DOI:
10.1039/D5SC04153H
(Edge Article)
Chem. Sci., 2025,
16, 18364-18371
Rapid nitrite reduction enabled by secondary sphere hydrogen bonds within non-heme iron complexes
Received
6th June 2025
, Accepted 28th August 2025
First published on 29th August 2025
Abstract
A non-heme iron(II) complex bearing a ligand with secondary sphere hydrogen bond (H-bond) donors, (tris(6-phenylaminopyridylmethyl)amine, TPANHPh, rapidly reduces nitrite (NO2−) to nitric oxide (NO) in the absence of exogenous additives, affording a Fe(III)2(μ-O)2 diamond core. An electronically analogous complex containing a ligand without H-bonds (tris(6-methylpyridylmethylamine), TPAMe, also reduces NO2− to NO and forms an Fe(III)2(μ-O)2 core, but is four orders of magnitude slower, highlighting the impact of H-bonds to promote NO2− reduction. We compare the structural and spectroscopic differences of the two Fe(III)2(μ-O)2 complexes and show that H-bonding interactions weaken the Fe–O bonds, perturb the electronic structure of the Fe2O2 cores, and thereby engender distinct reductive stability profiles.
Introduction
Nitrite (NO2−) is a source for physiological production of nitric oxide (NO), which is an important bioregulatory signalling molecule.1 NO has multiple functions as a vasodilator, neurotransmitter, and has important roles in mammalian immune responses.1,2 In these biological contexts, an array of Fe- and Cu-based metalloenzymes (e.g. cytochrome cd1- or cytochrome c nitrite reductases, cd1-NiR/cc-NiR, and Cu-NiR) catalyze the reduction of NO2− to NO or to NH3 (ccNiR).3 Mutagenesis and computational studies implicate a critical role of specific hydrogen-bonding (H-bonding) amino acids close to the active sites of these enzymes (i.e., in the secondary sphere) for enzymatic function.4 Removal of these H-bonding residues in cd1- or cc-NiR inhibits catalytic activity, affording up to a 99% decrease in catalytic rates, highlighting their roles to both enable and accelerate enzymatic NO2− reduction.5 Although the dependence on H-bonds for overall function is established, the molecular-level details of these interactions are challenging to clarify because removal induces larger structural changes, obfuscating their role(s) on individual reaction steps.4a
Inorganic model complexes can provide insight into the effects of secondary sphere H-bonds,6 and by extension, the mechanisms of enzymatic NO2− reduction. Complexes that do not contain H-bond donors often require the addition of an exogenous Brønsted acid and/or (electro)chemical reducing equivalents to promote NO release.7 In contrast, systems containing secondary sphere H-bond donors can induce spontaneous NO2− reduction without exogenous acids.8 Seminal work by the Fout group reported facile NO2− reduction mediated by Fe in a tripodal azafulvine-imine ligand scaffold, producing NO and a monomeric Fe(III)–O(H), a result that was attributed to the pendent-imino groups acting as H-bond donors (Fig. 1a).8b–d,9 Related reports from the Gilbertson group highlighted Fe-pyridinediimine complexes with tethered amines that reduced NO2− to afford an {Fe(NO)2}9,10 and the Hung group reported an N-confused porphyrin that similarly reduced NO2− to form an {Fe(NO)}6/7.11
 |
| Fig. 1 (a) Previously reported Fe-based systems with secondary sphere H+/H-bond donors that spontaneously reduce NO2−,8b–d,9,10 (b) this work. | |
While these prior studies demonstrated synthetic examples of H-bond promoted NO2− reduction, direct comparisons between ligands that either contain or omit H-bonds are notably absent. Modification of the secondary sphere of a complex can impact its primary coordination sphere via perturbations to ligand field environments, redox potentials, or spin states.5a,12 These competing effects present challenges when assessing the role(s) of H-bonds during NO2− reduction. Our group has attempted to decouple these parameters by comparing reaction outcomes with electronically similar ligands that differ in their secondary sphere,13 and we previously showed that a scaffold containing aniline groups as H-bond donors, (tris(6-phenylaminopyridy-lmethyl)amine, TPANHPh), promoted capture/activation of O2 (with Fe, Cu, and Zn)13b,f,g in addition to tandem ClO4− reduction/C–H oxygenation (with Fe).13a,c We also reported a Cu(I) complex with a related ligand containing appended OH groups, tris(6-hydroxypyridylmethyl)amine TPAOHCu(I), which reduced NO2− to NO, an example of ligand-promoted H+/e− transfer.14 In this report, we expand these efforts to Fe-mediated NO2− reduction.
Results and discussion
Introduction of excess (ca. 10 equiv.) [Bu4N][NO2] to an MeCN solution of TPANHPhFe(II) bis(bis-trifluoromethylsulfonylazanide), (TPANHPhFe(NTf2)2, I-NTf2, Fig. 2a), under ambient conditions afforded a two-step reaction sequence: a rapid color change from colorless to yellow (seconds), followed by a gradual transition to reddish-brown (hours). Removal of MeCN in vacuo followed by precipitation from CH2Cl2/Et2O provided a single product, as assessed by 1H-NMR spectroscopy. A solution-phase IR spectrum of the product contained no Fe–NO stretches between 1600–1850 cm−1, but did exhibit broadened and bathochromically shifted N–H stretches (νNH = 3232 and 3188 cm−1) relative to I-NTf2 (νNH = 3361 and 3278 cm−1), consistent with strengthened H-bond interactions.15 To examine whether NO was released during this reaction, we allowed I-NTf2 to react with excess [Bu4N][NO2] in the presence of CoTPP (TPP = tetraphenyl porphyrin), which provided quantitative yield of CoTPP(NO) after 5 h.16 These data are consistent with spontaneous NO2− reduction to NO by I-NTf2, potentially with concomitant formation of an Fe–O unit, as we did not observe Fe–NO bond formation.
 |
| Fig. 2 (a) Reaction of TPANHPhFe(X)2 (X = OTf− or NTf2−; I-OTf or I-NTf2) with NO2− to form NO(g) and TPANHPh2Fe(III)2(μ-O)22+ (II). (b) Reaction of TPAMeFe(NTf2)2 (III-NTf1) with NO2− to form TPAMe2Fe(III)2(μ-O)22+(IV) and NO(g), (c) molecular structures of II and IV (NO2−-derived); 50% probability ellipsoids, short H-bond contacts highlighted in blue, phenyl groups are wireframed, H-atoms not involved in H-bonds, and outer sphere anions omitted for clarity (d). Bond metrics of II, IVa and IVb (a: this work, b: ref. 19g, *denotes an average of 3 bond distances). | |
To determine the molecular composition of the product, we performed a single-crystal X-ray diffraction (XRD) experiment on crystals grown from CH2Cl2/Et2O. Rather than a monomeric Fe–O, the refined structure revealed a [Fe(III)2(μ-O)2]2+ diamond core (II, Fig. 2c) enveloped by moderate strength H-bonding interactions (average O1–N3 distance of 2.9 ± 0.2 Å).17 This Fe2O2 structural motif is reminiscent of higher valent intermediates within non-heme di-iron enzymes (e.g. soluble methane monooxygenase (sMMO-Q) in the Fe(IV)2 state and ribonucleotide reductase (RNR-X) in the Fe(III)Fe(IV) state)18 but is comparatively rare in synthetic systems.19 Que and coworkers reported the first isolated Fe(III)2(μ-O)2 core, supported by two tris(6-methylpyridylmethyl)amine) ligands, (TPAMe2Fe(III)2(μ-O)22+IV), which was prepared via more standard oxygenation reagents (tBuOOH and NEt3).19g Another related example, reported by Masuda, is a TPANH2 analogue that formed an H-bonded Fe(III)2(μ-O)2 species from O2.19c We found that IV could also be prepared directly from [Bu4N][NO2], albeit over a longer time frame (1 week, Fig. 2b). Importantly, TPAMe provides an analogous primary sphere environment as TPANHPh, but does not contain pendent H-bond donors.13b Thus, we propose that differences in the chemical properties of II and IV can be attributed to the effects of secondary sphere H-bonds.
The structural metrics in II are similar to those in IV, with a few notable exceptions (Fig. 2d).20 The Fe2O2 diamond core in II displays two distinct Fe–O bonds (Fe1–O1: 1.886(3) Å, and Fe1–O2: 1.925(3) Å), an Fe1–Fe2 separation of 2.729(1) Å, and an Fe1–O1–Fe2 angle of 91.5(2)°. In comparison, IV has a 0.038(4) Å shorter Fe1–O1 bond (1.848(2) Å) and a 0.034(2) Å shorter Fe–Fe distance (2.695(1) Å), but the other Fe2O2 core metrical parameters are similar (Fe1–O2: 1.925(2) Å, and ∠Fe1–O1–Fe2: 91.1(1)°). The elongated Fe1–O1 bond likely exhibits a weaker trans-influence, which is manifest by the contracted Fe1–N1 distance of II (2.178(3) Å) compared to IV (2.215(2) Å). We attribute these structural differences to the trifurcated H-bonds surrounding each μ-O2− ligand in II, which are proposed to reduce the basicity of oxo ligands, thereby weakening the Fe–O bonds.19c,21,22
To clarify the electronic differences that are imparted by H-bonding interactions, we examined the electronic absorption spectra of II and IV. Their respective spectra are distinct: II exhibits a prominent shoulder band at 433 nm (ε = 2.6 mM−1 cm−1) while IV displays a similarly intense shoulder band at 375 nm (ε = 2.0 mM−1 cm−1), both assigned to LMCT transitions.19b These spectral shifts suggest that H-bonding interactions alter the electronic structure of the Fe2(μ-O)2 cores, which we interrogated further by Mössbauer spectroscopy. The zero-field Mössbauer spectrum (Fig. 3, top) of II as a solid (at 298 K) displays a symmetric doublet with an isomer shift (δ) of 0.33 mm s−1 and quadrupole splitting (ΔEq) of 1.422 mm s−1, consistent with two identical high spin Fe(III) centers.19a,23 These data are distinct from the reported Mössbauer parameters of IV: (symmetric, δ = 0.50 mm s−1; ΔEq = 1.93 mm s−1). We attribute the differences in δ (lower δ for II) to more oxidized/Lewis acidic Fe(III) centers,22 and the differences in ΔEq values to H-bond induced charge redistribution and/or orbital rehybridization of the μ-O ligands.24 These effects may also be partially responsible for the decrease in antiferromagnetic coupling in II (J = −28 cm−1) relative to IV (J = − 54 cm−1).19g,25
 |
| Fig. 3 Top: 298 K Mössbauer spectrum of II and associated spectral data for both II and IV.26,27 Bottom: cyclic voltammograms of II and IV (6 mM [Fe2(μ-O)2, 0.1 M [Bu4N][NTf2] in MeCN, 100 mV s−1). | |
The differences in electronic structures of the Fe(III)2(μ-O)2 cores within II and IV allude to distinct redox behaviors, which we investigated by cyclic voltammetry (CV, Fig. 3, bottom). The CV of II exhibited three principal features: a reversible 1e− reduction at −0.78 V (vs. Fc0/+, assigned as a Fe(III)2/Fe(III)Fe(II) couple), and two irreversible events at −1.2 V and +0.92 V, see SI. The well-behaved reduction event of II is in stark contrast to IV, where the CV was not well-defined, and instead, displayed broad features with an Eonset = −0.80 V (Fig. 3, see SI).28,29
To probe differences in reductive stability, we evaluated chemical reductions of II and IV. The reaction of II with 1 equiv. Na0 (as 5% Na/NaCl) in frozen THF afforded a mixture of products (as assessed by 1H-NMR spectroscopy) but did not induce demetalation. An analogous reduction of IV underwent immediate demetalation, generating free ligand as the sole TPAMe containing product. In contrast, introduction of milder reagents capable of delivering H+/e− equivalents (i.e. H-atom donors) provided tractable reactivity with both II and IV. Addition of 1.5 equiv. 1,2-diphenylhydrazine (DPH, N–HBDFE = 68 kcal mol−1)30 to II (23 °C, MeCN) provided quantitative formation of TPANHPhFe(II)(OH)+ (I-OH, Fig. 4) and azobenzene after 24 h, a net 2H+/2e− reduction of II.31 In comparison, addition of 1.5 equiv. DPH to IV immediately (ca. 15 min, 23 °C) produced a new species (assigned as TPAMeFe(II)(H2O)x(MeCN)y2+, III-H2O),32 with a low conversion to azobenzene (∼20%). The slower reduction of II by DPH relative to IV is consistent with distinct reductive stability profiles of these Fe(III)2(μ-O)2 cores as a result of secondary sphere H-bonding interactions.
 |
| Fig. 4 Reductions of II and IV with 1,2-diphenylhydrazine (DPH). | |
The H-bond dependent differences in reaction times for reduction of II and IV noted above are opposite from observations in Fig. 2 (NO2− reactivity). To investigate the role of H-bonds during the initial NO2− reduction step, we examined reactions between NO2− and the control compound TPAMeFe(NTf2)2 (III-NTf2), which maintains a similar primary sphere environment to I-NTf2 but does not contain secondary sphere H-bonds (Fig. 1b).13b Introduction of excess (ca. 10 equiv.) [Bu4N][NO2] to a colorless solution of III-NTf2 in MeCN immediately produced a yellow solution that exhibited five broad 1H-NMR resonances (assigned as NO2− binding). After 7 days, the solution turned orange and developed a characteristic UV-vis shoulder feature at 470 nm, corresponding to IV in 93% yield (see SI), and a separate NO trapping experiment provided CoTPP(NO) in 80% yield.
Formation of NO from NO2− can occur through multiple distinct pathways (inner- or outer-sphere 2H+/1e− reduction or through H+-mediated disproportionation);33 thus, we executed additional control experiments to provide further clarity into the mechanism for NO2− reduction in this system. To probe a disproportionation process, we introduced an exogenous acid, 1-methyl-2-(phenylamino)pyridinium34 ([HA], Fig. 5), which has a similar structure and charge, and therefore acidity/H-bond donor strength as the appended NHPh groups in I-NTf2, to excess [Bu4N][NO2] in MeCN. NO did not form in appreciable amounts after 5 h (3% yield of NO (via CoTPP(NO)). To examine outer-sphere reduction, we introduced ferrocene35 to a mixture of 10 equiv. [Bu4N][NO2], TPANHPh, and Zn(OTf)2, and again did not observe NO formation after 5 h. Collectively, these control experiments suggest that neither disproportionation nor outer-sphere reduction pathways proceed at rates comparable to those occurring during NO2− reduction with I-NTf2 or III-NTf2. Because the O-atoms from NO2− reduction are incorporated into the terminal Fe-containing products (II and IV), we propose that an inner-sphere reduction pathway is most likely.
 |
| Fig. 5 Assessment of alternative pathways for NO production from NO2−. | |
Since both I-NTf2 and III-NTf2 enabled spontaneous reduction of NO2− to NO, we quantified the effects of H-bonds on their kinetic profiles (Fig. 6). Evolution of NO (monitored via CoTPP trapping) from a reaction between I-NTf2 and 10 equiv. [Bu4N][NO2] fit well to a first-order exponential model with kobs = (2.3 ± 0.2) × 10−4 s−1. In contrast, with III-NTf2, we observed a slow reaction (linear fit) with kobs ≈ (3) × 10−8 s−1, four orders of magnitude slower than I-NTf2. To clarify the extent to which the N–O bond cleavage step is promoted by weak Brønsted acids (including the appended NHPh H-bond donors), we performed a control experiment with [HA] and III-NTf2. Introduction of 3 equiv. [HA] to a mixture of III-NTf2 and 10 equiv. [Bu4N][NO2] marginally enhanced NO2− reduction, affording kobs = (7.0 ± 0.9) × 10−8 s−1 and a concomitant increase in NO yield to 21% (5 h). These results contrast with the rapid reaction rate and quantitative NO yield (5 h) afforded by I-NTf2, suggesting a proximity requirement for the weakly acidic-HNPh groups to provide large rate accelerations for NO2− reduction activity as an H-bond donor.
 |
| Fig. 6 Kinetic profiles of NO evolution and lines of best fit for NO2− reduction with I-NTf2 (red circles, 1st order exponential fit), III-NTf2 (orange triangles, linear fit) or III-NTf2 + 3 equiv. [HA] and 10 equiv. [NO2−] (green squares, linear fit). Error bars represent the range of yields from reactions executed in triplicate. | |
Conclusions
In conclusion, we have reported the first examples of spontaneous NO2− reduction to afford Fe(III)2(μ-O)22+ cores, both in the presence and absence of appended H-bonds (II and IV respectively). The appended H-bonds within I-NTf2 provide a four-order of magnitude rate acceleration for NO2− reduction, relative to III-NTf2, which does not contain H-bonds. Control reactions using exogenous reagents of similar H+/e− strengths as I-NTf2 illustrate the requirement of preorganization of H-bonding units to facilitate rapid nitrite reduction.
The H-bond interactions surrounding the resulting Fe2(μ-O)2 core of II imparts distinct electronic and chemical properties relative to IV, and as a consequence, II is more challenging to reduce with H-atom donors. These observations illustrate the interplay between Fe–O stabilization and subsequent H+/e− transfer necessary to promote a net reductive transformation, of particular relevance to nitrite reductases and reaction intermediates containing Fe2(μ-O)2 cores of dioxygenases, as those found in sMMO-Q or RNR-X. Ongoing work in our lab is focusing on the effects of H-bonds on the electronic structure and reactivity of Fe(III)2(μ-O)2 cores, as well as their application toward catalytic nitrogen-oxyanion reduction.
Author contributions
The manuscript was written through the contributions of all authors. Project conceptualization: J. D. G. and N. K. S. collection of experimental data: A. R. L., J. E. G., W. S., and J. D. G. Research supervision: N. K. S.
Conflicts of interest
There are no conflicts of interest to declare.
Data availability
CCDC 2448814 and 2448815 contain the supplementary crystallographic data for this paper.36a,b
The data supporting this article have been included as part of the SI. Supplementary information: Experimental procedures and characterization of all species and reaction products. See DOI: https://doi.org/10.1039/d5sc04153h.
Acknowledgements
This work was supported by the NIGMS of the NIH under awards R35GM136360 (N. K. S.), and 2R15GM123380-02 (J. D. G). A. R. L. is an NSF-GRFP Fellow (DGE-2241144). Prof. Takele Seda is acknowledged for assistance in obtaining the Mössbauer spectrum of II, Dr Reza Loloee at Michigan State University for collection of SQUID magnetometry data, and Prof. Nicolai Lehnert and Claire Patterson for discussions relating to analysis of magnetometry data.
Notes and references
-
L. Ignarro, Nitric Oxide: Biology and Pathology, 2000 Search PubMed.
-
(a) J. Heinecke and P. C. Ford, Mechanistic studies of nitrite reactions with metalloproteins and models relevant to mammalian physiology, Coord. Chem. Rev., 2010, 254(3), 235–247 CrossRef CAS;
(b) S. A. Omar and A. J. Webb, Nitrite reduction and cardiovascular protection, J. Mol. Cell. Cardiol., 2014, 73, 57–69 CrossRef CAS PubMed.
-
(a)
S. L. Rose; S. V. Antonyuk; D. Sasaki; K. Yamashita; K. Hirata; G. Ueno; H. Ago; R. R. Eady; T. Tosha; M. Yamamoto; et al., An unprecedented insight into the catalytic mechanism of copper nitrite reductase from atomic-resolution and damage-free structures. Sci. Adv., 2021, 7(1), eabd8523 Search PubMed;
(b) S. Rinaldo, G. Giardina, N. Castiglione, V. Stelitano and F. Cutruzzolà, The catalytic mechanism of Pseudomonas aeruginosa cd1 nitrite reductase, Biochem. Soc. Trans., 2011, 39(1), 195–200 CrossRef CAS PubMed;
(c) S. Besson, M. G. Almeida and C. M. Silveira, Nitrite reduction in bacteria: a comprehensive view of nitrite reductases, Coord. Chem. Rev., 2022, 464, 214560 CrossRef CAS.
-
(a) F. Cutruzzolà, K. Brown, E. K. Wilson, A. Bellelli, M. Arese, M. Tegoni, C. Cambillau and M. Brunori, The nitrite reductase from Pseudomonas aeruginosa: essential role of two active-site histidines in the catalytic and structural properties, Proc. Natl. Acad. Sci. U. S. A., 2001, 98(5), 2232–2237 CrossRef PubMed;
(b) R. Silaghi-Dumitrescu, Linkage Isomerism in Nitrite Reduction by Cytochrome cd1 Nitrite Reductase, Inorg. Chem., 2004, 43(12), 3715–3718 CrossRef CAS PubMed;
(c) K. Kobayashi, S. Tagawa, Deligeer and S. Suzuki, The pH-Dependent Changes of Intramolecular Electron Transfer on Copper-Containing Nitrite Reductase1, J. Biochem., 1999, 126(2), 408–412 CrossRef CAS PubMed;
(d) K. Kataoka, H. Furusawa, K. Takagi, K. Yamaguchi and S. Suzuki, Functional Analysis of Conserved Aspartate and Histidine Residues Located Around the Type 2 Copper Site of Copper-Containing Nitrite Reductase1, J. Biochem., 2000, 127(2), 345–350 CrossRef CAS PubMed;
(e)
M. J. Boulanger; M. Kukimoto; M. Nishiyama; S. Horinouchi; M. E. P. MurphyCatalytic Roles for Two Water Bridged Residues (Asp-98 and His-255) in the Active Site of Copper-containing Nitrite ReductaseJ. Biol. Chem. 2000, 275 (31), 23957–23964. DOI:10.1074/jbc.M001859200, acccessed 2025/05/04;
(f) N. Xu, J. Yi and G. B. Richter-Addo, Linkage Isomerization in Heme–NOx Compounds: Understanding NO, Nitrite, and Hyponitrite Interactions with Iron Porphyrins, Inorg. Chem., 2010, 49(14), 6253–6266 CrossRef CAS PubMed.
-
(a) Z. Gordon, M. J. Drummond, J. A. Bogart, E. J. Schelter, R. L. Lord and A. R. Fout, Tuning the Fe(II/III) Redox Potential in Nonheme Fe(II)–Hydroxo Complexes through Primary and Secondary Coordination Sphere Modifications, Inorg. Chem., 2017, 56(9), 4852–4863 CrossRef CAS PubMed;
(b) C. W. J. Lockwood, B. Burlat, M. R. Cheesman, M. Kern, J. Simon, T. A. Clarke, D. J. Richardson and J. N. Butt, Resolution of Key Roles for the Distal Pocket Histidine in Cytochrome c Nitrite Reductases, J. Am. Chem. Soc., 2015, 137(8), 3059–3068 CrossRef CAS PubMed;
(c) E. T. Judd, N. Stein, A. A. Pacheco and S. J. Elliott, Hydrogen Bonding Networks Tune Proton-Coupled Redox Steps during the Enzymatic Six-Electron Conversion of Nitrite to Ammonia, Biochemistry, 2014, 53(35), 5638–5646 CrossRef CAS PubMed.
- A. S. Borovik, Bioinspired Hydrogen Bond Motifs in Ligand Design: The Role of Noncovalent Interactions in Metal Ion Mediated Activation of Dioxygen, Acc. Chem. Res., 2005, 38(1), 54–61 CrossRef CAS PubMed.
-
(a) S. Atta, A. Mandal, R. Saha and A. Majumdar, Reduction of nitrite to nitric oxide and generation
of reactive chalcogen species by mononuclear Fe(II) and Zn(II) complexes of thiolate and selenolate, Dalton Trans., 2024, 53(3), 949–965, 10.1039/D3DT03768A;
(b) S. Karmakar, S. Patra, R. Halder, S. Karmakar and A. Majumdar, Reduction of Nitrite in an Iron(II)-Nitrito Compound by Thiols and Selenol Produces Dinitrosyl Iron Complexes via an {FeNO}7 Intermediate, Inorg. Chem., 2024, 63(49), 23202–23220 CrossRef CAS PubMed;
(c) Kulbir, S. Das, T. Devi, S. Ghosh, S. Chandra Sahoo and P. Kumar, Acid-induced nitrite reduction of nonheme iron(ii)-nitrite: mimicking biological Fe–NiR reactions, Chem. Sci., 2023, 14(11), 2935–2942, 10.1039/D2SC06704H;
(d) C.-C. Tsou, W.-L. Yang and W.-F. Liaw, Nitrite Activation to Nitric Oxide via One-fold Protonation of Iron(II)-O,O-nitrito Complex: Relevance to the Nitrite Reductase Activity of Deoxyhemoglobin and Deoxyhemerythrin, J. Am. Chem. Soc., 2013, 135(50), 18758–18761 CrossRef CAS PubMed;
(e) G. Cioncoloni, I. Roger, P. S. Wheatley, C. Wilson, R. E. Morris, S. Sproules and M. D. Symes, Proton-Coupled Electron Transfer Enhances the Electrocatalytic Reduction of Nitrite to NO in a Bioinspired Copper Complex, ACS Catal., 2018, 8(6), 5070–5084 CrossRef CAS;
(f) J. R. Stroka, B. Kandemir, E. M. Matson and K. L. Bren, Electrocatalytic Multielectron Nitrite Reduction in Water by an Iron Complex, ACS Catal., 2020, 10(23), 13968–13972 CrossRef CAS;
(g) B. Tran and J. Smith, Nitrogen oxyanion deoxygenation: redox chemistry and oxygen atom transfer reactions, Coord. Chem. Rev., 2025, 530, 216490 CrossRef CAS.
-
(a) A. Sarkar, S. Bhakta, S. Chattopadhyay and A. Dey, Role of distal arginine residue in the mechanism of heme nitrite reductases, Chem. Sci., 2023, 14(29), 7875–7886, 10.1039/D3SC01777J;
(b) J. M. Moore, T. J. Miller, M. Mu, M. N. Peñas-Defrutos, K. L. Gullett, L. S. Elford, S. Quintero, M. García-Melchor and A. R. Fout, Selective Stepwise Reduction of Nitrate and Nitrite to Dinitrogen or Ammonia, J. Am. Chem. Soc., 2025, 147(10), 8444–8454 CrossRef CAS PubMed;
(c) Y. J. Park, M. N. Peñas-Defrutos, M. J. Drummond, Z. Gordon, O. R. Kelly, I. J. Garvey, K. L. Gullett, M. García-Melchor and A. R. Fout, Secondary Coordination Sphere Influences the Formation of Fe(III)–O or Fe(III)–OH in Nitrite Reduction: A Synthetic and Computational Study, Inorg. Chem., 2022, 61(21), 8182–8192 CrossRef CAS PubMed;
(d) E. M. Matson, Y. J. Park and A. R. Fout, Facile Nitrite Reduction in a Non-heme Iron System: Formation of an Iron(III)-Oxo, J. Am. Chem. Soc., 2014, 136(50), 17398–17401 CrossRef CAS PubMed;
(e) K. T. Burns, W. R. Marks, P. M. Cheung, T. Seda, L. N. Zakharov and J. D. Gilbertson, Uncoupled Redox-Inactive Lewis Acids in the Secondary Coordination Sphere Entice Ligand-Based Nitrite Reduction, Inorg. Chem., 2018, 57(16), 9601–9610 CrossRef CAS PubMed;
(f) A. Mondal, K. P. Reddy, S. Som, D. Chopra and S. Kundu, Nitrate and Nitrite Reductions at Copper(II) Sites: Role of Noncovalent Interactions from Second-Coordination-Sphere, Inorg. Chem., 2022, 61(50), 20337–20345 CrossRef CAS PubMed;
(g) J. M. Moore and A. R. Fout, Synthetic strategies for oxyanion reduction: metal-based insights and innovations, Coord. Chem. Rev., 2025, 541, 216692 CrossRef CAS.
- C. L. Ford, Y. J. Park, E. M. Matson, Z. Gordon and A. R. Fout, A bioinspired iron catalyst for nitrate and perchlorate reduction, Science, 2016, 354(6313), 741–743 CrossRef CAS PubMed.
-
(a) Y. M. Kwon, M. Delgado, L. N. Zakharov, T. Seda and J. D. Gilbertson, Nitrite reduction by a pyridinediimine complex with a proton-responsive secondary coordination sphere, Chem. Commun., 2016, 52(73), 11016–11019, 10.1039/C6CC05962G;
(b) P. M. Cheung, K. T. Burns, Y. M. Kwon, M. Y. Deshaye, K. J. Aguayo, V. F. Oswald, T. Seda, L. N. Zakharov, T. Kowalczyk and J. D. Gilbertson, Hemilabile Proton Relays and Redox Activity Lead to {FeNO}x and Significant Rate Enhancements in NO2− Reduction, J. Am. Chem. Soc., 2018, 140(49), 17040–17050 CrossRef CAS PubMed.
-
(a) W.-M. Ching, C.-H. Chuang, C.-W. Wu, C.-H. Peng and C.-H. Hung, Facile Nitrite Reduction and Conversion Cycle of {Fe(NO)}6/7 Species: Chemistry of Iron N-Confused Porphyrin Complexes via Protonation/Deprotonation, J. Am. Chem. Soc., 2009, 131(23), 7952–7953 CrossRef CAS PubMed;
(b) W.-M. Ching, P. P.-Y. Chen and C.-H. Hung, A mechanistic study of nitrite reduction on iron(II) complexes of methylated N-confused porphyrins, Dalton Trans., 2017, 46(43), 15087–15094, 10.1039/C7DT02869E.
-
(a) Y. Zang, J. Kim, Y. Dong, E. C. Wilkinson, E. H. Appelman and L. Que, Models for Nonheme Iron Intermediates: Structural Basis for Tuning the Spin States of Fe(TPA) Complexes, J. Am. Chem. Soc., 1997, 119(18), 4197–4205 CrossRef CAS;
(b) K. A. Jesse and J. S. Anderson, Leveraging ligand-based proton and electron transfer for aerobic reactivity and catalysis, Chem. Sci., 2024, 15(40), 16409–16423, 10.1039/D4SC03896G;
(c) M. R. Elsby, A. Kumar, L. M. Daniels, M. Z. Ertem, N. Hazari, B. Q. Mercado and A. H. Paulus, Linear Free Energy Relationships Associated with Hydride Transfer From [(6,6′-R2-bpy)Re(CO)3H]: A Cautionary Tale in Identifying Hydrogen Bonding Effects in the Secondary Coordination Sphere, Inorg. Chem., 2024, 63(41), 19396–19407 CrossRef CAS PubMed;
(d) J. C. M. Rivas, S. L. Hinchley, L. Metteau and S. Parsons, The strength of hydrogen bonding to metal-bound ligands can contribute to changes in the redox behaviour of metal centres, Dalton Trans., 2006,(19), 2316–2322, 10.1039/B516234C.
-
(a) W. Sarkar, A. LaDuca, J. R. Wilson and N. K. Szymczak, Iron-Catalyzed C–H Oxygenation Using Perchlorate Enabled by Secondary Sphere Hydrogen Bonds, J. Am. Chem. Soc., 2024, 146(15), 10508–10516 CrossRef CAS PubMed;
(b) A. R. LaDuca, J. R. Wilson, W. Sarkar, M. Zeller and N. K. Szymczak, Impact of Secondary Sphere Hydrogen Bonds on O2 Reactivity within a Nonheme Iron Complex, J. Am. Chem. Soc., 2025, 147(6), 5099–5105 CrossRef CAS PubMed;
(c) W. Sarkar and N. K. Szymczak, Expanding Perchlorate Use for C–H Oxidative Transformations: A Tandem Photo- and Iron-Catalytic Strategy, Organometallics, 2025, 44(7), 777–782 CrossRef CAS PubMed;
(d) J. R. Wilson, M. Zeller and N. K. Szymczak, Hydrogen-bonded nickel(i) complexes, Chem. Commun., 2021, 57(6), 753–756 RSC;
(e) J. P. Shanahan, D. M. Mullis, M. Zeller and N. K. Szymczak, Reductively Stable Hydrogen-Bonding Ligands Featuring Appended CF2–H Units, J. Am. Chem. Soc., 2020, 142(19), 8809–8817, DOI:10.1021/jacs.0c01718;
(f) E. W. Dahl, J. J. Kiernicki, M. Zeller and N. K. Szymczak, Hydrogen Bonds Dictate O2 Capture and Release within a Zinc Tripod, J. Am. Chem. Soc., 2018, 140(32), 10075–10079 CrossRef CAS PubMed;
(g) E. W. Dahl, H. T. Dong and N. K. Szymczak, Phenylamino derivatives of tris(2-pyridylmethyl)amine: hydrogen-bonded peroxodicopper complexes, Chem. Commun., 2018, 54(8), 892–895, 10.1039/C7CC08619A;
(h) E. W. Dahl and N. K. Szymczak, Hydrogen Bonds Dictate the Coordination Geometry of Copper: Characterization of a Square-Planar Copper(I) Complex, Angew. Chem., Int. Ed., 2016, 55(9), 3101–3105 CrossRef CAS PubMed.
- C. M. Moore and N. K. Szymczak, Nitrite reduction by copper through ligand-mediated proton and electron transfer, Chem. Sci., 2015, 6(6), 3373–3377, 10.1039/C5SC00720H.
- A. V. Iogansen, Direct proportionality of the hydrogen bonding energy and the intensification of the stretching ν(XH) vibration in infrared spectra, Spectrochim. Acta, Part A, 1999, 55(7), 1585–1612 CrossRef.
- Relative to I-NTf2, the reaction evolved 1 equivalent of NO gas.
- We propose that the dimeric structure of II stays intact in solution because variable temperature 1H-NMR spectroscopy of II in MeCN did not display significant changes between +80 to −35 °C.
-
(a) A. J. Jasniewski and L. Que Jr, Dioxygen Activation by Nonheme Diiron Enzymes: Diverse Dioxygen Adducts, High-Valent Intermediates, and Related Model Complexes, Chem. Rev., 2018, 118(5), 2554–2592 CrossRef CAS PubMed;
(b) R. Banerjee, J. C. Jones and J. D. Lipscomb, Soluble Methane Monooxygenase, Annu. Rev. Biochem., 2019, 88, 409–431 CrossRef CAS PubMed;
(c) A. B. Jacobs, R. Banerjee, D. E. Deweese, A. Braun, J. T. Babicz Jr, L. B. Gee, K. D. Sutherlin, L. H. Böttger, Y. Yoda and M. Saito,
et al., Nuclear Resonance Vibrational Spectroscopic Definition of the Fe(IV)2 Intermediate Q in Methane Monooxygenase and Its Reactivity, J. Am. Chem. Soc., 2021, 143(39), 16007–16029 CrossRef CAS PubMed;
(d) B. E. Sturgeon, D. Burdi, S. Chen, B.-H. Huynh, D. E. Edmondson, J. Stubbe and B. M. Hoffman, Reconsideration of X, the Diiron Intermediate Formed during Cofactor Assembly in E. coli Ribonucleotide Reductase, J. Am. Chem. Soc., 1996, 118(32), 7551–7557 CrossRef CAS.
-
(a) D. Kass, S. Yao, K. B. Krause, T. Corona, L. Richter, T. Braun, S. Mebs, M. Haumann, H. Dau and T. Lohmiller,
et al., Spectroscopic Properties of a Biologically Relevant [Fe2(μ-O)2] Diamond Core Motif with a Short Iron-Iron Distance, Angew. Chem., Int. Ed., 2023, 62(10), e202209437 CrossRef CAS PubMed;
(b) H. Zheng, Y. Zang, Y. Dong, V. G. Young and L. Que, Complexes with FeIII2(μ-O)(μ-OH), FeIII2(μ-O)2, and [FeIII3(μ2-O)3] Cores: Structures, Spectroscopy, and Core Interconversions, J. Am. Chem. Soc., 1999, 121(10), 2226–2235 CrossRef CAS;
(c) Y. Honda, H. Arii, T. Okumura, A. Wada, Y. Funahashi, T. Ozawa, K. Jitsukawa and H. Masuda, Complexes with FeIII2(μ-O)(μ-OH) Core Surrounded by Hydrogen-Bonding Interaction, Bull. Chem. Soc. Jpn., 2007, 80(7), 1288–1295 CrossRef CAS;
(d) G. T. Rohde, G. Xue and L. Que, Explorations of the nonheme high-valent iron-oxo landscape: crystal structure of a synthetic complex with an [FeIV2(μ-O)2] diamond core relevant to the chemistry of sMMOH, Faraday Discuss., 2022, 234, 109–128, 10.1039/D1FD00066G;
(e) H.-F. Hsu, Y. Dong, L. Shu, V. G. Young and L. Que, Crystal Structure of a Synthetic High-Valent Complex with an Fe2(μ-O)2 Diamond Core. Implications for the Core Structures of Methane Monooxygenase Intermediate Q and Ribonucleotide Reductase Intermediate X, J. Am. Chem. Soc., 1999, 121(22), 5230–5237 CrossRef CAS;
(f) L. Fohlmeister, K. R. Vignesh, F. Winter, B. Moubaraki, G. Rajaraman, R. Pöttgen, K. S. Murray and C. Jones, Neutral diiron(iii) complexes with Fe2(μ-E)2 (E = O, S, Se) core structures: reactivity of an iron(I) dimer towards chalcogens, Dalton Trans., 2015, 44(4), 1700–1708, 10.1039/C4DT03081H;
(g) Y. Zang, Y. Dong, L. Que Jr, K. Kauffmann and E. Muenck, The First Bis(.mu.-oxo)diiron(III) Complex. Structure and Magnetic Properties of [Fe2(.mu.-O)2(6TLA)2](ClO4)2, J. Am. Chem. Soc., 1995, 117(3), 1169–1170 CrossRef CAS;
(h) Y. Mikata, Y. Aono, C. Yamamoto, H. Nakayama, A. Matsumoto, F. Kotegawa, M. Harada, H. Katano, Y. Kobayashi and S. Yanagisawa,
et al., A Synthetic Model for the Possible
FeIV2(μ-O)2 Core of Methane Monooxygenase Intermediate Q Derived from a Structurally Characterized FeIIIFeIV(μ-O)2 Complex, Inorg. Chem., 2022, 61(2), 786–790 CrossRef CAS PubMed;
(i) P. Zhao, H. Lei, C. Ni, J.-D. Guo, S. Kamali, J. C. Fettinger, F. Grandjean, G. J. Long, S. Nagase and P. P. Power, Quasi-Three-Coordinate Iron and Cobalt Terphenoxide Complexes {AriPr8OM(μ-O)}2 (AriPr8 = C6H-2,6-(C6H2-2,4,6-iPr3)2-3,5-iPr2; M = Fe or Co) with M(III)2(μ-O)2 Core Structures and the Peroxide Dimer of 2-Oxepinoxy Relevant to Benzene Oxidation, Inorg. Chem., 2015, 54(18), 8914–8922 CrossRef CAS PubMed.
- The structure of IV reported in this work is used for comparing structural metrics because it is of higher quality than the previous polymorph (R1 = 3.95% versus R1 = 6.5% in ref. 19g).
- A. Dey, R. K. Hocking, P. Larsen, A. S. Borovik, K. O. Hodgson, B. Hedman and E. I. Solomon, X-ray Absorption Spectroscopy and Density Functional Theory Studies of [(H3buea)FeIII–X]n− (X = S2−, O2−, OH−): Comparison of Bonding and Hydrogen Bonding in Oxo and Sulfido Complexes, J. Am. Chem. Soc., 2006, 128(30), 9825–9833 CrossRef CAS PubMed.
- V. F. Oswald, J. L. Lee, S. Biswas, A. C. Weitz, K. Mittra, R. Fan, J. Li, J. Zhao, M. Y. Hu and E. E. Alp,
et al., Effects of Noncovalent Interactions on High-Spin Fe(IV)–Oxido Complexes, J. Am. Chem. Soc., 2020, 142(27), 11804–11817 CrossRef CAS PubMed.
-
(a)
G. J. Long and F. Grandjean, 2.20 – Mössbauer Spectroscopy: Introduction, in Comprehensive Coordination Chemistry II, ed. McCleverty, J. A. and Meyer, T. J., Pergamon, 2003, pp. 269–277 Search PubMed;
(b) We tabulated the bond metrics and Mössbauer parameters of biological and synthetic Fe2(μ-O)2 cores (see Table S1). The spectral parameters for II are similar to the complexes in ref. 19a, f, and i..
- E. Zars, L. Gravogl, M. R. Gau, P. J. Carroll, K. Meyer and D. J. Mindiola, Isostructural bridging diferrous chalcogenide cores [FeII(μ-E)FeII] (E = O, S, Se, Te) with decreasing antiferromagnetic coupling down the chalcogenide series, Chem. Sci., 2023, 14(24), 6770–6779, 10.1039/D3SC01094E.
-
(a) The temperature-dependent magnetic susceptibility of II (SQUID) was modelled with the exchange Hamiltonian Ĥex = −J × S1S2 using Easyspin’s curry function.;
(b) S. Stoll and A. Schweiger, EasySpin, a comprehensive software package for spectral simulation and analysis in EPR, J. Magn. Reson., 2006, 178(1), 42–55 CrossRef CAS PubMed.
- The Mössbauer spectrum of II (as a polycrystalline solid) was collected at 298 K using a constant-acceleration spectrometer (Wissel GMBH, Germany) in a horizontal transmission mode using a 50 mCi 57Co source was utilized for Mössbauer spectra collection. All Mössbauer spectra were collected at room temperature at 4 mm s−1 velocity with approximately 80 mg of sample loaded into an acrylic sample holder with minimal Paratone-N oil to prevent oxidation. Lorentzian line shape with the NORMOS (Wissel GMBH) least-squares fitting program was used to fit spectra, and isomer shifts were normalized to metallic iron. We attribute the slight asymmetry in the two peaks to the random orientation of the polycrystalline sample. See ref. 27.
- E. Kreber and U. Gonser, The Problem of Texture in Mössbauer Spectroscopy, Texture, Stress, Microstruct., 1974, 1(4), 869318 Search PubMed.
- At −40 °C, the CV of IV has been reported to feature a reversible Fe(III)2 to Fe(III)Fe(IV) wave (E1/2 = 0.41 V) that we did not observe at ambient temperature. See ref. 29.
- H. Zheng, S. J. Yoo, E. Münck and L. Que, The Flexible Fe2(μ-O)2 Diamond Core: A Terminal Iron(IV)−Oxo Species Generated from the Oxidation of a Bis(μ-oxo)diiron(III) Complex, J. Am. Chem. Soc., 2000, 122(15), 3789–3790 CrossRef CAS.
- J. J. Warren, T. A. Tronic and J. M. Mayer, Thermochemistry of Proton-Coupled Electron Transfer Reagents and its Implications, Chem. Rev., 2010, 110(12), 6961–7001 CrossRef CAS PubMed.
- Treatment of II with substrates containing weak C–H bonds (i.e. 9,10-dihydroanthracene or Ph3CH) as reductants did not afford reduction of II nor C–H oxidation (to form anthracene, anthraquinone, or Ph3COH). See SI for further details.
- S. V. Kryatov, E. V. Rybak-Akimova, V. L. MacMurdo and L. Que, A Mechanistic Study of the Reaction between a Diiron(II) Complex [FeII2(μ-OH)2(6-Me3-TPA)2]2+ and O2 to Form a Diiron(III) Peroxo Complex, Inorg. Chem., 2001, 40(10), 2220–2228 CrossRef CAS PubMed.
-
(a) M. Knipp and C. He, Nitrophorins: nitrite disproportionation reaction and other novel functionalities of insect heme-based nitric oxide transport proteins, IUBMB Life, 2011, 63(5), 304–312 CrossRef CAS PubMed;
(b) M. S. Rayson, J. C. Mackie, E. M. Kennedy and B. Z. Dlugogorski, Accurate Rate Constants for Decomposition of Aqueous Nitrous Acid, Inorg. Chem., 2012, 51(4), 2178–2185 CrossRef CAS PubMed;
(c) V. Hosseininasab, I. M. DiMucci, P. Ghosh, J. A. Bertke, S. Chandrasekharan, C. J. Titus, D. Nordlund, J. H. Freed, K. M. Lancaster and T. H. Warren, Lewis acid-assisted reduction of nitrite to nitric and nitrous oxides via the elusive nitrite radical dianion, Nat. Chem., 2022, 14(11), 1265–1269 CrossRef CAS PubMed.
- 1-Methyl-2-(phenylamino)pyridinium [HA] features an ν(NH) band at 3272 cm−1, while I-NTf2's ν(NH) is at 3278 cm−1 (see SI). This further supports our assertion that [HA] and I-NTf2 have similar acidity/H-bond donor strengths..
- We selected ferrocene as a reductant in this control reaction because it approximates the reducing potential of I-NTf2 and III-NTf2 in the presence of 10 equiv. [Bu4N][NO2] (0.05 V vs. Fc0/+) see SI..
-
(a)
A. R. LaDuca, J. E. Gonder, W. Sarkar, J. D. Gilbertson and N. Szymczak, CCDC 2448814: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2n6603;
(b)
A. R. LaDuca, J. E. Gonder, W. Sarkar, J. D. Gilbertson and N. Szymczak, CCDC 2448815: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2n6614.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.