Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Mar 27.
Published in final edited form as: J Am Chem Soc. 2024 Mar 15;146(12):7915–7921. doi: 10.1021/jacs.3c12337

A Nonheme Iron(III) Superoxide Complex Leads to Sulfur Oxygenation

Sudha Yadav 1, Vishal Yadav 2, Maxime A Siegler 3, Pierre Moënne-Loccoz 4, Guy N L Jameson 5, David P Goldberg 6
PMCID: PMC11318076  NIHMSID: NIHMS2013043  PMID: 38488295

Abstract

A new alkylthiolate-ligated nonheme iron complex, FeII(BNPAMe2S)Br (1), is reported. Reaction of 1 with O2 at −40 °C, or reaction of the ferric form with O2•− at −80 °C, gives a rare iron(III)-superoxide intermediate, [FeIII(O2)(BNPAMe2S)]+ (2), characterized by UV–vis, 57Fe Mössbauer, ATR-FTIR, EPR, and CSIMS. Metastable 2 then converts to an S-oxygenated FeII(sulfinate) product via a sequential O atom transfer mechanism involving an iron-sulfenate intermediate. These results provide evidence of the feasibility of proposed intermediates in thiol dioxygenases.


The activation of dioxygen by nonheme iron enzymes is critical to a range of biological processes.15 One such class of nonheme iron enzymes are the thiol dioxygenases (TDOs), which convert thiol substrates into S-oxygenated compounds and are found in a wide range of organisms, from bacteria,6,7 to plants,8,9 to mammals.10,11 The TDOs play critical roles in biochemical processes,1216 and their misfunctioning has been implicated in neurodegenerative diseases,17,18 autoimmune disorders,19,20 and cancer.2123 The TDOs utilize a mononuclear iron center bound by an unusual 3-His structural motif, as opposed to the canonical 2-His-1-carboxylate iron site found in other nonheme iron enzymes2429. Despite their widespread biochemical significance, the mechanism of S-oxygenation for TDOs remains poorly understood.3032 For example, two hypothetical mechanisms have been proposed for cysteine dioxygenase (Scheme 1), a mammalian TDO that converts cysteine to cysteine sulfinate.1,3335 However, experimental evidence for either mechanism remains scarce.3640 In synthetic systems, significant efforts have been made to bind and activate O2 with nonheme iron complexes.4149 These efforts have yielded only three examples of Fe(O2) adducts, and none of these adducts give dioxygenated sulfinate products.41,45,46

Scheme 1.

Scheme 1.

Possible Mechanisms for S-Oxygenation in CDO, an Example of the Broader Class of Enzymes Known as TDOs

Herein, a new tetradentate ligand, (bis((6-(neopentylamino)pyridinyl)methyl)amino)-2-methylpropane-2-thiol (BNPAMe2SH) is reported, designed to provide three neutral nitrogen donors and one anionic thiolate donor, as found in the TDOs. The second coordination sphere is engineered to stabilize metal–oxygen intermediates through the inclusion of hydrogen bonding substituents. An iron(II) complex, FeII(BNPAMe2S)(Br) (1), reacts with O2 to give a new FeIII(O2•−) species, which was trapped and characterized at low temperature. This ferric superoxide complex converts to an FeII(sulfinate) product upon warming, and mechanistic analyses indicate that a mono-oxygenated sulfenate intermediate must be on the pathway to the dioxygenated product.

An abbreviated synthesis of the new tetradentate ligand BNPAMe2SH, together with the preparation of the nonheme iron complex FeII(BNPAMe2S)Br (1), is shown in Scheme 2. The secondary amine was synthesized by a Fukuyama amine synthesis,50 followed by the installation of an alkylthiolate donor from dimethylthiirane, and finally LAH reduction to the target ligand BNPAMe2SH. Deprotonation of the thiol followed by metalation gives FeII(BNPAMe2S)Br (1). Complex 1 was crystallized from pentane/THF at −35 °C, giving yellow blocks suitable for single-crystal X-ray diffraction. The structure of 1 (Figure 1) reveals a 5-coordinate, iron(II) complex with a bromide ligand occupying the fifth site. The neopentyl amine groups are oriented toward the bromide ligand, forming two intramolecular hydrogen bonds with N1(H)–Br = 3.404(5)°, N1–H···Br1 = 161(5)°, and N5(H)–Br = 3.361(5) Å, N5–H···Br1 = 154(6)°. The 1H nuclear magnetic resonance (NMR) spectrum of 1 in CD3CN reveals paramagnetically shifted peaks ranging from +100 to −20 ppm, consistent with a high-spin (S = 2) FeII complex. Mössbauer spectroscopy on 57Fe-labeled 1 in THF at 80 K shows a sharp quadrupole doublet with δ = 0.96 and |ΔEQ|=2.80 mm s−1, confirming a high-spin ferrous ion (Figure 2).51

Scheme 2.

Scheme 2.

Synthesis of 1

Figure 1.

Figure 1.

Displacement ellipsoid plot (50% probability level) for 1 at 110(2) K. Hydrogen atoms (except for N–H) have been omitted for the sake of clarity.

Figure 2.

Figure 2.

Zero-field 57Fe Mössbauer spectra of (a) 1, 80 K, (b) 2, 80 K, (c) 2, 5 K, and (d) final reaction mixture after warming up 2 to 298 K in THF, 80 K. Isomer shift (δ) and quadrupole splitting (|ΔEQ|) values are given in mm s−1. Solid lines represent the best fits.

The pale-yellow solution of 1 in THF exhibits a UV–vis absorption band at 330 nm (ε = 9600 M−1 cm−1). Exposure of 1 to O2 in THF, 2-MeTHF, or CH3CN at −40 °C generates a dark green species (2) that features an absorption band at 620 nm (ε = 1400 M−1 cm−1), which decays upon warming to 23 °C. Spectral titration of 1 with an O2-saturated solution of THF at −40 °C shows an Fe/O2 1:1 binding stoichiometry, consistent with the rapid formation of a mononuclear Fe(O2) adduct, 2 (Figure 3). Bubbling argon for 1 h through a solution of 2 results in no spectral change, and the formation of 2 cannot be reversed by repeated vacuum/purge cycles. Thus, the binding of the aqueous O2 to 1 is irreversible at −40 °C.

Figure 3.

Figure 3.

UV–vis spectra showing the conversion of 1 (0.5 mM, black line) to 2 (red line) after stoichiometric O2 addition in THF at −40 °C.

Examination of 57Fe-labeled 2 in THF by Mössbauer spectroscopy reveals a single quadruple doublet with parameters δ = 0.45 mm s−1 and |ΔEQ|=1.25 mm s−1, accounting for >90% of the total area (Figure 2b). Analysis at 5 K in the absence of a magnetic field shows a quadrupole doublet with the same parameters but narrower line widths (Figure 2c). The Mössbauer data are consistent with an integer spin ground state arising from a spin-coupled iron(III)-superoxide species.52,53 The X-band EPR spectrum for 2 (15 K) is silent, providing further evidence of an integer spin ground state. Solution-phase magnetic measurements (Evans method) for 2 in CD3CN indicate an apparent S = 2 spin ground state with μeff = 4.36 μB, assuming a concentration of 2 equal to the total iron content. This ground state can be explained by antiferromagnetic coupling between a high-spin FeIII ion (S = 5/2) and an O2•− ion (S = 1/2).

Analysis of 2 by cryospray ionization mass spectrometry (CSI-MS) shows an intense cluster at 544.23 m/z (Figure S6), consistent with the addition of two oxygen atoms and loss of one bromine atom from 1. Upon addition of the isotopically pure 18O2 (99% 18O) to 1, the center of the cluster shifts to 548.22 m/z, and fitting of the isotope pattern indicates 63% incorporation of 18O.

An alternative pathway to intermediate 2 involves potassium superoxide (Scheme 3). Reaction of 1 with the one-electron oxidant FcBArF4 in THF at −80 °C leads to a new UV–vis feature at 710 nm (Figure S9), representative of the new species 3. Mössbauer spectroscopy on 57Fe-labeled 3 (Figure S10) reveals a broad quadrupole doublet with δ = 0.39 mm s−1 and |ΔEQ|=1.04 mm s−1, typical for a high-spin iron(III) ion, and assigned to [FeIII(BNPAMe2S)(Br)]+. Addition of a slight excess of KO2 and Kryptofix 222 in THF/CH3CN (~20/1) to 3 at −80 °C gives a new UV–vis feature at 620 nm, which matches the intermediate 2 generated from 1/O2. Mössbauer analysis following addition of KO2 to 57Fe-labeled 3 also matches that seen for 2.

Scheme 3.

Scheme 3.

Formation of 2 from 1 and 3

Characterization of 2 by resonance Raman spectroscopy was attempted with both red and blue excitations, but no 16O/18O sensitive bands were observed. However, we were able to successfully isolate 2 in the solid state at a low temperature, which allowed for interrogation by ATR-FTIR spectroscopy. Precipitation of 2 at −80 °C as a green solid, followed by rapid transfer to the ATR stage under cold N2 gas led to the spectrum in Figure 4. This spectrum revealed intense bands in the 1300–1500 cm−1 region for the pyridyl groups,54 as well as an isotopically sensitive peak at 1061 cm−1, which shifts to 1008 cm−1 with 18O2 (99%). The band at 1061 cm−1 is assigned to an O–O stretch which downshifts by 53 cm−1 with 18O substitution, in good agreement with a harmonic oscillator. The frequency of the O–O stretch matches that for a metal-superoxide species (Table S3).5557

Figure 4.

Figure 4.

ATR-FTIR spectra of 1 (dotted line), 2(16O2 natural abundance) (black line), 2(18O2 (99%)) (red line), difference spectrum for 2(16O218O2) (blue line); 4(16O2 natural abundance) (green line), and 4(18O2 (99%)) (purple line).

The density functional theory (DFT) optimized geometry for 1 matches the crystal structure and was employed as the starting point for calculations on 2. The η2-O2 (side-on) and η1-O2 (end-on) structures were evaluated, both in an S = 2 ground state, as seen experimentally. The optimized end-on structure is 6-coordinate, with the bromide ion coordinated in the open site, whereas for the side-on geometry, the bromide ion is expelled from the coordination sphere during optimization. If the bromide ligand is removed from the starting structure with the η1-O2 binding mode, then the structure converges to the side-on geometry. Mössbauer parameters calculated for the side-on structure give a significantly better match with experimental results (Figure 5). Time-dependent DFT calculations predict a thiolate-to-Fe(O2) charge transfer transition at 627 nm for the side-on structure, as opposed to 500 nm for the end-on structure, providing further support for an η2 binding mode (Table S4). These calculations support the formation of a side-on iron(III)-superoxide intermediate; however, further experimental evidence is needed to confirm the binding mode.

Figure 5.

Figure 5.

DFT geometry optimized structures and Mössbauer parameters calculations (δ, |ΔEQ|(mms1)) for side-on vs end-on binding modes of 2 in the S = 2 spin state.

Complex 2 in THF is stable at −40 °C, but upon warming to 23 °C, this species decays into two new iron products, as observed by Mössbauer spectroscopy (Figure 2d). Fitting of the spectrum is achieved with two quadrupole doublets: one with δ = 0.47 mm s−1, |ΔEQ|=0.77 mm s−1, consistent with high-spin FeIII, and one with δ = 1.23 mm s−1, |ΔEQ|=2.96 mm s−1, consistent with high-spin FeII. One of the products was identified by crystallization, FeIII(BNPAMe2S)(OH)(Br) (5) (Figure S20, Scheme 4). Mössbauer spectroscopy on crystalline 5 in THF gives a single quadrupole doublet that matches the ferric component of the final reaction mixture. The EPR spectrum for 5 also matches that of high-spin iron(III) (Figure S21). A possible pathway for the formation of 5 involves release of O2 upon warming, and then reaction of the iron(II) complex with O2 at room temperature leads to the FeIII(OH) product, as seen in a previous system.5860 Addition of O2 to 1 at room temperature leads to the FeIII(OH) product in approximately 60% yield, as well as an unidentified iron(II) species that has similar Mössbauer parameters to 4, but does not show any S═O vibrational modes in FTIR (Figure S23).

Scheme 4.

Scheme 4.

Thermal Decay of 2

Further analysis of the final products by ATR-FTIR shows the loss of ν(O–O) at 1061 cm−1 and the appearance of two new absorption bands at 1233 and 1207 cm−1, which fall in the range of metal-sulfinate stretches (Figure 4).6169 Warm up of the 18O2-labeled 2 leads to downshifts in these vibrations to 1215 and 1189 cm−1. These shifts are consistent with the DFT calculated shifts for RSO2–M vibrations.44 The Mössbauer and IR data together point to the formation of an FeII(sulfinate) complex as one of the final products following oxygenation. Laser desorption ionization mass spectrometry (LDI-MS) of the final reaction mixture revealed a prominent isotopic cluster centered at 544.5 m/z, which matches that of [FeII(BNPAMe2SO2)]+ (4). Addition of 18O2 leads to a shift in the isotopic cluster, and the best simulations of the isotopic pattern involve a mixture of S16O2 (40%), S16O18O (40%), and S18O2 (20%). These results suggest that O2 is a source of O atoms for the S-oxygenated product, but there is significant exchange with exogenous H2O prior to formation of the final product. Addition of H218O to the reaction with natural abundance O2 leads to S18O2 (85%) and S16O18O (13%), while addition of H218O to the final Fe(sulfinate) product does not lead to incorporation of any 18O isotope. Addition of H218O also does not lead to any 18O-labeling in the FeIII(O2•−) species. These data support a mechanism which involves the formation of one or more intermediates after FeIII(O2•−) that are susceptible to H218O exchange.

A proposed mechanism for S-oxygenation is shown in Scheme 4 and follows similar proposals for the TDOs. After FeIII(O2•−) formation, S–O bond formation and O–O homolytic cleavage occur, giving a putative FeIV(O)(sulfenate) intermediate. This intermediate can be susceptible to exchange with H218O; both FeIV(O) and sulfenate species70,71 are known to exchange with H2O in synthetic systems.72,73 Some evidence for CDO indicates H218O does not lead to isotopic labeling in the sulfinate product, but in this case, the nature of the enclosed active site may prevent this exchange.38,74 To test this mechanism further, we added the known sulfenate trap dimedone. LDI-MS analysis shows a peak at 650.4 m/z, corresponding to the expected dimedone adduct (Figure S17). Addition of 4-chloro-7-nitrobenzofurazan (NBD-Cl), another sulfenate trap,75 gives a peak at 727.5 m/z, matching with the S-oxygenated NBD adduct, which is further confirmed with ATR-FTIR experiments (Figure S19).67,76 Control reactions show that these sulfenate adducts only form during the reaction of 1 with O2 and can be specifically labeled using 18O2 gas (see Supporting Information).

In conclusion, a rare sulfur-ligated Fe(O2) adduct can be obtained from either 1) FeII/O2 or 2) FeIII/O2•−. The spectroscopic and computational results for this species satisfyingly converge, leading to the conclusion that this complex is a mononuclear iron(III)-superoxide species. Warming of this metastable intermediate leads to the production of the S-oxygenated FeII(sulfinate) product, in direct analogy with the reactivity seen in the TDOs. A second intermediate containing a mono-oxygenated sulfenate moiety was also captured and, combined with the superoxide complex, provides the first full set of experimental data for a pathway 1 mechanism (Scheme 1), in either synthetic or enzymatic S-oxygenation. Future studies will focus on this and other mechanistic questions.

Supplementary Material

SI

ACKNOWLEDGMENTS

The NIH (R01GM119374 and R35GM149233 to D.P.G.) is gratefully acknowledged for financial support. We thank Dr. T. Albert for resonance Raman experiments. We also thank Dr. J.B. Gordon for useful discussions. Computer time was provided by the Maryland Advanced Research Computing Center (MARCC).

Footnotes

The authors declare no competing financial interest.

ASSOCIATED CONTENT

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c12337.

Mössbauer, EPR, ATR-FTIR and UV–vis spectra, mass spectrometry data, crystallographic information, DFT coordinates (PDF)

Accession Codes

CCDC 2294620–2294621 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Contributor Information

Sudha Yadav, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States.

Vishal Yadav, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States.

Maxime A. Siegler, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States

Pierre Moënne-Loccoz, Department of Chemical Physiology and Biochemistry, Oregon Health & Science University, Portland, Oregon 97239, United States.

Guy N. L. Jameson, School of Chemistry, Bio21 Molecular Science and Biotechnology Institute, The University of Melbourne, Parkville, Victoria 3010, Australia

David P. Goldberg, Department of Chemistry, The Johns Hopkins University, Baltimore, Maryland 21218, United States

REFERENCES

  • (1).Gordon JB; Goldberg DP Sulfur-Ligated, Oxidative Nonheme Iron Enzymes and Related Complexes. In Comprehensive Coordination Chemistry III, Elsevier: Oxford, 2021; pp 333–377. [Google Scholar]
  • (2).Sahu S; Goldberg DP Activation of Dioxygen by Iron and Manganese Complexes: A Heme and Nonheme Perspective. J. Am. Chem. Soc. 2016, 138, 11410–11428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Kal S; Que L Dioxygen activation by nonheme iron enzymes with the 2-His-1-carboxylate facial triad that generate high-valent oxoiron oxidants. J. Biol. Inorg. Chem. 2017, 22, 339–365. [DOI] [PubMed] [Google Scholar]
  • (4).Bollinger JM; Krebs C Enzymatic C-H activation by metal-superoxo intermediates. Curr. Opin. Chem. Biol. 2007, 11, 151–158. [DOI] [PubMed] [Google Scholar]
  • (5).Chen J; Jiang ZK; Fukuzumi S; Nam WW; Wang B Artificial nonheme iron and manganese oxygenases for enantioselective olefin epoxidation and alkane hydroxylation reactions. Coord. Chem. Rev. 2020, 421, 213443. [Google Scholar]
  • (6).Tchesnokov EP; Fellner M; Siakkou E; Kleffmann T; Martin LW; Aloi S; Lamont IL; Wilbanks SM; Jameson GNL The Cysteine Dioxygenase Homologue from Pseudomonas aeruginosa Is a 3 Mercaptopropionate Dioxygenase. J. Biol. Chem. 2015, 290, 24424–24437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Brandt U; Schürmann M; Steinbüchel A Mercaptosuccinate Dioxygenase, a Cysteine Dioxygenase Homologue, from Strain B4 Is the Key Enzyme of Mercaptosuccinate Degradation. J. Biol. Chem. 2014, 289, 30800–30809. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Weits DA; Giuntoli B; Kosmacz M; Parlanti S; Hubberten HM; Riegler H; Hoefgen R; Perata P; van Dongen JT; Licausi F Plant cysteine oxidases control the oxygen-dependent branch of the N-end-rule pathway. Nat. Commun. 2014, 5, 3425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).White MD; Kamps JJAG; East S; Taylor Kearney LJ; Flashman E The plant cysteine oxidases from Arabidopsis thaliana are kinetically tailored to act as oxygen sensors. J. Biol. Chem. 2018, 293, 11786–11795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Yamaguchi K; Hosokawa Y Cysteine Dioxygenase. Methods Enzymol. 1987, 143, 395–403. [DOI] [PubMed] [Google Scholar]
  • (11).Fraser-Pitt DJ; Mercer DK; Smith D; Kowalczuk A; Robertson J; Lovie E; Perenyi P; Cole M; Doumith M; Hill RLR; Hopkins KL; Woodford N; O’Neil DA Cysteamine, an Endogenous Aminothiol, and Cystamine, the Disulfide Product of Oxidation, Increase Sensitivity to Reactive Oxygen and Nitrogen Species and Potentiate Therapeutic Antibiotics against Bacterial Infection. Infect. Immun. 2018, 86, e00947–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).Sarkar B; Kulharia M; Mantha AK Understanding human thiol dioxygenase enzymes: structure to function, and biology to pathology. Int. J. Exp. Pathol. 2017, 98, 52–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Stipanuk MH; Ueki I; Dominy JE; Simmons CR; Hirschberger LL Cysteine dioxygenase: a robust system for regulation of cellular cysteine levels. Amino Acids 2009, 37, 55–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Dominy JE; Hirschberger LL; Coloso RM; Stipanuk MH Regulation of cysteine dioxygenase degradation is mediated by intracellular cysteine levels and the ubiquitin-26 S proteasome system in the living rat. Biochem. J. 2006, 394, 267–273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Gunawardana DM; Heathcote KC; Flashman E Emerging roles for thiol dioxygenases as oxygen sensors. Febs J. 2022, 289, 5426–5439. [DOI] [PubMed] [Google Scholar]
  • (16).Morrow WP; Sardar S; Thapa P; Hossain MS; Foss FW; Pierce BS Thiol dioxygenase turnover yields benzothiazole products from 2-mercaptoaniline and O-dependent oxidation of primary alcohols. Arch. Biochem. Biophys. 2017, 631, 66–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Jameson GNL Iron, cysteine and Parkinson’s disease. Monatsh. Chem. 2011, 142, 325–329. [Google Scholar]
  • (18).Heafield MT; Fearn S; Steventon GB; Waring RH; Williams AC; Sturman SG Plasma Cysteine and Sulfate Levels in Patients with Motor-Neuron, Parkinsons and Alzheimers-Disease. Neurosci. Lett. 1990, 110, 216–220. [DOI] [PubMed] [Google Scholar]
  • (19).Gordon C; Bradley H; Waring RH; Emery P Abnormal Sulfur Oxidation in Systemic Lupus-Erythematosus. Lancet 1992, 339, 25–26. [DOI] [PubMed] [Google Scholar]
  • (20).Perry TL; Norman MG; Yong VW; Whiting S; Crichton JU; Hansen S; Kish SJ Hallervorden-Spatz Disease - Cysteine Accumulation and Cysteine Dioxygenase Deficiency in the Globus Pallidus. Ann. Neurol. 1985, 18, 482–489. [DOI] [PubMed] [Google Scholar]
  • (21).Brait M; Ling SZ; Nagpal JK; Chang XF; Park HL; Lee J; Okamura J; Yamashita K; Sidransky D; Kim MS Cysteine Dioxygenase 1 Is a Tumor Suppressor Gene Silenced by Promoter Methylation in Multiple Human Cancers. PLoS One 2012, 7, e44951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Jeschke J; O’Hagan HM; Zhang W; Vatapalli R; Calmon MF; Danilova L; Nelkenbrecher C; Van Neste L; Bijsmans ITGW; Van Engeland M; Gabrielson E; Schuebel KE; Winterpacht A; Baylin SB; Herman JG; Ahuja N Frequent Inactivation of Contributes to Survival of Breast Cancer Cells and Resistance to Anthracyclines. Clin. Cancer Res. 2013, 19, 3201–3211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Ueno K; Kumagai T; Kijima T; Kishimoto T; Hosoe S Cloning and tissue expression of cDNAs from chromosome 5q21–22 which is frequently deleted in advanced lung cancer. Hum. Genet. 1998, 102, 63–68. [DOI] [PubMed] [Google Scholar]
  • (24).McCoy JG; Bailey LJ; Bitto E; Bingman CA; Aceti DJ; Fox BG; Phillips GN Structure and mechanism of mouse cysteine dioxygenase. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 3084–3089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Chai SC; Bruyere JR; Maroney MJ Probes of the catalytic site of cysteine dioxygenase. J. Biol. Chem. 2006, 281, 15774–15779. [DOI] [PubMed] [Google Scholar]
  • (26).Fernandez RL; Dillon SL; Stipanuk MH; Fox BG; Brunold TC Spectroscopic Investigation of Cysteamine Dioxygenase. Biochemistry 2020, 59, 2450–2458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Bruland N; Wübbeler JH; Steinbüchel A 3-Mercaptopropionate Dioxygenase, a Cysteine Dioxygenase Homologue, Catalyzes the Initial Step of 3-Mercaptopropionate Catabolism in the 3,3-Thiodipropionic Acid-degrading Bacterium. J. Biol. Chem. 2009, 284, 660–672. [DOI] [PubMed] [Google Scholar]
  • (28).Miller JR; Brunold TC Chapter Four - Spectroscopic analysis of the mammalian enzyme cysteine dioxygenase. In Methods Enzymol., Shukla AK, Ed.; Academic Press, 2023; Vol. 682, pp 101–135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (29).Forbes DL; Meneely KM; Chilton AS; Lamb AL; Ellis HR The 3-His Metal Coordination Site Promotes the Coupling of Oxygen Activation to Cysteine Oxidation in Cysteine Dioxygenase. Biochemistry 2020, 59, 2022–2031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Stampfli AR; Seebeck FP The catalytic mechanism of sulfoxide synthases. Curr. Opin. Chem. Biol. 2020, 59, 111–118. [DOI] [PubMed] [Google Scholar]
  • (31).Wang YJ; Yan LJ; Li XY; Zhang SQ; Wei JJ; Liu YJ Formation Mechanism of Cofactor Cys-Tyr in the Cysteine Dioxygenases (CDO and F-CDO) and Its Influence on Catalysis: A QM/MM Study. Inorg. Chem. 2021, 60, 7844–7856. [DOI] [PubMed] [Google Scholar]
  • (32).Fernandez RL; Brunold TC; Juntunen ND Differences in the Second Coordination Sphere Tailor the Substrate Specificity and Reactivity of Thiol Dioxygenases. Acc. Chem. Res. 2022, 55, 2480–2490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (33).Blaesi EJ; Fox BG; Brunold TC Spectroscopic and Computational Investigation of Iron(III) Cysteine Dioxygenase: Implications for the Nature of the Putative Superoxo-Fe(III) Intermediate. Biochemistry 2014, 53, 5759–5770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Kumar D; Thiel W; de Visser SP Theoretical Study on the Mechanism of the Oxygen Activation Process in Cysteine Dioxygenase Enzymes. J. Am. Chem. Soc. 2011, 133, 3869–3882. [DOI] [PubMed] [Google Scholar]
  • (35).Aluri S; de Visser SP The mechanism of cysteine oxygenation by cysteine dioxygenase enzymes. J. Am. Chem. Soc. 2007, 129, 14846. [DOI] [PubMed] [Google Scholar]
  • (36).Koehntop KD; Emerson JP; Que L The 2-His-1-carboxylate facial triad: a versatile platform for dioxygen activation by mononuclear non-heme iron(II) enzymes. J. Biol. Inorg. Chem. 2005, 10, 87–93. [DOI] [PubMed] [Google Scholar]
  • (37).Crawford JA; Li W; Pierce BS Single Turnover of Substrate-Bound Ferric Cysteine Dioxygenase with Superoxide Anion: Enzymatic Reactivation, Product Formation, and a Transient Intermediate. Biochemistry 2011, 50, 10241–10253. [DOI] [PubMed] [Google Scholar]
  • (38).Simmons CR; Krishnamoorthy K; Granett SL; Schuller DJ; Dominy JE; Begley TP; Stipanuk MH; Karplus PA A Putative Fe-Bound Persulfenate Intermediate in Cysteine Dioxygenase. Biochemistry 2008, 47, 11390–11392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (39).Souness RJ; Kleffmann T; Tchesnokov EP; Wilbanks SM; Jameson GB; Jameson GNL Mechanistic Implications of Persulfenate and Persulfide Binding in the Active Site of Cysteine Dioxygenase. Biochemistry 2013, 52, 7606–7617. [DOI] [PubMed] [Google Scholar]
  • (40).Tchesnokov EP; Faponle AS; Davies CG; Quesne MG; Turner R; Fellner M; Souness RJ; Wilbanks SM; de Visser SP; Jameson GNL An iron-oxygen intermediate formed during the catalytic cycle of cysteine dioxygenase. Chem. Commun. 2016, 52, 8814–8817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (41).Pan HR; Chen HJ; Wu ZH; Ge P; Ye SF; Lee GH; Hsu HF Structural and Spectroscopic Evidence for a Side-on Fe(III)- Superoxo Complex Featuring Discrete O-O Bond Distances. Jacs Au 2021, 1, 1389–1398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (42).Oddon F; Chiba Y; Nakazawa J; Ohta T; Ogura T; Hikichi S Characterization of Mononuclear Non-heme Iron(III)-Superoxo Complex with a Five-Azole Ligand Set. Angew. Chem., Int. Ed. 2015, 54, 7336–7339. [DOI] [PubMed] [Google Scholar]
  • (43).Hong S; Sutherlin KD; Park J; Kwon E; Siegler MA; Solomon EI; Nam W Crystallographic and spectroscopic characterization and reactivities of a mononuclear non-haem iron-(III)-superoxo complex. Nat. Commun. 2014, 5, 5440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (44).Müller L; Hoof S; Keck M; Herwig C; Limberg C Enhancing Tris(pyrazolyl)borate-Based Models of Cysteine/Cysteamine Dioxygenases through Steric Effects: Increased Reactivities, Full Product Characterization and Hints to Initial Superoxide Formation. Chem.—Eur. J. 2020, 26, 11851–11861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (45).Fischer AA; Lindeman SV; Fiedler AT A synthetic model of the nonheme iron-superoxo intermediate of cysteine dioxygenase. Chem. Commun. 2018, 54, 11344–11347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Blakely MN; Dedushko MA; Yan Poon PC; Villar-Acevedo G; Kovacs JA Formation of a Reactive, Alkyl Thiolate-Ligated Fe-Superoxo Intermediate Derived from Dioxygen. J. Am. Chem. Soc. 2019, 141, 1867–1870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Chiang CW; Kleespies ST; Stout HD; Meier KK; Li PY; Bominaar EL; Que L; Münck E; Lee WZ Characterization of a Paramagnetic Mononuclear Nonheme Iron-Superoxo Complex. J. Am. Chem. Soc. 2014, 136, 10846–10849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Winslow C; Lee HB; Field MJ; Teat SJ; Rittle J Structure and Reactivity of a High-Spin, Nonheme Iron(III)-Superoxo Complex Supported by Phosphinimide Ligands. J. Am. Chem. Soc. 2021, 143, 13686–13693. [DOI] [PubMed] [Google Scholar]
  • (49).Anandababu K; Ramasubramanian R; Wadepohl H; Comba P; Johnee Britto N; Jaccob M; Mayilmurugan R A Structural and Functional Model for the Tris-Histidine Motif in Cysteine Dioxygenase. Chem.—Eur. J. 2019, 25, 9540–9547. [DOI] [PubMed] [Google Scholar]
  • (50).Drewry JA; Fletcher S; Hassan H; Gunning PT Novel asymmetrically functionalized bis-dipicolylamine metal complexes: peripheral decoration of a potent anion recognition scaffold. Org. Biomol. Chem. 2009, 7, 5074–5077. [DOI] [PubMed] [Google Scholar]
  • (51).Gordon JB; McGale JP; Prendergast JR; Shirani-Sarmazeh Z; Siegler MA; Jameson GNL; Goldberg DP Structures, Spectroscopic Properties, and Dioxygen Reactivity of 5- and 6-Coordinate Nonheme Iron(II) Complexes: A Combined Enzyme/Model Study of Thiol Dioxygenases. J. Am. Chem. Soc. 2018, 140, 14807–14822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (52).Tamanaha E; Zhang B; Guo Y; Chang W.-c.; Barr EW; Xing G; St. Clair J; Ye S; Neese F; Bollinger, J. M.; Krebs C Spectroscopic Evidence for the Two C-H-Cleaving Intermediates of Isopenicillin Synthase. J. Am. Chem. Soc. 2016, 138, 8862–8874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (53).Mbughuni MM; Chakrabarti M; Hayden JA; Bominaar EL; Hendrich MP; Münck E; Lipscomb JD Trapping and spectroscopic characterization of an Fe-superoxo intermediate from a nonheme mononuclear iron-containing enzyme. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 16788–16793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Yang Y; Guan JQ; Qiu PP; Kan QB Synthesis, characterization and catalytic properties of heterogeneous iron(III) tetradentate Schiff base complexes for the aerobic epoxidation of styrene. Transition Met. Chem. 2010, 35, 263–270. [Google Scholar]
  • (55).Nakamoto K Applications in Inorganic Chemistry. In Infrared and Raman Spectra of Inorganic and Coordination Compounds, Wiley: New York, 2008; pp 149–354. [Google Scholar]
  • (56).Fischer AA; Lindeman SV; Fiedler AT Spectroscopic and computational studies of reversible O2 binding by a cobalt complex of relevance to cysteine dioxygenase. Dalton Trans. 2017, 46, 13229–13241. [DOI] [PubMed] [Google Scholar]
  • (57).Gordon JB; Vilbert AC; Siegler MA; Lancaster KM; Moënne-Loccoz P; Goldberg DP A Nonheme Thiolate-Ligated Cobalt Superoxo Complex: Synthesis and Spectroscopic Characterization, Computational Studies, and Hydrogen Atom Abstraction Reactivity. J. Am. Chem. Soc. 2019, 141, 3641–3653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Yadav V; Gordon JB; Siegler MA; Goldberg DP Dioxygen-Derived Nonheme Mononuclear Fe(OH) Complex and Its Reactivity with Carbon Radicals. J. Am. Chem. Soc. 2019, 141, 10148–10153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Yadav V; Rodriguez RJ; Siegler MA; Goldberg DP Determining the Inherent Selectivity for Carbon Radical Hydroxylation versus Halogenation with Fe(OH)(X)Complexes: Relevance to the Rebound Step in Non-heme Iron Halogenases. J. Am. Chem. Soc. 2020, 142, 7259–7264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (60).Yadav V; Siegler MA; Goldberg DP Temperature-Dependent Reactivity of a Non-heme Fe(OH)(SR) Complex: Relevance to Isopenicillin N Synthase. J. Am. Chem. Soc. 2021, 143, 46–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (61).Buonomo RM; Font I; Maguire MJ; Reibenspies JH; Tuntulani T; Darensbourg MY Study of Sulfinate and Sulfenate Complexes Derived from the Oxygenation of Thiolate Sulfur in [1,5-Bis(2-Mercapto-2 Methylpropyl)-1,5-Diazacyclooctanato(2-)]Nickel-(II). J. Am. Chem. Soc. 1995, 117, 963–973. [Google Scholar]
  • (62).Mullins CS; Grapperhaus CA; Frye BC; Wood LH; Hay AJ; Buchanan RM; Mashuta MS Synthesis and Sulfur Oxygenation of a (NS)Ni Complex Related to Nickel-Containing Superoxide Dismutase. Inorg. Chem. 2009, 48, 9974–9976. [DOI] [PubMed] [Google Scholar]
  • (63).McQuilken AC; Jiang YB; Siegler MA; Goldberg DP Addition of Dioxygen to an NS(thiolate) Iron(II) Cysteine Dioxygenase Model Gives a Structurally Characterized Sulfinato-Iron(II) Complex. J. Am. Chem. Soc. 2012, 134, 8758–8761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (64).Nguyen VH; Chew HQ; Su B; Yip JH Synthesis and spectroscopy of anionic cyclometalated iridium(III)-dithiolate and -sulfinates–effect of sulfur dioxygenation on electronic structure and luminescence. Inorg. Chem. 2014, 53, 9739–50. [DOI] [PubMed] [Google Scholar]
  • (65).Zhang D; Bin Y; Tallorin L; Tse F; Hernandez B; Mathias EV; Stewart T; Bau R; Selke M Sequential photooxidation of a Pt(II) (diimine)cysteamine complex: intermolecular oxygen atom transfer versus sulfinate formation. Inorg. Chem. 2013, 52, 1676–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (66).Farmer PJ; Solouki T; Mills DK; Soma T; Russell DH; Reibenspies JH; Darensbourg MY Isotopic Labeling Investigation of the Oxygenation of Nickel-Bound Thiolates by Molecular-Oxygen. J. Am. Chem. Soc. 1992, 114, 4601–4605. [Google Scholar]
  • (67).Lin-Vien D; Colthup NB; Fateley WG; Grasselli JG Chapter 14 - Organic Sulfur Compounds. In The Handbook of Infrared and Raman Characteristic Frequencies of Organic Molecules, Lin-Vien D; Colthup NB; Fateley WG; Grasselli JG, Eds.; Academic Press: San Diego, 1991; pp 225–250. [Google Scholar]
  • (68).McDonald AR; Bukowski MR; Farquhar ER; Jackson TA; Koehntop KD; Seo MS; De Hont RF; Stubna A; Halfen JA; Münck, E.; Nam, W.; Que, L. Sulfur versus Iron Oxidation in an Iron-Thiolate Model Complex. J. Am. Chem. Soc. 2010, 132, 17118–17129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (69).Vitzthum G; Lindner E Sulfinato Complexes. Angew. Chem., Int. Ed. 1971, 10, 315. [Google Scholar]
  • (70).Wu YR; Chang CM; Wang CC; Hsieh CC; Horng YC C = N Bond Activation and Hydration by an Iron(III) Complex with Asymmetric Sulfur Oxygenation. Eur. J. Inorg. Chem. 2017, 2017, 840–843. [Google Scholar]
  • (71).Villar-Acevedo G; Lugo-Mas P; Blakely MN; Rees JA; Ganas AS; Hanada EM; Kaminsky W; Kovacs JA Metal-Assisted Oxo Atom Addition to an Fe(III) Thiolate. J. Am. Chem. Soc. 2017, 139, 119–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (72).Heinrich L; Mary-Verla A; Li Y; Vaissermann J; Chottard JC Cobalt(III) complexes with carboxamido and sulfenato or sulfinato ligands suggest that a coordinated sulfenate is essential for the catalytic activity of nitrile hydratases. Eur. J. Inorg. Chem. 2001, 2001, 2203–2206. [Google Scholar]
  • (73).Yamanaka Y; Kato Y; Hashimoto K; Iida K; Nagasawa K; Nakayama H; Dohmae N; Noguchi K; Noguchi T; Yohda M; Odaka M Time-Resolved Crystallography of the Reaction Intermediate of Nitrile Hydratase: Revealing a Role for the Cysteine-sulfenic Acid Ligand as a Catalytic Nucleophile. Angew. Chem., Int. Ed. 2015, 54, 10763–10767. [DOI] [PubMed] [Google Scholar]
  • (74).Lombardini JB; Singer TP; Boyer PD Cysteine Oxygenase 0.2. Studies on Mechanism of Reaction with 18-Oxygen. J. Biol. Chem. 1969, 244, 1172. [PubMed] [Google Scholar]
  • (75).Lindhoud S; van den Berg WAM; van den Heuvel RHH; Heck AJR; van Mierlo CPM; van Berkel WJH Cofactor Binding Protects Flavodoxin against Oxidative Stress. PLoS One 2012, 7, e41363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (76).Korang J; Grither WR; McCulla RD Photodeoxygenation of Dibenzothiophene S-Oxide Derivatives in Aqueous Media. J. Am. Chem. Soc. 2010, 132, 4466–4476. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES