Molybdoenzymes-dependent Nitric Oxide Formation: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor:

Nitric oxide radical (NO) is a signalling molecule involved in several physiological and pathological processes and a new nitrate-nitrite-NO pathway has emerged as a physiological salvage pathway that operates under challenging conditions, when the "classic" L-arginine-dependent NO synthases are hindered. To catalyse the reduction of nitrite to NO, mammalian cells can use different metalloproteins that are present in cells to perform other functions (moonlighting). Among the so far identified ''non-dedicated nitrite reductases'', the molybdenum-containing enzymes stand out as very efficient NO synthases due to their well know ability to catalyse oxygen atom transfer reactions.

  • xanthine oxidase
  • aldehyde oxidase
  • molybdenum enzyme
  • nitrite
  • cell signalling
  • moonlighting

1. Context—I: The Molybdenum Side

Molybdenum is essential to the great majority of organisms [1][2][3][4], from archaea and bacteria to higher plants and mammals, being present in the active site of enzymes that catalyse redox reactions targeting nitrogen, sulfur and carbon atoms of key metabolites [5][6][7][8]. Molybdenum-containing enzymes are responsible for the biological handling of dinitrogen (Nitrogenase, Equation (1)), nitrate (Nitrate reductase (NaR), Equation (2)), sulfite (Sulfite oxidase (SO), Equation (3)), dimethylsulfoxide (dimethylsulfoxide reductase (DMSOR), Equation (4)), aldehydes (Aldehyde oxidase (AO), Equation (5)), xanthine (Xanthine oxidase (XO), Equation (6)) and carbon dioxide (Carbon dioxide dehydrogenase, Equation (7), and Formate dehydrogenase, Equation (8)), just to mention a few examples from the more than 50 enzymes presently known.
NN + 8H+ + 8e + 16MgATP → 2NH3 + H2 + 16MgADP + 16Pi
ONO2 + 2e + 2H+ → NO2 + H2O
SO22− + H2OOSO22− + 2e + 2H+ Molecules 28 05819 i002
+ 2e + 2H+Molecules 28 05819 i001

+ H2O

Molecules 28 05819 i003 + H2OMolecules 28 05819 i004 (H+) + 2e + 2H+
Molecules 28 05819 i005 + H2O Molecules 28 05819 i006 (H+) + 2e + 2H+
CO + H2O → OCO + 2e + 2H+
HCOO → CO2 + 2e + H+
With the single (as far as is presently known) exception of nitrogenase [9][10][11][12], molybdenum is found coordinated by the cis-dithiolene group (–S–C=C–S–) of one or two molecules of a pyranopterin cofactor (Figure 1). In a parallel situation to the haem ring, this unique cofactor is not an “innocent scaffold” and it is considered to be co-responsible to modulate the active site reactivity, besides acting as a “wire” to conduct the electrons to, or from, the other redox-active centres of the enzyme (intramolecular electron transfer, when this is the case) [13][14][15][16][17][18][19][20]. In addition to the pyranopterin cofactor, the molybdenum ion is coordinated by oxygen and/or sulfur and/or selenium terminal atoms and/or by enzyme-derived amino acid residues (Figure 1), which are also expected to have key roles in catalysis (although their individual roles are not yet understood in many molybdoenzymes). To introduce order in such a diversity of molybdenum compositions and (try to) rationalise the catalytic features, the molybdoenzymes were classified into three large families, denominated after one benchmark enzyme, as indicated in Figure 1 [5]: XO family, SO family and DMSOR family.
Figure 1. Structures of the molybdenum centres of the three families of molybdoenzymes. (a) Pyranopterin cofactor structure. The molecule is formed by pyrano (green)-pterin (blue)-dithiolene (orange) moieties. In eukaryotes, the cofactor is found in the simplest monophosphate form (R is a hydrogen atom; in gray), while in prokaryotes, it is most often found esterified with several nucleotides (R can be one cytidine monophosphate, guanosine monophosphate or adenosine monophosphate). (b) Structures of the molybdenum centres, in the oxidised form, of the three families of molybdoenzymes. For simplicity, only the cis-dithiolene group of the pyranopterin cofactor is represented. The labile oxygen position is indicated in red. It should be noted that the XO family member carbon monoxide dehydrogenase (X = S-Cu-S(Cys)) is suggested to harbour a labile Mo=O group and not Mo-OH. (c) Selected examples of enzymes from each family (mARC, mitochondrial amidoxime reducing component).
These molybdenum centres have been exploited by living organisms to carry out different reaction types [5][6][7][8], many of which are oxygen atom transfer reactions (OAT), where an oxygen atom is transferred from water to product -oxygen atom insertion (OAT-I) (Figure 2, violet arrows)- or from substrate to water -oxygen atom abstraction (OAT-A) (Figure 2, red arrows).
Figure 2. General mechanism of molybdoenzymes-catalysed OAT. The molybdenum role as electron acceptor/donor and as direct oxygen donor to substrate/oxygen acceptor from product is highlighted. OAT-I is represented in violet; e.g., for SO (Equation (3)), R would be sulfite and RO is sulfate; reduction of the enzyme physiological partner/regeneration of the enzyme active site is represented in dashes. OAT-A is represented in red; e.g., for NaR (Equation (2)), QO would be nitrate and Q is nitrite; oxidation of the enzyme physiological partner/regeneration of the enzyme active site is represented in dashes. It should be noted that catalysis can take place without the occurrence of the dashed reduction/oxidation.
The reaction mechanism of molybdoenzymes-catalysed OAT-I and OAT-A is presently well established and can be illustrated with the eukaryotic SO (Equation (3)) and NaR (Equation (2)) catalytic cycles (two enzymes from the SO family). Briefly, in SO [21][22][23][24][25][26][27][28][29][30][31][32][33][34] (Figure 3a), catalysis is initiated at the oxidised molybdenum centre, where its equatorial labile oxido group (Mo6+=Oequatorial) is (nucleophilically) attacked by the sulfite lone-pair of electrons, resulting in the formation of a reduced, covalent Mo4+-O-SO3 intermediate. After cleavage of the Mo-O(substrate)equatorial bond, sulfate is released (oxidation half-reaction). The Mo4+-OH(2) centre thus formed is then re-oxidised by two electrons (with the eventual reduction of the SO physiological partner—reduction half-reaction) to regenerate the initial Mo6+=O centre. The mechanism of OAT-A of NaR is suggested to be the reverse of the SO one (Figure 3b) [35][36][37][38][39][40][41], starting with the reduction of the molybdenum centre (two electrons provided by the NaR physiological partner—oxidation half-reaction). Nitrate binding yields the covalent intermediate Mo4+-O-NO2, which, after cleavage of the O-N bond and release of nitrite (reduction half-reaction), regenerates the Mo6+=O core. Hence, in a simplistic way, the molybdenum role in OAT is to accept/donate the necessary electrons for these redox reactions and to act as the direct oxygen donor to substrate/oxygen acceptor from product (Figure 2).
Figure 3. Simplified mechanism of OAT catalysed by eukaryotic SO (a), NaR (b), XO (c) and prokaryotic DMSOR (d).
This general OAT mechanism (Figure 2) is suggested to be followed by other molybdoenzymes, from other families, but with some variations imposed by the respective substrate nature. This is the case of the XO-catalysed xanthine hydroxylation to urate (Equation (6); OAT-I) a reaction more “complex” than the “simple” OAT of SO and NaR, because it requires the cleavage of the stable C(8)-H bond of xanthine for the oxygen to be inserted. To catalyse the xanthine hydroxylation, the molybdenum active site of XO holds an equatorial labile oxido group that is the direct source of the oxygen atom to be inserted into the xanthine C(8) (in parallel to SO). However, in the place of the S-coordinated cysteine residue of SO, XO active site harbours a terminal sulfido group to accept the xanthine H(8) and make its labile oxide group capable of carrying out a nucleophilic attack (instead of undergo a nucleophilic attack as in SO). Briefly, XO catalysis (Figure 3c) [42][43][44][45][46][47][48][49][50][51][52][53][54][55][56] is initiated with the activation of the equatorial labile oxido group by a conserved glutamate residue, to form the catalytically competent Mo6+(=S)-O centre. This makes a nucleophilic attack into xanthine C(8), while the sulfido group accepts the xanthine H(8) as a hydride, leading to the formation of a reduced, covalent intermediate Mo4+-O-C-R(-SH) (where R represents the remainder of the xanthine molecule). After hydrolysis of the Mo-O(xanthine)equatorial bond, urate is released (oxidation half-reaction) and the Mo4+(SH)-OH(2) centre formed is re-oxidised by two electrons (with the eventual reduction of the XO physiological partner—reduction half-reaction) to regenerate the initial Mo6+(=S)-O centre.
The enzymes from the DMSOR family, in spite of having its active site molybdenum ion coordinated by two molecules of the pyranopterin cofactor (Figure 1), are also suggested to follow the same general OAT mechanism (Figure 2), as illustrated, for example, by bacterial DMSOR itself (Figure 3d) [57][58][59][60][61][62][63][64][65][66][67] or bacterial NaRs [64][66][67][68][69][70][71][72][73][74][75][76][77][78][79][80][81]. Bacterial nitrite oxidoreductases, that catalyse the oxidation of nitrite to nitrate (NaR reverse reaction), also follow a typical OAT-A mechanism [82][83][84][85][86][87][88][89][90][91][92]. Noteworthy, to date, no parallel molybdenum-dependent nitrite reductase has been identified.

2. Context—II: The Nitric Oxide Side

Nitric oxide radical (NO, herein abbreviated as NO) is a signalling molecule involved in several physiological processes and, consequently, its deficit or excess is associated with many pathological conditions. In humans, NO is produced, in a tightly regulated manner, by three isoforms of NO synthase (NOS; neuronal, endothelial and inducible NOS), using L-arginine and dioxygen as the source of the NO nitrogen and oxygen atoms, respectively (Equation in Figure 4) [93][94][95][96]. The “NO signal” is transmitted intra- and extracellularly mainly through post-translational modifications of transition metal centres and thiols. The targets are mostly labile [4Fe-4S] centres and hemes (as is the case of the well-known activation of guanylate cyclase) and cysteine residues, that are converted into the respective nitrosyl (–metal-N=O) and nitrosothiol (-S-N=O) derivates. In addition, to limit the NO toxicity and ensure the fast “on/off signal” response, the NO life time is controlled through its rapid oxidation to nitrate (by reaction with oxy-hemoglobin and oxy-myoglobin [97][98][99][100][101][102][103][104]) and also to nitrite (by ceruloplasmin [105], cytochrome c oxidase [98][106][107] or dioxygen [108][109][110]) (Figure 4).
Figure 4. Schematic summary of NO formation–function–extinction. NO formation through NOS and nitrate and nitrite-dependent pathways (black arrows/gray shadowed area and violet arrows, respectively). NO consumption in signalling/function (arrow and shadowed area in green). NO extinction in reactions to limit its toxicity and control the signal specificity (oxidation to nitrate and nitrite; indigo arrows). Hb, haemoglobin; Mb, myoglobin; Cb, cytoglobin; Nb, neuroglobin; Cc, cytochrome c; CcO, Cc oxidase. See text for details.
This NO metabolic flux, formation–function–extinction (Figure 4), where nitrate and nitrite are considered end-products, with no physiological function, was firmly established by the end of the 20th century, when new ideas emerged. It began to become clear that nitrite can be reduced back to NO under anoxic conditions (Equation (9)) and exert a cytoprotective role during in vivo ischaemia and other pathological conditions [111][112][113][114][115][116][117][118][119][120][121][122][123][124][125][126][127][128][129][130][131][132][133][134][135][136][137][138][139][140][141][142][143][144][145]. In accordance, nitrite began to stop being seen as a useless end-metabolite, to become a “storage form” of NO. This new, alternative, NO source would be key to maintain the NO formation and ensure cell functioning under conditions of hypoxia/anoxia, precisely when the dioxygen-dependent NOS activity is impaired (Equation in Figure 4) and a “rescue” pathway is needed.
ONO + 1e + 2H+NO + H2O
At the same time, numerous studies began to be carried out to identify the pathway(s) responsible for nitrite reduction in mammals (humans), and, later, also in higher plants. The quest did not lead to the identification of any mammalian NO-forming nitrite reductase. Instead, the mammalian nitrite reduction activity was assigned to (already well-known) proteins that are present in cells to perform other (well-known) functions [146][147][148][149][150][151][152][153][154][155][156][157][158][159][160][161][162][163][164][165][166][167][168][169][170][171][172][173][174][175], including, at first, mainly the haem proteins haemoglobin (Hb) and myoglobin (Mb) and the molybdoenzyme XO (Figure 4). However, like all new ideas, the ability of those and other proteins to use nitrite to form and deliver NO faced severe criticism. Many authors questioned the physiological relevance of the NO fluxes generated from the physiologically available (very low) nitrite and called attention to the inhibition by other metabolites. For the “haem community”, the hurdle was the NO delivery, because of the (for long) known ability of haems to easily bind NO and hardly dissociate it. With XO, the main problem was its ability to form NO. The reaction feasibility was controversial, because XO is a well-established OAT-I enzyme (Equation (6)), rather than the OAT-A catalyst needed to reduce nitrite to NO (Equation (9)), and dioxygen was believed to completely inhibit the reaction.

This entry is adapted from the peer-reviewed paper 10.3390/molecules28155819

References

  1. Zhang, Y.; Gladyshev, V.N. Molybdoproteomes and evolution of molybdenum utilization. J. Mol. Biol. 2008, 379, 881–899.
  2. Zhang, Y.; Rump, S.; Gladyshev, V.N. Comparative genomics and evolution of molybdenum utilization. Coord. Chem. Rev. 2011, 255, 1206–1217.
  3. Gladyshev, V.N.; Zhang, Y. Chapter 2: Abundance, ubiquity and evolution of molybdoenzymes. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; pp. 81–99.
  4. Wells, M.; Kanmanii, N.J.; Al Zadjali, A.M.; Janecka, J.E.; Basu, P.; Oremland, R.S.; Stolz, J.F. Methane, arsenic, selenium and the origins of the DMSO reductase family. Sci. Rep. 2020, 10, 10946.
  5. Hille, R. The mononuclear molybdenum enzymes. Chem. Rev. 1996, 96, 2757–2816.
  6. Hille, R. The molybdenum oxotransferases and related enzymes. Dalton Trans. 2013, 42, 3029–3042.
  7. Hille, R.; Hall, J.; Basu, P. The mononuclear molybdenum enzymes. Chem. Rev. 2014, 114, 3963–4038.
  8. Maia, L.; Moura, J.J.G. Chapter 1: Molybdenum and tungsten-containing enzymes: An overview. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; pp. 1–80.
  9. Seefeldt, L.C.; Dean, D.R.; Hoffman, B.M. Nitrogenase Mechanism: Electron and Proton Accumulation and N2 Reduction. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; Chapter 8; pp. 274–296.
  10. Burén, S.; Jiménez-Vicente, E.; Echavarri-Erasun, C.; Rubio, L.M. Biosynthesis of Nitrogenase Cofactors. Chem. Rev. 2020, 120, 4921–4968.
  11. Einsle, O.; Rees, D.C. Structural Enzymology of Nitrogenase Enzymes. Chem. Rev. 2020, 120, 4969–5004.
  12. Stappen, V.; Decamps, L.; Cutsail, G.E.; Bjornsson, R.; Henthorn, J.T.; Birrell, J.A.; DeBeer, S. The Spectroscopy of Nitrogenases. Chem. Rev. 2020, 120, 5005–5081.
  13. Fischer, B.; Enemark, J.H.; Basu, P. A chemical approach to systematically designate the pyranopterin centers of molybdenum and tungsten enzymes and synthetic models. J. Inorg. Biochem. 1998, 72, 13–21.
  14. Basu, P.; Burgmayer, S.N.J. Pterin chemistry and its relationship to the molybdenum cofactor. Coord. Chem. Rev. 2011, 255, 1016–1038.
  15. Williams, B.R.; Fu, Y.; Yap, G.P.A.; Burgmayer, S.J.N. Structure and Reversible Pyran Formation in Molybdenum Pyranopterin Dithiolene Models of the Molybdenum Cofactor. J. Am. Chem. Soc. 2012, 134, 1958–1987.
  16. Dong, C.; Yang, J.; Leimkühler, S.; Kirk, M.L. Pyranopterin Dithiolene Distortions Relevant to Electron Transfer in Xanthine Oxidase/Dehydrogenase. Inorg. Chem. 2014, 53, 7077–7079.
  17. Rothery, R.A.; Weiner, J.H. Shifting the metallocentric molybdoenzyme paradigm: The importance of pyranopterin coordination. J. Biol. Inorg. Chem. 2015, 20, 349–372.
  18. Maiti, B.K.; Maia, L.B.; Moro, A.J.; Lima, J.C.; Cordas, C.M.; Moura, I.; Moura, J.J.G. Unusual Reduction Mechanism of Copper in Cysteine-Rich Environment. Inorg. Chem. 2018, 57, 8078–8088.
  19. Enemark, J.H. n, (n = 0–8): A general formalism for describing the highly covalent molybdenum cofactor of sulfite oxidase and related Mo enzymes. J. Inorg. Biochem. 2022, 231, 111801.
  20. Yang, J.; Enemark, J.H.; Kirk, M.L. Metal-Dithiolene Bonding Contributions to Pyranopterin Molybdenum Enzyme Reactivity. Inorganics 2020, 8, 19.
  21. Kisker, C.; Schindelin, H.; Pacheco, A.; Wehbi, W.A.; Garrett, R.M.; Rajagopalan, K.V.; Enemark, J.H.; Rees, D.C. Molecular basis of sulfite oxidase deficiency from the structure of sulfite oxidase. Cell 1997, 91, 973–983.
  22. Brody, M.S.; Hille, R. The kinetic behavior of chicken liver sulfite oxidase. Biochemistry 1999, 38, 6668–6677.
  23. Pacheco, A.; Hazzard, J.T.; Tollin, G.; Enemark, J.H. The pH dependence of intramolecular electron transfer rates in sulfite oxidase at high and low anion concentrations. J. Biol. Inorg. Chem. 1999, 4, 390–401.
  24. Feng, C.; Kedia, R.V.; Hazzard, J.T.; Hurley, J.K.; Tollin, G.; Enemark, J.H. Effect of solution viscosity on intramolecular electron transfer in sulfite oxidase. Biochemistry 2002, 41, 5816–5821.
  25. Peariso, K.; McNaughton, R.L.; Kirk, M.L. Active-site stereochemical control of oxygen atom transfer reactivity in sulfite oxidase. J. Am. Chem. Soc. 2002, 124, 9006–9007.
  26. Wilson, H.L.; Rajagopalan, K.V. The role of tyrosine 343 in substrate binding and catalysis by human sulfite oxidase. J. Biol. Chem. 2004, 279, 15105–15113.
  27. Peariso, K.; Helton, M.E.; Duesler, E.N.; Shadle, S.E.; Kirk, M.L. Sulfur K-edge spectroscopic investigation of second coordination sphere effects in oxomolybdenum-thiolates: Relationship to molybdenum-cysteine covalency and electron transfer in sulfite oxidase. Inorg. Chem. 2007, 46, 1259–1267.
  28. Bailey, S.; Rapson, T.; Johnson-Winters, K.; Astashkin, A.V.; Enemark, J.H.; Kappler, U. Molecular basis for enzymatic sulfite oxidation: How three conserved active site residues shape enzyme activity. J. Biol. Chem. 2009, 284, 2053–2063.
  29. Johnson-Winters, K.; Nordstrom, A.R.; Emesh, S.; Astashkin, A.V.; Rajapakshe, A.; Berry, R.E.; Tollin, G.; Enemark, J.H. Effects of interdomain tether length and flexibility on the kinetics of intramolecular electron transfer in human sulfite oxidase. Biochemistry 2010, 49, 1290–1296.
  30. Maiti, B.K.; Maia, L.B.; Pal, K.; Pakhira, B.; Avilés, T.; Moura, I.; Pauleta, S.R.; Nuñez, J.L.; Rizzi, A.C.; Brondino, C.D.; et al. One electron reduced square planar bis(benzene-1,2-dithiolato) copper dianionic complex and redox switch by O2/HO(-). Inorg. Chem. 2014, 53, 12799–12808.
  31. Kappler, U.; Enemark, J.H. Sulfite-oxidizing enzymes. J. Biol. Inorg. Chem. 2015, 20, 253–264.
  32. Enemark, J.H. Consensus structures of the Mo(v) sites of sulfite-oxidizing enzymes derived from variable frequency pulsed EPR spectroscopy, isotopic labelling and DFT calculations. Dalton Trans. 2017, 46, 13202–13210.
  33. Kappler, U.; Schwarz, G. The Sulfite Oxidase Family of Molybdenum Enzymes. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; Chapter 7; pp. 240–273.
  34. Caldararu, O.; Feldt, M.; Cioloboc, D.; van Severen, M.C.; Starke, K.; Mata, R.A.; Nordlander, E.; Ryde, U. QM/MM study of the reaction mechanism of sulfite oxidase. Sci. Rep. 2018, 8, 4684.
  35. Gutteridge, S.; Bray, R.C.; Notton, B.A.; Fido, R.J.; Hewitt, E.J. Studies by electron-paramagnetic-resonance spectroscopy of the molybdenum centre of spinach (Spinacia oleraceaSpinacia oleracea) nitrate reductase. Biochem. J. 1983, 213, 137–142.
  36. Solomonson, L.P.; Barber, M.J.; Howard, W.D.; Johnson, J.L.; Rajagopalan, K.V. Electron paramagnetic resonance studies on the molybdenum center of assimilatory NADH:nitrate reductase from Chlorella vulgaris. J. Biol. Chem. 1984, 259, 849–853.
  37. Campbell, W.H. Nitrate reductase structure, function and regulation: Bridging the Gap between Biochemistry and Physiology. Annu. Rev. Plant. Physiol. Plant. Mol. Biol. 1999, 50, 277–303.
  38. George, G.N.; Mertens, J.A.; Campbell, W.H. Structural Changes Induced by Catalytic Turnover at the Molybdenum Site of Arabidopsis Nitrate Reductase. J. Am. Chem. Soc. 1999, 121, 9730–9731.
  39. Fischer, K.; Barbier, G.G.; Hecht, H.J.; Mendel, R.R.; Campbell, W.H.; Schwarz, G. Structural basis of eukaryotic nitrate reduction: Crystal structures of the nitrate reductase active site. Plant Cell 2005, 17, 1167–1179.
  40. Chi, J.C.; Roeper, J.; Schwarz, G.; Fischer-Schrader, K. Dual binding of 14-3-3 protein regulates Arabidopsis nitrate reductase activity. J. Biol. Inorg. Chem. 2015, 20, 277–286.
  41. Sanz-Luque, E.; Chamizo-Ampudia, A.; Llamas, A.; Galvan, A.; Fernandez, E. Understanding nitrate assimilation and its regulation in microalgae. Front. Plant Sci. 2015, 6, 899.
  42. Murray, K.N.; Watson, J.G.; Chaykin, S. Catalysis of the direct transfer of oxygen from nicotinamide N-oxide to xanthine by xanthine oxidase. J. Biol. Chem. 1966, 241, 4798–4801.
  43. Hille, R.; Sprecher, H. On the mechanism of action of xanthine oxidase. Evidence in support of an oxo transfer mechanism in the molybdenum-containing hydroxylases. J. Biol. Chem. 1987, 262, 10914–10917.
  44. Enroth, C.; Eger, B.T.; Okamoto, K.; Nishino, T.; Nishino, T.; Pai, E.F. Crystal structures of bovine milk xanthine dehydrogenase and xanthine oxidase: Structure-based mechanism of conversion. Proc. Natl. Acad. Sci. USA 2000, 97, 10723–10728.
  45. Kuwabara, Y.; Nishino, T.; Okamoto, K.; Nishino, T. Unique amino acids cluster for switching from the dehydrogenase to oxidase form of xanthine oxidoreductase. Proc. Natl. Acad. Sci. USA 2003, 100, 8170–8175.
  46. Okamoto, K.; Matsumoto, K.; Hille, R.; Eger, B.T.; Pai, E.F.; Nishino, T. The crystal structure of xanthine oxidoreductase during catalysis: Implications for reaction mechanism and enzyme inhibition. Proc. Natl. Acad. Sci. USA 2004, 101, 7931–7936.
  47. Nishino, T.; Okamoto, K.; Kawaguchi, Y.; Hori, H.; Matsumura, T.; Eger, B.T.; Pai, E.F.; Nishino, T. Mechanism of the conversion of xanthine dehydrogenase to xanthine oxidase: Identification of the two cysteine disulfide bonds and crystal structure of a non-convertible rat liver xanthine dehydrogenase mutant. J. Biol. Chem. 2005, 280, 24888–24894.
  48. Nishino, T.; Okamoto, K.; Eger, B.T.; Pai, E.F.; Nishino, T. Mammalian xanthine oxidoreductase—mechanism of transition from xanthine dehydrogenase to xanthine oxidase. FEBS J. 2008, 275, 3278–3289.
  49. Pauff, J.M.; Zhang, J.; Bell, C.E.; Hille, R. Substrate orientation in xanthine oxidase: Crystal structure of enzyme in reaction with 2-hydroxy-6-methylpurine. J. Biol. Chem. 2008, 283, 4818–4824.
  50. Sempombe, J.; Stein, B.; Kirk, M.L. Spectroscopic and Electronic Structure Studies Probing Covalency Contributions to C-H Bond Activation and Transition-State Stabilization in Xanthine Oxidase. Inorg. Chem. 2011, 50, 10919–10928.
  51. Ishikita, H.; Eger, B.T.; Okamoto, K.; Nishino, T.; Pai, E.F. Protein conformational gating of enzymatic activity in xanthine oxidoreductase. J. Am. Chem. Soc. 2012, 134, 999–1009.
  52. Okamoto, K.; Kusano, T.; Nishino, T. Chemical nature and reaction mechanisms of the molybdenum cofactor of xanthine oxidoreductase. Curr. Pharm. Des. 2013, 19, 2606–2614.
  53. Cao, H.; Hall, J.; Hille, R. Substrate orientation and specificity in xanthine oxidase: Crystal structures of the enzyme in complex with indole-3-acetaldehyde and guanine. Biochemistry 2014, 53, 533–541.
  54. Nishino, T.; Okamoto, K.; Kawaguchi, Y.; Matsumura, T.; Eger, B.T.; Pai, E.F.; Nishino, T. The C-terminal peptide plays a role in the formation of an intermediate form during the transition between xanthine dehydrogenase and xanthine oxidase. FEBS J. 2015, 282, 3075–3090.
  55. Stein, B.W.; Kirk, M.L. Electronic structure contributions to reactivity in xanthine oxidase family enzymes. J. Biol. Inorg. Chem. 2015, 20, 183–194.
  56. Nishino, T.; Okamoto, K.; Leimkuhler, S. Enzymes of the Xanthino Oxidase Family. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; Chapter 6; pp. 192–239.
  57. Schultz, B.E.; Hille, R.; Holm, R.H. Direct oxygen atom transfer in the mechanism of action of Rhodobacter sphaeroides dimethyl sulfoxide reductase. J. Am. Chem. Soc. 1995, 117, 827–828.
  58. George, G.N.; Hilton, J.; Rajagopalan, K.V. X-ray Absorption Spectroscopy of Dimethyl Sulfoxide Reductase from Rhodobacter sphaeroides. J. Am. Chem. Soc. 1996, 118, 1113–1117.
  59. Schindelin, H.; Kisker, C.; Hilton, J.; Rajagopalan, K.V.; Rees, D.C. Crystal structure of DMSO reductase: Redox-linked changes in molybdopterin coordination. Science 1996, 272, 1615–1621.
  60. Schneider, F.; Löwe, J.; Huber, R.; Schindelin, H.; Kisker, C.; Knäblein, J. Crystal structure of dimethyl sulfoxide reductase from Rhodobacter capsulatus at 1.88 A resolution. J. Mol. Biol. 1996, 263, 53–69.
  61. 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, 12906–12916.
  62. George, G.N.; Hilton, J.; Temple, C.; Prince, R.C.; Rajagopalan, K.V. Structure of the Molybdenum Site of Dimethyl Sulfoxide Reductase. J. Am. Chem. Soc. 1999, 121, 1256–1266.
  63. Li, H.K.; Temple, C.; Rajagopalan, K.V.; Schindelin, H. The 1.3 Å Crystal Structure of Rhodobacter sphaeroides Dimethyl Sulfoxide Reductase Reveals Two Distinct Molybdenum Coordination Environments. J. Am. Chem. Soc. 2000, 122, 7673–7680.
  64. Magalon, A.; Ceccaldi, P.; Schoepp-Cothenet, B. The Prokaryotic Mo/W-bisPGD Family. In Molybdenum and Tungsten Enzymes; Hille, R., Schulzke, C., Kirk, M.L., Eds.; Royal Society of Chemistry: Cambridge, UK, 2017; Chapter 5; pp. 143–191.
  65. Pacheco, J.; Niks, D.; Hille, R. Kinetic and spectroscopic characterization of tungsten-substituted DMSO reductase from Rhodobacter sphaeroides. J. Biol. Inorg. Chem. 2018, 23, 295–301.
  66. Kirk, M.L.; Hille, R. Spectroscopic Studies of Mononuclear Molybdenum Enzyme Centers. Molecules 2022, 27, 4802.
  67. Kirk, M.L.; Lepluart, J.; Yang, J. Resonance Raman spectroscopy of pyranopterin molybdenum enzymes. J. Inorg. Biochem. 2022, 235, 111907.
  68. Dias, J.M.; Than, M.E.; Humm, A.; Huber, R.; Bourenkov, G.P.; Bartunik, H.D.; Bursakov, S.; Calvete, J.; Caldeira, J.; Carneiro, C.; et al. Crystal structure of the first dissimilatory nitrate reductase at 1.9 A solved by MAD methods. Structure 1999, 7, 65–79.
  69. Arnoux, P.; Sabaty, M.; Alric, J.; Frangioni, B.; Guigliarelli, B.; Adriano, J.-M.; Pignol, D. Structural and redox plasticity in the heterodimeric periplasmic nitrate reductase. Nat. Struct. Biol. 2003, 10, 928–934.
  70. Jepson, B.J.; Anderson, L.J.; Rubio, L.M.; Taylor, C.J.; Butler, C.S.; Flores, E.; Herrero, A.; Butt, J.N.; Richardson, D.J. Tuning a nitrate reductase for function. The first spectropotentiometric characterization of a bacterial assimilatory nitrate reductase reveals novel redox properties. J. Biol. Chem. 2004, 279, 32212–32218.
  71. Jormakka, M.; Richardson, D.; Byrne, B.; Iwata, S. Architecture of NarGH reveals a structural classification of Mo-bisMGD enzymes. Structure 2004, 12, 95–104.
  72. Bertero, M.G.; Rothery, R.A.; Boroumand, N.; Palak, M.; Blasco, F.; Ginet, N.; Weiner, J.H. Structural and biochemical characterization of a quinol binding site of Escherichia coli nitrate reductase A. J. Biol. Chem. 2005, 280, 14836–14843.
  73. Coelho, C.; González, P.J.; Trincão, J.; Carvalho, A.L.; Najmudin, S.; Hettman, T.; Dieckman, S.; Moura, J.J.G.; Moura, I.; Romão, M.J. Heterodimeric nitrate reductase (NapAB) from Cupriavidus necator H16: Purification, crystallization and preliminary X-ray analysis. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2007, 63, 516–519.
  74. Jepson, B.J.N.; Mohan, S.; Clarke, T.A.; Gates, A.J.; Cole, J.A.; Butler, C.S.; Butt, J.N.; Hemmings, A.M.; Richardson, D.J. Spectropotentiometric and structural analysis of the periplasmic nitrate reductase from Escherichia coli. J. Biol. Chem. 2007, 282, 6425–6437.
  75. Coelho, C.; González, P.J.; Moura, J.G.; Moura, I.; Trincão, J.; Romão, M.J. The crystal structure of Cupriavidus necator nitrate reductase in oxidized and partially reduced states. J. Mol. Biol. 2011, 408, 932–948.
  76. Sparacino-Watkins, C.; Stolz, J.F.; Basu, P. Nitrate and periplasmic nitrate reductases. Chem. Soc. Rev. 2014, 43, 676–706.
  77. Rendon, J.; Biaso, F.; Ceccaldi, P.; Toci, R.; Seduk, F.; Magalon, A.; Guigliarelli, B.; Grimaldi, S. Elucidating the Structures of the Low- and High-pH Mo(V) Species in Respiratory Nitrate Reductase: A Combined EPR, 14,15N HYSCORE, and DFT Study. Inorg. Chem. 2017, 56, 4423–4435.
  78. Mintmier, B.; McGarry, J.M.; Sparacino-Watkins, C.E.; Sallmen, J.; Fischer-Schrader, K.; Magalon, A.; McCormick, J.R.; Stolz, J.F.; Schwarz, G.; Bain, D.J.; et al. Molecular cloning, expression and biochemical characterization of periplasmic nitrate reductase from Campylobacter jejuni. FEMS Microbiol. Lett. 2018, 365, fny151.
  79. Seif Eddine, M.; Biaso, F.; Rendon, J.; Pilet, E.; Guigliarelli, B.; Magalon, A.; Grimaldi, S. 1,2H hyperfine spectroscopy and DFT modeling unveil the demethylmenasemiquinone binding mode to E. coli nitrate reductase A (NarGHI). Biochim. Biophys. Acta Bioenerg. 2020, 1861, 148203.
  80. Mintmier, B.; McGarry, J.M.; Bain, D.J.; Basu, P. Kinetic consequences of the endogenous ligand to molybdenum in the DMSO reductase family: A case study with periplasmic nitrate reductase. J. Biol. Inorg. Chem. 2021, 26, 13–28.
  81. Al-Attar, S.; Rendon, J.; Sidore, M.; Duneau, J.P.; Seduk, F.; Biaso, F.; Grimaldi, S.; Guigliarelli, B.; Magalon, A. Gating of Substrate Access and Long-Range Proton Transfer in Escherichia coli Nitrate Reductase A: The Essential Role of a Remote Glutamate Residue. ACS Catal. 2021, 11, 14303–14318.
  82. Kumar, S.; Nicholas, D.J.D.; Williams, E.H. Definitive 15N NMR evidence that water serves as a source of ‘O’ during nitrite oxidation by Nitrobacter agilis. FEBS Lett. 1983, 152, 71–74.
  83. Sundermeyer-Klinger, H.; Meyer, W.; Warninghoff, B.; Bock, E. Membrane-bound nitrite oxidoreductase of Nitrobacter: Evidence for a nitrate reductase system. Arch. Microbiol. 1984, 140, 153–158.
  84. DiSpirito, A.A.; Hooper, A.B. Oxygen exchange between nitrate molecules during nitrite oxidation by Nitrobacter. J. Biol. Chem. 1986, 261, 10534–10537.
  85. Friedman, S.H.; Massefski, W.; Hollocher, T.C. Catalysis of intermolecular oxygen atom transfer by nitrite dehydrogenase of Nitrobacter agilis. J. Biol. Chem. 1986, 261, 10538–10543.
  86. Meincke, M.; Bock, E.; Kastrau, D.; Kroneck, P.M.H. Nitrite oxidoreductase from Nitrobacter hamburgensis: Redox centers and their catalytic role. Arch. Microbiol. 1992, 158, 127–131.
  87. Kirstein, K.; Bock, E. Close genetic relationship between Nitrobacter hamburgensis nitrite oxidoreductase and Escherichia coli nitrate reductases. Arch. Microbiol. 1993, 160, 447–453.
  88. Spieck, E.; Muller, S.; Engel, A.; Mandelkow, E.; Patel, H.; Bock, E. Two-dimensional structure of membrane-bound nitrite oxidoreductase from Nitrobacter hamburgensis. J. Struct. Biol. 1996, 117–123.
  89. Spieck, E.; Ehrich, S.; Aamand, J.; Bock, E. Isolation and immunocytochemical location of the nitrite-oxidizing system in nitrospira moscoviensis. Arch. Microbiol. 1998, 169, 225–230.
  90. Martinez-Espinosa, R.M.; Dridge, E.J.; Bonete, M.J.; Butt, J.N.; Butler, C.S.; Sargent, F.; Richardson, D. Look on the positive side! The orientation, identification and bioenergetics of 'Archaeal' membrane-bound nitrate reductases. J. FEMS Microbiol. Lett. 2007, 276, 129–139.
  91. Lücker, S.; Wagner, M.; Maixner, F.; Pelletier, E.; Koch, H.; Vacherie, H.; Rattei, T.; Damstë, J.S.S.; Spieck, E.; Le Paslier, D.; et al. Nitrospira metagenome illuminates the physiology and evolution of globally important nitrite-oxidizing bacteria. Proc. Natl. Acad. Sci. USA 2010, 107, 13479–13484.
  92. Chicano, T.M.; Dietrich, L.; de Almeida, N.M.; Akram, M.; Hartmann, E.; Leidreiter, F.; Leopoldus, D.; Mueller, M.; Sánchez, R.; Nuijten, G.H.L. Structural and functional characterization of the intracellular filament-forming nitrite oxidoreductase multiprotein complex. Nat. Microbiol. 2021, 6, 1129–1139.
  93. Moncada, S.; Palmer, R.M.; Higgs, E.A. Nitric oxide: Physiology, pathophysiology, and pharmacology. Pharmacol Rev. 1991, 43, 109–142.
  94. Pfeiffer, S.; Mayer, B.; Hemmens, B. Nitric Oxide: Chemical Puzzles Posed by a Biological Messenger. Angew. Chem. Int. Edn. Engl. 1999, 38, 1714–1731.
  95. Stuehr, D.J. Mammalian nitric oxide synthases. Biochim. Biophys. Acta 1999, 1411, 217–230.
  96. Alderton, W.K.; Cooper, C.E.; Knowles, R.G. Nitric oxide synthases: Structure, function and inhibition. Biochem. J. 2001, 357, 593–615.
  97. Gow, A.J.; Luchsinger, B.P.; Pawloski, J.R.; Singel, D.J.; Stamler, J.S. The oxyhemoglobin reaction of nitric oxide. Proc. Natl. Acad. Sci. USA 1999, 96, 9027–9032.
  98. Brunori, M. Nitric oxide, cytochrome-c oxidase and myoglobin. Trends Biochem. Sci. 2001, 26, 21–23.
  99. Flögel, U.; Merx, M.W.; Gödecke, A.; Decking, U.K.M.; Schrader, J. Myoglobin: A scavenger of bioactive NO. Proc. Natl. Acad. Sci. USA 2001, 98, 735–740.
  100. Herold, S.; Exner, M.; Nauser, T. Kinetic and mechanistic studies of the NO*-mediated oxidation of oxymyoglobin and oxyhemoglobin. Biochemistry 2001, 40, 3385–3395.
  101. Huang, K.-T.; Han, T.H.; Hyduke, D.R.; Vaughn, M.W.; Herle, H.V.; Hein, T.W.; Zhang, C.; Kuo, L.; Liao, J.C. Modulation of nitric oxide bioavailability by erythrocytes. Proc. Natl. Acad. Sci. USA 2001, 98, 11771–11776.
  102. Witting, P.K.; Douglas, D.J.; Mauk, A.G. Reaction of human myoglobin and nitric oxide. Heme iron or protein sulfhydryl (s) nitrosation dependence on the absence or presence of oxygen. J. Biol. Chem. 2001, 276, 3991–3998.
  103. Joshi, M.S.; Ferguson, T.B., Jr.; Han, T.H.; Hyduke, D.R.; Liao, J.C.; Rassaf, T.; Bryan, N.; Feelisch, M.; Lancaster, J.R., Jr. Nitric oxide is consumed, rather than conserved, by reaction with oxyhemoglobin under physiological conditions. Proc. Natl. Acad. Sc. USA 2002, 99, 10341–10346.
  104. Gardner, P.R.; Gardner, A.M.; Brashear, W.T.; Suzuki, T.; Hvitved, A.N.; Setchell, K.D.R.; Olson, J.S. Hemoglobins dioxygenate nitric oxide with high fidelity. J. Inorg. Biochem. 2006, 100, 542–550.
  105. Shiva, S.; Wang, X.; Ringwood, L.-A.; Xu, X.; Yuditskaya, S.; Annavajjhala, V.; Miyajima, H.; Hogg, N.; Harris, Z.L.; Gladwin, M.T. Ceruloplasmin is a NO oxidase and nitrite synthase that determines endocrine NO homeostasis. Nat. Chem. Biol. 2006, 2, 486–493.
  106. Brunori, M.; Giuffre, A.; Forte, E.; Mastronicola, D.; Barone, M.C.; Sarti, P. Control of cytochrome c oxidase activity by nitric oxide. Biochim. Biophys. Acta 2004, 1655, 365–371.
  107. Poyton, R.O.; Castello, P.R.; Ball, K.A.; Woo, D.K.; Pan, N. Mitochondria and hypoxic signaling: A new view. Ann. N. Y. Acad. Sci. 2009, 1177, 48–56.
  108. Wink, D.A.; Darbyshire, J.F.; Nims, R.W.; Saavedra, J.E.; Ford, P.C. Reactions of the bioregulatory agent nitric oxide in oxygenated aqueous media: Determination of the kinetics for oxidation and nitrosation by intermediates generated in the NO/O2 reaction. Chem. Res. Toxicol. 1993, 6, 23–27.
  109. Goldstein, S.; Czapski, G. Kinetics of Nitric Oxide Autoxidation in Aqueous Solution in the Absence and Presence of Various Reductants. The Nature of the Oxidizing Intermediates. J. Am. Chem. Soc. 1995, 117, 12078–12084.
  110. Liu, X.; Liu, Q.; Gupta, E.; Zorko, N.; Brownlee, E.; Zweier, J.L. Quantitative measurements of NO reaction kinetics with a Clark-type electrode. Nitric Oxide 2005, 13, 68–77.
  111. Johnson, G., 3rd; Tsao, P.S.; Mulloy, D.; Lefer, A.M. Cardioprotective effects of acidified sodium nitrite in myocardial ischemia with reperfusion. J. Pharmacol. Exp. Ther. 1990, 252, 35–41.
  112. Demoncheaux, E.A.; Higenbottam, T.W.; Foster, P.J.; Borland, C.D.; Smith, A.P.; Marriott, H.M.; Bee, D.; Akamine, S.; Davies, M.B. Circulating nitrite anions are a directly acting vasodilator and are donors for nitric oxide. Clin. Sci. 2002, 102, 77–83.
  113. Hunter, C.J.; Dejam, A.; Blood, A.B.; Shields, H.; Kim-Shapiro, D.B.; Machado, R.F.; Tarekegn, S.; Mulla, N.; Hopper, A.O.; Schechter, A.N.; et al. Inhaled nebulized nitrite is a hypoxia-sensitive NO-dependent selective pulmonary vasodilator. Nat. Med. 2004, 10, 1122.
  114. Webb, A.; Bond, R.; McLean, P.; Uppal, R.; Benjamin, N.; Ahluwalia, A. Reduction of nitrite to nitric oxide during ischemia protects against myocardial ischemia–reperfusion damage. Proc. Natl. Acad. Sci. USA 2004, 101, 13683–13688.
  115. Duranski, M.R.; Greer, J.J.; Dejam, A.; Jaganmohan, S.; Hogg, N.; Langston, W.; Patel, R.P.; Yet, S.F.; Wang, X.; Kevil, C.G. Cytoprotective effects of nitrite during in vivo ischemia-reperfusion of the heart and liver. J. Clin. Investig. 2005, 115, 1232–1240.
  116. Lu, P.; Liu, F.; Yao, Z.; Wang, C.Y.; Chen, D.D.; Tian, Y.; Zhang, J.H.; Wu, Y.H. Nitrite-derived nitric oxide by xanthine oxidoreductase protects the liver against ischemia–reperfusion injury. Hepatobiliary Pancreat Dis. Int. 2005, 4, 350.
  117. Lundberg, J.O.; Weitzberg, E. NO Generation From Nitrite and Its Role in Vascular Control. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 915.
  118. Pluta, R.M.; Dejam, A.; Grimes, G.; Gladwin, M.T.; Oldfield, E.H. Nitrite infusions to prevent delayed cerebral vasospasm in a primate model of subarachnoid hemorrhage. JAMA 2005, 293, 1477–1484.
  119. Tsuchiya, K.; Kanematsu, Y.; Yoshizumi, M.; Ohnishi, H.; Kirima, K.; Izawa, Y.; Shikishima, M.; Ishida, T.; Kondo, S.; Kagami, S.; et al. Nitrite is an alternative source of NO in vivo. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H2163–H2170.
  120. Jung, K.-H.; Chu, K.; Ko, S.-Y.; Lee, S.-T.; Sinn, D.-I.; Park, D.-K.; Kim, J.-M.; Song, E.-C.; Kim, M.; Roh, J.K. Early Intravenous Infusion of Sodium Nitrite Protects Brain Against In Vivo Ischemia-Reperfusion Injury. Stroke 2006, 37, 2744.
  121. Baker, J.E.; Su, J.; Fu, X.; Hsu, A.; Gross, G.J.; Tweddell, J.S.; Hogg, N. Nitrite confers protection against myocardial infarction: Role of xanthine oxidoreductase, NADPH oxidase and K(ATP) channels. J. Mol. Cell. Cardiol. 2007, 43, 437.
  122. Bryan, N.S.; Calvert, J.W.; Elrod, J.W.; Gundewar, S.; Ji, S.Y.; Lefer, D.J. Dietary nitrite supplementation protects against myocardial ischemia-reperfusion injury. Proc. Natl. Acad. Sci. USA 2007, 104, 19144.
  123. Dezfulian, C.; Raat, N.; Shiva, S.; Gladwin, M.T. Role of the anion nitrite in ischemia-reperfusion cytoprotection and therapeutics. Cardiovasc. Res. 2007, 75, 327–338.
  124. Oldfield, E.H.; Cannon, R.O., 3rd; Schechter, A.N.; Gladwin, M.T. Nitrite infusion in humans and nonhuman primates: Endocrine effects, pharmacokinetics, and tolerance formation. Circulation 2007, 116, 1821–1832.
  125. Shiva, S.; Sack, M.N.; Greer, J.J.; Duranski, M.; Ringwood, L.A.; Burwell, L.; Wang, X.; MacArthur, P.H.; Shoja, A.; Raghavachari, N.; et al. Nitrite augments tolerance to ischemia/reperfusion injury via the modulation of mitochondrial electron transfer. J. Exp. Med. 2007, 204, 2089–2102.
  126. Bryan, N.S.; Calvert, J.W.; Gundewar, S.; Lefer, D.J. Dietary nitrite restores NO homeostasis and is cardioprotective in endothelial nitric oxide synthase-deficient mice. Free Radic. Biol. Med. 2008, 45, 468.
  127. Gonzalez, F.M.; Shiva, S.; Vincent, P.S.; Ringwood, L.A.; Hsu, L.Y.; Hon, Y.Y.; Aletras, A.H.; Cannon, R.O., 3rd; Gladwin, M.T.; Arai, A.E. Nitrite anion provides potent cytoprotective and antiapoptotic effects as adjunctive therapy to reperfusion for acute myocardial infarction. Circulation 2008, 117, 2986.
  128. Maher, A.R.; Milsom, A.B.; Gunaruwan, P.; Abozguia, K.; Ahmed, I.; Weaver, R.A.; Thomas, P.; Ashrafian, H.; Born, G.V.; James, P.E.; et al. Hypoxic modulation of exogenous nitrite-induced vasodilation in humans. Circulation 2008, 117, 670–677.
  129. Sinha, S.S.; Shiva, S.; Gladwin, M.T. Myocardial protection by nitrite: Evidence that this reperfusion therapeutic will not be lost in translation. Trends Cardiovasc. Med. 2008, 18, 163.
  130. Raat, N.J.; Shiva, S.; Gladwin, M.T. Effects of nitrite on modulating ROS generation following ischemia and reperfusion. Adv. Drug. Deliv. Rev. 2009, 61, 339–350.
  131. Zuckerbraun, B.S.; Shiva, S.; Ifedigbo, E.; Mathier, M.A.; Mollen, K.P.; Rao, J.; Bauer, P.M.; Choi, J.J.; Curtis, E.; Choi, A.M.; et al. Nitrite potently inhibits hypoxic and inflammatory pulmonary arterial hypertension and smooth muscle proliferation via xanthine oxidoreductase-dependent nitric oxide generation. Circulation 2010, 121, 98–109.
  132. Alef, M.J.; Vallabhaneni, R.; Carchman, E.; Morris, S.M., Jr.; Shiva, S.; Wang, Y.; Kelley, E.E.; Tarpey, M.M.; Gladwin, M.T.; Tzeng, E.; et al. Nitrite generated NO circumvents dysregulated arginine/NOS signaling to protect against intimal hyperplasia in Sprague-Dawley rats. J. Clin. Investig. 2011, 121, 1646.
  133. Blood, A.B.; Schroeder, H.J.; Terry, M.H.; Merrill-Henry, J.; Bragg, S.L.; Vrancken, K.; Liu, T.; Herring, J.L.; Sowers, L.C.; Wilson, S.M.; et al. Inhaled nitrite reverses hemolysis-induced pulmonary vasoconstriction in newborn lambs without blood participation. Circulation 2011, 123, 605–612.
  134. Gilchrist, M.; Shore, A.C.; Benjamin, N. Inorganic nitrate and nitrite and control of blood pressure. Cardiovasc. Res. 2011, 89, 492–498.
  135. Kevil, C.G.; Kolluru, G.K.; Pattillo, C.B.; Giordano, T. Inorganic nitrite therapy: Historical perspective and future directions. Free Radic. Biol. Med. 2011, 51, 576–593.
  136. Larsen, F.J.; Schiffer, T.A.; Borniquel, S.; Sahlin, K.; Ekblom, B.; Lundberg, J.O.; Weitzberg, E. Dietary inorganic nitrate improves mitochondrial efficiency in humans. Cell Metab. 2011, 13, 149–159.
  137. Lundberg, J.O.; Carlstrom, M.; Larsen, F.J.; Weitzberg, E. Roles of dietary inorganic nitrate in cardiovascular health and disease. Cardiovasc. Res. 2011, 89, 525–532.
  138. Murillo, D.; Kamga, C.; Mo, L.; Shiva, S. Nitrite as a mediator of ischemic preconditioning and cytoprotection. Nitric Oxide 2011, 25, 70–80.
  139. Pattillo, C.B.; Bir, S.; Rajaram, V.; Kevil, C. G Inorganic nitrite and chronic tissue ischaemia: A novel therapeutic modality for peripheral vascular diseases. Cardiovasc. Res. 2011, 89, 533–541.
  140. Sindler, A.L.; Fleenor, B.S.; Calvert, J.W.; Marshall, K.D.; Zigler, M.L.; Lefer, D.J.; Seals, D.R. Nitrite supplementation reverses vascular endothelial dysfunction and large elastic artery stiffness with aging. Aging Cell 2011, 10, 429–437.
  141. Baliga, R.S.; Milsom, A.B.; Ghosh, S.M.; Trinder, S.L.; Macallister, R.J.; Ahluwalia, A.; Hobbs, A.J. Dietary nitrate ameliorates pulmonary hypertension: Cytoprotective role for endothelial nitric oxide synthase and xanthine oxidoreductase. Circulation 2012, 125, 2922–2932.
  142. Sparacino-Watkins, C.E.; Lai, Y.C.; Gladwin, M.T. Nitrate-nitrite-nitric oxide pathway in pulmonary arterial hypertension therapeutics. Ther. Circ. 2012, 125, 2824–2826.
  143. Bueno, M.; Wang, J.; Mora, A.L.; Gladwin, M.T. Nitrite signaling in pulmonary hypertension: Mechanisms of bioactivation, signaling, and therapeutics. Antioxid. Redox Signal. 2013, 18, 1797–1809.
  144. Ghosh, S.M.; Kapil, V.; Fuentes-Calvo, I.; Bubb, K.J.; Pearl, V.; Milsom, A.B.; Khambata, R.; Maleki-Toyserkani, S.; Yousuf, M.; Benjamin, N.; et al. Enhanced vasodilator activity of nitrite in hypertension: Critical role for erythrocytic xanthine oxidoreductase and translational potential. Hypertension 2013, 61, 1091–1102.
  145. Omar, S.A.; Webb, A. Nitrite reduction and cardiovascular protection. J. Mol. Cell Cardiol. 2014, 73, 57–69.
  146. Millar, T.M.; Stevens, C.R.; Benjamin, N.; Eisenthal, R.; Harrison, R.; Blake, D.R. Xanthine oxidoreductase catalyses the reduction of nitrates and nitrite to nitric oxide under hypoxic conditions. FEBS Lett. 1998, 427, 225–228.
  147. Zhang, Z.; Naughton, D.; Winyard, P.G.; Benjamin, N.; Blake, D.R.; Symons, M.C. Generation of nitric oxide by a nitrite reductase activity of xanthine oxidase: A potential pathway for nitric oxide formation in the absence of nitric oxide synthase activity. Biochem. Biophys. Res. Commun. 1998, 249, 767–772.
  148. Kozlov, A.V.; Staniek, K.; Nohl, H. Nitrite reductase activity is a novel function of mammalian mitochondria. FEBS Lett. 1999, 454, 127–130.
  149. Godber, H.L.J.; Doel, J.J.; Sapkota, G.P.; Blake, D.R.; Stevens, C.R.; Eisenthal, R.; Harrison, R. Reduction of nitrite to nitric oxide catalyzed by xanthine oxidoreductase. J. Biol. Chem. 2000, 275, 7757–7763.
  150. Li, H.; Samouilov, A.; Liu, X.; Zweier, J.L. Characterization of the magnitude and kinetics of xanthine oxidase-catalyzed nitrite reduction. Evaluation of its role in nitric oxide generation in anoxic tissues. J. Biol. Chem. 2001, 276, 24482–24489.
  151. Li, H.; Samouilov, A.; Liu, X.; Zweier, J.L. Characterization of the effects of oxygen on xanthine oxidase-mediated nitric oxide formation. J. Biol. Chem. 2004, 279, 16939–16946.
  152. Castello, P.R.; David, P.S.; McClure, T.; Crook, Z.; Poyton, R.O. Mitochondrial cytochrome oxidase produces nitric oxide under hypoxic conditions: Implications for oxygen sensing and hypoxic signaling in eukaryotes. Cell Metab. 2006, 3, 277–287.
  153. Gautier, C.; van Faassen, E.; Mikula, I.; Martasek, P.; Slama-Schwok, A. Endothelial nitric oxide synthase reduces nitrite anions to NO under anoxia. Biochem. Biophys. Res. Commun. 2006, 341, 816–821.
  154. Li, H.; Liu, X.; Cui, H.; Chen, Y.R.; Cardounel, A.J.; Zweier, J.L. Characterization of the mechanism of cytochrome P450 reductase-cytochrome P450-mediated nitric oxide and nitrosothiol generation from organic nitrates. J. Biol. Chem. 2006, 281, 12546–12554.
  155. Vanin, A.F.; Bevers, L.M.; Slama-Schwok, A.; van Faassen, E.E. Nitric oxide synthase reduces nitrite to NO under anoxia. Cell Mol. Life Sci. 2006, 64, 96–103.
  156. Basu, S.; Azarova, N.A.; Font, M.D.; King, S.B.; Hogg, N.; Gladwin, M.T.; Shiva, S.; Kim-Shapiro, D.B. Nitrite reductase activity of cytochrome c. J. Biol. Chem. 2008, 283, 32590–32597.
  157. Benamar, A.; Rolletschek, H.; Borisjuk, L.; Avelange-Macherel, M.-H.; Curien, G.; Mostefai, H.A.; Andriantsitohaina, R.; Macherel, D. Nitrite-nitric oxide control of mitochondrial respiration at the frontier of anoxia. Biochim. Biophys. Acta 2008, 1777, 1268–1275.
  158. Castello, P.R.; Woo, D.K.; Ball, K.; Wojcik, J.; Liu, L.; Poyton, R.O. Oxygen-regulated isoforms of cytochrome c oxidase have differential effects on its nitric oxide production and on hypoxic signaling. Proc. Natl. Acad. Sci. USA 2008, 105, 8203–8208.
  159. Li, H.; Cui, H.; Kundu, T.K.; Alzawahra, W.; Zweier, J.L. Nitric oxide production from nitrite occurs primarily in tissues not in the blood: Critical role of xanthine oxidase and aldehyde oxidase. J. Biol. Chem. 2008, 283, 17855–17863.
  160. Li, H.; Kundu, T.K.; Zweier, J.L. Characterization of the magnitude and mechanism of aldehyde oxidase-mediated nitric oxide production from nitrite. J. Biol. Chem. 2009, 284, 33850–33858.
  161. Badejo, A.M., Jr.; Hodnette, C.; Dhaliwal, J.S.; Casey, D.B.; Pankey, E.; Murthy, S.N.; Nossaman, B.D.; Hyman, A.L.; Kadowitz, P. Mitochondrial aldehyde dehydrogenase mediates vasodilator responses of glyceryl trinitrate and sodium nitrite in the pulmonary vascular bed of the rat. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H819–H826.
  162. Tiso, M.; Tejero, J.; Basu, S.; Azarov, I.; Wang, X.; Simplaceanu, V.; Frizzell, S.; Jayaraman, T.; Geary, L.; Shapiro, C.; et al. Human neuroglobin functions as a redox-regulated nitrite reductase. J. Biol. Chem. 2011, 286, 18277–18289.
  163. Li, H.; Hemann, C.; Abdelghany, T.M.; El-Mahdy, M.A.; Zweier, J.L. Characterization of the mechanism and magnitude of cytoglobin-mediated nitrite reduction and nitric oxide generation under anaerobic conditions. J. Biol. Chem. 2012, 278, 36623–36633.
  164. Sparacino-Watkins, C.E.; Tejero, J.; Sun, B.; Gauthier, M.C.; Thomas, J.; Ragireddy, V.; Merchant, B.A.; Wang, J.; Azarov, I.; Basu, P.; et al. Nitrite reductase and nitric-oxide synthase activity of the mitochondrial molybdopterin enzymes mARC1 and mARC2. J. Biol. Chem. 2014, 289, 10345–10358.
  165. Wang, J.; Krizowski, S.; Fischer-Schrader, K.; Niks, D.; Tejero, J.; Sparacino-Watkins, C.; Wang, L.; Ragireddy, V.; Frizzell, S.; Kelley, E.E.; et al. Sulfite Oxidase Catalyzes Single-Electron Transfer at Molybdenum Domain to Reduce Nitrite to Nitric Oxide. Antioxid. Redox. Signal. 2015, 23, 283–294.
  166. Chamizo-Ampudia, A.; Sanz-Luque, E.; Llamas, Á.; Ocaña-Calahorro, F.; Mariscal, V.; Carreras, A.; Barroso, J.B.; Galván, A.; Fernández, E. A dual system formed by the ARC and NR molybdoenzymes mediates nitrite-dependent NO production in Chlamydomonas. Plant Cell Environ. 2016, 39, 2097–2107.
  167. Bender, D.; Tobias Kaczmarek, A.; Niks, D.; Hille, R.; Schwarz, G. Mechanism of nitrite-dependent NO synthesis by human sulfite oxidase. Biochem. J. 2019, 476, 1805–1815.
  168. Kaczmarek, A.T.; Strampraad, M.J.F.; Hagedoorn, P.L.; Schwarz, G. Reciprocal regulation of sulfite oxidation and nitrite reduction by mitochondrial sulfite oxidase. Nitric Oxide 2019, 89, 22–31.
  169. Mohn, M.A.; Thaqi, B.; Fischer-Schrader, K. Isoform-Specific NO Synthesis by Arabidopsis thaliana Nitrate Reductase. Plants 2019, 8, 67.
  170. Mutus, B. The catalytic mechanism for NO production by the mitochondrial enzyme, sulfite oxidase. Biochem. J. 2019, 476, 1955–1956.
  171. Tejada-Jimenez, M.; Llamas, A.; Galván, A.; Fernández, E. Role of Nitrate Reductase in NO Production in Photosynthetic Eukaryotes. Plants 2019, 8, 56.
  172. Bender, D.; Kaczmarek, A.T.; Kuester, S.; Burlina, A.B.; Schwarz, G. Oxygen and nitrite reduction by heme-deficient sulphite oxidase in a patient with mild sulphite oxidase deficiency. J. Inherit. Metab. Dis. 2020, 43, 748–757.
  173. Costa-Broseta, Á.; Castillo, M.; León, J. Post-Translational Modifications of Nitrate Reductases Autoregulates Nitric Oxide Biosynthesis in Arabidopsis. Int. J. Mol. Sci. 2021, 22, 549.
  174. Gupta, K.J.; Kaladhar, V.C.; Fitzpatrick, T.B.; Fernie, A.R.; Møller, I.M.; Loake, G.J. Nitric oxide regulation of plant metabolism. Mol. Plant. 2022, 15, 228–242.
  175. Timilsina, A.; Dong, W.; Hasanuzzaman, M.; Liu, B.; Hu, C. Nitrate-Nitrite-Nitric Oxide Pathway: A Mechanism of Hypoxia and Anoxia Tolerance in Plants. Int. J. Mol. Sci. 2022, 23, 11522.
More
This entry is offline, you can click here to edit this entry!
ScholarVision Creations