Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Oct 13.
Published in final edited form as: J Inorg Biochem. 2022 Jun 23;235:111907. doi: 10.1016/j.jinorgbio.2022.111907

Resonance Raman Spectroscopy of Pyranopterin Molybdenum Enzymes

Martin L Kirk a, Jesse Lepluart a, Jing Yang a
PMCID: PMC10575615  NIHMSID: NIHMS1931982  PMID: 35932756

Abstract

Resonance Raman spectroscopy (rR) is a powerful spectroscopic probe that is widely used for studying the geometric and electronic structure of metalloproteins. In this focused review, we detail how resonance Raman spectroscopy has contributed to a greater understanding of electronic structure, geometric structure, and the reaction mechanisms of pyranopterin molybdenum enzymes. The review focuses on the enzymes sulfite oxidase (SO), dimethyl sulfoxide reductase (DMSOR), xanthine oxidase (XO), and carbon monoxide dehydrogenase. Specifically, we highlight how Mo-Ooxo, Mo-Ssulfido, Mo-Sdithiolene, and dithiolene C=C vibrational modes, isotope and heavy atom perturbations, resonance enhancement, and associated Raman studies of small molecule analogs have provided detailed insight into the nature of these metalloenzyme active sites.

Keywords: Resonance Raman spectroscopy, pyranopterin molybdenum enzymes, xanthine oxidase, dimethyl sulfoxide reductase, sulfite oxidase, carbon monoxide dehydrogenase, molybdenum cofactor

1. Introduction and Scope

This contribution focuses on how resonance Raman spectroscopy has contributed to our current understanding of pyranopterin enzyme geometric structure, electronic structure, and mechanism over the past 30 years. The review begins with a brief background of the canonical pyranopterin Mo enzymes sulfite oxidase (SO), xanthine oxidase (XO), and dimethyl sulfoxide reductase (DMSOR), in addition to the pyranopterin dithiolene ligand, which is the organic component of the Mo cofactor (Moco) and common to all these enzymes. Where appropriate, examples from resonance Raman studies of small model analogs have been included to enhance our understanding of the enzyme spectroscopic data.

2. Mo Enzyme Background

Pyranopterin Mo enzymes catalyze a wide variety of redox transformations that are based on oxygen atom transfer, hydroxyl/oxyl radical transfer, hydride transfer, and hydroxylation reactions involving their substrate molecules.18 All these enzymes possess a unique molybdenum cofactor, or Moco, that is comprised of a Mo ion coordinated by the ene-1,2-ditholate linkage of the pyranopterin dithiolene (PDT, or MPT molybdopterin) ligand (Figure 1).9 A pyran ring connects the dithiolene chelate of the PDT with a pterin ring to yield an electronically complex ligand that has the potential to display a remarkable electronic flexibility.1015 Although a complete understanding of how the PDT functions in catalysis is still unknown, there is compelling evidence that the PDT functions to anchor Moco in the active sites of the enzymes, facilitate electron transfer reactivity, and modulate the reduction potential of the Mo site during the course of catalysis.10

Figure 1.

Figure 1.

Structure of Moco derived from X-ray crystallographic and XAS/EXAFS studies of the Moco insertase. Moco is shown coordinated by a single pyranopterin dithiolene with a closed pyran ring.

Pyranopterin Mo enzymes are categorized as belonging to either the sulfite oxidase (SO) family, the xanthine oxidase (XO) family, or the dimethylsulfoxide reductase (DMSOR) family based on the nature of their protein folds, active site coordination geometry, or the nature of the reactions they catalyze.3, 7, 16 The relationship between these enzyme families and the PDT is underscored by the observation that out-of-plane PDT distortions appear to be related to enzyme function.17 Determining how the geometric structure of the PDT correlates with the Mo oxidation state,18, 19 and how the PDT contributes to the unique active site electronic structures of the enzyme active sites remains a very active area of research.

SO family enzymes.

The active site coordination geometry of SO family enzymes has been probed by EXAFS20, 21 and X-ray crystallography2227 to reveal a square pyramidal [(PDT)MoO2(SRCys)]1- geometry for SOox where the Mo(VI) ion is coordinated by a single PDT dithiolene, two terminal oxo ligands oriented cis to one another, and a S donor from a cysteine residue (Figure 2). The structure of SOred also possesses a square pyramidal geometry. In the oxygen atom transfer mechanism (Figure 3), it is the water-derived equatorial oxo (Oeq) ligand that is transferred to the substrate. Oxygen atom transfer is believed to occur via an associative mechanism, with the apical oxo (Oap) functioning as a “spectator oxo”28 in catalysis.1, 16, 2935 The Mo=Oeq bond is oriented toward the substrate access channel, and the two-electron oxidation of the sulfite substrate is initiated by the attack of a sulfite S lone pair on Oeq to yield an initial E-P complex with sulfate bound to the reduced Mo(IV) center. This Mo-bound sulfate intermediate subsequently breaks down by product dissociation from the Mo ion and is followed by the binding of a water molecule to the reduced Mo(IV) site. The Mo(VI) form of the enzyme is regenerated from the reduced [(PDT)MoO(SRCys)(OH2)]1- Mo(IV) site by two sequential 1e/1H+ transfers.

Figure 2.

Figure 2.

The active site structure for Arabidopsis thaliana plant sulfite oxidase. PDB 1ogp.

Figure 3.

Figure 3.

The generally accepted oxygen atom transfer mechanism proposed for sulfite oxidase.

DMSOR family enzymes.

Pyranopterin Mo enzymes that belong to the DMSO reductase family are commonly divided into three classes (i.e. Type I, II, and III) based on the nature of the non-PDT ligands bound the Mo ion.6, 7, 16 Unlike the SO family enzymes, which possess a single PDT ligand bound to Mo, the DMSO reductase family enzymes possess two PDT ligands that remain coordinated to the metal throughout the course of catalysis. The enzymes that comprise the DMSO reductase family are very diverse in nature and catalyze a wide range of chemical transformations. DMSO reductase family enzymes that lack additional redox chromophores are found among the Type III DMSO reductases, making them highly amenable to optical spectroscopies that can intimately probe their electronic structures.36, 37 As such, resonance Raman studies have focused on the Type III enzymes from Rhodobacter. Type III DMSO reductases have been structurally characterized by a combination of EXAFS3842 and X-ray crystallography4346 and display a [(PDT)2MoO(OSer)]1-distorted trigonal prismatic active site geometry in the oxidized state with two coordinated PDTs, a terminal oxo, and a coordinated serinate oxygen (Figure 4). Reduced forms of this active site have been structurally characterized by EXAFS and display Mo(IV) des-oxo [(PDT)2Mo(OH2)(OSer)]1- and Mo(V) [(PDT)2Mo(OH)(OSer)]1- active site coordination spheres.47 The source of the terminal oxo ligand derives from the substrate in an oxygen atom transfer reaction (Figure 5) that has been addressed both experimentally and computationally.31, 36, 42, 4853 Here, catalysis is initiated in the reduced Mo(IV) state, with the substrate (DMSO) binding to Mo and replacing the labile water ligand. Oxygen atom transfer to Mo results in the two-electron oxidation of the metal to form product bound [(PDT)2Mo(O-DMS)(OSer)]1-, which is followed by product release that generates an oxidized [(PDT)2MoO(OSer)]1- site. The catalytically competent reduced form of the enzyme is generated by coupled e−/H+ transfer steps. The first of these converts [(PDT)2MoO(OSer)]1- to Mo(V) [(PDT)2Mo(OH)(OSer)]1-, which is also known as the high-g split intermediate due to its unique hydroxyl proton hyperfine interaction with 95,97Mo.36, 37, 42, 54 A second e−/H+ transfer step converts high-g split to the fully reduced des-oxo [(PDT)2Mo(OH2)(OSer)]1- state, with the loss of water generating five-coordinate [(PDT)2Mo(OSer)]1- as a putative resting state. A six-coordinate des-oxo [(PDT)2Mo(OSMe2)(OSer)]1- results upon the binding of substrate, leading to the next catalytic cycle.36

Figure 4.

Figure 4.

The active site structures for oxidized (DMSORox) and reduced (DMSORred) Rhodobacter sphaeroides DMSO reductase, each showing two PDTs coordinated to the Mo ion. Top: PDB 1eu1, Bottom: PDB 1h5n.

Figure 5.

Figure 5.

The generally accepted oxygen atom transfer mechanism proposed for dimethylsulfoxide reductase.

XO family enzymes

The XO family enzymes xanthine oxidase/dehydrogenase (XO/XDH) and aldehyde oxidase (AO) have been structurally characterized by EXAFS and X-ray crystallography (Figure 6), and they are found to adopt 5-coordinate square pyramidal coordination geometries in both their oxidized and reduced forms.5563 The oxidized forms of these enzymes possess a catalytically essential terminal sulfido ligand, which is converted to a sulfhydryl donor following substrate oxidation. Importantly, metal activated water present as a hydroxide donor is bound to the oxidized Mo(VI) form of these enzymes. This coordinated hydroxide ligand is directed toward the substrate access channel and is in close proximity to a catalytically essential glutamate residue, which serves as an active site base to promote Mo-OH base-assisted nucleophilic attack on a carbon atom of a wide variety of substrates that include purines, various N-heterocycles, and aldehydes. The mechanism of the oxidation reaction catalyzed by XO family enzymes is markedly different than that of typical Fe and Cu monooxygenases.64, 65 Namely, the mechanism is not radical-based, the enzymes generate rather than consume redox equivalents, and water rather than dioxygen is the source of the oxygen atom incorporated into product. The nature of the cis-[MoOS]2+ unit and its function in catalysis has been studied both spectroscopically66, 67 and by various computational approaches,4, 29, 6873 and this has contributed greatly to our understanding of the reaction mechanism of these enzymes. It is found that catalysis begins with base-assisted nucleophilic attack of the coordinated hydroxyl ligand on a substrate carbon center resulting in the formation of a tetrahedral intermediate and transition state that breaks down by formal hydride transfer to the terminal sulfido (Figure 7). For heterocyclic substrates, this yields a Mo(IV) E-P complex with the product bound to Mo as the enolate tautomer prior to product release. In the electron transfer half reaction, two sequential 1e/1H+ transfers regenerate the oxidized [(PDT)MoOS(OH)]1- form of the enzyme.

Figure 6.

Figure 6.

The active site structure for bovine milk xanthine dehydrogenase with bound FYX-051 inhibitor. PDB 1v97.

Figure 7.

Figure 7.

The generally accepted hydroxylation mechanism proposed for xanthine oxidase/dehydrogenase.

Another member of the XO enzyme family is the Mo-dependent carbon monoxide dehydrogenase,2, 5, 7479 which converts CO to CO2. Similar to XO/XDH and AO, this enzyme also adopts a square pyramidal coordination geometry with a coordinated PDT dithiolene, apical and equatorial oxo ligands, and a µ-sulfido ligand that bridges the Mo ion to a Cu(I) center that is also ligated to a cysteine thiolate (Figure 8).78, 79 Additionally, a water molecule may be weakly bound to the Cu site. Mechanistically, the Mo-dependent CODH catalyzes the oxidation of CO in a manner that is different than that of the Ni-Fe CODHs.4, 5, 74, 76, 78, 8082 One of the proposed mechanisms for Mo-CODH is based on a detailed DFT and natural bond orbital (NBO) study, and begins with CO binding to Cu(I), followed by nucleophilic attack of the Mo-oxo oxygen on the carbonyl carbon to yield the cyclic intermediate shown in Figure 9.4, 5 Metal activated water coordinated to the Cu ion is poised for nucleophilic attack on the µ22 CO2 carbon center to yield bicarbonate bound to Mo(IV), which subsequently breaks down to release CO2 as the ultimate product of the reaction. EPR spectroscopy using 13C labeled bicarbonate indicates that bicarbonate is not likely to be coordinated to the Mo(V) form of the enzyme.75 Alternative mechanisms have been suggested that involve either the direct release of CO2,7, 75, 76, 80 or the formation of a C-S bond along the reaction coordinate.74, 78, 82 The latter mechanism is based upon the crystal structure of an inhibited enzyme species.79

Figure 8.

Figure 8.

The active site structure for carbon monoxide dehydrogenase. PDB 1n63.

Figure 9.

Figure 9.

Top: A mechanism proposed for carbon monoxide dehydrogenase that does not include an intermediate with a C-S bond. Bottom: A proposed µ- thiocarbonate intermediate suggested from the X-ray structure of an inhibited enzyme species.

3. Resonance Raman Spectroscopy

Background.

Resonance Raman spectroscopy is an important probe of metalloenzyme geometric and electronic structure,83 and has played pivotal roles in increasing our understanding of pyranopterin containing molybdenum enzymes.1, 16, 84 In infrared spectroscopy, the selection rules are such that transitions are observed when there is a change in the dipole moment of a molecule as a function of the normal mode of vibration. Raman spectroscopy is complementary to infrared in that they both are used to probe molecular vibrations, but the Raman selection rules are such that vibrational transitions are observed when there is a change in the polarizability of the molecule as function of the normal mode of vibration. Key advantages of Raman over infrared spectroscopy for metalloproteins include its ease of use for aqueous samples, since water is a weak Raman scatterer, and the observed signal enhancement due to the resonance Raman effect. However, the interpretation of resonance Raman data can be complicated when the sample is heterogeneous, there are other strongly absorbing chromophores present in the sample, or the sample suffers from background fluorescence. Raman intensity enhancement derives from Franck-Condon scattering and Herzberg-Teller vibronic coupling.83, 85 In the case of Franck-Condon scattering (Albrecht A-term85) Raman enhancement derives from the displacement of an excited state potential energy surface relative to that of the ground state along a normal coordinate of vibration, Qk. Herzberg-Teller (Albrecht B-term85) Raman enhancement occurs when the resonant excited state can mix with another excited state by vibronic coupling. Typically, the Franck-Condon A-term is the dominant contributor to the scattering under resonant conditions, leading to a large resonance enhancement of totally symmetric normal modes that collectively describe the nature of the excited state distortion coordinate. Detailed analyses of resonance Raman spectra have traditionally utilized a Kramers-Heisenberg sum over states formalism,85 but time-dependent wavepacket approaches86 have increased in popularity and have recently been used to understand resonance Raman enhancement in xanthine oxidase.87

Model studies of Mo-S and Mo-oxo vibrations.

Perhaps the most characterized benchmark oxomolybdenum dithiolene complexes are Tp*MoO(bdt),88 Tp*MoO(tdt),88 and Tp*MoO(qdt)89 (Tp* = hydrotris-(3,5-dimethyl-1-pyrazolyl)borate; bdt = 1,2-benzenedithiolate; tdt = 3,4-toluenedithiolate; qdt = quinoxaline dithiolate). These paramagnetic d1 Mo(V) complexes are important for understanding the electronic structure of molybdoenzymes that possess a single terminal Mo≡O bond oriented cis to a dithiolene chelate, since this coordination geometry is encountered in many pyranopterin containing Mo enzyme forms.13, 6, 16, 19, 88, 90 The [MoO(dithiolene)]−1 unit in these molecules possesses idealized Cs symmetry, with the mirror plane bisecting the two S atoms of the dithiolene and containing the Mo≡O unit. From a resonance Raman spectroscopy perspective, the presence of a symmetry element simplifies the analysis of the spectra since there are a’ (symmetric) and a” (asymmetric) vibrational modes. Within this 4-atom core, there are 6 vibrational modes. These include the symmetric Mo≡O (νMo≡O) and S-Mo-S (νMo-S) stretching vibrations, the symmetric bend (δS-Mo-S), the asymmetric stretch (νMo-S), the symmetric wag (ω), and the asymmetric twist (τ). A combination of qualitative rR depolarization ratios88 and rR excitation profiles for Tp*MoO(bdt) showed that excitation into dithiolene → Mo(xy) LMCT transitions, which directly probe the redox orbital, resulted in resonance enhancement of two totally symmetric vibrations, the symmetric δS-Mo-S bend at 362 cm−1 and the νMo-S symmetric stretch at 393 cm−1 (S-Mo-S stretch), while excitation into dithiolene → Mo(xz,yz) LMCT transitions resulted in resonance enhancement of the symmetric νMo≡O stretch at 932 cm−1 (Figure 10). The resonance Raman results for Tp*MoO(qdt) were similar to those of Tp*MoO(bdt), displaying vibrational bands at 348 and 407 cm−1. However, a computational analysis of the normal modes showed that the S-Mo-S (νMo-S) stretch and bend (δS-Mo-S) vibrations were mixed, leading to a different normal mode description for the MoOS2 core.89 A key observation from these model studies is the small number of vibrational modes in the low-frequency (~200–1000 cm−1) region of the Raman spectra, which is in stark contrast to what is observed in many of the enzymes, where low-frequency Mo-S vibrations can be coupled with other low-frequency modes of the pyranopterin or even kinematically coupled with other protein modes. Considering bis-dithiolene complexes, the symmetrized Mo(QAd)(dithiolene)2 (Ad = 2-adamantyl; Q = O, S, Se) displayed only a single vibration at ~400 cm−1 in the 250 – 600 cm−1 region of the Raman spectrum that was assigned as the totally symmetric MoS4Mo-S) stretching vibration.91 The Raman spectra of the related MoO(SPh)4 and MoO(SPh-PhS)2 complexes were more complex, displaying five vibrations below 600 cm−1 in addition to the νMo≡O stretch. Mo-S modes were tentatively assigned at 366 cm−1 and 425 cm−1 for MoO(SPh)4 and 375 cm−1 and 423 cm−1 for MoO(SPh-PhS)2.92

Figure 10.

Figure 10.

Resonance Raman spectrum of Tp*MoO(bdt) and associated symmetry coordinates. Adapted from ref. 80.

4. Resonance Raman Spectroscopy of Sulfite Oxidase Family Enzymes

Resonance Raman spectra were initially collected on both the wt and C207S trioxo variant forms of the human SO Mo domain in the late 1990s.93, 94 To obtain Mo domain specific Raman spectra without the strong interference that originates from the highly absorbing cytochrome b5-type heme, the heme domain was removed by tryptic cleavage of the K108R and K108R/C207S mutants.95 This seminal study provided the first Mo-oxo stretching vibrations for a pyranopterin Mo enzyme. Specifically, two Mo-oxo stretching vibrations were observed in SOox at 903 cm−1 and 881 cm−1 that shifted to 890 cm−1 and 848 cm−1 with redox cycling in H218O buffer. The assignment of these bands as the symmetric νs (903 cm−1) and asymmetric νas (881 cm−1) O-Mo-O stretches was supported by prior studies on cis-dioxomolybdenum model compounds.96100 Based on a reduced mass argument, the observed isotopic shift was approximately half of that anticipated if both of the oxo groups were labeled by 18O, and this supports the conclusion that only one of the oxo groups is labile and capable of being labelled under these conditions. These two Mo-oxo vibrational modes were observed to be resonantly enhanced with excitation into the 480 nm charge transfer band that has been assigned as a SCys → Mo(VI) LMCT due to the absence of this absorption feature in the C207S variant.95 Additionally, low-frequency rR bands were observed at 289 cm−1 and 362 cm−1 with excitation into this LMCT band, and they were assigned as the coordinated cysteine Sγ-Cβ-Cα bending vibration and a Mo-S stretching vibration, respectively. High-frequency vibrations in resonance with the LMCT band were assigned to modes associated with the dithiolene chelate. These band were observed at 1006 cm−1 and 1161 cm−1 (coupled C-S and C-C stretches), and at 1532 cm−1 (C=C stretch).

Later, resonance Raman studies were performed on sulfite oxidase as part of a study aimed at understanding electronic structure contributions to equatorial Mo-oxo bond activation in sulfite oxidizing enzymes.35, 101 In addition to the previously assigned 20,833 cm−1 SCys → Mo LMCT band,93, 94, 101 the electronic absorption spectrum of oxidized wt plant sulfite oxidase (A. thaliana pSO) displays a higher energy band at 27,778 cm−1 that can be assigned as a Sdithiolene → Mo LMCT transition.101 The one-electron promotions associated with these LMCT bands originate from filled ligand-based orbitals and are promoted to the LUMO. Early DFT computations on a small [Mo(VI)O2(S2C2Me2)(SCH3)]- computational model for the SO0x active site showed the dramatic inequivalence of the apical (Oap) and equatorial (Oeq) oxo ligands due to a strong trans influence on Oeq.35 This observation is in stark contrast to more symmetric cis-dioxo MoO2 model complexes where the two oxo ligands are effectively related by a symmetry element.96100 This work also emphasized the critical importance of the SOox LUMO, since it possesses dominant Mo(xy)-Oeq π* antibonding character (52% Mo, 20% Oeq) with effectively no Oap contribution.35 The importance of the LUMO in the reductive half reaction of sulfite oxidizing enzymes stems from the fact that it is the two-electron acceptor orbital in oxygen atom transfer reactions with oxo acceptor substrates. Thus, the unique electronic structure of the site lowers the activation energy for Mo-Oeq bond cleavage along the reaction coordinate and promotes product release.35

Due to the Mo(xy)-Oeq π* character of the SO LUMO, LMCT excitations that result in population of the LUMO are expected to produce an excited state distortion (i.e. bond lengthening) along the Mo-Oeq bond. Raman spectroscopic studies on A. thaliana plant SO (pSO) yielded strong experimental support for this hypothesis,101 confirming the critically important Mo(xy)-Oeq π* antibonding interaction in the LUMO of SO0x suggested by the earlier computational study.35 Low temperature (30K) Raman spectra collected on resonance (488 nm) with the SCys → Mo LMCT band displayed prominent vibrational bands at 896, 877, and 864 cm−1 (Figure 11). A combination of DFT frequency computations on both large and small molecule active site models, combined with an 18O isotope perturbation, allowed for the higher frequency peak at 896 cm−1 to be assigned at the symmetric O-Mo-O stretch (νs) and the 864 cm−1 band to be assigned as the antisymmetric O-Mo-O stretch (νas). Both of these stretches possess noticeable mode localization due the low-symmetry of the oxidized active site. In a small computational model of the active site, the 877 cm−1 vibration was shown to possesses a dominant combination of Mo-Oeq and symmetric C-S dithiolene stretching character. In contrast, for cis-MoO2 complexes with symmetry related oxo donors, one observes strong resonance enhancement of the totally symmetric O-Mo-O stretch, but the antisymmetric Ooxo-Mo-Ooxo stretch possesses markedly reduced Raman intensity. The rationale for resonance enhancement of both the νs and νas Ooxo-Mo-Ooxo stretches in SO0x derives from the unique nature of the active site LUMO. As mentioned previously, one-electron promotions to the LUMO result in an excited state distortion only along the Mo-Oeq bond, which can be described as a linear combination of the νs and νas Mo-oxo stretches, and the observation of resonance enhancement for both modes. By analogy, the 864 cm−1 mode is also observed to be resonantly enhanced since there is Mo-Oeq stretching character in this vibration. Thus, the rR results are consistent with the 20,833 cm−1 band in SOox being assigned as a S → Mo(xy)-Oeq π* LMCT transition, contributing to a greater understanding of the SOox LUMO and Mo-Oeq bond activation along the reaction coordinate when this orbital becomes doubly occupied in the reductive half reaction.101

Figure 11.

Figure 11.

Resonance Raman spectrum of as-prepared A. thaliana plant SO. The spectrum is virtually identical the that of oxidized A. thaliana pSO redox cycled in H218O. Adapted from ref. 94.

Regarding electron transfer reactivity in the oxidative half reaction of SO, a combined cyclic voltammetry and surface enhanced resonance Raman (SERR) study of human sulfite oxidase (hSO) was used to probe heterogeneous electron transfer and catalysis in this enzyme, and it was shown that this activity was a function of the local protein environment.102 Specifically, electron transfer from the reduced Mo site to the electrode requires the Cyt b5 domain to change its orientation from one where it is interacting with the Mo domain to a new orientation where it can interact with the self-assembled monolayer (SAM) surface. The Mo and Cyt b5 domains are connected via a flexible polypeptide loop, which facilitates intraprotein electron transfer between these redox centers when the heme domain interacts with the surface of the Mo domain.103 Interprotein electron transfer is facilitated by docking of the heme domain with its electron transfer partner Cyt c. In this study, where the heme domain interacts with the electrode, a very fast heterogeneous electron transfer rate (ks = 440 s−1) was measured that displayed a dependence on ionic strength. Fast electron transfer was determined to be a result of a shorter contact time with the SAM surface at higher ionic strength.102

5. Resonance Raman Spectroscopy of DMSO Reductase Family Enzymes

DMSO reductase family enzymes can be divided into three basic types based on their structure and the nature of the non-PDT ligands that are bound to the Mo ion.1, 6, 7, 16 However, only the type III DMSOR family enzymes, which possess a general [MoO(PDT)2(OR)]1- (OR = serinate, aspartate) oxidized enzyme active site structure, have been probed in detail by resonance Raman spectroscopy. These Type III enzymes generally lack the strongly absorbing flavin, heme, and Fe-S cluster redox chromophores that are present in many other enzymes. These redox chromophores obscure the weaker electronic absorption features of the Mo active site and effectively render resonance Raman interrogation of the active site difficult or impossible. R. sphaeroides DMSOR is a typical type III enzyme and possesses distinct low energy charge transfer features at 720 nm for DMSORox and 640 nm for DMSORred.37, 93, 94 These isolated absorption bands conveniently allow for Raman spectra to be collected on-resonance with these low-energy transitions using the 676.4 nm line of a Kr+ laser. Initial studies by Spiro and coworkers104 showed that excitation into these low energy charge transfer bands of DMSORox produced rR enhancement of vibrations in the PDT dithiolene C=C stretching region centered at 1575 cm−1. They showed that the C=C stretching frequency is downshifted by ~ 7 cm−1 to 1568 cm−1 for DMSORred, consistent with a change in Mo-dithiolene covalency as a function of the Mo ion oxidation state. The dithiolene C=C stretching vibration is expected to be observed in the 1300 – 1600 cm−1 region of the Raman spectrum100, 105 with frequencies that are anticipated to vary as a function of the oxidation state of the dithiolene,12, 13, 106, 107 the oxidation state of the metal, and the magnitude of the “sulfur-fold” distortion within the five-membered chelate ring.18, 89, 106 In 1995 Spiro and coworkers noted that the 1568 cm−1 band that had previously been observed for DMSORred, and assigned as a C=C stretching vibration, was likely an experimental artifact.108 This latter study assigned the 1576 cm−1 mode observed for DMSORox as a C=C stretch from a 5,8-dihydro form of the pterin component of the PDT, with a 1526 cm−1 band being assigned as the PDT dithiolene C=C stretch.

R. sphaeroides DMSOR isolated from bacteria grown on 34S was used to understand the nature of low-frequency vibrational modes associated with the dithiolene component of the PDT.104 Two 34S sensitive bands were observed for DMSORox at 350 cm−1 and 370 cm−1, which shift to 352 cm−1 and 383 cm−1 in DMSORred. These bands were assigned as arising from the symmetric and asymmetric S-Mo-S stretches of the dithiolene chelate. Additional low-frequency Raman bands were also observed to be affected by the 34S isotope perturbation. This work was important, since the observation of both C=C and Mo-S stretching vibrations in resonance with a low-energy CT transition were used to confirm that the dithiolene component of the PDT was directly coordinated in a bidentate fashion to the Mo ion. 34S labeling of the enzyme indicated that, in addition to Mo-S stretches, other vibrational modes in the 250 – 420 cm−1 region are present that couple with the Mo-S stretching vibrations leading to their resonance enhancement.108 Resonance Raman studies of both pterin and quinoxoline containing model compounds show isotope sensitive bands that shift by 7–8 cm−1, supporting the idea of extensive low-frequency vibrational mode mixing in DMSOR.108 Terminal Mo-oxo stretches are typically observed between 850 – 1000 cm−1.88, 108 Although weak rR bands had been observed in this spectral region for R. sphaeroides DMSOR, these early studies indicated that redox cycling in the presence of 18OH2 did not result any frequency shifts associated with the isotopic labeling.108

Later, a more complete resonance Raman study of four R. spaeroides DMSOR enzyme forms was performed that allowed for all of the vibrational modes observed in the 200 – 1700 cm−1 region to be assigned as originating from Mo≡O, dithiolene, Mo-OSer, Mo-S, and DMSO vibrational modes or non-resonantly enhanced modes of the protein (Figure 12).93, 109 This study, conducted after the initial crystal structure was published for DMSO reductase,46 probed DMSORox, DMSORred, DMS reduced DMSORred, and a glycerol inhibited Mo(V) form that is structurally similar to the high-g split paramagnetic intermediate that has been studied extensively by MCD, electronic absorption, and EPR spectroscopies,36 in addition to EXAFS.42, 54 An 18O isotopic perturbation study on the Mo(VI) form of the enzyme was used to assign the νMo≡O stretch at 862 cm−1 and show that DMSOR0x is a mono-oxo species, consistent with oxygen atom transfer from the substrate to a des-oxo Mo(IV) species. The vibrational frequencies and vibrational mode assignments for the four different enzyme forms have been summarized.109 DMS reduced DMSOR was shown to possess a bound DMSO product molecule, as evidenced by νMo-O and νS=O stretches observed at 497 cm−1 and 862 cm−1 respectively, again consistent with an associative oxygen atom transfer mechanism.36, 109 The X-ray structures of DMSORs at that time yielded markedly different descriptions of the first coordination sphere ligation to the Mo ion. The structure of the R. sphaeroides enzyme46 indicated a mono-oxo Mo(VI) species with one of the PDT dithiolenes being asymmetrically coordinated to the metal with a long Mo-Sdithiolene bond length of 3.1Å. The reduced des-oxo R. sphaeroides structure revealed partial dissociation of one of the PDT dithiolene ligands. The DMSORox structure from R. capsulatus indicated a di-oxo Mo(VI) site with one of the dithiolene ligands being completely de-coordinated from Mo.44 Thus, the resonance Raman work109 is of considerable importance since it definitively showed that all four of the Mo-Sdithiolene bonds remained coordinated in the different enzyme forms studied, the Mo(VI) enzyme form was a mono-oxo species, and the Mo(IV) form was a des-oxo species clearly supporting an oxygen atom transfer mechanism.

Figure 12.

Figure 12.

Resonance Raman spectrum of R. sphaeroides DMSOR. Adapted with permission from Garton, S. D.; Hilton, J.; Oku, H.; Crouse, B. R.; Rajagopalan, K. V.; Johnson, M. K., Active Site Structures and Catalytic Mechanism of Rhodobacter sphaeroides Dimethyl Sulfoxide Reductase as Revealed by Resonance Raman Spectroscopy. J. Am. Chem. Soc. 1997, 119 (52), 12906–12916.

Biotin sulfoxide reductase (BSOR) is another DMSO reductase family enzyme that is closely related to the R. capsulatus and R. sphaeroides DMSORs. Resonance Raman data for BSOR have been collected for as-prepared DMSORox, product-associated DMSORox, and dithionite-reduced DMSORred.110 The Mo-S stretching vibrations (200 – 600 cm−1) for R. sphaeroides biotin sulfoxide reductase have been analyzed in the context of an idealized square pyramidal C4v site. Resonance Raman spectra for as-prepared BSORox, using excitation wavelengths between 457–647 nm allowed for assignment of the totally symmetric Mo-S4 breathing mode at 355 cm−1. This is the dominant spectral feature in the Mo-S stretching region and can be compared with the corresponding vibrational mode in DMSORox at 350 cm−1. Similarly, for reduced BSORred and DMSORred these modes were assigned at 363 cm−1 and 367 cm−1, respectively. The observation of strong resonance enhancement for these breathing modes with laser excitation across the visible and NIR spectral region was used to suggest the optical bands in this region could be assigned as dithiolene→Mo(VI) LMCT transitions, in agreement with spectral assignments for DMSOR.93, 109

A comparison of the observed and assigned Mo-S stretching vibrations for DMSOR and BSOR clearly show that active sites of these prokaryotic oxotransferase enzymes are indeed very similar. Additional vibrational bands at 288, 416, and 447 cm−1 were not observed to shift significantly as a function of oxidation state or dithiolene 34S isotopic substitution and have therefore been assigned as originating from deformations within the PDT dithiolene and pyran rings. Similar assignments were previously made for DMSOR.109 Raman bands at 463 cm−1 and 543 cm−1 were suggested as potential Mo-OSer stretching modes in an as-prepared form of the enzyme (i.e. BSORox), and Raman bands at 425 cm−1 and 495 cm−1 were suggested for the Mo-OSer stretch in BSORred. For comparison, the Mo-OSer stretching vibration was tentatively assigned at 513 cm−1 for DMSORred and at 536 cm−1 for DMSORox.109 Serine coordination is of current interest in DMSOR and related Type III enzymes since this ligand has been suggested to be labile. Current evidence for the coordination of serine in the Mo(V) state derives from a combined electronic absorption, MCD, EPR, and computational study of the high-g split catalytic intermediate.42 Obtaining direct evidence from EXAFS for a Mo(V)-OSer scatterer is more problematic, but the recent model studies suggest that serine does remain coordinated throughout the catalytic cycle.42

The Mo≡O stretch for BSOR0x is observed at 860 cm−1 and this band shifts to 818 cm−1 after redox cycling in H218O buffer, as predicted using the reduced mass for a diatomic Mo≡O oscillator.110 This is consistent with a mono-oxo Mo(VI) site in BSOR0x. As expected, the mono-oxo Mo≡O stretching vibration is not observed in reduced enzyme, in full accord with a des-oxo Mo(IV) site. Dithionite reduced BSOR that was subsequently oxidized using 16O DMSO results in a mono-oxo Mo site that possesses peaks at 840 cm−1 and 860 cm−1 that were assigned as Mo≡O stretching vibrations, with the former being more intense than the latter using 568 nm excitation. The observation of two Mo≡O stretching vibrations was used to support an argument for active site heterogeneity that results from interactions between the protein (e.g. Trp-90 in BSOR) and the terminal oxo ligand, indicating a degree of active site conformational flexibility.110 Although redox cycling did not seem to affect this active site heterogeneity, the observation of the Mo≡O stretching band sharpening during catalytic cycling showed that this heterogeneity could be reduced to yield an active site with less conformational flexibility. Due to the difficulty in synthesizing 18O labeled BSO, Me2S18O was used to label BSOR under turnover conditions. An elegant set of 16O/18O labeling experiments showed that during catalytic cycling with the non-physiological Me2SO substrate, the terminal oxo ligand in the Mo(VI) state can exchange with both water and the oxygen atom of Me2SO. Furthermore, analysis of the BSOR Mo≡O stretch allowed for a description of how the enzyme interacts with various product molecules that appear to be weakly bound to the Mo(VI) ion in the active site (Figure 13).

Figure 13.

Figure 13.

R. sphaeroides BSOR structures derived from resonance Raman spectroscopy. Adapted from ref. 106.

Raman spectroscopy is a very sensitive probe of the nature of the dithiolene that is bound to the Mo ion in both the proteins and models. Namely, the ν(C-S) and ν(C=C) modes are found to be inversely correlated,84, 105, 110, 111 with high ν(C=C) and low ν(C-S) frequencies indicative of dithiolate character and low ν(C=C) and high ν(C-S) frequencies signifying a greater degree of dithione character in the chelate ring.110 For BSOR, vibrations in the ν(C=C) stretching region were used to propose that the enzyme possesses two different types of PDT dithiolene ligands, as was also noted for as-prepared wild-type DMSOR.109 Similarly, the resonance Raman spectra of the DMSOR family enzyme arsenite oxidase was interpreted in the context of one PDT dithiolene possessing a dithiolate structure with the second having more π-delocalization.112 Model studies of oxo-Mo(IV) dithiolene complexes display evidence of thiol/thione character in a single dithiolene,12, 13 complicating the issue of ν(C-S) and ν(C=C) vibrational assignments and highlighting aspects of the non-innocent nature of the dithiolene when bound to Mo.1, 10, 1214, 19, 90, 107

Raman data for Type III DMSOR family enzymes can be compared with the Raman spectra of good structural models (e.g. [MoVIO(OSiPr3)(S2C2(COOMe)2)2]-) for the DMSOR0x site, and these studies113 suggest an alternative proposal to the original idea that DMSO reductase possesses one PDT dithiolene that has dithiolate character with the second being more π-delocalized.109 The MoVIO(OSiPr3)(S2C2(COOMe)2)2]- model possess terminal oxo ligation, a serine mimic (i.e. OSiPr3), and two coordinated true olefinic dithiolene chelates.114 Remarkably, these complexes also are good electronic structural models for the DMSORox site since the electronic absorption spectrum of [MoVIO(OSiPr3)(S2C2(COOMe)2)2]- possesses low energy charge transfer bands at 13 533 cm−1 (ε=1350 M−1 cm−1) and 17 549 cm−1 (ε=2800 M−1 cm−1), that are very similar in intensity and transition energy to those observed for DMSORox.93, 109 The resonance Raman spectrum of [MoVIO(OSiPr3)(S2C2(COOMe)2)2]1- shows dithiolene C=C stretches at 1554 and 1489 cm−1,114 which are slightly lower than the frequencies observed for R. sphaeroides DMSO reductase (1578 and 1527 cm–1).93 Vibrational frequency computations performed on [MoVIO(OSiPr3)(S2C2(COOMe)2)2]1- show dithiolene C=C stretches at 1577 and 1505 cm−1, which were assigned as arising from each dithiolene ligand (Figure 14). The two dithiolenes are not symmetry related, with one of the dithiolenes (A) oriented such that both S donors are in the equatorial plane of the complex, while the second dithiolene (B) is oriented such that only one S donor is in the equatorial plane with the second oriented trans to the apical oxo ligand. This geometry results in a unique Mo-S π*-bonding interaction114, 115 in the model system that leads to a LUMO that possesses ~18% S character from the (B) dithiolene. This same Mo-S π*-bonding interaction can be inferred to occur in DMSORs and BSOR due to the similarity of their electronic absorption and resonance Raman spectra with those of the model. The implication of this type of covalent bonding interaction in DMSO reductase suggests that the PDT with the lower frequency dithiolene C=C stretch likely possesses a dithiolene with a single S donor in the equatorial plane and an ~180° Ooxo-Mo-Sdithiolene-C dihedral angle. Thus, this dithiolene could function to provide a hole superexchange pathway for electron transfer regeneration of the reduced catalytically competent DMSORred site involving a π-type bonding scheme similar to that observed in blue copper proteins.116, 117 Although there is still much to be learned regarding the nature of the PDT in DMSOR family enzymes, the resonance Raman spectra for DMSOR, BSOR, and model systems support a high degree of Mo coordination sphere structural similarity between these two enzymes and a similar oxygen atom transfer reaction mechanism with the Mo ion cycling between the Mo(IV) and Mo(VI) oxidation states.

Figure 14.

Figure 14.

Resonance Raman spectrum of the DMSOR/BSOR model compounds [MoVIO(OSiBuPh2)(S2C2(COOMe)2)2]1- (a) and [MoVIO(OSiPr3)(S2C2(COOMe)2)2]1- (b). Adapted with permission from Sugimoto, H.; Tatemoto, S.; Suyama, K.; Miyake, H.; Mtei, R. P.; Itoh, S.; Kirk, M. L., Monooxomolybdenum(VI) Complexes Possessing Olefinic Dithiolene Ligands: Probing Mo-S Covalency Contributions to Electron Transfer in Dimethyl Sulfoxide Reductase Family Molybdoenzymes. Inorg. Chem. 2010, 49 (12), 5368–5370.

6. Raman of Xanthine oxidase family enzymes

In the early 1980s, Palmer and coworkers,118, 119 used XO to catalyze the oxidation of lumazine to violapterin and form a stable charge transfer complex that possesses an intense long wavelength absorption feature at 650nm (15,385 cm−1).120 Initial resonance Raman experiments on XO were obtained by optically exciting into the long-wavelength absorption of the enzyme complexed with violopterin.120 This work was important since resonance Raman studies on XO family enzymes are made difficult by the presence of FAD and 2Fe2S ferredoxin clusters. These redox chromophores conspire to mask the Mo active site absorbance due to their greater extinction coefficients and contributions to background fluorescence. The low energy absorption band in the charge transfer complex was originally assigned as a Mo→violapterin charge transfer transition, with violapterin being oriented cis to the Mo≡O bond. This assignment was based on the νMo≡O stretch not being resonantly enhanced with excitation into this band. Although the νMo≡O stretch was not resonantly enhanced, numerous violapterin vibrational modes were observed to be resonantly enhanced using 676.4 nm excitation. Additionally, vibrational bands between 250–1100 cm−1 were also observed and hypothesized to originate from the Mo coordination sphere. A 16,18O isotope perturbation was used to tentatively assign the 276, 517, and 853 cm−1 bands as originating from the Mo-O-R unit of the coordinated violapterin. In subsequent work, Hille and coworkers again used 676.4 nm laser excitation to excite into this long wavelength charge transfer band and observed strong resonance enhancement of Raman vibrations deriving from violopterin product modes.121 A combination of 16,18O and 1,2H isotope perturbations, coupled with DFT frequency calculations, were used to make extensive vibrational frequency assignments in the 350–1750 cm−1 region of the spectra. These data convincingly showed that the violopterin product was bound to Mo in an end-on geometry, with the oxygen atom of the Mo-O-Rviolopterin bridging unit being introduced by direct hydroxylation of the lumazine substrate. Due to the large number of vibrational modes that possess a bridging oxygen contribution, a side-on binding mode of product was eliminated as a possibility.

This work was followed by additional resonance Raman spectroscopic studies by Kirk and coworkers87, 122124 who employed the reducing substrates 4-thiolumazine and 2,4-dithiolumazine in place of lumazine. Enzymatic turnover of these thiolumazine substrates generated new Mo(IV)-product complexes (Mo(IV)-4TV and Mo(IV)-2,4TV) that possessed a single intense long wavelength absorption feature in the near-infrared region of the spectrum (758–778 nm) using both bovine xanthine oxidase (XO) and Rhodobacter capsulatus xanthine dehydrogenase (XDH). Using these heavy thiolumazine congeners of lumazine, the MLCT band was shifted to markedly lower energy relative to Mo(IV)-product complexes that used lumazine as the reducing substrate.121, 125 This allowed for the acquisition of high quality resonance Raman spectra using 780nm excitation that were completely devoid of contributions from the 2Fe2S and FAD chromophores, and the ability to separate low-frequency violapterin product-based vibrational modes from those of the enzyme active site. The 200–600 cm−1 Raman spectra for XO and XDH were virtually identical and displayed a combination of Mo-PDT modes and in-plane bending modes of the bound product. Additional electronic absorption and resonance Raman studies were performed using 4-thiolumazine and 2,4-dithiolumazine on wt XDH from R. capsulatus and the Q197A and Q102G variants.124 Q197A and Q102 potentially hydrogen bond to the terminal oxygen of the Mo≡O group and the amino terminus of the PDT, respectively. It was observed that both the energy and the band shape of the Mo(IV) → product charge transfer transition are virtually unchanged as a function of the Q197A and Q102G variants, providing evidence that these residues do not dramatically affect the bonding in these Mo(IV)-thiolumazine product complexes. Moco specific vibrational modes were observed for Mo(IV)-4TV at 234, 290, 326, and 351 cm−1 (Bands A-D) that were also observed in Mo(IV)-2,4TV) at 236, 286, 326, and 351 cm−1 (Figure 15). DFT frequency computations were used to assist in the frequency assignments, with Band A being assigned as arising from a dithiolene envelope fold with Mo≡O rocking and pyranopterin contributions, and Band B being assigned as an asymmetric dithiolene ring distortion with Mo-SH stretching and pyranopterin contributions. Band C possessed a dominant symmetric S-Mo-S dithiolene core stretching and bending component with some Mo≡O rocking and pyranopterin contributions, while Band D was described as having a combination of S-Mo-S dithiolene and Mo-SH stretching contributions. These low frequency Moco vibrations derive from the nature of the Mo(IV) → product excited state distortion, which is driven by the fact that this charge transfer excited state possesses a large degree of hole character on the Mo ion. The observation of these resonantly enhanced Moco modes provide evidence that the PDT serves as an electron transfer conduit in the oxidative half-reactions of these enzymes and that the PDT is quite sensitive to redox changes at the Mo center.88, 123, 124

Figure 15.

Figure 15.

Low frequency resonance Raman spectra of wt XDH from R. capsulatus and the Q197A and Q102G variants. Reproduced with permission from Dong, C.; Yang, J.; Reschke, S.; Leimkühler, S.; Kirk, M. L., Vibrational Probes of Molybdenum Cofactor–Protein Interactions in Xanthine Dehydrogenase. Inorg. Chem. 2017, 56 (12), 6830–6837.

A combination of bonding and spectroscopic calculations have been used to understand the electronic origin of resonantly enhanced, high-frequency, in-plane product stretching vibrations in bovine Mo(IV)-4TV and Mo(IV)-2,4TV enzyme-product complexes (Figure 16).87 Here, the low-energy absorption band was assigned as a Mo(xy) → product π* MLCT transition involving a one-electron promotion to the thiolumazine LUMO. The intensity of this MLCT band, coupled with the observation of strong resonance enhancement for in-plane 4TV and 2,4TV modes, was used to show that the apical Mo≡O bond must be oriented in the same plane as the product molecules. This allows for the doubly occupied Mo(xy) redox orbital to be involved in a π-bonding interaction with the π-LUMO of the product molecule. The resonance Raman enhancement of these thiolumazine modes results from excited state distortions within the Mo–O–Cproduct linkage and are therefore of key mechanistic importance since they can be correlated with a covalent pathway for electron flow between Mo and substrate.87 More specifically, the tetrahedral intermediate and transition state57, 64, 72, 73, 126 possesses a Mo-Oeq-Cproduct linkage, which provides a direct π-orbital pathway for the two-electron oxidation of the thiolumazine and other heterocyclic substrates. This work formed the basis for later spectroscopic and computational studies on XO that showed specific electron delocalizations in the cis-[MoOS]2+ unit and the Mo-Oeq-Csubstrate linkage make substantial contributions to C-H bond activation and transition state stabilization along the reaction coordinate.127

Figure 16.

Figure 16.

High frequency resonance Raman spectrum of bovine XO Mo(IV)-4-TV. Adapted from ref. 79.

Resonance Raman studies on both sulfo and desulfo forms of bovine XOox were used to assign metal-ligand vibrations in the oxidized enzyme.128 Raman data collected using 400 – 650 nm excitation allowed identification of the Mo(VI)=S stretch at 474 cm−1 in sulfo XOox, which was downshifted by 12 cm−1 in the 34S sulfo form. By comparison, the Mo(VI)≡O stretch was observed as a weak band at 899 cm−1 in sulfo XOox. Raman excitation profiles indicated that both the Mo(VI)=S and Mo(VI)≡O stretches were maximally resonantly enhanced at 400 nm, and this was used to suggest that the 425 nm absorption feature, obtained by spectral subtraction of the desulfo electronic absorption spectrum from the sulfo spectrum, was a Ssulfido→ Mo LMCT band. However, it was acknowledged that the 2Fe2S centers and FAD dominantly contribute to the electronic absorption spectrum in the 360 – 470 nm region, making this charge transfer assignment tentative. Using 568.2 nm excitation, low frequency vibrations observed at 334, 348, 365, and 408 cm−1 were suggested as possible Mo-PDT vibrational modes or Fe-S vibrations arising from the 2Fe2S clusters. These Mo ≡ Ooxo and Mo=S vibrational assignments can be compared with the model compound TpiPrMoVIOS(OPh),67 which possesses an apical oxo ligand oriented cis to the terminal sulfido donor in the equatorial plane as observed in XO0x. The solution Raman spectrum of TpiPrMoVIOS(OPh) revealed the Mo ≡Ooxo stretch at 912 cm−1 and the Mo=S stretch at 486 cm−1. It is important to note that νMoS is markedly less than that observed for Tp*MoSCl2 (525 cm−1),129 which possesses a formally triple bonded terminal sulfido ligand. Thus, there is a significant reduction of the Mo-S bond order in the cis-[MoOS] unit due to the presence of the strong field terminal oxo ligand.67

Resonance Raman Spectroscopy of CODH.

Resonance Raman spectra of CODHox from Oligotropha carboxidovorans were collected using 514.5 nm excitation and displayed vibrational modes that can be associated with the Mo active site, FAD (1300–1400 cm−1) and Fe/S clusters (300–350 cm−1) (Figure 17).76 Vibrational modes that may be assigned as originating from the [(PDT)MoO2-SCu(SCys)]2- bimetallic site are observed at 895, 879, and 861 cm−1, in remarkable agreement with the Mo-Ooxo stretching vibrations observed for A. thaliana sulfite oxidase.101 Using this spectral comparison between enzymes that possess di-oxo Mo(VI) centers, the 895 cm−1 vibration in CODH can be assigned as the symmetric O-Mo-O stretch (νs) and the 861 cm−1 vibration as the asymmetric O-Mo-O stretch (νas). Since the intensity of the assigned asymmetric O-Mo-O stretch is greater than that of νs, it is possible that the low symmetry of the active site, coupled with mode localization and an excited state distortion predominantly along the equatorial Mo-Oeq bond, are responsible for the nature of the unusual resonance enhancement pattern. Also, by analogy to plant sulfite oxidase,101 the 879 cm−1 vibration may be assigned as a PDT dithiolene C-S stretch that also possesses some Mo-Oeq stretching character. The observation of multiple Mo-Ooxo stretching vibrations in the Raman spectrum of CODH is consistent with a [(PDT)MoO2-SCu(SCys)]2- dioxo site for CODH0x, and supports the mechanistic hypothesis of nucleophilic attack by the equatorial oxo ligand on the C≡O carbon atom of the substrate bound to Cu(I).5, 76, 130

Figure 17.

Figure 17.

Resonance Raman spectrum of O. carboxidovorans compared to the Raman spectra of FAD, desulfo-XO, and SO. Reproduced from ref. 68.

Conclusions

Our understanding of pyranopterin Mo enzymes has benefitted greatly from the resonance Raman spectroscopic studies detailed here. The highest information content is obtained with the use of isotope or heavy atom congener perturbations, the construction of resonance Raman excitation profiles, parallel studies on accurate structural models, and DFT frequency computations. Critical pyranopterin dithiolene Mo-S vibrational modes are now available for direct comparison between Type III DMSOR family enzymes and xanthine oxidases, indicating differences in Mo-S bonding between these two different enzymes. Raman spectroscopy has increased our current understanding of Mo-Ooxo stretching modes in the low symmetry environment of SO family sulfite oxidizing enzymes, and shows that the equatorial Mo-Ooxo bond that is directed toward the substrate access channel is specifically activated for oxygen atom transfer reactivity. The observation of Mo-product vibrational modes in DMSOR and XO have contributed greatly to our understanding of oxygen atom transfer reactivity and the unique hydroxylation mechanism posited for xanthine oxidase family enzymes. New interpretations of older Raman data have allowed for new insight into molybdoenzyme electronic structure, and this has improved our understanding of electronic structure contributions to reactivity. We expect new resonance Raman spectroscopic studies of pyranopterin Mo enzymes to further increase our understanding of enzymes beyond the canonical SO, XO, and DMSOR family enzymes that have been studied to date.

ACKNOWLEDGMENT

M. L. K. (R01-GM-057378) acknowledge the NIH for continued financial support of work performed by his group that is detailed in this manuscript. The authors thank the UNM Center for Advanced Research Computing, supported in part by the NSF, for providing high performance computing resources in support of their work detailed in this manuscript.

Table of Abbreviations

rR

resonance Raman

SO

sulfite oxidase

DMSOR

dimethyl sulfoxide reductase

XO

xanthine oxidase

PDT

pyranopterin dithiolene

MPT

molybdopterin

EXAFS

extended X-ray absorption fine structure

AO

aldehyde oxidase

E-P

enzyme-product

XDH

xanthine dehydrogenase

DFT

density functional theory

NBO

natural bond order

CODH

carbon monoxide dehydrogenase

Qk

normal coordinate vibration along mode q

bdt

1,2-benzenedithiolate

tdt

3,4-toluenedithiolate

qdt

quinoxaline dithiolate

Tp*

hydrotris-(3,5-dimethyl-1-pyrazolyl)borate

LMCT

ligand-to-metal charge transfer

LUMO

lowest unoccupied molecular orbital

SERR

surface enhanced resonance Raman

SAM

self-assembled monolayer

MCD

magnetic circular dichroism

EPR

electron paramagnetic resonance

BSOR

biotin sulfoxide reductase

DMS

dimethyl sulfide

NIR

near infrared

BSO

biotin sulfoxide

FAD

flavin adenine dinucleotide

MLCT

metal-to-ligand charge transfer

References

  • 1.Kirk ML, Spectroscopic and Electronic Structure Studies of Mo Model Compounds and Enzymes. The Royal Society of Chemistry: Cambridge, UK, 2017; p 13–67. [Google Scholar]
  • 2.Ingersol LJ; Kirk ML, Structure, Function, and Mechanism of Pyranopterin Molybdenum and Tungsten Enzymes. In Comprehensive Coordination Chemistry III, Constable EC; Parkin G; Que L Jr, Eds. Elsevier: Oxford, 2021; pp 790–811. [Google Scholar]
  • 3.Kirk ML; Kc K, Molybdenum and Tungsten Cofactors and the Reactions They Catalyze Transition Metals and Sulfur – A Strong Relationship for Life. In Metal Ions in Life Sciences, Sosa Torres M; Kroneck P, Eds. De Gruyter: Berlin, 2020; Vol. 20, pp 313–342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Electronic structure contributions to reactivity in xanthine oxidase family enzymes. Stein BW; Kirk ML, Journal of Biological Inorganic Chemistry 2015, 20, (2), 183–194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Orbital contributions to CO oxidation in Mo-Cu carbon monoxide dehydrogenase. Stein BW; Kirk ML, Chem. Commun 2014, 50, 1104–1106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Kirk ML; Stein B, The Molybdenum Enzymes. In Comprehensive Inorganic Chemistry II (Second Edition), Editors-in-Chief: Jan R; Kenneth P, Eds. Elsevier: Amsterdam, 2013; pp 263–293. [Google Scholar]
  • 7.The Mononuclear Molybdenum Enzymes. Hille R; Hall J; Basu P, Chemical Reviews 2014, 114, (7), 3963–4038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Crawford AM; Cotelesage JJH; Prince RC; George GN, The Catalytic Mechanisms of the Molybdenum and Tungsten Enzymes. In Metallocofactors That Activate Small Molecules: With Focus on Bioinorganic Chemistry, Ribbe MW, Ed. 2019; Vol. 179, pp 63–100. [Google Scholar]
  • 9.Mechanism of molybdate insertion into pterin-based molybdenum cofactors. Probst C; Yang J; Krausze J; Hercher TW; Richers CP; Spatzal T; Khadanand KC; Giles LJ; Rees DC; Mendel RR; Kirk ML; Kruse T, Nature Chemistry 2021, 13, (8), 758–765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Burgmayer SJN; Kirk ML, The Role of the Pyranopterin Dithiolene Component of Moco in Molybdoenzyme Catalysis. In Metallocofactors that Activate Small Molecules: With Focus on Bioinorganic Chemistry, Ribbe MW, Ed. Springer International Publishing: Cham, 2019; Vol. 179, pp 101–151. [Google Scholar]
  • 11.Synthesis, characterization, and spectroscopy of model molybdopterin complexes. Burgmayer SJN; Kim M; Petit R; Rothkopf A; Kim A; BelHamdounia S; Hou Y; Somogyi A; Habel-Rodriguez D; Williams A; Kirk ML, J. Inorg. Biochem 2007, 101, (11–12), 1601–1616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Noninnocent Dithiolene Ligands: A New Oxomolybdenum Complex Possessing a Donor Acceptor Dithiolene Ligand. Matz KG; Mtei RP; Leung B; Burgmayer SJN; Kirk ML, J. Am. Chem. Soc 2010, 132, (23), 7830–7831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Study of Molybdenum(4+) Quinoxalyldithiolenes as Models for the Noninnocent Pyranopterin in the Molybdenum Cofactor. Matz KG; Mtei RP; Rothstein R; Kirk ML; Burgmayer SJN, Inorg. Chem 2011, 50, (20), 9804–9815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Implications of Pyran Cyclization and Pterin Conformation on Oxidized Forms of the Molybdenum Cofactor. Gisewhite DR; Yang J; Williams BR; Esmail A; Stein B; Kirk ML; Burgmayer SJN, J. Am. Chem. Soc 2018, 140, (40), 12808–12818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Pyranopterin conformation defines the function of molybdenum and tungsten enzymes. Rothery RA; Stein B; Solomonson M; Kirk ML; Weiner JH, Proc. Natl. Acad. Sci. U. S. A 2012, 109, 14773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Hille R; Schulzke C; Kirk ML, Molybdenum and Tungsten Enzymes. The Royal Society of Chemistry: Cambridge, UK, 2017. [Google Scholar]
  • 17.Pyranopterin conformation defines the function of molybdenum and tungsten enzymes. Rothery RA; Stein B; Solomonson M; Kirk ML; Weiner JH, Proc. Natl. Acad. Sci. U. S. A 2012, 109, (37), 14773–14778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Vibrational Control of Covalency Effects Related to the Active Sites of Molybdenum Enzymes. Stein BW; Yang J; Mtei R; Wiebelhaus NJ; Kersi DK; LePluart J; Lichtenberger DL; Enemark JH; Kirk ML, J. Am. Chem. Soc 2018, 140, (44), 14777–14788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Metal-Dithiolene Bonding Contributions to Pyranopterin Molybdenum Enzyme Reactivity. Yang J; Enemark JH; Kirk ML, Inorganics 2020, 8, (3). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.High-Resolution EXAFS of the Active Site of Human Sulfite Oxidase: Comparison with Density Functional Theory and X-ray Crystallographic Results. Harris HH; George GN; Rajagopalan KV, Inorg. Chem 2006, 45, 493–495. [DOI] [PubMed] [Google Scholar]
  • 21.Coordination chemistry at the molybdenum site of sulfite oxidase: Redox-induced structural changes in the cysteine 207 to serine mutant. George GN; Garrett RM; Prince RC; Rajagopalan KV, Inorg. Chem 2004, 43, (26), 8456–8460. [DOI] [PubMed] [Google Scholar]
  • 22.Kinetic and structural evidence for the importance of Tyr236 for the integrity of the Mo active site in a bacterial sulfite dehydrogenase. Kappler U; Bailey S; Feng CJ; Honeychurch MJ; Hanson GR; Bernhardt PV; Tollin G; Enemark JH, Biochemistry 2006, 45, (32), 9696–9705. [DOI] [PubMed] [Google Scholar]
  • 23.Molecular basis of intramolecular electron transfer in sulfite-oxidizing enzymes is revealed by high resolution structure of a heterodimeric complex of the catalytic molybdopterin subunit and a c-type cytochrome subunit. Kappler U; Bailey S, J. Biol. Chem 2005, 280, (26), 24999–25007. [DOI] [PubMed] [Google Scholar]
  • 24.Molecular Basis of Sulfite Oxidase Deficiency from the Structure of Sulfite Oxidase. Kisker C; Schindelin H; Pacheco A; Wehbi WA; Garrett RM; Rajagopalan KV; Enemark JH; Rees DC, Cell 1997, 91, (7), 973–983. [DOI] [PubMed] [Google Scholar]
  • 25.The molybdenum-cofactor: a crystallographic perspective. Schindelin H; Kisker C; Rees DC, Journal of Biological Inorganic Chemistry 1997, 2, 773–781. [Google Scholar]
  • 26.Structural Insights into Sulfite Oxidase Deficiency. Karakas E; Wilson HL; Graf TN; Xiang S; Jaramillo-Busquets S; Rajagopalan KV; Kisker C, J. Biol. Chem 2005, 280, (39), 33506–33515. [DOI] [PubMed] [Google Scholar]
  • 27.The crystal structure of plant sulfite oxidase provides insights into sulfite oxidation in plants and animals. Schrader N; Fischer K; Theis K; Mendel RR; Schwarz G; Kisker C, Structure 2003, 11, (10), 1251. [DOI] [PubMed] [Google Scholar]
  • 28.Bivalent spectator oxo bonds in metathesis and epoxidation alkenes. Rappe AK; Goddard WA, Nature 1980, 285, (5763), 311. [Google Scholar]
  • 29.Theoretical studies on the reactivity of molybdenum enzymes. Metz S; Thiel W, Coord. Chem. Rev 2011, 255, (9–10), 1085–1103. [Google Scholar]
  • 30.A theoretical study of the primary oxo transfer reaction of a dioxo molybdenum(VI) compound with imine thiolate chelating ligands: A molybdenum oxotransferase analogue. Thomson LM; Hall MB, J. Am. Chem. Soc 2001, 123, (17), 3995–4002. [DOI] [PubMed] [Google Scholar]
  • 31.Ryde U; Dong G; Li J; Feldt M; Mata RA, Computational Studies of Molybdenum and Tungsten Enzymes. In Molybdenum and Tungsten Enzymes: Spectroscopic and Theoretical Investigations, The Royal Society of Chemistry: 2017; pp 275–321.
  • 32.QM/MM study of the reaction mechanism of sulfite oxidase. Caldararu O; Feldt M; Cioloboc D; van Severen MC; Starke K; Mata RA; Nordlander E; Ryde U, Scientific Reports 2018, 8. [DOI] [PMC free article] [PubMed]
  • 33.Theoretical studies of molybdenum oxo-transfer enzymes. Ryde U; Li J; van Severen M; Nordlander E, Journal of Biological Inorganic Chemistry 2014, 19, S80–S80. [Google Scholar]
  • 34.A quantum-mechanical study of the reaction mechanism of sulfite oxidase. van Severen MC; Andrejic M; Li JL; Starke K; Mata RA; Nordlander E; Ryde U, Journal of Biological Inorganic Chemistry 2014, 19, (7), 1165–1179. [DOI] [PubMed] [Google Scholar]
  • 35.Active-site stereochemical control of oxygen atom transfer reactivity in sulfite oxidase. Peariso K; McNaughton RL; Kirk ML, J. Am. Chem. Soc 2002, 124, (31), 9006–9007. [DOI] [PubMed] [Google Scholar]
  • 36.Spectroscopic and Electronic Structure Studies of a Dimethyl Sulfoxide Reductase Catalytic Intermediate: Implications for Electron- and Atom-Transfer Reactivity. Mtei RP; Lyashenko G; Stein B; Rubie N; Hille R; Kirk ML, J. Am. Chem. Soc 2011, 133, (25), 9762–9774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Mechanistic studies of Rhodobacter sphaeroides Me2SO reductase. Cobb N; Conrads T; Hille R, J. Biol. Chem 2005, 280, (12), 11007. [DOI] [PubMed] [Google Scholar]
  • 38.X-ray absorption spectroscopy of a quantitatively Mo(V) dimethyl sulfoxide reductase species Pushie MJ; Cotelesage JJ; Lyashenko G; Hille R; George GN, Inorg. Chem 2013, 52, 2830. [DOI] [PubMed] [Google Scholar]
  • 39.Structure of the Molybdenum Site of Dimethyl Sulfoxide Reductase. George GN; Hilton J; Temple C; Prince RC; Rajagopalan KV, J. Am. Chem. Soc 1999, 121, 1256. [Google Scholar]
  • 40.Interaction of Product Analogues with the Active Site of Rhodobacter Sphaeroides Dimethyl Sulfoxide Reductase. George GN; Nelson KJ; Harris HH; Doonan CJ; Rajagopalan KV, Inorg. Chem 2007, 46, (8), 3097–3104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.X-Ray Absorption Spectroscopy of Dimethyl Sulfoxide Reductase from Rhodobacter sphaeroides. George GN; Hilton J; Rajagopalan KV, J. Am. Chem. Soc 1996, 118, (5), 1113–1117. [Google Scholar]
  • 42.Addressing Serine Lability in a Paramagnetic Dimethyl Sulfoxide Reductase Catalytic Intermediate. Kc K; Yang J; Kirk ML, Inorg. Chem 2021, 60, (13), 9233–9237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.The 1.3 Å Crystal Structure of Rhodobacter sphaeroides Dimethyl Sulfoxide Reductase Reveals Two Distinct Molybdenum Coordination Environments. Li HL; Temple C; Rajagopalan KV; Schindelin H, J. Am. Chem. Soc 2000, 122, 7673. [Google Scholar]
  • 44.Crystal structure of dimethyl sulfoxide reducatase from Rhodobacter capsulatus at 1.88 Å resolution. Schneider F; Löwe J; Huber R; Schindelin H; Kisker C; Knäblein J, J. Mol. Biol 1996, 263, 53–69. [DOI] [PubMed] [Google Scholar]
  • 45.Dimethylsulfoxide Reductase: An Enzyme Capable of Catalysis with Either Molybdenum or Tungsten at the Active Site. Stewart L; Bailey S; Bennett B; Charnock J; Garner C; McAlpine A, J. Biol. Chem 2000, 299, 593. [DOI] [PubMed] [Google Scholar]
  • 46.Crystal structure of DMSO reductase: redox-linked changes in molybdopterin coordination. Schindelin H; Kisker C; Hilton J; Rajagopalan KV; Rees DC, Science 1996, 272, (5268), 1615–1621. [DOI] [PubMed] [Google Scholar]
  • 47.Structure of the Molybdenum Site of Dimethyl Sulfoxide Reductase. George GN; Hilton J; Temple C; Prince RC; Rajagopalan KV, J. Am. Chem. Soc 1999, 121, 1256–1266. [Google Scholar]
  • 48.Large Density-Functional and Basis-Set Effects for the DMSO Reductase Catalyzed Oxo-Transfer Reaction. Li JL; Mata RA; Ryde U, Journal of Chemical Theory and Computation 2013, 9, (3), 1799–1807. [DOI] [PubMed] [Google Scholar]
  • 49.A Computational Comparison of Oxygen Atom Transfer Catalyzed by Dimethyl Sulfoxide Reductase with Mo and W. Li JL; Andrejic M; Mata RA; Ryde U, European Journal of Inorganic Chemistry 2015, (21), 3580–3589. [Google Scholar]
  • 50.Effect of the protein ligand in DMSO reductase studied by computational methods. Dong G; Ryde U, J. Inorg. Biochem 2017, 171, 45. [DOI] [PubMed] [Google Scholar]
  • 51.Substrate and Metal Control of Barrier Heights for Oxo Transfer to Mo and W Bis-dithiolene Sites. Tenderholt AL; Hodgson KO; Hedman B; Holm RH; Solomon EI, Inorg. Chem 2012, 51, (6), 3436–3442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Sulfur K-Edge X-ray Absorption Spectroscopy and Density Functional Calculations on Mo(IV) and Mo(VI)=O Bis-dithiolenes: Insights into the Mechanism of Oxo Transfer in DMSO Reductase and Related Functional Analogues. Tenderholt AL; Wang JJ; Szilagyi RK; Holm RH; Hodgson KO; Hedman B; Solomon EI, J. Am. Chem. Soc 2010, 132, (24), 8359–8371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.A kinetic study of dimethyl sulfoxide reductase based on density functional theory. Hernandez-Marin E; Ziegler T, Canadian Journal of Chemistry-Revue Canadienne De Chimie 2010, 88, (8), 683–693. [Google Scholar]
  • 54.X-ray Absorption Spectroscopy of a Quantitatively Mo(V) Dimethyl Sulfoxide Reductase Species. Pushie MJ; Cotelesage JJH; Lyashenko G; Hille R; George GN, Inorg. Chem 2013, 52, (6), 2830–2837. [DOI] [PubMed] [Google Scholar]
  • 55.The crystal structure of xanthine oxidoreductase during catalysis: Implications for reaction mechanism and enzyme inhibition. Okamoto K; Matsumoto K; Hille R; Eger BT; Pai EF; Nishino T, Proc. Nat. Acad. Sci. USA 2004, 101, (21), 7931–7936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Crystal structures of the active and alloxanthine-inhibited forms of xanthine dehydrogenase from Rhodobacter capsulatus. Truglio J; Theis K; Leimkuhler S; Rappa R; Rajagopalan K; Kisker C, Structure 2002, 10, (1), 115–125. [DOI] [PubMed] [Google Scholar]
  • 57.Substrate Orientation and Catalysis at the Molybdenum Site in Xanthine Oxidase Crystal Structures in Complex with Xanthine and Lumazine Pauff JM; Cao H; Hille R, J. Biol. Chem 2009, 284, (13), 8751–8758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Substrate orientation in xanthine oxidase - Crystal structure of enzyme in reaction with 2-hydroxy-6-methylpurine. Pauff JM; Zhang JJ; Bell CE; Hille R, J. Biol. Chem 2008, 283, (8), 4818–4824. [DOI] [PubMed] [Google Scholar]
  • 59.Crystal structures of bovine milk xanthine dehydrogenase and xanthine oxidase: Structure-based mechanism of conversion. Enroth C; Eger BT; Okamoto K; Nishino T; Nishino T; Pai EF, Proc. Natl. Acad. Sci. U. S. A 2000, 97, 10723–10728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.A structure-based catalytic mechanism for the xanthine oxidase family of molybdenum enzymes. Huber R; Hof P; Duarte R; Moura J; Moura I; Liu M; LeGall J; Hille R; Archer M; Romao M, Proc. Natl. Acad. Sci. U. S. A 1996, 93, (17), 8846–8851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.X-ray crystal structure and EPR spectra of "arsenite-inhibited" Desulfovibrio gigas aldehyde dehydrogenase: A member of the xanthine oxidase family. Boer DR; Thapper A; Brondino CD; Romao MJ; Moura JJG, J. Am. Chem. Soc 2004, 126, (28), 8614. [DOI] [PubMed] [Google Scholar]
  • 62.Crystal Structure of the Xanthine Oxidase Related Aldehyde Oxidoreductase from D. gigas. Romao MJ; Archer M; Moura I; Moura JJG; Legall J; Engh R; Schneider M; Hof P; Huber R, Science 1995, 270, (5239), 1170–1176. [DOI] [PubMed] [Google Scholar]
  • 63.Molybdenum and tungsten enzymes: a crystallographic and mechanistic overview. Romao MJ, Dalton Transactions 2009, (21), 4053–4068. [DOI] [PubMed] [Google Scholar]
  • 64.Molybdenum-containing hydroxylases. Hille R, Arch. Biochem. Biophys 2005, 433, (1), 107–116. [DOI] [PubMed] [Google Scholar]
  • 65.High-valent iron in chemical and biological oxidations. Groves JT, J. Inorg. Biochem 2006, 100, (4), 434–447. [DOI] [PubMed] [Google Scholar]
  • 66.Molybdenum(V) Sites in Xanthine Oxidase and Relevant Analogue Complexes: Comparison of Molybdenum-95 and Sulfur-33 Hyperfine Coupling. Wilson GL; Greenwood RJ; Pilbrow JR; Spence JT; Wedd AG, J. Am. Chem. Soc 1991, 113, 6803–6812. [Google Scholar]
  • 67.Electronic structure description of the cis-MoOS unit in models for molybdenum hydroxylases. Doonan CJ; Rubie ND; Peariso K; Harris HH; Knottenbelt SZ; George GN; Young CG; Kirk ML, J. Am. Chem. Soc 2008, 130, (1), 55–65. [DOI] [PubMed] [Google Scholar]
  • 68.Reductive half-reaction of aldehyde oxidoreductase toward acetaldehyde: Ab initio and free energy quantum mechanical/molecular mechanical calculations. Dieterich JM; Werner HJ; Mata RA; Metz S; Thiel W, J. Chem. Phys 2010, 132, (3), 035101. [DOI] [PubMed] [Google Scholar]
  • 69.Substrate oxidation in the active site of xanthine oxidase and related enzymes. A model density functional study. Voityuk A; Albert K; Romao M; Huber R; Rosch N, Inorg. Chem 1998, 37, (2), 176–180. [Google Scholar]
  • 70.QM/MM Studies of Xanthine Oxidase: Variations of Cofactor, Substrate, and Active-Site Glu802. Metz S; Thiel W, J. Phys. Chem. B 2010, 114, (3), 1506–1517. [DOI] [PubMed] [Google Scholar]
  • 71.Reductive Half-Reaction of Aldehyde Oxidoreductase toward Acetaldehyde: A Combined QM/MM Study. Metz S; Wang DQ; Thiel W, J. Am. Chem. Soc 2009, 131, (13), 4628–4640. [DOI] [PubMed] [Google Scholar]
  • 72.A Combined QM/MM Study on the Reductive Half-Reaction of Xanthine Oxidase: Substrate Orientation and Mechanism. Metz S; Thiel W, J. Am. Chem. Soc 2009, 131, (41), 14885–14902. [DOI] [PubMed] [Google Scholar]
  • 73.Kirk ML, Knottenbelt S, Habtegabre A, The Reaction Coordinate of Pyranopterin Molybdenum Enzymes. In Computational Inorganic and Bioinorganic Chemistry, Solomon EI, Scott RA, King BR, Ed. Wiley: 2009; pp 277–286. [Google Scholar]
  • 74.Revisiting the catalytic mechanism of Mo-Cu carbon monoxide dehydrogenase using QM/MM and DFT calculations. Xu K; Hirao H, Physical Chemistry Chemical Physics 2018, 20, (28), 18938–18948. [DOI] [PubMed] [Google Scholar]
  • 75.Studies of carbon monoxide dehydrogenase from Oligotropha carboxidovorans. Dingwall S; Wilcoxen J; Niks D; Hille R, Journal of Molecular Catalysis B-Enzymatic 2016, 134, 317–322. [Google Scholar]
  • 76.Kinetic and Spectroscopic Studies of the Molybdenum-Copper CO Dehydrogenase from Oligotropha carboxidovorans. Zhang B; Hemann CF; Hille R, J. Biol. Chem 2010, 285, (17), 12571–12578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Paramagnetic Active Site Models for the Molybdenum−Copper Carbon Monoxide Dehydrogenase. Gourlay C; Nielsen DJ; White JM; Knottenbelt SZ; Kirk ML; Young CG, J. Am. Chem. Soc 2006, 128, (7), 2164–2165. [DOI] [PubMed] [Google Scholar]
  • 78.A novel binuclear [CuSMo] cluster at the active site of carbon monoxide dehydrogenase: Characterization by X-ray absorption spectroscopy. Gnida M; Ferner R; Gremer L; Meyer O; Meyer-Klaucke W, Biochemistry 2003, 42, (1), 222. [DOI] [PubMed] [Google Scholar]
  • 79.Catalysis at a dinuclear [CuSMo (=O)OH] cluster in a CO dehydrogenase resolved at 1.1-Å resolution. Dobbek H; Gremer L; Kiefersauer R; Huber R; Meyer O, Proceedings of the National Academy of Sciences 2002, 99, (25), 15971–15976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.A W/Cu Synthetic Model for the Mo/Cu Cofactor of Aerobic CODH Indicates That Biochemical CO Oxidation Requires a Frustrated Lewis Acid/Base Pair. Ghosh D; Sinhababu S; Santarsiero BD; Mankad NP, J. Am. Chem. Soc 2020, 142, (29), 12635–12642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.A realistic in silico model for structure/function studies of molybdenum-copper CO dehydrogenase. Rokhsana D; Large TAG; Dienst MC; Retegan M; Neese F, Journal of Biological Inorganic Chemistry 2016, 21, (4), 491–499. [DOI] [PubMed] [Google Scholar]
  • 82.Quantum chemical modeling of CO oxidation by the active site of Molybdenum CO Dehydrogenase. Siegbahn PEM; Shestakov AF, J. Comput. Chem 2005, 26, (9), 888–898. [DOI] [PubMed] [Google Scholar]
  • 83.Czernuszewicz RS; Spiro TG, Raman IR, and Resonance Raman Spectroscopy. In Inorganic Electronic Structure and Spectroscopy, Solomon EI; Lever ABP, Eds. John Wiley and Sons, Inc.: New York, 1999; Vol. I, pp 353–442. [Google Scholar]
  • 84.Johnson MK, Vibrational Spectra of Dithiolene Complexes. In Prog. Inorg. Chem, Stiefel EI, Ed. John Wiley and Sons, Inc.: Hoboken, New Jersey, 2004; Vol. 52, pp 213–266. [Google Scholar]
  • 85.On the Theory of Raman Intensities. Albrecht AC, J. Chem. Phys 1961, 34, (5), 1476–1484. [Google Scholar]
  • 86.The Semiclassical Way to Molecular Spectroscopy. Heller EJ, Acc. Chem. Res 1981, 14, (12), 368–375. [Google Scholar]
  • 87.Xanthine oxidase-product complexes probe the importance of substrate/product orientation along the reaction coordinate. Yang J; Dong C; Kirk ML, Dalton Transactions 2017, 46, (39), 13242–13250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Spectroscopic evidence for a unique bonding interaction in oxo-molybdenum dithiolate complexes: Implications for sigma electron transfer pathways in the pyranopterin dithiolate centers of enzymes. Inscore FE; McNaughton R; Westcott BL; Helton ME; Jones R; Dhawan IK; Enemark JH; Kirk ML, Inorg. Chem 1999, 38, (7), 1401–1410. [Google Scholar]
  • 89.Understanding the origin of metal-sulfur vibrations in an oxo-molybdenurn dithiolene complex: Relevance to sulfite oxidase. Inscore FE; Knottenbelt SZ; Rubie ND; Joshi HK; Kirk ML; Enemark JH, Inorg. Chem 2006, 45, (3), 967. [DOI] [PubMed] [Google Scholar]
  • 90.Kirk ML; McNaughton RL; Helton ME, The Electronic Structure and Spectroscopy of Metallo-Dithiolene Complexes. In Progress in Inorganic Chemistry: Synthesis, Properties, and Applications, Stiefel EI; Karlin KD, Eds. John Wiley and Sons: Hoboken, New Jersey, 2004; Vol. 52, pp 111–212. [Google Scholar]
  • 91.Spectroscopic and electronic structure studies of symmetrized models for reduced members of the dimethylsulfoxide reductase enzyme family. McNaughton RL; Lim BS; Knottenbelt SZ; Holm RH; Kirk ML, J. Am. Chem. Soc 2008, 130, (14), 4628–4636. [DOI] [PubMed] [Google Scholar]
  • 92.Electronic structure studies of oxomolybdenum tetrathiolate complexes: Origin of reduction potential differences and relationship to cysteine-molybdenum bonding in sulfite oxidase. McNaughton RL; Tipton AA; Rubie ND; Conry RR; Kirk ML, Inorg. Chem 2000, 39, (25), 5697–5706. [DOI] [PubMed] [Google Scholar]
  • 93.Resonance Raman as a Direct Probe for the Catalytic Mechanism of Molybdenum Oxotransferases. Johnson MK; Garton SD; Oku H, J. Biol. Inorg. Chem 1997, 2, (6), 797–803. [Google Scholar]
  • 94.Resonance Raman Characterization of the Molybdenum Center in Sulfite Oxidase: Identification of Mo=O Stretching Modes. Garton SD; Garrett RM; Rajagopalan KV; Johnson MK, J. Am. Chem. Soc 1997, 119, (10), 2590–2591. [Google Scholar]
  • 95.Site-Directed Mutagenesis of Recombinant Sulfite Oxidase: Identification of Cysteine 207 as a Ligand of Molybdenum. Garrett RM; Rajagopalan KV, J. Biol Chem 1996, 271, (13), 7387–7391. [PubMed] [Google Scholar]
  • 96.Dioxomolybdenum(VI) and Dioxotungsten(VI) Structure with Mono(dithiolato) Coordination: Crystal Structures, Properties and Reactivity of [{MoVIO2(dithiolato)}2(m-O)]2- (M = Mo and W) Complexes. Oku H; Ueyama N; Nakamura A, Bull. Chem. Soc. Jpn 1999, 72, (10), 2261–2270. [Google Scholar]
  • 97.Raman and Resonance Raman Spectra of Oxomolybdenum(VI) and -(V) Complexes of Cysteine and Related Thiolate Ligands. Ueyama N; Nakata M; Araki T; Nakamura A; Yamashita S; Yamashita T, Inorg. Chem 1981, 20, 1934–1937. [Google Scholar]
  • 98.Normal Coordinate Analysis of the Molybdenum Oxygen Stretching Frequencies in the Infrared and Raman-Spectra of O-16 Dioxomolybdenum(VI) and O-18 Dioxomolybdenum(VI) Complexes. Willis LJ; Loehr TM, Spectrochimica Acta Part a-Molecular and Biomolecular Spectroscopy 1987, 43, (1), 51–58. [Google Scholar]
  • 99.Raman and Infrared Spectroscopic Studies of Dioxomolybdenum(Vi) Complexes With Cysteamine Chelates. Willis LJ; Loehr TM; Miller KF; Bruce AE; Stiefel EI, Inorg. Chem 1986, 25, (23), 4289–4293. [Google Scholar]
  • 100.Resonance Raman Signatures of Oxomolybdenum Thiolate and Dithiolene Models of Molybdenum Proteins. Subramanian P; Burgmayer S; Richards S; Szalai V; Spiro TG, Inorg. Chem 1990, 29, 3849–3853. [Google Scholar]
  • 101.Spectroscopic and kinetic studies of Arabidopsis thaliana sulfite oxidase: Nature of the redox-active orbital and electronic structure contributions to catalysis. Hemann C; Hood BL; Fulton M; Hansch R; Schwarz G; Mendel RR; Kirk ML; Hille R, J. Am. Chem. Soc 2005, 127, (47), 16567. [DOI] [PubMed] [Google Scholar]
  • 102.Redox properties and catalytic activity of surface-bound human sulfite oxidase studied by a combined surface enhanced resonance Raman spectroscopic and electrochemical approach. Sezer M; Spricigo R; Utesch T; Millo D; Leimkuehler S; Mroginski MA; Wollenberger U; Hildebrandt P; Weidinger IM, Physical Chemistry Chemical Physics 2010, 12, (28), 7894–7903. [DOI] [PubMed] [Google Scholar]
  • 103.Pulsed ELDOR spectroscopy of the Mo(V)/Fe(III) state of sulfite oxidase prepared by one-electron reduction with Ti(III) citrate. Codd R; Astashkin AV; Pacheco A; Raitsimring AM; Enemark JH, Journal of Biological Inorganic Chemistry 2002, 7, (3), 338. [DOI] [PubMed] [Google Scholar]
  • 104.Dithiolene Coordination in the Molybdopterin Cofactor of DMSO Reductase: In Situ Evidence From Resonance Raman Spectroscopy. Gruber S; Kilpatrick L; Bastian N; Rajagopalan K; Spiro T, J. Am. Chem. Soc 1990, 112, (22), 8179–8180. [Google Scholar]
  • 105.Resonance-Raman Spectroscopy of Tris(1,2-Dithiolene) Complexes of Vanadium, Molybdenum, and Tungsten. Clark RJH; Turtle PC, J. Chem. Soc., Dalton Trans 1978, 1978, (12), 1714–1721. [Google Scholar]
  • 106.Large Ligand Folding Distortion in an Oxomolybdenum Donor Acceptor Complex. Yang J; Mogesa B; Basu P; Kirk ML, Inorg. Chem 2016, 55, (2), 785–793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.A Valence Bond Description of Dizwitterionic Dithiolene Character in an Oxomolybdenum-Bis(dithione) Complex. Mtei RP; Perera E; Mogesa B; Stein B; Basu P; Kirk ML, European Journal of Inorganic Chemistry 2011, (36), 5467–5470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Resonance Raman Spectroscopic Characterization of the Molybdopterin Active Site of DMSO Reductase. Kilpatrick L; Rajagopalan KV; Hilton J; Bastian NR; Stiefel EI; Pilato RS; Spiro TG, Biochemistry 1995, 34, (9), 3032–3039. [DOI] [PubMed] [Google Scholar]
  • 109.Active Site Structures and Catalytic Mechanism of Rhodobacter sphaeroides Dimethyl Sulfoxide Reductase as Revealed by Resonance Raman Spectroscopy. Garton SD; Hilton J; Oku H; Crouse BR; Rajagopalan KV; Johnson MK, J. Am. Chem. Soc 1997, 119, (52), 12906–12916. [Google Scholar]
  • 110.Resonance Raman Characterization of Biotin Sulfoxide Reductase: Comparing Oxomolybdenum Enzymes in the Me2SO Reductase Family. Garton SD; Temple CA; Dhawan IK; Barber MJ; Rajagopalan KV; Johnson MK, J. Biol. Chem 2000, 275, (10), 6798–6805. [DOI] [PubMed] [Google Scholar]
  • 111.Infrared Spectra and Normal Coordinate Analysis of 1,2—Dithiolate Complexes with Nickel. Schlapfer CW; Nakamoto K, Inorg. Chem 1975, 14, (6), 1338–1344. [Google Scholar]
  • 112.The active site of arsenite oxidase from Alcaligenes faecalis. Conrads T; Hemann C; George GN; Pickering IJ; Prince RC; Hille R, J. Am. Chem. Soc 2002, 124, (38), 11276–11277. [DOI] [PubMed] [Google Scholar]
  • 113.A Model for the Active-Site Formation Process in DMSO Reductase Family Molybdenum Enzymes Involving Oxido Alcoholato and Oxido Thiolato Molybdenum(VI) Core Structures. Sugimoto H; Sato M; Asano K; Suzuki T; Mieda K; Ogura T; Matsumoto T; Giles LJ; Pokhrel A; Kirk ML; Itoh S, Inorg. Chem 2016, 55, (4), 1542–1550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Monooxomolybdenum(VI) Complexes Possessing Olefinic Dithiolene Ligands: Probing Mo-S Covalency Contributions to Electron Transfer in Dimethyl Sulfoxide Reductase Family Molybdoenzymes. Sugimoto H; Tatemoto S; Suyama K; Miyake H; Mtei RP; Itoh S; Kirk ML, Inorg. Chem 2010, 49, (12), 5368–5370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Nature of the oxomolybdenum-thiolate pi-bond: Implications for Mo-S bonding in sulfite oxidase and xanthine oxidase. McNaughton RL; Helton ME; Cosper MM; Enemark JH; Kirk ML, Inorg. Chem 2004, 43, (5), 1625–1637. [DOI] [PubMed] [Google Scholar]
  • 116.Recent advances in understanding blue copper proteins. Solomon EI; Hadt RG, Coord. Chem. Rev 2011, 255, (7–8), 774–789. [Google Scholar]
  • 117.Modeling Blue Copper Protein Resonance Raman Spectra with Thiolate-Cu(II) Complexes of a Sterically Hindered Tris(pyrazolyl)borate. Qiu D; Kilpatrick LT; Kitajima N; Spiro TG, J. Am. Chem. Soc 1994, 116, (6), 2585–2590. [Google Scholar]
  • 118.The Reaction of Xanthine Oxidase with Lumazine: Characterization of the Reductive Half-reaction. Davis M; Olson J; Palmer G, J. Biol. Chem 1984, 259, (6), 3526–3533. [PubMed] [Google Scholar]
  • 119.Charge-transfer complexes between pteridine substrates and the active-center molybdenum of xanthine oxidase. Davis MD; Olson JS; Palmer G, J. Biol. Chem 1982, 257, (24), 4730–4737. [PubMed] [Google Scholar]
  • 120.Resonance Enhanced Raman Scattering from the Molybdenum Center of Xanthine Oxidase. Oertling W; Hille R, J. Biol. Chem 1990, 265, (29), 17446–17450. [PubMed] [Google Scholar]
  • 121.Resonance Raman studies of xanthine oxidase: the reduced enzyme - Product complex with violapterin. Hemann C; Ilich P; Stockert AL; Choi EY; Hille R, J. Phys. Chem. B 2005, 109, (7), 3023–3031. [DOI] [PubMed] [Google Scholar]
  • 122.Electronic structure contributions to reactivity in xanthine oxidase family enzymes. Stein B; Kirk M, J Biol Inorg Chem 2015, 20, (2), 183–194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Pyranopterin Dithiolene Distortions Relevant to Electron Transfer in Xanthine Oxidase/Dehydrogenase. Dong C; Yang J; Leimkühler S; Kirk ML, Inorg. Chem 2014, 53, (14), 7077–7079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Vibrational Probes of Molybdenum Cofactor–Protein Interactions in Xanthine Dehydrogenase. Dong C; Yang J; Reschke S; Leimkühler S; Kirk ML, Inorg. Chem 2017, 56, (12), 6830–6837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Resonance-Enhanced Raman-Scattering From the Molybdenum Center of Xanthine-Oxidase. Oertling WA; Hille R, J. Biol. Chem 1990, 265, (29), 17446–17450. [PubMed] [Google Scholar]
  • 126.A theoretical study on the mechanism of the reductive half-reaction of xanthine oxidase. Zhang XH; Wu YD, Inorg. Chem 2005, 44, (5), 1466–1471. [DOI] [PubMed] [Google Scholar]
  • 127.Spectroscopic and Electronic Structure Studies Probing Covalency Contributions to C-H Bond Activation and Transition-State Stabilization in Xanthine Oxidase. Sempombe J; Stein B; Kirk ML, Inorg. Chem 2011, 50, (21), 10919–10928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Resonance Raman studies on xanthine oxidase: observation of Mo-VI-ligand vibrations. Maiti NC; Tomita T; Kitagawa T; Okamoto K; Nishino T, Journal of Biological Inorganic Chemistry 2003, 8, (3), 327–333. [DOI] [PubMed] [Google Scholar]
  • 129.Raman-Scattering Properties of MoV=O and MoV=S Complexes as Probes For Molybdenum-Containing Enzymes. Backes G; Enemark JH; Loehr TM, Inorg. Chem 1991, 30, (8), 1839–1842. [Google Scholar]
  • 130.C-13 and Cu-63,Cu-65 ENDOR studies of CO Dehydrogenase from Oligotropha carboxidovorans. Experimental Evidence in Support of a Copper-Carbonyl Intermediate. Shanmugam M; Wilcoxen J; Habel-Rodriguez D; Cutsail GE; Kirk ML; Hoffman BM; Hille R, J. Am. Chem. Soc 2013, 135, (47), 17775–17782. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES