔGdissoc(CPCM) of ~6 kcal mol−1 and ~7 kcal mol−1 for Cl and Br NiI(IB)X2 complexes, respectively. The positive ΔGdissoc may be associated with inaccuracies in halide solvation energy when using a simple CPCM. The high electrolyte concentration in electrochemical experiments, which is not accounted for in the computations, may also further shift the equilibrium toward dissociation. For comparison, the calculated ΔGdissoc of halide loss from NiII(IB)X2 is significantly higher in energy (ΔGdissoc(CPCM) = ~30 kcal mol−1 and ~29 kcal mol−1 for Cl and Br, respectively), indicating much stronger ligand–metal bonds for the NiII species.
Table 5.
Computed formal potentials for NiII(IB)Cl2 and NiII(IB)Br2. All potentials referenced to Fc+/0 with computed absolute potential of –4.55 eV to compare the effect of solvent coordination without the shift induced by Fc+/0 in different solvents. Note that MeCN and DMA are essentially indistinguishable in the CPCM and produce the same computed value of Fc+/0 absolute potential.
Coordinated Solvent | Reaction | V vs. Fc+/0 |
---|---|---|
None | NiII(IB)Cl2 + e− → NiI(IB)Cl2 | −1.24 |
NiII(IB)Cl + e− → NiI(IB)Cl | −0.20 | |
NiII(IB)Cl2 + e− → NiI(IB)Cl + Cl− | −1.49 | |
NiIINiI(IB)2Cl4 Dimer + e− → [NiI(IB)Cl2]2 Dimer | −1.28 | |
NiII(IB)Br2 + e− → NiI(IB)Br2 | −1.11 | |
NiII(IB)Br + e− → NiI(IB)Br | −0.13 | |
NiII(IB)Br2 + e− → NiI(IB)Br + Br− | −1.39 | |
NiIINiI(IB)2Br4 Dimer + e− → [NiI(IB)Br2]2 Dimer | −1.26 | |
MeCN | NiII(IB)Cl(MeCN) + e− → NiI(IB)Cl(MeCN) | −0.59 |
NiII(IB)Cl2(MeCN) + e− → NiI(IB)Cl(MeCN) + Cl− | −1.65 | |
NiII(IB)Cl2(MeCN) + e− → NiI(IB)Cl2 + MeCN | −1.03 | |
NiII(IB)Br(MeCN) + e− → NiI(IB)Br(MeCN) | −0.90 | |
NiII(IB)Br2(MeCN) + e− → NiI(IB)Br(MeCN) + Br− | −1.48 | |
NiII(IB)Br2(MeCN) + e− → NiI(IB)Br2 + MeCN | −0.85 | |
DMA | NiII(IB)Cl(DMA) + e− → NiI(IB)Cl(DMA) | −1.17 |
NiII(IB)Cl2(DMA) + e− → NiI(IB)Cl(DMA) + Cl− | −1.66 | |
NiII(IB)Cl2(DMA) + e− → NiI(IB)Cl2 + DMA | −1.06 | |
NiII(IB)Br(DMA) + e− → NiI(IB)Br(DMA) | −1.09 | |
NiII(IB)Br2(DMA) + e− → NiI(IB)Br(DMA) + Br− | −1.44 | |
NiII(IB)Br2(DMA) + e− → NiI(IB)Br2 + DMA | −0.86 |
The computed one-electron reduction potentials coupled to halide loss are −1.49 V and −1.39 V vs. Fc+/0 for NiII(IB)Cl2 and NiII(IB)Br2, respectively, with no additional coordinated solvent ligand (Table 5). The absolute calculated values compare well with those measured experimentally (Table 3), as does the calculated potential difference between the two complexes (Δ = ~0.10 V (~805 cm−1) (calculated) vs. Δ = ~0.24 V (~1960 cm−1) (experiment – average for all three solvents)). Thus, DFT calculations support the idea that stronger ligand fields generally lead to a more negative potential for NiI formation. In that regard, more facile reduction of NiII(IB)Br2 is attributed to the lower energy of the β-LUMOs by ~0.2 eV (Figure 7).
The computed one-electron reduction potentials coupled to halide loss also shift negatively upon solvent coordination. For the Cl complex, calculated values are −1.49 V, −1.65 V, and −1.66 V vs. Fc+/0 for no solvent, MeCN, and DMA coordination, respectively. For the Br complex, analogous calculated values are −1.39 V, −1.48 V, and −1.44 V vs. Fc+/0. These negative potential shifts due to solvent coordination are also ascribed to modifications of the NiII β-LUMOs. For example, average energy destabilizations of 0.40 eV and 0.47 eV are observed for NiII(IB)Cl2 and NiII(IB)Br2 complexes, respectively, upon MeCN or DMA coordination. Together, these calculations and the ligand field spectroscopy both indicate for the first time that the RAMO energy is an excellent descriptor of the redox properties of these metal-based cross-coupling catalysts, and DFT calculations provide a useful approach for analyzing ligand contributions to potentials.
Finally, further extending these calculations to additional species, the computed reduction of NiI(IB)X to Ni0(IB)X or Ni0(IB) is calculated to occur at exceedingly negative formal potentials (~−3 V to −4 V). These calculated potentials for Ni0 formation are significantly more negative than the reduction potentials of typical reductants used in cross-coupling reactions (e.g., ~−1.11 V vs. Fc+/0 in DMA for tetrakis(dimethylamino)ethylene (TDAE) or ~−1.94 V vs Fc+/0 in DMF for Mn0).47,48 Similarly, the disproportionation of NiI(IB)X to NiII(IB)X2 and Ni0(IB) is thermodynamically highly unfavorable, with a ΔG(CPCM) of ~+57 kcal mol−1 for both X = Cl and Br. Thus, Ni0 is unlikely to be a catalytically relevant redox state. Bis(1,5-cyclooctadiene)nickel0 (Ni(COD)2), when used as the nickel source, likely first generates NiII(IB)X2 to commence the catalytic cycle. It is possible that Ni0(IB) can be formed by mixing IB and Ni(COD)2, and that this Ni0 complex might oxidatively add one of the electrophiles to give NiII(IB)RX and initiate the first turnover. However, it is unlikely that the catalytic cycle involves reduction back to Ni0(IB).
3.4. Resting State Electronic Structure and its Contributions to Reactivity
The electronic structure and electrochemical properties determined herein for NiII(IB)X2 complexes support a NiI/III cross-coupling reaction cycle. A low-spin square planar NiII phenanthroline (phen) aryl halide (NiII(phen)(Ar)X) resting state has been proposed by Diao et al. for Ni-catalyzed reductive 1,2–dicarbofunctionalization of alkenes.49 This resting state is thought to form from NiI(phen)X bimolecular oxidative addition of the aryl halide. The NiII(phen)(Ar)X is anticipated to react rapidly with C(sp3) radicals to produce a five-coordinate NiIII(phen)(Ar)X intermediate.31 The ‘R−Ar’ cross-coupled product can then form via reductive elimination. The analogous reactivity considerations for IB complexes and vinyl halides are given in Figure 9.44
Figure 9.
RAMO of the resting NiII(IB)(Vn)X for the NiII/III redox couple and radical capture with Vn = 1-methyl-4-vinylbenzene and X = Br.
Given the ability to accurately calculate and reproduce experimental formal potentials, the same approach can be used to investigate the bonding and redox properties of the NiII(IB)(Vn)Br (Vn = 1-methyl-4-vinylbenzene; Figure 9) resting state, which we were unable to isolate and study electrochemically due to instability.44 The calculated reduction potential of this species coupled to halide loss is −1.82 V vs. Fc+/0 in DMA. Interestingly, this is ~0.4 V more negative than that calculated for reduction of NiII(IB)Br2. The less facile reduction of NiII(IB)(Vn)Br relative to NiII(IB)Br2 likely stems from the increased donor strength of the Vn co-ligand (relative to Br). In the low-spin state of NiII(IB)(Vn)Br, the NiII d(x2-y2) orbital is the RAMO for the NiII/I couple. Strong antibonding interactions within the RAMO will destabilize the energy of this orbital, which will decrease the electron affinity and, thus, result in a more negative reduction potential. In this respect, it is also noted that, across a series of various para-substituted styrenyl derivatives, the calculated NiII/I reduction potentials of the NiII(IB)(Vn)Br species (coupled to halide dissociation) trend with the Hammett parameters of the substituent and the NiII character in the d(x2-y2) RAMO obtained from Löwdin population analysis (Figure S90).
It is also instructive to consider the role of the Vn ligand in the benzyl-C(sp3) radical recombination with NiII(IB)(Vn)X. The high donation of the Vn co-ligand in NiII(IB)(Vn)X would make this intermediate more oxidizable and readily available for the less-exergonic radical coupling; this variation in donation should also be correlated to changes in NiII–ligand bond covalency. Indeed, the calculated ligand–metal covalency of NiII(IB)(Vn)Br varies linearly with the calculated energy of the RAMO (Figure 10A). NiII(IB)(Vn)Br styrenyl systems with either dimethylamino or carbomethoxy substituents at the four position feature ~62% 3d character vs. ~10% 3d character in the RAMO, which correlate to more negative and less negative RAMO energies, respectively. Previous studies have also observed a linear correlation between redox potential and covalency in metalloproteins and model complexes.50–53
Figure 10.
Electronic structure and reactivity considerations for precatalysts and resting states. (A) Correlation between the total 3d orbital character in the RAMO of the NiII(IB)(Vn)Br and the RAMO energy. (B) Comparison between the NiII/III(IB)(Vn)Br oxidation potentials vs. reaction yields of the Vn/ethylbenzene C−C coupling with benzyl-C(sp3) radicals generated by reduction of NHP esters. Blue line indicates the oxidation potentials of NiII/III(IB)X2 precatalysts to demonstrate insufficient reactivity to compete with NiII(IB)(Vn)Br to quench benzyl-C(sp3) radicals. (C) Reaction yields of various Vn/benzyl-C(sp3) C−C coupling reactions plotted against the difference between computed oxidation potentials of NiII/III(IB)(Vn)Br and benzyl-C(sp3) radicals. Note all values are also tabulated in Table S8. Reaction yields plotted from Reference 54.
The less-donating halide ligands (relative to the Vn ligand) could electronically disfavor trapping of NiII(IB)X2 with benzylic radicals; this interpretation is supported by DFT calculations with NiII/III(IB)X2 oxidation potentials that are >1 V negative of oxidation potentials of NiII/III(IB)(Vn)Br (Figure 10B). The electronic influence of the Vn co-ligand can thus be envisioned as enhancing the kinetically competent radical recombination relative to deleterious side reactivity of the free benzylic radical. Indeed, there is a reasonable match between the calculated oxidation energies of the benzylic radicals and NiII(IB)(Vn)Br species, leading to favorable (near-ergoneutral) radical recombination reactions (Table S9). Additionally, there is a direct correlation between calculated oxidation potential gaps (ΔEox between the benzyl-C(sp3) radical and NiII(IB)(Vn)Br species) and reaction yields obtained from coupling the benzylic radicals generated from the NHP esters (NHP = N-hydroxyphthalimide),54 with the highest yields observed for reactions between benzyl-C(sp3) radicals and NiII(IB)(Vn)Br species with calculated oxidation potential gaps (ΔEox) of ~−0.4 V (Figure 10C). Although care must be taken drawing mechanistic conclusions from reaction yields, given that 1) NiII(IB)(Vn)X is the catalyst resting state44 (vide infra), and 2) Ni is not involved in generating the radical from the NHP ester, the correlation between reaction yield and ΔEox reflects the favorability of the radical addition to NiII(IB)(Vn)Br.
4. Discussion
Due to their disparate electronic structures relative to precious metal analogs, first-row transition metal catalysts unlock a rich area of exploration for the discovery of new organic reactions. Indeed, many novel bond constructions have been discovered recently.2 Exciting mechanistic studies have also highlighted a variety of inorganic species of relevance to reactivity, many of which feature interesting metal- and ligand-based redox events. Furthermore, ground state cross-coupling reaction mechanisms of first-row transition metal complexes have direct relevance to the thermal components of photoredox catalysis. Unlike their bipyridine and terpyridine analogues, which feature intense MLCT transitions, non-conjugated ligands (e.g., IB) exhibit rich ligand field spectral features that can be utilized to interrogate Ni-based electronic structure, and just as transient spectroscopic methods can provide insights into the light-induced components of photoredox mechanisms,55 steady state spectroscopic methods can also provide direct insights into electronic structure contributions to transition metal reactivity.12,56,57
Here we have provided several new insights into the area of cross-coupling catalysis: 1) a combination of electronic and magneto-optical spectroscopic methods has allowed for a detailed analysis and definition of the ligand field excited state manifolds of two previously employed cross-coupling catalysts, NiII(IB)Cl2 and NiII(IB)Br2. This analysis has allowed for the quantification of relative ligand field strengths. As discussed further below, we demonstrate how ligand field strength is linked directly to changes in the energy of the Ni-based RAMO and, thus, to reduction potentials. 2) VT UV-vis-NIR spectroscopy has uncovered and identified two equilibrium processes that we believe to be significant for cross-coupling reaction mechanisms and yields. These equilibria feature a) four-coordinate and five-coordinate NiII complexes when in DMA, and b) a monomeric and trimeric species in the Cl complex when in DCM. 3) The Vn ligand involved in cross-coupling plays a critical role in directing the reactivity of the putative resting NiII complex. Specifically, the electronic nature of the Vn ligand tunes the Ni-based covalency over a large range. In doing so, the covalent character modulates the reactivity of this species toward C(sp3) radicals, and we have observed a direct correlation between previously reported reaction yields58 and the difference in oxidation potential between the NiII resting state and the C(sp3) radicals, which reflects the driving force for the radical recombination reaction between these species.
Averaging all assignable, spin-allowed ligand field excited states, it was demonstrated in Section 2.1 that the ligand field strength of the Cl complex is ~390 cm−1 stronger than the Br complex. In Section 2.5, it was shown that one-electron reduction of NiII(IB)X2 complexes is coupled to rapid halide dissociation, and it is on average (over three solvents) ~0.24 V (~1960 cm−1) harder to reduce the Cl complex relative to the Br complex. In Section 3.3, it was also demonstrated that electronic structure calculations can accurately reproduce the experimental one-electron reduction potentials and that the origin of the relative reduction potential arises from ligand field (de)stabilization of the NiII-based RAMO. Thus, combining ligand field spectroscopies with electrochemistry and electronic structure calculations provides a direct means to dissect geometric and electronic structure contributions to catalytically-relevant redox potentials and allows for an experimentally calibrated approach to extend computations to reaction intermediates.
While the ligand perturbations studied here are halide-based, the ligand field spectral features of Ni-catalysts will also be highly sensitive to ligand perturbations. For example, in Table 6, we have compiled electronic transitions for the IB complexes and a series of previously reported pseudo-Td NiII complexes, with structures shown in Figure 11.59–63 In general, the second 3B1 → 3A2(F) transition, which is the most intense observed in the lower-energy ligand field manifold, tracks roughly with the overall ligand field strength, with NiII(PPh3)2Cl2 exhibiting the highest transition energy. Another noticeable trend, the variability in 3B1/3B2 transition energies, stems from high variability in geometric distortions. We also note the similarity in blueshifts in transition energies for NiII(IB)X2 and NiII(L+)X3 as a function of halide identity. Thus, measured ligand field transitions can provide a means to estimate specific ligand contributions to catalyst electronic structure and ultimately redox potentials and reactivity in future studies.
Table 6.
Comparison between spin-allowed transition energies for NiII(IB)Cl2 and NiII(IB)Br2 in DCM and previously studied pseudo-Td NiII complexes (units of cm−1).
Complex | 3B1(F) → 3A2(F) | 3B1(F) → 3B1/3B2 (F) | 3B1(F) → 3A1(F) | 3B1(F) → 3A2(F) | 3B1(F) → 3B1(P) | 3B1(F) → 3B2(P) | 3B1(F) → 3A2(P) |
---|---|---|---|---|---|---|---|
Ni II (IB)Cl 2 a | 2170 | 6640 | 8240 | 10 270 | 14 940 | 18 210 | 20 130 |
Ni II (IB)Br 2 a | 2210 | 6140 | 7970 | 10 140 | 14 570 | 17 420 | 19 380 |
Ni II (PPh 3 ) 2 Cl 2 b | 4500 | 10 200 | 8000 | 11 200 | 18 100 | 14 600 – 16 400* | 17 600 |
Ni II (biquinoline)Br 2 c | 3740 | 4941 | 6020 | 9810 | 18 450 | 17 300 | 19 430 |
[Ni II Cl 4 ] 2- d | 4440 | 3970 | 4440 | 6870 | 14 250 | 14 250 | 15 240 |
Ni II (L+)Cl 3 e | 4660* | 5260 | 6410 | 9090 | 16 000 | 16 000 | 17 240 |
Ni II (L+)Br 3 e | 4530 | 5100 | 6410 | 8700 | 15 200 – 16 100 | 15 200 – 16 100 | 16 250 |
Figure 11.
Structures of pseudo-Td NiII complexes used for comparison in Table 6.
VT electronic absorption spectra reported here are also particularly illuminating and provide 1) direct evidence of variable speciation for IB complexes as a function of solvent and temperature and 2) direct thermodynamic parameters describing these equilibria. For NiII(IB)Cl2 and NiII(IB)Br2 in DMA, a significant fraction of DMA coordinated complex is present at room temperature (~37% and 49%, respectively, at 3.6 mM (Cl) and 3.5 mM (Br)). At lower temperatures (273 K), DMA coordination is strongly favored (~53% and 69%, respectively). Generally, amide-based solvents can coordinate to metal centers through the carbonyl oxygen, preference for which is further influenced by steric hindrance at the nitrogen center. The nN → π*CO delocalization of the amide results in increased electron density on oxygen (Figure 5), leading to a strong donor and a partial negative charge near the NiII center.69 As demonstrated herein for the first time by a combination of electrochemistry and electronic structure calculations, DMA coordination can lead to a more difficult to reduce NiII center, with negatively shifted potentials of ~200 mV in DMA vs. MeCN for the Cl and Br complexes. The single electron reduction leads to rupture of the halide–metal bond; we propose that halide solvation and solvent coordination to Ni can also contribute significantly to the measured reduction potential (e.g., –1.47 V vs. Fc+/0 in DMA/DCM for NiII(IB)Cl2). Of further note, yields for reductive alkenylation are maximized when 1) DMA or another amide solvent is utilized, and 2) the reaction temperature is lowered to 0–5 °C, with both considerations favoring nickel coordination by the solvent.
Scheme 2 provides boron trifluoride (BF3) donor numbers, acceptor numbers, substrate conversion percentage, and cross-coupled yield percentages for the reaction between (1-chloroethyl)benzene and (E)-1-(2-bromovinyl)-4-methoxybenzene. Previous studies determined the optimized conditions based on systematic screening of reaction parameters.7 We conducted triplicate measurements of this reaction in several solvents with a range of Lewis basicities. Generally, as the donicity of the solvent decreases, the substrate consumption and cross-coupled product yield decrease significantly. As discussed further below, we believe this reflects a significant steric effect at play, which can protect NiII and/or reduced NiI species from off-cycle reactivity. Based on these considerations, we also purport that amide coordination results in a more oxidizable NiI species, which increases oxidative addition reactivity.
Scheme 2.
Conversion and cross-coupled yields for a representative reductive cross-coupling of (1-chloroethyl)benzene and (E)-1-(2-bromovinyl)-4-methoxybenzene using NiII(IB)Cl2, along with donor and acceptor numbers for solvents tested in optimization studies. Standard errors were calculated from three trials. Experimental procedure given in Supporting Information Section S1. D-BF3 donor numbers taken from References 32 and 33. Acceptor numbers (ANs) taken from Reference 33. All ANs measured with respect to the affinity of antimony pentachloride towards triethylphosphine oxide.
The VT UV-vis-NIR and spectroelectrochemistry studies, when compared with experimental conditions often utilized for nickel-catalyzed reductive cross-coupling (i.e., low temperatures, use of coordinating solvents, relatively mild reducing agents, etc.),2,7,8 are consistent with the mechanistic investigation of this reaction by Turro and coworkers,44 which suggests this reaction proceeds via a NiI/III redox cycle (Scheme 3), similar to that proposed by Diao and coworkers for the related cross-coupling of aryl halides.20 Additional aspects of the optimized reaction conditions can be commented on, however, as derived from the findings in the present study.
Scheme 3.
Representative computed NiI/III catalytic cycle for the cross-coupling of C(sp2) vinyl and C(sp3) benzyl electrophiles. Relative Gibbs free energy values were computed at the TPSSh (CPCM) level for X = Br. Reactions involving electron transfers were corrected by adding the energy of Mn0/2+ redox couple or TDAE0/+, which were used as the external reducing agents in refs. 54 and 7. Note that we have adopted values of E° = −1.94 V vs. Fc+/0 for Mn2+/0 obtained from the experiments in refs. 48 and 70 and E° = −1.34 V vs. Fc+/0 for TDAE+/0 from computations in this article.
To better understand their role in the cross-coupling of (E)-1-(2-bromovinyl)-4-methylbenzene (Vn C(sp2) component) and (1-bromoethyl)benzene (C(sp3) radical), DFT computed stable intermediates along the nickel catalytic cycle are shown in Scheme 3 for NiII(IB)Br2. Free energies of individual steps are energetically accessible, suggesting favorable cross-coupling reactivity. Note that the C–C bond forming cycle is identical regardless of the C(sp3) benzyl radical source. However, two distinct pathways to generating these radicals have been proposed by Turro et al., which differ in terms of substrate (i.e., benzyl halide, pathway A, or NHP ester, pathway B) and reductant (i.e., Mn0 vs. TDAE reductants) (Scheme 3).44 We also note Scheme 3 only features computed free energies, and we have restricted our discussion of the possible effects of DMA to these steps/energies. It is also possible for DMA coordination to influence reaction barriers. More detailed mechanistic aspects and reaction kinetics are discussed in reference 44 by Turro et al.. More broadly, results described herein suggest future computational studies should also consider explicit solvent coordination in calculated reaction mechanisms.
With benzyl bromide and a Mn0 reductant, the only computed endergonic reaction in the catalytic cycle is the radical recombination of the C(sp3) radical with NiII(IB)(Vn)Br to form NiIII(IB)(Vn)(R)Br (Scheme 3, reaction 3, ΔG0(CPCM) = 4.6 kcal/mol). Notably, NiII(IB)(Vn)Br and related species have been assigned as the catalytic resting state.49,20,44 The generation of C(sp3) benzyl radical from benzyl bromide is a favorable process through Br-atom abstraction (reaction ii.A) with either NiI(IB)Vn (ΔG0(CPCM) = −19.1 kcal mol−1) or NiI(IB)Br (−9.3 kcal mol−1), thus suggesting there might be a balance between this reactivity pathway and reactions i.A and ii.A in Scheme 3. The detailed kinetic analysis presented in ref. 44, however, suggests that XAT from NiII(IB)X is a more likely alternative, as also evidenced by the observed reactivity of NiI(IB)X with both C(sp2) vinyl- and C(sp3) benzyl-halides.
With the NHP ester substrate and TDAE as a reductant, the NiI(IB)Br formation from NiII(IB)Br2 becomes slightly disfavored with ΔG0(CPCM) = 1.1 kcal/mol, in addition to the only endergonic reaction 3 in pathway A. In this scenario, the C(sp3) benzyl radical can be generated from NHP ester without Ni involvement through a direct reduction with TDAE (reaction i.B, E1/2 = −1.39 V vs. Fc+/0)) followed by a highly exergonic decomposition, yielding CO2, phthalimide anion, and C(sp3) benzyl radical (pathway i.B).
As discussed, higher yields of cross-coupled product are typically obtained at lower temperatures and in coordinating solvents, supporting solvent coordination as a key component for the catalytic cycle.2,7,8 The computed free energies indeed suggest an equilibrium might exist between the NiI(IB)Br and the [NiI(IB)Br]2 μ-Br dimeric species (ΔG0(CPCM) = +4.1 kcal/mol). We speculate low temperature, which entropically favors solvent coordination, might prevent highly reactive NiI intermediates from entering off-cycle pathways, especially toward forming oligomeric (e.g., dimeric/trimeric) species. The same could be true for NiII species. Coordinating solvents can similarly protect NiI from dimerization by forming a more favorable equilibrium between NiI(IB)Br ↔ NiI(IB)(S)Br, where S = solvent, since formation of the dimer from NiI(IB)(S)Br intermediate may be disfavored sterically. Although the ΔG0 for DMA coordination to NiI(IB)Br is computed to be higher than ΔG0 for dimerization (~6.9 kcal mol−1 vs. ~4.1 kcal mol−1), the large excess of DMA (especially relative to the rather low concentration of NiI(IB)Br under catalytic conditions) will favor solvent coordination and further perturb the relative concentrations/equilibria for both processes.
DMA coordination to NiI(IB)X could also lower the kinetic barrier for bimolecular oxidative addition reactivity in reaction 2 (Scheme 3): 1) the NiI(IB)X oxidation to NiII(IB)X2 is favored in DMA, as observed by the negative shift in reduction potential in the experimental cyclic voltammetry, and 2) oxidative addition of vinyl-halide to NiI(IB)X forming NiII(IB)(Vn)X is favored due to increased electron density on the Ni center and destabilized frontier molecular orbitals upon DMA coordination.
In addition to insights into the role of solvent coordination, the Vn ligand itself can strongly contribute to the reactivity of the resting state. Indeed, there is a linear trend between the amount of metal character in the RAMO and its energy (Figure 10A). In general, ligand contributions to the destabilization of the energy of the RAMO will result in greater reactivity and easier oxidation. Notably, we observed a correlation between the reaction yields and the difference in oxidation potentials of NiII(IB)(Vn)Br and benzyl-C(sp3) radical (ΔEox). The same metric was recently proposed by Okamoto et al. as a measure of the probability of intramolecular redox reactions, showing the most favorable radical recombination for ΔEox ~0 V.71 In our case, the coupling efficiency and resulting yields maximize when the benzyl-C(sp3) radical oxidation potential is slightly higher (by ~0.4 V) than the oxidation potential of NiII(IB)(Vn)Br. A higher reactivity of the benzyl-C(sp3) radical might be necessary to overcome the reorganization energy required for the change in coordination geometry of the Ni catalyst upon formation of a five-coordinate NiIII(IB)(R)(Vn)Br intermediate. Lower yields at higher oxidation potentials of the benzyl-C(sp3) radicals could be similarly justified by accessing side-reaction radical pathways (e.g., homocoupling). The calculated mismatch in the oxidation potentials may also be a computational artifact. Although we have demonstrated a reasonable correlation between computed and experimental electrochemical data, the nature of Ni catalysts and small organic radicals is quite different and could result in varied relative accuracy of the computed oxidation potentials. The computational model also neglects explicit solvation, whereas DMA solvation and/or coordination to the Ni catalyst (as demonstrated in the present study) could shift the observed oxidation potentials. Regardless, there is a clear trend between the calculated ΔEox and reaction yields, which indicates the relative oxidation potentials and frontier molecular orbitals of the NiII(IB)(Vn)Br resting state and the benzyl-C(sp3) radicals may be a useful predictor for reactivity and can be utilized to computationally screen and guide the synthesis of new Ni complexes for reductive cross-coupling catalysis.
Conclusions
We have presented the first comprehensive spectroscopic, electrochemical, and computational study of NiII-based chiral reductive cross-coupling catalysts, NiII(IB)Cl2 and NiII(IB)Br2. This combination of methods has established 1) there is a direct connection between ligand field strength and catalytically relevant Ni-based reduction potentials, 2) there exists variable speciation for IB complexes as a function of solvent and temperature, and VT spectroscopy provides a means to directly measure thermodynamic parameters describing these equilibria, and 3) the Vn ligand involved in cross-coupling plays a critical role in directing the reactivity of the putative resting NiII complex through strong modulation of ligand–metal covalency. These experimental findings strongly support a critical steric and electronic role that amide-based solvents play in the cross-coupling catalytic cycle. Through this electronic structural characterization, we suggest that dimerization or oligomerization of the catalyst is inhibited by solvent coordination. Furthermore, the modification of electronic structure through solvent coordination increases the driving force for oxidative addition, as evidenced by formal potentials from voltammetry. These results highlight the importance of reaction conditions on catalytic efficiencies and will guide future methodology studies, as well as the development of new ligand scaffolds for Ni-based reductive cross-coupling catalysis.
Supplementary Material
Acknowledgment
We acknowledge the X-ray Crystallography Facility in the Beckman Institute at Caltech, and the Dow Next Generation Instrumentation Grant for X-ray structure collection. Some computations presented here were conducted in the Resnick High Performance Computing Center, a facility supported by Resnick Sustainability Institute at the California Institute of Technology. We are grateful for assistance from Alexander Q. Cusumano in collecting vibrational circular dichroism data and from Michael Zott in collecting variable temperature UV-vis-NIR spectra. We thank Harry B. Gray and Jay R. Winkler for helpful discussions on equilibrium processes in spectroscopy and David E. Hill and David A. Cagan for helpful general discussions. N.P.K. acknowledges support from the Hertz Fellowship and from the National Science Foundation Graduate Research Fellowship under Grant No. DGE-1745301. This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No. 883987 (D.B.). Support has been provided by the Grant Agency of the Czech Republic (20–06451Y to J.C.). Support has been provided by the National Institutes of Health (National Institute of General Medical Sciences, R35-GM142595) (R.G.H.). S.E.R. acknowledges financial support from the NIH (R35GM118191).
Footnotes
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: Materials and methods, NMR spectra, additional UV-vis-NIR, CD, and MCD spectra, additional voltammetry and spectroelectrochemistry data, X-ray crystallographic parameters, and DFT and CASSCF/CASPT2 input parameters and results.
References
- (1).Semmelhack MF; Helquist PM; Jones LD Synthesis with Zerovalent Nickel. Coupling of Aryl Halides with Bis(1,5-Cyclooctadiene)Nickel(0). J. Am. Chem. Soc. 1971, 93 (22), 5908–5910. 10.1021/ja00751a062. [DOI] [Google Scholar]
- (2).Poremba KE; Dibrell SE; Reisman SE Nickel-Catalyzed Enantioselective Reductive Cross-Coupling Reactions. ACS Catal. 2020, 10 (15), 8237–8246. 10.1021/acscatal.0c01842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Durandetti M; Nédélec J-Y; Périchon J Nickel-Catalyzed Direct Electrochemical Cross-Coupling between Aryl Halides and Activated Alkyl Halides. J. Org. Chem. 1996, 61 (5), 1748–1755. 10.1021/jo9518314. [DOI] [PubMed] [Google Scholar]
- (4).Durandetti M; Gosmini C; Périchon J Ni-Catalyzed Activation of α-Chloroesters: A Simple Method for the Synthesis of α-Arylesters and β-Hydroxyesters. Tetrahedron 2007, 63 (5), 1146–1153. 10.1016/j.tet.2006.11.055. [DOI] [Google Scholar]
- (5).Cherney AH; Kadunce NT; Reisman SE Catalytic Asymmetric Reductive Acyl Cross-Coupling: Synthesis of Enantioenriched Acyclic α,α-Disubstituted Ketones. J. Am. Chem. Soc. 2013, 135 (20), 7442–7445. 10.1021/ja402922w. [DOI] [PubMed] [Google Scholar]
- (6).Hofstra JL; Cherney AH; Ordner CM; Reisman SE Synthesis of Enantioenriched Allylic Silanes via Nickel-Catalyzed Reductive Cross-Coupling. J. Am. Chem. Soc. 2018, 140 (1), 139–142. 10.1021/jacs.7b11707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Cherney AH; Reisman SE Nickel-Catalyzed Asymmetric Reductive Cross-Coupling Between Vinyl and Benzyl Electrophiles. J. Am. Chem. Soc. 2014, 136 (41), 14365–14368. 10.1021/ja508067c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).DeLano TJ; Reisman SE Enantioselective Electroreductive Coupling of Alkenyl and Benzyl Halides via Nickel Catalysis. ACS Catal. 2019, 9 (8), 6751–6754. 10.1021/acscatal.9b01785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Cagan DA; Bím D; Silva B; Kazmierczak NP; McNicholas BJ; Hadt RG Elucidating the Mechanism of Excited-State Bond Homolysis in Nickel–Bipyridine Photoredox Catalysts. J. Am. Chem. Soc. 2022, 144 (14), 6516–6531. 10.1021/jacs.2c01356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Cagan DA; Bîm D; McNicholas BJ; Kazmierczak NP; Oyala PH; Hadt RG Photogenerated Ni(I)–Bipyridine Halide Complexes: Structure-Function Relationships for Competitive C(sp2)–Cl Oxidative Addition and Dimerization Reactivity Pathways. Inorg. Chem. 2023, 62, 9538–9551. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Gao Y; Hill DE; Hao W; McNicholas BJ; Vantourout JC; Hadt RG; Reisman SE; Blackmond DG; Baran PS Electrochemical Nozaki–Hiyama–Kishi Coupling: Scope, Applications, and Mechanism. J. Am. Chem. Soc. 2021, 143 (25), 9478–9488. 10.1021/jacs.1c03007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).Daifuku SL; Al-Afyouni MH; Snyder BER; Kneebone JL; Neidig ML A Combined Mössbauer, Magnetic Circular Dichroism, and Density Functional Theory Approach for Iron Cross-Coupling Catalysis: Electronic Structure, In Situ Formation, and Reactivity of Iron-Mesityl-Bisphosphines. J. Am. Chem. Soc. 2014, 136 (25), 9132–9143. 10.1021/ja503596m. [DOI] [PubMed] [Google Scholar]
- (13).Day CS; Somerville RJ; Martin R Deciphering the Dichotomy Exerted by Zn(Ii) in the Catalytic sp2 C–O Bond Functionalization of Aryl Esters at the Molecular Level. Nature Catalysis 2021, 4 (2), 124–133. 10.1038/s41929-020-00560-3. [DOI] [Google Scholar]
- (14).Jensen AE; Knochel P Nickel-Catalyzed Cross-Coupling between Functionalized Primary or Secondary Alkylzinc Halides and Primary Alkyl Halides. J. Org. Chem. 2002, 67 (1), 79–85. 10.1021/jo0105787. [DOI] [PubMed] [Google Scholar]
- (15).Dorval C; Gosmini C Low-Valent Cobalt Complexes in C–X Coupling and Related Reactions. In Cobalt Catalysis in Organic Synthesis; 2020; pp 163–205. 10.1002/9783527814855.ch5. [DOI] [Google Scholar]
- (16).Cahiez G; Marquais S Highly Chemo- and Stereoselective Fe-Catalyzed Alkenylation of Organomanganese Reagents. Tetrahedron Letters 1996, 37 (11), 1773–1776. 10.1016/0040-4039(96)00116-5. [DOI] [Google Scholar]
- (17).Cahiez G; Marquais S Copper-Catalyzed Alkylation of Organomanganese Chloride Reagents. Synlett 2002, 1993 (01), 45–47. [Google Scholar]
- (18).Cahiez G; Avedissian H Cobalt-Catalyzed Alkenylation of Organomagnesium Reagents. Tetrahedron Letters 1998, 39 (34), 6159–6162. 10.1016/S0040-4039(98)01266-0. [DOI] [Google Scholar]
- (19).Yanagi T; Somerville RJ; Nogi K; Martin R; Yorimitsu H Ni-Catalyzed Carboxylation of C(Sp2)–S Bonds with CO2: Evidence for the Multifaceted Role of Zn. ACS Catal. 2020, 10 (3), 2117–2123. 10.1021/acscatal.9b05141. [DOI] [Google Scholar]
- (20).Ju L; Lin Q; LiBretto NJ; Wagner CL; Hu CT; Miller JT; Diao T Reactivity of (Bi-Oxazoline)Organonickel Complexes and Revision of a Catalytic Mechanism. J. Am. Chem. Soc. 2021, 143 (36), 14458–14463. 10.1021/jacs.1c07139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).Xu J; Li Z; Xu Y; Shu X; Huo H Stereodivergent Synthesis of Both Z- and E-Alkenes by Photoinduced, Ni-Catalyzed Enantioselective C(sp3)–H Alkenylation. ACS Catal. 2021, 11 (21), 13567–13574. 10.1021/acscatal.1c04314. [DOI] [Google Scholar]
- (22).Lu Q; Guan H; Wang Y-E; Xiong D; Lin T; Xue F; Mao J Nickel/Photoredox-Catalyzed Enantioselective Reductive Cross-Coupling between Vinyl Bromides and Benzyl Chlorides. J. Org. Chem. 2022, 87 (12), 8048–8058. 10.1021/acs.joc.2c00707. [DOI] [PubMed] [Google Scholar]
- (23).Liu J; Gong H; Zhu S Nickel-Catalyzed, Regio- and Enantioselective Benzylic Alkenylation of Olefins with Alkenyl Bromide. Angewandte Chemie International Edition 2021, 60 (8), 4060–4064. 10.1002/anie.202012614. [DOI] [PubMed] [Google Scholar]
- (24).Zhu Z; Lin L; Xiao J; Shi Z Nickel-Catalyzed Stereo- and Enantioselective Cross-Coupling of Gem-Difluoroalkenes with Carbon Electrophiles by C−F Bond Activation. Angewandte Chemie International Edition 2022, 61 (6), e202113209. 10.1002/anie.202113209. [DOI] [PubMed] [Google Scholar]
- (25).Hu X; Cheng-Sánchez I; Cuesta-Galisteo S; Nevado C Nickel-Catalyzed Enantioselective Electrochemical Reductive Cross-Coupling of Aryl Aziridines with Alkenyl Bromides. J. Am. Chem. Soc. 2023. 10.1021/jacs.2c12869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Jiang Y; Yang K; Wei Y; Wang Q; Li S-J; Lan Y; Koh MJ Catalytic Multicomponent Synthesis of C-Acyl Glycosides by Consecutive Cross-Electrophile Couplings. Angewandte Chemie International Edition 2022, 61 (46), e202211043. 10.1002/anie.202211043. [DOI] [PubMed] [Google Scholar]
- (27).Geng J; Sun D; Song Y; Tong W; Wu F Ni-Catalyzed Asymmetric Reductive Alkenylation of α-Chlorosulfones with Vinyl Bromides. Org. Lett. 2022, 24 (9), 1807–1811. 10.1021/acs.orglett.2c00217. [DOI] [PubMed] [Google Scholar]
- (28).Duan M; Wang Y; Zhu S Nickel-Catalyzed Asymmetric 1,2-Alkynylboration of Vinylarenes. Tetrahedron Letters 2023, 114, 154247. 10.1016/j.tetlet.2022.154247. [DOI] [Google Scholar]
- (29).Sun D; Ma G; Zhao X; Lei C; Gong H Nickel-Catalyzed Asymmetric Reductive Arylation of α-Chlorosulfones with Aryl Halides. Chem. Sci. 2021, 12 (14), 5253–5258. 10.1039/D1SC00283J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Ye Y; Liu J; Xu B; Jiang S; Bai R; Li S; Xie T; Ye X-Y Nickel-Catalyzed Enantioselective 1,2-Vinylboration of Styrenes. Chem. Sci. 2021, 12 (39), 13209–13215. 10.1039/D1SC04071E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Lin Q; Spielvogel EH; Diao T Carbon-Centered Radical Capture at Nickel(II) Complexes: Spectroscopic Evidence, Rates, and Selectivity. Chem 2023, 9 (5), 1295–1308. 10.1016/j.chempr.2023.02.010. [DOI] [Google Scholar]
- (32).Laurence C; Gal JF Lewis Basicity and Affinity Scales: Data and Measurements; Wiley, New York, 2010. [Google Scholar]
- (33).Classification of Solvents. In Solvents and Solvent Effects in Organic Chemistry; 2010; pp 65–106. 10.1002/9783527632220.ch3. [DOI] [Google Scholar]
- (34).Best SP; Clark RJH The Identification of an Electronic Raman Transition for the Hexa-Aquavanadium(III) Ion. A Direct Spectroscopic Determination of the Trigonal Field Splitting of the 3T1g Ground Term. Chemical Physics Letters 1985, 122 (4), 401–405. 10.1016/0009-2614(85)80245-1. [DOI] [Google Scholar]
- (35).Pescitelli G; Lüdeke S; Chamayou A-C; Marolt M; Justus V; Górecki M; Arrico L; Di Bari L; Islam MA; Gruber I; Enamullah M; Janiak C Broad-Range Spectral Analysis for Chiral Metal Coordination Compounds: (Chiro)Optical Superspectrum of Cobalt(II) Complexes. Inorg. Chem. 2018, 57 (21), 13397–13408. 10.1021/acs.inorgchem.8b01932. [DOI] [PubMed] [Google Scholar]
- (36).Lu R; Yang T; Chen X; Fan W; Chen P; Lin Z; Liu G Enantioselective Copper-Catalyzed Radical Cyanation of Propargylic C–H Bonds: Easy Access to Chiral Allenyl Nitriles. J. Am. Chem. Soc. 2021, 143 (36), 14451–14457. 10.1021/jacs.1c07190. [DOI] [PubMed] [Google Scholar]
- (37).Hofstra JL Development and Mechanistic Studies of Ni-Catalyzed Asymmetric Reductive Cross-Coupling Reactions, California Institute of Technology, 2019. https://resolver.caltech.edu/CaltechTHESIS:06012019-034543690. [Google Scholar]
- (38).Kazmierczak NP; Chew JA; Vander Griend DA Bootstrap Methods for Quantifying the Uncertainty of Binding Constants in the Hard Modeling of Spectrophotometric Titration Data. Analytica Chimica Acta 2022, 1227, 339834. 10.1016/j.aca.2022.339834. [DOI] [PubMed] [Google Scholar]
- (39).Vander Griend DA; Bediako DK; DeVries MJ; DeJong NA; Heeringa LP Detailed Spectroscopic, Thermodynamic, and Kinetic Characterization of Nickel(II) Complexes with 2,2’-Bipyridine and 1,10-Phenanthroline Attained via Equilibrium-Restricted Factor Analysis. Inorg. Chem. 2008, 47 (2), 656–662. 10.1021/ic700553d. [DOI] [PubMed] [Google Scholar]
- (40).Espinoza EM; Clark JA; Soliman J; Derr JB; Morales M; Vullev VI Practical Aspects of Cyclic Voltammetry: How to Estimate Reduction Potentials When Irreversibility Prevails. J. Electrochem. Soc. 2019, 166 (5), H3175–H3187. 10.1149/2.0241905jes. [DOI] [Google Scholar]
- (41).Tang T; Jones E; Wild T; Hazra A; Minteer SD; Sigman MS Investigating Oxidative Addition Mechanisms of Allylic Electrophiles with Low-Valent Ni/Co Catalysts Using Electroanalytical and Data Science Techniques. J. Am. Chem. Soc. 2022, 144 (43), 20056–20066. 10.1021/jacs.2c09120. [DOI] [PubMed] [Google Scholar]
- (42).Bard AJ; Faulkner LR; White HS Electrochemical Methods: Fundamentals and Applications, 3rd ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2022. [Google Scholar]
- (43).Zhou Y-Y; Uyeda C Reductive Cyclopropanations Catalyzed by Dinuclear Nickel Complexes. Angewandte Chemie International Edition 2016, 55 (9), 3171–3175. 10.1002/anie.201511271. [DOI] [PubMed] [Google Scholar]
- (44).Turro RF; Wahlman JLH; Tong ZJ; Chen X; Yang M; Chen EP; Hong X; Hadt RG; Yang Y-F; Houk KN; Reisman SE Mechanistic Investigation of Ni-Catalyzed Reductive Cross-Coupling of Alkenyl and Benzyl Electrophiles. J. Am. Chem. Soc. 2023, 145 (27), 14705–14715. 10.1021/jacs.3c02649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (45).Li G; Brady MD; Meyer GJ Visible Light Driven Bromide Oxidation and Ligand Substitution Photochemistry of a Ru Diimine Complex. J. Am. Chem. Soc. 2018, 140 (16), 5447–5456. 10.1021/jacs.8b00944. [DOI] [PubMed] [Google Scholar]
- (46).Harding MJ; Mason SF; Robbins DJ; Thomson AJ Magnetic Circular Dichroism Spectra of Some Nickel(II) Complexes. Part I. The Temperature Dependence of the Spectrum of the Nickel(II) Hexaquo-Ion. J. Chem. Soc. A 1971, No. 0, 3047–3058. 10.1039/J19710003047. [DOI] [Google Scholar]
- (47).Burkholder C; Dolbier WR; Médebielle M Tetrakis(Dimethylamino)Ethylene as a Useful Reductant of Some Bromodifluoromethyl Heterocycles. Application to the Synthesis of New Gem-Difluorinated Heteroarylated Compounds. J. Org. Chem. 1998, 63 (16), 5385–5394. 10.1021/jo980201+. [DOI] [Google Scholar]
- (48).Gaur JN; Goswami NK Kinetics of the Reduction of Mn2+ at the Dropping Mercury Electrode in Non-Aqueous Media. Electrochimica Acta 1967, 12 (11), 1489–1493. 10.1016/0013-4686(67)80064-1. [DOI] [Google Scholar]
- (49).Lin Q; Diao T Mechanism of Ni-Catalyzed Reductive 1,2-Dicarbofunctionalization of Alkenes. J. Am. Chem. Soc. 2019, 141 (44), 17937–17948. 10.1021/jacs.9b10026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Anxolabéhère-Mallart E; Glaser T; Frank P; Aliverti A; Zanetti G; Hedman B; Hodgson KO; Solomon EI Sulfur K-Edge X-Ray Absorption Spectroscopy of 2Fe–2S Ferredoxin: Covalency of the Oxidized and Reduced 2Fe Forms and Comparison to Model Complexes. J. Am. Chem. Soc. 2001, 123 (23), 5444–5452. 10.1021/ja010472t. [DOI] [PubMed] [Google Scholar]
- (51).Dey A; Jenney FE; Adams MWW; Babini E; Takahashi Y; Fukuyama K; Hodgson KO; Hedman B; Solomon EI Solvent Tuning of Electrochemical Potentials in the Active Sites of HiPIP Versus Ferredoxin. Science 2007, 318 (5855), 1464–1468. 10.1126/science.1147753. [DOI] [PubMed] [Google Scholar]
- (52).Dey A; Okamura T; Ueyama N; Hedman B; Hodgson KO; Solomon EI Sulfur K-Edge XAS and DFT Calculations on P450 Model Complexes: Effects of Hydrogen Bonding on Electronic Structure and Redox Potentials. J. Am. Chem. Soc. 2005, 127 (34), 12046–12053. 10.1021/ja0519031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Hadt RG; Sun N; Marshall NM; Hodgson KO; Hedman B; Lu Y; Solomon EI Spectroscopic and DFT Studies of Second-Sphere Variants of the Type 1 Copper Site in Azurin: Covalent and Nonlocal Electrostatic Contributions to Reduction Potentials. J. Am. Chem. Soc. 2012, 134 (40), 16701–16716. 10.1021/ja306438n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Suzuki N; Hofstra JL; Poremba KE; Reisman SE Nickel-Catalyzed Enantioselective Cross-Coupling of N-Hydroxyphthalimide Esters with Vinyl Bromides. Org. Lett. 2017, 19 (8), 2150–2153. 10.1021/acs.orglett.7b00793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Arias-Rotondo DM; McCusker JK The Photophysics of Photoredox Catalysis: A Roadmap for Catalyst Design. Chem. Soc. Rev. 2016, 45 (21), 5803–5820. 10.1039/C6CS00526H. [DOI] [PubMed] [Google Scholar]
- (56).Fillman KL; Przyojski JA; Al-Afyouni MH; Tonzetich ZJ; Neidig ML A Combined Magnetic Circular Dichroism and Density Functional Theory Approach for the Elucidation of Electronic Structure and Bonding in Three- and Four-Coordinate Iron(ii)–N-Heterocyclic Carbene Complexes. Chem. Sci. 2015, 6 (2), 1178–1188. 10.1039/C4SC02791D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (57).Kawamata Y; Vantourout JC; Hickey DP; Bai P; Chen L; Hou Q; Qiao W; Barman K; Edwards MA; Garrido-Castro AF; deGruyter JN; Nakamura H; Knouse K; Qin C; Clay KJ; Bao D; Li C; Starr JT; Garcia-Irizarry C; Sach N; White HS; Neurock M; Minteer SD; Baran PS Electrochemically Driven, Ni-Catalyzed Aryl Amination: Scope, Mechanism, and Applications. J. Am. Chem. Soc. 2019, 141 (15), 6392–6402. 10.1021/jacs.9b01886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).Cherney AH Development of Nickel-Catalyzed Asymmetric Reductive Cross-Coupling of Benzylic Electrophiles, California Institute of Technology, 2015. https://resolver.caltech.edu/CaltechTHESIS:06012015-102412340. [Google Scholar]
- (59).Venanzi LM 140. Tetrahedral Nickel(II) Complexes and the Factors Determining Their Formation. Part I. Bistriphenylphosphine Nickel(II) Compounds. J. Chem. Soc. 1958, No. 0, 719–724. 10.1039/JR9580000719. [DOI] [Google Scholar]
- (60).Garton G; Henn DE; Powell HM; Venanzi LM 682. Tetrahedral Nickel(II) Complexes and the Factors Determining Their Formation. Part V. The Tetrahedral Co-Ordination of Nickel in Dichlorobistriphenylphosphinenickel. J. Chem. Soc. 1963, No. 0, 3625–3629. 10.1039/JR9630003625. [DOI] [Google Scholar]
- (61).Cotton FA; Goodgame DML New Tetrahedral Complexes of Nickel(II). J. Am. Chem. Soc. 1960, 82 (22), 5771–5774. 10.1021/ja01507a001. [DOI] [Google Scholar]
- (62).Venanzi LM Tetrahedral Complexes of Nickel (II) and the Factors Determining Their Formation. Journal of Inorganic and Nuclear Chemistry 1958, 8, 137–142. 10.1016/0022-1902(58)80175-X. [DOI] [Google Scholar]
- (63).Cotton FA; Faut OD; Goodgame DML Preparation, Spectra and Electronic Structures of Tetrahedral Nickel(II) Complexes Containing Triphenylphosphine and Halide Ions as Ligands. J. Am. Chem. Soc. 1961, 83 (2), 344–351. 10.1021/ja01463a021. [DOI] [Google Scholar]
- (64).Davies JE; Gerloch M; Phillips DJ Phosphine π-Acceptor Properties in Dihalogenobis(Triphenylphosphine)-Nickel(II) and -Cobalt(II). J. Chem. Soc., Dalton Trans. 1979, No. 11, 1836–1842. 10.1039/DT9790001836. [DOI] [Google Scholar]
- (65).Gerloch M; Hanton LR; Manning MR Tetrahedral Complexes of Nickel(II): Electronic Spectra, γ and π Bonding, and the Electroneutrality Principle. Inorganica Chimica Acta 1981, 48, 205–214. 10.1016/S0020-1693(00)90092-6. [DOI] [Google Scholar]
- (66).Koester VJ; Dunn TM Electronic Spectrum of the Tetrachloronickelate(II) Complex at 2.2.Deg.K. Inorg. Chem. 1975, 14 (8), 1811–1817. 10.1021/ic50150a014. [DOI] [Google Scholar]
- (67).Lever AB P. Inorganic Electronic Spectroscopy, 2nd ed.; Elsevier, 1984. [Google Scholar]
- (68).Gerloch M; Manning MR Structural and Ligand Field Parameters of Some Trigonally Distorted Tetrahedral Cobalt(II) and Nickel(II) Complexes. Inorg. Chem. 1981, 20 (4), 1051–1056. 10.1021/ic50218a020. [DOI] [Google Scholar]
- (69).Sigel H; Martin RB Coordinating Properties of the Amide Bond. Stability and Structure of Metal Ion Complexes of Peptides and Related Ligands. Chem. Rev. 1982, 82 (4), 385–426. 10.1021/cr00050a003. [DOI] [Google Scholar]
- (70).Connelly NG; Geiger WE Chemical Redox Agents for Organometallic Chemistry. Chem. Rev. 1996, 96 (2), 877–910. 10.1021/cr940053x. [DOI] [PubMed] [Google Scholar]
- (71).Okamoto K; Shida N; Morizumi H; Kitano Y; Chiba K Oxidation Potential Gap (ΔEox): The Hidden Parameter in Redox Chemistry. Angewandte Chemie International Edition 2022, 61 (30), e202206064. 10.1002/anie.202206064. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.