Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Jul 24.
Published in final edited form as: Inorg Chem. 2023 Jul 10;62(29):11487–11499. doi: 10.1021/acs.inorgchem.3c01068

Investigating Metal–Metal Bond Polarization in a Heteroleptic Tris-ylide Diiron System

Ariana Z Spentzos 1, Sam R May 1, Alex M Confer 1, Michael R Gau 1, Patrick J Carroll 1, David P Goldberg 1,, Neil C Tomson 1,*
PMCID: PMC11071007  NIHMSID: NIHMS1980647  PMID: 37428000

Abstract

This article describes the synthesis, characterization, and S-atom transfer reactivity of a series of C3v-symmetric diiron complexes. The iron centers in each complex are coordinated in distinct ligand environments, with one (FeN) bound in a pseudo-trigonal bipyramidal geometry by three phosphinimine nitrogens in the equatorial plane, a tertiary amine, and the second metal center (FeC). FeC is coordinated, in turn, by FeN, three ylidic carbons in a trigonal plane, and in certain cases, by an axial oxygen donor. The three alkyl donors at FeC form through the reduction of the appended N=PMe3 arms of the monometallic parent complex. The complexes were studied crystallographically, spectroscopically (NMR, UV-Vis, and Mössbauer), and computationally (DFT, CASSCF) and found to be high-spin throughout, with short Fe–Fe distances that belie weak orbital overlap between the two metals. Further, the redox nature of this series allowed for the determination that oxidation is localized to the FeC. S-atom transfer chemistry resulted in the formal insertion of a S atom into the Fe–Fe bond of the reduced diiron complex to form a mixture of Fe4S and Fe4S2 products.

Graphical Abstract

Spectroscopic and computational analyses on a neutral diiron complex supported by an N4 binding pocket for one metal (FeN) and a C3 binding pocket for the other (FeC) supported assignments of FeN(II) and FeC(I) physical oxidation states. The chemical oxidation of this complex occurred at FeC, with the Fe–Fe bond polarity governed by axial, not equatorial, interactions. Treatment of the neutral complex with Ph3SH led to a mixture of Fe4Sx (x = 1,2) clusters.

graphic file with name nihms-1980647-f0010.jpg

INTRODUCTION

From enzymatic systems to heterogeneous catalysts, the study of metal-metal interactions has far reaching implications and provides insight into the activity and properties of myriad synthetic and naturally occurring systems.14 Since Cotton’s seminal reports on Re-based metal-metal bonding,5 bimetallic complexes have emerged as an important area of synthetic inorganic chemistry, due to their ability to perform cooperative multi-electron transformations. Bimetallic 3d transition metal complexes containing metal-metal bonds are few in comparison to their heavier congeners, partly due to the diminished orbital overlap between the smaller 3d manifolds,6 but the potential for performing multi-electron chemistry with two metals that readily undergo single-electron transfer, as is the case for 3d transition metals, motivates a desire to develop ligands capable of controlling the coordination environments about dinuclear cluster cores.

The vast majority of diiron complexes contain carbonyl or thiolate bridges and are not constructed with ligands that support and control the bimetallic core. Of the more systematically designed complexes, a small subset exhibit C3-symmetry, typically containing nitrogen and/or phosphorus donors,7 with one metal serving as a metalloligand that has been used to modulate the electronics of a second, reactive metal. This approach has led to interesting electronic structures for the complexes as the Fe–Fe interaction is tuned by the nature of the supporting ligand framework.

The most well characterized examples of these trigonal diiron complexes are shown in Figure 1.812 All share short Fe–Fe distances (2.20–2.46 Å) and high-spin electron configurations (5-7 unpaired electrons), despite the varied ligand donor types. For a symmetric complex in this trigonal field, the entire d-manifold may participate in Fe–Fe bonding, leading to the possibility of one σ-, two π-, and two δ-symmetry bonds. Even in the highest symmetry cases, however, the bond orders remain fairly low, with limited Fe–Fe multiple bonding due to occupation of the π*/δ* manifolds. The complexes of this type are primarily differentiated by the extent of metal-metal bond delocalization. With more asymmetric ligand sets, the energy gap between the d-manifolds increases, leading to more polarized σ bonding and little to no π/δ bonding.7 It should be noted that the ligand types do not solely determine the degree of localization.9 However, across a series of homologous diiron complexes, ligand effects are expected to determine the amount of Fe–Fe bond polarization.

Figure 1.

Figure 1.

Literature examples of C3-symmetric diiron complexes.

Herein, we describe the synthesis and characterization of axially pseudo-trigonal diiron complexes that result from the unexpected methyl group C–H activation of P3tren, a phosphinimine-substituted tris(2-aminoethyl)amine ligand. Previously, we have used the zwitterionic character of the phosphinimines in this ligand system to create a cationic electrostatic field in the secondary coordination sphere that is capable of tuning the valence manifolds of CuI coordination complexes.13 During the course of related explorations of P3tren iron complexes targeted at making use of the C3-symmetric binding pocket for small molecule activation chemistry, a chemical reduction led to an unexpected diiron product with an unusual trialkyl coordination environment about one Fe center.

We have used spectroscopic and computational methods to evaluate the effect of the differing primary coordination spheres of the two Fe centers (N4Fe vs. C3(O)Fe) on the metal-metal bonding relative to related diiron complexes. One of these complexes was then investigated for its ability to reduce a sulfur-atom transfer reagent to form bridging sulfide products.

RESULTS AND DISCUSSION

Synthesis and characterization.

The treatment of FeCl2 with P3tren in acetonitrile led to the clean formation of the mononuclear, trigonal bipyramidal complex [(P3tren)FeCl][Cl] (1, Scheme 1) in 95% yield as a white solid. Crystallographic analysis revealed the complex to be a 1:1 ion pair with an outer sphere chloride. The cation was found to exhibit local C3v symmetry in the solid state, with unexceptional bond metrics (Table 1) considering the characterization of the metal center as high-spin (S = 2) FeII (solution phase, Evans’ method).

Scheme 1.

Scheme 1.

Synthesis of complexes 1-4

Table 1.

Selected and averaged bond distances for all compounds.

Compound Fe-Nax (Å) Fe-Fe (Å) P-N (Å) Fe-Cl (Å) Fe-C (Å) Fe-Neq (Å) FeN-S (Å) FeC-S (Å) Fe-O (Å)
1 2.334(1) - 1.596(2) 2.4761(5) - 2.075(1) - - -
2 2.279(2) - 1.627(1) 2.2852(6) - 1.995(2) - - -
3 2.258(1) 2.2808(3) 1.612(2) - 2.126(2) 2.028(2) - - -
[4-OTf]0 2.241(8) 2.538(2) 1.611(9) - 2.129(1) 2.012(8) - - 2.264(8)
[4-THF]+ 2.218(2) 2.5537(4) 1.617(2) - 2.115(3) 2.009(2) - - 2.377(2)
5 2.427(4) 2.6908(8) a
3.6188(8) b
1.603(4) - 2.118(5) 2.073(4) 2.443(1) 2.301(1) -
6 2.374(6) 3.485(1) 1.606(5) - 2.157(7) 2.095(5) 2.446(2) 2.354(1) -
a

Refers to the Fe-Fe distance within the Fe-S core.

b

Refers to the FeN-FeC distances.

A cyclic voltammogram of 1 revealed one reversible oxidation at −0.6 V vs. Fc0/+ (see Supporting Information). Upon chemical oxidation with excess AgOTf, the high-spin FeIII complex [(P3tren)FeCl][OTf]2 (2, Scheme 1) was formed as a bright orange solid. Crystallographic characterization of 2 revealed a slight elongation of the P–N bond lengths compared to those in 1 and a ~0.2 Å decrease in the Fe–Cl bond distance. Similarly, the Fe–Nax distance decreased from 2.334(1) Å for 1 to 2.279(2) Å; both values are within the range of previously reported Fe(II)/Fe(III) TBP complexes.14 Paramagnetically shifted resonances between 150 and 0 ppm were observed in the 1H NMR spectra for both complexes (see Supporting Information), though unequivocal assignments could not be made from these data.

While the only electrochemical feature observed by cyclic voltammetry within the solvent window of MeCN for 1 was a reversible one-electron oxidation, the addition of 2 equiv of KC8 to 1 in THF resulted in a color change to deep green along with gas evolution. Optimization of the synthetic procedure found that the use of toluene as a solvent with 4 equiv of KC8 led to reproducible crystallization of large green blocks of a new product in 30-40% yield following crystallization from concentrated hexane solutions at −20 °C. In an attempt to account for the Fe-based stoichiometry of the reaction, an additional equivalent of FeCl2 was introduced prior to the reductant. Doing so had a marginal effect on the outcome of the reaction, increasing the yield to a maximum of 42%. The use of either 3 or 5+ equiv of KC8 was not found to lead to alternative, isolable products but instead affected the yield of 3. A dark-brown, insoluble material forms in these reactions, and the only species observed in solution were 3, P3tren, and PMe3.

A crystal structure of the reduction product 3 (Figure 2) revealed an unexpected C–H activation on each of the PMe3 groups to give a diiron complex with three alkyl-like CH2-coordinated ylide ligands on the “top” iron (FeC) and the N4 coordination environment of the tren backbone on the “bottom” Fe (FeN, Scheme 1). Repeating the reaction in a sealed tube allowed us to identify by 1H NMR spectroscopy that H2 was a byproduct of this reaction. Despite the relatively low yields, no other organic or organometallic byproducts were observed during the reduction, and analytically pure material was isolated following crystallization.

Figure 2.

Figure 2.

Thermal ellipsoid plots of complexes 1, 3, [4-OTf]0, and [4-THF]+. Thermal ellipsoids are shown at the 50% probability level. Hydrogen atoms, counter anions, solvates and disordered atoms are omitted for clarity.

A close analysis of the crystallographic data revealed important features of the coordination environments about the metal centers. FeN has a distorted trigonal bipyramidal geometry enforced by the tetradentate ligand and the second iron atom. The Fe–Fe distance of 2.2808(3) Å is short but not unprecedented, and it is consistent with some degree of metal-metal bonding. The formal shortness ratio (FSR) for this interaction is 0.97, which is indicative of a M–M single bond.15 FeN lies slightly above the plane created by the phosphinimine nitrogens, with the NP–Fe–NP angles summing to 354°. FeC exhibits a similarly idealized trigonal planar environment, with the sum of C–Fe–C angles totaling 358°. If the Fe–Fe interaction is taken into account, FeC has a trigonal pyramidal geometry with a τ4 value of 0.857 (ideal trigonal pyramid τ4 is 0.851).16

Electrochemical investigation of the redox properties of 3 using cyclic voltammetry revealed two ostensibly reversible oxidation features at −0.70 V and −2.06 V. An additional feature was observed at ca. −1.0 V, but its intensity was inconsistent over multiple trials (see Supporting Information). Chemical oxidation of 3 with [Cp2Co][OTf] led to the formation of a one-electron oxidation product: [4-THF]+ or [4-OTf]0, depending on the crystallization solvent (THF and toluene, respectively; Scheme 1). Both are largely isostructural to 3, except for an added axial oxygen-based donor (THF or OTf) to FeC. [4-THF]+ and [4-OTf]0 display identical spectroscopic features when dissolved in non-coordinating solvents, signifying the facile formation of [4-OTf]0 from [4-THF]+ through a weakly bound THF ligand. When dissolved in THF, the triflate appeared to be coordinated or in equilibrium with the solvent, based on the broadened and shifted triflate signal in the 19F NMR spectrum (see Supporting Information). The Fe–Fe distance in [4-THF]+ elongated to 2.5537(4) Å (FSR 1.10), indicating a weakening of the metal-metal interaction. This expansion occurred at the expense of the FeN-Nax distance, which decreased from 2.258(1) Å in 3 to 2.218(2) Å in [4-THF]+. The rest of the analysis will be based on [4-OTf]0, due to the similar NMR and UV-Vis spectroscopic data and similar computational results for it and [4-THF]+.

[4-OTf]0 was found to be paramagnetic in solution, owing to a 1H NMR spectrum that shows three broad signals between 80 and 350 ppm. The solution-phase effective magnetic moment was determined to be 9.1 ± 0.1 μB (Evans’ method), corresponding to an S = 4 spin state with modest contributions from spin-orbit coupling. These data may be compared to those for complex 3, which was also found to be paramagnetic in solution. Two broad signals were visible in its 1H NMR spectrum, and the solution-phase effective magnetic moment was determined to be 8.2 ± 0.1 μB (Evans’ method), which is most readily interpreted as resulting from an S = 7/2 system, again with a modest amount of spin-orbit coupling. These spin state values track closely with those observed for complexes C and D (isoelectronic to 3) and B (isoelectronic to [4-OTf]0).

DFT Modelling.

Density functional theory (DFT) calculations were used to model the structures of complexes 1-4, using the structural and magnetochemical data for guidance. The geometry optimized structures for the cationic portions of complexes 1 and 2 closely matched the experimental data, and their electronic structures were consistent with high-spin Fe(II) and Fe(III) centers, respectively, in trigonal bipyramidal ligand fields. As expected, the Fe–Neq/Fe–Nax distances contract on oxidation. The experimental data show a contraction of ca. 0.06 Å for each type of interaction, consistent with the decrease in ionic radius following an oxidation event that involves depopulation of a primarily non-bonding orbital. The most notable deviation between the experimental and computed bond lengths was found in the Fe–Nax distance for 2, for which the predicted value of 2.102 Å was 0.177 Å shorter than the experimental distance of 2.279(2) Å.

The electronic structures of the dinuclear complexes 3 and [4-OTf]0 were more complex than those of the mononuclear species and invited comparison to related C3-symmetric diiron systems that have been described in the literature (Figure 1). All known C3-symmetric diiron systems have N- or P-atom donors coordinated to the iron centers, making the carbon-based ligands in compounds 3 and [4-OTf]0 unique and worth further investigation. Not only are C-based donors rare in diiron complexes, but even mononuclear iron complexes coordinated by three alkyl groups are scarce, with only a handful of known examples.1721 Notably, Smith and co-workers recently described several iron complexes coordinated by bis(ylide)diphenylborate ligands, which feature chelation of two ylides per chelate,22 and Neidig and coworkers have provided recent examples of homoleptic 3- and 4-coordinate iron alkyl complexes.17, 21 Both the ylide and phosphinimine ligand sets in 3 and [4-OTf]0 are expected to serve as strong σ-donors.23, 24 Related ylides have actually been shown to be even stronger donors than both NHCs and phosphinimines. This was demonstrated by César and co-workers, who synthesized a series of rhodium carbonyl complexes bound by various neutral donors that included a phosphinimine, an NHC, and an ylide. The latter showed the most activated CO stretching frequency from the series.25 We thus sought to investigate the effect of these ligand environments around each iron on the complexes’ spectroscopic properties and extents of M–M bonding, with a particular interest in the ability of ligand inequivalence to create spectroscopic inequivalence and polarization of the metal-metal interaction.

The DFT-optimized geometry of 3 closely matched the experimental data for this neutral diiron species. The optimized structure retained the pseudo-C3v geometry observed crystallographically, and the Fe–C and Fe–N distances all lay within 2% of the experimental values. DFT calculations also reproduced the Fe–Fe distance, yielding a calculated value of 2.265 Å, within 0.02 Å of the experimental distance. DFT was similarly robust in predicting the Fe–Fe (−0.039 Å), Fe–C, and Fe–Neq distances in [4-OTf]0 but lagged for the Fe–O (−0.124 Å) and Fe–Nax (+0.120 Å) distances. For [4-THF]+, the Fe–O distance (+0.015 Å), Fe–Nax (+0.09 Å), and Fe–Fe (−0.066 Å) distances were all predicted with excellent agreement to the experimental data.

UV-vis-NIR and Mössbauer Spectroscopic Characterization.

The UV-Vis-NIR absorption data for complexes 1-[4-OTf]0 provided useful insight into the electronic structures of each (see Figure 3 and the Supporting Information). Both 1 and [4-OTf]0 are nearly featureless in the visible region, while 2 and 3 have several intense visible transitions, as expected based on their colors. In the NIR region, where we would expect to see features that result from both inter- and intra-metal transitions, the mononuclear HS FeII complex 1 shows low-intensity d-d transitions (ε < 30 M−1cm−1), while the HS FeIII complex 2 is featureless. Compounds 3 and [4-OTf]0 exhibited more intense features in this region, which may similarly reflect intrametal d-d transition or transitions within a weakly interacting Fe2 bonding manifold. 3 has a weak absorption at 1044 nm (ε = 157 M−1cm−1), and [4-OTf]0 exhibits a comparable feature at 1410 nm (ε = 164 M−1cm−1) (Figure 3). These absorption bands are similar to the features observed by Lu for a series of complexes bearing the ligand in C (Figure 1). Dunbar and coworkers also observed features in the same range for a series of thiolate bridged dirhodium complexes.26 They characterized those transitions as intervalence charge transfer (IVCT) bands; however, their observed extinction coefficients are nearly an order of magnitude larger than those for 3 and [4-OTf]0. TD-DFT computational analyses were not able to accurately reproduce the experimental results, likely due to the significant ground state static correlation in the Fe–Fe interaction (see below), limiting our ability to model the origin of these features.(Figure 4)

Figure 3.

Figure 3.

NIR spectra of 1, 2, 3, and [4-OTf]0.

Figure 4.

Figure 4.

57Fe Mössbauer spectra of 1 (left) and [4-OTf]0 (right) collected at 80 K with a parallel 50 mT applied field. Experimental data are shown with open circles; simulations are shown in solid red and dashed lines. The difference between the experimental data and the simulations are given with grey lines.

Mössbauer spectroscopic data were collected at 80 K. The experimental Mössbauer spectrum of 1 showed a doublet at 1.00±0.03 mm/s (|ΔEQ| =2.57±0.06 mm/s). This isomer shift is typical of high-spin Fe(II) and was used to aid in validating our computationally predicted isomer shift data (Table 2). The DFT calculated isomer shift for 1 was determined to be 0.98 mm/s (|ΔEQ| = 2.70 mm/s), in good agreement with the experimental data.

Table 2.

Experimental and calculated Mössbauer parameters, electron count, and oxidation state assignment information.

Complex δ (mm/s)a |ΔEQ| (mm/s)b Total d-Electron Count EANc
1 Exp
Calc
1.00
0.98
2.57
2.70
6 16
2 Exp
Calc
~0
0.45

0.78
5 15
3
(FeC, FeN)
Exp
Calc
~0.5, ~0.8
0.47, 0.70
~1.4, ~0.8
0.27, 0.73
13 27
[4-OTf]0
(FeC, FeN)
Exp
Calc
0.42, 0.62
0.52, 0.73
1.46, 0.99
1.00, 0.67
12 28
[4]+
(FeC, FeN)
Calc 0.35, 0.66 0.99, 2.44 12 26
A Exp 0.65 0.32 13 25
B
(FeN, Fepy)
Exp 0.48, 0.58 1.31, 0.38 12 28
C
(FeN,FeP)
Exp 0.42, 0.55 0.13, 0.12 13 27
D
(FeN, FeP)
Exp 0.56, 0.44 1.11, 1.41 13 27
a

Estimated uncertainty of ±0.03 mm/s.

b

Estimated uncertainty of ±0.06 mm/s.

c

Effective atomic number.

Compared to the well-defined doublet for 1, the spectra for 2 and 3 were broadened and showed small quadupole splitting values that complicated the fits to the experimental data (see Supporting Information). This quality is common for half-integer-spin iron complexes under no/low field conditions.10 Despite our inability to extract quantitative data from the spectra for 2 and 3, qualitative isomer shift and quadrupole splitting information can be determined to a degree that allows for useful comparisons both across the series and to the computational data.

The experimental spectrum of 2 shows a single, broad feature at ~0 mm/s. The observation of only a single feature in a no/low field Mössbauer spectrum indicates the presence of minimal quadrupole splitting. The DFT calculated isomer shift for 2 was determined to be 0.46 mm/s (|ΔEQ| = 0.78 mm/s), which is consistent with oxidation of the metal center to Fe(III). The poor quantitative agreement with the experimental data may be explained in part by the use of the dicationic unit of 2 in the computational experiments. Holland and co-workers have recently demonstrated the sensitivity of 57Fe Mössbauer features to the electrostatic environment created by the specific locations of counterions.27 Furthermore, the low quadrupole splitting that is apparent from the experimental data is consistent with the presence of a minimal electric field gradient about the Fe nucleus that commonly results from a high-spin S = 5/2 electron configuration.

All complexes in Figure 1, except the symmetric complex A, have been reported to display distinct quadrupole doublets by Mössbauer spectroscopy, indicating that the metal centers in these homobimetallic systems are spectroscopically differentiable. For example, while the Mössbauer spectrum for A exhibits a single doublet at 0.65±0.03 mm/s for the symmetric Fe centers, the isoelectronic but dissymmetric species C has isomer shifts at 0.55±0.03 (FeP) and 0.42±0.03 (FeN) mm/s. The accumulated Mössbauer data for complexes A-D are listed in Table 2.

As was the case with 2, the Mössbauer data for 3 were found to be broad with poor signal-to-noise ratios (see Supporting Information, Figure S22), despite extended data collection times. Still, the spectrum for 3 did reveal two apparent doublets and could be assigned approximate values when informed by the computational data given below. This result is consistent with the presence of inequivalent iron environments, though we acknowledge that other fits to the data may be possible. These features were identified at ~0.8 and ~0.5 mm/s (|ΔEQ| = ~0.8, ~1.4 mm/s, respectively), with the former being more positive than the typical range of 0.4–0.6 mm/s for C3-symmetric diiron systems. We note that the LiCl adduct of C, though less structurally analogous to 3 than C itself, displays isomer shift values of 0.77±0.03 (FeP) and 0.50±0.03 (FeN) mm/s that are similar to the observed values for 3.10 The Mössbauer spectrum for [4-OTf]0 exhibited two inequivalent, but well resolved, quadrupole doublets, with isomer shifts of 0.62±0.03 and 0.42±0.03 mm/s (|ΔEQ| = 1.46±0.06, 0.99±0.06 mm/s, respectively). These Mossbauer parameters are similar to those reported for complex B, which is isoelectronic to [4-OTf]0 (see Table 2 for d-electron count and effective atomic numbers (EAN)) and displays isomer shifts of 0.58±0.03 (ΔEQ = 0.38±0.06 mm/s) and 0.48±0.03 (ΔEQ = 1.31±0.06 mm/s) mm/s.9

The computationally informed Mössbauer data for 3 and the unambiguous data for [4-OTf]0 indicate both that the isomer shifts move toward more negative values on oxidation and that oxidation of 3 to [4-OTf]0 has a more substantive impact on the more positive isomer shift. The former point is consistent with what is typically observed for Werner-type complexes; the isomer shift would be expected to become more negative as the complex is oxidized. This trend results from diminished shielding of the nucleus by the d-manifold on oxidation, thereby allowing for greater nuclear penetration by electron density in the s-manifolds. This is worth noting because this trend can be reversed by changes in spin-state, coordination number, and M–L backbonding, where the shorter M–L bonds lead to higher s-electron density and a more negative isomer shifts.2830 Strong M–M interactions also tend to result in more negative isomer shifts, due to decreased d-electron density on Fe and less shielding of the s-manifolds.31 These two effects can be in competition, with the oxidation event causing the isomer shift to become more negative and the weakened M–M interaction causing it to become more positive.32 It does not appear, however, that the M–M interaction is a dominant factor in the present case. The typical-to-low quadrupole splitting values are much lower than those observed for some multiply bonded metal complexes (4–5 mm/s).33 Instead, the more negative isomer shifts for [4-OTf]0 compared to those for 3 are consistent with an increase in the oxidation state of the diiron core being the dominant effect and a perturbation of an already weak metal-metal bonding interaction having a lesser impact.

DFT modelling of Mössbauer spectroscopic data.

Still unresolved at this stage is an assignment of which iron center, FeN or FeP, is giving rise to each of the quadrupole doublets observed in the Mössbauer spectra for 3 and [4-OTf]0. To aid in these assignments, we performed extensive computational modelling of our compounds and analogs from the literature. We found that the use of several known computational methods34, 35 and their associated fitting parameters led to unsatisfactory correlations with the experimental results. We thus formed a new test set composed of seven C3v diiron complexes that have reported experimental Mössbauer parameters (see Supporting Information). Their ρ(0) values were calculated using a methodology similar to that developed by the Holland group,34 and a regression analysis yielded a new set of fitting parameters that were then used to calculate isomer shifts for 3 and [4-OTf]0.

The computational results suggested that the more positive isomer shift values for 3 and [4-OTf]0 (δ = 0.6–0.8 mm/s) are attributable to FeN, with the more negative isomer shifts resulting from FeC (0.35–0.5 mm/s). Both sets of values lie more positive than what would be expected based on available literature precedent. In related diiron complexes containing an iron center coordinated in a TBP geometry by four nitrogens and an iron metalloligand, the N4Fe-ligated Fe centers exhibit isomer shifts between 0.42–0.50 mm/s.10 In the available literature on trialkyl iron species, Neidig and co-workers have characterized homoleptic [FeR3] complexes (R = Me, Et) that are structurally related to FeC.17, 21 Neidig’s Fe(II)-containing ions exhibit trigonal planar geometries with Fe–C distances of ca. 2.06 Å. The zero-field Mössbauer spectra for these salts reveal quadrupole doublets for R = Me at δ = 0.25±0.03 mm/s (ΔEQ = 1.36±0.06 mm/s) and for R = Et at δ = 0.19±0.03 mm/s (ΔEQ = 1.43±0.06 mm/s). Without more examples that span a larger oxidation state range, it is difficult to pinpoint a reason for the positive movement in the isomer shifts for both FeN and FeC in 3 and [4-OTf]0, so we next turned to in-depth electronic structure analyses in an effort to parse the origin of the trends in the Mössbauer spectra.

CASSCF modelling.

The DFT computational results used for the modelling the Mössbauer spectroscopic data indicated that the spin state energetics for 3 and [4-OTf]0 were consistent with the solution-phase experimental data, which had identified high-spin ground states for both complexes: Stot = 7/2 for 3 and Stot = 4 for [4-OTf]0. However, the single-determinant nature of DFT prevented evaluation of the extent to which multiple configurations contribute to the ground state electronic structures of these M–M bonded systems. Given the extensive static correlation that is possible for systems with weak-overlap covalent bonds, we employed complete active space self-consistent field (CASSCF) calculations on 3 and [4-OTf]0, with the natural orbital output from DFT as a starting point. The active spaces for each complex were chosen to include the valence d-manifolds for each species. Doing so resulted in CAS(13,10) and CAS(12,10) active spaces for 3 and [4-OTf]0, respectively; CASSCF natural orbital representations are shown in Figure 5. The ground state of 3 has a leading configuration (61% of the ground state) that is best described as (σ)2(π)2(π)2(π*)1(π*)1(σ*)1(FeC-dx2-y2)1(FeC-dxy)1(FeN-dx2-y2)1(FeN-dxy)1. This is similar to the ground state of C.10 Here, the δ-symmetry components of the d-manifold are expressed as atomic orbitals, since their half-occupation negates any net bonding interaction. The Fe–Fe formal bond order (FBO) for this leading configuration is 1.5, with the bonding combinations exhibiting polarization toward FeC. This polarization is reflected in the antibonding components, which show majority FeN character. The secondary configuration accounts for 16% of the ground state wavefunction and formally represents an excitation from π to π*, while the remaining configurations each contribute < 5% to the ground state and result from formal excitations within the β-spin manifold. When the entire ground state is considered, which includes partial occupations of the σ (1.8127), π (3.4409), σ* (1.1170), and π* (2.5541) systems, an effective bond order (EBO) of 0.79 is determined. Applying this same analysis to [4-OTf]0 (and [4-THF]+, see Supporting Information) revealed that oxidation results in depopulation of the π-bonding manifold. This and inclusion of the axial donor leads to an EBO of ~0, consistent with the substantial 0.26 Å increase in Fe–Fe distance observed experimentally. From here, we used the orthogonal valence bond (OVB) method to gain quantitative predictions of the relative electron densities at each metal center in the dinuclear complexes 3 and [4-OTf]0. This method uses an active space that is transformed through an orbital localization procedure into a new set of orbitals that are more easily interpreted by way of a valence bond analysis. This orbital localization yields a new wavefunction that is energetically indistinguishable from the original, making it an equally valid interpretation of the CASSCF results. Upon orbital localization of the active space of 3, an isolated d-manifold on each iron is observed. Examination of the electron distributions and configuration weights reveals that the oxidation states of FeN and FeC in 3 are best described as Fe(II) (6.04 e) and Fe(I) (6.96 e), respectively. A similar analysis on [4-OTf]0 suggests that the oxidation occurs primarily at FeC to give an Fe(II)/Fe(II) (6.00 e for each) oxidation state distribution within the diiron core. Nearly identical results were obtained for [4-THF]+ and [4]+, the latter of which is a theoretical complex that has no axial ligand coordinated to FeC (see Supporting Information).

Figure 5.

Figure 5.

Natural orbital representations of the CASSCF active spaces of 3 (top, CAS(13,10)) and [4-OTf]0 (bottom, CAS(12,10)), along with orbital occupation numbers and orbital assignments.

Physical Oxidation State Assignments.

The picture that emerges from the combined experimental and computational data may be summarized as follows. The crystallographic data for 1 indicate that high-spin Fe(II) supported by the N4 coordination environment of P3tren will exhibit Fe–Neq bond distances of ca. 2.07 Å and a Mössbauer isomer shift of δ = +1.0 mm/s. The shortening of the Fe–Neq distances to ca. 2.03 Å and, though the spectral data are somewhat ambiguous, the apparent modest movement of δ to ~0.8 mm/s on formation of 3 indicate that FeN within the complex’s [Fe2]3+ core is in a comparable oxidation state to the metal center within 1, albeit with some added covalency through binding to FeC in 3. The computational data similarly indicated that FeN within 3 is best described as an Fe(II) center, which is found to take part in weak-overlap covalent bonding to FeC. This latter attribution results from the observed low-energy electronic transitions for 3max = 1044 nm), which are made possible by the presence of a compact valence manifold for this species. The CASSCF data for this compound support the notion that a wide number of electron configurations lie within a small energy range and contribute to the complex’s ground-state electronic structure.

Since the [Fe2]3+ core of 3 contains a ferrous FeN, one is left to assume that FeC is best described as Fe(I); this description is supported by the data. The Fe–C distances (ca. 2.13 Å) are lengthened from known Fe(II) alkyl complexes in a trigonal geometry (ca. 2.08 Å),19, 21 and the computationally-informed isomer shift assignment ascribed to FeC in 3 (~0.5 mm/s) is more positive than the values for these same [FeR3] Fe(II) complexes (δ = ca. 0.23 mm/s). Further, the CV data for 3 reveal an electrochemical event at −0.70 V that is close to the Fe(II/III) couple in 1 (−0.60 V), which suggests that the −0.70 V feature for 3 may be attributable to an FeN(II/III) electrochemical event. The feature at −2.06 V for 3 could then be assigned to the Fe(I/II) couple at FeC, which is consistent with the data accumulated for the 1 electron oxidation products [4-OTf]0 and [4-THF]+. Finally, the assignment of an Fe(I) physical oxidation state to FeC in 3 is further supported by the computational data, in that an OVB analysis of the CASSCF data found that FeC was best described as Fe(I).

The computational data further suggest that the polarization of the Fe–Fe bond in 3 may be accounted for by the inequivalent axial coordination environments between FeC and FeN. The axial nitrogen bound to FeN serves to raise the energy of the d-manifold on FeN with respect to FeC, allowing it to attain a higher oxidation state despite its formally neutral coordination environment. This difference is represented in the polarization of the calculated σ-bonding interaction toward FeC, suggesting either that the equatorial ligands are only able to play a minor role in the polarization of the Fe–Fe interaction in these compounds or, more likely, that the phosphinimine-ylide ligands show strongly anionic character at both the N- and C-based donors, resulting in little effective polarization of the Fe–Fe interaction as a result of the equatorial ligand fields. The shorter P–CFe distances (ca. 1.74 Å) compared to those for the P–CMe bonds (ca. 1.81 Å) support the notion that resonance pictures like those depicted in Figure 6 provide a non-negligible contribution to the ground state of this ligand.

Figure 6.

Figure 6.

Resonance structures of 3 showing the zwitterionic character of the phosphinimines and the ylidic character of the P–CH2–Fe linkages.

The impact of the axial coordination was investigated further through a comparison of 3 to the oxidized species [4-OTf]0. The combination of the lowered electron count and the presence of an axial donor in [4-OTf]0 leads to an increase in the Fe–Fe distance by 0.26 Å. CASSCF calculations were consistent with disruption of the Fe–Fe bond; this outcome was supported by the NIR electronic absorption data for [4-OTf]0, which were red-shifted compared to those in 3 (Figure 3). The changing coordination number upon oxidizing 3 to [4-OTf]0 was also found to play a role in the Mössbauer parameters. The calculated isomer shift assigned to FeN in 3 (+0.70 mm/s) is close to the calculated values for FeN in [4-OTf]0 (+0.73 mm/s) and [4]+ (+0.66 mm/s). With the locus of oxidation at FeC, we anticipated that the isomer shift for FeC would move significantly toward a less positive number from the calculated value for 3 of +0.47 mm/s, but the value for [4-OTf]+ was found to be more positive (+0.53 mm/s). This appears to be a coordination number effect, in that the isomer shift of FeC decreased to +0.35 mm/s in [4]+, a species that is isomorphic to 3.

S-atom Transfer to 3.

Finally, we performed a preliminary investigation into the reactivity of 3 toward the sulfur atom transfer reagent HSCPh3 as a way of both demonstrating the reducing capacity of Fec and probing the reactivity of the ylide-based ligands. The Lu and Thomas compounds shown in Figure 1 generally retain their architectures through a range of chemical transformations. While many reactivity studies remain to be done with 3 and 4, the results presented below reveal the ready ability of this system to dissociate on group addition chemistry.

Upon slow addition of the thiol at low temperature and subsequent warming to room temperature, a vibrant red solution was formed. Two products, 5 and 6 (Scheme 2), were consistently co-crystallized from this solution in low yield. The low conversion and inability to separate the two precluded extensive characterization but nonetheless provided an example of how this weak M-M bond and unique ligand set may be used to support group transfer reactivity.

Scheme 2.

Scheme 2.

Synthesis of Fe-S insertion products

Intriguingly, both complexes involve both the formal insertion of a sulfur atom into the Fe–Fe bond and the formation of a dimer (Scheme 2 and Figure 7). Complex 5 is an Fe2S2 complex in which one ylidic arm has gained an H-atom, presumably from the thiol. The FeN–FeC distances average 3.6188 Å, demonstrating cleavage of the Fe–Fe bond, and the FeN–S bond distances (2.443(1) Å) are also elongated. The Fe2S2 core is planar with an Fe–Fe distance of 2.6964(8) Å. The core contains a mirror plane and has Fe–S distances of 2.301(1) and 2.443(1) Å, which are on the upper end of the expected range.36 Complex 6 is the product of the formal addition of 0.5 equivalents of an S-atom into 3 and has a μ4-S moiety bound to the four iron atoms. Again, disruption of the Fe–Fe bond is required to form this complex, and the product appears to contain four ferrous centers. The sulfur is in a distorted tetrahedral configuration with Fe-S-Fe angles ranging from 80.9° to 106.5°. The Fe–S bond lengths are again long, with average FeN–S and FeC–S bond lengths of 2.446(2) Å and 2.354(1) Å, respectively. Both complexes are paramagnetic, and like the parent complex, NMR spectroscopic peak assignments cannot definitively be made, especially as they are only observed as a mixture. The weak M–M bond observed in 3 along with its apparent propensity to dissociate an ylidic arm can thus lead to interesting S atom inserted products. This reactivity contrasts with that of compounds B, C, and D (Figure 1), for which the dinuclear core is maintained and chemistry primarily occurs on the distal metal. Future work will focus on ways to expand and exploit this dinuclear participation.

Figure 7.

Figure 7.

Thermal ellipsoid plots of complexes 5 and 6. Thermal ellipsoids are shown at the 50% probability level. Hydrogen atoms, solvates, and disordered atoms are omitted for clarity.

CONCLUSION

In conclusion, the reduction of a phosphinimine-supported mononuclear Fe(II) complex led to the serendipitous formation of an unusual trigonal diiron complex following C–H activation and H2 extrusion. This result is notable given the unusual set of three ylidic carbon donors coordinated the “top” metal (FeC) and the strong phosphinimine and amine nitrogens supporting the “bottom” iron (FeN). This complex and its one electron oxidation product, [4-OTf]0/[4-THF]+, exhibit high-spin ground states that allow for a modest amount of Fe–Fe bonding. However, the varying coordination environments of the metal centers lead to polar metal-metal interactions that reflect distinct oxidation states for the two metal centers in 3. Importantly, it appears that the difference in axial donor, instead of the C- vs. N-equatorial donor identity, bears significant responsibility for the polarization of the Fe–Fe interaction. Upon addition of a sulfur atom donor to 3, cleavage of the weak M–M bond occurred, and two bridging sulfide complexes were formed. Both 3 and [4-OTf]0 were characterized using spectroscopic and computational methods and add a unique ligand architecture to the repertoire of trigonal, high-spin diiron complexes that have applications ranging from small molecule catalysis to single-molecule magnets.

EXPERIMENTAL

General Considerations

All reactions containing transition metals were performed under an inert atmosphere of N2, using standard Schlenk line or glovebox techniques. Glassware, stir bars, filter aid (Celite), and 3 Å molecular sieves were dried in an oven at 150 °C for at least 12 h prior to use. All solvents (THF, acetonitrile, toluene, n-pentane, and diethyl ether) were dried by passage through a column of activated alumina, deoxygenated by passage through a copper Q5 column where applicable, sparged with N2, and stored over activated 3 Å molecular sieves under an inert atmosphere. Deuterated solvents were purchased from Cambridge Isotope Laboratories, Inc., and dried over either Na/benzophenone (THF-d8, C6D6) or CaH2 (MeCN-d3), isolated by vacuum transfer (THF-d8, MeCN-d3) or distillation (C6D6), and stored under an inert atmosphere over 3 Å sieves. P3tren was synthesized following previously published procedures.13 KC8 was synthesized according to a literature procedure37 and stored under nitrogen at −35 °C in a glovebox prior to use. PMe3 (98%) was either purchased from Strem Chemicals or synthesized according to a literature procedure.13 AgOTf was dried over P2O5 and stored in the dark. [nBu4N][PF6] (98%) was purchased from MilliporeSigma and recrystallized twice from hot ethanol, followed by drying at 60 °C under 30 mbar for 10 h and stored in a glovebox under dry nitrogen prior to use.38 Ferrocene (98%) was purchased from Acros Organics, purified by sublimation three times,39 and stored in a glovebox under dry nitrogen prior to use. Cobaltocenium triflate was prepared using a modified literature procedure, replacing indium triflate with silver triflate.40 All other chemicals were used as received. 1H and 19F{1H} NMR spectra were recorded on a UNI 400 spectrometer. All chemical shifts (δ) are reported in units of ppm, with references to the residual protio-solvent resonance for proton chemical shifts. Internal PhF was used for referencing 19F NMR spectra.5 Elemental analysis was performed by Midwest Microlab, LLC or at the CENTC Elemental Analysis Facility, Department of Chemistry, University of Rochester. IR spectra (KBr pellet) were collected on a JASCO FT/IR-480 Plus spectrometer FTIR, and the IR spectra are given in the Supporting Information. Solution phase effective magnetic moment data were determined using Evans’ method.41 P3tren’ refers to the activated methyl group variant of P3tren.

Cyclic voltammetry experiments were performed in a VAC OMNI-LAB glovebox with an Epsilon E2 Potentiostat. The data were processed with BASi Epsilon-EC software version 2.13.77. All experiments were performed under an N2 atmosphere in a glovebox using an electrochemical cell that consists of a glassy carbon (3 mm outer diameter) working electrode, a platinum wire counter electrode, and a Ag/AgNO3 (0.01 M in acetonitrile) reference electrode. All experiments were conducted in acetonitrile, with 1 mM analyte and 100 mM [nBu4N][PF6] as the supporting electrolyte. Potentials were reported versus Cp2Fe+/0.

UV–Vis absorption spectra were collected from 200 to 1000 nm using an Agilent Cary 60 UV–vis spectrophotometer at room temperature. NIR spectra were recorded on a Perkin Elmer 950 UV-Vis/NIR spectrophotometer at room temperature. The samples were prepared under an N2 atmosphere in a glovebox. Stock solutions were prepared by dissolving a known mass (ca. 10 mg) of sample in 10.00 mL of solvent (3, [4-OTf]0 (THF); 1 and 2 (CH3CN)). Varying amounts of the stock solutions (50.0–400.0 μL) were then diluted to 5.00 mL and transferred to a 10 mm path-length quartz cuvette with a screw cap for data collection. For each species, at least four spectra at different concentrations were collected. The absorption intensities of various peak maxima were plotted vs. concentration to ensure that absorption data were being collected in the linear response range of the spectrometer. Linear regression fits to the plotted data resulted in R2 values ≥ 0.994.

57Fe Mossbauer spectra were collected on a spectrometer from SEE Co. (Edina, MN) operating in the constant-acceleration mode in a transmission geometry. Samples were cooled in an SVT-400 cryostat from Janis Research Co. (Wilmington, MA), using liquid N2 as a cryogen for 80 K measurements. A parallel applied magnetic field of ca. 50 mT was applied. Samples were prepared under nitrogen in custom Delrin cups equipped with tight-fitting Delrin inserts. The zero velocity of each Mossbauer spectrum refers to the centroid of a 25 micron metallic iron (α-Fe) foil collected at room temperature. The WMOSS program (SEE Co., formerly WEB Research Co., Edina, MN) was used to analyze the data, which were fit to simple quadrupole doublets with Lorentzian line shapes.

Crystallographic X-ray intensity data were collected on a Bruker D8QUEST CMOS (2, [4-THF]+, 6) or a Bruker APEXII CCD (1, 3, [4-OTf]0) area-detector diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å) at 100 K or a Rigaku XtaLAB Synergy-S HyPix-6000HE HPC (5) area-detector diffractometer with confocal multilayer optic-monochromated Cu-Kα radiation (λ=1.54184Å) at a temperature of 100K. Rotation frames were integrated using CrysalisPro42 for 5 and SAINT43 producing a listing of unaveraged F2 and σ(F2) values. The intensity data were corrected for Lorentz and polarization effects and for absorption using SCALE3 ABSPACK44 or SADABS.45 The structures were solved by direct methods by using SHELXT46 and refined by full-matrix least-squares, based on F2 using SHELXL-2017.47 For [4-OTf]0 the asymmetric unit consists of four crystallographically-independent molecules. Non-hydrogen atoms were refined anisotropically and hydrogen atoms were refined using a riding model. CCDC entries 2154284-2154290 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

DFT geometry optimizations were performed with unrestricted Kohn-Sham DFT using the B97-D3 pure GGA functional4850 with Orca v4.2.1.51 The initial geometries for all optimizations were obtained from crystallographic coordinates. This functional was chosen due to its reliable ability to predict bond lengths and angles within this system. The def2-TZVP basis sets and def2-TZVP/J auxiliary basis sets (used to expand the electron density in the resolution-of-identity (RI) approach) were used for all Fe and Fe-bound atoms.5254 The def2-SV(P) basis sets were used for all other atoms. The RIJCOSX approximation was used for computational feasibility. The SCF calculations were tightly converged (1x10−8 Eh in energy, 1x10−7 Eh in the density change, and 5x10−7 in the maximum element of the DIIS error vector). In all cases the geometries were considered converged after the energy change was less than 1x10−6 Eh, the gradient norm and maximum gradient element were smaller than 3x10−4 Eh-Bohr−1 and 1x10−4 Eh-Bohr−1, respectively, and the root-mean square and maximum displacements of all atoms were smaller than 6x10−4 Bohr and 1x10−3 Bohr, respectively.

CASSCF calculations were carried out in Orca v4.2.1 to characterize the ground state across the diiron series. The same basis sets were used as above. RIJCOSX approximation was once again used to reduce the computational cost. All orbitals in the 3d manifold were included for both iron centers, consisting of 13 electrons for 3, and 12 electrons for both [4-THF]+ and [4-OTf]0.

The process of modelling Mössbauer isomer shifts with DFT requires computation of the nuclear electron density, ρ(0), followed by the use of fitting parameters (α, β, and C) that allow ρ(0) to be translated into a calculated isomer shift. The values for α and β are obtained by performing a linear regression analysis between calculated ρ(0) values and experimental isomer shifts for a test set of complexes, while C is an unrefined constant that is used to prevent β from becoming unreasonably large. These values are dependent on the level of theory but often translate well to novel compounds, provided the new species are structurally and electronically related to those in the test set. Isomer shifts for each iron center were estimated using single point calculation data carried out in Orca v4.2.1. EPR/NMR analysis was carried out to get accurate values for ρ(0) for each center. Quadrupole splitting parameters and isomer shifts were then estimated based on the ρ(0) values for each Fe using a published method with parameters from test set 1.55

Synthetic Procedures

[(P3tren)FeCl]Cl (1):

FeCl2 (353.5 mg, 2.79 mmol) was dissolved in ca. 15 mL of acetonitrile with stirring. An acetonitrile solution (10 mL) of P3tren (1.0310g, 2.80 mmol) was added to the FeCl2 solution via cannula at room temperature. The solution turned light yellow and was allowed to stir for 20 min before being filtered through Celite. Volatile materials were removed in vacuo, and the resulting white solid was washed with THF before being dried again in vacuo for 10 min to yield a white powder (1.2433g, 90%). Large, colorless blocks suitable for single-crystal X-ray diffraction were grown by vapor diffusion of diethyl ether into a saturated acetonitrile solution of 1. 1H NMR (400 MHz, CD3CN, 300 K): δ = 119.21 (br s, 6H), 106.42 (br s, 6H), −1.77 (br s, 27H) ppm. Anal. Calcd. for C15H39Cl2Fe1N4P3 (495.17 g/mol): C, 36.38; H, 7.94; N, 11.31%. Found: C, 35.81; H, 7.89; N, 11.16%.

[(P3tren)FeCl][OTf]2 (2):

A solution of AgOTf (217.4 mg, 0.85 mmol) in 2-3 mL of acetonitrile was added at room temperature to a stirred solution of 1 (204.5 mg, 0.41 mmol) in 5 mL of acetonitrile, resulting in a rapid color change from light yellow to bright orange-red with the formation of a dark precipitate. The mixture was stirred for 5 min before all volatile materials were removed in vacuo. The reddish solid was washed with THF (2 x 2 mL) and then extracted into a minimal amount of acetonitrile (ca. 2 mL). Vapor diffusion of ether into the saturated acetonitrile solution at room temperature yielded large red blocks after 2 days (244.4 mg, 78%). 1H NMR (400 MHz, CD3CN, 300 K): δ = 14.05 (br s) ppm. 19F{1H} NMR (376 MHz, CD3CN, 300 K): δ = −78.53 ppm (s). Anal. Calcd. For C17H39Cl1Fe1N4O6P3S2 (757.85 g/mol): C, 26.94; H, 5.19; N, 7.39%. Found: C, 26.91; H, 4.88; N, 7.28%.

(P3tren’)FeFe (3):

Safety note: The filtrand from this reaction is pyrophoric. 1 (1.2437g, 2.51 mmol) and FeCl2 (322.0 mg, 2.54 mmol) were suspended in 15 mL of toluene with stirring. A slurry of KC8 (1.3697g, 10.1 mmol) in 5 mL of toluene was added to the iron suspension at room temperature. After ~30 min the mixture became a deep green color and was stirred overnight with venting to prevent build-up of H2. The mixture was filtered through Celite and the volatiles were removed in vacuo. The resulting green solid was extracted into 20 mL of n-hexane, filtered through Celite, and concentrated to 10 mL. Storage of the resulting solution overnight at −35 °C yielded large, dark-green blocks (387.1 mg, 32%). The crystals were washed with cold HMDSO (3 x 1 mL). The yield may be marginally increased upon further concentration and cooling of the mother liquor to yield additional crops of 3. Typical total yields ranged from 30-40%. 1H NMR (400 MHz, C6D6, 300 K): δ = 217.07 (br s), 83.55 (br s) ppm. Anal. Calcd. for C15H36Fe2N4P3 (477.09 g/mol): C, 37.76; H, 7.61; N, 11.74%. Found: C, 38.13; H, 7.45; N, 11.57%.

[(P3tren’)FeFe(thf)][OTf] ([4-THF]+):

3 (113.4 mg, 0.24 mmol) was dissolved in 5 mL of THF, and a suspension of [Cp2Co][OTf] (75.0 mg, 0.22 mmol) in 3 mL THF was added dropwise over 1 min. The solution turned dark brownish-green and was stirred for 5 min. The volatile materials were removed in vacuo, giving a red-brown solid. The solid was washed with pentane (4 x 2 mL) to remove cobaltocene, and the remaining yellow-green solid was extracted into THF (3 mL), and the solution was filtered through Celite. The THF solution was then layered with pentane (10 mL) and stored overnight at −25 °C. The yellow-green needles that grew during this time were washed with pentane and dried in vacuo to yield a crystalline product (81.7 mg, 49%). 1H NMR (400 MHz, THF-d8, 300 K): δ = 349.75 (br s), 152.62 (br s), 84.13 (br s) ppm. 19F{1H} NMR (376 MHz, THF-d8, 300 K): δ = −34.53 ppm (br s). Anal. Calcd. For C20H44F3Fe2N4O4P3S1 (698.26 g/mol): C, 34.40; H, 6.35; N, 8.02%. Found: C, 33.63; H, 5.98; N, 7.74%. We could not obtain passing elemental analysis due to the thermal instability of [4-THF]+.

(P3tren’)FeFeOTf ([4-OTf]0):

3 (79.4 mg, 0.17 mmol) was dissolved in 2 mL of THF and a suspension of [Cp2Co][OTf] (56.3 mg, 0.17 mmol) in 3 mL THF was added dropwise over 1 min. The solution turned dark brownish green and was stirred for 5 min. The volatile materials were removed in vacuo, giving a red-brown solid. The solid was washed with pentane (4x2 mL) to remove cobaltocene, and the remaining yellow-green solid was extracted into toluene (5x1 mL). The resulting solution was filtered through Celite, then layered with pentane (10 mL). Yellow-green needles grew after cooling to −25ºC overnight. The crystals were washed with pentane and dried in vacuo (12.6 mg, 12%). 1H NMR (400 MHz, C6D6, 300 K): δ = 351.86 (br s), 147.58 (br s), 86.51 (br s) ppm. 19F{1H} NMR (376 MHz, C6D6, 300 K): δ = −18.54 ppm (br s). 1H NMR (400 MHz, THF-d8, 300 K): δ = 349.84 (br s), 152.76 (br s), 84.19 (br s) ppm. 19F{1H} NMR (376 MHz, THF-d8, 300 K): δ = −36.90 ppm (br s). Anal. Calcd. for C16H36F3Fe2N4O4P3S1 (626.16 g/mol): C, 30.69; H, 5.80; N, 8.95%. Found: C, 26.81; H, 5.26; N, 7.41%. We could not obtain passing elemental analysis due to the thermal instability of [4-OTf]0.

[(P3tren’)FeFe]2(S2) (5) and [(P3tren’)FeFe]2S (6):

3 (23.0 mg, 0.05 mmol) was dissolved in 2 mL of Et2O and frozen in a cold well. HSCPh3 (13.2 mg, 0.05mmol) was dissolved in 3 mL of Et2O and partially frozen in a cold well. The solution of 3 was removed from the cold well and the chilled solution of HSCPh3 was added to the solution of 3 with stirring in 3 aliquots over 1–2 minutes. A precipitate formed and the solution turned brown-green. Upon further warming to room temperature the solution turned deep red with a red precipitate and was stirred for 2 minutes. The volatile materials were removed in vacuo, giving a red solid. The solid was sequentially extracted with hexane (1 mL), Et2O (1 mL), and THF (1 mL), with each extract filtered through Celite. X-ray quality yellow plates of 6 were grown from the concentrated light-pink Et2O or hexane solutions at −25ºC after several days. Vapor diffusion of pentane into the saturated red THF solution at −25ºC for at least 1 week yielded a mixture of dark red blocks of 5 and small amount of 6 (6.4 mg combined, ~26%). Despite varying the rate of addition and equivalents of HSCPh3 (0.5–1.0 equiv), the separability issues and product distribution precluded isolation of each product. 1H NMR (400 MHz, C6D6) δ 137.18 (s, 1H), 121.49 (s, 1H), 96.67 (s, 1H), 90.65 (s, 1H), 70.19 (s, 1H), 61.31 (s, 1H), 58.64 (s, 1H), 41.25 (s, 1H), 21.12 (s, 3H), −2.63 (s, 4H).

Supplementary Material

Supporting Information

ACKNOWLEDGMENT

We thank the National Institute of General Medical Sciences of the National Institutes of Health for financial support through grant numbers R35GM128794 (NCT) and R01GM119374 (DPG). NCT also thanks the University of Pennsylvania for financial support.

Footnotes

Supporting Information

NMR spectra, UV-vis spectra, IR spectra, Mössbauer spectra, cyclic voltammetry traces, crystallographic details, and computational data (PDF). The Supporting Information is available free of charge on the ACS Publications website.

REFERENCES

  • (1).Campbell CT The Energetics of Supported Metal Nanoparticles: Relationships to Sintering Rates and Catalytic Activity. Acc. Chem. Res 2013, 46 (8), 1712–1719. DOI: 10.1021/ar3003514. [DOI] [PubMed] [Google Scholar]
  • (2).Blackburn NJ; Barr ME; Woodruff WH; van der Ooost J; de Vries S Metal-Metal Bonding in Biology: EXAFS Evidence for a 2.5 .ANG. Copper-Copper Bond in the CuA Center of Cytochrome Oxidase. Biochemistry 1994, 33 (34), 10401–10407. DOI: 10.1021/bi00200a022. [DOI] [PubMed] [Google Scholar]
  • (3).Lindahl PA Metal–metal bonds in biology. J. Inorg. Biochem 2012, 106 (1), 172–178. DOI: 10.1016/j.jinorgbio.2011.08.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).Liu L; Corma A Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev 2018, 118 (10), 4981–5079. DOI: 10.1021/acs.chemrev.7b00776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (5).Cotton FA Metal—Metal Multiple Bonds and Metal Clusters. In Reactivity of Metal-Metal Bonds, ACS Symposium Series, Vol. 155; AMERICAN CHEMICAL SOCIETY, 1981; pp 1–16. [Google Scholar]
  • (6).McGrady JE Introduction and General Survey of Metal–Metal Bonds. In Molecular Metal-Metal Bonds, 2015; pp 1–22. [Google Scholar]
  • (7).Krogman JP; Thomas CM Metal–metal multiple bonding in C3-symmetric bimetallic complexes of the first row transition metals. Chem. Commun 2014, 50 (40), 5115–5127. DOI: 10.1039/C3CC47537A. [DOI] [PubMed] [Google Scholar]
  • (8).Zall CM; Zherebetskyy D; Dzubak AL; Bill E; Gagliardi L; Lu CC A Combined Spectroscopic and Computational Study of a High-Spin S = 7/2 Diiron Complex with a Short Iron–Iron Bond. Inorg. Chem 2012, 51 (1), 728–736. DOI: 10.1021/ic202384b. [DOI] [PubMed] [Google Scholar]
  • (9).Tereniak SJ; Carlson RK; Clouston LJ; Young VG; Bill E; Maurice R; Chen Y-S; Kim HJ; Gagliardi L; Lu CC Role of the Metal in the Bonding and Properties of Bimetallic Complexes Involving Manganese, Iron, and Cobalt. J. Am. Chem. Soc 2014, 136 (5), 1842–1855. DOI: 10.1021/ja409016w. [DOI] [PubMed] [Google Scholar]
  • (10).Miller DL; Siedschlag RB; Clouston LJ; Young VG; Chen Y-S; Bill E; Gagliardi L; Lu CC Redox Pairs of Diiron and Iron–Cobalt Complexes with High-Spin Ground States. Inorg. Chem 2016, 55 (19), 9725–9735. DOI: 10.1021/acs.inorgchem.6b01487. [DOI] [PubMed] [Google Scholar]
  • (11).Greer SM; Gramigna KM; Thomas CM; Stoian SA; Hill S Insights into Molecular Magnetism in Metal–Metal Bonded Systems as Revealed by a Spectroscopic and Computational Analysis of Diiron Complexes. Inorg. Chem 2020, 59 (24), 18141–18155. DOI: 10.1021/acs.inorgchem.0c02605. [DOI] [PubMed] [Google Scholar]
  • (12).Kuppuswamy S; Powers TM; Johnson BM; Bezpalko MW; Brozek CK; Foxman BM; Berben LA; Thomas CM Metal–Metal Interactions in C3-Symmetric Diiron Imido Complexes Linked by Phosphinoamide Ligands. Inorg. Chem 2013, 52 (9), 4802–4811. DOI: 10.1021/ic302108k. [DOI] [PubMed] [Google Scholar]
  • (13).Weberg AB; McCollom SP; Thierer LM; Gau MR; Carroll PJ; Tomson NC Using internal electrostatic fields to manipulate the valence manifolds of copper complexes. Chem. Sci 2021, 12 (12), 4395–4404, 10.1039/D0SC06364A. DOI: 10.1039/D0SC06364A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Mandon D; Machkour A; Goetz S; Welter R Trigonal Bipyramidal Geometry and Tridentate Coordination Mode of the Tripod in FeCl2 Complexes with Tris(2-pyridylmethyl)amine Derivatives Bis-α-Substituted with Bulky Groups. Structures and Spectroscopic Comparative Studies. Inorg. Chem 2002, 41 (21), 5364–5372. DOI: 10.1021/ic011104t. [DOI] [PubMed] [Google Scholar]
  • (15).Chipman JA; Berry JF Paramagnetic Metal–Metal Bonded Heterometallic Complexes. Chem. Rev 2020, 120 (5), 2409–2447. DOI: 10.1021/acs.chemrev.9b00540. [DOI] [PubMed] [Google Scholar]
  • (16).Yang L; Powell DR; Houser RP Structural variation in copper(i) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, (9), 955–964. DOI: 10.1039/B617136B. [DOI] [PubMed] [Google Scholar]
  • (17).Sears JD; Muñoz III SB; Daifuku SL; Shaps AA; Carpenter SH; Brennessel WW; Neidig ML The Effect of β-Hydrogen Atoms on Iron Speciation in Cross-Couplings with Simple Iron Salts and Alkyl Grignard Reagents. Angew. Chem. Int. Ed 2019, 58 (9), 2769–2773. DOI: 10.1002/anie.201813578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Bedford RB; Brenner PB; Carter E; Cogswell PM; Haddow MF; Harvey JN; Murphy DM; Nunn J; Woodall CH TMEDA in Iron-Catalyzed Kumada Coupling: Amine Adduct versus Homoleptic “ate” Complex Formation. Angew. Chem. Int. Ed 2014, 53 (7), 1804–1808. DOI: 10.1002/anie.201308395. [DOI] [PubMed] [Google Scholar]
  • (19).Bedford RB; Brenner PB; Elorriaga D; Harvey JN; Nunn J The influence of the ligand chelate effect on iron-amine-catalysed Kumada cross-coupling. Dalton Trans. 2016, 45 (40), 15811–15817. DOI: 10.1039/C6DT01823H. [DOI] [PubMed] [Google Scholar]
  • (20).Serrano CB; Less RJ; McPartlin M; Naseri V; Wright DS The First-Row Transition Metal Interstitial Hydride Anion [{PhP(CH2)3Fe}4(μ4-H)]–. Organometallics 2010, 29 (22), 5754–5756. DOI: 10.1021/om100838u. [DOI] [Google Scholar]
  • (21).Muñoz III SB; Daifuku SL; Sears JD; Baker TM; Carpenter SH; Brennessel WW; Neidig ML The N-Methylpyrrolidone (NMP) Effect in Iron-Catalyzed Cross-Coupling with Simple Ferric Salts and MeMgBr. Angew. Chem. Int. Ed 2018, 57 (22), 6496–6500. DOI: 10.1002/anie.201802087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Gao Y; Pink M; Smith JM Iron(II) Complexes of an Anionic Bis(ylide)diphenylborate Ligand. Inorg. Chem 2020, 59 (23), 17303–17309. DOI: 10.1021/acs.inorgchem.0c02575. [DOI] [PubMed] [Google Scholar]
  • (23).Lee HB; Ciolkowski N; Winslow C; Rittle J High Spin Cobalt Complexes Supported by a Trigonal Tris(Phosphinimide) Ligand. Inorg. Chem 2021, 60 (16), 11830–11837. DOI: 10.1021/acs.inorgchem.1c01400. [DOI] [PubMed] [Google Scholar]
  • (24).Scharf LT; Gessner VH Metalated Ylides: A New Class of Strong Donor Ligands with Unique Electronic Properties. Inorg. Chem 2017, 56 (15), 8599–8607. DOI: 10.1021/acs.inorgchem.7b00099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Azouzi K; Duhayon C; Benaissa I; Lugan N; Canac Y; Bastin S; César V Bidentate Iminophosphorane-NHC Ligand Derived from the Imidazo[1,5-a]pyridin-3-ylidene Scaffold. Organometallics 2018, 37 (24), 4726–4735. DOI: 10.1021/acs.organomet.8b00733. [DOI] [Google Scholar]
  • (26).Coll RP; Dunbar KR Three Reversible Redox States of Thiolate-Bridged Dirhodium Complexes without Metal–Metal Bonds. J. Am. Chem. Soc 2020, 142 (38), 16313–16323. DOI: 10.1021/jacs.0c06205. [DOI] [PubMed] [Google Scholar]
  • (27).Skubi KL; Hooper RX; Mercado BQ; Bollmeyer MM; MacMillan SN; Lancaster KM; Holland PL Iron Complexes of a Proton-Responsive SCS Pincer Ligand with a Sensitive Electronic Structure. Inorg. Chem 2022, 61 (3), 1644–1658. DOI: 10.1021/acs.inorgchem.1c03499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Ye S; Bill E; Neese F Electronic Structures of the [Fe(N2)(SiPiPr3)]+1/0/–1 Electron Transfer Series: A Counterintuitive Correlation between Isomer Shifts and Oxidation States. Inorg. Chem 2016, 55 (7), 3468–3474. DOI: 10.1021/acs.inorgchem.5b02908. [DOI] [PubMed] [Google Scholar]
  • (29).Clouston LJ; Bernales V; Cammarota RC; Carlson RK; Bill E; Gagliardi L; Lu CC Heterobimetallic Complexes That Bond Vanadium to Iron, Cobalt, and Nickel. Inorg. Chem 2015, 54 (24), 11669–11679. DOI: 10.1021/acs.inorgchem.5b01631. [DOI] [PubMed] [Google Scholar]
  • (30).Eisenhart RJ; Rudd PA; Planas N; Boyce DW; Carlson RK; Tolman WB; Bill E; Gagliardi L; Lu CC Pushing the Limits of Delta Bonding in Metal–Chromium Complexes with Redox Changes and Metal Swapping. Inorg. Chem 2015, 54 (15), 7579–7592. DOI: 10.1021/acs.inorgchem.5b01163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (31).Wu B; Wilding MJT; Kuppuswamy S; Bezpalko MW; Foxman BM; Thomas CM Exploring Trends in Metal–Metal Bonding, Spectroscopic Properties, and Conformational Flexibility in a Series of Heterobimetallic Ti/M and V/M Complexes (M = Fe, Co, Ni, and Cu). Inorg. Chem 2016, 55 (23), 12137–12148. DOI: 10.1021/acs.inorgchem.6b01543. [DOI] [PubMed] [Google Scholar]
  • (32).Moore JT; Chatterjee S; Tarrago M; Clouston LJ; Sproules S; Bill E; Bernales V; Gagliardi L; Ye S; Lancaster KM; et al. Enhanced Fe-Centered Redox Flexibility in Fe–Ti Heterobimetallic Complexes. Inorg. Chem 2019, 58 (9), 6199–6214. DOI: 10.1021/acs.inorgchem.9b00442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (33).Eisenhart RJ; Clouston LJ; Lu CC Configuring Bonds between First-Row Transition Metals. Acc. Chem. Res 2015, 48 (11), 2885–2894. DOI: 10.1021/acs.accounts.5b00336. [DOI] [PubMed] [Google Scholar]
  • (34).McWilliams SF; Brennan-Wydra E; MacLeod KC; Holland PL Density Functional Calculations for Prediction of 57Fe Mössbauer Isomer Shifts and Quadrupole Splittings in β-Diketiminate Complexes. ACS Omega 2017, 2 (6), 2594–2606. DOI: 10.1021/acsomega.7b00595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Römelt M; Ye S; Neese F Calibration of Modern Density Functional Theory Methods for the Prediction of 57Fe Mössbauer Isomer Shifts: Meta-GGA and Double-Hybrid Functionals. Inorg. Chem 2009, 48 (3), 784–785. DOI: 10.1021/ic801535v. [DOI] [PubMed] [Google Scholar]
  • (36).Jarrett JT Iron–Sulfur Clusters. In Encyclopedia of Biophysics, Roberts GCK Ed.; Springer Berlin Heidelberg, 2013; pp 1153–1156. [Google Scholar]
  • (37).Schwindt MA; Lejon T; Hegedus LS Improved synthesis of (aminocarbene)chromium(0) complexes with use of C8K-generated Cr(CO)52-. Multivariant optimization of an organometallic reaction. Organometallics 1990, 9 (10), 2814–2819. DOI: 10.1021/om00160a033. [DOI] [Google Scholar]
  • (38).Creager S 3 - Solvents and Supporting Electrolytes. In Handbook of Electrochemistry, Zoski CG Ed.; Elsevier, 2007; pp 57–72. [Google Scholar]
  • (39).Kaplan L; Kester WL; Katz JJ Some Properties of Iron Biscyclopentadienyl. J. Am. Chem. Soc 1952, 74 (21), 5531–5532. DOI: 10.1021/ja01141a518. [DOI] [Google Scholar]
  • (40).Andrews CG; Macdonald CLB The unusual reactions of indium(I) trifluoromethanesulfonate with some first row metallocenes and the structure of “indium(II) cyclopentadienide”. J. Organomet. Chem 2005, 690 (23), 5090–5097. DOI: 10.1016/j.jorganchem.2005.03.028. [DOI] [Google Scholar]
  • (41).Evans DF 400. The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance. J. Chem. Soc 1959, (0), 2003–2005, 10.1039/JR9590002003. DOI: 10.1039/JR9590002003. [DOI] [Google Scholar]
  • (42).CrysAlisPro 1.171.40.53: Rigaku Oxford Diffraction, Rigaku Corporation, Oxford, UK: (2019). [Google Scholar]
  • (43).Bruker SAINT v8.38A; Bruker AXS Inc.: Madison, WI, 2014. [Google Scholar]
  • (44).SCALE3 ABSPACK v1.0.7: an Oxford Diffraction program; Oxford Diffraction Ltd: Abingdon, UK, 2005. [Google Scholar]
  • (45).Krause L; Herbst-Irmer R; Sheldrick GM; Stalke D Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Crystallogr 2015, 48, 3–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Sheldrick G SHELXT - Integrated space-group and crystal-structure determination. Acta Crystallogr. A 2015, 71, 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Sheldrick G Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Grimme S; Antony J; Ehrlich S; Krieg H A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys 2010, 132 (15), 154104. DOI: 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  • (49).Grimme S; Ehrlich S; Goerigk L Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem 2011, 32 (7), 1456–1465. DOI: 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
  • (50).Grimme S Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem 2006, 27 (15), 1787–1799. DOI: 10.1002/jcc.20495. [DOI] [PubMed] [Google Scholar]
  • (51).Neese F Software update: the ORCA program system, version 4.0. WIREs Comput. Mol. Sci 2018, 8 (1), e1327. DOI: 10.1002/wcms.1327. [DOI] [Google Scholar]
  • (52).Schäfer A; Huber C; Ahlrichs R Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys 1994, 100 (8), 5829–5835. DOI: 10.1063/1.467146. [DOI] [Google Scholar]
  • (53).Schäfer A; Horn H; Ahlrichs R Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys 1992, 97 (4), 2571–2577. DOI: 10.1063/1.463096. [DOI] [Google Scholar]
  • (54).Weigend F; Ahlrichs R Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys 2005, 7 (18), 3297–3305, 10.1039/B508541A. DOI: 10.1039/B508541A. [DOI] [PubMed] [Google Scholar]
  • (55).Bjornsson R; Neese F; DeBeer S Revisiting the Mössbauer Isomer Shifts of the FeMoco Cluster of Nitrogenase and the Cofactor Charge. Inorg. Chem 2017, 56 (3), 1470–1477. DOI: 10.1021/acs.inorgchem.6b02540. [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