Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Apr 6.
Published in final edited form as: Inorg Chem. 2021 Aug 18;60(17):13400–13408. doi: 10.1021/acs.inorgchem.1c01720

Using JPP to Identify Ni Bidentate Phosphine Complexes In Situ

Matthew D Hannigan 1, Anne J McNeil 2, Paul M Zimmerman 3
PMCID: PMC8937619  NIHMSID: NIHMS1784631  PMID: 34405991

Abstract

Identifying intermediates of Ni-containing reactions can be challenging due to the high reactivity of Ni complexes and their sensitivity toward air and moisture. Many Ni bidentate phosphine complexes are diamagnetic and can be analyzed in situ via 31P NMR spectroscopy, but the oxidation state of Ni is difficult to determine using 31P chemical shift analysis alone. The J-coupling between P atoms, JPP, has been proposed to correlate with oxidation state, but few investigations have looked at how JPP is affected by parameters such as length of the linker or identity of the phosphine or other ligands. The present investigation into the JPP values of Ni bidentate phosphine complexes with two-carbon and three-carbon linkers shows that the JPP values observed in 31P NMR spectra, ∣JPP∣, are competent indicators of the oxidation state at Ni. For complexes with two-carbon linkers, ∣JPP∣ > 40 Hz is typical of Ni0 while ∣JPP∣ < 30 Hz is typical of NiII; this trend is reversed for complexes with three-carbon linkers. Additionally, the Lewis acidity of the Ni and Lewis basicity of the phosphine ligand affect JPP predictably. For example, increased P-to-Ni donation arising from more-donating phosphines or more-withdrawing ligands trans to the P atoms causes a more negative JPP. These results should enable the oxidation state of Ni and properties of ligands in Ni bidentate phosphine complexes to be determined in situ during reactions containing these species.

Graphical Abstract

graphic file with name nihms-1784631-f0001.jpg

INTRODUCTION

Nickel complexes bearing bidentate phosphine ligands (Scheme 1a) are ubiquitous in many synthetically relevant transformations.1 The oxidation state of Ni typically determines the reactions available to the complex, with electron-rich Ni0 complexes reacting via addition and insertion, while NiII complexes commonly undergo elimination.2. Kumada,3-11 Suzuki,12-20 Negishi,21-30 and Stille31 couplings and the Buchwald–Hartwig amination32-41 utilize these reactivity differences to activate C─X bonds (where X = Cl, Br, I) by addition to Ni0 and form C─C and C─N bonds by elimination from NiII. Given the relevance of Ni complexes to these synthetically important bond-forming reactions, many investigations involving Ni complexes have utilized ligand design to impart control over reactivity. Bidentate phosphine complexes are useful in this regard, because the linker and other groups bound to P atoms are easily varied. Short linkers between the P atoms enforce a cis geometry at the metal center, enabling facile reductive elimination in comparison to complexes with a trans geometry.2,42 The other groups bound to the P atoms enable control over the coordination environment, imparting stereoselectivity or modified reactivity.1 Additionally, due to the high abundance of the spin-active 31P isotope,43 obtaining spectra via nuclear magnetic resonance (NMR) spectroscopy is easier in comparison to complexes of C-, N-, O-, and S-based ligands.

Scheme 1.

Scheme 1.

Ni Bidentate Phosphine Complexes and Their JPP Values

Ni bidentate phosphine complexes can be characterized by NMR spectroscopy because many are diamagnetic,44 giving well-resolved spectra that are free of paramagnetic shifts and associated line broadening.45 Ciofini and coworkers noted in a recent study, however, that predicting and rationalizing 31P chemical shifts of metal complexes is complicated by the contrasting effects of ligand donation and back-bonding.46 This challenge in rationalizing 31P chemical shifts makes analyzing organometallic reaction mixtures less straightforward in comparison to using 1H and 13C chemical shifts for organic molecules.47,48 Additionally, this influence of donation and back-bonding prevents identifying the oxidation state of the metal by chemical shift because σ-donation and π-back-bonding in P─M bonds can change dramatically with oxidation state.47 The utility of NMR spectroscopy—which otherwise is a convenient means for the in situ analysis of reactions—is therefore limited by its inability to assign the oxidation state. Instead, the oxidation state must be identified via other methods, which often require isolation or purification of the desired complex. This ex situ characterization can be especially challenging for reactive intermediates, which may be short-lived or difficult to isolate. Given the ease of obtaining NMR spectra from reactions of Ni bidentate phosphines, and the ability to easily keep NMR samples moisture- and air-free,49,50 a protocol for understanding the structure of Ni bidentate phosphine complexes via 31P NMR spectroscopy would be useful.

In place of the chemical shift, J-coupling constants have been investigated to understand the structure and properties of bidentate phosphine complexes.51 J-coupling arises when spin-active nuclei in different electronic environments interact through bonds via magnetic interactions with electrons.52 The strength of this interaction is described using a J-coupling constant denoted as nJXY, where X and Y are the nuclei that are coupled and n is the number of bonds separating the nuclei. J-coupling constants involving protons, such as JHH, have been useful for obtaining chemically meaningful information. For example, the sign of J-coupling (positive or negative) identifies whether JHH arises from two-bond coupling (2JHH) or three-bond coupling (3JHH) because the former tends to be negative and the latter is positive.53-58 2JHH of CH2 groups can give electronic information about the substituents on the carbon atom, with more σ-withdrawing substituents or π-donating substituents making 2JHH less negative.59 The 3JHH coupling gives structural information on the dihedral angle between the two H atoms via the Karplus equation, enabling chemists to understand the geometry of pairs of H atoms from the J-coupling alone.60 For this reason, analyses based on three-bond J-couplings have proven useful for the structural characterization of small molecules61,62 as well as polymers,63 peptides,64,65 and proteins.66,67

Similar to the case for JHH, JPP can convey important electronic and structural information about pairs of P atoms and the metals they are bound to.68 For metal complexes with multiple monodentate phosphines, the magnitude of 2JPP can indicate a cis or trans geometry, with 2JPP being larger for complexes with trans phosphine ligands.51 2JPP can also provide electronic information about the P atoms because 2JPP depends on the electronegativity of the groups bound to P.48,69 Additionally, complexes with cis-monodentate phosphine ligands tend to have different JPP values in comparison to complexes with bidentate phosphine ligands. For example, in Mo and W phosphine complexes, J-coupling between P atoms through the metal (the only JPP coupling pathway for monodentate phosphine complexes) results in a negative 2JPP while coupling through the metal and the carbon linker (only available in bidentate phosphine complexes) results in a positive JPP.70

For Ni complexes, systematic studies into the structural effects on JPP are lacking, but many authors have noted that the JPP of Ni bidentate phosphine complexes correlates with the oxidation state of Ni.71-79 Desnoyer et al. state that bidentate phosphine NiII complexes often have a low JPP (less than 30 Hz), and Ni0 complexes have a higher JPP (45–80 Hz; Scheme 1b)80 and propose that the dependence of JPP on oxidation state is influenced by the electronegativity of ligands trans to the phosphines (trans ligands). Similar insights have aided in identifying Ni complexes from reaction mixtures,81-83 but structural parameters that are commonly varied, such as linker length and R groups, have not been explored with respect to their effect on J-coupling. Additionally, most 1-D NMR spectra give only the magnitude of JPP (denoted herein as ∣JPP∣);89 the sign of JPP, which can be obtained experimentally by 2-D NMR spectroscopy,70 has not yet been explored but may contain useful chemical information. A deeper understanding of J-coupling in Ni bidentate phosphine complexes is needed to fill these knowledge gaps and improve our ability to identify Ni complexes in situ.

Herein a systematic survey of Ni bidentate phosphine complexes with two-carbon linkers (2C) and three-carbon linkers (3C) is used to understand how the JPP value varies with the chemical structure. Notably, we show that JPP trends with the oxidation state and linker length, where 2C linker complexes have JPP distinct from that of 3C linker complexes. Additionally, calculations show the sign of JPP is related to the chemical structure, with Ni0 complexes having a more positive JPP in comparison to NiII complexes and 2C complexes having a more positive JPP in comparison to 3C complexes (Scheme 1c). Finally, using charge transfer analysis, JPP is shown to vary with P-to-Ni donation, with more-donating phosphines or more-withdrawing trans ligands causing a more negative JPP. This study gives a more complete framework for identifying Ni complexes via 31P NMR spectra and for interpreting JPP in Ni bidentate phosphine complexes.

RESULTS AND DISCUSSION

Literature Survey of Measured ∣JPP∣ Values.

To understand the relationship between the complex structure and JPP, a Reaxys84 search was performed to find Ni bidentate phosphine complexes with reported NMR spectra (see the Supporting Information). For the 2C data set, the results were limited to species that also had single-crystal X-ray structures, to verify the oxidation state of Ni. This search yielded a sizable data set of 449 complexes. A similar search for three-carbon-linker complexes yielded a data set of only 67 complexes, the majority of which did not contain a resolvable ∣JPP∣ due to paramagnetism or symmetry. Therefore, 3C complexes without reported X-ray crystal structures were added to the data set, giving 243 complexes (see the Supporting Information). Afterward, the 2C and 3C data sets were screened to remove species that did not have an observable JPP, such as fluxional complexes or C2-symmetric complexes. Complexes of other linkers (e.g., one-carbon, four-carbon, etc.) were not included because few examples were available. The 2C and 3C data sets both had more NiII complexes in comparison to Ni0 complexes. This difference in representation may be due to survivorship bias, because Ni0 complexes have high sensitivity toward air and moisture,85,86 making isolation or characterization more difficult. Additionally, Ni0 π-complexes often form fluxional structures in solution, in equilibrium with other π-complexes, causing peak broadening in NMR spectra and reventing a clear assignment of ∣JPP∣.87,88 Therefore, information provided by these data sets is limited to complexes that are stable enough and rigid enough to show ∣JPP∣ and may not represent all possible ∣JPP∣ from all possible Ni0 and NiII bidentate phosphine complexes.

The data sets enabled examining the relationships between oxidation state and J-coupling (Figure 1a). The distributions of ∣JPP∣ were compared statistically by analyzing the variance in ∣JPP∣ for each oxidation state/linker length pair. The distribution of ∣JPP∣ between oxidation states for each linker are significantly different, as indicated by the small p-values for the 2C Ni0/NiII pair (p < 0.0001) and the 3C Ni0/NiII pair (p = 0.0034). This statistical difference between oxidation states and the small overlap of the data ranges (Table 1) indicate that it is possible to accurately assign oxidation states via experimentally observed ∣JPP∣ when the linker length is known. For the 2C data set, the Ni0 complexes have a higher ∣JPP∣ in comparison the NiII complexes, matching the relationship proposed by previous authors.71-80 For the 3C linker data set, however, we observed the opposite relationship, with Ni0 complexes giving a lower ∣JPP∣ in comparison to the NiII complexes, albeit with considerable overlap. Information about the 31P chemical shift was also obtained from these data sets and analyzed, but we did not observe a relationship with the oxidation state (Figure S8) or JPP (Figures S9-S11).

Figure 1.

Figure 1.

(a) Violin plots showing the distributions of experimentally observed ∣JPP∣ for different Ni oxidation states (Ni0, NiII) and phosphine linker lengths (two-carbon, three-carbon). Bold lines represent medians, and fine lines represent the 25th and 75th percentiles. Probability that an observed ∣JPP∣ represents a Ni0 or NiII complex for (b) 2C and (c) 3C data sets. The gap at the 60–70 Hz range is due to the lack of 3C complexes with ∣JPP∣ in that range.

Table 1.

Distribution of ∣JPP∣ Values (in Hz) Reported in the 2C and 3C Data Sets

2C Ni0 2C NiII 3C Ni0 3C NiII
90th percentile 73.6 47.6 37.1 95.0
75th percentile 66.0 27.1 32.7 83.0
median 57.3 17.7 28.0 44.7
25th percentile 48.0 8.9 12.1 38.9
10th percentile 40.7 4.0 3.7 33.4

On the basis of the data sets, we also determined the probability of an experimentally observed ∣JPP∣ arising from a Ni0 or NiII complex of a given linker length. When experimental ∣JPP∣ < 30 Hz or ∣JPP∣ > 40 Hz, the oxidation state can be assigned on the basis of JPP and the number of atoms in the linker. For example, if a Ni bidentate phosphine complex with a two-carbon linker is present in a reaction mixture, and the 31p NMR spectrum reveals a set of doublets with ∣JPP∣ = 57 Hz, there is a high likelihood (89%) that it is a Ni0 complex (Figure 1b). If ∣JPP∣ is instead 23 Hz, then there is a higher likelihood (93%) of it being a NiII complex. For 2C complexes with ∣JPP∣ within the range of 30–40 Hz, there is nearly an equal likelihood of the complex being Ni0 (52%) or NiII (48%). Similar results are observed for the 3C complexes (Figure 1c), where ∣JPP∣ in the range of 30–40 Hz has a 60% chance of arising from a NiII complex and a 40% chance of arising from a Ni0 complex. Therefore, ∣JPP∣ in the range of 30–40 Hz cannot reliably identify the oxidation state of Ni in a bidentate phosphine complex, while ∣JPP∣ values outside of this 30–40 Hz range can provide an oxidation state assignment. The reason ∣JPP∣ increases with increasing oxidation state for 2C complexes and decreases with increasing oxidation state for 3C complexes was not immediately obvious; therefore, quantum chemical calculations were used to elucidate the physical origin.

Signs of JPP for Ni Bidentate Phosphine Complexes.

JPP, like all J-couplings, can have a positive or negative sign. A negative J corresponds to a more stable antiparallel nuclear spin alignment, while a positive J corresponds to a more stable parallel nuclear spin alignment.89 Although most 31P NMR experiments can give the magnitude of JPP (i.e., ∣JPP∣), few can give the sign. We sought to understand the signs of the experimentally observed ∣JPP∣ in our data sets because the sign of JPP was shown to relate to the structure in other studies of JPP in metal complexes.70 To obtain signs for our data set, quantum chemical simulations were used (see the Supporting Information for computational details). The absolute values of the computed JPP trend linearly with the experimental values (JPPexperimental = 0.913∣JPPcomputed∣ + 2.24 Hz; R2 = 0.852), indicating that the coupling can be computed with reasonably good accuracy.

Two-bond couplings, such as the JPP values arising from the through-Ni coupling pathway, are expected to have a negative sign.47,70,89,90 Simulations show that this is not always true for Ni bidentate phosphine complexes, with many complexes having a positive JPP (Figure 2). Generally, the 3C complexes have more a negative JPP in comparison to 2C complexes by ~40 Hz, and NiII complexes have a more negative JPP in comparison to Ni0 by ~60 Hz regardless of linker length. The observation that NiII complexes have a more negative JPP in comparison to Ni0 regardless of linker length appears to contrast with the relationships observed with ∣JPP∣, where 2C complexes had the smallestJPP∣ for NiII and 3C complexes had the largestJPP∣ for NiII (Figure 1). These differences between JPP and ∣JPP∣ become consistent when it is recognized that a more negative JPP results in a smaller ∣JPP∣ when JPP is positive and a larger ∣JPP∣ when JPP is negative. Therefore, ∣JPP∣ is largest when both the linker length and oxidation state are associated with the same shift in JPP, such as in 2C Ni0 complexes (both shift JPP more positive) and 3C NiII complexes (both shift JPP more negative), explaining the odd relationship among JPP, linker, and oxidation state derived from Figure 1. To relate the effects of structure and oxidation state on JPP, the sign of a measured ∣JPP∣ therefore needs to be considered (vide infra).

Figure 2.

Figure 2.

Violin plots showing distributions of computed JPP values for different Ni oxidation states (Ni0, NiII) and phosphine linker lengths (two-carbon, three-carbon). Bold lines represent medians, and fine lines represent the 25th and 75th percentiles.

After we identified that JPP correlates with the ligand structure and oxidation state, an explanation is still needed to answer why JPP becomes more negative upon oxidation from Ni0 to NiII or upon increasing the linker length from 2C to 3C. Given the observed dependence of ∣JPP∣ on linker length and oxidation state, JPP may be described as

JPP=2JPP+m+1JPP (1)

where the 2JPP term arises from through-Ni coupling and the m+1JPP term arises from through-linker coupling (m = number of carbons in the linker). The effect of changes in oxidation state and linker length on the values of through-Ni and through-linker coupling was therefore explored further with this model in mind.

Effect of Ligand Donation on JPP.

J-coupling links pairs of nuclei through bonding electrons;52 thus, it is expected that changes to the electron density in bonds along the coupling pathway can affect J. This dependence is evidenced by the electronic effects observed for many 2JHH and 2JCH couplings.59,91 For Ni bidentate phosphine complexes, the 2JPP (i.e., through-Ni) coupling term in eq 1 would likely depend on the electron density in the P─Ni bonds. Therefore, charge transfer analysis using absolutely localized molecular orbitals92-95 was used to understand P─Ni bonds, including the effects of P-donation and back-bonding (see the Supporting Information for computational details).

The charge transfer analysis showed a strong correlation of JPP with the σ-bonding P-to-Ni charge transfer (CTP-to-Ni) (Figure 3), whereas the back-bonding Ni to P charge transfer showed a worse correlation with JPP (Figure S12). The relationship of JPP with CTP-to-Ni indicates that JPP should become more negative with increased charge donated to Ni. Notably, the slopes of the 2C and 3C lines of best fit do not differ significantly (p-value = 0.3066), indicating that CTP-to-Ni is independent of the linker length. This independence of CTP-to-Ni from linker length signifies that P-to-Ni charge transfer solely affects the through-Ni 2JPP term in eq 1. In principle, increased P-to-Ni donation can arise from an increasing Lewis basicity of P, an increasing Lewis acidity of Ni, or an increasing overlap of Ni and P orbitals. The relationship between JPP and P-to-Ni charge transfer also provides a rationale for why JPP varies with oxidation state, because NiII is generally more Lewis acidic than Ni0, inducing more charge donation from P atoms.

Figure 3.

Figure 3.

Plot of JPP versus P-to-Ni charge transfer with regression lines shown in black. 2C line of best fit: JPP = −1.648 Hz/me CTP-to-Ni + 108.6 Hz. 3C line of best fit: JPP = −1.542 Hz/me CTP-to-Ni + 50.22 Hz. Gray areas indicate the 95% confidence interval for lines of best fit. A single outlier, complex 2-2329, is shown as a gray circle and was excluded from the analysis.

While JPP trends with P-to-Ni charge transfer, the physical origin of this dependence needed further elucidation. According to Ramsey, J-coupling occurs via four different mechanisms.52 The Fermi contact mechanism is the dominant J-coupling mechanism for JPP in other compounds96,97 and was therefore expected to be dominant for complexes in our data set. Indeed, calculations of the four different Ramsey contributors to JPP indicated that the Fermi contact contribution dominated the total JPP (see the Supporting Information). In the Fermi contact mechanism, J-coupling arises via spin-pairing of s electrons with a nucleus due to their nonzero electron density at the nuclei of atoms.89,52 Given this, and the high 3s character of the lone pairs of trivalent P atoms, it follows that the value of JPP should be strongly affected by the donation of P lone pairs into Ni.

This relationship between JPP and CTP-to-Ni tracks with the presence of electron-withdrawing groups on ligands trans to P atoms, likely by affecting the Lewis acidity of Ni. Structurally related NiII complexes that differ in the number (e.g., 2-2025, 2-2057, 2-2087) or strength (e.g., 2-2046e, 2-2056, 2-2035)98 of withdrawing groups on the trans ligands demonstrate this relationship between CTP-to-Ni and JPP (Chart 1), with complexes containing more-withdrawing groups having a more negative JPP and a larger ∣JPP∣. Complexes of Ni0 (e.g., 2-2046d, 2-2046f, 2-2291) also exhibit the trend of a more negative JPP with an increased number withdrawing groups on the trans ligand, but because Ni0 complexes generally have a positive JPP, this is manifested as a less positive JPP and smaller ∣JPP∣.

Chart 1.

Chart 1.

Three Series of Complexes That Differ in the Identity of trans Ligands

The relationship between JPP and CTP-to-Ni also tracks with the donating ability of the phosphine ligand (e.g., 2-2110 and 2-2296 of Chart 2), with more-donating phosphines99 causing a more negative JPP. However, for complexes with large phosphines, JPP values were less negative and CTP-to-Ni values were smaller than those of similar complexes with smaller phosphines. This trend is opposite of what is expected, given that large alkyl substituents (such as tert-butyl) are more electron rich than smaller alkyl substituents and should therefore impart greater P-to-Ni charge transfer. This deviation might arise from the increased steric bulk of large ligands, causing P─Ni bonds to elongate, as observed on comparing complexes 2-2020 and 2-2143c (Chart 2). The increased P─Ni bond length in complexes containing sterically bulky ligands may cause a decrease in the effective overlap of P and Ni orbitals, causing a lower CTP-to-Ni and a less negative JPP than expected on the basis of the Lewis structure.

Chart 2.

Chart 2.

Series of Structurally Similar Complexes with Varied Phosphine and trans Ligands

While the effect of P-to-Ni charge transfer on JPP is useful for understanding the NMR spectra of Ni bidentate phosphine complexes, the trends observed with P-to-Ni charge transfer do not explain the differences in JPP for 2C in comparison to 3C complexes. This limitation is evidenced by the ranges of CTP-to-Ni, which span similar values for 2C complexes (28.2–133.0 me) and 3C complexes (28.4–112.8 me) but have trend lines with significantly different y intercepts (p-value <0.0001). On the basis of eq 1, the negative shift of JPP of 3C complexes in comparison to 2C complexes may arise from differences in the through-linker m+1JPP coupling due to the difference in linker identity. Alternatively, the negative shift of JPP of 3C complexes may be due to the through-Ni 2JPP term, most likely due to differences in P─Ni─P bite angle caused by the linker.

Effect of Linker Length on JPP.

Many 2J-coupling constants are dependent on the bond angle at the central atom;100,101 therefore, the relationship of JPP with P─Ni─P bite angle was evaluated. If the linker affects JPP via the through-Ni (2JPP) coupling term in eq 1, there would be a clear dependence of JPP on bite angle. However, a low correlation was observed between JPP and bite angle for the combined data sets (Figure S16). The poor correlations of JPP with bite angle indicate that the linker does not affect the through-Ni 2JPP coupling term. This finding suggests that differences in JPP between 2C and 3C complexes instead arise from differences in the through-linker m+1JPP coupling terms in eq 1.

The through-linker m+1JPP coupling describes 3JPP and 4JPP for the 2C and 3C complexes, respectively, which should obey Karplus-type relationships that depend on the linker geometry.100,102 The range of 3JPP values for complexes in the 2C data set is expected to be small because there is little conformational flexibility available to bidentate ligands with such short linkers when they are bound to metals.103 Similarly, the range of 4JPP values is expected to be small, and therefore the average values of m+1JPP are sufficient for eq 1 to be useful.

With these considerations in mind, a three-parameter model was found to be sufficient to describe JPP for 2C and 3C complexes. The best fit for the three parameters (one slope, one intercept, and one correction for 3C linkers) together enables eq 1 to become

JPP=1.594Hzme(CTP-to-Ni)+105.2Hz+LL={0Hzfor 2C complexes48.47Hzfor 3C complexes} (2)

where CTP-to-Ni is the numerical value of charge transfer from P-to-Ni from a charge transfer analysis and L is a parameter that describes the difference in average through-linker m+1JPP. Equation 2 shows a good R2 = 0.9267 and RMS error of 11.8 Hz for the combined data set of 203 complexes for which JPP and CTP-to-Ni values could be calculated (Figure 4).

Figure 4.

Figure 4.

Parity plot comparing JPP values computed from DFT calculations to the JPP values predicted by eq 2: slope, 1.00; y intercept, 0.00 Hz.

Implications for the Analysis of 31P NMR Spectra.

Equation 2 can be used to interpret 31P NMR spectra of Ni bidentate phosphine complexes, even when the sign of JPP is not directly available through measurements. For example, in the case where quantum chemical simulations are available, a computed value of CTP-to-Ni can be used with eq 2 to find JPP, for a comparison to ∣JPP∣ observed in spectra. Except for ∣JPP∣ ≲ 10 Hz, where the sign would be ambiguous, the sign will be apparent if the absolute value of the computed JPP for the simulated Ni complex is near the experimentally observed ∣JPP∣.

The further utility of eq 2 comes from its description of the behavior of JPP. The negative dependence of JPP on P-to-Ni charge transfer enables one to deduce the sign of JPP in a series of Ni bidentate phosphine complexes with varying ligand electronics. For example, for a sequence of Ni bidentate phosphine complexes with varying P-donating abilities, 1-D 31P NMR spectroscopy will give a series of JPP values. If ∣JPP∣ increases as the phosphine ligands become more donating, the observed JPP values are likely negative. If ∣JPP∣ decreases with more-donating phosphines, the JPP values are likely positive. A similar screening can be performed by keeping the phosphine ligand the same and incorporating withdrawing groups on the trans ligands, because a more negative JPP is expected for an increased withdrawing ability of trans ligands. Once the sign is deduced, the trends in Figure 2 suggest that the sign of JPP can be used to make an oxidation state assignment. Given the complexities involved in determining the sign of JPP via NMR spectroscopy, this approach—while indirect—may prove to be the simplest option in many cases. Additionally, the incorporation of a term (i.e., L) accounting for differences in through-linker coupling may enable a description of other complexes with varied linker lengths or linker identities, such as BINAP104 and SPIRAP,105 provided a value of L can be determined for the linker.

CONCLUSION

JPP is a convenient NMR parameter for elucidating the structure of Ni bidentate phosphine complexes. For 2C complexes, ∣JPP∣ < 30 Hz and ∣JPP∣ > 40 Hz are indicative of NiII and Ni0, respectively, while the reverse is true for 3C complexes. This seemingly odd relationship among ∣JPP∣, oxidation state, and linker length becomes more easily understood after examining the sign of JPP, where a more negative JPP is observed with a higher oxidation state for both linker lengths. The relationships between CTP-to-Ni and JPP will enable assignment of the sign of JPP by examining how ∣JPP∣ changes with structures that induce P-to-Ni donation. Alternatively, at least for complexes that are structurally similar to any of the 232 complexes of the present data set, the sign may be elucidated by comparison to the JPP values given in Tables S4 and S5 of the Supporting Information.

Given the ease of obtaining 31P NMR spectra of reaction mixtures, the correlations between JPP and ligand electronics are expected to be useful for identifying reaction intermediates in situ. Similar relationships of JPP with charge transfer are likely present in other phosphine complexes, but more analysis will be needed to understand the effects of metal identity or metal geometry. Put together, these insights should enable a deeper understanding of the structure and properties of key intermediates in reactions of Ni bidentate phosphine complexes.

Supplementary Material

Supporting Information

ACKNOWLEDGMENTS

The authors thank Jessica L. Tami and Dr. Eugenio Alvarado for help in understanding the limitations of 31P NMR spectroscopy. The authors also thank David L. Braun for computational cluster support and the U.S. Department of Defense for a National Defense Science and Engineering Graduate Fellowship for M.D.H. This work was supported in part by a grant from the National Institutes of Health to P.M.Z. (R35-GM-128830).

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.1c01720.

2C and 3C data set construction and processing, relationships with 31P chemical shift, computational details, and relationships with bite angle (PDF)

2C complexes and associated data (PDF)

3C complexes and associated data (PDF)

Cartesian coordinates of all two-carbon complexes (PDF)

Cartesian coordinates of all three-carbon complexes (PDF)

The authors declare no competing financial interest.

Contributor Information

Matthew D. Hannigan, Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109-1055, United States.

Anne J. McNeil, Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109-1055, United States; Macromolecular Science and Engineering Program, University of Michigan, Ann Arbor, Michigan 48109-2800, United States.

Paul M. Zimmerman, Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109-1055, United States.

REFERENCES

  • (1).Clevenger AL; Stolley RM; Aderibigbe J; Louie J Trends in the Usage of Bidentate Phosphines as Ligands in Nickel Catalysis. Chem. Rev 2020, 120, 6124–6196. [DOI] [PubMed] [Google Scholar]
  • (2).Hartwig JF Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books: 2010; p 324. [Google Scholar]
  • (3).Tamao K; Sumitani K; Kumada M Selective Carbon-Carbon Bond Formation by Cross-Coupling of Grignard Reagents with Organic Halides. Catalysis by Nickel-Phosphine Complexes. J. Am. Chem. Soc 1972, 94, 4374–4376. [Google Scholar]
  • (4).Taylor BLH; Swift EC; Waetzig JD; Jarvo ER Stereospecific Nickel-Catalyzed Cross-Coupling Reactions of Alkyl Ethers: Enantioselective Synthesis of Diarylethanes. J. Am. Chem. Soc 2011, 133, 389–391. [DOI] [PubMed] [Google Scholar]
  • (5).Erickson LW; Lucas EL; Tollefson EJ; Jarvo ER Nickel Catalyzed Cross-Electrophile Coupling of Alkyl Fluorides: Stereospecific Synthesis of Vinylcyclopropanes. J. Am. Chem. Soc 2016, 138, 14006–14011. [DOI] [PubMed] [Google Scholar]
  • (6).Tollefson EJ; Dawson DD; Osborne CA; Jarvo ER Stereospecific Cross-Coupling Reactions of Aryl-Substituted Tetrahydrofurans, Tetrahydropyrans, and Lactones. J. Am. Chem. Soc 2014, 136, 14951–14958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Dawson DD; Jarvo ER Stereospecific Nickel-Catalyzed Cross-Coupling Reactions of Benzylic Ethers with Isotopically-Labeled Grignard Reagents. Org. Process Res. Dev 2015, 19, 1356–1359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Taylor BLH; Harris MR; Jarvo ER Synthesis Of Enantioenriched Triarylmethanes By Stereospecific Cross-Coupling Reactions. Angew. Chem., Int. Ed 2012, 51, 7790–7793. [DOI] [PubMed] [Google Scholar]
  • (9).Nomura N; Rajanbabu TV Nickel-Catalyzed Asymmetric Allylation Of Alkyl Grignard Reagents. Effect Of Ligands, Leaving Groups And A Kinetic Resolution With A Hard Nucleophile. Tetrahedron Lett. 1997, 38, 1713–1716. [Google Scholar]
  • (10).Soni V; Sharma DM; Punji B Nickel-Catalyzed Regioselective C(2)-H Difluoroalkylation of Indoles with Difluoroalkyl Bromides. Chem. - Asian J 2018, 13, 2516–2521. [DOI] [PubMed] [Google Scholar]
  • (11).Guan B-T; Xiang S-K; Wang B-Q; Sun Z-P; Wang Y; Zhao K-Q; Shi Z-J Direct Benzylic Alkylation via Ni-Catalyzed Selective Benzylic sp3 C-O Activation. J. Am. Chem. Soc 2008, 130, 3268–3269. [DOI] [PubMed] [Google Scholar]
  • (12).Han FS Transition-Metal-Catalyzed Suzuki-Miyaura Cross-Coupling Reactions: A Remarkable Advance From Palladium to Nickel Catalysts. Chem. Soc. Rev 2013, 42, 5270–5298. [DOI] [PubMed] [Google Scholar]
  • (13).Kuwano R; Shimizu R An Improvement of Nickel Catalyst For Cross-Coupling Reaction of Aryl Boronic Acids With Aryl Carbonates by Using a Ferrocenyl Bisphosphine Ligand. Chem. Lett 2011, 40, 913–915. [Google Scholar]
  • (14).Trost BM; Spagnol MD Nickel-Catalyzed Coupling of Allylamines and Boronic Acids. J. Chem. Soc., Perkin Trans 1 1995, 2083–2096. [Google Scholar]
  • (15).Li Y; Wang K; Ping Y; Wang Y; Kong W Nickel-Catalyzed Domino Heck Cyclization/Suzuki Coupling for the Synthesis of 3,3-Disubstituted Oxindoles. Org. Lett 2018, 20, 921–924. [DOI] [PubMed] [Google Scholar]
  • (16).Okuda Y; Xu J; Ishida T; Wang C-A; Nishihara Y Nickel-Catalyzed Decarbonylative Alkylation of Aroyl Fluorides Assisted by Lewis-Acidic Organoboranes. ACS Omega 2018, 3, 13129–13140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Zhao Y-L; Li Y; Li Y; Gao L-X; Han F-S Aryl Phosphoramides: Useful Electrophiles for Suzuki-Miyaura Coupling Catalyzed by a NiCl2/dppp System (dppp = 1,3-bis-(diphenylphosphino)propane). Chem. - Eur. J 2010, 16, 4991–4994. [DOI] [PubMed] [Google Scholar]
  • (18).Zeng Z; Yang D; Long Y; Pan X; Huang G; Zuo X; Zhou W Nickel-Catalyzed Asymmetric Ring Opening of Oxabenzonorbornadienes with Arylboronic Acids. J. Org. Chem 2014, 79, 5249–5257. [DOI] [PubMed] [Google Scholar]
  • (19).Hu F; Lei X A Nickel Precatalyst for Efficient Cross-Coupling Reactions of Aryl Tosylates with Aryl Boronic Acids: Vital Role of dppf. Tetrahedron 2014, 70, 3854–3858. [Google Scholar]
  • (20).Sylvester KT; Wu K; Doyle AG Mechanistic Investigation of the Nickel-Catalyzed Suzuki Reaction of N,O-Acetals: Evidence for Boronic Acid Assisted Oxidative Addition and an Iminium Activation Pathway. J. Am. Chem. Soc 2012, 134, 16967–16970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).Johnson JB; Yu RT; Fink P; Bercot EA; Rovis T Selective Substituent Transfer from Mixed Zinc Reagents in Ni Catalyzed Anhydride Alkylation. Org. Lett 2006, 8, 4307–4310. [DOI] [PubMed] [Google Scholar]
  • (22).Tarui A; Shinohara S; Sato K; Omote M; Ando A Nickel Catalyzed Negishi Cross-Coupling of Bromodifluoroacetamides. Org. Lett 2016, 18, 1128–1131. [DOI] [PubMed] [Google Scholar]
  • (23).Yu D; Wang C-S; Yao C; Shen Q; Lu L Nickel-Catalyzed α-Arylation of Zinc Enolates with Polyfluoroarenes via C-F Bond Activation under Neutral Conditions. Org. Lett 2014, 16, 5544–5547. [DOI] [PubMed] [Google Scholar]
  • (24).Yamamoto T; Yamakawa T Nickel-Catalyzed Vinylation of Aryl Chlorides and Bromides with Vinyl ZnBr·MgBrCl. J. Org. Chem 2009, 74, 3603–3605. [DOI] [PubMed] [Google Scholar]
  • (25).Melzig L; Metzger A; Knochel P Room Temperature Crosscoupling of Highly Functionalized Organozinc Reagents with Thiomethylated N-heterocycles by Nickel Catalysis. J. Org. Chem 75, 2131–2133. [DOI] [PubMed] [Google Scholar]
  • (26).Wisniewska HM; Swift EC; Jarvo ER Functional-Group Tolerant, Nickel-Catalyzed Cross-Coupling Reaction for Enantioselective Construction of Tertiary Methyl-Bearing Stereocenters. J. Am. Chem. Soc 2013, 135, 9083–9090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Gavryushin A; Kofink C; Manolikakes G; Knochel P An Efficient Negishi Cross-Coupling Reaction Catalyzed by Nickel(II) and Diethyl Phosphite. Tetrahedron 2006, 62, 7521–7533. [Google Scholar]
  • (28).Metzger A; Melzig L; Knochel P Scaled-up Transition-Metal Catalyzed Cross-Coupling Reactions of Thioether-Substituted N-Heterocycles with Organozinc Reagents. Synthesis 2010, 16, 2853–2858. [Google Scholar]
  • (29).Melzig L; Dennenwaldt T; Gavryushin A; Knochel P Direct Aminoalkylation of Arenes, Heteroarenes, and Alkenes via Ni Catalyzed Negishi Cross-Coupling Reactions. J. Org. Chem 2011, 76, 8891–8906. [DOI] [PubMed] [Google Scholar]
  • (30).Melzig L; Metzger A; Knochel P Pd- and Ni-Catalyzed Cross-Coupling Reactions of Functionalized Organozinc Reagents with Unsaturated Thioethers. Chem. - Eur. J 2011, 17, 2948–2956. [DOI] [PubMed] [Google Scholar]
  • (31).Shirakawa E; Yamasaki K; Hiyama T Nickel-Catalyzed Cross Coupling Reactions of Aryl Halides with Organostannanes. J. Chem. Soc., Perkin Trans. 1 1997, 1, 2449–2450. [Google Scholar]
  • (32).Clark JSK; McGuire RT; Lavoie CM; Ferguson MJ; Stradiotto M Examining the Impact of Heteroaryl Variants of PAd-DalPhos on Nickel-Catalyzed C(sp2)-N Cross-Couplings. Organometallics 2019, 38, 167–175. [Google Scholar]
  • (33).Clark JSK; Voth CN; Ferguson MJ; Stradiotto M Evaluating 1,1’-Bis(phosphino)ferrocene Ancillary Ligand Variants in the Nickel-Catalyzed C-N Cross-Coupling of (Hetero)aryl Chlorides. Organometallics 2017, 36, 679–686. [Google Scholar]
  • (34).Iwai T; Harada T; Shimada H; Asano K; Sawamura M A Polystyrene-Cross-Linking Bisphosphine: Controlled Metal Monochelation and Ligand-Enabled First-Row Transition Metal Catalysis. ACS Catal. 2017, 7, 1681–1692. [Google Scholar]
  • (35).Omar-Amrani R; Thomas A; Brenner E; Schneider R; Fort Y Efficient Nickel-Mediated Intramolecular Amination of Aryl Chlorides. Org. Lett 2003, 5, 2311–2314. [DOI] [PubMed] [Google Scholar]
  • (36).Harada T; Ueda Y; Iwai T; Sawamura M Nickel-Catalyzed Amination of Aryl Fluorides with Primary Amines. Chem. Commun 2018, 54, 1718–1721. [DOI] [PubMed] [Google Scholar]
  • (37).Ackermann L; Song W; Sandmann R Nickel-Catalyzed, Base Mediated Amination/Hydroamination Reaction Sequence for a Modular Synthesis of Indoles. J. Organomet. Chem 2011, 696, 195–201. [Google Scholar]
  • (38).Pawlas J; Nakao Y; Kawatsura M; Hartwig JF A General Nickel-Catalyzed Hydroamination of 1,3-Dienes by Alkylamines: Catalyst Selection, Scope, and Mechanism. J. Am. Chem. Soc 2002, 124, 3669–3679. [DOI] [PubMed] [Google Scholar]
  • (39).Green RA; Hartwig JF Nickel-Catalyzed Amination of Aryl Chlorides with Ammonia or Ammonium Salts. Angew. Chem., Int. Ed 2015, 54, 3768–3772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (40).Borzenko A; Rotta-Loria NL; MacQueen PM; Lavoie CM; McDonald R; Stradiotto M Nickel-Catalyzed Monoarylation of Ammonia. Angew. Chem., Int. Ed 2015, 54, 3773–3777. [DOI] [PubMed] [Google Scholar]
  • (41).Yang P; Zhang C; Ma Y; Zhang C; Li A; Tang B; Zhou JS Nickel-Catalyzed N-Alkylation of Acylhydrazines and Arylamines Using Alcohols and Enantioselective Examples. Angew. Chem., Int. Ed 2017, 56, 14702–14706. [DOI] [PubMed] [Google Scholar]
  • (42).Birkholz M-N; Freixa Z; van Leeuwen PWNM Bite angle effects of diphosphines in C-C and C-X bond forming cross coupling reactions. Chem. Soc. Rev 2009, 38, 1099–1118. [DOI] [PubMed] [Google Scholar]
  • (43).Rosman KJR; Taylor PDP Isotopic Compositions of The Elements 1997. Pure Appl. Chem 1998, 70, 217–235. [Google Scholar]
  • (44).Shriver DF; Atkins PW; Langford CH Inorganic Chemistry; Oxford University Press: 1992; p 212. [Google Scholar]
  • (45).Bren KL NMR Analysis of Spin Densities; Wiley-VCH: 2015. [Google Scholar]
  • (46).Payard P; Perego LA; Grimauad L; Ciofini I A DFT Protocol for the Prediction of 31P NMR Chemical Shifts of Phosphine Ligands in First-Row Transition-Metal Complexes. Organometallics 2020, 39, 3121–3130. [Google Scholar]
  • (47).Kühl O Phosphorus-31 NMR Spectroscopy; Springer-Verlag, 2008; pp 7–12, 16–22. [Google Scholar]
  • (48).Pregosin PS; Kunz R Coupling Constants in 31P and 13C NMR of Transition Metal Phosphine Complexes; NMR Basic Principles and Progress Series; Springer-Verlag: 1979; pp 16–46. [Google Scholar]
  • (49).Carlson AW; Primka DA; Douma ED; Bowring MA Evaluation of Air-Free Glassware Using the Ketyl Test. Dalton Trans. 2020, 49, 15213–15218. [DOI] [PubMed] [Google Scholar]
  • (50).The air tolerance and moisture tolerance of NMR primarily derives from the ability to prepare samples in air- and moisture-free environments, such as on a nitrogen line or in a glovebox. Air-free NMR tubes, such as J. Young tubes, have been shown to be airtight for days, such as in ref 49, far longer than the time scale of most NMR spectroscopy.
  • (51).Pregosin PS NMR in Organometallic Chemistry; Wiley-VCH: 2012; pp 258–264. [Google Scholar]
  • (52).Ramsey NF Electron Coupled Interactions between Nuclear Spins in Molecules. Phys. Rev 1953, 91, 303–307. [Google Scholar]
  • (53).Fraser RR; Lemieux RV; Stevens JD Opposite Relative Signs of Geminal and Vicinal Proton-Proton Coupling Constants in Saturated Organic Molecules. J. Am. Chem. Soc 1961, 83, 3901–3902. [Google Scholar]
  • (54).Finegold H Signs of Nuclear Resonance Coupling Constants in Saturated Aliphatic Systems. Proc. Chem. Soc 1962, 213–214. [Google Scholar]
  • (55).Karplus M Comments on the Signs of Proton Coupling Constants. J. Am. Chem. Soc 1962, 84, 2458–2460. [Google Scholar]
  • (56).Freeman R; Bhacca NS Relative Signs of Nuclear Spin Coupling Constants: A Refinement of the Double Irradiation Experiment. J. Chem. Phys 1963, 38, 1088–1093. [Google Scholar]
  • (57).Freeman R; McLauchlan KA; Musher JI; Pachler KGR The Relative Signs of Geminal and Vicinal Proton Spin Coupling Constants. Mol. Phys 1962, 5, 321–327. [Google Scholar]
  • (58).Kaplan F; Roberts JD Nuclear Magnetic Resonance Spectroscopy. Abnormal Splitting of Ethyl Groups Due to Molecular Asymmetry. J. Am. Chem. Soc 1961, 83, 4666–4668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Cookson RC; Crabb TJ; Frankel JJ; Hudec J Geminal Coupling Constants in Methylene Groups. Tetrahedron 1966, 22, 355–390. [Google Scholar]
  • (60).Karplus M Vicinal Proton Coupling in Nuclear Magnetic Resonance. J. Am. Chem. Soc 1963, 85, 2870–2871. [Google Scholar]
  • (61).Thureau P; Carvin I; Ziarelli F; Viel S; Mollica G A Karplus Equation for the Conformational Analysis of Organic Molecular Crystals. Angew. Chem., Int. Ed 2019, 58, 16047–16051. [DOI] [PubMed] [Google Scholar]
  • (62).Thibaudeau C; Plavec J; Chattopadhyaya J A New Generalized Karplus-Type Equation Relating Vicinal Proton-Fluorine Coupling Constants to H-C-C-F Torsion Angles. J. Org. Chem 1998, 63, 4967–4984. [Google Scholar]
  • (63).Fukuda Y; Miyamae K; Sasanuma Y Computational Design of Polymers: Poly(Esteramide) and Polyurethane. RSC Adv. 2017, 7, 38387–38398. [Google Scholar]
  • (64).Stangler T; Hartmann R; Willbold D; Koenig BW Modern High Resolution NMR for the Study of Structure, Dynamics and Interactions of Biological Macromolecules. Z. Phys. Chem 2006, 220, 567–613. [Google Scholar]
  • (65).Bystrov VF Spin-Spin Coupling and the Conformational States Of Peptide Systems. Prog. Nucl. Magn. Reson. Spectrosc 1976, 10, 41–82. [Google Scholar]
  • (66).Lee JH; Li F; Grishaev A; Bax A Quantitative Residue-Specific Protein Backbone Torsion Angle Dynamics from Concerted Measurement of 3J Couplings. J. Am. Chem. Soc 2015, 137, 1432–1435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (67).Wang A; Bax A Determination of the Backbone Dihedral Angles φ in Human Ubiquitin from Reparametrized Empirical Karplus Equations. J. Am. Chem. Soc 1996, 118, 2483–2494. [Google Scholar]
  • (68).Verkade JG Spectroscopic Studies of Metal-Phosphorus Bonding in Coordination Complexes. Coord. Chem. Rev 1972, 9, 1–106. [Google Scholar]
  • (69).Bertrand RD; Ogilvie FB; Verkade JG Signs of Phosphorus-Phosphorus Coupling Constants in Coordination Compounds. J. Am. Chem. Soc 1970, 92, 1908–1915. [Google Scholar]
  • (70).Fontaine XLR; Kennedy JD; Shaw BL; Vila JM Determination of the Relative Signs of 2J(31P-31P) in complexes of Tungsten(0) and Molebdenum(0) using Two-dimensional [31P-31P]-COSY-45 Nuclear Magnetic Resonance Chemical Shift Correlation. J. Chem. Soc., Dalton Trans 1987, 2401–2405. [Google Scholar]
  • (71).Lejkowski ML; Lindner R; Kageyama T; Bódizs GÉ; Plessow PN; Müller IB; Schäfer A; Rominger F; Hofmann P; Futter C; Schunk SA; Limbach M The First Catalytic Synthesis of an Acrylate from CO2 and an Alkene -A Rational Approach. Chem. - Eur. J 2012, 18, 14017–14025. [DOI] [PubMed] [Google Scholar]
  • (72).Plessow PN; Weigel L; Lindner R; Schäfer A; Rominger F; Limbach M; Hofmann P Mechanistic Details of the Nickel-Mediated Formation of Acrylates from CO2, Ethylene and Methyl Iodide. Organometallics 2013, 32, 3327–3338. [Google Scholar]
  • (73).García JJ; Brunkan NM; Jones WD Cleavage of Carbon-Carbon Bonds in Aromatic Nitriles Using Nickel(0). J. Am. Chem. Soc 2002, 124, 9547–9555. [DOI] [PubMed] [Google Scholar]
  • (74).García JJ; Jones WD Reversible Cleavage of Carbon-Carbon Bonds in Benzonitrile Using Nickel(0). Organometallics 2000, 19, 5544–5545. [Google Scholar]
  • (75).Waterman R; Hillhouse GL η2-Organoazide Complexes of Nickel and Their Conversion to Terminal Imido Complexes via Dinitrogen Extrusion. J. Am. Chem. Soc 2008, 130, 12628–12629. [DOI] [PubMed] [Google Scholar]
  • (76).Mindiola DJ; Waterman R; Jenkins DM; Hillhouse GL Synthesis of 1,2-bis(di-tert-butylphosphino)ethane (dtbpe) Complexes of Nickel: Radical Coupling and Reduction Reactions Promoted by the Nickel(I) Dimer [(dtbpe)NiCl]2. Inorg. Chim. Acta 2003, 345, 299–308. [Google Scholar]
  • (77).Ateşin TA; Li T; Lachaize S; Brennessel WW; García JJ; Jones WD Experimental and Theoretical Examination of C-CN and C-H Bond Activations of Acetonitrile Using Zerovalent Nickel. J. Am. Chem. Soc 2007, 129, 7562–7569. [DOI] [PubMed] [Google Scholar]
  • (78).Curley JJ; Kitiachvili KD; Waterman R; Hillhouse GL Sequential Insertion Reactions of Carbon Monoxide and Ethylene into the Ni-C Bond of a Cationic Nickel(II) Alkyl Complex. Organometallics 2009, 28, 2568–2571. [Google Scholar]
  • (79).Desnoyer AN; Bowes EG; Patrick BO; Love JA Synthesis of 2-Nickela(II)oxetanes from Nickel(0) and Epoxides: Structure, Reactivity, and a New Mechanism of Formation. J. Am. Chem. Soc 2015, 137, 12748–12751. [DOI] [PubMed] [Google Scholar]
  • (80).Desnoyer AN; He W; Behyan S; Chiu W; Love JA; Kennepohl P The Importance of Ligand-Induced Backdonation in the Stabilization of Square Planar d10 Nickel π-Complexes. Chem. - Eur. J 2019, 25, 5259–5268. [DOI] [PubMed] [Google Scholar]
  • (81).Smith ML; Leone AK; Zimmerman PM; McNeil AJ Impact of Preferential π-Binding in Catalyst-Transfer Polycondensation of Thiazole Derivatives. ACS Macro Lett. 2016, 5, 1411–1415. [DOI] [PubMed] [Google Scholar]
  • (82).Lanni EL; McNeil AJ Mechanistic Studies on Ni(dppe)Cl2-Catalyzed Chain-Growth Polymerizations: Evidence for Rate-Determining Reductive Elimination. J. Am. Chem. Soc 2009, 131, 16573–16579. [DOI] [PubMed] [Google Scholar]
  • (83).Willot P; Koeckelberghs G Evidence for Catalyst Association in the Catalyst Transfer Polymerization of Thieno[3,2-b]thiophene. Macromolecules 2014, 47, 8548–8555. [Google Scholar]
  • (84).Reaxys: https://www.reaxys.com/ (data retrieved on Dec 22, 2020).
  • (85).Nett AJ; Cañellas S; Higuchi Y; Robo MT; Kochkodan JM; Haynes MT; Kampf JW; Montgomery J Stable, Well-Defined Nickel(0) Catalysts for Catalytic C-C and C-N Bond Formation. ACS Catal 2018, 8, 6606–6611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (86).Nattmann L; Saeb R; Nöthling N; Cornella J An Air-Stable Binary Ni(0)-Olefin Catalyst. Nature Catal. 2020, 3, 6–13. [Google Scholar]
  • (87).Bilbrey JA; Bootsma AN; Bartlett MA; Locklin J; Wheeler SE; Allen WD Ring-Walking of Zerovalent Nickel on Aryl Halides. J. Chem. Theory Comput 2017, 13, 1706–1711. [DOI] [PubMed] [Google Scholar]
  • (88).D’ Accriscio F; Ohleier A; Nicolas E; Demange M; Du Boullay OT; Saffon-Merceron N; Fustier-Boutignon M; Rezabal E; Frison G; Nebra N; Mézailles N [(dcpp)Ni(η2-Arene)] Precursors: Synthesis, Reactivity, and Catalytic Application to the Suzuki-Miyaura Reaction. Organometallics 2020, 39, 1688–1699. [Google Scholar]
  • (89).Levitt M Spin Dynamics: Basics of Nuclear Magnetic Resonance; Wiley-VCH: 2008. [Google Scholar]
  • (90).Chalmers BA; Nejman PS; Llewellyn AV; Felaar AM; Griffiths BL; Portman EI; Gordon E-JL; Fan KJH; Woolins JD; Bühl M; Malinka OL; Cordes DB; Salwin AMZ; Kilian P A Study of Through-Space and Through-Bond JPP Coupling in a Rigid Nonsymmetrical Bis(phosphine) and its Metal Complexes. Inorg. Chem 2018, 57, 3387–3398. [DOI] [PubMed] [Google Scholar]
  • (91).Pedersoli S; Tormena CF; dos Santos FP; Contreras RH; Rittner R Stereochemical Behavior of 1JCH and 2JCH NMR Coupling Constants in α-Substituted Acetamides. J. Mol. Struct 2008, 891, 508–513. [Google Scholar]
  • (92).Khaliullin RZ; Cobar EA; Lochan RC; Bell AT; Head-Gordon M Unravelling the Origin of Intermolecular Interactions Using Absolutely Localized Molecular Orbitals. J. Phys. Chem. A 2007, 111, 8753–8765. [DOI] [PubMed] [Google Scholar]
  • (93).Horn PR; Head-Gordon M Polarization contributions to intermolecular interactions revisited with fragment electric-field response functions. J. Chem. Phys 2015, 143, 114111. [DOI] [PubMed] [Google Scholar]
  • (94).Khaliullin RZ; Bell AT; Head-Gordon M Analysis of charge transfer effects in molecular complexes based on absolutely localized molecular orbitals. J. Chem. Phys 2008, 128, 184112. [DOI] [PubMed] [Google Scholar]
  • (95).Horn PR; Mao Y; Head-Gordon M Probing non-covalent interactions with a second generation energy decomposition analysis using absolutely localized molecular orbitals. Phys. Chem. Chem. Phys 2016, 18, 23067. [DOI] [PubMed] [Google Scholar]
  • (96).Huynh K; Lough AJ; Forgeron MAM; Bendle M; Pressa Soto A; Wasylishen RE; Manners I Synthesis and Reactivity of Phosphine-Stabilized Phosphoranimine Cations, [R3P·PR′2═NSiMe3]+. J. Am. Chem. Soc 2009, 131, 7905–7916. [DOI] [PubMed] [Google Scholar]
  • (97).Burck S; Gotz K; Kaupp M; Nieger M; Weber J; Schmedt auf der Gunne J; Gudat D Diphosphines with Strongly Polarized P-P Bonds: Hybrids between Covalent Molecules and Donor-Acceptor Adducts with Flexible Molecular Structures. J. Am. Chem. Soc 2009, 131, 10763–10774. [DOI] [PubMed] [Google Scholar]
  • (98).pKa values of the parent arenes were used to quantify the withdrawing strength of each aryl ligand: Shen K; Fu Y; Li J-N; Liu L; Guo Q-X What are the pKa values of C-H bonds in aromatic heterocyclic compounds in DMSO? Tetrahedron 2007, 63, 1568–1576. [Google Scholar]
  • (99).Tolman CA Steric Effects of Phosphorus Ligands in Organometallic Chemistry and Homogeneous Catalysis. Chem. Rev 1977, 77, 313–348. [Google Scholar]
  • (100).Contreras RH; Peralta JE Angular Dependence of Spin-Spin Coupling Constants. Prog. Nucl. Magn. Reson. Spectrosc 2000, 37, 321–425. [Google Scholar]
  • (101).Balci M Basic 1H- and 13C-NMR Spectroscopy, 1st ed.; Elsevier: 2005; pp 87–133. [Google Scholar]
  • (102).Grossmann G; Lang R; Ohms G; Scheller D Dependence of Vicinal 31P-31P and 31P-13C Coupling Constants on the Dihedral Angle of P-Diphosphonates. Magn. Reson. Chem 1990, 28, 500–504. [Google Scholar]
  • (103).Vitek AK; Jugovic TME; Zimmerman PMZ Revealing the Strong Relationships between Ligand Conformers and Activation Barriers: A Case Study of Bisphosphine Reductive Elimination. ACS Catal. 2020, 10, 7136–7145. [Google Scholar]
  • (104).Noyori R; Takaya H BINAP: An Efficient Chiral Element for Asymmetric Catalysis. Acc. Chem. Res 1990, 23, 345–350. [Google Scholar]
  • (105).Argüelles AJ; Sun S; Budaitis BG; Nagorny P Design, Synthesis, and Application of Chiral C2-Symmetric Spiroketal-Containing Ligands in Transition-Metal Catalysis. Angew. Chem., Int. Ed 2018, 57, 5325–5329. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

RESOURCES