Abstract
SCS pincer ligands have an interesting combination of strong-field and weak-field donors, a coordination environment that is also present in the nitrogenase active site. Here, we explore the electronic structures of iron(II) and iron(III) complexes with such a pincer ligand, bearing a monodentate phosphine, thiolate S donor, amide N donor, ammonia, or CO. The ligand scaffold features a proton-responsive thioamide site, and the protonation state of the ligand greatly influences the reduction potential of iron in the phosphine complex. The N–H bond dissociation free energy, derived from a square scheme, is 56 ± 2 kcal/mol. EPR spectroscopy and SQUID magnetometry measurements show that the iron(III) complexes with S and N as the fourth donors have an intermediate spin (S = 3/2) ground state with large zero field splitting, and X-ray absorption spectra show high Fe–S covalency. The Mössbauer spectrum changes drastically with the position of a nearby alkali metal cation in the iron(III) amido complex, and DFT calculations explain this phenomenon through a change between having the doubly-occupied orbital as dz2 or dyz, as the former is more influenced by the nearby positive charge.
Graphical Abstract
INTRODUCTION
The organometallic chemistry of iron has been dominated by strong-field supporting ligands such as CO, CN, and phosphines, and by macrocyclic N ligands like porphyrins.1–6 The active sites of hydrogenase enzymes incorporate S-based ligands, and these are low-spin due to the influence of carbonyl and cyanide donors.7 However, the interesting reactions of nitrogenase enzymes, and a new generation of low-valent iron catalysts, have predominantly weak-field ligands and have led to increasing attention to organometallic iron complexes with higher spin states.8–11 We focus here on iron coordination environments that result from a mixture of C and S donors – choices that are particularly compelling since the six “belt” iron atoms in the iron-molybdenum cofactor (FeMoco) of nitrogenase have a mixed C/S coordination sphere.12–14 This unusual combination of potentially strong-field C donors and weak-field S donors could lead to changes in spin states during catalysis, which has been linked to changes in barriers and selectivity.15–18 Thus, the study of C- and S-ligated iron has relevance for both fundamental coordination chemistry and bioinorganic mechanisms.19–21 However, CO-free iron complexes with supporting ligands that coordinate through only carbon and sulfur donors are rare.22–28 Of these examples, only one multidentate C/S ligand is known to support N2 binding.26 In addition, Qu has provided important studies on Cp*-supported iron dimers bridged by dithiolates, and this C/S ligand sphere can bind and facilitate the reduction of nitrogenase-relevant NxHy substrates.23–25, 29
Our group’s recent work on SCS systems began with a dithiolate ligand having a central arene that can interact with iron through backbonding into its π-system in low oxidation states.26, 30 While this hemilabile interaction stabilizes the reduced complexes, complete dissociation of the SCS ligand occurred in the presence of Brønsted acids, and backbonding into N2 competed with backbonding into the arene. In order to address this instability, we then moved to tridentate pincer scaffolds in which the carbon forms a σ-bond to the iron center. Namely, we synthesized a new SCS pincer comprised of an N-heterocyclic carbene (NHC) ligand with two bulky thiolate arms, and isolated a number of iron complexes with this scaffold.28 The geometries at iron sites with this supporting ligand were sometimes tetrahedral and sometimes square planar, indicating significant flexibility in the core. However, this NHC ligand was also sensitive to acids, and the ligand synthesis involved time-consuming protection and deprotection steps.
Here, we study the first iron complexes of pincer ligands containing a central aryl C donor and two S-donor arms. SCS pincers of this type have been primarily studied in complexes of group 10 metals (Chart 1, top), but only have precedent in group 8 metal complexes of ruthenium.31–35 Thioamide arms are compelling, because they are easily prepared from readily available precursors and offer the potential for proton responsiveness.36–38 There is no precedent for iron complexes with thioamide-aryl ligands, although there is a related iron SNS pincer complex with a central pyridine (Chart 1, lower left).39–40
Chart 1.
Typical palladium SCS pincer complexes and closely-related iron complexes with SNS and SCS pincers. In the bottom pictures, the square represents a coordination site with various ligands.
Thioamides have two tautomers, which form the same anion upon deprotonation (Figure 1, right). Most often, the thioamide coordinates as an iminothiolate through the S donor rather than the potential N donor as a result of the relative weakness of the C=S π bond and the better overlap of the S atom with metal orbitals.41 After coordination, protonation at the N atom gives a formally neutral thione donor that maintains the M–S interaction (Figure 1, brackets). Even as a formally neutral ligand, the thione C–S bond is significantly polarized toward a Cδ+–Sδ– form, which affords some anionic character at S.42 The ability to tune the donor properties of iron-coordinated sulfur via protonation/deprotonation of the supporting ligand could be used to probe how protonation state affects the redox potential, spectroscopic characteristics, and reactivity of the iron center.
Figure 1.
Illustration of thioamide tautomerization and deprotonation to form an anionic donor.
The potential for proton responsiveness is also relevant to nitrogenase mechanisms. The Thorneley-Lowe kinetic scheme for nitrogenases proposes that one proton is transferred to the FeMoco with each reduction step,43 but the location of these protons has been controversial. Computational studies have showed that there are many potential sites of protonation on the FeMoco, including the sulfides,44–47 the Mo-coordinated homocitrate,45, 48 and even the carbide.46–47, 49 Studies on multi-site proton-coupled electron transfer (PCET) have demonstrated that various protonation sites can influence the redox potentials and bond dissociation free energies (BDFE).50–53 These inspirations motivated us to pursue thermochemical studies on well-characterized (SCS)Fe complexes with proton-responsive ligands to examine the effects of distant protonation events on potentially biologically-relevant iron sites.
In the present study, we describe a new ligand scaffold that incorporates thioamides and an anionic aryl group to mimic the sulfur and carbon coordination environment found in the FeMoco. The ligand is easy to prepare on a useful scale, and the stability of the iron complexes is highlighted by the ability of an iron(III) complex to be handled in air and water. Although no iron complex of the new SCS ligand was observed to bind N2, we use EPR and Mössbauer spectroscopy alongside magnetometry measurements to elucidate the unusual electronic structures of complexes with different exogenous donors, including biologically-relevant amide, thiolate, ammonia, and CO. In the amide complex, the electronic structure is influenced by the presence of nearby cations, as demonstrated by computations and spectroscopy. We also demonstrate the ability of the thioamide arm to be proton responsive, and quantify the connection between protonation state and redox state of iron.
RESULTS
Ligand Synthesis and Metalation.
Using a procedure modified from the literature,54 isophthalic acid was treated with thionyl chloride followed by 2,6-diisopropylaniline, which provided the diamide (Scheme 1). Reaction of this compound with P2S5 in toluene at 100 °C yielded the bis(thioamide) 1, in an overall yield of 69% in two steps from commercial starting materials. Treating 1 with a slight excess of Fe(PMe3)4 in Et2O at room temperature led to effervescence and a color change from yellow-brown to dark green, and the iron(II) complex 2 was isolated in 96% yield (Scheme 1).55–56
Scheme 1.
Synthesis and metalation of SCS pincer ligand 1. Dipp = 2,6-diisopropylphenyl.
Crystallization of 2 gave green blocks suitable for X-ray diffraction (Figure 2, top). The diffraction data showed an octahedral iron site with a meridional pincer ligand as expected. The Fe−C bond length is 1.9502(18) Å, and the average Fe−S length is 2.268(9) Å. The three Fe−P bonds are only slightly different, with a length of 2.2484(3) Å for the phosphine trans to C and lengths of 2.2553(4) Å and 2.2608(4) Å for the phosphines cis to C. There was disorder at the thioamide N−H site, which was satisfactorily modeled with each thioamide having 50% hydrogen atom occupancy, consistent with single protonation of the supporting ligand in 2.
Figure 2.
Molecular structures of monoprotonated iron(II) complex 2 and fully deprotonated iron(III) complex 4, plotted as 50% probability thermal ellipsoids. All hydrogen atoms except those bonded to N are hidden for clarity, and each N−H site is modeled as having 50% hydrogen atom occupancy.
The Mössbauer spectrum of 2 has a doublet with δ = 0.21 mm/s and |ΔEQ| = 1.25 mm/s (Figure S46). Proton NMR spectra in either C6D6 or THF-d8 show signals in the range expected for a diamagnetic complex of low-spin iron(II). The most downfield resonance integrates to one proton and disappears upon addition of D2O, and therefore it is assigned to the thioamide proton (Figure S5). Based on the NMR spectra, 2 is unsymmetric in solution, and it is best described has having one neutral thioamide S donor and one anionic iminothiolate donor. The solid-state IR spectrum of 2 shows a broad peak centered at 3336 cm–1 (Figure 3, red trace), which is distinct from the analogous signal in free ligand 1 at 3136 cm–1 (Figure 3, gray trace), consistent with an N–H stretching vibration.
Figure 3.
Solid state IR spectra of neutral ligand 1 and monoprotonated ferrous complex 2, ferric iminothiolate complex 4, and ferrous iminothiolate complex 3-Na in the range of 4000 to 2000 cm−1. The bands assigned to N−H stretches (3136 and 3336 cm−1) in the spectra of 1 and 2 are labeled.
Electron and Proton Transfer in Phosphine Complexes.
We serendipitously discovered that neutral, fully-deprotonated iron(III) compound 4 could be prepared in one pot from 1 and Fe(PMe3)4 by simply exposing the crude product 2 to air after metalation was complete. Proton NMR and UV-visible spectroscopy of air-exposed 2 in situ demonstrates rapid and complete conversion to 4, with no monoprotonated iron(III) transient species observed. This one-pot procedure afforded 4 in 88% yield from 2, and this method was used to prepare 4 for subsequent experiments.
Additional proton and electron transfer reactions were carried out on 2 and 4 (Scheme 2). Treating the neutral compound 2 with stoichiometric sodium bis(trimethylsilyl)amide provided anionic 3-Na. Oxidation of 3-Na using AgPF6 regenerated 4. Alternatively, 4 could be formed in one step by hydrogen atom abstraction from 2 using TEMPO. It was also possible to generate the potassium analogue of 3-Na (3-K) by treatment of neutral iron(III) compound 4 with stoichiometric KC8. Both 3-Na and 3-K have sharp peaks in their 1H NMR spectra suggesting low-spin iron(II), and their structures were inferred based on these spectra (Figures S7 and S9).
Scheme 2.
Conversion of 2 to 4 using stepwise and concerted H atom removal. Yields were determined by 1H NMR using an internal standard.
The X-ray crystal structure of the fully-deprotonated iron(III) complex 4 (Figure 2, bottom) shows an Fe−C bond length of 1.9413(12) Å, which is indistinguishable from the analogous distance in the monoprotonated iron(II) complex 2. The average Fe−S length is 2.23(2) Å in 4, which is slightly shorter than in 2. Again, all three Fe−P bond distances are different, with the phosphine trans to C having the shortest bond of 2.2692(6) Å while the cis phosphines have lengths of 2.3007(6) Å and 2.3050(5) Å. The average Fe−P bond in 4 is shorter than in 2 (2.23(1) Å vs. 2.255(4) Å). Thus, there is a general contraction of bonds in ferric 4 when compared to ferrous 2, consistent with 4 being more oxidized.
As expected for complexes lacking protonated thioamides, 3-Na and 4 do not have any IR bands above 3050 cm–1 that correspond to N–H stretching vibrations (Figure 3, blue and green traces). Proton NMR spectra of ferric 4 contain very broad peaks between −20 and +12 ppm (Figure S11), and the 31P NMR spectrum is featureless, as expected for nuclei directly bound to a paramagnetic metal. The relative integrations of the evident peaks in the 1H NMR spectra of 3-M in THF-d8 are consistent with C2v symmetry. Taken together, the IR and NMR data support double deprotonation of the thioamide groups of the supporting ligand in 3-Na and 4.
Cyclic voltammetry (CV) studies of 4 showed a reversible Fe3+/2+ couple at E1/2 = −1.42 V vs. Fc+/Fc in THF (Figures S60–61). Next, we tested the ability of different organic bases to deprotonate 2, using changes in the 31P NMR chemical shifts to determine the ratio of 2 and deprotonated 2 since the conjugate base and acid are in rapid equilibrium. Addition of neither Et3N (pKaH = 12.5 in THF) nor DBU (pKaH = 16.9 in THF) gave changes in the NMR spectra of 2.57 Addition of 1.10 equiv of triazabicyclodecene (TBD, pKaH = 21.0 in THF)57 to 2, on the other hand, gave a 31P NMR spectrum after equilibration with peaks at 12.1 ppm and 10.8 ppm. These shifts are 4.8 ppm upfield from the 31P resonances of 3-K but 1.4 ppm downfield of 31P resonances of 2 (Figure S77). The presence of both the protonated and deprotonated species in this equilibrium indicates that the pKa of 2 is 22 ± 1 in neat THF-d8. An equivalent NMR experiment in 0.3 M [NnBu4][PF6] in THF-d8 (the electrolytic solution used for CV) indicated a pKa of 21 ± 1 (Figure S78). These data enable us to define the bond dissociation free energy (BDFE) of the N–H bond using the Bordwell equation, with CG(THF) = 60.4 ± 2 kcal/mol.58–59 This analysis gave a BDFE(N–H) of 56 ± 2 kcal/mol (Scheme 3).
Scheme 3.
Thermodynamic values used to calculate the BDFE.
Other redox and protonation events were tested as well. Electrochemical oxidation of 2 is irreversible under the same CV conditions used for 4 (Figure S59). To test whether 4 could be protonated, it was treated with one equiv of [H(OEt2)2][BF4] in THF. A dark brown solid precipitated, and its IR spectrum showed a band at 3250 cm−1, which is much lower than the N–H stretching frequency of 2 at 3336 cm−1 (Figure S79). Though further studies on putative protonated 4 was prevented by its insolubility, the IR spectrum of acid-treated 4 may indicate the feasibility of thioamide protonation.
Spectroscopy of Phosphine Complexes.
The Mössbauer signal for 2 (δ = 0.21 mm/s and |ΔEQ| = 1.25 mm/s, Figure S46) is similar to those for 3-Na (δ = 0.24 mm/s and |ΔEQ| = 0.94 mm/s, Figure S47) and 3-K (δ = 0.25 mm/s and |ΔEQ| = 1.05 mm/s, Figure S48), consistent with low-spin iron(II) complexes. Solid 4 has a lower isomer shift of 0.16 mm/s and much larger quadrupole splitting of 3.45 mm/s, and its doublets are highly asymmetric, with ΓL = 0.57 mm/s, ΓR = 0.32 mm/s. To determine the spin state of 4, solid state and C6D6 solution magnetic moments were obtained. Both measurements gave μeff = 1.6 μB at 298 K, which indicates a ground-state spin of S = 1/2.
Spectroscopy and DFT calculations using B3LYP/ZORA-def2-TZVP were further employed to reveal the electronic structure of 4. The EPR spectrum of 4 in 2-methyltetrahydrofuran (MeTHF) at 77 K was modeled as an S = 1/2 spin system with g = [2.200, 2.120, 1.998] and coupling to three 31P nuclei: two with |A| = [71, 83, 71] MHz and one with |A| = [0, 37, 37] MHz (Figure 4). Mulliken spin population analysis showed spin populations of −0.036 on the axial phosphorus atoms and −0.019 on the equatorial phosphorus, in agreement with the fit to two equivalent 31P nuclei having larger A than a third 31P nucleus. The spin population analysis also showed an averaged spin density of 0.031 on the two ligand nitrogen atoms, though the smaller nuclear magnetic moment of 14N relative to 31P likely contributed to smaller, unresolvable 14N hyperfine splitting, which was modeled by anisotropic H-strain of [66, 27, 0] MHz in the simulation.
Figure 4.
X-Band EPR spectrum of 4 at 77 K (1.0 mM in 2-MeTHF) measured with a microwave frequency of 9.384 GHz and power of 1.779 mW. Parameters for simulation: g1 = 2.200, |A1(31P)| = 71, 71, 0 MHz; g2 = 2.120, |A2(31P)| = 83, 83, 37 MHz; g3 = 1.998, |A3(31P)| = 71, 71, 37 MHz; HStrain = [66, 27, 0] MHz. Inset: Mulliken spin densities are shown next to the respective phosphorus atoms (shown in magenta). Hydrogen atoms and dipp groups are omitted for clarity.
Synthesis of Phosphine-Free SCS Iron Complexes.
When iron(III) species 4 was treated with a slight excess of triphenylborane in THF, the color of the reaction mixture changed from green to orange, and precipitation of Me3P–BPh3 was observed (Scheme 4). Crystallization from Et2O gave 5-Et2O in 58% yield (Figure S70). Dissolving 5-Et2O in THF and removing solvent under vacuum several times led to the formation of 5-THF (Figure S69). We refer to these compounds collectively as 5-Solv because of their similarity and ability to interconvert as a function of solvent.
Scheme 4.
Phosphine abstraction with triphenylborane. Dipp = 2,6-diisopropylphenyl.
Both compounds 5-Solv are dimeric in the solid state (Figure 5), with each iron center in a square pyramidal geometry with a coordinated solvent molecule occupying the axial site. The dimers adopt a butterfly shape in which the two iron centers are bridged by one sulfur from each ligand. In 5-Et2O, the planes containing each pincer intersect at 75°. Comparing the two arms of the pincer ligand in 5-Et2O, the bridging Fe–S bond is longer than the non-bridging Fe–S by 0.08(1) Å. Other bond metrics of 5-Et2O are comparable to those in phosphine precursor 4, and the long C–S (1.78(3) Å average) and short C–N bonds (1.26(1) Å average) indicate two anionic iminothiolate donors. The metrical parameters of 5-THF are nearly identical to those of 5-Et2O, with Fe−C bond lengths of 1.961(2) Å (5-THF) and 1.966(3) Å (5-Et2O), Fe−Sbridging bond lengths of 2.2557(7) Å (5-THF) and 2.278(1) Å (5-Et2O), and Fe−Snon-bridging bond lengths of 2.1797(7) Å (5-THF) and 2.187(1) Å (5-Et2O).
Figure 5.
Molecular structure of 5-Et2O plotted at 50% probability thermal ellipsoids, showing its dimeric butterfly shape from side and top views. Dipp groups and Et2O are shown in wireframe, and hydrogen atoms are omitted.
Mössbauer spectra of solid 5-Et2O and 5-THF are also comparable, with parameters of δ = 0.34 mm/s and |ΔEQ| = 3.91 mm/s for 5-Et2O and δ = 0.34 mm/s, |ΔEQ| = 3.73 mm/s for 5-THF (Figures S50 and S51). Mössbauer spectra of 5-THF differ, however, between the solid state and a frozen THF solution. Specifically, there is a higher isomer shift of 0.45 mm/s for the frozen-solution spectrum, which may suggest that the dimer is broken up with THF replacing the bridging sulfur (Figure S80). The frozen solution also shows an increased quadrupole splitting and marked doublet asymmetry (|ΔEQ| = 4.26 mm/s, ΓL = 0.73 mm/s, ΓR = 0.49 mm/s). The 1H NMR spectrum of 5-Et2O in THF-d8 shows six broad signals between 12 and −81 ppm (Figure S12). The number and integration of the peaks is consistent with C2v symmetry; a seventh expected peak integrating to 4H was not resolved, likely due to broadening. This is also consistent with the dimer breaking up, although it is also possible that there is some other dynamic phenomenon. The 1H NMR spectra of 5-Solv in C6D6 have at least 12 peaks with considerable broadness, which prevented further analysis but points to a lowered symmetry in the dimer.
Cyclic voltammetry of 5-Solv showed three irreversible reduction events (Figure S62). An attempt to reference the reduction potentials to a ferrocene internal standard was unsuccessful due to reactivity with Fc and Fc+. The lack of electrochemical reversibility may be due to large structural changes occurring at different oxidation states. Undeterred, we proceeded to stoichiometrically reduce 5-Solv with 1 equiv of KC8 per iron at −78 °C in THF, which resulted in a deep purple solution. Crystallization from diethyl ether at −40 °C gave iron(II) complex 6 in 46% yield. Its X-ray crystal structure, shown in Figure 6, shows a tetrameric assembly in the solid state, with K+ bridging S atoms and Dipp groups.
Figure 6.
Molecular structure of 6. Thermal ellipsoids are displayed at the 50% probability level. Dipp groups are shown in wireframe and hydrogen atoms are omitted. Two-thirds of the asymmetric unit is shown.
The 1H NMR spectrum of 6 in THF-d8 shows at least 19 resonances between −50 and +15 ppm, reflecting C1 symmetry in solution. We were unable to assign resonances to specific proton environments due to peak broadness and overlapping in the 0 to 8 ppm region (Figure S13). Despite the complexity of the 1H NMR spectrum and crystal structure, the solid-state Mössbauer spectrum of 6 (Figure S52) has a single quadrupole doublet with an isomer shift similar to that of the starting material (0.34 mm/s for 6 vs. 0.35 mm/s for 5-Et2O), but a smaller quadrupole splitting (1.87 mm/s for 6 vs. 3.91 mm/s for 5-Et2O). The Mössbauer spectrum suggests that all iron sites are equivalent.
Next, we tested 6 for the ability to bind N2 at low temperatures using variable-temperature UV-visible spectroscopy. Negligible spectral changes were observed between 25 and −100 °C in THF, Et2O, or toluene (Figure S81a–S81c). Addition of stoichiometric amounts of 18-crown-6 slightly shifted the absorption maxima, but the resulting species also did not exhibit notable temperature-dependent spectral changes (Figure S81d–S81f). Thus, it does not appear that 6 undergoes speciation changes or N2 binding at these concentrations and temperatures. Further reduction of the iron(II) tetrameter 6 led to a silent NMR spectrum and a broad, featureless UV-Vis spectrum. The products formed in this reaction are unknown, and attempts to isolate or identify the species present were unsuccessful.
Synthesis of Monomeric Iron Complexes With S and N Donors.
To study monomeric complexes of our SCS iron framework in the absence of strong phosphine donors, we next treated 5-Et2O with ligands containing S and N donors, which gave the products in Scheme 5 within 1 hour at room temperature. Their syntheses, NMR spectra, and IR spectra are described here, and their magnetism, EPR spectra and electronic structures are described in following sections.
Scheme 5.
Synthesis of complexes 7–10. L = triply deprotonated trianionic ligand 1; Ar = 2,6-dimesitylphenyl.
Addition of a bulky thiolate (potassium 2,6-dimesitylphenylthiolate) to 5-Et2O in THF generated complex 7 (Scheme 5, top). Addition of 18-crown-6 followed by crystallization furnished 7-crown (Figure S72). 7-Crown displays an axially coordinated THF molecule and a nearly square pyramidal geometry with τ5 = 0.08. The Fe−C, average Fe−SPincer, and Fe−SAr, bond lengths are 2.0006(17) Å, 2.244(6) Å, and 2.2758(5) Å, respectively. Addition of Et2O, arenes, or alkanes to solid 7 prior to the addition of 18-crown-6 led to a color change from orange to green, indicating the formation of the THF-free analogue (8) in non-coordinating solvents in 97% yield. Compound 8 was crystallized from toluene in the presence of 18-crown-6 as 8-crown (Figure 7, top, and Figure S73). The geometry is distorted from square planar, with τ4 = 0.20. The Fe−SAr bond distance is 2.2448(8) Å, which is shorter than the Fe−SAr bond in 7-crown (Table 1). Unlike the structure of the THF adduct 7-crown, one mesityl group of THF-free 8-crown is twisted to cover an axial iron site; however, the long 3.06 Å distance between iron and the arene centroid is inconsistent with a direct electronic interaction. We previously observed a similar orientation of this thiolate ligand in another SCS iron complex.28 We were unable to crystallize four-coordinate thiolate 8 without crown ether. The 1H NMR spectrum of 8 in THF-d8 has ten broad peaks between 15 and −63 ppm (Figure S14). Upon dissolution in THF-d8, the color of 8 changes from dark brown to red, indicating that the THF-adduct 7 is likely formed. The number and integration of the signals suggests C2v symmetry, which implies that the crystallographically-observed lack of symmetry caused by the bulky aryl thiolate is not preserved in THF-d8 solution.
Figure 7.
Molecular structures of 8-crown, 9, and 10 at 50% probability thermal ellipsoids. Hydrogen atoms are excluded for clarity, with exception of the NH3 hydrogen atoms of 10. The outer-sphere potassium ion in 8-crown is not shown.
Table 1.
Comparison of Fe−C, Fe−S, and Fe−X (X = SAr, N(TMS)2, NH3) bond lengths (in Å) in 7-crown, 8-crown, 9-crown, and 10. Ar = 2,6-dimesitylphenyl. Two sets of lengths for one parameter represent the two molecules in the unit cell.
Complex | Exogenous Ligand (X) | Fe−C | Average Fe−SPincer | Fe−X |
---|---|---|---|---|
7-crown | SAr | 2.0006(17) | 2.244(6) | 2.2758(5) |
8-crown | SAr | 1.999(3) | 2.250(5) | 2.2448(8) |
9-crown | N(TMS)2 | 2.002(6) 2.010(5) |
2.234(5) 2.241(7) |
1.882(5) 1.902(4) |
10 | NH3 | 1.973(3) 1.971(3) |
2.24(1) 2.25(1) |
2.065(3) 2.084(3) |
Amide complex 9 was synthesized in 87% yield by treating 5-solv with KN(TMS)2, and it was crystallized with and without 18-crown-6 to give structures of 9-crown (Figure S75) and 9 (Scheme 5, middle, and Figure S74), respectively. Crown-free 9 is unstable as a solid at room temperature and in solution even at −40 °C, as evidenced by the appearance of multiple new peaks between 11 and −43 ppm in its NMR spectrum after a few days (Figure S17). Once this unknown impurity formed, we have been unable to remove it by recrystallization or washing. 9-Crown, however, is more stable. The crystallographic structures of 9 and 9-crown lack axial solvent coordination (despite 9 being crystallized from THF), and their color is the same in coordinating and non-coordinating solvents. These observations suggest that solvent coordination to the N(TMS)2 adduct is unfavorable, possibly because the TMS groups block the axial sites.
The crystallographic structure of 9 displays a square planar molecular geometry at iron with τ4 = 0.09–0.11. Fe–N bond lengths in 9 are Fe1–N14 = 1.915(3) and Fe2–N18 = 1.899(3), which are distinguishable. Interestingly, the structure of 9 shows a potassium cation lying in the pincer plane, with a K1−S11 distance of 3.133(1) Å and K2−S15 distance of 3.126(1) Å. The Fe−S bond lengths on the side with potassium are 2.2459(9) Å (Fe1) and 2.2463(9) Å (Fe2), which are indistinguishable. The Fe−S bond length is, however, significantly shorter on the side of the molecule without the close S−K contact, with a difference of 0.0065(12) Å at Fe1 and 0.0080(12) Å at Fe2. Metrical parameters for 9-crown are shown in Table 1. The 1H NMR spectra of 9 and 9-crown in THF-d8 have eight and nine, respectively, paramagnetically-shifted peaks between 19 and −68 ppm (Figures S16 and S18), consistent with C2v symmetry. Their chemical shifts are nearly the same.
Treating 5-solv with 1 equiv of NH3 gas in THF furnished ammine complex 10 in 86% yield (Scheme 5, bottom). In the crystallographic structure, there are two molecules of 10 in the asymmetric unit, and each has an axially-coordinated THF that gives a distorted square pyramidal geometry with τ5 = 0.09–0.14. The Fe−N bond lengths are 2.084(3) Å (Fe1) and 2.065(3) Å (Fe2). The molecule containing Fe2 shows a hydrogen bond between one NH3 proton and one uncoordinated THF. Hence, the shorter Fe−N length in Fe2 may be due to the hydrogen bond, which increases the basicity of the NH3. While the Fe−S bond lengths of 10 are similar to those in 8-crown and 9-crown, the Fe−C bond length is significantly shorter in 10 (Table 1). This difference may arise because the thiolate and amido ligands exert a larger trans influence than the ammine.60,61 The solid-state IR spectrum of 10 shows four weak bands at 3352, 3299, 3236, and 3159 cm−1 that are assigned as N–H stretches (Figure S44). Their frequencies are not suggestive of any significant coordination-induced N−H bond weakening. The 1H NMR spectrum of 10 in THF-d8 has eight resonances between 175 ppm and −78 ppm, consistent with C2v symmetry (Figure S19). Since the complex is Cs symmetric with THF coordination on one face, the spectrum in THF-d8 could indicate fast THF exchange on both sides of the complex. Its spectrum in C6D6 has severe broadening compared to THF-d8, and there are at least 15 resonances (Figure S20). Assuming that coordinated THF has two different proton environments and that NH3 rotation on the NMR timescale renders its protons equivalent, a Cs symmetric geometry (with the mirror plane perpendicular to the pincer) would predict 14 peaks. Thus, the number of peaks in the C6D6-solvated NMR spectrum suggests 10 has no symmetry in C6D6, though the reason for the low symmetry is not obvious.
Synthesis and Characterization of an Iron(II) Carbonyl Complex.
Treating a THF solution of iron(III) complex 5-Solv with 1 atm of CO gave little reaction, with 85% unreacted 5-Solv remaining after 2 hours at room temperature, as judged by NMR spectroscopy. In contrast, stirring iron(II) complex 6 under 1 atm of CO afforded 11 in 79% yield after 1 hour (Scheme 6). Analytically pure material was isolated by removing THF under vacuum and washing the pale-yellow solids with Et2O. Material generated in this way gave a Mössbauer spectrum with a quadrupole doublet at δ = 0.01 mm/s and |ΔEQ| = 0.99 mm/s (Figure S58). Low isomer shifts such as that of 11 are frequently observed in iron carbonyl complexes, reflecting the withdrawal of d-electron density from the iron nucleus by backbonding.62 The 1H and 13C NMR spectra of 11 show 7 and 14 narrow peaks, respectively (Figures S21 and S22). These spectra are consistent with a C2v-symmetric diamagnetic species. The two most-downfield 13C signals have a relative integration of 2:1 and are assigned to the CO ligands. They have comparable chemical shifts to the CO signals in an SNS-supported iron(II) tricarbonyl complex.39
Scheme 6.
Synthesis of tricarbonyl complex 11. L = triply deprotonated trianionic ligand 1.
The solid-state IR spectrum of 11 shows three bands in the region expected for CO stretching at 2078, 2013, and 1988 cm–1 (Figure S45). To compare our experimental spectrum with theoretical IR spectra of SCS-supported complexes with either two or three CO ligands, we used DFT calculations (BP86/ZORA-def2-TZVP) to optimize and predict the CO stretching frequencies of [LFe(CO)2]− and [LFe(CO)3]−. The model with two CO ligands predicted two IR-active bands of equal intensity, while the model with three CO ligands predicted stretching frequencies whose number and relative intensities matched the experimental spectrum (Figures S84 and S85). Thus, we assign the bands at 2078, 2013, and 1988 cm–1 to CO stretching vibrations in an octahedral tricarbonyl complex.
X-Ray Absorption Spectra of Thiolate Complex 8.
To gain insight into the Fe–S bonding in 8-crown, which has iron(III) in the most biomimetic S3C coordination sphere, Fe K-edge X-ray absorption spectra were measured. The spectrum showed two pre-edge features at ca. 7112 and 7114 eV. TD-DFT calculations reproduced the pre-edge features of the experimental spectrum (Figure 8, bottom). Frontier quasi-restricted orbital (QRO) analysis showed three singly-occupied orbitals of primarily Fe 3d character with < 10% S 3p mixing and a LUMO composed of 39.3% Fe 3d character and 21.3% S 3p character (Figure 8, top). The greater covalency between the Fe and S ligands in the LUMO is responsible for its increased energy. Based on the QRO analysis, the first pre-edge feature at 7112 eV is assigned to excitations of Fe 1s electrons into the three nearly-degenerate singly-occupied orbitals, while the second feature at 7114 eV is assigned to an Fe 1s to LUMO excitation. As expected for iron in the +3 oxidation state, the large contributions of ligand S orbitals to the frontier QROs indicate a large degree of Fe–S bond covalency.
Figure 8.
Top: Frontier MO diagram of 8. Energies were obtained from the QROs (calculated using the B3LYP level of theory with the CP(PPP) basis set on Pd and ZORA-def2-TZVP(-f) on all other atoms). Ligand orbitals are labeled in red. Orbitals are plotted at an isovalue of 0.03 au with hydrogen atoms removed for clarity. Bottom: Overlay of the experimental (red) and TD-DFT calculated (black) Fe K-edge spectra for 8-crown. Black sticks represent individual transitions. The calculation employed the CP(PPP) basis set for Fe and the ZORA-def2-TZVP(-f) basis set for all other atoms.
Electronic Structures of S and N Donor Complexes from Magnetometry and EPR Spectroscopy.
The solution magnetic moments of the iron(III) complexes were measured at 298 K in THF-d8, giving μeff(7/8) = 4.0 ± 0.1 μB, μeff(9) = 4.1 ± 0.1 μB, and μeff(10) = 4.1 ± 0.1 μB. These moments are consistent with a well-isolated S = 3/2 ground state. Note that in THF solution, thiolate complex 8 is in equilibrium with its THF adduct 7. The solid-state magnetic susceptibilities of 8, 9, and 10 are each consistent with an S = 3/2 ground state and little mixing of excited states (Figures S63–S65). To further evaluate the zero-field splitting (ZFS) parameters, we measured the low-temperature solution X-band EPR spectra of each in MeTHF and magnetization curves at ≤ 10 K under field strengths of 1 T to 7 T. The data were fit within the constraints of the usual spin Hamiltonian for an S = 3/2 spin system,
Variable-temperature (3 K to 10 K) magnetization data of thiolate complex 8 showed nesting of curves obtained at different fields, with saturation around 1.6 μB (Figure S66, top). These data were fit to g = [2.30, 2.30, 1.89] (giso = 2.16) and D = −28(1) cm−1; an adequate fit could not be obtained with positive D. At 10 K, complex 8 has an extremely broad EPR spectrum with poorly-resolved signals at g ≈ 5.7, 5.1, 4.5, 4.3, 2.2, 2.1, and 1.2 (Figure 9, top). The sharp set of signals near g = 2.00 is assigned to a S = 1/2 impurity (1%) with g = [2.08, 2.03, 2.00]. An acceptable simulation of the remainder of the EPR spectrum was obtained using S = 3/2, g and D values derived from the magnetization data, and E/D of 0.20. D strain of 28 cm−1 (100% of D) was required to account for the broad spectrum, indicating that a wide distribution of ZFS parameters were sampled in the spectrum. EPR spectra with similarly strain-dominated line shapes have been observed for other S = 3/2 iron complexes.63,64 In our system, the crystallographically-observed flexibility of the thiolate ligand may result in a mixture of conformers with varying ZFS (E/D between 0.1 and 0.2).
Figure 9.
X-band (frequency = 9.37 GHz) EPR spectra from 50 to 500 mT of thiolate complex 8 (top, green) at ~10 K, amide complex 9 (middle, purple) at ~5 K, and ammonia complex 10 (bottom, red) at ~10 K as 1 mM solutions in MeTHF. Each spectrum was collected using a 19 G modulation amplitude and a microwave power of either 0.064 mW (8) or 0.020 mW (9 and 10). Fits are plotted as black traces beneath their respective experimental spectrum. Signals marked by asterisks are assigned to impurities, which were modeled as S = 1/2 species for 8 and 10 in 1% relative concentration to the S = 3/2 species. The simulations with +D and −D for 10 are shown; the latter is preferred because the lowest-field component has an intensity that fits better.
Isofield lines from 2 to 10 K for amide complex 9 also show a field dependence, with saturation near 2.0 μB. These magnetization curves were fit to g = [2.15, 2.54, 2.33] (giso = 2.34) and D = −30(2) cm−1 (Figure S66, middle). No reasonable fit was found using D > 0. The EPR spectrum at 5 K shows a sharp, intense resonance at geff = 6.35 with broader signals at geff = 2.55 and 1.57 (Figure 9, middle). These features were simulated as an S = 3/2 spin system with the g and D values from the magnetization, also introducing E/D = 0.333. Similar to the spectrum of 8, D strain of 5 cm−1 (17% of D) was necessary to account for line shapes. Unresolved hyperfine coupling to the amido 14N nucleus may also contribute to broadening. There are small signals at g = 4.4 and 2 that may come from the unknown decomposition product discussed in the synthesis section.
The magnetization curves for ammonia complex 10 show a pronounced field dependence with saturation occurring around 2.7 μB. They were fit to g = [2.01, 2.01, 2.07] (giso = 2.03) and D = ±3.5(1) cm−1 (Figure S66, bottom); in this case the magnetization data did not distinguish the sign of D. At 10 K, the EPR spectrum of 10 shows features at geff = 6.15, 4.75, and 3.40, while signals at 2.36, 2.07, 2.05, and 1.99 are assigned to impurities (Figure 9, bottom). The spectrum was simulated as S = 3/2 with E/D = 0.106 using g from the magnetization fit, and the use of the negative D value gave a better fit to the intensity of the geff = 6.15 component (see both simulations in Figure 9). An S = 1/2 impurity with g = [2.07, 2.05, 1.99] was included in the simulation in 1% relative abundance to the major S = 3/2 signal. A spectrum similar to that of 10 was reported for a square pyramidal ferric intermediate spin complex with E/D = 0.107.65
The magnetization and EPR fit parameters are summarized in Table 2. Our magnetization data suggest that D < 0 for 8 and 9, and the intensity of the component at g = 6.15 in the EPR spectrum of 10 suggests that the sign of D is negative in this compound as well.
Table 2.
Key parameters for the simulations of the EPR spectra and magnetization data of complexes 8-10.
g x , g y , g z | g iso | D | E/D | |
---|---|---|---|---|
8 | 2.30, 2.30, 1.89 | 2.16 | − 28(1) | 0.200 |
9 | 2.15, 2.54, 2.33 | 2.34 | − 30(2) | 0.333 |
10 | 2.01, 2.01, 2.07 | 2.03 | − 3.5(1) | 0.106 |
Mössbauer Spectroscopy of S and N Donor Complexes.
Zero-field Mössbauer spectra of solid thiolate complexes 8 and 8-crown, N(TMS)2 amide complexes 9 and 9-crown, and ammonia complex 10 were recorded at 80 K. The spectra of 8, 9, and 10 are shown in Figure 10, and fit parameters in Table 3. The isomer shifts and large quadrupole splitting values of 8, 8-crown, and 10 are within the expected ranges for intermediate spin iron(III).62 The spectrum of 9, however, shows an isomer shift of only 0.18 mm/s, a markedly low quadrupole splitting of 0.93 mm/s, and narrower lines with Γ = 0.3 mm/s. The Mössbauer parameters of 8-crown (δ = 0.30 mm/s, |ΔEQ| = 4.03 mm/s) and 9-crown (δ = 0.17 mm/s, |ΔEQ| = 0.88 mm/s) are nearly the same as their crown-free versions 8 and 9.
Figure 10.
Zero field Mössbauer spectra of solid 8, 9, and 10 at 80 K. Data are plotted as black circles and residual as gray trace. 8: δ = 0.31 mm/s, |ΔEQ| = 3.88 mm/s, ΓL = 0.61 mm/s, ΓR = 0.54 mm/s. 9: δ = 0.18 mm/s, |ΔEQ| = 0.93 mm/s, ΓL = 0.31 mm/s, ΓR = 0.33 mm/s; there is a 4% impurity with parameters matching 5-Et2O. 10: δ = 0.35 mm/s, |ΔEQ| = 4.28 mm/s, ΓL = 0.55 mm/s, ΓR = 0.43 mm/s.
Table 3.
Comparison between the experimental and calculated Mössbauer parameters for anionic 8, anionic 9, and neutral 10. The sign of the experimental quadrupole splitting is undetermined.
Complex | δ (mm/s) | ΔEQ (mm/s) | ||
---|---|---|---|---|
Exp. | Calc’d | Exp. | Calc’d | |
8 | 0.31 | 0.28 | |3.88| | + 3.97 |
9 | 0.18 | 0.29 | |0.93| | − 4.24 |
10 | 0.35 | 0.33 | |4.28| | + 4.34 |
In order to understand this difference, we calculated the Mössbauer parameters of 8, 9, and 10 (omitting the countercations) using DFT at the B3LYP/ZORA-def2-TZVP level of theory.66–68 These potassium-free calculated parameters were in agreement with the experimental Mössbauer parameters of thiolate complex 8 and ammonia complex 10, but the calculated and experimental values for δ and |ΔEQ| of amide complex 9 diverged by 0.11 mm/s and 3.18 mm/s, respectively (Table 3).
Since the crystal structure of 9 shows a potassium cation only 3.13 Å from one S atom in the SCS pincer, we hypothesized that including K+ may be necessary to properly describe its electronic structure. Thus, we started from the crystallographic structure of 9 and performed Mössbauer calculations on an optimized structure that includes the close K+ (Figure 11A). In the crystal structure, the potassium is also coordinated to a Dipp aryl group of a neighboring molecule of 9, and this second arene was modeled as benzene in the calculations. Calculations on this model gave predicted Mössbauer parameters in agreement with experiment for 9 despite a significant distortion of the pincer plane (Figure S86). Thus, replication of the Mössbauer parameters requires the inclusion of the potassium ion (Figure 11, comparing parts A and B).
Figure 11.
Structures viewed along the N−Fe−C bonds of three models of 9 after BP86/ZORA-def2-TZVP optimization, with their predicted Mössbauer parameters stated below. A: optimized structure with K+. B: optimized structure of the anion without K+. C: optimized structure with [K+(18-crown-6)(THF)2]. Potassium is represented as a purple sphere.
Analogously, we performed Mössbauer calculations on an optimized model of 9-crown, which has a formally outer-sphere potassium with an Fe···K distance of 8.026 Å (Figure 11C). This structure gave δ = 0.16 mm/s and ΔEQ = −0.91 mm/s, in agreement with experimental values for 9-crown (and very close to those of 9). The Löwdin charge on potassium was found to be −0.06 in this structure, suggesting significant delocalization of positive charge onto the coordinated THF and crown ether components.
We were surprised that the effect of potassium was maintained despite the crown ether. To test the isolated effect of a distant and free potassium ion, we changed the computational model. We removed the two THF molecules and the 18-crown-6 on the potassium in optimized 9-crown and repeated the Mössbauer calculations without changing the geometry (Figure S86). This model gave δ = 0.33 mm/s and ΔEQ = −4.08 mm/s, which is much closer to that calculated for potassium-free, anionic 9. In contrast to the structure containing THF and 18-crown-6, the Löwdin charge on potassium was found to be +0.99 in this structure. These results indicate that in 9-crown, the THF and 18-crown-6 may be accepting most of the positive charge of the potassium cation, which then places enough positive charge close to the iron to alter its electronic structure. Charge reorganization, albeit in varying degrees, has been implicated as a stabilizing factor in the formation of alkali-crown complexes in previous theoretical studies.69–70
To compare the electronic structures of the computational models whose Mössbauer parameters match experiment to those that do not, we plotted the frontier quasi-restricted orbitals (QROs) for models A (reproduces experimental Mössbauer parameters; optimized with K+) and B (does not reproduce experimental Mössbauer parameters; optimized without K+) in Figure 12. The QROs show that the shape of the doubly-occupied d orbital (DOMO) for model A resembles dz2 while the doubly-occupied d orbital for model B resembles dyz. The dz2 DOMO has 3.7% Fe s character, while the dyz DOMO has 0.0% Fe s character, explaining the significant change in the isomer shift. The ordering of the dxz and dxy orbitals is also switched, but the dx2−y2 orbital is always higher in energy than the remaining four d orbitals due to its σ antibonding interactions with the SCS pincer. QRO analysis of the 18-crown-6 species (9-crown, Figure 11C) shows the same d orbital ordering as model A (optimized with K+), though dz2 is 0.90 eV more stabilized than dxy in 9-crown (Figure S87).
Figure 12.
QRO plots at an isovalue of 0.03 au of the d-assigned orbitals for amide complex 9 from model A (left, optimized with K+) and model B (right, optimized without K+). Energies relative to the doubly-occupied orbital are listed below the d orbital assignment. The structures of the complex are in gray wireframe, and potassium is represented as a purple sphere.
Additional experiments were performed to probe the influence of the cation on the electronic structure of 9. First, we computationally sampled a wide range of potential locations for a positive charge to learn about the orientational dependence, but the results were not conclusive (details in Supporting Information, Figure S88). In addition, we sought to experimentally “remove” the potassium completely by reacting starting material 5-Et2O with N(TMS)2 in the presence of an excess (2 equiv per N(TMS)2 group) of cryptand-2.2.2 (Scheme S1). The reaction conditions were otherwise identical to the syntheses of 9 and 9-crown. The Mössbauer spectrum of the dried crude reaction mixtures showed a mixture of two components: one with parameters close to 9 (δ = 0.17 mm/s, |ΔEQ| = 0.82 mm/s) and one with δ = 0.37 mm/s and |ΔEQ| = 3.27 mm/s (FigureS83). The parameters of the component with a larger isomer shift and quadrupole splitting than 9 are more similar to those predicted for anionic 9, which represents the extreme in which iron and K+ are too far apart to interact. However, we were unable to isolate this species for further study.
DISCUSSION
Relationship between Protonation State and Reduction Potential.
To our knowledge, bond dissociation free energies of thioamide SCS pincer compounds have not been reported. The N–H BDFE of 2 (56 kcal/mol) is substantially lower than the computed BDEs for free thioamides (90–100 kcal/mol), and its pKa of ~21 indicates that the N−H proton is much less acidic than free thioamides, which lie in the range 11–15 (Figure 13).71–73 It is seen from the Bordwell equation that the high pKa of 2 raises the BDFE relative to organic thioamides, but the very negative reduction potential (E1/2 = −1.42 V vs. Fc+/Fc) of the iron(II) center lowers the BDFE to a much larger extent. This potential is 1.0 V more negative than E1/2 for the Fe3+/2+ couple in a four-coordinate SNS pincer complex with an NHC as the fourth ligand, and is 1.4 V more negative than the analogous five-coordinate SNS scaffold with two THF molecules.40 The more negative reduction potential in our system may result from the negatively-charged aryl pincer in place of the pyridine pincer, and also from the coordination of three strong phosphine donors.
Figure 13.
Summary of the effect of iron coordination on the bond energy and pKa of the thioamide N−H site.
Although the N–H BDFE of 2 is significantly lower than those of free thioamides, it is comparable to other iron systems with N–H bonds in the supporting ligand. In a recent example, the N–H BDFEs of aminobenzenethiolate iron(II) complexes supported by a TACN ligand were determined by measurement of pKa and E1/2.74 In the parent system, the Fe3+/2+ redox couple was found at E1/2 = − 0.84 V, and the pKa was determined to be ≥ 19.5, giving a BDFE of ≥ 62 kcal/mol; bracketing showed an exact BDFE near 68 kcal/mol. Studies of nickel, platinum, and ruthenium SCS compounds also corroborate decreased N–H bond acidity and more negative reduction potential upon metalation, though these examples do not report bond energies.38, 75–77
It is interesting that deprotonating a distant thioamide site gives a highly-reducing iron(II) species, because of the potential for creating electron-rich active sites for small molecule activation using mild methods. This supports a growing body of evidence that reduction potentials may be modulated by the protonation state of sites outside the vicinity of metals in metallocofactors like the nitrogenase FeMoco.50, 78
Electronic Structure.
The EPR spectrum of 9 is notable for its high rhombicity. While the spectrum resembles those for some other S = 3/2 examples of well-defined iron complexes with large E/D,79–80 the most salient comparison is to a square planar S = 3/2 cobalt(II) ONO pincer complex with a chloride as the fourth donor.81 Structurally, this dianionic complex is reminiscent of 9, with the ONO pincer coordinating through a central amido N donor and lithium counterions flanking the alkoxide arms of the pincer ligand. The EPR spectrum of this complex was also highly rhombic, with E/D ≥ 0.25, maximum geff of 7.35, and gy > gx, gz. DFT calculated d orbital energies revealed an unusual doubly-occupied dz2 orbital, and the EPR spectrum was rationalized on the basis of strong spin-orbit coupling of the ground state with the dz2(β) → dyz and dxy(β) → dxz excited states. In our case, mixing of the ground state with an excited state created by promotion of dz2(β) into the half-filled dyz orbital could also account for the large deviation in g from ge in 9.
Sulfur, nitrogen, or mixed sulfur/nitrogen donors like in 8-11 are found in several other non-porphyrin ferric intermediate spin complexes.79, 82–87 These examples all also have near square planar or square pyramidal geometries. Mössbauer spectra of ferric S = 3/2 complexes generally have isomer shifts around 0.3 mm/s and large quadrupole splitting.62 Though Mössbauer spectra of complexes 8 (thiolate) and 10 (ammine) show normal δ and |ΔEQ| for this class of compounds, the spectrum of 9 (amide) is an outlier in line width, isomer shift, and quadrupole splitting. In regard to line widths, the 80 K spectra of both 8 and 10 consist of broadened asymmetric doublets. Broad, asymmetric doublets at 80 K in intermediate-spin iron(III) complexes are a marker of slow magnetic relaxation in the range of 10−8 s to 10−11 s; i.e., faster than the nuclear excited state lifetime of 57Fe (~10−8 s) but slower than the ~10−12 s that would give negligible paramagnetic broadening.62, 86, 88 The smaller line widths of 9 relative to those in 8 and 10 indicate that paramagnetic relaxation is faster, which is attributable to its larger E/D that mixes the Kramers doublets.89–90 To our knowledge, the quadrupole splitting values of 9 and 9-crown are the lowest reported for any intermediate-spin iron(III) complex, though anomalously small quadrupole splitting has also been reported in a few S = 2 iron(II) complexes with planar ligand environments.91–92 Planar ligand scaffolds can induce large positive contributions to the electric field gradient (EFG) on the axis perpendicular to the ligand plane (usually assigned as the z axis), and the Mössbauer quadrupole splitting is proportional to the z-component of the EFG.62, 91 Interestingly, computational analysis of the cited systems with anomalously small ΔEQ also revealed a doubly-occupied dz2 orbital, as we found for 9 (Figure 12). Thus, the small ΔEQ values imply that the large positive contribution to the z-component of the electric field gradient by the planar ligand is counteracted by a large negative contribution from the dz2 ground state. Double occupation of the dz2 orbital also accounts for the decreased isomer shift (which decreases with increasing iron s-electron density) of 9 by greater mixing of the doubly-occupied 3dz2 with the 4s orbital,93 which is corroborated by our DFT calculations.
The question that arises is then: why do amide complexes 9 and 9-crown show evidence for a dz2 ground state electronic structure while thiolate complexes 8 and 8-crown as well as ammonia complex 10 do not? First, we propose that the π-donating orbitals of the N(TMS)2 ligand in 9 exert a stronger destabilizing influence on the π-symmetry iron dxy and dyz orbitals than the thiolate donor in 8 and certainly more than the ammine donor in 10. Our proposition of a stronger interaction in 9 than 8 is tied to its structure, which shows a short Fe−N bond and a nearly 90° angle between the N(TMS)2 plane and the pincer plane, permitting excellent overlap between dxy and the N px orbital and between the dyz and sp2-like orbitals aligned roughly along the N−Si bonds. This highly π-donating ligand field closes the energy gap between dz2 and dxz/dyz in the typical square planar splitting pattern. This explanation alone does not account, however, for the extreme variability in the calculated Mössbauer parameters for 9 depending on whether there is a positive charge near iron. On the basis of our DFT results, we therefore also suggest that the dz2 orbital can be stabilized by the presence of a nearby positive charge. Though the complete description of the cation interaction with each of the d orbitals is not elucidated by our work, selective stabilization of the dz2 orbital may be a result of its shape. A positive charge in the plane of the pincer could interact with the torus-shaped portion of the dz2 orbital, and one along the z axis can interact with the lobes pointing above and below the pincer plane. The stabilizing interaction of charge with the lobes of dz2 is illustrated by the comparison between the QRO energies of model A in Figure 12 with optimized 9-crown (Figure S87). Model A has positive charge with non-zero projections both onto the z axis and xy-plane while 9-crown has positive charge that is along the z axis. Accordingly, the dz2 orbital is 0.35 eV lower in energy than the DOMO+1 in 9-crown versus model A.
The influence of the alkali metal cation is a distinctive aspect of the work reported here, and it emphasizes the potential tuning role of cations near a transition-metal center. In a relevant example, Yang has synthesized systems in which appended cations cause substantial electric fields,94 which can have beneficial influences on catalysis.95–96 Electronic structure calculations in these systems did not show that the orbital energies were differentially affected by the cation, leading to the conclusion that the cation mainly exerts an electrostatic effect. Tolman has shown the influences of nearby cations on O–H bond dissociation energies.97 In more recent work, Tomson observed shifts in redox potentials ascribed to selective stabilization of a dz2 orbital, and supported this idea with DFT calculations.98 Cation effects in these examples were typically assessed using redox potentials, but in our complexes we utilize Mössbauer spectroscopy, which can report the electronic structure of iron. Through the combination of spectroscopy and DFT, we are able to substantiate the presence of cation-dependent changes in orbital energetics. While the overall spin state is conserved throughout, a β electron shifts to a cation-stabilized dz2 orbital, manifested in Mössbauer spectra by low δ and |ΔEQ|. In the future, this effect could be utilized to control reactivity that depends on the ordering of orbital energies.
The conclusions from our electronic structure studies of complexes 8 (thiolate), 9 (amide), and 10 (ammine) utilizing EPR spectroscopy, SQUID magnetometry, Mössbauer spectroscopy, and DFT calculations are summarized as follows. EPR spectroscopy in combination with variable-temperature magnetometry measurements demonstrated large negative D with E ≠ 0 in each case. Mössbauer spectra of 8 and 10 showed isomer shifts and quadrupole splitting values consistent with other known intermediate spin iron(III) complexes, and DFT could accurately predict these parameters without including a counterion in 8. The quadrupole splitting of 9, however, was anomalously low, and DFT predicted a large |ΔEQ| when no counterion was included. Inclusion of potassium in the model of 9 gave calculated Mössbauer parameters in agreement with experiment, and the energetic ordering of dz2 relative to the other d orbitals is likely to determine whether a typical large |ΔEQ| versus an atypical small |ΔEQ| is observed. Complex 9 is responsive because it is poised near a transition between two electronic states, which arises from the relatively low distortion from planarity, strong π donation from the amide ligand, and short Fe−N bond. Clearly, the SCS scaffold gives access to a range of electronic structures depending on the identity of the fourth donor ligand and on the cation location.
Comparisons to the Nitrogenase FeMoco.
The thioamide supporting ligand has the same donor atoms as the iron sites in the FeMoco, and we briefly explore the comparison of molecular and electronic structures. The most thoroughly-characterized state of the FeMoco is the enzyme resting state (E0), and in this structure, the high-spin belt irons are bridged by sulfide ligands and share a central carbide, leading to a 3S/1C environment.12 Our complex 8 likewise has an iron center ligated exclusively by three S donors and one C donor. Its Fe−C bond length of 2.00 Å is very close to the 2.01 Å Fe−C distance in the FeMoco.13–14, 99 The Fe–S lengths, which range from 2.24 to 2.25 Å, are also highly similar to Fe−S bond lengths in the FeMoco (2.25 to 2.27 Å). Despite being in an intermediate spin state, these bond distances are closer to those in the resting state than the two previously-reported (SCS)iron complexes, which were in lower oxidation states and thus had longer Fe−C/S bonds.26, 28
The final state of the enzyme during catalysis is an iron-ammonia adduct (E8).100 Because turnover from E8 to E0 is redox neutral, ammonia would be expected to bind to either the iron(II) or iron(III) oxidation states.100 We were able to isolate the stable ammonia adduct 10 in high yield in an intermediate-spin iron(III) oxidation state, thus demonstrating the feasibility of an iron(III) ammonia adduct in a C and S-ligand field.
CONCLUSIONS
A new pincer ligand bearing only sulfur and carbon donor atoms can support new iron complexes. The ligand scaffold has a proton-responsive site, gives access to unusual electronic structures, and mimics the sulfide and carbide donors found in the cofactor of nitrogenase enzymes. Adducts of thiolate, amide, ammonia, and CO demonstrate the versatility of this SCS pincer system to support a variety of nitrogenase-relevant donors, and induce an electronic structure that is highly sensitive to the iron surroundings.
Supplementary Material
Synopsis:
Iron(III) complexes of a proton-responsive SCS pincer ligand have unusual properties, which include finely balanced d-orbital energies in the electronic structure.
ACKNOWLEDGMENTS
We gratefully acknowledge funding from the National Institutes of Health (R01-GM065313 to P.L.H.; F32-GM126656 to K.L.S.; R35-GM124908 to K.M.L.). We thank Majed Fataftah for assistance with magnetism interpretation. We thank Nicholas Bingham for collecting SQUID magnetometry measurements. We thank the Yale Center for Research Computing for guidance and use of the research computing infrastructure, specifically Thomas Langford. We thank Jessica Freeze for assistance with scripts for computational searching. XAS data were obtained at the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, which is supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the DOE Office of Biological and Environmental Research, and by the National Institutes of Health, National Institute of General Medical Sciences (P30GM133894).
a preprint of this article was posted on ChemRxiv.101
Supporting Information. The following files are available free of charge. Synthetic procedures, spectroscopic data, and computational details (PDF) Crystallographic files (CIF)
REFERENCES
- 1.Bolm C; Legros J; Le Paih J; Zani L Iron-Catalyzed Reactions in Organic Synthesis. Chem. Rev 2004, 104, 6217–6254. [DOI] [PubMed] [Google Scholar]
- 2.Walker FA; Simonis U, Iron Porphyrin Chemistry. In Encyclopedia of Inorganic Chemistry, John Wiley & Sons: New York, 2006. [Google Scholar]
- 3.Bauer I; Knölker H-J Iron Catalysis in Organic Synthesis. Chem. Rev 2015, 115, 3170–3387. [DOI] [PubMed] [Google Scholar]
- 4.Fürstner A Iron Catalysis in Organic Synthesis: A Critical Assessment of What It Takes To Make This Base Metal a Multitasking Champion. ACS Cent. Sci 2016, 2, 778–789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Shang R; Ilies L; Nakamura E Iron-Catalyzed C–H Bond Activation. Chem. Rev 2017, 117, 9086–9139. [DOI] [PubMed] [Google Scholar]
- 6.Wei D; Darcel C Iron Catalysis in Reduction and Hydrometalation Reactions. Chem. Rev 2019, 119, 2550–2610. [DOI] [PubMed] [Google Scholar]
- 7.Tard C; Pickett CJ Structural and Functional Analogues of the Active Sites of the [Fe]-, [NiFe]-, and [FeFe]-Hydrogenases. Chem. Rev 2009, 109, 2245–2274. [DOI] [PubMed] [Google Scholar]
- 8.Bart SC; Chłopek K; Bill E; Bouwkamp MW; Lobkovsky E; Neese F; Wieghardt K; Chirik PJ Electronic Structure of Bis(imino)pyridine Iron Dichloride, Monochloride, and Neutral Ligand Complexes: A Combined Structural, Spectroscopic, and Computational Study. J. Am. Chem. Soc 2006, 128, 13901–13912. [DOI] [PubMed] [Google Scholar]
- 9.Volbeda J; Meetsma A; Bouwkamp MW Electron-Deficient Iron Alkyl Complexes Supported by Diimine Ligand (Ph2CN)2C2H4: Evidence for Reversible Ethylene Binding. Organometallics 2009, 28, 209–215. [Google Scholar]
- 10.Holland PL Distinctive Reaction Pathways at Base Metals in High-Spin Organometallic Catalysts. Acc. Chem. Res 2015, 48, 1696–1702. [DOI] [PubMed] [Google Scholar]
- 11.Sears JD; Neate PGN; Neidig ML Intermediates and Mechanism in Iron-Catalyzed Cross-Coupling. J. Am. Chem. Soc 2018, 140, 11872–11883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lancaster KM; Römelt M; Ettenhuber P; Hu Y; Ribbe MW; Neese F; Bergmann U; DeBeer S X-ray Emission Spectroscopy Evidences a Central Carbon in the Nitrogenase Iron-Molybdenum Cofactor. Science 2011, 334, 974–977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Spatzal T; Aksoyoglu M; Zhang L; Andrade SLA; Schleicher E; Weber S; Rees DC; Einsle O Evidence for Interstitial Carbon in Nitrogenase FeMo Cofactor. Science 2011, 334, 940–940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Einsle O; Rees DC Structural Enzymology of Nitrogenase Enzymes. Chem. Rev 2020, 120, 4969–5004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Poli R Open-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and Structure. Chem. Rev 1996, 96, 2135–2204. [DOI] [PubMed] [Google Scholar]
- 16.Schröder D; Shaik S; Schwarz H Two-State Reactivity as a New Concept in Organometallic Chemistry. Acc. Chem. Res 2000, 33, 139–145. [DOI] [PubMed] [Google Scholar]
- 17.Shaik S; Hirao H; Kumar D Reactivity of High-Valent Iron–Oxo Species in Enzymes and Synthetic Reagents: A Tale of Many States. Acc. Chem. Res 2007, 40, 532–542. [DOI] [PubMed] [Google Scholar]
- 18.Lutz SA; Hickey AK; Gao Y; Chen C-H; Smith JM Two-State Reactivity in Iron-Catalyzed Alkene Isomerization Confers σ-Base Resistance. J. Am. Chem. Soc 2020, 142, 15527–15535. [DOI] [PubMed] [Google Scholar]
- 19.Čorić I; Holland PL Insight into the Iron–Molybdenum Cofactor of Nitrogenase from Synthetic Iron Complexes with Sulfur, Carbon, and Hydride Ligands. J. Am. Chem. Soc 2016, 138, 7200–7211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Speelman AL; Holland PL, Sulfur-Supported Iron Complexes for Understanding N2 Reduction. In Nitrogen Fixation, Nishibayashi Y, Ed. Springer International Publishing: Cham, 2017; pp 197–213. [Google Scholar]
- 21.Tanifuji K; Ohki Y Metal–Sulfur Compounds in N2 Reduction and Nitrogenase-Related Chemistry. Chem. Rev 2020, 120, 5194–5251. [DOI] [PubMed] [Google Scholar]
- 22.Imamura Y; Kubo K; Mizuta T; Miyoshi K Reactions of Ring-Slipped Iron Complexes Derived from P(S)Ph-Bridged [1]Ferrocenophane: Synthesis of Bis(half-sandwich) Heterodinuclear Complexes. Organometallics 2006, 25, 2301–2307. [Google Scholar]
- 23.Chen Y; Zhou Y; Chen P; Tao Y; Li Y; Qu J Nitrogenase Model Complexes: Synthesis, Structure, and Catalytic N-N Bond Cleavage of Hydrazines on Diiron Centers. J. Am. Chem. Soc 2008, 130, 15250–15251. [DOI] [PubMed] [Google Scholar]
- 24.Chen Y; Liu L; Peng Y; Chen P; Luo Y; Qu J Unusual Thiolate-Bridged Diiron Clusters Bearing the cis-HN═NH Ligand and Their Reactivities with Terminal Alkynes. J. Am. Chem. Soc 2011, 133, 1147–1149. [DOI] [PubMed] [Google Scholar]
- 25.Li Y; Li Y; Wang B; Luo Y; Yang D; Tong P; Zhao J; Luo L; Zhou Y; Chen S; Cheng F; Qu J Ammonia formation by a thiolate-bridged diiron amide complex as a nitrogenase mimic. Nat. Chem 2013, 5, 320–326. [DOI] [PubMed] [Google Scholar]
- 26.Čorić I; Mercado BQ; Bill E; Vinyard DJ; Holland PL Binding of dinitrogen to an iron–sulfur–carbon site. Nature 2015, 526, 96–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Yang D; Li Y; Wang B; Zhao X; Su L; Chen S; Tong P; Luo Y; Qu J Synthesis and Electrocatalytic Property of Diiron Hydride Complexes Derived from a Thiolate-Bridged Diiron Complex. Inorg. Chem 2015, 54, 10243–10249. [DOI] [PubMed] [Google Scholar]
- 28.Speelman AL; Skubi KL; Mercado BQ; Holland PL Synthesis and Reactivity of Iron Complexes with a Biomimetic SCS Pincer Ligand. Inorg. Chem 2021, 60, 1965–1974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Sun T; Xu S; Yang D; Su L; Wang B; Qu J Catalytic Disproportionation of Hydrazine Promoted by Biomimetic Diiron Complexes with Benzene-1,2-Dithiolate Bridge Modified by Different Substituents. Eur. J. Inorg. Chem 2020, 2020, 4263–4269. [Google Scholar]
- 30.Speelman AL; Čorić I; Van Stappen C; DeBeer S; Mercado BQ; Holland PL Nitrogenase-Relevant Reactivity of a Synthetic Iron–Sulfur–Carbon Site. J. Am. Chem. Soc 2019, 141, 13148–13157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Bergbreiter DE; Osburn PL; Liu Y-S Tridentate SCS Palladium(II) Complexes: New, Highly Stable, Recyclable Catalysts for the Heck Reaction. J. Am. Chem. Soc 1999, 121, 9531–9538. [Google Scholar]
- 32.Manen H.-Jan v.; Nakashima K; Shinkai S; Kooijman H; Spek Anthony L.; van Veggel, Frank CJM; Reinhoudt DN Coordination Chemistry of (SCS)PdII Pincer Systems. Eur. J. Inorg. Chem 2000, 2000, 2533–2540. [Google Scholar]
- 33.Begum RA; Powell D; Bowman-James K Thioamide Pincer Ligands with Charge Versatility. Inorg. Chem 2006, 45, 964–966. [DOI] [PubMed] [Google Scholar]
- 34.Mehendale NC; Lutz M; Spek AL; Klein Gebbink RJM; van Koten G Self-assembly of para-OH functionalized ECE-metalated pincer complexes. J. Organomet. Chem 2008, 693, 2971–2981. [Google Scholar]
- 35.Kruithof CA; Berger A; Dijkstra HP; Soulimani F; Visser T; Lutz M; Spek AL; Gebbink RJMK; van Koten G Sulfato-bridged ECE-pincer palladium(II) complexes: structures in the solid-state and in solution, and catalytic properties. Dalton Trans. 2009, 3306–3314. [DOI] [PubMed] [Google Scholar]
- 36.Okamoto K; Kanbara T; Yamamoto T; Wada A Preparation and Characterization of Luminescent SCS and NCN Pincer Platinum Complexes Derived from 3,5-Bis(anilinothiocarbonyl)toluene. Organometallics 2006, 25, 4026–4029. [Google Scholar]
- 37.Kuwabara J; Kanbara T Synthesis and Optical Properties of Pincer Palladium and Platinum Complexes having Thioamide Units. J. Photopolym. Sci. Technol 2008, 21, 349–353. [Google Scholar]
- 38.Komiyama Y; Kuwabara J; Kanbara T Deprotonation-Induced Structural Changes in SNS-Pincer Ruthenium Complexes with Secondary Thioamide Groups. Organometallics 2014, 33, 885–891. [Google Scholar]
- 39.Suzuki T; Kajita Y; Masuda H Deprotonation/protonation-driven change of the σ-donor ability of a sulfur atom in iron(II) complexes with a thioamide SNS pincer type ligand. Dalton Trans. 2014, 43, 9732–9739. [DOI] [PubMed] [Google Scholar]
- 40.Suzuki T; Matsumoto J; Kajita Y; Inomata T; Ozawa T; Masuda H Nitrosyl and carbene iron complexes bearing a κ3-SNS thioamide pincer type ligand. Dalton Trans. 2015, 44, 1017–1022. [DOI] [PubMed] [Google Scholar]
- 41.Okamoto K; Kuwabara J; Kanbara T Secondary Thioamides as Multidentate Ligands for Functional Metal Complexes. Chem. Lett 2014, 44, 102–110. [Google Scholar]
- 42.Lee CM; Kumler WD The Dipole Moment and Structure of Thiolactams. J. Org. Chem 1962, 27, 2052–2054. [Google Scholar]
- 43.Hoffman BM; Lukoyanov D; Yang Z-Y; Dean DR; Seefeldt LC Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage. Chem. Rev 2014, 114, 4041–4062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Raugei S; Seefeldt LC; Hoffman BM Critical computational analysis illuminates the reductive-elimination mechanism that activates nitrogenase for N2 reduction. Proc. Natl. Acad. Sci. U.S.A 2018, 115, E10521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Dance I The controlled relay of multiple protons required at the active site of nitrogenase. Dalton Trans. 2012, 41, 7647–7659. [DOI] [PubMed] [Google Scholar]
- 46.McKee ML A New Nitrogenase Mechanism Using a CFe8S9 Model: Does H2 Elimination Activate the Complex to N2 Addition to the Central Carbon Atom? J. Phys. Chem. A 2016, 120, 754–764. [DOI] [PubMed] [Google Scholar]
- 47.Rao L; Xu X; Adamo C Theoretical Investigation on the Role of the Central Carbon Atom and Close Protein Environment on the Nitrogen Reduction in Mo Nitrogenase. ACS Catal. 2016, 6, 1567–1577. [Google Scholar]
- 48.Thorhallsson AT; Benediktsson B; Bjornsson R A model for dinitrogen binding in the E4 state of nitrogenase. Chem. Sci 2019, 10, 11110–11124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Siegbahn PEM Model Calculations Suggest that the Central Carbon in the FeMo-Cofactor of Nitrogenase Becomes Protonated in the Process of Nitrogen Fixation. J. Am. Chem. Soc 2016, 138, 10485–10495. [DOI] [PubMed] [Google Scholar]
- 50.Lampret O; Duan J; Hofmann E; Winkler M; Armstrong FA; Happe T The roles of long-range proton-coupled electron transfer in the directionality and efficiency of [FeFe]-hydrogenases. Proc. Natl. Acad. Sci. U.S.A 2020, 117, 20520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Tyburski R; Liu T; Glover SD; Hammarström L Proton-Coupled Electron Transfer Guidelines, Fair and Square. J. Am. Chem. Soc 2021, 143, 560–576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Warren JJ; Menzeleev AR; Kretchmer JS; Miller TF; Gray HB; Mayer JM Long-Range Proton-Coupled Electron-Transfer Reactions of Bis(imidazole) Iron Tetraphenylporphyrins Linked to Benzoates. J. Phys. Chem. Lett 2013, 4, 519–523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Darcy JW; Koronkiewicz B; Parada GA; Mayer JM A Continuum of Proton-Coupled Electron Transfer Reactivity. Acc. Chem. Res 2018, 51, 2391–2399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Bisson AP; Hunter CA; Morales JC; Young K Cooperative Interactions in a Ternary Mixture. Chem. Eur. J 1998, 4, 845–851. [Google Scholar]
- 55.Karsch HH Funktionelle Trimethylphosphinderivate, V. Kovalente Methyleisen(II)-Phosphinkomplexe. Chem. Ber 1977, 110, 2699–2711. [Google Scholar]
- 56.Bhattacharya P; Krause JA; Guan H Iron Hydride Complexes Bearing Phosphinite-Based Pincer Ligands: Synthesis, Reactivity, and Catalytic Application in Hydrosilylation Reactions. Organometallics 2011, 30, 4720–4729. [Google Scholar]
- 57.Tshepelevitsh S; Kütt A; Lõkov M; Kaljurand I; Saame J; Heering A; Plieger PG; Vianello R; Leito I On the Basicity of Organic Bases in Different Media. Eur. J. Org. Chem 2019, 2019, 6735–6748. [Google Scholar]
- 58.Wise CF; Agarwal RG; Mayer JM Determining Proton-Coupled Standard Potentials and X–H Bond Dissociation Free Energies in Nonaqueous Solvents Using Open-Circuit Potential Measurements. J. Am. Chem. Soc 2020, 142, 10681–10691. [DOI] [PubMed] [Google Scholar]
- 59.Warren JJ; Tronic TA; Mayer JM Thermochemistry of Proton-Coupled Electron Transfer Reagents and its Implications. Chem. Rev 2010, 110, 6961–7001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Cao L; Jennings MC; Puddephatt RJ Amine−Amide Equilibrium in Gold(III) Complexes and a Gold(III)−Gold(I) Aurophilic Bond. Inorg. Chem 2007, 46, 1361–1368. [DOI] [PubMed] [Google Scholar]
- 61.Hunt AP; Lehnert N The Thiolate Trans Effect in Heme {FeNO}6 Complexes and Beyond: Insight into the Nature of the Push Effect. Inorg Chem 2019, 58, 11317–11332. [DOI] [PubMed] [Google Scholar]
- 62.Gütlich P; Bill E; Trautwein AX, Mössbauer Spectroscopy and Transition Metal Chemistry. Springer-Verlag: Berlin, Heidelberg, 2011. [Google Scholar]
- 63.Sik Min K; Weyhermüller T; Wieghardt K Coordination chemistry of 2-(8-aminoquinolino)-4,6-di-tert-butylphenol with manganese(IV), iron(III), and cobalt(II/III): N,O-coordinated o-iminobenzosemiquinonate(1−) π radical monoanions vs.o-iminophenolate(2−) dianions. Dalton Trans. 2004, 178–186. [DOI] [PubMed] [Google Scholar]
- 64.Bittner MM; Lindeman SV; Fiedler AT A Synthetic Model of the Putative Fe(II)-Iminobenzosemiquinonate Intermediate in the Catalytic Cycle of o-Aminophenol Dioxygenases. J. Am. Chem. Soc 2012, 134, 5460–5463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.De Hont RF; Xue G; Hendrich MP; Que L Jr.; Bominaar EL; Münck E Mössbauer, electron paramagnetic resonance, and density functional theory studies of synthetic S = 1/2 Fe(III)-O-Fe(IV)═O complexes. Superexchange-mediated spin transition at the Fe(IV)═O site. Inorg. Chem 2010, 49, 8310–8322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.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, 2594–2606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Bochevarov AD; Friesner RA; Lippard SJ Prediction of 57Fe Mössbauer Parameters by Density Functional Theory: A Benchmark Study. J. Chem. Theory Comput 2010, 6, 3735–3749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Zhang Y; Oldfield E An Investigation of the Unusual 57Fe Mössbauer Quadrupole Splittings and Isomer Shifts in 2 and 3-Coordinate Fe(II) Complexes. J. Phys. Chem. B 2003, 107, 7180–7188. [Google Scholar]
- 69.Glendening ED; Feller D; Thompson MA An Ab Initio Investigation of the Structure and Alkali Metal Cation Selectivity of 18-Crown-6. J. Am. Chem. Soc 1994, 116, 10657–10669. [Google Scholar]
- 70.Hori K; Yamada H; Yamabe T Theoretical study on the nature of the interaction between crown ethers and alkali cations: Relation of interaction energy and ion selectivity. Tetrahedron 1983, 39, 67–73. [Google Scholar]
- 71.Kaur D; Kaur RP; Kohli R Substituent effect on N-H bond dissociation enthalpies of amines and amides: A theoretical study. Int. J. Quantum Chem 2009, 109, 559–568. [Google Scholar]
- 72.Sifferlen T; Rueping M; Gademann K; Jaun B; Seebach D β-Thiopeptides: Synthesis, NMR Solution Structure, CD Spectra, and Photochemistry. Helv. Chim. Acta 1999, 82, 2067–2093. [Google Scholar]
- 73.Bordwell FG Equilibrium acidities in dimethyl sulfoxide solution. Acc. Chem. Res 1988, 21, 456–463. [Google Scholar]
- 74.Gordon JB; McGale JP; Siegler MA; Goldberg DP Proton-Coupled Electron-Transfer Reactivity Controls Iron versus Sulfur Oxidation in Nonheme Iron–Thiolate Complexes. Inorg Chem 2021, 60, 6255–6265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Teratani T; Koizumi T-A; Yamamoto T; Kanbara T Changes in redox potential of a nickel-pincer complex bearing reactive secondary thioamide units: Changes caused by deprotonation/protonation reactions on addition of NEt3 and DBU. Inorg. Chem. Commun 2011, 14, 836–838. [Google Scholar]
- 76.Okamoto K; Yamamoto T; Akita M; Wada A; Kanbara T Chemical Stimuli Induced Phosphorescence Modulation of Secondary Thioamide-Based Pincer Platinum Complexes. Organometallics 2009, 28, 3307–3310. [Google Scholar]
- 77.Teratani T; Koizumi T-A; Yamamoto T; Tanaka K; Kanbara T Deprotonation/protonation of coordinated secondary thioamide units of pincer ruthenium complexes: Modulation of voltammetric and spectroscopic characterization of the pincer complexes. Dalton Trans. 2011, 40, 8879–8886. [DOI] [PubMed] [Google Scholar]
- 78.Hickey DP; Cai R; Yang Z-Y; Grunau K; Einsle O; Seefeldt LC; Minteer SD Establishing a Thermodynamic Landscape for the Active Site of Mo-Dependent Nitrogenase. J. Am. Chem. Soc 2019, 141, 17150–17157. [DOI] [PubMed] [Google Scholar]
- 79.Keutel H; Käpplinger I; Jäger EG; Grodzicki M; Schünemann V; Trautwein AX Structural, Magnetic, and Electronic Properties of a Pentacoordinated Intermediate-Spin (S = 3/2) Iron(III) Complex with a Macrocyclic [N4]2- Ligand. Inorg Chem 1999, 38, 2320–2327. [Google Scholar]
- 80.Maier TM; Gawron M; Coburger P; Bodensteiner M; Wolf R; van Leest NP; de Bruin B; Demeshko S; Meyer F Low-Valence Anionic α-Diimine Iron Complexes: Synthesis, Characterization, and Catalytic Hydroboration Studies. Inorg Chem 2020, 59, 16035–16052. [DOI] [PubMed] [Google Scholar]
- 81.Pascualini ME; Stoian SA; Ozarowski A; Di Russo NV; Thuijs AE; Abboud KA; Christou G; Veige AS Synthesis and characterization of a family of M2+ complexes supported by a trianionic ONO3− pincer-type ligand: towards the stabilization of high-spin square-planar complexes. Dalton Trans. 2015, 44, 20207–20215. [DOI] [PubMed] [Google Scholar]
- 82.Roy N; Sproules S; Bill E; Weyhermüller T; Wieghardt K Molecular and Electronic Structure of the Square Planar Bis(o-amidobenzenethiolato)iron(III) Anion and Its Bis(o-quinoxalinedithiolato)iron(III) Analogue. Inorg Chem 2008, 47, 10911–10920. [DOI] [PubMed] [Google Scholar]
- 83.Jain SK; Garg BS; Bhoon YK Iron(III) complexes of 2-acetylpyridine-4-phenyl-3-thiosemicarbazones: Magnetic, E.S.R. and spectral studies. Transition Met. Chem 1986, 11, 89–95. [Google Scholar]
- 84.Lugo-Mas P; Taylor W; Schweitzer D; Theisen RM; Xu L; Shearer J; Swartz RD; Gleaves MC; DiPasquale A; Kaminsky W; Kovacs JA Properties of Square-Pyramidal Alkyl−Thiolate FeIII Complexes, Including an Analogue of the Unmodified Form of Nitrile Hydratase. Inorg. Chem 2008, 47, 11228–11236. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Begum A; Sarkar S An Iron(III) Dithiolene Complex as a Functional Model of Iron Hydrogenase. Eur. J. Inorg. Chem 2012, 2012, 40–43. [Google Scholar]
- 86.Niarchos D; Kostikas A; Simopoulos A; Coucouvanis D; Piltingsrud D; Coffman RE Mössbauer, magnetic susceptibility, and EPR studies of intermediate spin iron (III) dithiooxalato halides. J. Chem. Phys 1978, 69, 4411–4418. [Google Scholar]
- 87.Fallon GD; Gatehouse BM; Marini PJ; Murray KS; West BO Iron complexes of N-substituted thiosalicylideneimines. Structure, magnetism, and Mössbauer spectra of the complex chloro[N,N′-ethylenebis(thiosalicylideneiminato)]iron(III), an intermediate-spin complex which exists in three crystalline forms. J. Chem. Soc., Dalton Trans 1984, 2733–2739. [Google Scholar]
- 88.Blume M Magnetic Relaxation and Asymmetric Quadrupole Doublets in the Mössbauer Effect. Phys. Rev. Lett 1965, 14, 96–98. [Google Scholar]
- 89.Fataftah MS; Coste SC; Vlaisavljevich B; Zadrozny JM; Freedman DE Transformation of the coordination complex [Co(C3S5)2]2− from a molecular magnet to a potential qubit. Chem. Sci 2016, 7, 6160–6166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Gatteschi D; Sessoli R Quantum Tunneling of Magnetization and Related Phenomena in Molecular Materials. Angew. Chem., Int. Ed 2003, 42, 268–297. [DOI] [PubMed] [Google Scholar]
- 91.Sanakis Y; Power PP; Stubna A; Münck E Mössbauer Study of the Three-Coordinate Planar FeII Thiolate Complex [Fe(SR)3]−: Model for the Trigonal Iron Sites of the MoFe7S9:Homocitrate Cofactor of Nitrogenase. Inorg. Chem 2002, 41, 2690–2696. [DOI] [PubMed] [Google Scholar]
- 92.Pascualini ME; Di Russo NV; Thuijs AE; Ozarowski A; Stoian SA; Abboud KA; Christou G; Veige AS A high-spin square-planar Fe(II) complex stabilized by a trianionic pincer-type ligand and conclusive evidence for retention of geometry and spin state in solution. Chem. Sci 2015, 6, 608–612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Deeth RJ Ligand field and density functional descriptions of the d-states and bonding in transition metal complexes. Faraday Discuss. 2003, 124, 379–391. [DOI] [PubMed] [Google Scholar]
- 94.Kang K; Fuller J; Reath AH; Ziller JW; Alexandrova AN; Yang JY Installation of internal electric fields by non-redox active cations in transition metal complexes. Chem. Sci 2019, 10, 10135–10142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Chantarojsiri T; Reath AH; Yang JY Cationic Charges Leading to an Inverse Free-Energy Relationship for N−N Bond Formation by MnVI Nitrides. Angew. Chem., Int. Ed 2018, 57, 14037–14042. [DOI] [PubMed] [Google Scholar]
- 96.Chantarojsiri T; Ziller JW; Yang JY Incorporation of redox-inactive cations promotes iron catalyzed aerobic C–H oxidation at mild potentials. Chem. Sci 2018, 9, 2567–2574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Dhar D; Yee GM; Tolman WB Effects of Charged Ligand Substituents on the Properties of the Formally Copper(III)-Hydroxide ([CuOH]2+) Unit. Inorg. Chem 2018, 57, 9794–9806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.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, 4395–4404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Einsle O; Tezcan FA; Andrade SLA; Schmid B; Yoshida M; Howard JB; Rees DC Nitrogenase MoFe-Protein at 1.16 Å Resolution: A Central Ligand in the FeMo-Cofactor. Science 2002, 297, 1696–1700. [DOI] [PubMed] [Google Scholar]
- 100.Lukoyanov D; Yang Z-Y; Barney BM; Dean DR; Seefeldt LC; Hoffman BM Unification of reaction pathway and kinetic scheme for N2 reduction catalyzed by nitrogenase. Proc. Natl. Acad. Sci. U.S.A 2012, 109, 5583. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Skubi K; Hooper R; Mercado B; Bollmeyer M; MacMillan S; Lancaster K; Holland P ChemRxiv 2021, DOI: 10.33774/chemrxiv-2021-6qjz6. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.