Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Nov 1.
Published in final edited form as: Polyhedron. 2021 Jul 30;208:115384. doi: 10.1016/j.poly.2021.115384

Synthesis and characterization of trigonal bipyramidal FeIII complexes and their solution behavior

Adriana Lugosan a, Sophi R Todtz a, Andrew Alcázar a, Matthias Zeller b, James J Devery III a, Wei-Tsung Lee a
PMCID: PMC8462799  NIHMSID: NIHMS1731167  PMID: 34566234

Abstract

A series of air-stable trigonal bipyramidal FeIII complexes supported by a redox non-innocent NNN pincer ligand, CztBu(PyrR)2 (R = iPr, Me, or H), were synthesized, fully characterized, and utilized for the investigation of the interaction between acetone and the FeIII center. The magnetic moments determined from the paramagnetic 1H NMR spectra in conjunction with EPR and Mössbauer spectroscopy indicate the presence of a high-spin ferric center. Cyclic voltammetry studies feature two quasi-reversible events corresponding to a metal-centered FeIII/II reduction around −0.40 V (vs. Fc) and a ligand-centered CztBu(PyrR)2/CztBu(PyrR)2•+ oxidation at potentials near +0.70 V (vs. Fc). UV-Visible spectroscopy in CH2Cl2 showcases ligand-metal charge transfer (LMCT) bands, as well as coordination of acetone to CztBu(PyrH)2FeCl2. In situ IR spectroscopy and solution conductivity (κ) measurements of CztBu(PyrR)2FeCl2 with varied equivalents of acetone reveal that acetone is weakly associated with the iron center.

Keywords: Iron complex, Carbonyl, Pincer ligand, Lewis acid, Solution spectroscopy

Introduction

Iron is the most abundant transition metal in the earth’s crust, making it a highly attractive element from which reagents and catalysts can be readily developed [1]. In particular, a great deal of effort has been focused on the development of FeCl3 as both a stoichiometric reagent and a catalyst [2]. Often, reactions with FeCl3 are performed in chlorinated solvents (i.e. CHCl3, CH2Cl2, ClCH2CH2Cl (DCE), etc.). Under anhydrous conditions, FeCl3 is largely insoluble in these solvents, and systems can require the addition of silica gel as a solid support [3], Lewis basic additives like nitromethane [4], or ligand systems [5]. Previous studies from our lab have shown that in the absence of these additives, the substrate must interact directly with heterogeneous FeCl3, resulting in the substrate increasing the solubility of the Lewis acid via the formation of a Lewis pair [6]. However, this interaction changes dramatically when FeCl3 is employed as a catalyst, allowing for the formation of complex FeIII-centered aggregates. We report herein the characterization of complexes derived from FeCl3 and NNN pincer ligands. Full ground state characterization data are provided to analyze the impact of the ligand system via electrochemical and spectroscopic analysis, as well as how the ligand perturbs the behavior of these systems in the presence of acetone in solution.

Previous efforts from our lab have focused on the examination of the impact of NNN pincer ligands on the behavior of first-row transition metals. We have examined, in depth, complexes generated with NiII [7], VIII/IV [8], as well as an FeII complex that was able to undergo C–H activation to yield an unusual pyrazolide-bridged FeII complex at elevated temperatures [9]. We sought to examine the impact of our ligand system on FeIII because of a report by Nishiyama and coworkers, who employed a structurally analogous NNN pincer ligand to facilitate the asymmetric hydrosilylation of ketones to alcohols in THF [10]. What further intrigued us about their results was that the starting material, product, and solvent are all competent Lewis bases that can compete for access to the Lewis acid. With these observations in mind, we sought to study how our NNN pincer system impacts the solubility of FeIII chloride complexes and how it may facilitate similar intermolecular interactions.

Experimental Section

Materials and Methods.

All manipulations were performed under a nitrogen atmosphere using standard Schlenk techniques or in an M. Braun UNIlab Pro glovebox. Glassware was dried at 150 °C overnight. Diethyl ether, n-pentane, tetrahydrofuran, and toluene were purified using a Pure Process Technology solvent purification system. Before use, an aliquot of each solvent was tested with a drop of sodium benzophenone ketyl in THF solution. All reagents were purchased from commercial vendors and used as received. HCztBu(PyriPr)2 was prepared according to a modified literature procedure [7]. 1H NMR data were recorded on a Varian Inova 500 MHz spectrometer at 22 °C. Resonances in the 1H NMR spectra are referenced to residual CD2Cl2 at δ = 5.32 ppm or CDCl3 at δ = 7.27 ppm. Solution magnetic susceptibilities were determined using the Evans method [11]. Continuous-wave (CW) EPR spectra were recorded at 77 K on a Bruker EMX plus X-band EPR spectrometer equipped with a liquid N2 cold-finger Dewar flask. Cyclic voltammetry was conducted via a CH-Instruments electrochemical analyzer (model 620E), employing a 3 mm glassy carbon working electrode, a silver wire pseudo reference electrode, and a platinum coiled wire counter electrode. All measurements were performed using either CH2Cl2 or THF solutions containing 1 mM analyte and 0.1 M n-Bu4NPF6 as the supporting electrolyte. The potentials were referenced to a ferrocene/ferrocenium redox couple. Elemental analyses were conducted by Midwest Microlab, LLC (Indianapolis, IN).

Synthesis of HCztBu(PyrH)2.

10.0 g (23 mmol) of 1,8-dibromo-3,6-di-tert-butyl-9H-carbazole, 7.8 g (114 mmol) of 1H-pyrazole, 12.8 g (114 mmol) of potassium tert-butoxide (KOtBu), and 2.3 g (20 mmol) of N,N,N,N-tetra-methyl-ethylenediamine (TMEDA) were dissolved in 300 mL dimethylformamide (DMF). The yellow DMF slurry was degassed by three freeze-pump-thaw cycles. 2.9 g (20 mmol) of copper(I) oxide (Cu2O) was added and the reaction was refluxed at ~150 °C for 2 days under inert N2 atmosphere. The resulting brown slurry was dissolved in 200 mL dichloromethane, washed with 1 M HCl (3×200 mL), and then vigorously stirred with 200 mL 1 M HCl overnight. The organic layer was again washed with 3×200 mL of 1M HCl, 3×200 mL of 1 M NH4OH, and 2×200 mL of 1 M NH4Cl solutions. The dichloromethane layer was collected, dried using MgSO4, and solvent was removed by rotary evaporation. The resulting light brown solids were crystallized from a concentrated n-hexane solution at −20 °C to yield the product as off-white solids (4.9 g, 52%). 1H NMR (500 MHz, CDCl3, δ): 1.54 (C(CH3)3, s, 9H), 6.57 (ArH, s, 1H), 7.61 (ArH, s, 1H), 7.93 (ArH, s, 1H), 8.09 (ArH, s, 1H), 8.17 (ArH, s, 1H), 11.31 (NH, s, 1H). Anal. Calcd for molecular formula C26H29N5: C 75.88, H 7.10, N 17.02. Found: C 75.93, H 7.06, N 16.85.

Synthesis of HCztBu(PyrMe)2.

15.0 g (34 mmol) of 1,8-dibromo-3,6-di-tert-butyl-9H-carbazole, 8.44 g (103 mmol) of 3-methylpyrazole, 11.54 g (103 mmol) of KOtBu, and 3.50 g (30 mmol) of TMEDA were dissolved in 370 mL of DMF. The yellow DMF slurry was degassed by 3 freeze-pump-thaw cycles. 5.39 g (38 mmol) of Cu2O was added, and the reaction was refluxed at ~150 °C for 3 days under inert N2 atmosphere. The resulting red/brown slurry was dissolved in 500 mL diethyl ether (Et2O) and washed with 3×200 mL of 1 M HCl, 3×200 mL of 1 M NH4OH, and 2×200 mL of 1M NH4Cl solutions. The organic layer was dried with MgSO4, and volatiles were removed in vacuo. The resulting solids were crystallized from a hot n-hexane solution to yield fluffy off-white solids (9.0 g, 60%). 1H NMR (500 MHz, CDCl3, δ): 1.50 (C(CH3)3, s, 9H), 2.54 (CH3, s, 3H), 6.33 (ArH, d, J = 2.4, 1H), 7.52 (ArH, d, J = 1.5, 1H), 8.00 (ArH, d, J = 1.0, 1H), 8.04 (ArH, d, J = 2.4, 1H), 11.56 (NH, s, 1H). Anal. Calcd for molecular formula C28H33N5: C 76.50, H 7.57, N 15.93. Found: C 76.46, H 7.51, N 16.02.

Synthesis of CztBu(PyriPr)2FeCl2 (1).

To 1.00 g (0.21 mmol) of HCztBu(PyriPr)2 dissolved in 5 mL of THF at room temperature under inert N2 atmosphere was added 0.23 g (0.21 mmol) of lithium diisopropylamide (LDA) in 5 mL of THF. The resulting fluorescent yellow mixture was stirred for 1 h. The fluorescent mixture was then added to 0.36 g (0.22 mmol) of FeCl3 in 50 mL of THF and stirred overnight at ambient temperature. Volatiles were removed by rotary evaporation to afford dark green solids, which were washed with 30 mL of n-hexane and collected by vacuum filtration over a celite pad. The remaining green solids were dissolved in 100 mL of CH2Cl2 and dried to yield the final dark green product. Crystals suitable for X-ray diffraction were grown from a concentrated toluene solution at room temperature. Yield: 1.2 g, 98%. 1H NMR (500 MHz, CDCl3, δ): 6.23 (C(CH3)3), 16.71 (CH(CH3)2), 36.87 (ArH), 69.24 (ArH), 75.16 (ArH), 89.05 (ArH). μeff (CDCl3) = 5.5(3) μB. Anal. Calcd for molecular formula C32H40Cl2FeN5: C 61.85, H 6.49, N 11.27. Found: C 62.14, H 6.44, N 11.37. UV–Vis (CH2Cl2) λmax, nm (ε): 356 (12390), 427 (sh, 2230), 743 (3450). UV–Vis (acetone) λmax, nm (ε): 356 (10480), 420 (sh, 2120), 728 (2590).

Synthesis of CztBu(PyrMe)2FeCl2 (2).

The procedure was adapted from the synthesis of 1 using HCztBu(PyrMe)2 (1.00 g, 0.23 mmol), LDA (0.26 g, 0.24 mmol), and FeCl3 (0.41 g, 0.25 mmol). Crystals suitable for X-ray diffraction were grown from a concentrated toluene solution at room temperature. Yield: 1.2 g, 92%. 1H NMR (500 MHz, CDCl3, δ): 6.30 (C(CH3)3), 37.39 (ArH), 68.47 (ArH), 78.41(ArH), 88.71 (ArH). μeff (CDCl3) = 5.4(2) μB. Anal. Calcd for molecular formula C28H32Cl2FeN5: C 59.49, H 5.71, N 12.39. Found: C 59.71, H 5.50, N 12.19. UV–Vis (CH2Cl2) λmax, nm (ε): 357 (14060), 422 (sh, 2590), 739 (3790). UV–Vis (acetone) λmax, nm (ε): 351 (13590), 427 (sh, 2650), 714 (3890).

Synthesis of CztBu(PyrH)2FeCl2 (3).

The procedure was adapted from the synthesis of 1, but a more dilute concentration is required. To 1.00 g (0.24 mmol) of HCztBu(PyrH)2 dissolved in 10 mL of THF at room temperature under an inert N2 atmosphere was added 0.27 g (0.25 mmol) of LDA in 10 mL of THF. The resulting fluorescent yellow mixture was stirred for 1 hour. The fluorescent mixture was then added to 0.43 g (0.26 mmol) of FeCl3 in 100 mL of THF and stirred overnight at ambient temperature. Volatiles were removed by rotary evaporation to afford a dark green solid. The dried solids were washed with 30 mL of n-hexane and collected by vacuum filtration over a celite pad. The remaining green solids were dissolved in 150 mL of CH2Cl2 and dried to yield the final dark green product. Crystals suitable for X-ray diffraction were grown from a concentrated toluene solution at room temperature. Yield: 1.3 g, 99%. 1H NMR (500 MHz, CD2Cl2, δ): 5.53 (C(CH3)3), 32.37 (ArH), 66.33 (ArH), 80.65 (ArH), 90.81 (ArH). μeff (CD2Cl2) = 5.3(2) μB. Anal. Calcd for molecular formula C26H28Cl2FeN5: C 58.12, H 5.25, N 13.03. Found: C 58.17, H 5.17, N 12.99. UV–Vis (CH2Cl2) λmax, nm (ε): 349 (11260), 374 (sh, 8670), 405 (sh, 3490), 680 (3480). UV–Vis (acetone) λmax, nm (ε): 359 (7470), 486 (sh, 790), 931 (1650).

Crystallography.

Single crystal data were collected on either a Mo Kα wavelength (λ = 0.71073 Å) Bruker Quest diffractometer with a fixed chi angle, a sealed tube fine focus X-ray tube, single crystal curved graphite incident beam monochromator, and a Photon100 CMOS area detector (complex 1) or on a Cu Kα wavelength (λ = 1.54178 Å) Bruker Quest diffractometer with kappa geometry, a I-μ-S microsource X-ray tube, laterally graded multilayer (Goebel) mirror for monochromatization, a Photon2 CMOS area detector (complex 3). Both instruments were equipped with Oxford Cryosystems low temperature devices and examination and data collection were performed at 100 K (for 1) 150 K (for 3). Data were collected, reflections were indexed and processed, and the files scaled and corrected for absorption using APEX3 [12] and SADABS [13] or TWINABS [13]. Space groups were assigned, and the structures were solved by direct methods using XPREP within the SHELXTL suite of programs [14] and refined by full matrix least squares against F2 with all reflections using Shelxl2014 [15] or Shelxl2018 [14] using the graphical interface Shelxle [16]. Additional data collection and refinement details, including hydrogen atom treatment, description of disorder, twinning and pseudosymmetry (where present) can be found in the Supporting Information. Complete crystallographic data, in CIF format, have been deposited with the Cambridge Crystallographic Data Centre. CCDC 1996921–1996922 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

Computational Details.

Geometry optimization was carried out using Spartan ‘16 [17]. The DFT modelling method, using the hybrid B3LYP functional with the 6–31G* basis set, was used to calculate the model complexes 1’3’ and [3’-acetone]+, where the tBu groups were truncated to an H atom. Geometry optimization was carried out until global minima were achieved. The results of the optimized structures are reported in Tables S4S7 and are comparable with X-ray diffraction results. The atomic contributions to frontier molecular orbitals and simulated electronic spectra were calculated using the Chemissian program [18].

Results and Discussion

Synthesis and Characterization of Complexes 13

Dark green complexes 13 can be prepared in high yield (>90%) by deprotonation of HCztBu(PyrR)2 (R = iPr (1), Me (2), H (3)) using LDA followed by transmetallation with FeCl3 at ambient temperature in a N2 atmosphere (Scheme 1). Unlike 1 and 2, formation of 3 is dependent upon concentration during synthesis based on 1H NMR spectroscopy; since the coordination site of 3 is more exposed than in 1 or 2, the use of dilute concentrations prevents the formation of undesired byproducts, e.g. [{CztBu(PyrH)2}2Fe]Cl. 1H NMR spectroscopy of 13 shows broad, paramagnetically shifted resonances ranging from 5.53 to 90.81 ppm in various deuterated solvents. Note that one resonance is missing for each complex, which could be due to the fast relaxation of protons that are close to the FeIII center.

Scheme 1.

Scheme 1.

Synthesis of 13.

Solution magnetic measurements in CDCl3 (for 1 and 2) and CD2Cl2 (for 3) at room temperature (298 K) using Evans method gave μeff = 5.3(2), 5.4(2), and 5.5(3) μB for 13, respectively, all of which are slightly lower than the expected value for high-spin FeIII systems and may be caused by mixing of another spin state. Complexes 13 are highly soluble in various organic solvents and 1 is even soluble in n-pentane. In the solid state, 13 are all dark green powders. Interestingly, 1 and 2 exhibit the same color when dissolved in both coordinating (acetone) and non-coordinating (CH2Cl2) solvents; whereas, we observed dark green and teal solutions when dissolving 3 in CH2Cl2 and acetone, respectively (vide infra). Finally, complexes 13 are air-stable indefinitely and resistant to decomposition upon heating up to 80 °C.

Crystal Structures of 1 and 3

We obtained single crystals suitable for X-ray structure determination for complexes 1 and 3 (Fig. 1) [19]. The solid-state structures confirm complexes 1 and 3 as mononuclear iron complexes with one NNN pincer ligand in association with two Cl atoms. Both 1 and 3 have a slightly distorted trigonal bipyramidal geometry (τ5 values are 0.86 and 0.83, respectively) with two neutral pyrazole-nitrogen atoms (Npyr) in the axial positions and three anionic atoms in the equatorial plane [20]. The structural parameters of 1 and 3 agree well with those of closely related carbazolide-based pincer-supported FeIII chloride complexes, except for slightly shorter iron to carbazolide nitrogen (Ncz) distances [21]. This could be attributed to the stronger electron-donating tert-butyl groups, compared to previously reported methyl and phenyl groups on the carbazolide backbone. Different from complex 3, the carbazolide and two flanking pyrazole rings in 1 are not coplanar. This may be caused by steric hindrance resulting from the bulkier iso-propyl substituents or close contact of the iPr C−H groups to the Cl ligands [22].

Fig. 1.

Fig. 1.

Molecular structure of 1 with thermal ellipsoids at the 50% probability level. Hydrogen atoms and solvent molecules are omitted for clarity. Color key: orange = Fe, blue = N, gray = C, green = Cl [19].

The dihedral angles between the pyrazole groups and the carbazolide ring are 28.7(9)° and 0° for 1 and 3, respectively (Fig. 2). Surprisingly, we observed a lesser distortion (17.7(2)°) for the VIII congener, CztBu(PyriPr)2VCl2,[8] despite comparable ionic radii for both FeIII and VIII [23].

Fig. 2.

Fig. 2.

Overlaid structure of complexes 1 (blue) and 3 (red). Spheres represent Fe atoms. Hydrogen atoms and tert-butyl groups are omitted for clarity [19].

Electrochemical Studies

Cyclic voltammetry analysis of the FeIII complexes carried out in CH2Cl2 and acetone solutions using 0.1 M n-Bu4NPF6, a platinum wire auxiliary electrode, glassy carbon working electrode, and Ag wire reference electrode gave insight into the electrochemical behavior. We varied scan rates from 100 mV s−1 to 500 mV s−1 and referenced to [FeCp2]/[FeCp2]+. In CH2Cl2, two distinct quasi-reversible events represent a diffusion-controlled process for complexes 13 (Fig. 3): the peak current is proportional to the square root of the scan rate (Fig. 4) [19].

Fig. 3.

Fig. 3.

Cyclic voltammograms of 0.1 mM of 1 (blue), 2 (green), and 3 (red) in CH2Cl2 (0.1 M n-Bu4PF6) at scan rates of 100 mV s−1 [19]. The arrow indicates the initial scan direction.

Fig. 4.

Fig. 4.

Cyclic voltammograms of 0.1 mM 3 in CH2Cl2 (0.1 M n-Bu4PF6) at scan rates of 100, 200, 300, 400, and 500 mV s−1. Inset: Plot of anodic, Ipa, and cathodic, Ipc, peak current, versus square root of scan rate for the first (•, blue; R2 = 0.9992) and second oxidation (◆, red; R2 = 0.9650) [19]. The arrow indicates the initial scan direction.

Complex 3 exhibited E1/2 values of −0.41 and +0.72 V, which are assigned as metal-centered (FeIII/II) and ligand-centered CztBu(PyrH)2/CztBu(PyrH)2•+ redox couples, respectively. The plots of v1/2 vs. the Ipa and Ipc for both events suggest these electron transfer processes are governed by diffusion control.

Voltammograms similar to those for 3 obtained for 1 and 2 yielded less negative redox potentials (−0.38 V for 1 and −0.39 V for 2) for the metal-centered redox step, which are conversely related to the increasing electron donation from the pyrazole substituents to the Fe center. The same unusual trend shared by 1 and 2 is present for the ligand-centered oxidation, where the redox potentials of 1 (0.70 V) and 2 (0.69 V) are less than 3 (0.72 V). This contradiction could be attributed to the poor overlap between the p orbital of Ncz and d orbital of Fe, resulting from the increased out-of-plane distortion of the Fe from the carbazolide plane in 1. Complex 3 is the most reversible while 1 is the least. The decrease of reversibility of 1 is consistent with out-of-plane distortion. In acetone, complexes 13 exhibit more complicated redox events, which could be due to multiple species in solution, owing to the solvent-specific coordination of acetone to the Fe center.

Electron Paramagnetic Resonance and Mössbauer Spectroscopy

EPR spectroscopy of 13 was conducted in both the solid and solution state at room temperature and 77 K [19]. In the solid state, complex 1 gives spectra with g values near 4.3 and 2.0 at room temperature and 77 K. When a CH2Cl2 solution of 1 is measured at room temperature, a large g = 2.0 signal is accompanied by a diffuse wing stretching at g = 4.3 (Fig. 5); whereas, the intensity of the signal at g = 4.3 is larger than the g = 2.0 signal at 77 K implying the presence of a high-spin FeIII center.

Fig. 5.

Fig. 5.

EPR spectra of 1 in CH2Cl2 at room temperature (top) and 77 K (bottom) [19].

Room temperature EPR measurement of complexes 2 and 3 as solids and solutions exhibit spectra with a g value near 2. The solid-state spectra are considerably broader in comparison to the solution state spectra, which may be due to multiple transitions coinciding at the same time resulting in an imperfect rhombic shaped signal. Measurement of 2 and 3 as solids and CH2Cl2 glasses at 77 K gave spectra with similar g ≈ 2 values with an additional small signal around g ≈ 4, which is typical of a high-spin FeIII center [19].

The differences in EPR spectra of 1, 2, and 3 in CH2Cl2 can be attributed to differences in overall steric bulk and the corresponding out-of-plane distortion imparted by the -R groups on the pyrazole ring. -iPr analogue 1 has the most steric bulk and least free rotation, which decreases the zero-field splitting average. When the -iPr substituent is replaced by -CH3 (2) or -H (3), the out-of-plane distortion decreases and allows the free rotation and zero-field splitting averaging to increase. This may account for the differences observed in the EPR spectra and consequent trends observed in cyclic voltammetry and UV-Visible spectroscopy. The range of EPR spectra for 1-3 observed in the solid state may arise from the presence of intermolecular interactions and/or an imperfect rhombic environment, causing several transitions to coincide at the same time and make precise interpretation of the solid-state spectra challenging.

Mössbauer spectroscopy of 1 at 80 K gives rise to an asymmetric spectrum with an isomer shift (δ) of 0.42 mm/s and a quadrupole splitting (ΔEQ) of 1.18 mm/s (Fig. 6), typical of non-integer spin systems. Increasing the temperature to 294 K produces a spectrum with a slightly lower δ value (0.32 mm/s) and a slightly larger ΔEQ (1.27 mm/s). The δ value agrees with previously reported quantum spin-admixed systems, while the ΔEQ values are consistent with the presence of a high-spin ferric iron center [24].

Fig. 6.

Fig. 6.

Mössbauer spectrum of 1 at 80 K (top) and 294 K (bottom).

At low temperatures, complexes 13 can be assigned as high-spin FeIII species, although they do possess some similarities to previously reported pincer and porphyrin-based spin-admixed (S = 3/2, 5/2) and spin crossover systems [24c, 25, 26]. Overall, the results of solution magnetic susceptibility measurements, EPR, and Mössbauer spectroscopy lead us to conclude 13 are most likely high-spin ferric complexes.

Electronic Absorption Spectra

The UV-Vis spectra of complexes 13 in CH2Cl2 and acetone are shown in Fig. 7. “CztBu(PyriPr)2 ” pincer complexes typically have π → π* transitions below the 400 nm region, and the lowest absorption bands are best described as ligand-to-metal charge transfer bands (LMCT) [7].

Fig. 7.

Fig. 7.

Electronic spectra of 6.66 × 10−2 mM of 1 (blue), 2 (green), and 3 (red) in CH2Cl2 (top) and acetone (bottom) [19].

The position of the LMCT bands shifts toward shorter wavelengths (3 (680 nm) < 2 (739 nm) < 1 (743 nm)), when there are less electron donating substituents on the pyrazole groups. Although it is well understood that the ligand field is enhanced with electron donating substituents, the same kind of reverse trend has been observed in the cyclic voltammograms that led to conclude that the extent of out-of-plane distortion of Fe dictates the degree of metal-ligand orbital overlap.

In acetone, complexes 1 and 2 exhibit a slight blue shift in the visible range, e.g. 743 → 725 nm and 739 → 712 nm for 1 and 2, respectively; whereas, the spectrum of 3 shows a drastic red shift from the visible to the near infrared region 680 → 939 nm. We believe that the drastic red-shift is not simply due to the polarity of solvents, but rather the effect of acetone coordination to the Fe center in 3, under forcing conditions. However, when 3 is exposed to 10 equiv acetone in CH2Cl2 solution, this spectral feature is not observed, consistent with a lack of acetone binding. The observed disparate spectra for complex 3 are consistent with a concentration-dependent interaction of acetone with 3.

Solution Behavior

To examine the intermolecular interactions further, we turned to spectroscopic investigation under synthetically relevant conditions. A variety of methods have been developed to measure the binding ability of Lewis acid catalysts such as NMR spectroscopy [27], UV-Vis spectroscopy [28], selectivity [29], and mass spectrometry [30]. More specifically, chemists have employed infrared (IR) spectroscopy when examining the interactions of paramagnetic Lewis acids and carbonyl-containing compounds, allowing for the determination of Lewis acidity [31]. To probe the solution interactions of Lewis acids and carbonyls, we have developed a titration-based method, utilizing in situ IR as a detector [32]. With this method, we benchmarked a range of Lewis acid/carbonyl interactions in solution [33] and were able to employ these benchmarks to gain insight into the mechanism of FeIII-catalyzed carbonyl-olefin metathesis [6]. Because of these benchmarks, we chose to examine Lewis acidity of our FeIII complexes via their interactions with acetone in DCE.

Complexes 1, 2, and 3 are highly soluble in DCE, and we performed titrations into homogeneous mixtures. When acetone is added to a solution of 1 from 0–4 equiv, we observe two vibrations: one at 1714 cm−1, which is consistent with uncoordinated acetone, and an additional shoulder at approximately 1690 cm−1 (Fig. 8). This extra vibration at lower wavenumbers, in the direction of the known C–O vibration, is suggestive of a slight decrease in the π-character of the C=O of acetone, consistent with previous observations of carbonyl coordination to a Lewis acid (Fig. 8) [6, 25]. When analogous titrations were carried out with 2 and 3, only unbound acetone at 1714 cm−1 is observed in the IR spectra, consistent with no interaction [19].

Fig. 8.

Fig. 8.

Solution IR data for titration of 1 (1 mmol in 6 mL DCE) with 0–4.2 equiv acetone. Titration proceeds from black to violet with increasing amounts of acetone ([acetone] = 0 M, 0.089 M, 0.178 M, 0.243 M, 0.330 M, 0.499 M [19].

We have previously demonstrated that when ≤1 equiv acetone is added to FeCl3, a monomeric complex forms in which one molecule of acetone is bound to one molecule of FeCl3 and no uncoordinated acetone is observed [6]. This 1:1 acetone/FeCl3 coordination complex exhibits a single characteristic vibration in the carbonyl region at 1633 cm−1. The observation of exclusive formation of coordination complex in the absence of unbound acetone is representative of a high affinity Lewis acid/base interaction; whereas, the simultaneous observation of a coordination complex and unbound acetone is consistent with a low affinity Lewis acid/base interaction [6, 33]. When the titration of FeCl3 proceeds beyond 1 equiv, the 1:1 acetone/FeCl3 coordination complex is consumed and higher wavenumber vibrations are observed in the region between the initial Lewis pair and unbound acetone. This change in IR is consistent with the formation of a mixture of two highly ligated complexes, with multiple Lewis basic carbonyls coordinating to Fe [6, 33, 34]. Using the analogy of our previous observations, the simultaneous growth of uncoordinated acetone with the shoulder at 1690 cm−1 is consistent with a low affinity interaction between 1 and acetone. Further, we have previously shown that this low affinity interaction can arise from a trigonal bipyramidal complex with a more distantly associated acetone (Fig. 9, left) [34]. Because of the analogous geometry and observed IR band, we propose an analogous solution interaction (Fig. 9, right).

Fig. 9.

Fig. 9.

Theoretically predicted C=O IR stretching of 4:1 FeCl3:acetone solution aggregate (left) [34]. Hypothesized analogous solution interaction for 1 and acetone (right).

We also examined the solution conductivity (κ) of our Fe complexes as compared to FeCl3 [6, 33]. When each of the respective complexes are dissolved in DCE, we measured the κ for the resulting solution. When FeCl3, 1, 2, and 3 are added to DCE, we observe values of κ equal to 1.5, 84.4, 27.4, and 827.3 μS cm−1, respectively. A value of 827.3 μS cm-1 for 3 suggests that this structure is spontaneously forming a solvent-separated ion pair, which likely occurs when a chloride ion dissociates from the complex. Interestingly, we do not observe this behavior for the other two complexes. This unique behavior of 3 in DCE is consistent with the unique behavior we observed via our UV-Vis and DFT analysis (vide infra), suggesting that the origin of the spectral features we see in the UV-Vis may arise from the change in the solvation sphere of 3.

We next examined the combination of 1 and acetone via conductivity, using concentrations identical to our IR investigation. We previously demonstrated that when acetone is added to FeCl3, a negligible conductivity results between 0–1 equiv, with a conductivity of 96 μS cm−1 at the equivalence point (red, Fig. 10) [6]. However at 2 equiv acetone, κ increases to 733 μS cm−1, which continues up to 1244 μS cm−1 at 5 equiv acetone. Over the course of titration of 1 with acetone, κ doubles from 84.4 μS cm−1 to 163.8 μS cm−1. When compared with the three order of magnitude increase in κ for the addition of acetone to FeCl3, our titration of 1 with acetone is inconsistent with acetone-facilitated displacement of one of the chloride ligands [6].

Fig. 10.

Fig. 10.

Conductivity of FeCl3 (red, 2 mmol in 12 mL DCE) [6] and 1 (blue, 2 mmol in 12 mL DCE) with increasing equivalents of acetone. Solution conductivity of each Fe complex in DCE (inset) [19].

Theoretical Calculations

To find additional support for the shift of the LMCT band in the electronic spectra and proposed solution behavior, TD-DFT calculations were performed [19]. The shift of the LMCT band found in the spectra of complexes 1-3 and 3-acetone agrees well with the calculated absorption spectra of model complexes 1’−3’ and [3’-acetone]+. For example, the experimental band at 743 nm corresponds to the transition calculated at 726 nm originating from HOMO (β) → LUMO+3 (β) (96%) transition with LMCT character in complex 1 [19]. The same HOMO (β) → LUMO+3 (β) LMCT transitions were observed for complexes 2 and 3 with a blue-shift [19].

In acetone, the drastic red-shift observed in the experimental spectrum of 3 (Fig. 7, bottom) was also found in the calculated spectrum [3’-acetone]+. Structural models of [3’-acetone]+ indicate that the formation of an octahedral Fe complex resulting from coordination of the NNN pincer ligand, two chloride atoms, and one acetone ligand is too sterically congested to be considered as a possible structure. The most reasonable model of [3’-acetone]+ was generated when one chloride ion was displaced by acetone, retaining the original trigonal bipyramidal geometry. These results agree well solution conductivity experiments.

Conclusion

Three FeIII complexes bearing a CztBu(PyrR)2 (R = iPr (1), Me (2), or H (3)) ligand were synthesized and characterized in the ground state. We tentatively assign these complexes as high-spin FeIII systems as suggested by solution magnetic susceptibility and EPR spectroscopy. Complexes 13 exhibit quasi-reversible redox activity at both the metal center and organic ligand scaffold. We propose that the degree of metal d orbital and Ncz p orbital overlap is directly influenced by the degree of out-of-plane movement of the Fe atom and plays a major role in the redox activity. This behavior is well illustrated by the difference in metal-center reduction potentials of 1 and 3, where the reduced metal-ligand orbital overlap of 1 may inhibit electron-donation from the ligand, therefore reducing the overall effect that the electronic environment of the ligand has on reduction potential. Solvation of 13 in non-coordinating (CH2Cl2) and coordinating (acetone) solvents produces noticeable changes within the UV-Vis spectra, corresponding to interactions between the coordinating carbonyl of acetone and the Lewis-acidic Fe atom of CztBu(PyrR)2FeCl2. Indeed, in situ IR spectroscopy revealed that 1 (in DCE) possesses a low affinity interaction with added acetone indicated by concomitant observation of acetone and a new vibration at ~1690 cm−1 at all concentrations examined, a significant departure from what is observed for acetone and FeCl3 [6, 33]. As a result of these studies, we propose that complexes 13 can serve as geometrically constrained models for further investigation of the complex interactions between Lewis-acids, such as Fe, and various carbonyl substrates.

Supplementary Material

1

Acknowledgements

We thank Loyola University Chicago for financial support. J.J.D. thanks Merck & Co., Inc. and the NIH/National Institute of General Medical Sciences (GM128126) for financial support. We thank Dr. Joshua Telser for his insightful EPR suggestions and Dr. Yafei Gao for his assistance with Mössbauer spectroscopy. The X-ray diffractometer was funded by the United States National Science Foundation through the Major Research Instrumentation Program under Grant No. DMR 1337296 and CHE 1625543.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1] a).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]; b) Gopalaiah K Chiral Iron Catalysts for Asymmetric Synthesis. Chem. Rev. 2013, 113, 3248–3296; [DOI] [PubMed] [Google Scholar]; c) Bauer I; Knölker H-J Iron Catalysis in Organic Synthesis. Chem. Rev. 2015, 115, 3170–3387; [DOI] [PubMed] [Google Scholar]; d) 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]
  • [2] a).Diaz D; Miranda P; Padron J; Martin V Recent Uses of Iron (III) Chloride in Organic Synthesis. Curr. Org. Chem. 2006, 10, 457–476; [Google Scholar]; b) Sarhan AAO; Bolm C Iron(III) Chloride in Oxidative C–C Coupling Reactions. Chem. Soc. Rev. 2009, 38, 2730. [DOI] [PubMed] [Google Scholar]
  • [3].Jempty TC; Gogins KAZ; Mazur Y; Miller LL Iron Trichloride/Silicon Dioxide Reacts as Oxidant or Lewis Acid with Phenol Ethers. J. Org. Chem. 1981, 46, 4545–4551. [Google Scholar]
  • [4].Wadumethrige SH; Rathore R A Facile Synthesis of Elusive Alkoxy-Substituted Hexa-Peri-Hexabenzocoronene. Org. Lett. 2008, 10, 5139–5142. [DOI] [PubMed] [Google Scholar]
  • [5] a).Khiar N; Fernández I; Alcudia F C2-Symmetric Bis-Sulfoxides as Chiral Ligands in Metal Catalysed Asymmetric Diels-Alder Reactions. Tetrahedron Lett. 1993, 34, 123–126; [Google Scholar]; b) Matsukawa S; Sugama H; Imamoto T Synthesis of P-Chirogenic Diphosphine Oxides and Their Use in Catalytic Asymmetric Diels–Alder Reaction. Tetrahedron Lett. 2000, 41, 6461–6465; [Google Scholar]; c) Usuda H; Kuramochi A; Kanai M; Shibasaki M Challenge toward Structural Complexity Using Asymmetric Catalysis: Target-Oriented Development of Catalytic Enantioselective Diels−Alder Reaction. Org. Lett. 2004, 6, 4387–4390; [DOI] [PubMed] [Google Scholar]; d) Yang L; Zhu Q; Guo S; Qian B; Xia C; Huang H Chiral Brønsted Acid Directed Iron-Catalyzed Enantioselective Friedel-Crafts Alkylation of Indoles with β-Aryl Α′-Hydroxy Enones. Chem. - Eur. J. 2010, 16, 1638–1645. [DOI] [PubMed] [Google Scholar]
  • [6].Hanson CS; Psaltakis MC; Cortes JJ; Devery JJ III Catalyst Behavior in Metal-Catalyzed Carbonyl-Olefin Metathesis. J. Am. Chem. Soc. 2019, 141, 11870–11880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7] a).Ghannam J; Al Assil T; Pankratz TC; Lord RL; Zeller M; Lee W-T A Series of 4- and 5-Coordinate Ni(II) Complexes: Synthesis, Characterization, Spectroscopic, and DFT Studies. Inorg. Chem. 2018, 57, 8307–8316; [DOI] [PubMed] [Google Scholar]; b) Ghannam J; Sun Z; Cundari TR; Zeller M; Lugosan A; Stanek CM; Lee W-T Intramolecular C–H Functionalization Followed by a [2σ + 2π] Addition via an Intermediate Nickel–Nitridyl Complex. Inorg. Chem. 2019, 58, 7131–7135. [DOI] [PubMed] [Google Scholar]
  • [8].Lugosan A; Cundari T; Fleming K; Dickie DA; Zeller M; Ghannam J; Lee W-T Synthesis, Characterization, DFT Calculations, and Reactivity Study of a Nitrido-Bridged Dimeric Vanadium(IV) Complex. Dalton Trans. 2020, 49, 1200–1206. [DOI] [PubMed] [Google Scholar]
  • [9].Lin H-J; Lutz S; O’Kane C; Zeller M; Chen C-H; Al Assil T; Lee W-T Synthesis and Characterization of an Iron Complex Bearing a Hemilabile NNN-Pincer for Catalytic Hydrosilylation of Organic Carbonyl Compounds. Dalton Trans. 2018, 47, 3243–3247. [DOI] [PubMed] [Google Scholar]
  • [10].Inagaki T; Ito A; Ito J; Nishiyama H Asymmetric Iron-Catalyzed Hydrosilane Reduction of Ketones: Effect of Zinc Metal upon the Absolute Configuration. Angew. Chem. Int. Ed. 2010, 49, 9384–9387. [DOI] [PubMed] [Google Scholar]
  • [11].Schubert EM Utilizing the Evans Method with a Superconducting NMR Spectrometer in the Undergraduate Laboratory. J. Chem. Educ. 1992, 69, 62. [Google Scholar]
  • [12].Bruker (2018). Apex3, Saint, Bruker AXS Inc.: Madison (WI), USA. [Google Scholar]
  • [13].Krause L; Herbst-Irmer R; Sheldrick GM; Stalke D Comparison of Silver and Molybdenum Microfocus X-Ray Sources for Single-Crystal Structure Determination. J. Appl. Crystallogr. 2015, 48, 3–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14] a).SHELXTL Suite of Programs, Version 6.14, 2000–2003, Bruker Advanced X-Ray Solutions, Bruker AXS Inc., Madison, Wisconsin: USA; [Google Scholar]; b) Sheldrick GM A Short History of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [DOI] [PubMed] [Google Scholar]
  • [15] a).Sheldrick GM University of Göttingen, Germany, 2018;; b) Sheldrick GM Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Hübschle CB; Sheldrick GM; Dittrich B ShelXle: A Qt Graphical User Interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281–1284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Spartan 16; Wavefunction, Inc.: Irvine, CA. [Google Scholar]
  • [18].Skripnikov L, Chemissian, version 4.43, 2017, http://www.chemissian.com.
  • [19].See Supporting Information for more details.
  • [20].Addison AW; Rao TN; Reedijk J; van Rijn J; Verschoor GC Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds Containing Nitrogen-Sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[l,7-Bis(N-Methylbenzimidazol-2’-Yl)-2,6-Dithiaheptane]Copper(II) Perchlorate t. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar]
  • [21] a).Niwa T; Nakada M A Non-Heme Iron(III) Complex with Porphyrin-like Properties That Catalyzes Asymmetric Epoxidation. J. Am. Chem. Soc. 2012, 134, 13538–13541; [DOI] [PubMed] [Google Scholar]; b) Gibson VC; Spitzmesser SK; White AJP; Williams DJ Synthesis and Reactivity of 1,8-Bis(Imino)Carbazolide Complexes of Iron, Cobalt and Manganese. Dalton Trans. 2003, 32, 2718. [Google Scholar]
  • [22].Spaniel T; Gorls H; Scholz J, (1,4-Diaza-1.3-diene)titanium and -niobium Halides: Unusual Structures with Intramolecular C-H´´´Halogen Hydrogen Bonds. Angew. Chem. Int. Ed. 1998, 37, 1862–1865. [Google Scholar]
  • [23].Shannon RD Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar]
  • [24] a).Gupta GP; Lang G; Lee YJ; Scheidt WR; Shelly K; Reed CA Spin Coupling in Admixed Intermediate-Spin Iron(III) Porphyrin Dimers: Crystal Structure, Mössbauer, and Susceptibility Study of Fe(TPP)(B11CH12)·C7H8. Inorg. Chem. 1987, 26, 3022–3030. [Google Scholar]; b) Scheidt WR; Osvath SR; Lee YJ; Reed CA; Shaevitz B; Gupta GP Control of Spin State in Iron(III) Porphyrinates. The Admixed Intermediate-Spin Case. Crystal Structure, Moessbauer, and Susceptibility Study of Bis(3,5-Dichloropyridine)(Octaethylporphyrinato)Iron(1+) Perchlorate. Inorg. Chem. 1989, 28, 1591–1595. [Google Scholar]; c) Ohya T; Takeda J; Sato M Spin States of Iron(III) in Highly Saddled Dodecaphenylporphyrin Complexes. Hyperfine Interactions. 2004, 156, 265–272. [Google Scholar]
  • [25] a).Ikezaki A; Nakamura M Models for Cytochromes c’: Spin States of Mono(Imidazole)-Ligated (Meso -Tetramesitylporphyrinato)Iron(III) Complexes as Studied by UV−Vis, 13C NMR, 1H NMR, and EPR Spectroscopy. Inorg. Chem. 2002, 41, 6225–6236; [DOI] [PubMed] [Google Scholar]; b) Weiss R; Gold A; Terner J Cytochromes c’: Biological Models for the S = 3/2, 5/2 Spin-State Admixture? Chem. Rev. 2006, 106, 2550–2579. [DOI] [PubMed] [Google Scholar]; c) Gibson VC; Spitzmesser SK; White AJP; Williams DJ Synthesis and reactivity of 1.8-(bis)iminocarbazolide complexes of iron, cobalt and manganese. Dalton. Trans. 2003, 2718–2727. [Google Scholar]; d) Maltempo MM; Moss TH The Spin 3/2 State and Quantum Spin Mixtures in Haem Proteins. Q. Rev. Biophys. 1976, 9, 181–215. [DOI] [PubMed] [Google Scholar]; e) Mossin S; Tran BL; Adhikari D; Pink M; Heinemann FW; Sutter J; Szilagyi RK; Meyer K; Mindiola DJ A Mononuclear Fe(III) Single Molecule Magnet with a 3/2 5/2 Spin Crossover. J. Am. Chem. Soc. 2012, 134, 13651–13661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26] a).Boersma AD; Goff HM Multinuclear Magnetic Resonance Spectroscopy of Spin-Admixed S = 5/2, 3/2 Iron(III) Porphyrins. Inorg. Chem. 1982, 21, 581–586; [Google Scholar]; b) Boersma AD; Goff HM Electrochemical Oxidation of Spin-Admixed S = 5/2, 3/2 Iron(III) Porphyrins: In Situ Characterization by Deuterium NMR Spectroscopy. Inorg. Chem. 1984, 23, 1671–1676; [Google Scholar]; c) Maltempo MM Magnetic State of an Unusual Bacterial Heme Protein. J. Chem. Phys. 1974, 61, 2540–2547; [Google Scholar]; d) Reed CA; Mashiko T; Bentley SP; Kastner ME; Scheidt WR; Spartalian K; Lang G The Missing Heme Spin State and a Model for Cytochrome c’. The Mixed S = 3/2, 5/2 Intermediate Spin Ferric Porphyrin: Perchlorato(Meso-Tetraphenylporphinato)Iron(III). J. Am. Chem. Soc. 1979, 101, 2948–2958. [Google Scholar]; e) Alonso PJ; Arauzo AB; Forniés J; García-Monforte MA; Martín A; Martínez JI; Menjón B; Rillo C; Sáiz-Garitaonandia JJ A Square-Planar Organoiron(III) Compound with a Spin-Admixed State. Angew. Chem. Int. Ed. 2006, 45, 6707–6711. [DOI] [PubMed] [Google Scholar]
  • [27] a).Morozova V; Mayer P; Berionni G Scope and Mechanisms of Frustrated Lewis Pair Catalyzed Hydrogenation Reactions of Electron-Deficient C=C Double Bonds. Angew. Chem. Int. Ed. 2015, 54, 14508–14512; [DOI] [PubMed] [Google Scholar]; b) Gutmann V Solvent Effects on the Reactivities of Organometallic Compounds. Coord. Chem. Rev. 1976, 18, 225–255; [Google Scholar]; c) Beckett MA; Strickland GC; Holland JR; Sukumar Varma K A Convenient n.m.r. Method for the Measurement of Lewis Acidity at Boron Centres: Correlation of Reaction Rates of Lewis Acid Initiated Epoxide Polymerizations with Lewis Acidity. Polymer 1996, 37, 4629–4631; [Google Scholar]; d) Beckett MA; Brassington DS; Coles SJ; Hursthouse MB Lewis Acidity of Trispentafluorophenyl/ Borane: Crystal and Molecular Structure of BC6F5/ 3POPEt3. 2000, 4; [Google Scholar]; e) Britovsek GJP; Ugolotti J; White AJP From B(C6F5)3 to B(OC6F5)3: Synthesis of (C6F5)2 BOC6F5 and C6F5B(OC6F5)2 and Their Relative Lewis Acidity. Organometallics 2005, 24, 1685–1691; [Google Scholar]; f) Childs RF; Mulholland DL; Nixon A The Lewis Acid Complexes of α,β-Unsaturated Carbonyl and Nitrile Compounds. A Nuclear Magnetic Resonance Study. Can. J. Chem. 1982, 60, 801–808; [Google Scholar]; g) Hilt G; Pünner F; Möbus J; Naseri V; Bohn MA A Lewis Acidity Scale in Relation to Rate Constants of Lewis Acid Catalyzed Organic Reactions. Eur. J. Org. Chem. 2011, 5962–5966; [Google Scholar]; h) Hilt G; Nödling A The Correlation of Lewis Acidities of Silyl Triflates with Reaction Rates of Catalyzed Diels-Alder Reactions. Eur. J. Org. Chem. 2011, 7071–7075; [Google Scholar]; i) Santelli M; Pons J-M Lewis Acids and Selectivity in Organic Synthesis; New directions in organic and biological chemistry; CRC Press: Boca Raton, 1996. [Google Scholar]
  • [28].Hirano T; Sekiguchi T; Hashizume D; Ikeda H; Maki S; Niwa H Colorimetric and Fluorometric Sensing of the Lewis Acidity of a Metal Ion by Metal-Ion Complexation of Imidazo[1,2-a]Pyrazin-3(7H)-Ones. Tetrahedron 2010, 66, 3842–3848. [Google Scholar]
  • [29] a).Kobayashi S; Busujima T; Nagayama S A Novel Classification of Lewis Acids on the Basis of Activity and Selectivity. Chem. Eur. J. 2000, 6, 3491–3494; [DOI] [PubMed] [Google Scholar]; b) Ludwig JR; Zimmerman PM; Gianino JB; Schindler CS Iron(III)-Catalysed Carbonyl–Olefin Metathesis. Nature 2016, 533, 374–379. [DOI] [PubMed] [Google Scholar]
  • [30] a).Tsuruta H; Yamaguchi K; Imamoto T Tandem Mass Spectrometric Analysis of Rare Earth(III) Complexes: Evaluation of the Relative Strength of Their Lewis Acidity. Tetrahedron 2003, 59, 10419–10437; [Google Scholar]; b) Gal J-F; Iacobucci C; Monfardini I; Massi L; Duñach E; Olivero S Metal Triflates and Triflimides as Lewis “Superacids”: Preparation, Synthetic Application and Affinity Tests by Mass Spectrometry: Catalytic and Mass Spectrometry Studies of Metal Triflates and Triflimides. J. Phys. Org. Chem. 2013, 26, 87–97. [Google Scholar]
  • [31] a).Satchell DPN; Satchell RS Quantitative Aspects of the Lewis Acidity of Covalent Metal Halides and Their Organo Derivatives. Chem. Rev. 1969, 69, 251–278; [Google Scholar]; b) Mohammad A; Satchell DPN; Satchell RS Quantitative Aspects of Lewis Acidity. Part VIII. The Validity of Infrared Carbonyl Shifts as Measures of Lewis Acid Strength. The Interaction of Lewis Acids and Phenalen-1-One(Perinaphthenone). J. Chem. Soc. B Phys. Org. 1967, 723–725; [Google Scholar]; c) Satchell DPN; Satchell RS Quantitative Aspects of Lewis Acidity. Q. Rev. Chem. Soc. 1971, 25, 171–199; [Google Scholar]; d) Driessen WL; Groeneveld WL Complexes with Ligands Containing the Carbonyl Group. Part I: Complexes with Acetone of Some Divalent Metals Containing Tetrachloro-Ferrate(III) and -Indate(III) Anions. Recl. Trav. Chim. Pays-Bas 1969, 88, 977–988; [Google Scholar]; e) Driessen WL; Groeneveld WL Complexes with Ligands Containing the Carbonyl Group. Part III: Metal (II) Acetaldehyde, Propionaldehyde and Benzaldehyde Solvates. Recl. Trav. Chim. Pays-Bas 1971, 90, 87–96; [Google Scholar]; f) Driessen WL; Groeneveld WL Complexes with Ligands Containing the Carbonyl Group. Part IV Metal(II) Butanone, Acetophenone, and Chloroacetone Solvates. Recl. Trav. Chim. Pays-Bas 1971, 90, 258–264; [Google Scholar]; g) Cook D Infrared Spectra of Xanthone: Lewis Acid Complexes. Can. J. Chem. 1963, 41, 522–526. [Google Scholar]
  • [32].Hanson CS; Devery JJ III Characterizing Lewis Pairs Using Titration Coupled with In Situ Infrared Spectroscopy. J. Vis. Exp. 2020, 156, e60745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Hanson CS; Psaltakis MC; Cortes JJ; Siddiqi SS; Devery JJ III Investigation of Lewis Acid-Carbonyl Solution Interactions via Infrared-Monitored Titration. J. Org. Chem. 2020, 85, 820–832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Malakar T; Hanson CS; Devery JJ III; Zimmerman PM Combined Theoretical and Experimental Investigation of Lewis Acid-Carbonyl Interactions for Metathesis. ACS Catal. 2021, 11, 4381. [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.

Supplementary Materials

1

RESOURCES