Sean F.
McWilliams
a,
Brandon Q.
Mercado
a,
K. Cory
MacLeod
a,
Majed S.
Fataftah
a,
Maxime
Tarrago
b,
Xiaoping
Wang
c,
Eckhard
Bill
b,
Shengfa
Ye
*bd and
Patrick L.
Holland
*a
aDepartment of Chemistry, Yale University, New Haven, Connecticut, USA. E-mail: patrick.holland@yale.edu
bMax Planck Institute for Chemical Energy Conversion, Mülheim an der Ruhr, Germany
cNeutron Sciences Directorate, Oak Ridge National Laboratory, Oak Ridge, Tennessee, USA
dState Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian, China. E-mail: shengfa.ye@dicp.ac.cn
First published on 8th February 2023
Hydride complexes are important in catalysis and in iron–sulfur enzymes like nitrogenase, but the impact of hydride mobility on local iron spin states has been underexplored. We describe studies of a dimeric diiron(II) hydride complex using X-ray and neutron crystallography, Mössbauer spectroscopy, magnetism, DFT, and ab initio calculations, which give insight into the dynamics and the electronic structure brought about by the hydrides. The two iron sites in the dimer have differing square-planar (intermediate-spin) and tetrahedral (high-spin) iron geometries, which are distinguished only by the hydride positions. These are strongly coupled to give an Stotal = 3 ground state with substantial magnetic anisotropy, and the merits of both localized and delocalized spin models are discussed. The dynamic nature of the sites is dependent on crystal packing, as shown by changes during a phase transformation that occurs near 160 K. The change in dynamics of the hydride motion leads to insight into its influence on the electronic structure. The accumulated data indicate that the two sites can trade geometries by rotating the hydrides, at a rate that is rapid above the phase transition temperature but slow below it. This small movement of the hydrides causes large changes in the ligand field because they are strong-field ligands. This suggests that hydrides could be useful in catalysis not only due to their reactivity, but also due to their ability to rapidly modulate the local electronic structure and spin states at metal sites.
Given this broad range of one- and two-electron reactions in diverse coordination environments, the spin states of hydride complexes have been much less studied.10 Open-shell hydride complexes are strongly implicated in unobserved catalytic intermediates, for example nitrogenase cofactors (which have iron–sulfur clusters in weak-field environments)11–14 and HAT catalysts for alkene hydrofunctionalization (which often have only weak-field acac or salen ligands).15 Thus there is strong motivation for understanding the electronic structure of hydride ligands in open-shell systems. However, very few isolated hydride complexes are known that are paramagnetic.10,16
In the work described here, we focus on iron(II), which is found in many of the above catalysts. The conventionally found ground spin states of iron(II) are low-spin (S = 0, often the ground state in octahedral geometry) or high-spin (S = 2, often the ground state in tetrahedral or octahedral geometries). An intermediate-spin electronic configuration (S = 1) is less commonly the ground state, but this can occur in a square-planar ligand field that raises the energy of the dx2–y2 orbital enough that it remains unoccupied. A d6 system with an intermediate-spin ground state was first observed and explained in (tetraphenylporphyrin)iron(II).17,18 More recently, Chirik et al. used a pincer ligand and varying steric demands to explore the balance between square-planar and tetrahedral geometries, and similarly observed S = 1 ground states for square-planar iron(II) and S = 2 ground states for tetrahedral iron(II).19 However, a square-planar geometry does not necessarily give an intermediate-spin ground state.20–25
Different spin states engender different metal–ligand bond lengths, with bonds being shortest in low-spin iron complexes and longer in open-shell iron complexes because the population of metal–ligand antibonding orbitals in the higher spin states weakens the bonding interactions. However, M–H distances are not known for the few open-shell hydrides, because locating hydrides accurately requires a neutron crystal structure. Correlating these differences with reactivity is also challenging because the spin state can change during a reaction.26–31
In the quest to understand open-shell hydride species, β-diketiminate ligands have been beneficial, because they are modular and tunable weak-field ligands.32 We have isolated iron(II) hydride complexes with three related N,N′-diaryldiketiminate ligands, and each was crystallographically characterized as a dimer with two bridging hydrides.33–35 In related work, Murray used a macrocyclic β-diketiminate to create complexes with a triiron core bridged by three hydrides.36 In each β-diketiminate-supported iron hydride complex, Mössbauer and magnetism studies have led the authors to conclude that the ground state has high-spin electronic configurations at the iron sites. This is reasonable since a wide variety of related mononuclear β-diketiminate-supported iron(II) complexes have S = 2 ground states with related σ-donor ligands such as alkyl, aryl, alkoxide, amide, and halide.37 However, in this manuscript we reveal that the motion of hydrides in the hydride dimer 1 (Fig. 1) can surprisingly cause the iron(II) sites to flip between high-spin and intermediate-spin electronic configurations rapidly. The intermolecular interactions, ligand-field changes, and hydride motions that bring about this spin-state change are elucidated using X-ray and neutron crystallography, magnetism, spectroscopy, and computations. These data illustrate that hydride ligands are unique because of their rapid motions and strong ligand field, a combination that has implications for the roles of hydrides in catalysis.
Fig. 1 (a) Ball-and-stick model showing the core of the X-ray crystal structure of 1 at 223 K,35 emphasizing that the β-diketiminate supporting ligands are perpendicular to one another. The pairs of Fe centers and the pairs of hydrides are rendered crystallographically equivalent by the operator . (b) Chemical structure of 1, showing the composition of the diketiminate ligands. (c) Two views of the previously reported X-ray crystal structure of 1 at 223 K (left: along the 2-fold axis; right: along the Fe–Fe axis), with the Fe and N atoms as thermal ellipsoids. The residual electron density (Fobs–Fcalc) is shown as a green wireframe. |
Our previous analysis using the refined hydride positions had been that each iron(II) site has a geometry that lies between square planar and tetrahedral.35,381H NMR spectra of 1 show the expected number of peaks for equivalent diketiminate environments, and cooling a sample in toluene-d8 to 188 K showed no evidence of dynamic behavior. The solution effective magnetic moment at room temperature in C6D6 is 7.0(4) Bohr magnetons (μB) per dimer, which was consistent with the previous assessment that 1 had two equivalent S = 2 subsites with little to no exchange coupling (two uncoupled S = 2 ions predict a spin-only value of 6.9 μB).39 This model was supported by the Mössbauer spectrum of a solid at 170 K (Fig. 2, top), which showed a doublet with an isomer shift (δ) of 0.48 mm s−1 and a quadrupole splitting (ΔEQ) of 1.25 mm s−1, which are similar to those in low-coordinate iron(II) alkyl complexes that have high-spin ground states.40
However, the measurements at lower temperature are inconsistent with equivalent iron sites. Namely, cooling single crystals of 1 below 170 K leads to splitting in the X-ray diffraction data and in the Mössbauer spectrum (Fig. 2). The crystallographic data indicate a change in the unit cell, causing the peaks in the diffraction pattern to gradually split as the temperature is lowered (orange insets in Fig. 2). At 130 K, the splitting is complete, and the data from a sample measured at 93 K (1-LT) fit best to a twinned structure in space group P (Z = 2). The unit cell parameters of this smaller unit cell are similar to the primitive setting of the C2/c space group of the higher-temperature model (crystallographic details in Fig. S1–S5‡).
Importantly, in 1-LT, the molecule has lost the crystallographic 2-fold axis that related the two halves of the molecule (Fig. 3). Accordingly, there is a change in the positions of the small Fourier peaks corresponding to the two bridging hydrides. However, since hydride locations are unreliable in X-ray crystallography, we sought to corroborate this structural shift using other techniques. Unfortunately, we were unable to grow a crystal of this packing form that had sufficient size for neutron crystallography.
Fig. 3 (Left) Overlay of X-ray crystal structure models of 1 at 223 K (red)35 and 1-LT at 93 K (blue). (Right) Overlay of X-ray crystal structure models of 1-LT at 93 K (blue) and 1-toluene at 100 K (grey). In 1, the hydrides are in (incorrect) averaged positions, whereas in 1-LT and in the structure with co-crystallized toluene (1-toluene), the hydrides are in the plane of the left Fe-diketiminate plane. Mercury files containing the overlays are available in the ESI.‡ |
The 1:1 ratio of signals at low temperature suggests that the two iron atoms in 1 are inequivalent, and the difference in isomer shifts (δ = 0.43 mm s−1vs. 0.60 mm s−1) suggests that the iron environments have different electronic structures. Further, the change in the Mössbauer spectrum at the same temperature as the change in the crystal structure indicates that this electronic structure change is linked to the crystallographic packing. We tested this idea by measuring Mössbauer spectra of flash-frozen benzene solutions. A frozen benzene solution of 1 displays only one doublet, even at 80 K. When the benzene is evaporated from this solution, the spectrum of the resulting solid at 80 K again shows two equal doublets (Fig. 4). The lack of dynamic behavior in the absence of the crystal lattice indicates that somehow the twinning of the crystals leads to the splitting of the doublets in the Mössbauer spectrum.
We further fit the dynamic behavior by considering the quadrupole splitting values, which arise from the electric field gradients (EFGs) at the two iron sites. At temperatures well below the phase transition (<100 K) they are different, and above the phase transition (>180 K) they are the same on the Mössbauer time scale (ca. 10−7 s). The coalescence behavior did not fit to a physically meaningful Boltzmann model (incoherent interconversion of sites), but they did fit to a cooperative transition (see ESI‡ for discussion). We used the Sorai–Seki domain model42 for a change from a static low-temperature phase to a dynamic high-temperature phase, the latter of which has fast interconversion of different possible electronic environments of the iron sites. This model has been used for entropy-driven phase transitions of mononuclear spin crossover complexes.43 When we adapted it for “hopping” between the environments of two iron sites, the observed temperature-dependent quadrupole splittings fit well to this model with an enthalpy factor nΔH ∼1500 cm−1 (n represents the size of the domain; see ESI‡). In this model, the barrier for “hopping” arises from the elastic coupling of molecules in the solid, which requires a phonon coupled to the molecular motion in order to overcome the barrier. If there is no coupling between molecules (n = 1), the Sorai–Seki model predicts a very low barrier of ∼4 kcal mol−1.
The crystals of 1-toluene are substantially larger, enabling the use of neutron diffraction to verify the positions of the hydrogen atoms unambiguously. The neutron crystal structure of 1-toluene (Fig. 5) shows the marked difference between the coordination geometries of the two inequivalent iron atoms. Fe1 has a square-planar geometry (5.6(1)° between the NFeN and HFeH planes), and the Fe1–H distances of 1.62(1) and 1.64(1) Å are significantly shorter than the Fe2–H distances to pseudotetrahedral (89.7(1)° between planes) Fe2 at 1.71(1) and 1.73(1) Å. The parameter τ4, which quantifies the geometry of a metal center on a scale from 0 (square planar) to 1 (tetrahedral) is 0.06 for Fe1 and is 0.80 for Fe2. The Fe–N distances at Fe1 are also shorter (1.924(4) and 1.925(3) Å) than to Fe2 (1.936(3) and 1.967(3) Å), suggesting that Fe1 has a lower spin state than Fe2.
The crystal structures of 1-toluene show that the toluene of crystallization is stacked above the β-diketiminate of the square-planar Fe1, with its methyl group wedged between the xylyl groups of the supporting ligand (methyl carbon 3.7 Å from plane, and H 2.7 Å from plane). This toluene molecule in the crystal does not change the electronic structure, as shown by the similarity of the Mössbauer parameters between crystals of 1-LT and 1-toluene (Fig. S9‡). Instead, the toluene molecule in the crystal influences the relative energies of the two geometries, as shown in the bottom of Fig. 6. In this model, the presence of a flat toluene molecule near only one of the two iron sites “freezes out” one of the two conformational possibilities. This differs from toluene-free crystals of 1, where the two conformations (Fe1SP/Fe2Td and Fe1Td/Fe2SP) can interconvert with very little motion of the supporting ligands (Fig. 3).
Fig. 6 A model of hydride motions in 1 that explains the accumulated spectroscopic, crystallographic, and magnetic data. |
This logic leads to the following overall explanation for both the Mössbauer and crystallographic observations in 1 (Fig. 6). At higher temperatures in the higher-symmetry space group C2/c, the iron and hydride positions are crystallographically required to be equivalent due to the 2-fold rotation axis, but the “observed” positions are an artifact arising from multiple geometries of the molecular cores in these crystals. Near 150 K, there is a phase change to a lower-symmetry space group, which removes the crystallographic 2-fold axis. Half of the domains rotate the hydrides in each direction, which leads to twinning in 1-LT (Fig. S4 and S5‡). The observation of 50:50 twinning suggests that this desymmetrization follows the same rotation within domains, rather than alternating molecules in the crystal. This domain-wide motion requires coherent reorganizations (phonons), which enable the molecules to overcome the barrier.
In 1-toluene, the two Fe sites are crystallographically inequivalent because of the packing of the toluene molecule above one of the two iron-diketiminate units. There is no symmetry breaking of the molecule with temperature in crystals of 1-toluene, since the iron sites are already distinct at all temperatures in the crystal due to the location of the toluene. This explains why 1-toluene crystals show two quadrupole doublets in Mössbauer spectra at all temperatures.
In contrast, in the flash-frozen benzene solution the molecules have disorganized molecular surroundings. Because there is no cooperativity in the interconversion of the conformers in the crystal, the barrier is low for fast interchange of the local electronic environment between Td and SP geometries (τ ≪ 10−8 s) and therefore averaged Mössbauer spectra are seen at all temperatures. This model is consistent with DFT computations below, which show that the barrier to rotation of the Fe2H2 plane within the ligand scaffold is only 3 kcal mol−1 in the absence of intermolecular interactions.
Finally, to evaluate the dependence of the dynamics on the isotope of hydrogen, we compared the temperature dependence of the Mössbauer spectrum for crystals of 1 and its deuteriated analogue 1-D (Fig. S12 and S13‡). The observed coalescence temperatures were the same for the hydride and deuteride within 3 K. The lack of an observable kinetic isotope effect supports the idea that the phase change is the driver of the temperature-dependent changes, whereas motion of the hydrides is rapid in the absence of the twinned 1-LT crystal packing. We also collected a neutron structure of the toluene crystals from 1-D, and the metrical parameters match the protiated structure within standard uncertainty limits, despite the potential for differences between the size of H and D.44
Fig. 7 Temperature dependence of the molar magnetic susceptibility times temperature, χMT, of 1 recorded under an applied field of 1 T, and inverse temperature dependence of the magnetization recorded with 1, 4 and 7 T (inset). The experimental data were corrected for a TIP-like contribution to χ of 600 × 10−6 cm3 mol−1. The colored lines represent a global SH simulation using STd = 2, SSP = 1, J = +63 cm−1, gTd = (2.47, 2.47, 2.61), gSP = (3.65, 3.65, 1.0), DTd = −26.1 cm−1, and DSP = +40 cm−1. The rhombicity parameters were constrained to and the D and g matrices of the FeSP site are rotated with respect to the principal axes for the FeTd site by an Euler angle β = 98° in this model. See ESI‡ for details of the model. |
We could fit these data to three models: (a) two S = 2 sites with ferromagnetic coupling and a diamagnetic impurity, (b) one S = 2 subsite and one S = 1 subsite, or (c) one delocalized S = 3 system. The first model did not fit our magnetic Mössbauer data; see ESI‡ for details and description of the alternative model. The second model yielded STd = 2, gTd = 2.47, 2.47, 2.61, and DTd = −26 cm−1, and SSP = 1, gSP = 3.65, 3.65, 1.0, and DSP = +40 cm−1, with exchange coupling constant J = +63 cm−1 (using −2J STd·SSP). The model required different orientations of the two local D and g matrices, as shown in Fig. 8. Spin projection arguments show that the zero-field splitting of the tetrahedral site DTd is the main source of magnetic anisotropy (see ESI‡ for details).
Aligning the gTd and gSP matrices and the local D-tensors as shown in Fig. 8 reproduced the experimental over-shooting of χMT(T) around 20 K (red fit in Fig. 7). The large g components along zTd strengthen the magnetization of the (z-polarizable) high ms levels, which are exclusively populated at low temperatures. In contrast, lower gTd values in the x/y direction reduce the magnetization of the other levels with lower ms that become considerably populated above 20 K. This difference in magnetization of the ms levels causes the observed overshooting of χMT(T) with temperature, rather than the more typical assignment of overshooting to ferromagnetic coupling.
We also modeled the magnetization in the context of an isolated Stotal = 3 system. This model also provides a reasonable global fit of magnetic susceptibility and magnetic Mössbauer data (Fig. S17‡). The zfs parameter of the S = 3 model, D = −11 cm−1, is consistent with DTd = −26 cm−1 in the localized model (due to the projection coefficient 2/5). In the delocalized model, the over-shooting of χMT(T) around 20 K is caused entirely by strong g anisotropy, g = (1.4, 1.4, 2.6). The same behavior was previously observed for an S = 1 {FeNO}8 species45 and a two-coordinate S = 2 imidoiron(II) species,46 both of which feature large negative D and gz larger than 2. Large and anisotropic shifts in g values have been observed previously in mononuclear S = 1 square-planar iron(II) compounds, such as [FeII(TPP)].47 Further details of the various models may be found in the ESI.‡
To elucidate the physical origin of the strong magnetic anisotropy of 1, it is necessary to explicitly consider spin–orbit coupling (SOC) of its low-lying electronic states. To this end, we carried out wavefunction-based CASSCF/NEVPT2 calculations (Fig. 9). The reference frame was chosen with the z axis along the Fe–Fe vector, and the xz and yz planes as the NTdFeTdNTd and NSPFeSPNSP planes, respectively. The active space spanned both Fe sites, and included all Fe 3d orbitals of the FeTd and FeSP sites as well as the σ-bonding counterpart of the high-lying FeSP dyz based orbital (not shown in Fig. 9b). These CASSCF(14,11) calculations predicted that 1 has three S = 3 states within 700 cm−1, which indicates that the system possesses an orbitally near-degenerate ground state. (As a consequence, we did not succeed in converging single-root ground-state CASSCF calculations as CASSCF computations suffered from severe convergence problems, but had to average the three low-lying septet states.) The septet ground state of 1 has substantial multireference character, because each of the electron configurations accounts for less than 16% of the wavefunction.
Fig. 9 (a) Orientation of the D tensor, and (b) active orbitals of 1, obtained from CASSCF(14,11) calculations averaging three low-lying S = 3 states. The orbital labels follow the localized axes of Fig. 8. |
The CASSCF(14,11)/NEVPT2 computations give D = −12.9 cm−1 and E/D = 0.29, which is in reasonable agreement with experiment (Dtotal = −11 cm−1). The largest component (Dzz) is along the z axis (Fe–Fe vector), and the smallest Dxx and medium Dyy components as defined in Fig. 9a. The principal axis systems of both D and g matrices are calculated to be almost collinear and gx,y,z = 1.97, 2.29, 2.60 with respect to the reference frame. The three lowest-energy excited states (Table S5‡) make the dominant contributions to D. The first excited state lies 390 cm−1 higher in energy and largely stems from the single excitation FeTd dxy → dx2–y2. Hence, the FeTd site is the main contributor to an orbitally nearly degenerate ground state. The in-state SOC introduces a negative D value at the FeTd site, and therefore leads to its easy-axis magnetic anisotropy along the z axis in agreement with the fitting of the magnetic data presented above. A closely related mononuclear pseudotetrahedral high spin ferrous complex, LFeCl2Li(THF)2 (L = β-diketiminate), also features an analogous electronic structure and hence similar magnetic properties.49 The second and third excited states were computed to lie 770 and 1250 cm−1 above the ground state, respectively, and predominantly arise from single excitations of FeSP dx2–y2 → dxz and dx2–y2 → dxy. The SOC of these excited states with the ground state leads to considerable orbital angular momentum in the yz plane, in agreement with the rotation of the FeSPD tensor presented above. Furthermore, because the aforementioned excitations are all DOMO-to-SOMO transitions (DOMO = doubly occupied molecular orbital, SOMO = singly occupied molecular orbital), the resulting g-shifts are positive and agree with the experimentally observed gy and gz values that are considerably greater than 2. The fact that the three dominant excitations are each localized to one of the two sites adds weight to a localized model of the electronic structure, as discussed below.
The DFT calculations also give information about how the electronic structure responds to geometric changes. At a dihedral angle θ near zero, the lowest-energy unoccupied molecular orbital (LUMO) of 1 is mostly derived from a FeSP dyz orbital with limited contribution from the FeTd dyz orbital. With increasing θ, the FeTd dyz parentage gradually increases at the cost of FeSP dyz. This change culminates when θ = 45°, at which both fragment orbitals make approximately equal contributions to the LUMO. Because of this smooth transition, the exchange of the local electronic configurations between the two Fe centers does not require electronic excitation, and proceeds with a low barrier.
However, this picture assumes a spin Hamiltonian model with ferromagnetic coupling between two discrete spins (model b from "Spin states of the iron sites" above). This model fit all of the available data with J = +63 cm−1. However, given the strong electronic coupling between the sites, it is also possible to view the septet ground state as a single S = 3 spin system that is delocalized across both metal sites (model c above). This alternative model did not fit the magnetic susceptibility data quite as well, as it corresponds to an unmeasurably large value of J, but there could also be limitations due to the zfs model we used. Though the CASSCF natural orbitals above appear delocalized, the in-plane FeSP dyz orbital is unpopulated in the ten lowest-energy S = 3 states within 3000 cm−1, as expected from qualitative ligand field theory for a square-planar FeSP site. As noted above, the lowest-lying excitations were localized on the tetrahedral FeTd site. Thus, aspects of the electronic structure are localized. Overall, this system is at the borderline of systems that are best treated with a delocalized model, which explains spin alignment through Hund's rule as in the multi-iron complexes of Betley.50 Here, mixing is facilitated by the short distance between the metals (2.45 Å, close to Pauling's Fe–Fe single bond distance of 2.48 Å).51 Berry has described a dinickel complex in which two bridging hydrides similarly enforce close metal–metal contacts that lead to a non-intuitive ground spin state.52
The geometry, electronic structure and magnetism of 1 bear a close similarity to a bis(carbene)-supported diiron system recently reported by Smith and coworkers (2), which has two hydride bridges.53 Though neutron crystallography was not used, an X-ray crystal structure suggested one square-planar iron(II) site and one tetrahedral iron(II) site as shown here, and the Mössbauer spectra strongly supported this interpretation. Like 1, their compound had an S = 3 ground state that was described through a ferromagnetic coupling model, but the fitted exchange constant was J = +110 cm−1. This corresponds even more closely to a fully delocalized septet. They also used high-field EPR spectroscopy to derive an overall D = −7 cm−1 which is somewhat less than in 1 (−11 cm−1), presumably because the stronger-field carbene ligands more effectively quench the orbital angular momentum. However, the frontier orbitals (derived from DFT in their work) bear great resemblance to the ones derived here. In the bis(carbene) supported system, the structure and Mössbauer doublets showed distinct iron sites at all investigated temperatures, and thus were not subject to the phase change dependence described here.
The most unusual aspect of our findings is the topotactic transition54,55 around 160 K, in which the low-symmetry twinned unit cell at low temperature changes to a higher-symmetry untwinned unit cell with crystallographic equivalence between the two iron sites. At low temperature, the crystallographic differentiation of the iron sites gives an “extra” barrier for hydride motion, because trading the geometries of the two sites cannot be done one molecule at a time. Rather, movement of the hydrides to trade site geometries requires a coordinated motion of an entire domain! This may seem surprising since the hydrides are small and the supporting ligands do not change orientations, but there are substantial differences between the Fe–N(diketiminate) bond lengths of the tetrahedral and square planar sites, which are apparently sufficient to influence the intermolecular interactions. Then, the relaxation of crystal packing constraints above 160 K removes the intermolecular forces, giving a low barrier for hydride movement and coalescence on the Mössbauer timescale.§ Both DFT and extrapolation of the Sorai–Seki model suggested a very low barrier of <4 kcal mol−1 for hydride rotation around the Fe–Fe axis in the absence of crystal packing forces. Quantum tunneling would be expected to affect the barrier, but we were unable to assess it experimentally because the rates measured here arose solely from the phase change; without this constraint, the hydride motion was too rapid to measure using methods at our disposal.
It is interesting to speculate about a connection to recent work on the FeMoco of nitrogenase, in which reduced forms E2 and E4 have been shown to have bridging hydride ligands and have accordingly been termed E2(2H) and E4(4H).11–14 In particular, the E4(4H) state responds to N2 binding by undergoing reductive elimination of H2 from these hydrides, which provides an essential driving force to move along the pathway toward ammonia formation.13 However, research on nitrogenase rarely considers the ligand field that is induced by these hydrides.14,58–61 We suggest that hydrides on the FeMoco might offer a powerful combination of (a) extreme mobility56 and (b) ability to strongly modulate the relative energies of iron orbitals (perhaps even changing the iron sites from the high-spin configuration of the resting state to a lower-spin configuration).59 In both nitrogenase and in future catalytic systems, we suggest that hydrides have the potential to be used as supporting ligands for rapid access to diverse electronic structures and spin, and thus their role may transcend the reactions that hydrides themselves perform.
Footnotes |
† This article is dedicated to the remarkable Eckhard Bill (1953–2022), who will be sorely missed. |
‡ Electronic supplementary information (ESI) available: Experimental and computational details, crystallographic files, Mercury overlay, further analysis of magnetic and electronic structure models. CCDC 1940222–1940225. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2sc06412j |
§ The shapes of the Mössbauer signals were the same when approaching a given temperature from the warmer or colder direction. Thus the phase change in 1 is fully reversible. |
This journal is © The Royal Society of Chemistry 2023 |