Kritika
Keshari
a,
Moumita
Bera
a,
Lucía
Velasco
b,
Sandip
Munshi
c,
Geetika
Gupta
a,
Dooshaye
Moonshiram
*b and
Sayantan
Paria
*a
aDepartment of Chemistry, Indian Institute of Technology Delhi, Hauz Khas, New Delhi 110016, India. E-mail: sparia@chemistry.iitd.ac.in
bInstituto Madrileño de Estudios Avanzados en Nanociencia (IMDEA Nanociencia, Calle Faraday, 9, 28049 Madrid, Spain. E-mail: dooshaye.moonshiram@imdea.org
cSchool of Chemical Sciences, Indian Association for the Cultivation of Science, 2A & 2B Raja S. C. Mullick Road, Jadavpur, Kolkata 700032, India
First published on 27th January 2021
A terminal FeIIIOH complex, [FeIII(L)(OH)]2− (1), has been synthesized and structurally characterized (H4L = 1,2-bis(2-hydroxy-2-methylpropanamido)benzene). The oxidation reaction of 1 with one equiv. of tris(4-bromophenyl)ammoniumyl hexachloroantimonate (TBAH) or ceric ammonium nitrate (CAN) in acetonitrile at −45 °C results in the formation of a FeIIIOH ligand radical complex, [FeIII(L˙)(OH)]− (2), which is hereby characterized by UV-visible, 1H nuclear magnetic resonance, electron paramagnetic resonance, and X-ray absorption spectroscopy techniques. The reaction of 2 with a triphenylcarbon radical further gives triphenylmethanol and mimics the so-called oxygen rebound step of Cpd II of cytochrome P450. Furthermore, the reaction of 2 was explored with different 4-substituted-2,6-di-tert-butylphenols. Based on kinetic analysis, a hydrogen atom transfer (HAT) mechanism has been established. A pKa value of 19.3 and a BDFE value of 78.2 kcal/mol have been estimated for complex 2.
Scheme 1 General mechanism for the hydroxide rebound of (a) α-keto glutarate-dependent oxygenase and (b) cytochrome P450. |
High-valent (Por)FeIV–OH species have been identified in compound II of cytochrome P450 and HRP-II, CcP.8 A computational investigation by Shaik revealed that (Por)FeIV–OH can also exist as a (Por˙+)FeIII–OH during the desaturation pathway of P450cam.32 Therefore, characterization and reactivity studies of high-valent Fe–OH complexes constitute an important topic of research to bioinorganic chemists. The artificial analogue of such species is rare and synthetically very challenging, as the mononuclear Fen+O complex has an inherent tendency to form thermodynamically more stable Fe–O–Fe-type complexes. Nonetheless, there remain only a few elusive analogues of FeIV–OH complexes reported in the literature.33–36 Borovik has described a protonated FeIVO complex of a tris-urea tripodal ligand, where the protonation likely occurred at the ligand backbone.37 Very recently, Nam et al., spectroscopically characterized and studied the reactivity of a ligand-protonated FeVO species of a Tetra-Amido Macrocyclic Ligand (TAML).38 Goldberg has reported a structurally characterized FeIV–OH complex of a corrole ligand, which was shown to participate in OH re-bound reactions similar to the enzymatic systems, the only reported functional model of compound II.33,36 Very recently, Hendrich described an FeIV–OHn (n = 1 or 2) of a TAML through various spectroscopic techniques.35
Herein, we report the synthesis, structural characterization, and reactivity study of a terminal FeIII–OH complex, [FeIII(L)(OH)]2− (1). One-electron oxidation of complex 1 with one equiv. of tris(4-bromophenyl)ammoniumyl hexachloroantimonate (TBAH) or ceric ammonium nitrate (CAN) results in the formation of an FeIII–OH ligand radical complex, [FeIII(L˙)(OH]− (2, Scheme 2) which is hereby characterized by UV-visible (UV-vis), 1H-Nuclear Magnetic Resonance (NMR), and X-ray absorption spectroscopy (XAS) techniques. The latter complex spontaneously reacts with Gomberg's dimer forming Ph3COH and mimics the C–OH bond formation step of Cpd II (Scheme 2).
Fig. 1 Crystal structure of complex 1 with 50% ellipsoid probability. Counter cations and all of the hydrogen atoms except those attached to O1 have been omitted for clarity. |
In contrast, the X-band EPR spectrum of complex 1 in acetonitrile at 100 K shows a rhombic spectrum having g values at 2.78, 2.21, and 1.36 (Fig. 2, inset), typical for low-spin iron(III) complexes having (dxy)2(dxz,dyz)3 electron configurations.44 We speculate that upon decreasing the temperature, the spin-state of 1 changes from S = 3/2 to S = 1/2.
Fig. 2 UV-vis spectra of 1 and 2 at −45 °C in acetonitrile (0.32 mM). Inset: EPR spectrum of complex 1 (0.5 mM) at 100 K in frozen acetonitrile. |
The redox properties of 1 were evaluated by cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in acetonitrile at 25 °C, which revealed a quasi-reversible redox process at E1/2 = 0.487 V (versus NHE), assigned to (L˙)FeIII/(L)FeIII (Fig. S7†).
Given the electrochemical properties, we envisioned that the chemical oxidation of 1 could form a high-valent Fe–OH complex. We utilized one-electron oxidizing agents like TBAH (EOX = 1.321 V versus NHE) or CAN (EOX = 1.61 V versus NHE) to oxidize 1. The reaction of 1 with TBAH was monitored by UV-vis spectroscopy in acetonitrile at −45 °C (Fig. 2). The addition of one equiv. of TBAH to 1 resulted in the formation of a new species 2 having absorbance maxima at 360 nm (5100 M−1 cm−1), 470 nm (3700 M−1 cm−1) and ca. 680 nm (1000 M−1 cm−1). A similar spectral feature was obtained upon the addition of one equiv. of CAN to the solution of complex 1 (Fig. S8†). We propose that such peaks originated from the π–π* transitions of the ligand radical.45 Formation of 2 was also monitored by 1H NMR spectroscopy. The NMR spectrum of 2, formed upon addition of TBAH to 1 in CD3CN, was measured at −30 °C and showed shifted peaks with the appearance of new spectral features, indicating the formation of a new intermediate complex (Fig. S9†).
In order to determine the spin-state of complex 2, solution magnetic moment was measured in CD3CN by Evans' method and revealed μeff = 3.05 μB, which corresponds to a spin-state of S = 1.46 The X-band EPR spectrum of 2 was further monitored in frozen acetonitrile at 100 K, which revealed a close to silent EPR spectrum, thus ascertaining its S = 1 spin state and the large zero-field splitting tensor (Fig. S10†).47
In order to determine the electronic and local structural conformations of 1 and 2, both complexes were studied in solution state by X-ray absorption near edge structure (XANES) and Extended X-ray absorption fine structure (EXAFS) spectroscopy (Fig. 3). Both 1 and 2 display a rising edge from 7120 to ∼7135 eV, corresponding to the 1s to 4p electronic transition. The distinct ligand versus metal-centered oxidation in complex 2 is confirmed by the missing shift of the rising edge towards the higher energy of the XANES spectra at 0.6 normalized absorption and at 7125.15 eV (Fig. 3A). By contrast, upon oxidation with 1 equiv. TBAH, changes at higher photon energies at ∼7133–7135 eV (Fig. 3A) and in the pre-edge region (Fig. 3A inset) are observed. The presence of pre-edge features corresponds to 1s to 3d quadrupole transitions and dipole excitations of the core electrons into the valence 3d states hybridized with ligand p orbitals.48,49 The changes observed in the pre-edge and rising edge features hereby reflect the variations and local symmetrical and geometrical differences expected between both complexes.50 The pre-edge peaks of complexes 1 and 2 were further fitted and yielded maximum peak energies of 7113.97 eV and 7113.71 eV, respectively. A pre-edge peak area of 19.3 units was, on one hand, obtained for complex 1 (Fig. 3B, Table S5†), which is close and consistent with that of previously studied five-coordinated ferric centers, demonstrating pre-edge areas of ∼17 units.51,52 The decreased pre-edge area of 16.1 units for complex 2 is however in sharp contrast to that of Fe(IV)oxo51,53 high-valent complexes, exhibiting typical areas of 41–52 units,51,53 thus confirming the lack of metal-centered oxidation.
The EXAFS spectra of 1 and 2 are shown in Fig. 3C. Two prominent peaks are observed in complex 1's spectrum corresponding to the distinctive Fe–N and Fe–O bond distances. The EXAFS fits for the extraction of actual bond distances of the two complexes are shown in Fig. 3C inset, Table S3 and Fig. S11.† Analysis of EXAFS spectra in complex 1 in the solution state clearly resolved 3 Fe–O distances at 1.88 ± 0.01 Å and 2 Fe–N distances at 2.04 ± 0.01 Å, in close agreement with the obtained XRD data (Tables S3 and S4, Fig. S11†). Upon oxidation with 1 equiv. of TBAH, complex 2 demonstrates a prominent sharp peak denoted in red as III (Fig. 3C) and an elongated shoulder peak feature shown as Peak IV in the first coordination sphere (Fig. 3C). Analysis of the EXAFS region of complex 2 resolves 3 Fe–O distances at 1.89 ± 0.02 Å and comparatively lengthened Fe–N distances at 2.09 ± 0.02 Å (Table S3, Fig. S11†). The obtained local structural conformation not only ascertains the ligand-centered oxidation process in complex 2 but also additionally indicates that the Fe metal center likely lies slightly above the plane of the connected ligating atoms upon oxidation in 2 in comparison with 1. Here, it is important to note that the Fe–O distance observed in 2 is close to the Fe–O distance (1.857 Å) reported in [FeIV(ttppc)(OH)].33
The reactivity of complex 2 with Gomberg's dimer was examined by UV-vis spectroscopy in 3:2 acetonitrile/THF (v/v) at −40 °C. The addition of 10-fold excess of Gomberg's dimer to 2 resulted in immediate decay of peaks at 470 and 680 nm (Fig. 4). The reaction products were further analyzed by gas chromatography-mass spectrometry (GC-mass) and 1H-NMR spectroscopy techniques, which confirmed the formation of Ph3COH with 80% yield. The source of oxygen atom in Ph3COH was evaluated by 18O labeling experiments. In the GC-mass spectrum, a peak was observed at m/z = 260.1 (Fig. S12†), which corresponds to the composition of [Ph3COH]+. This peak displays a two-mass unit shift to m/z = 262.1 (Fig. S13†) when the sample was prepared in the presence of [FeIII(L˙)(18OH)]− and Gomberg's dimer and confirms that the source of OH in the product is the Fe–OH moiety. The experiment demonstrated 75% incorporation of 18O into triphenylmethanol. The study thus shows that complex 2 mimics the reactivity of Cpd II and supports an alternative structure and reactivity pathway for Cpd II, as documented by Shaik.32
Fig. 4 Change in the UV-vis spectrum of 2 (0.1 mM) upon addition of Gomberg's dimer in 3:2 acetonitrile/THF (v/v) at −40 °C (a) and time trace of the decay of 2 at 470 nm (b). |
The reaction of 2 was further explored with different 4-X-2,6-di-tert-butylphenols (X = –OCH3, –OCH3 (D), –CH3, –CH2CH3, –C(CH3)3, and –H) in acetonitrile at −25 °C. Addition of one equiv. of 4-methoxy-2,6-di-tert-butylphenol (4-OMe-DTBP) to 2 resulted in immediate decomposition of peaks at 470 and 680 nm in the UV-vis spectrum with concomitant appearance of a new peak at 405 nm, corresponding to the formation of a 4-methoxy-2,6-di-tert-butylphenoxyl radical (Fig. 5a).
The second-order rate constant of the reaction was obtained from the slope of a plot of (A0 − A)/C0(A − A∞) vs. time (t) and yielded kH2 = 71.15 M−1 s−1 (Fig. S14†). The formation of 88% phenoxyl radical was additionally confirmed by EPR spectroscopy (Fig. S15†).
The reaction of 2 was subsequently explored in the presence of 4-OMe-DTBP-D, which yielded a kD2 value of 50.8 M−1 cm−1 in acetonitrile at −25 °C, thus displaying a kinetic isotope effect (KIE) of 1.4. The reactivity of 2 was further explored with 2,6-di-tert-butyl-4-methylphenol (Fig. S17–19†), 2,6-di-tert-butyl-4-ethylphenol (Fig. S20–22†), 2,4,6-tri-tert-butylphenol (Fig. S23–26†) and 2,6-di-tert-butylphenol (Fig. S27–29†) under pseudo-first-order reaction conditions resulting in k2 values of 0.417, 0.416, 0.375, and 0.026 M−1 cm−1, respectively. A plot of logk2versus O–H bond dissociation energy of phenols follows a linear correlation and resulted in a slope of −0.77 (Table S9 and Fig. S30†). A Hammett correlation plot was obtained by plotting logk2versus σp+ of 4-X-2,6-di-tert-butylphenols (Fig. 5b and Table S10†) and yielded a value of σ+ = −4.4, which hereby indicates the strong electrophilic nature of 2. It is important to remark that a σ+ value of −2.6(8) was reported for [FeIV(ttpc)(OH)]33 for similar reactions (ttppc = 5,10,15-tris(2,4,6-triphenyl)-phenyl corrole), which indicates a greater charge separation in the transition state of the reaction of 2 with O–H bonds. The σ+ value of 2 is higher than those of [FeIV(O)(TMC)(X)]n+ complexes for the reaction with phenolic O–H bonds (X = N3, σ+ = −1.5; X = CF3COO−, σ+ = −2.3; X = CH3CN, σ+ = −3.2; TMC = 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane).54 Furthermore, the Marcus correlation of the reaction of 2 with different 4-X-2,6-di-tert-butylphenols was explored. A plot of (RT/F) lnk2versus Eox of the phenols follows a linear correlation and resulted in a slope of −0.31 (Fig. 5c and Table S11†). The obtained value is close to the values reported for the hydrogen atom transfer reactions of [FeIV(ttpc)(OH)] (slope = −0.19(6)) and [MnIV(ttpc)(OH)] (slope = −0.12) with p-substituted 2,6-di-tert-butyl phenols.33
Eyring analysis of the reaction of 2 with 4-OMe-2,6-DTBP was also performed over a temperature range of 248 to 228 K (Table S7† and Fig. 5d). Activation enthalpy, ΔH# = 7.23 ± 0.28 kcal mol−1, and activation entropy, ΔS# = −20.54 ± 1.6 cal mol−1 K−1, were estimated from the slope and intercept of a plot of ln(k2/T) vs. 1/T, respectively. Such a large negative activation entropy suggests a well-ordered transition state in the rate-determining step. Although no similar values of high-valent Fe–OH complexes are available for comparison purposes, our results are close to the activation parameters reported for the hydrogen atom transfer type reaction of with 4-OMe-2,6-DTBP and other Cu(III), Mn(V) and Ru(VI) complexes (Table S8†).55 The activation parameters of 2 are very close to the values reported for the HAT type reaction of FeIVO complexes with 1,4-cyclohexadiene or xanthene.56,57 Based on the linear correlations obtained between the rates of the reaction of 2 and the BDE of phenols, the value of KIE, activation parameters of the reaction of 2 with 4-OMe-DTBP, and Marcus plot analysis, we propose that the cleavage of the O–H bond by 2 occurs in the rate-determining step.
We additionally attempted to estimate the pKa value of 2. However, no spectral change was observed upon addition of equiv. amount of 2-aminopyridine (pKa = 14.3) or triethylamine (pKa = 18.82) to 2. Reversible deprotonation of 2 with pyrrolidine (pKa = 19.56)58 allowed us to estimate a pKa value of ∼19.3 for 2 (Fig. S31 and Scheme S1†). This value is higher than that of the FeIIIOH (pKa = 16.5(0.4)) complex stabilized by secondary coordination sphere hydrogen bonding interactions reported by Fout et al.30 Based on the obtained pKa and E0 value of the (L˙)FeIII/(L)FeIII redox couple, we determined the O–H bond dissociation free energy (BDFE) of 78.2 kcal mol−1 of 2 (eqn (1), CH = 54.9 kcal mol−1 in acetonitrile) from the thermodynamic square scheme described in Scheme 3.59 The BDFE value is further verified experimentally as 2 spontaneously reacts with different 4-substituted-2,6-di-tert-butylphenols having similar BDFE values.
BDFE(O–H) = 1.37pKa + 23.06E0 + CH | (1) |
The experimentally obtained BDFE value of 2 is higher than those of FeIII–OH (70 kcal mol−1),30 FeIIIO (66 kcal mol−1),60 and MnIIIO (72 kcal mol−1)61 complexes, but lower than those of FeIVO complexes (87 kcal mol−1).62
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2018882. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0sc07054h |
This journal is © The Royal Society of Chemistry 2021 |