Biochemistry of Peroxynitrite and Protein Tyrosine ... - ACS Publications

Feb 5, 2018 - Bioquímica, Facultad de Medicina,. ∥. Laboratorio de Enzimología, Facultad de Ciencias, Universidad de la República, Montevideo,. U...
0 downloads 0 Views 7MB Size
Review pubs.acs.org/CR

Cite This: Chem. Rev. 2018, 118, 1338−1408

Biochemistry of Peroxynitrite and Protein Tyrosine Nitration Gerardo Ferrer-Sueta,†,‡,⊥ Nicolás Campolo,‡,§,⊥ Madia Trujillo,‡,§ Silvina Bartesaghi,‡,§ Sebastián Carballal,‡,§ Natalia Romero,‡,§,@ Beatriz Alvarez,‡,∥ and Rafael Radi*,‡,§ Laboratorio de Fisicoquímica Biológica, Facultad de Ciencias, ‡Center for Free Radical and Biomedical Research, §Departamento de Bioquímica, Facultad de Medicina, ∥Laboratorio de Enzimología, Facultad de Ciencias, Universidad de la República, Montevideo, Uruguay

Downloaded via UNIV OF SUNDERLAND on October 8, 2018 at 17:40:26 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



ABSTRACT: Peroxynitrite is a short-lived and reactive biological oxidant formed from the diffusion-controlled reaction of the free radicals superoxide (O2•−) and nitric oxide (•NO). In this review, we first analyze the biochemical evidence for the formation of peroxynitrite in vivo and the reactions that lead to it. Then, we describe the principal reactions that peroxynitrite undergoes with biological targets and provide kinetic and mechanistic details. In these reactions, peroxynitrite has roles as (1) peroxide, (2) Lewis base, and (3) free radical generator. Physiological levels of CO2 can change the outcome of peroxynitrite reactions. The second part of the review assesses the formation of protein 3-nitrotyrosine (NO2Tyr) by peroxynitrite-dependent and -independent mechanisms, as one of the hallmarks of the actions of •NO-derived oxidants in biological systems. Moreover, tyrosine nitration impacts protein structure and function, tyrosine kinase signal transduction cascades and protein turnover. Overall, the review is aimed to provide an integrated biochemical view on the formation and reactions of peroxynitrite under biologically relevant conditions and the impact of this stealthy oxidant and one of its major footprints, protein NO2Tyr, in the disruption of cellular homeostasis.

CONTENTS 1. Introduction to Peroxynitrite History and Biochemistry 1.1. Early Evidence of Peroxynitrite in Biological Systems 1.2. The Question of Homolysis 1.3. Biological Formation of Peroxynitrite from Nitric Oxide and Superoxide 1.4. Formation of Peroxynitrite from the Reaction of Nitroxyl (HNO) with Oxygen 1.5. Formation of Peroxynitrite on Metal Centers 1.6. Peroxynitrite Synthesis, Quantitation, and Handling in the Laboratory 1.6.1. Peroxynitrite Administration and Exposure in Experimental Setups 2. Reactions of Peroxynitrite in Biochemistry 2.1. Kinetically Competitive Reactions 2.2. Carbon Dioxide Is the Criteria of Kinetic Competitiveness 2.3. Metal Centers in Proteins 2.3.1. Heme Peroxidases and Heme Thiolate Proteins 2.3.2. Cytochrome c Oxidase 2.3.3. Iron−Sulfur Cluster Proteins 2.4. Reduction of Peroxynitrite by Thiols and Selenols 2.4.1. Kinetic Competitiveness of Thiol and Selenol Peroxidases 2.5. Other Targets of Peroxynitrite 2.5.1. Methionine 2.5.2. Tryptophan © 2018 American Chemical Society

2.5.3. Uric Acid 2.5.4. Ascorbate 2.5.5. NAD(P)H 2.5.6. Coenzymes 2.5.7. Carbonyl Compounds 2.5.8. Ubiquinol 2.6. Fate of Peroxynitrite 2.6.1. Isomerization to Nitrate 2.6.2. Reduction to Nitrite 2.6.3. Reduction to Nitrogen Dioxide 3. Radicals and Secondary Oxidants Derived from Peroxynitrite 3.1. Carbonate Radical Biochemistry 3.1.1. Formation, Properties, and Reactions 3.1.2. Role of CO3•− in Pathophysiological Processes 3.2. Nitrogen Dioxide Radical Biochemistry 3.2.1. Formation Pathways 3.2.2. Reactivity 3.2.3. Role of •NO2 in Pathophysiological Processes 4. Tyrosine Nitration 4.1. Historical Background 4.2. Mechanisms of Tyrosine Nitration 4.2.1. General Free Radical Mechanism

1339 1339 1340 1341 1343 1343 1344 1344 1346 1347 1347 1348 1348 1349 1349

1356 1356 1358 1358 1358 1359 1359 1359 1359 1360 1360 1360 1360 1361 1362 1362 1363 1363 1364 1364 1367 1367

1350 1354 1354 1354 1355

Special Issue: Posttranslational Protein Modifications Received: September 17, 2017 Published: February 5, 2018 1338

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews 4.2.2. Biological Pathways of Tyrosine Nitration 4.3. Selectivity of Protein Tyrosine Nitration 4.3.1. Nitration of Nonmetalloproteins 4.3.2. Nitration of Metalloproteins 4.4. Secondary Oxidative Products of Tyrosine 4.4.1. Dimerization to 3,3′-Dityrosine 4.4.2. Hydroxylation to 3-Hydroxytyrosine 4.5. Biological Consequences of Protein Tyrosine Nitration 4.5.1. Physicochemical Properties of 3-Nitrotyrosine 4.5.2. Effects of Tyrosine Nitration on Protein Structure and Function 4.5.3. Effects of Protein Tyrosine Nitration on Cell Signaling Cascades 4.5.4. Tyrosine Nitration in the Immune Response 4.6. Metabolic Fate of Tyrosine Nitrated Proteins 4.6.1. Proteolytic Degradation of Tyrosine Nitrated Proteins 4.6.2. Metabolism of Free 3-Nitrotyrosine 4.6.3. Denitration of Tyrosine Nitrated Proteins 4.7. Structural Biology of Tyrosine-Nitrated Proteins 4.7.1. X-ray Crystallography of Tyrosine-Nitrated Proteins 4.7.2. Computational Methods 4.7.3. Other Experimental Approaches 5. Conclusions and Perspectives Author Information Corresponding Author ORCID Present Address Author Contributions Notes Biographies Acknowledgments References

Review •

1368 1369 1369 1371 1372 1372 1372

NO + O2•− → ONOO−

ONOO− + H+ ⇌ ONOOH

(1) (2)

Despite being known in the chemical literature since the beginning of the 20th century, peroxynitrite emerged in the biomedical literature in the early 1990’s,2,7−15 and since then it has been progressively identified as an important biological oxidant, playing roles as mediator in signal transduction pathways as well as in the disruption of cellular redox homeostasis.4,16,17 Among the several reactions that peroxynitrite and peroxynitrite-derived species can undergo with biomolecular targets, the nitration of protein tyrosine residues has been revealed as one of their main hallmarks. Early seminal articles in the chemical literature showed that peroxynitrous acid was capable of nitrating phenolic compounds.18 Being peroxynitrite a short-lived transient species in biological systems (ca. 1−20 ms4), protein tyrosine nitration represents its premier footprint in cells and tissues. Moreover, protein tyrosine nitration constitutes an oxidative post-translational modification that can impact in protein structure and function. Thus, this review will be presented in two parts: the first part will analyze the biochemistry of peroxynitrite, and the second part will focus on the mechanisms and consequences of protein tyrosine nitration.

1373 1373 1373 1377 1379 1380 1381 1381 1382

1.1. Early Evidence of Peroxynitrite in Biological Systems

1383

During almost 90 years of the 20th century only a few articles about peroxynitrite had been published and almost all of them were based in its unusual oxidant properties and chemical synthesis or were related to atmospheric pollution chemistry, but without any relationship with its biological role.19 It was right after the discovery of the chemical nature of the endothelium-derived relaxing factor (EDRF), finally identified as •NO,20,21 that scientists started to consider peroxynitrite as a biologically relevant molecule. In fact, the first PubMed citation using peroxynitrite as a keyword was the work of Beckman et al.,2 a paper that is a landmark of the studies about the biological role of peroxynitrite. However, several previous pieces of evidence supported the formation of peroxynitrite in biological systems, although its name or chemical nature was not appreciated or mentioned. During the 1980’s, several groups were intensely working to define the identify of a relatively unstable compound described as the endothelium-derived relaxing factor or EDRF, a molecule responsible for promoting vessel relaxation.22−26 In 1986, Gryglewsky, Palmer, and Moncada showed that the addition of the enzyme superoxide dismutase (SOD), known to catalyze the fast dismutation of O2•− (eq 3), increased the EDRFinduced relaxation of aortic rings denuded of endothelium perfused with extracellular media of endothelial cell cultures,27 suggesting that O2•− contributed to the chemical instability of EDRF.

1383 1384 1384 1385 1386 1386 1386 1386 1386 1386 1386 1386 1387

1. INTRODUCTION TO PEROXYNITRITE HISTORY AND BIOCHEMISTRY In the field of biochemistry, the term “peroxynitrite” encompasses the sum of two closely related species: peroxynitrite anion (ONOO−) and its conjugated acid, peroxynitrous acid (ONOOH). The IUPAC recommended names are oxoperoxonitrate (1−) and hydrogen oxoperoxonitrate, respectively. The reason for including both species in the definition of peroxynitrite stands on the fact that they coexist and are in rapid acid−base equilibrium under biologically relevant conditions (pKa = 6.8). Due to the transient nature of its precursor species and the chemical characteristics of peroxynitrite itself, establishing its formation, reactions, biochemical consequences, and relevance in biological systems represented major efforts and debates from the chemical, biological, and biomedical research communities, which finally amalgamated and settled over the past decade. The main reaction leading to the formation of peroxynitrite in biological systems is the diffusion-controlled reaction of the two radicals nitric oxide (•NO) and superoxide (O2•−) (eq 1):1−6

SOD

2O2•− + 2H+ ⎯⎯⎯→ H 2O2 + O2

(3)

An increase in the actions of EDRF was obtained in the presence of cytochrome c3+, a known scavenger of O2•− (eq 4): Cyt c3 + + O2•− → Cyt c 2 + + O2

(4)

O2•−-mediated 28

Similar results on the EDRF inactivation were also obtained by Rubanyi et al., including the observation of enhanced instability of EDRF under hyperoxic conditions. Additionally, Moncada et al. showed that redox-cycling 1339

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

that cogenerate •NO and O2•− and that more closely resemble its biological formation. For example, while Radi et al.7 found that authentic peoxynitrite could induce lipid peroxidation, Hogg et al. reported that the product of O2•− and •NO could oxidize lipoproteins such as LDL and this process contributed to the pathogenesis of atherosclerosis in vivo.34,35 Another early piece of evidence of the formation of peroxynitrite in vivo came from studies about the oxidative mechanisms associated with inflammatory processes. Activation of macrophages or neutrophils promotes the assembly of the enzyme NADPH oxidase responsible for the oxidative burst and concomitant O2•− production at inflammation sites. Studies using rat lung demonstrated that vascular injury induced by immunocomplex is not only dependent on polymorphonuclear cell activation and O2•− formation but also is enhanced in the presence of L-arginine. In addition, the arginine analogue, Nγ-monomethyl-L-arginine, which blocks • NO formation, protected against the vascular injury.36 Soon after the initial proposal of its biological formation and actions,2,7,8 the discovery that peroxynitrite promotes nitration of tyrosine residues to form the stable product 3-nitrotyrosine (NO2Tyr)10,11 provided a framework to reveal peroxynitrite formation by cells and tissues. Indeed, immunochemical and bioanalytical methods allowed to rapidly advance on the knowledge of peroxynitrite formation in biological systems through the detection of protein NO 2 Tyr. In 1992, Ischiropoulos et al. demonstrated that the enzyme Cu,ZnSOD was able to catalyze peroxynitrite-mediated tyrosine nitration11 and used this readout to demonstrate that peroxynitrite can be formed by activated rat alveolar macrophages, suggesting that peroxynitrite formation may contribute to inflammatory cell-mediated tissue injury.10 A couple of years later, Beckman et al. provided another fundamental demonstration of peroxynitrite formation in vivo by detecting for the first time the presence of protein NO2Tyr in human aorta atherosclerotic lesions immunochemically.37 In the following years, and with the use of antibodies that specifically recognize NO2Tyr,38 numerous papers were published confirming the presence of NO2Tyr in diverse tissues; protein NO2Tyr was found even in healthy tissues and its levels were usually found to be significantly increased under a wide range of pathological conditions.39,40 In addition, the presence of NO2Tyr in human samples, both as a free amino acid and in proteins, was also demonstrated by mass spectrometry and other analytical techniques.16,41−45 Despite that protein NO2Tyr is established as the most relevant footprint of peroxynitrite in biological systems, it became apparent early in the field, that NO2Tyr can be also formed by nitrating species generated by peroxynitriteindependent pathways46,47 and that the sole detection of protein NO2Tyr does not necessarily equal peroxynitrite formation. There is currently total consensus that a combination of experimental approaches and measurements is required to ultimately reveal the contribution of peroxynitrite versus other nitrating systems during the generation of protein NO2Tyr in specific pathological conditions.44,48,49

compounds such as pyrogallol, that are able to generate O2•−, also inhibited EDRF-induced vascular relaxation.29 The O2•−mediated inactivation of EDRF led the groups of Moncada20 and Ignarro21 to ultimately identify EDRF as •NO. In the following years, several publications confirmed the relevance of reaction 1 in the context of vascular physiology using diverse experimental approaches. For example, when a sodium nitrite solution (containing glucose) was irradiated with UV-light immediately before perfusing a vessel, it induced vascular relaxation only if the irradiation was performed in the presence of SOD.30 The UV-irradiation induced the formation of •NO and hydroxyl radical (•OH).31 The reaction of glucose (Glc) with •OH produces then at least six different radicals by hydrogen abstraction (Glc•, eq 5),32 some of Glc• add O233 and eliminate hydroperoxyl radicals (HO2•) [i.e., the conjugate acid of O2•− also yielding a number of isomers of oxidized glucose (Glcox)]. These series of reactions provided the precursors for peroxynitrite (eq 1) and strongly diminished the half-life of • NO, preventing its arrival to the vessel-relaxation chamber. hv

glucose

NO2− + H+ → •NO + •OH ⎯⎯⎯⎯⎯⎯→ •NO + Glc•

(5)

Glc• + O2 → GlcO2 → Glcox + HO2

(6)

In these experiments, peroxynitrite was being formed secondary to photochemical rather than enzymatic reactions of •NO and O2•− production. The physiologically relevant role of O2•− in controlling EDRF half-life was clear even before knowing the chemical structure of EDRF. Once EDRF was demonstrated to be effectively •NO, the mechanism behind the inactivation observed was readily understood since reaction 1 had been previously characterized.1 However, none of these studies suggested that apart from diminishing •NO bioavailability, a product of this reaction could be biologically active. Indeed, at times the reaction of eq 1 was simplified in the physiological literature as directly leading to the formation of relatively inert nitrate (NO3−) originally regarded as the O2•−-mediated inactivation form of •NO,26 and the transient formation of peroxynitrite was not appreciated. The formation of peroxynitrite in biological systems and its potential cytotoxic role was originally postulated by Beckman et al.2 and further expanded and rationalized by Radi et al.7,8 In Beckman et al., the authors demonstrated that peroxynitrite decomposition can generate a strong oxidant with a reactivity similar to that of hydroxyl radical (•OH). They also proposed for the first time that SOD may protect vascular tissue under pathological conditions not only by preventing •NO consumption but also by preventing the formation of peroxynitrite.2 Immediately after this seminal first observation, the direct reactivity of peroxynitrite with biomolecules was established and a formal scheme showing alternative mechanisms of O2•−mediated toxicity in the presence and absence of •NO was presented.7,8 These studies also indicated that peroxynitrite could promote oxidation reactions in biological systems by direct bimolecular reactions or mediated by secondary radicals including •OH and nitrogen dioxide (•NO2). The proposed formation of •NO2 anticipated the possibility of nitration reactions associated with the biochemistry of peroxynitrite, processes that were in fact established in the next years. While the initial reports on the biochemistry of peroxynitrite implied the use of pure and concentrated peroxynitrite solutions, most of the initial results were recapitulated using chemical systems

1.2. The Question of Homolysis

The homolysis of peroxynitrous acid was the first reaction indicated as responsible for a cytotoxic role of peroxynitrite.2 The reaction happens spontaneously at neutral or acidic pH and more slowly in alkaline medium since peroxynitrous is a weak acid (pKa = 6.8). 1340

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

step in the formation of protein NO2Tyr from peroxynitrite, and until the late 1990’s homolysis was the main known source of radicals. The study of the reaction with CO2 and subsequent characterization of the reaction products as carbonate radical (CO3•−) and •NO267 greatly undermined the quantitative relevance of homolysis in biochemistry. Finally, the discovery of thiol and selenol (R-SeH) peroxidases48,68,69 as fast and specialized systems for the reduction of peroxynitrite further diminished the biological significance of homolysis. In summary, homolysis is too slow (1.13 s−1 at pH 7.4 and 37 °C12) to compete with the reactions of CO2, thiol and selenol peroxidases and some transition metal centers, as will be discussed in the section of peroxynitrite reactions and kinetic competitiveness. Additionally, the formation of radicals needed for tyrosine nitration can come from other sources, such as the reaction of peroxynitrite with CO2 or with a number of metal centers, which may become kinetically competitive under different circumstances. Still, there remains the possibility that a minor fraction of peroxynitrite-derived •OH may initiate free radical-dependent processes and promote oxidations through chain propagation processes; this may well be the case in hydrophobic structures such as of biomembranes and lipoproteins where polyunsaturated lipids can be oxidized by peroxynitrite in vitro and in vivo7,14,70−72 in a way that resembles the actions of •OH.

In the absence of targets, ONOOH isomerizes to nitrate, with pH-independent rate constants of 0.1, 1.2, and 4.5 s−1 at 0, 25, and 37 °C, respectively.12,50 The isomerization reaction goes through the formation of a geminate pair of •OH and nitrogen dioxide (•NO2) radicals that either recombine in the cage to yield nitrate or diffuse to the bulk solution (Scheme 1).51 The yield of •OH and •NO2 in solution has been a matter Scheme 1. Homolysis of Peroxynitrous Acid, the Ratio kB/kA Indicates the Fraction of Radicals That Diffuse out of the Solvent Cage

of a very long debate; in fact, a number of authors considered the proposed reaction inconsistent with experimental findings and proposed alternate “activated forms” of peroxynitrite (different from the geminate pair).12,51 The nature and extent of radical formation, and the implications in its biological chemistry was the topic of discussion of an open forum published in Chemical Research in Toxicology in 1998 where the main contributors of the area collected and discussed data from reactivity and scavenging experiments, computational chemistry, thermochemistry, and kinetics.52−58 The question of homolysis has been previously analyzed in Chemical Reviews, in an article dealing with the chemistry of peroxynitrite.59 Many of the experiments and calculations designed to find out the identity and yield of the oxidants formed upon peroxynitrite spontaneous isomerization to nitrate produced ambiguous results; nevertheless, the discussions around the extent of homolysis and the species formed have continued on a purely inorganic chemistry interest. A recent and excellent review can be found in ref 60. To date, all the authors agree that peroxynitrite undergoes homolysis producing •OH and •NO2 radicals to some extent. While most of the community agrees that yields are ca. 30%,51,59 one group has repeatedly supported that homolysis is a rather small percentage, accounting for ca. 5% of the proton-catalyzed decomposition of peroxynitrite.60 In any case, product formation and yields from several different • OH scavengers, kinetic competition data and electron paramagnetic resonance experiments2,61−64 support the formation of •OH from the homolysis of peroxynitrite. Of particular interest for this review, the formation of NO2Tyr (and other oxidized forms of tyrosine) requires the intermediacy of radicals arising from peroxynitrite such as • OH and •NO2, as initially proposed in the chemical18,65 and, later, in the biochemical literature.10 Despite a long chemical debate, the homolytic pathway of peroxynitrite decay turned, after all, to be a very minor route in biological contexts and therefore of lesser quantitative and biological relevance than originally thought.4,66 Indeed, several of the authors pointed out that many reactions consuming peroxynitrite were much faster than the isomerization to nitrate. As new and faster targets were discovered, the amount of possible radicals being formed from homolysis grew smaller only considering the kinetic competition. The potential relevance of homolysis as a process accounting for the oxidative biochemistry of peroxynitrite in biological systems diminished substantially when products of the reaction with CO2 were finally characterized in 1999.67 Radical formation was a required

1.3. Biological Formation of Peroxynitrite from Nitric Oxide and Superoxide

The formation of peroxynitrite in vivo from •NO and O2•− is kinetically favored as the reaction rate constant, determined by several methodologies, is in the range of 4−16 × 109 M−1 s−1.3,5,6,73 This very large kinetic constant explains how peroxynitrite can be formed at significant rates even when the concentration of its radical precursors is low, and they, in turn, can undergo alternative reactions. The rate of peroxynitrite formation can be established as d[ONOO−] = k[•NO][O•− 2 ] dt

(7) •

Due to the favored diffusion of NO across cellular membranes74 and its longer half-life in comparison to O2•−, it can be easily predicted that peroxynitrite will be formed in the proximity of O2•− formation sites upon •NO arrival. Indeed, the anionic character of O2•− (pKa = 4.875) and its low membrane permeability confine it to the compartment where it is formed. In contrast, •NO is a small, uncharged, and nearly nonpolar molecule that freely diffuses through cellular membranes. The main constraints to the formation of peroxynitrite in biological systems that cogenerate •NO and O2•− are the pathways that consume or degrade its radical precursors. The steady-state level of O2•− in any biological compartment will depend on its rate of production and also on its decay rate, the latter largely determined by the enzyme superoxide dismutase (SOD) (eq 3). SOD is found in practically all mammalian cell compartments, at concentrations of approximately 4−40 μM Cu,Zn-SOD and 1−30 μM MnSOD in the cytosol and mitochondria, respectively.76−78 In addition, there is an extracellular isoform (EC-SOD) that shows some sequence homology to the cytosolic Cu,Zn-SOD but is specialized to function in the extracellular spaces. EC-SOD is glycosylated and secreted by the cells that produce it and binds to the extracellular matrix.79 SODs are also largely distributed in microorganisms79,80 and may decrease microbicidal perox1341

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

ynitrite levels.81 All SODs catalyze the dismutation of O2•− with rate constants around (1−2) × 109 M−1 s−182−84 and effectively decrease its steady-state concentration.79,85 Although rate constant of eq 1 is larger than that of O2•− dismutation by SOD, the local SOD concentration is a key determinant of the amount of peroxynitrite that can be formed in biological systems. On the other hand, the fact that •NO can easily permeate through lipid membranes86,87 determines that its diffusion out of the compartment where O2•− is produced can become another limiting factor of peroxynitrite formation. Nitric oxide can be consumed intracellularly by reaction with other molecules or cell targets that strongly decrease the •NO half-life. In accordance with the measured rate of •NO consumption by parenchymal cells, the extravascular half-life of •NO ranges from 0.09 to >2 s, depending mainly on O2 concentration and the distance from a vessel.88 However, in the intravascular compartment •NO half-life is strongly decreased by its fast reaction with oxyhemoglobin (k = 8 × 107 M−1 s−189) that yields nitrate as the final product. The fact that hemoglobin is encapsulated slows the •NO reaction a thousand times; nevertheless, the half-life of •NO in whole blood is as short as 1.8 ms.90 In all tissues, the vasculature constitutes a major sink of •NO establishing a concentration gradient from the site of formation to the nearest vessel and hence preventing •NO accumulation in cells and tissues. The influence of the main draining pathways of •NO and O2•− (i.e., diffusion toward the vascular compartment and reaction with SOD, respectively) on peroxynitrite formation under different biological contexts is not always intuitive. For example, the effect of SOD on the steady-state concentration of • NO and on the rate of peroxynitrite production when O2•− and •NO are both being produced at comparable rates has been the subject of debate in the field for a long time. Intuition suggests that SOD will directly compete with •NO for O2•− and will therefore decrease the production of peroxynitrite.91 But, as SOD lowers the steady state level of O2•−, more •NO would be available and the product [•NO] × [O2•−] would not substantially be affected. The ability of SOD to decrease the rate of peroxynitrite production, from eq 3, will depend upon the availability of pathways that consume •NO independently of O2•−.92 Simple algebraic equations and steady state assumptions applied to model systems that considered only the presence of reactants and scavengers showed that, in general terms, the presence of SOD modestly limits the formation of peroxynitrite.91 These simplified systems did not take into account the impact of •NO diffusion, which prompted additional work emphasizing the effect of spatial distribution on peroxynitrite formation.93−96 In a mathematical model that considers a cylindrical geometry with five different concentric regions to simulate a blood vessel, the formation and chemical reactions of the peroxynitrite radical precursors •NO and O2•−93 and also the diffusion properties of •NO as well as oxygen partial pressure gradients,93 it was possible to quantitate how competition between O2•− scavenging by •NO and the SOD-catalyzed removal varied spatially. An interesting prediction of this work was that while in the vascular compartment, the highest •NO levels are expected to be near the endothelium; the maximum peroxynitrite concentration occurs deeper in the vascular wall or perivascular tissue. Another computer-assisted kinetic model to simulate peroxynitrite formation and peroxynitrite-dependent tyrosine

nitration under biological conditions was developed.96 In this model, •NO diffusion outside the compartment where it is being formed was also included as an additional kinetic process partially responsible for •NO disappearance. The rate of consumption due to diffusion depended mainly in the distance between the compartment and the nearest vessel where •NO concentration decayed to zero; but the model could also be applied to other conditions where •NO concentration decay was due to cell consumption. It was concluded that when SODcatalyzed O2•− dismutation and •NO decay due to diffusion are incorporated in the model, the steady-state concentrations of • NO and O2•− do not significantly change with variations in the relative fluxes of formation of the radicals. The absence of sensitive probes that directly react with peroxynitrite have made difficult the validation of these theoretical models until recently. In fact, most of the experimental studies analyzing peroxynitrite formation yields from fluxes of •NO and O2•− relied on the measurement of reactions of peroxynitrite derived radicals, such as NO2Tyr. The use of NO2Tyr as a marker of peroxynitrite formation has provoked a series of confounding interpretations regarding the yields of peroxynitrite formation under different fluxes of • NO and O2•−. Studies in biochemical systems varying the fluxes of either •NO or O2•− yielded a bell-shaped profile with maximun NO2Tyr yield at equimolar fluxes of both precursors. These observations were initially hard to reconcile with in vivo data where NO2Tyr yields seemed to increase under conditions where either O2•− or •NO formation was increased and even argued against peroxynitrite as the mediator of tyrosine nitration in vivo.97 However, this biphasic response with maximum yields of NO2Tyr at equimolar ratios is lost when the role of the draining pathways of •NO and O2•− present in cells and tissues are considered. Computer-assisted simulations demonstrated that the inhibition of NO2Tyr yields under excess of formation of any of its precursors was due to the competing reactions of both •NO and O2•− with radical intermediates of the nitration mechanism, mainly with tyrosyl radicals and •NO2. Nevertheless, when considering heterogeneous and compartmentalized biological systems in the presence of catalytic sinks and transmembrane diffusion that allow elimination of the excess precursor radicals, •NO and O2•− concentrations are much lower and resistant to variations. In those cases, competing reactions with nitration intermediates are not significant,96 and the decrease in nitration yields with an excess of precursor flux is not observed. Similar inhibitory responses of target oxidation in vitro by variation of O2•− or • NO fluxes were observed during lipid peroxidation98 and probe oxidation studies97,99−101 in processes that are initiated by peroxynitrite-derived radicals. The recent incorporation of boronate-derived probes that react fast and stoichiometrically with peroxynitrite102,103 (Scheme 2) provided strong evidence that peroxynitrite formation increased linearly and reached a plateau with a 1:1 ratio of cogenerated •NO and O2•− fluxes.102 In a recent study, using boronate-derived probes in a model of murine macrophages activated for •NO and O2•− production, the effect of • NO diffusion and O2•− compartmentalization on peroxynitrite formation was addressed experimentally104 and confirmed that the bell-shaped profiles obtained in vitro when NO2Tyr was used as a marker of peroxynitrite formation were a confounding observation not extrapolable to in vivo systems. In addition to the boronate-based chemical probes, genetically encoded 1342

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Since O2 has a triplet ground state, its reaction with 3NO− is spin-allowed and forms peroxynitrite with a rate constant of 2.7 × 109 M−1 s−1.115 Nevertheless, the slow deprotonation of HNO determines that, in biological systems, 3NO− will not be formed, and reactions of HNO with suitable targets will prevail. The situation could be different in the presence of systems with the potential to generate 3NO− in the presence of oxygen, but biological studies involving nitroxyl donors (e.g., Angeli’s salt) most likely involve the protonated HNO species and lead to reactions only related to HNO. In contrast to the fast reaction of 3O2 with 3NO−, the reaction with 1HNO is spin-forbidden and relatively slow. The rate constant was estimated to be 8 × 103 M−1 s−1118 and 3 × 103 M−1 s−1,119 while a more recent determination gave a slightly higher value of 1.8−2 × 104 M−1 s−1 at pH 7.4.120 The species that is formed from the reaction of HNO and O2 has been the subject of some controversy, and it has been argued that a species distinct from peroxynitrite is formed.119,121 Nevertheless, a recent study using boronate-based probes provided strong evidence for the formation of authentic peroxynitrite.120 Still, the biological relevance of the formation of peroxynitrite from the reaction between HNO and O2 can be questioned on kinetic grounds. From the rate constant and tissue O2 concentrations, which are relatively low, and from the presence of high concentrations of alternative targets (e.g., thiols and metal centers) that react with HNO with high rate constants, it becomes clear that formation of peroxynitrite in vivo from the reaction between HNO and O2 is negligible. For example, the reaction between HNO and glutathione, which is present in millimolar concentrations inside cells, has a rate constant of 3.1 × 106 M−1 s−1.120 In addition, HNO can dimerize yielding nitrous oxide (N2O) (8 × 106 M−1 s−1).117 Thus, the reaction between HNO and O2 to form peroxynitrite may only be of some significance in vivo in well-oxygenated systems with decreased thiol content; although, undoubtedly, this reaction needs to be taken into account in in vitro studies.120 In summary, although the reaction of O2 with triplet NO− to form peroxynitrite is very fast, the presence of triplet NO− in biological systems is improbable. The formation of singlet HNO is more probable, but its reaction with O2 is spinforbidden and relatively slow, making formation of peroxynitrite through this pathway in vivo unlikely.

Scheme 2. Boronate-Based Probes Detection of Peroxynitritea

a

The probe is an arylboronate that may be forming a cyclic 1,3,2 dioxaborolane structure through a bond between R1 and R2. Upon reaction with peroxynitrite, the reporter/tag moiety yields a phenolic compound suitable to spectroscopic or affinity detection. The secondorder rate constant of peroxynitrite with these probes is on the order of 106 M−1 s−1 at pH 7.4 and 37 °C. The probes also detect other oxidants such as hydroperoxides and hypochlorite. Importantly, the rate constant of peroxynitrite with, for example, fluorescein boronate is about a million times larger than that with H2O2 and 2700 times larger than that of HOCl; these significantly different rate constant values provide a large kinetic advantage for peroxynitrite detection.103,107,108

versions of green fluorescent protein-containing the unnatural amino acid p-boronophenylalanine have been recently reported and show high specificity for peroxynitrite detection.105,106 Two thorough analyses of boronates and other chemical probes for the detection of peroxynitrite (or peroxynitrite-derived radicals) have been recently published elsewhere.107,108 1.4. Formation of Peroxynitrite from the Reaction of Nitroxyl (HNO) with Oxygen

A second potential pathway for peroxynitrite formation could be the reaction of O2 with HNO (nitroxyl, nitrosyl hydride, hydrogen oxonitrate, and azanone). The chemistry involved is complex, and the reaction is dependent on deprotonation and on spin state. There are comprehensive reviews about HNO in the literature.109−113 HNO is the one-electron reduction product of •NO (E°′ • NO, H+/HNO ≈ − 0.55 V at pH 7).114,115 HNO can be exogenously administered with donor compounds such as Angeli’s salt and Piloty’s acid. It is also possible that it is formed endogenously and exerts signaling effects; this possibility still needs to be unequivocally demonstrated. The pathways proposed to be involved in HNO formation in vivo have been recently reviewed116 and include the incomplete oxidation of arginine by nitric oxide synthase, the oxidation of hydroxylamines by heme proteins, and the reduction of •NO by enzymes (SOD, xanthine oxidase), thiols, hydrogen sulfide, vitamin C, and aromatic alcohols. The ground state of HNO is a singlet, but its conjugated base [NO−, nitroxyl anion, oxonitrate (1−)] has a triplet ground state with two unpaired electrons and is isoelectronic with O2. The reduction of •NO to triplet NO− is even more unfavorable than the reduction to HNO, since the E° (NO•/3NO−) is ∼ −0.8 V.114,115 The pKa of HNO was reevaluated and is 11.4, not 4.7 as previously thought.114,115 This has important biological consequences, since it reveals that HNO is the only species that is present at physiological pH. Furthermore, the acid−base equilibrium of HNO is a slow process since it is spin forbidden; it can be represented by eq 8. HNO + OH− ⇌ 3 NO− + H 2O

1

1.5. Formation of Peroxynitrite on Metal Centers

One possible route of biological peroxynitrite formation that has attracted some attention over the last decades involves the reaction of the radicals •NO and O2•− with one of them being bound to a metal center. As early as 1981, it was proposed that the reaction of oxymyoglobin and oxyhemoglobin with •NO produced an O-bound peroxynitrito complex (Scheme 3a).122 Peroxynitrite thus formed has been invoked in several studies of • NO oxidation by these proteins.122−126 As a matter of fact, it has been proposed that the •NO dioxygenase activity of oxymyoglobin and oxyhemoglobin constitutes a primal function for the hemoglobin superfamily; this was suggested based on the evolution of hemoglobins from the microbial flavoHb that bacteria use as protection against •NO toxicity.127 Under this optic, the peroxynitrite formed would only be released after isomerization to nitrate which explains the strict 1:1 • NO:NO3− stoichiometry observed in the reaction.122 Spectrophotometric evidence of peroxynitrite formation was shown at pH = 9 in the reaction of oxymyoglobin with •NO, and even

(8)

In contrast to typical acid−base reactions, that are very fast, the deprotonation of HNO is relatively slow with forward and reverse rate constants for eq 8 of 4.9 × 104 M−1 s−1 and 1 × 102 s−1, respectively, because of the requirement for nuclear reorganization and spin change.117 Thus, contrary to typical conjugate pairs, HNO and NO− will not be in fast equilibrium. 1343

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Scheme 3. Metal Peroxynitrito Complexes Can Be Formed from O2 or •NO Bound to a Metal Center, Typically a Heme Protein, The Reactions Yield the (a) O-bound and (b) N-bound Complexes, Respectivelya

a In the case of proteins of the hemoglobin superfamily, the complex catalyzes the isomerization of O-bound peroxynitrite to nitrate prior to the dissociation. Other heme proteins can sustain the homolysis to •NO2 and even promote site specific protein nitration.

some •OH release was hinted by the oxidation of ethylene to ethylene glycol when the reaction was performed under argon.123 Another possible formation of peroxynitrite bound to a metal center involves the reaction of a nitrosyl complex with oxygen that would lead to a N-bound peroxynitrito complex (Scheme 3b) that has been proposed for nitrosylmyoglobin,128 nitrosylcobalamin,129 and more recently cystathionine β-synthase.130 Most of the reported peroxynitrito compounds have been spectroscopically characterized to different extents. Additionally, there are several examples of metal peroxynitrito compounds coming from the bioinorganic chemistry literature.131 One crucial point to consider in these potential alternative routes to peroxynitrite formation is that in many cases, metal bound peroxynitrite tends to isomerize or homolyze before dissociating from the metal, thus leading to a different kind of chemistry, as has been recently pointed out.132 Finally, there is at least one report of an enzymatic nitration reaction specifically catalyzed from peroxynitrite formed in situ, O-bound to a heme thiolate in a Streptomyces spp. cytochrome P450.133

eliminated by adding manganese dioxide (MnO2), which catalytically decomposes hydrogen peroxide. It is also important to control the preparations for the level of contaminant nitrite, because it can by itself generate confounding results; on one hand, nitrite can react with peroxynitrite-derived •OH and yield •NO2 (in fact excess NO2− can be used as a strategy to convert all radicals arising from peroxynitrite homolysis to •NO2144) or have bioactive effects.145−152 Thus, ideally, peroxynitrite preparations for biochemical and cellular experiments should contain negligible amounts of H2O2 (after MnO2 treatment), NO2− below 10% of peroxynitrite content and low carbonate contamination from air exposure of the alkaline stock solution. To control for the effect of contaminants, it is good practice to perform the so-called “reverse order of addition experiments”, where peroxynitrite is decomposed in the buffer prior to the experiments. One major issue in the experimental design involving peroxynitrite is the selection of the buffer system, phosphate being the first choice. Many of the buffers utilized for performing biochemical assays such as tris and hepes can either react with peroxynitrite or peroxynitrite-derived radicals and yield products (e.g., H2O2, nitroso compounds, •NO donors)153−156 that can secondarily affect the experimental outcome. Thus, the use of organic buffers should be avoided if at all possible. The concentration of peroxynitrite anion can be conveniently followed at 302 nm, and therefore alkaline stock solutions are utilized for peroxynitrite quantitation previous to experiments. The absorption coefficient of peroxynitrite anion was initially measured to be ε = 1670 ± 50 M−1 cm−1.157 The value was later confirmed to be precisely ε = 1700 ± 10 M−1 cm−1.143 1.6.1. Peroxynitrite Administration and Exposure in Experimental Setups. Peroxynitrite is labile, as discussed previously, even if alone in neutral solution it decays by homolysis/isomerization to nitrate (Scheme 1). In the laboratory, the exposure of biochemical systems to peroxynitrite can be performed through the addition of peroxynitrite itself, either as a bolus or as a continuous infusion, or through the simultaneous generation of its precursor radicals, O2•− and • NO. The final outcome may turn out to be different due to the possible involvement of free radicals in parallel processes. The bolus introduction of a high (sometimes millimolar) concentration of peroxynitrite can, in the absence of efficient scavengers, lead to relatively high concentrations of radicals and hence to high rates of radical−radical termination reactions in comparison to competing reactions of the free radicals with nonradical targets. In contrast, the slow infusion of peroxynitrite leads to lower steady-state free radical concentrations and hence to lower rates of radical−radical reactions.

1.6. Peroxynitrite Synthesis, Quantitation, and Handling in the Laboratory

Several procedures for synthesizing peroxynitrite in the laboratory have been designed (a) from acidified hydrogen peroxide and nitrite,2,134−136 (b) from the reaction of azide and ozone,137 (c) from the autoxidation of hydroxylamine,138 (d) from alkaline hydrogen peroxide and alkylnitrites,139,140 (e) from •NO gas and solid KO2 in sand,141 and (f) from •NO gas and tetramethylammonium superoxide in liquid ammonia.142,143 This last method stands out as the cleanest one. It produces solid tetramethylammonium peroxynitrite free of contaminants. The first method mentioned, the synthesis from acidified hydrogen peroxide and nitrite, is the method more often used. This method involves the nitrosation of hydrogen peroxide according to eqs 9 and 10. HNO2 + H+ ⇌ H 2O + NO+

NO+ + H 2O2 → ONOOH + H+

(9) (10)

The reaction is carried out by mixing hydrogen peroxide in hydrochloric acid with sodium nitrite in a flow reactor with efficient mixers. Since the peroxynitrous acid formed is unstable, the reaction is quenched with sodium hydroxide. Peroxynitrite anion is obtained in variable yields and contains sodium chloride, sodium hydroxide, hydrogen peroxide, nitrite, and nitrate as contaminants. Hydrogen peroxide can be 1344

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Figure 1. Different scenarios cause different exposure to peroxynitrite measured as the area under the curve (AUC) of the time course of the concentration. The decay of 1 mM peroxynitrite in phosphate buffer pH = 7.4 and 25 °C produces an exposure of 3.85 mM s in a spectator target (A). If the solution contains 1.3 mM CO2 then the decay is much faster and the exposure reduces to 25.7 μM s (B). A flux of 16.7 μM/s during 10 min, equivalent in total amount to the 1 mM bolus, produces a slightly larger exposure of 3.96 mM s (C). In microheterogeneous systems such as cells in suspension or liposomes, exposure inside and outside differ due to permeation of the membrane and different targets inside and outside the cell. A 1% (V/V) cell suspension mimicking the size of T. cruzi (2 μm average radius) subject to a 1 mM bolus of peroxynitrite in inert buffer at pH 7.4 has an outside exposure of 1.8 mM s (D), whereas the inside experiences an exposure of only 1.12 μM s, inset, detail of the buildup of peroxynitrite in the first milliseconds after the bolus (E). Numerical integration of the peroxynitrite time course inside vesicles of different sizes exposed to bolus of 1 mM (black ■) or 100 μM (blue ■) or fluxes of equivalent amount (1.67 μM/s during 10 min green ● or 0.167 μM/s during 100 min red ●). As a comparison, the measured values of peroxynitrite exposure inside an ionomycine-activated endothelial cell during 90 min (blue ▲) and inside T. cruzi cell phagocytosed by a macrophage during a 90 min respiratory burst (green ▼). The independent variable kD[Ves] is inversely proportional to the square root of the radius of the vesicle, representative sizes of cell types and liposomes are depicted as vertical lines.

μM/s during 10 min would be needed to approximately equalize the exposure of the 1 mM bolus, implying that a continuous flow is slightly more efficient than a bolus in terms of exposure (Figure 1C). That is for a homogeneous system, a microheterogeneous system such as a cell suspension represents additional complexity. Three processes are important in such a situation: diffusion, permeation, and chemical reactions. The rate of diffusion-controlled encounters between peroxynitrite and liposomes or cells in suspension can be estimated using the Einstein−Smoluchowski model159 (eq 13)

Measuring the exposure of a target to peroxynitrite depends substantially on the experimental setup, the pH, and the presence of other targets. To compare different scenarios, we can assume a spectator target that could consume a minimal fraction of total peroxynitrite and with a slow rate to avoid altering the decay. The simplest case would be a solution of peroxynitrite in pH = 7.4 buffer that only considers the decay via homolysis/isomerization to nitrate. Under those conditions, a single bolus of 1 mM peroxynitrite decays exponentially with k = 0.26 s−1 at 25 °C according to eq 11: [ONOO−]t = [ONOO−]0 exp(−kt )

(11)

kD = 4000πNrABDAB

The exposure (i.e., the concentration experienced by the spectator along the decay) can be calculated as the integral of peroxynitrite concentration as function on time158 (eq 12).

∫0



[ONOO−] dt =

−[ONOO−]0 exp(−kt ) k



= 0

(13)

where kD is the rate constant of diffusion-controlled encounters, N is the Avogadro constant, rAB is the sum of the radii of the cell and peroxynitrite (essentially equal to the average radius of the cell), and DAB is the sum of the diffusion coefficients of the cell and peroxynitrite (essentially equal to the diffusion coefficient of peroxynitrite). By multiplying kD times the concentration of vesicles (liposomes, cells, etc.) a pseudo-first order decay constant is obtained, kD[Ves]. Permeation is more complex, it implies the permeability constant PONOOH of peroxynitrous acid through a given barrier (e.g., a cell membrane) competing with desorption (i.e., the rate at which peroxynitrite can go back to the bulk solution once it reached the membrane interface). There are a few determinations of permeability on liposomes in the literature all around PONOOH = 1 × 10−3 cm/s,160−162 that, combined with a membrane width of 5 nm, yields a transmembrane diffusion

[ONOO−]0 k

(12)

The integral (i.e., the area under the curve of the time course) is then 3.85 mM s (Figure 1A). Just in order to compare, if the same experiment is made in the presence of 1.3 mM CO2 to intercept the peroxynitrite, then k = kapp[CO2] = 39 s−1 and the exposure of the spectator only 25.7 μM s (Figure 1B). If peroxynitrite is added as a continuous flow injection, the function describing the time course is not so simple but can be simulated in any computational kinetics program and numerically integrated. A simple calculation tells that an influx of 1.5 1345

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

constant kt of 2 × 103 s−1. The desorption rate constant kDes has been determined based on experimental data of permeation by the group of Hurst.160 Finally, as there are two compartments (outside and inside), two sets of chemical reactions need to be considered. For the sake of simplicity, the assumption is made that peroxynitrite only decays via homolysis/isomerization outside and via a pseudo-first-order process inside that involves the reaction with carbon dioxide and the most relevant catalytic antioxidants such as peroxiredoxins and glutathione peroxidases (see below). The overall scheme is depicted in Scheme 4 based on a similar approach previously published.160

Another interesting feature is that the exposure becomes independent of diffusion when the vesicles or cells are small enough and concentrated enough as to make kD[Ves] ≫ kiso. On the other hand, large cells such as granulocytes are in smaller numbers even when occupying 1% of the volume (ca. 5.6 × 106 cells/mL) and diffusion becomes a limiting factor. It has been repeatedly observed in the literature that extracellular CO2 actually protects cells from exposure to peroxynitrite.165−169 In the light of the diffusional-kinetic model, it is easy to see that anything that consumes peroxynitrite extracellularly and transforms it into a shorterlived oxidant (e.g., CO2 yielding CO3•−) would impose an additional competition to the diffusional encounter, thus diminishing the amount of ONOOH actually reaching the cell and permeating the membrane. Thus, CO2 in vivo helps focus the reactivity of ONOO− to a very narrow region. In this context, CO2 becomes an “extracellular protectant” in systems where diffusion is a contender, such as cell suspensions or bacteria in plasma. On the other hand, in phagocytosed cells, distances are too small for chemical reactions to compete with diffusion and, in those cases, extracellular CO2 would not protect from exposure to ONOOH and may change the type of extracellular oxidants to include CO3•−. A large number of caveats must be considered after this mathematical consideration of the diffusion−kinetic fate of peroxynitrite, the most important are (1) only peroxynitrite is taken into account, the exposure to its decay products (radicals, nitrite, etc.) is never considered. For example, •NO2 readily permeates membranes170 and may enhance transmembrane oxidation. (2) Transmembrane permeation of ONOOH is taken as equal to that measured in liposomes, permeation of ONOO− is not considered even when it is known that ONOO− can enter cells through anion channels;171,172 additionally, membranes are taken as symmetrical, with kt equal in both directions. (3) The consumption of peroxynitrite by antioxidant enzymes is considered to be constant even in cases where the oxidative insult is sustained and prolonged, (i.e., phagocytosed T. cruzi during 90 min of respiratory burst).173 (4) Damaging effect on the membrane that may alter is permeability, (e.g., as a byproduct of extracellular •OH and • NO2 production) is not considered. (5) This approximate model can only estimate the effect of ONOOH to cells in suspension, adherent cells constitute a different diffusional challenge that was not approached here. Aside from all these considerations, the numerical calculations provide a rough estimate of the exposure inside a suspended cell subject to a bolus or a flux of peroxynitrite and the values obtained are on the same order of magnitude of experimental values of cell-produced peroxynitrite.103 As a conclusion, a bolus of 1 mM peroxynitrite, that could be considered very large for more stable oxidants such as H2O2 or HOCl, is within the range of biological exposure to peroxynitrite experienced in cell models.

Scheme 4. Pathways of Peroxynitrite Diffusion, Permeation, and Reaction in and out of a Cell or Vesicle in Suspensiona

a kiso is the rate constant of homolysis/isomerization at a given pH, kD[Ves] is the diffusion-controlled encounter rate constant times the concentration of vesicles, kDes is the desorption rate constant,160 and kapp[T] is the pseudo-first order rate constant of peroxynitrite consumption inside a cell, a typical value being 500 s−1.66,163 kt is the transmembrane permeation rate constant, obtained as kt = PONOOH/d, where d is the bilayer width, typically 3−5 nm.

The kinetic simulations of the fate of peroxynitrite in such a system can be done considering different boluses and fluxes. To illustrate the point, a further assumption is made that a constant volume of the system (1%) is occupied by the cells or vesicles with variations in their size, to mimic different cell types. Several studies have exposed different cells in suspension to peroxynitrite boluses, for instance, red blood cells164 and Trypanosoma cruzi.158 As an in depth example, we will discuss the situation assuming a 2 μM-radius cell in suspension, mimicking T. cruzi, a model that has been studied in our laboratory.165 A 1% volume suspension of such cells equals a count of 3.2 × 108 cells per milliliter or 538 fM. When exposed to a 1 mM bolus of peroxynitrite (pH 7.4, 25 °C), approximately half is consumed outside by homolysis/ isomerization (outside exposure 1.81 mM s, Figure 1D). The other half is consumed inside by enzymatic antioxidants, mainly enzymatically by peroxiredoxins, and nonenzymatically by CO2, with an aggregate kapp[T] ≈ 500 s−1.66,103,163 Therefore, owing to the faster consumption inside, the exposure of a spectator target inside would be much lower (1.5 μM s) than outside (Figure 1E). Comparison of those estimates with previously measured intracellular data of peroxynitrite103 in terms of exposure shows that a 1 mM bolus of peroxynitrite in pH 7.4 inert buffer is within the same order of magnitude to the situation encountered in ionomycin-stimulated endothelial cells and well below the exposure that T. cruzi undergoes inside the phagosome of activated macrophages103 (Figure 1F).

2. REACTIONS OF PEROXYNITRITE IN BIOCHEMISTRY Three types of reactions can encompass the whole biochemistry of peroxynitrite. It can act as an oxidant by bimolecular nucleophilic substition (SN2) and electron transfer (ET) mechanisms, or it can act as a Lewis base to form transient Lewis adducts that have a reactivity of their own. This minimalistic summary veils all the details and complexity but serves as a starting point. As we will soon see, the chemical mechanisms of the reactions are not essential but the rates at 1346

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

2.1. Kinetically Competitive Reactions

which the reactions occur should be the cornerstone of the discussion of peroxynitrite biochemistry. Almost since its emergence as a biomolecule, targets for peroxynitrite reactions accumulated. A quick glance to the literature in the early 1990’s reveals numerous molecules that either reacted with or were somehow modified by peroxynitrite. This ever growing list of putative targets included also a large number of artifactually modified molecules (i.e., molecules that did not react at any measurable rate with peroxynitrite but that reacted with decay products formed from it in the experimental setup). Low molecular weight compounds,8,12,174−181 proteins,8,9,182−193 and even lipids,7,34,194,195 carbohydrates,196 and nucleic acid components197−199 appeared as possible targets constituting a plethora of modifiable biomolecules. Potential biotargets for peroxynitrite were accumulating in the literature, either as possible scavengers/protective agents or as ultimate recipients of oxidative damage through which peroxynitrite could “exert its toxic effects”. Very often, studies on molecular modifications due to peroxynitrite lacked contextualization; the knowledge of the chemical composition of the biological compartment where the putative reactions take place can indicate their likelihood of happening at all. It is clear that not every molecule that could react with peroxynitrite has the same quantitative relevance as a scavenger or target, and the discerning criterion is reaction kinetics. Early studies of peroxynitrite biochemistry evidenced that faster reactions would be more likely,166,200 and a few candidates to be the main peroxynitrite consumers soon stepped forward. We will describe the current knowledge of the biological targets of peroxynitrite based on a few simple assumptions. (1) Acid−base equilibria are fast. (2) Reactions take place with rates similar as those in dilute solution. (3) Reactions consuming peroxynitrite are fast, therefore its steady-state concentration is very low. Peroxynitrite disappears from a biological compartment either via chemical reactions or by diffusion and permeation through membranes. The chemical consumption of peroxynitrite can be understood as a sum of numerous parallel reactions, each contributing to the overall consumption rate according to the equation: −

d[ONOO−] = [ONOO−]ss dt

∑ knapp[Tn]

In accordance with the current knowledge of peroxynitrite reactions, a few targets account for the vast majority of its consumption, but their presence and concentrations depend greatly on the compartment being considered. Carbon dioxide, peroxidases (both heme and selenol/thiol-dependent), and a few metalloproteins are responsible for most of peroxynitrite consumption in vivo. In the next sections, we will provide details of these chemical reactions and we will describe parameters that allow assessing the quantitative importance of a given target for the consumption of peroxynitrite. 2.2. Carbon Dioxide Is the Criteria of Kinetic Competitiveness

The reaction between peroxynitrite and carbon dioxide was first hinted in 1969 by Keith and Powell50 and later substantiated through the change in reactivity that bicarbonate buffers imposed on peroxynitrite-mediated reactions.202 Later on, the reaction was found to occur strictly between ONOO− and CO2 and not other acid−base species.178 The pH-independent rate constant for the reaction was measured as k = 3 × 104 M−1 s−1 (25 °C)178 or 5.8 × 104 M−1 s−1 (37 °C).203 The first reaction product was proposed to be the short-lived adduct ONOOCO2−,202 but this species has never been detected because through rapid homolysis, it yields nitrogen dioxide and carbonate radicals (•NO2, CO3•−). The existence of two intermediates in the reaction, namely ONOOCO2− and the radical pair, was ascertained by kinetics and yield analyses using several probes such as metal complexes, tyrosine, and iodide.204,205 The short life of ONOOCO2− implies that it is not scavengeable because its transformation to the geminate pair [CO3•−, •NO2] is faster than any bimolecular reaction. The geminate pair could recombine in the solvent cage to produce carbon dioxide and nitrate, or diffuse out as free radicals;204,206,207 the reaction rates (kA and kB in Scheme 5) of Scheme 5. Sequence of Reactions between ONOO− and CO2a

(14)

a

The rate-limiting step in the sequence is the formation of ONOOCO2−. The ratio kA/kB is approximately 2, which implies that approximately one third of the radicals are scavengeable. The fact that carbon dioxide is reformed at the end of the reaction explains why some authors refer this sequence as the CO2-catalyzed isomerization of peroxynitrite.209 Such catalysis is only possible in vitro and in the absence of scavengers of CO3•− and •NO2.

where [Tn] is the concentration of a target that reacts with an apparent second-order rate constant knapp. Therefore, the product of target concentration times its respective rate constant will represent the quantitative significance of its reaction in the context of all the concurrent reactions. Alternatively, one can define the fraction of peroxynitrite consumed by a specific target Ti as the ratio:

ΦTi =

kiapp[Ti ] ∑ knapp[Tn]

the two processes have a ratio of approximately 2:1, therefore, at most one-third of the original peroxynitrite can be trapped as nitrogen dioxide and carbonate radical. Carbonate radical as a reaction intermediate has been unambiguously identified by continuous-flow EPR.67,208 The simplest reaction sequence accounting for all experimental findings is presented in Scheme 5. The reaction with CO2 is quantitatively one of the most important in accounting for peroxynitrite consumption. The product kapp[CO2] is relatively high, and additionally, carbon dioxide is an ubiquitous target, so this reaction is ideal to set a reference value for the rest of the kinetic contributions; we have

(15)

Calculating the overall rate of peroxynitrite consumption or the fraction corresponding to each target requires a thorough knowledge of the chemical composition of the compartment of interest and of the reaction kinetics for each target. This kind of kinetic analyses has been done a number of times in the peroxynitrite literature in the past 20 years with increasing number of targets.57,66,201 1347

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

previously proposed kapp[CO2] = 60 s−1 as a benchmark value that considers cytosolic-like conditions (Figure 2).66 Never-

Both routes of equilibration are rather slow at neutral pH, but carbonic anhydrases, which are widely distributed and extremely efficient enzymes,212 catalyze reaction 18 and help the system approach equilibrium in vivo. Additionally, different intracellular compartments have different pHs, but CO2 permeates easily across membranes. Finally, there is the question of CO2 gradients due to formation and consumption sites. The assumption that all species are in equilibrium yields the line plots shown in Figure 2 for kapp and [CO2]. Those allow a first approximation to kapp[CO2] values (blue ■) in which kapp[CO2] ranges from 17 s−1 for mitochondrial matrix to 129 s−1 for Golgi. Another set of assumptions considers that mitochondria are the main source of CO2, and they typically have a net outflux of CO2 through membrane permeation,213 thus the mitochondrial [CO2] should be higher than cytosolic [CO2] to establish a favorable gradient. Under those assumptions, we could calculate a mitochondrial kapp[CO2] ≥ 72 s−1. Overall, the mitochondrial [CO2], and therefore kapp[CO2], could have variations depending on the metabolic commitment of decarboxylation reactions, urea synthesis, and mitochondrial carbonic anhydrase activity but within the aforementioned range of values (i.e., 20−70 s−1). Other targets reacting with peroxynitrite should have kapp[T] not too far below the reference values for CO2 (Figure 2) in order to be considered kinetically competitive. In the following section, we will refer to the current knowledge of peroxynitrite reactions always considering that all reactions are concurrent with the reactions in Scheme 5 and thus competing kinetically with it.

Figure 2. pH effect on CO2 concentration and the rate constants of its reaction with ONOO − . Dotted line (left axis), equilibrium concentration of CO2 based on normocapnic values (PaCO2 40 mmHg, HCO3− 25 mM). Continuous line (right axis), calculated kapp values based on pKa′(CO2) = 6.02,210 pKa (ONOOH) = 6.88, k = 5.8 × 104 M−1 s−1 (37 °C).203 Squares and the values below them represent kapp[CO2] (s−1) for three intracellular compartments of different pH.214 The value in the light blue circle assumes that CO2 equilibrates across the mitochondrial membrane and thus [CO2] is the same in the mitochondrial and cytosolic compartment.

2.3. Metal Centers in Proteins

The chemistry of peroxynitrite reacting with transition metal centers in proteins is complex and intricate both in terms of reaction rate constants and of the numerous possible products formed. Since the first published articles of peroxynitrite reaction with metal proteins,183 it was evident that even closely related proteins such as myeloperoxidase and lactoperoxidase differed in rate constant and product distribution. Other hemeproteins with a nonperoxidatic main function, such as oxyhemoglobin171,215 and oxy-myoglobin,216 also react quite fast with peroxynitrite. 2.3.1. Heme Peroxidases and Heme Thiolate Proteins. The understanding of peroxynitrous acid reacting with heme peroxidases is far from complete. According to Uniprot (www. uniprot.org), there are six proteins belonging to the XPO subfamily of human peroxidases (Entries P11678, P22079, P05164, P07202, A1KZ92, and Q92626), but only myeloperoxidase (P05164) and lactoperoxidase (P22079) have been characterized in their reaction with peroxynitrite.183,217,218 The reactions of peroxynitrite with eosinophil peroxidase (P11678), thyroid peroxidase (P07202), and the recently described vascular peroxidases (VPO1 and VPO2, Q92626 and A1KZ92, respectively) have not been studied yet, despite their relative importance in specific cell types and tissues. Additional to XPO subfamily, the peroxidase-like domain of dual oxidases (Duox1 and Duox2) may bind heme and possess peroxidase activity,219 but their interaction with peroxynitrite has yet to be explored. Catalase reacts with peroxynitrous acid in a quite fast reaction220 contrary to the first report indicating no detectable reaction.183 Finally, of the two prostaglandin G/ H synthases, the reaction with the peroxidase domain of PGHS-1 has been described.221

theless, it has some drawbacks as there is not one single value for kapp[CO2] but a range that depends on the physical chemical conditions of the compartment where the reaction takes place. As shown by Lymar and Hurst,178 the reaction occurs strictly between ONOO− and CO2, and both have acid base equilibria that affect their availability at neutral pH. The dependence of the apparent reaction rate with proton concentration is given by eq 16. ⎛ ⎞⎛ ⎞ K aONOOH [H+] ⎟ ⎜ ⎟ kapp = k ⎜ ONOOH + [H+] ⎠⎝ K aCO2 + [H+] ⎠ ⎝ Ka

(16)

where k is the pH-independent rate constant that has values of 3 × 104 M−1 s−1 (25 °C)178 or 5.8 × 104 M−1 s−1 (37 °C);203 KaONOOH is the ionization constant of peroxynitrous acid (1.58 × 10−7),8 and KaCO2 is the apparent ionization constant of carbon dioxide (9.58 × 10−7).210 The pH dependence of kapp for the reaction of ONOO− with CO2 can thus be calculated at any desired pHs. Calculation of [CO2] is somewhat harder, as a first approach we could assume that CO2 is in equilibrium with bicarbonate, an equilibrium that can be reached either by reaction with hydroxide:211 CO2 + OH− → HCO3−

k = 8.5 × 104 (KW /[H+])

(17)

Or may proceed by hydration-ionization: CO2 + H 2O ⇌ H 2CO3

k f = 0.03 s−1; k r = 20 s−1

H 2CO3 ⇌ H+ + HCO3−(fast)

K a = 5.7 × 10−4

(18) (19) 1348

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Table 1. Reactions of Metal Proteins with Peroxynitrite rate constant with peroxynitrite metal enzyme horseradish peroxidase prostaglandin endoperoxide H syntase −1 myeloperoxidase lactoperoxidase chloroperoxidase catalase M. tuberculosis KatG S. typhimurium KatG endothelial nitric oxide synthase endothelial nitric oxide synthase (oxidase domain) inducible nitric oxide synthase neuronal nitric oxide synthase Bacillus subtilis nitric oxide synthase cytochrome P450 (BM-3) Campilobacter jejuni trHemoglobin P (FeIII) cystathionine β synthase cytochrome c Fe(II) oxyhemoglobin methemoglobin oxymyoglobin metmyoglobin aconitase MnSOD FeSOD Cu,Zn-SOD

kapp (M−1 s−1)

conditions

heme product

ref

3.2 × 10 1.02 × 106 1.7 × 107 2 × 107 6.8 × 106 3.3 × 105 1.96 × 106 1.7 × 106 1.4 × 105 4.2 × 104 7.5 × 105 2.7 × 106 2.1 × 105 2.3 × 105 1.3 × 105 1.0 × 106 4.1 × 106 5 × 104 1.3 × 104 2.4 × 104 3.9 × 104 5.4 × 104 7.7 × 104 1.4 × 105 1.3 × 105 4.3−4.6 × 104 1.88 × 104

25 °C, pH indep. 25 °C, pH 7.4 8 °C, pH 7 12 °C, pH indep. 25 °C, pH 7.4 12 °C, pH 7.4 23 °C, pH 5.1 23 °C, pH 7.1 37 °C, pH 7.4 25 °C, pH 7.4 37 °C, pH 7.4 37 °C, pH 7.4 20 °C, pH 7.4 20 °C, pH 7.4 20 °C, pH 7.4 12 °C, pH 6.8 25 °C, pH indep. 37 °C, pH 7.4 25 °C, pH 7.4 37 °C, pH 7.4 20 °C, pH 7.0 20 °C, pH 7.3 20 °C, pH 7.0 25 °C, pH 7.6 37 °C, pH 7.47 37 °C, pH 7.4 37 °C, pH = 7.53

compound I

183 225 221 183,217

6

compound I compound II compound II compound II compound I

183,218 223 220 226 227 224,228

compound I Fe(III)

229 230 231 193 171 232 216 232 185 78 81 233

Fe(III) Fe(III) Fe(III)

protein, including a decrease in Vmax and increase in Km and afterward irreversible loss of electron transfer activity for the catalyzed reduction of O2 to water.234,235 The functional and structural changes in the enzyme observed under a large excess of peroxynitrite are unlikely to be of any biological relevance. With regard to the catalytic activity of peroxynitrite reduction by cytochrome c oxidase at the expense of reducing equivalents provided by the electron transport chain (i.e., by cytochrome c2+), it has been correctly pointed out that the second-order rate constant with peroxynitrite is about two-orders of magnitude lower than that governing the reaction between the enzyme and O2.235 Thus, under physiological conditions oxygen is expected to typically outcompete peroxynitrite for the reaction with reduced cytochrome c oxidase. Also, under excess • NO (which binds and inhibits the enzyme with at least one order of magnitude higher affinity than O2), the cytochrome c oxidase-nitrosyl complex is not competent for peroxynitrite reduction.238 Still, under conditions of enhanced mitochondrial fluxes of peroxynitrite, cytochrome c oxidase may partially attenuate mitochondrial oxidative damage via peroxynitrite reduction to nitrite, including the reported inhibition of the tyrosine nitration of cytochrome c.235,239 Overall, under biologically relevant conditions cytochrome c oxidase may contribute, together with other enzymatic systems such as thiol peroxidases, to the mitochondrial reduction of peroxynitrite in the inner membrane region. 2.3.3. Iron−Sulfur Cluster Proteins. Peroxynitrite can readily react with some iron−sulfur-containing proteins; in particular, the peroxynitrite-sensitive ones are the group of dehydratases that have cubane [4Fe-4S] clusters in which one of the iron atoms (Feα) is not coordinated by polypeptide

A few heme-thiolate proteins have been studied in their reaction with peroxynitrite, most notably prostacyclin synthase,222 chloroperoxidase,223 and the oxygenase domain of nitric oxide synthase.224 Table 1 presents a summary of reaction rate constants of peroxynitrite with metal proteins. Overall it can be seen that hemeperoxidases and heme-thiolate proteins containing pentacoordinated iron tend to have larger rate constants in the 105−107 M−1 s−1 range, whereas hemeproteins containing hexacoordinated iron and nonheme metal proteins have smaller values. To complete the question of competitiveness, the concentration factor has to be considered, and that is a complex problem. Some metal proteins would be extremely concentrated in specialized cell types or subcellular compartments where they may become competitive targets in the consumption of peroxynitrite. 2.3.2. Cytochrome c Oxidase. Peroxynitrite reacts with the reduced form of cytochrome c oxidase at the binuclear heme a3-CuB site in a two-electron redox process to mainly yield nitrite (and some amounts of •NO) with a rate constant ≥106 M−1 s−1 at room temperature and physiological pH.234,235 At low and moderate concentrations of peroxynitrite, cytochrome c oxidase readily decomposes peroxynitrite and this process has been invoked to support the observation of the resistance of mitochondrial complex IV to peroxynitritemediated inactivation in isolated mitochondria and cells.190,236,237 However, in vitro experiments with purified cytochrome oxidase preparations at high peroxynitrite to cytochrome c oxidase ratio (>100:1) promote changes and oxidative processes in the protein, including heme degradation, which overall cause spectral and functional changes in the 1349

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

nitration of Y151 and Y472, overoxidation of C126 and C385 and also S-glutathionylation of aconitase. Still, in spite of these possible oxidative posttranslational modifications, the main cause of aconitase inactivation is the reaction of peroxynitrite at the Feα of the active site iron−sulfur cluster. In contrast to the sensitivity of the labile [4Fe-4S] clusters to peroxynitrite, electron transfer enzymes that utilize iron−sulfur clusters with Fe atoms fully coordinated to the polypeptide chain are resistant to peroxynitrite. Indeed, [4Fe-4S], [3Fe-4S], and [2Fe-2S]-containing bacterial and mamamalian succinate dehydrogenase and NADH dehydrogenase are, in general, resistant to peroxynitrite,256,257 despite becoming nitrated in tyrosine residues. However, in the case of succinate dehydrogenase, it may become labile to peroxynitrite once the enzyme is depleted or affected at its accessory subunits that shield the clusters.258 Indeed, these additional mechanisms may serve to explain why peroxynitrite in excess affects succinate dehydrogenase activity and complex II-dependent respiration.190,236 In any case, when considering the sensitivity of iron−sulfur-containing mitochondrial enzymes to peroxynitrite, it is clear that m-aconitase is the premier target and this results in an important effect in energy metabolism due to its role in the metabolic control of the Krebs cycle.259

cysteine residues but only ligated to inorganic sulfurs of the cluster, leaving a free coordination position. This particular arrangement permits the Feα to bind a hydroxyl leaving group of the substrate to be dehydrated. The absence of a fourth cysteine ligand and an overall positive charge (2+) of the cluster makes it accessible and vulnerable to anionic oneelectron oxidants such as peroxynitrite (eq 20). Indeed,

mammalian mitochondrial and cytosolic (IRP1) aconitases,185,240 bacterial aconitase, 6-phosphogluconate dehydratase, and fumarase A241,242 are selective and early targets of peroxynitrite in vitro and in vivo. These enzymes have been also previously shown to be exquisitely sensitive to O2•−, with rate constants ca. 107 M−1 s−1.241 The reaction of O2•− and peroxynitrite with this type of iron−sulfur cluster leads to the release of the labile iron (Feα), the disruption of the cluster to [3Fe-4S] state and enzyme inactivation. Aconitases catalyze the reversible isomerization of citrate and isocitrate via cis-aconitate in the Krebs cycle. The second-order rate constant of the reaction for the mitochondrial enzyme with peroxynitrite was determined as 1.4 × 105 M−1 s−1 at pH 7.6 and 25 °C.185,243 The resulting inactive enzyme could be reactivated by reincorporation of Fe(II) to the 3Fe-4S cluster under appropriate reducing conditions in vitro and in vivo.185,242 On the other hand, •NO turned out to be a poor reactant and inactivator of m-aconitase, unless sustained levels over time were present.244 Overall, the experiments reporting the reactivity of peroxynitrite with m-aconitase185 helped to rationalize previous data on the effect of •NO synthesis or exposure in different cell types, resulting in an early loss of maconitase activity.245−249 Indeed, most of the •NO-dependent inactivation of cellular aconitases is now known to be mediated by peroxynitrite. The early reports with isolated aconitase preparations185,241 were key to expand the understanding of the cellular toxicity of •NO in terms of the participation of peroxynitrite and other •NO-derived oxidants. Later work also indicated that peroxynitrite was able to disrupt the iron−sulfur cluster of mammalian cytosolic aconitase240 (c-aconitase), also known as iron regulatory protein 1 (IRP1). The cluster-free protein (apo-IRP1) participates in the regulation of the expression of genes involved in iron metabolism. Activity, resonance Raman spectra and EPR data on human recombinant holo-IRP1 (rhIRP1) are consistent with the conversion of the [4Fe-4S]2+ to a [3Fe4S]1+ center upon exposure to peroxynitrite.250,251 Thus, peroxynitrite reactions with bacterial and mammalian aconitases lead to enzyme inactivation,185,241 iron release,242 and can also influence iron-regulatory pathways.240,251 Peroxynitrite can also cause m-aconitase tyrosine nitration in vitro243,252,253 and in vivo254,255 through the reactions of its secondary radicals; however, the generation of NO2Tyr in aconitase is not related to enzyme inactivation.243 For example, while substrates such as citrate protect m-aconitase from peroxynitrite-mediated inactivation, they do not modify the extent of enzyme nitration. Peptide mapping and mass spectrometry analysis have shown that peroxynitrite can cause

2.4. Reduction of Peroxynitrite by Thiols and Selenols

Thiols were among the first biological targets of peroxynitrite that were described.8 The reaction is a two-electron exchange with an SN2 mechanism in which the thiolate attacks peroxynitrous acid producing nitrite and the corresponding sulfenic acid (eq 21).260 The pH profiles of the apparent rate constants are described by eq 22 and are bell-shaped (Figure 3). RS− + ONOOH → RSOH + NO2−

(21)

⎞ ⎛ ⎞⎛ K aSH [H+] app ⎟ ⎜ ⎟ k pH = k RS−⎜ ONOOH + [H+] ⎠⎝ K aSH + [H+] ⎠ ⎝ Ka

(22)

The reaction of peroxynitrite with low molecular weight thiols differing in pKa reveals that the pH-independent rate

Figure 3. Simulated pH profiles of the apparent rate constants of several thiols reacting with peroxynitrite at 37 °C based in experimental data8,260 and using eq 22. The curves represent data generated for cysteine ethyl ester (black), cysteine (red), glutathione (blue), and dihydrolipoic acid (green). The vertical line marks pH 7.4 where most of the experimental determinations are made. 1350

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Table 2. Rate Constants of Peroxynitrite Reduction by Thiol-Containing Proteins protein Homo sapiens PRDX5

267,275

Xylella fastidiosa Ohr273 Mycobacterium tuberculosis AhpE285 Trypanosoma cruzi mTXNPx286 Mycobacterium tuberculosis TPx287 Plamodium falciparum AOP288 Homo sapiens PRDX2289,290 Arenicola marina Prx6291 Mycobacterium tuberculosis AhpC68 Mycobacterium tuberculosis PrxQ B292 Populus trichocarpa GPx 5272 Salmonella tiphymurium AhpC68 Helicobacter pylori AhpC68 Xylella fastidiosa PrxQ293 Trypanosoma cruzi cTXNPx286,294 Saccharomyces cerevisiae Tsa1225 Saccharomyces cerevisiae Tsa2225 Rattus norvegicus Prx6295 Giardia intestinalis Prx1a296 Giardia intestinalis ΔPrx1b296 Homo sapiens CD45297 Homo sapiens LAR297 Homo sapiens PTP1B297 Homo sapiens creatine kinase298,299 Homo sapiens DJ1300,301 Homo sapiens PDI302 Oryctolagus cuniculus GAPDH303,304 Homo sapiens arylamine N-acetyl transferase 1305 Homo sapiens tryptophan hydroxylase306 Mycobacterium tuberculosis Trx C307 Trypanosoma brucei TXN48,308 Bos taurus serum albumin8 Homo sapiens serum albuminc309,310 Pseudomonas aeruginosa DksA311 Pseudomonas aeruginosa DksA2311 cysteine (free amino acid)8 a

kappa (M−1 s−1) 7 × 10 1.2 × 108 2 × 107 1.9 × 107 1.8 × 107 1.5 × 107 1.5 × 107 1.4 × 107 2 × 106 1.51 × 106 1.4 × 106 1.4 × 106 1.33 × 106 1.21 × 106 1.04 × 106 1 × 106 7.4 × 105 5.1 × 105 3.7 × 105 4 × 105 2 × 105 2 × 108 2.3 × 107 2.2 × 107 8.85 × 105 2.7 × 105 6.9 × 104 6.25 × 104 5 × 104 3.4 × 104 1 × 104 ∼3.5 × 103 2.6 × 103 3.8 × 103 2.42 × 103 1.43 × 103 4.5 × 103 7

conditions pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH pH

7.8, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 6.75, RT 7.4, 25 °C 7.4, 25 °C 6.75, RT 6.75, RT 7.4, 37 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7, 4 °C 7, 4 °C 7.4, 37 °C 7.4, 37 °C 7.4, 37 °C 6.9, 37 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.5, 37 °C 7.4, 25 °C 7.4, 25 °C 7.4, 25 °C 7.4, 37 °C 7.4, 37 °C 7.4, 25 °C 7.4, 25 °C 7.4, 37 °C

thiol pKa 5.2 NDd 5.2 NDd NDd NDd 5−6 5.1 one NO2Tyr/10,000 Tyr, under basal conditions; levels increase at least 1 order of magnitude under inflammatory conditions). However, this modification is favored in specific proteins, and, within proteins, in specific tyrosine residues and can cause structural and/or functional changes that are biologically relevant. For example, the nitration of Tyr34 in human MnSOD causes its inactivation in vitro and in vivo, and the sitespecificity in this case is explained by the Mn-catalyzed nitration reaction at the active site Tyr34; this modification provides a highly specific indication of peroxynitrite as the proximal •NOderived oxidant. In other cases, protein tyrosine nitration causes a gain-of-function, in some cases related to toxic activities. Thus, the formation of NO2Tyr constitutes not only a footprint of peroxynitrite but also in some cases can participate in its biological effects. In spite of a significant amount of literature, some aspects that require more precise definition are how peroxynitrite and protein tyrosine nitration participate in signal transduction cascades, how nitrated proteins are repaired and degraded, and how oxidative changes in biomolecules produced by peroxynitrite really translate to changes in function with quantitative and structural support. Recent work indicates that tyrosine and tryptophan nitration can be also catalyzed enzymatically, discoveries that expand the field of protein nitration biochemistry. In summary, peroxynitrite formation and protein tyrosine nitration are continuous events that occur in many organisms under physiological and pathological conditions; their role in human physiology, pathology, and aging has been largely established. There is plenty of room to use the biochemical knowledge obtained through over 25 years of research in these aspects of redox and free radical biochemistry in the translation into rational applications to the medical field.

could still help in defining the structural features of a given section of a protein. Some of these techniques include nuclear magnetic resonance (NMR), resonance Raman, and circular dichroism (CD). Given the particular spectroscopic characteristics in the visible range of NO2Tyr, resonance Raman represents a good way for detecting nitrated tyrosine residues in proteins.11 Furthermore, it can be used to obtain some information on the microenvironment of the nitrated residue, as well as to identify the protonation state of the nitrotyrosyl residue. For proteins containing chromophore prosthetic groups, like heme proteins, resonance Raman can also give structural information about that section of the protein, so it can be applied to study how its structure and interactions are affected by the nitration of tyrosine residues of the protein.807 The combination of resonance Raman with paramagnetic NMR was used to define the already mentioned structural changes induced in cytochrome c after nitration of solvent exposed Tyr74 and how such conformational change is coupled with the deprotonation of the nitrated tyrosine residue.688,689 The structural alterations induced by nitration of tyrosine residue 103 in myoglobin were studied also by a combination of spectroscopic techniques, including NMR and CD.808 Through the NMR approach, it was noted that nitrated Mb has a very similar overall tertiary structure than its native form but with a slightly more open conformation at the heme pocket with weaker iron-ligand interactions. At the secondary structure level, CD studies revealed that the α-helix content of nitrated Mb was slightly higher than for native Mb.808 For peroxynitritetreated LDL, in which apoB-100 contains several NO2Tyr residues, CD analysis indicated that nitrated apoB-100 contained a considerably lower content of α-helix than its native form, while it showed a greater content of β-sheet and random coil structure. These observations suggested that tyrosine nitration of apoB-100 by peroxynitrite, along with other nitrooxidative modifications, promoted the partial unfolding of the protein, as the major structural consequence of peroxynitrite promoted nitration and oxidation.809 The impact of tyrosine nitration on the secondary structure of αsynuclein was also analyzed by CD spectroscopy, in order to understand the effect of α-synuclein nitration on its aggregation events. Far-UV CD spectra showed that, after nitration of its four tyrosine residues, there was an increase on the ordered secondary structure content (mainly β-sheet) of α-synuclein in comparison with the natively unfolded monomer.810 These examples indicate that, independently of the elucidation of the three-dimensional structure of a nitrated protein or the development of computational simulations, several spectroscopic techniques, especially when used in combination, can provide information about the structural consequences of protein tyrosine nitration.

5. CONCLUSIONS AND PERSPECTIVES The occurrence of peroxynitrite in biological systems and its role as a potent and short-lived oxidant was originally proposed based on mechanistic and kinetic terms and later confirmed experimentally by detection of oxidative modifications in biomolecules (i.e., thiol oxidation, protein tyrosine nitration, and lipid peroxidation) and, more recently, via detection with chemical probes (e.g., with boronates); pharmacological and genetic-engineering approaches have been instrumental for unambiguously inferring the role of peroxynitrite in biological outcomes, including its participation as culprit in different 1385

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

AUTHOR INFORMATION

Sebastián Carballal received his Ph.D. (2011, in Chemistry) at the Universidad de la República, Uruguay. Currently, he is Associate Professor of Biochemistry at the Facultad de Medicina, Universidad de República, Uruguay, and also a researcher at the Center for Free Radical and Biomedical Research, Universidad de la República, Uruguay. His research interests have focused in the field of proteins and the biochemistry of free radicals and redox processes. He has experience in a variety of biochemical procedures (protein purification, enzyme kinetics, bioanalytical methods) and in specialized techniques such as rapid kinetics approaches (stopped-flow spectroscopy). His current research is focused on the characterization of synthetic molecules with antioxidant capacity and evaluation of their cytoprotective effects during cellular challenge to nitro-oxidative stress conditions.

Corresponding Author

*Address: Departamento de Bioquı ́mica, Facultad de Medicina, Universidad de la República, Avda. General Flores 2125, 11800 Montevideo, Uruguay. E-mail: [email protected]. Tel: +598 2924 9561. ORCID

Rafael Radi: 0000-0002-1114-1875 Present Address @

N.R.: Cell Analysis Division, Agilent Technologies, Lexington, MA, United States.

Author Contributions ⊥

Natalia Romero obtained her Ph.D. in Chemistry in May 2007 working at the Department of Biochemistry, Facultad de Medicina, Universidad de la República, Uruguay. She was a postdoctoral fellow at the Brigham and Women’s Hospital−Harvard Medical School, USA, working with Dr. Thomas Michel. She has been Assistant Professor at the Department of Biochemistry, Facultad de Medicina, Universidad de la República, Uruguay, and investigator at the Center for Free Radical and Biomedical Research, Universidad de la República, Uruguay. Since 2015, she has been working as a Scientist at the Cell Analysis Division of Agilent Technologies, USA. Her research is focused on the study of reactivity and diffusional properties of reactive oxygen and nitrogen species and their role in signaling process and disease states.

G.F.-S. and N.C. contributed equally to this work.

Notes

The authors declare no competing financial interest. Biographies Gerardo Ferrer-Sueta obtained his M.Sc. in Inorganic Chemistry (UNAM, México, 1995) and his Ph.D. in Chemistry (Universidad de la República, Uruguay, 2006). He is currently an Assistant Professor of Biological Physical Chemistry at the School of Sciences, Universidad de la República. His research interests are focused on kinetics and mechanisms of redox reactions involving peroxides and thioldependent redox enzymes. He is a member of the Center for Free Radical and Biomedical Research.

Beatriz Alvarez received her M.Sc. (1993) and Ph.D. (1999) degrees in Chemistry from the Universidad de la República, Montevideo, Uruguay. She is now Associate Professor of Enzymology at the School of Sciences, Universidad de la República. She is also an investigator at the Center for Free Radical and Biomedical Research. Her research interests span the areas of redox biochemistry, kinetics, and enzymology, especially in relation to biological thiols and hydrogen sulfide.

Nicolás Campolo received his B.Sc. in Biochemistry in 2014 from the School of Sciences of the Universidad de la República, Uruguay. Since then, he has been a Ph.D. student in Biological Sciences, performing his work at the Department of Biochemistry at the School of Medicine and Center for Free Radical and Biomedical Research, Universidad de la Repúbica, Uruguay. His research focuses on the formation and biological consequences of the oxidative modifications in proteins, with a special focus on 3-nitrotyrosine and its derivatives. More specifically, he is now studying the mechanisms of the oxidative inactivation of human glutamine synthetase and its relation with pathological processes.

Rafael Radi received his M.D. (1988) and Ph.D. (1991, in Biochemistry) at the Universidad de la República, Montevideo, Uruguay. He was a postdoctoral fellow at the University of Alabama at Birmingham, USA, working with Bruce A. Freeman and Joseph S. Beckman, where he generated his first works of the biochemistry of peroxynitrite. He has been tenured faculty at the Department of Biochemistry, Facultad de Medicina, Universidad de la República, Uruguay, for three decades and is now its Professor and Chairman. He is the Director of the Center for Free Radical and Biomedical Research at the same University. Radi is a foreign member of the U.S. National Academy of Sciences and Howard Hughes Medical Institute alumni. His research interests have focused in the biochemistry of free radical and redox processes in oxidative stress and signaling, mitochondrial dysfunction, and oxidative post-translational modifications of proteins, including protein tyrosine nitration. He has applied rapid kinetic techniques, bioanalytical and immunochemical methods, electron paramagnetic resonance studies, and structural biology approaches to unravel the molecular basis of oxidative processes in disease states and characterize and evaluate redox-based therapeutics.

Madia Trujillo received her M.D. (1995) and Ph.D. in Biochemistry (2005) from Universidad de la República, Uruguay. She has been a faculty member of the Department of Biochemistry, Facultad de Medicina, Universidad de la República since 1991, where she is now an Associate Professor. She is a member of the Center for Free Radical and Biomedical Research, in Universidad de la República, Uruguay, where she leads a research line on enzymatic antioxidant systems. Particularly, she is interested in the functional and structural characterization of antioxidant systems of pathogenic microorganisms as well as human host cells. For that purpose, she applies a combination of rapid kinetic spectroscopy, electron paramagnetic resonance, as well as biophysical and biochemical analytic methods. Silvina Bartesaghi obtained her degree in Biochemistry in 2004 at the School of Sciences, Universidad de la República, and her Ph.D. in Chemistry at the School of Chemistry in 2010. Her research is focused on peroxynitrite-mediated posttranslational modifications and has worked during the last years in the mechanisms of tyrosine nitration/ oxidation in membranes. At the moment, she is studying the molecular basis of the oxidative inactivation of glutamine synthetase. Currently, she is an Associate Professor in Facultad de Medicina and investigator at the Center for Free Radical and Biomedical Research, Universidad de la República, Uruguay. She is also a Level I Investigator in the National Research System.

ACKNOWLEDGMENTS The authors are grateful to Dr. Valeria Valez for assistance with the artwork. Financial support was provided by the Universidad de la República (CSIC and Espacio Interdisciplinario, UdelaR), Agencia Nacional de Investigació n e Innovació n (FCE_2014_104233). Additional support was obtained from 1386

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(19) Koppenol, W. H. 100 Years of Peroxynitrite Chemistry and 11 Years of Peroxynitrite Biochemistry. Redox Rep. 2001, 6, 339. (20) Palmer, R. M. J.; Ferrige, A. G.; Moncada, S. Nitric Oxide Release Accounts for the Biological Activity of Endothelium-Derived Relaxing Factor. Nature 1987, 327, 524. (21) Ignarro, L. J.; Buga, G. M.; Wood, K. S.; Byrns, R. E.; Chaudhuri, G. Endothelium-Derived Relaxing Factor Produced and Released from Artery and Vein Is Nitric Oxide. Proc. Natl. Acad. Sci. U. S. A. 1987, 84, 9265. (22) Furchgott, R. F. In Mechanisms of Vasodilatation; Vanhoute, P. M., Ed.; Raven Press: New York, 1988. (23) Moncada, S.; Higgs, E. A. The Discovery of Nitric Oxide and Its Role in Vascular Biology. Br. J. Pharmacol. 2006, 147, S193. (24) Furchgott, R. F. In Nobel Lectures in Physiology or Medicine 1996−2000; Jornvall, H., Ed.; World Scientific Publishing Company, 2003. (25) Murad, F.; Waldman, S.; Molina, C.; Bennett, B.; Leitman, D. Regulation and Role of Guanylate Cyclase-Cyclic Gmp in Vascular Relaxation. Prog. Clin. Biol. Res. 1987, 249, 65. (26) Ignarro, L. J. Biosynthesis and Metabolism of EndotheliumDerived Nitric Oxide. Annu. Rev. Pharmacol. Toxicol. 1990, 30, 535. (27) Gryglewski, R. J.; Palmer, R. M.; Moncada, S. Superoxide Anion Is Involved in the Breakdown of Endothelium-Derived Vascular Relaxing Factor. Nature 1986, 320, 454. (28) Rubanyi, G. M.; Vanhoutte, P. M. Superoxide Anions and Hyperoxia Inactivate Endothelium-Derived Relaxing Factor. Am. J. Physiol. 1986, 250, H822. (29) Moncada, S.; Palmer, R. M.; Gryglewski, R. J. Mechanism of Action of Some Inhibitors of Endothelium-Derived Relaxing Factor. Proc. Natl. Acad. Sci. U. S. A. 1986, 83, 9164. (30) Matsunaga, K.; Furchgott, R. F. Interactions of Light and Sodium Nitrite in Producing Relaxation of Rabbit Aorta. J. Pharmacol. Exp. Ther. 1989, 248, 687. (31) Mack, J.; Bolton, J. R. Photochemistry of Nitrite and Nitrate in Aqueous Solution: A Review. J. Photochem. Photobiol., A 1999, 128, 1. (32) Gilbert, B. C.; King, D. M.; Thomas, C. B. Radical Reactions of Carbohydrates. Part 2. An Electron Spin Resonance Study of the Oxidation of D-Glucose and Related Compounds with the Hydroxyl Radical. J. Chem. Soc., Perkin Trans. 2 1981, 1186. (33) Bothe, E.; Schulte-Frohlinde, D.; von Sonntag, C. Radiation Chemistry of Carbohydrates. Part 16. Kinetics of HO2• Elimination from Peroxyl Radicals Derived from Glucose and Polyhydric Alcohols. J. Chem. Soc., Perkin Trans. 2 1978, 416. (34) Darley-Usmar, V. M.; Hogg, N.; O’Leary, V. J.; Wilson, M. T.; Moncada, S. The Simultaneous Generation of Superoxide and Nitric Oxide Can Initiate Lipid Peroxidation in Human Low Density Lipoprotein. Free Radical Res. Commun. 1992, 17, 9. (35) Hogg, N.; Darley-Usmar, V. M.; Wilson, M. T.; Moncada, S. Production of Hydroxyl Radicals from the Simultaneous Generation of Superoxide and Nitric Oxide. Biochem. J. 1992, 281, 419. (36) Mulligan, M. S.; Hevel, J. M.; Marletta, M. A.; Ward, P. A. Tissue Injury Caused by Deposition of Immune Complexes Is LArginine Dependent. Proc. Natl. Acad. Sci. U. S. A. 1991, 88, 6338. (37) Beckmann, J. S.; Ye, Y. Z.; Anderson, P. G.; Chen, J.; Accavitti, M. A.; Tarpey, M. M.; White, C. R. Extensive Nitration of Protein Tyrosines in Human Atherosclerosis Detected by Immunochemistry. Biol. Chem. Hoppe-Seyler 1994, 375, 81. (38) Viera, L.; Ye, Y. Z.; Estevez, A. G.; Beckman, J. S. Immunohistochemical Methods to Detect Nitrotyrosine. Methods Enzymol. 1999, 301, 373. (39) van der Loo, B.; Labugger, R.; Skepper, J. N.; Bachschmid, M.; Kilo, J.; Powell, J. M.; Palacios-Callender, M.; Erusalimsky, J. D.; Quaschning, T.; Malinski, T.; et al. Enhanced Peroxynitrite Formation Is Associated with Vascular Aging. J. Exp. Med. 2000, 192, 1731. (40) Heijnen, H. F.; van Donselaar, E.; Slot, J. W.; Fries, D. M.; Blachard-Fillion, B.; Hodara, R.; Lightfoot, R.; Polydoro, M.; Spielberg, D.; Thomson, L.; et al. Subcellular Localization of Tyrosine-Nitrated Proteins Is Dictated by Reactive Oxygen Species Generating Enzymes

Programa de Desarrollo de Ciencias Básicas (PEDECIBA), Centro de Biologı ́a Estructural del Mercosur (CeBEM), Centro Argentino Brasileño de Biotecnologı ́a (CABBIO), and Ridaline through Fundación Manuel Pérez (Facultad de Medicina, Universidad de la República). N.C. was partially supported by a fellowship from Universidad de la República (CAP_Uruguay), respectively. R.R. wishes to thank all the national and international collaborators that have jointly contributed to the topics presented in this review over the last 25 years.

REFERENCES (1) Blough, N. V.; Zafiriou, O. C. Reaction of Superoxide with Nitric Oxide to Form Peroxonitrite in Alkaline Aqueous Solution. Inorg. Chem. 1985, 24, 3502. (2) Beckman, J. S.; Beckman, T. W.; Chen, J.; Marshall, P. A.; Freeman, B. A. Apparent Hydroxyl Radical Production by Peroxynitrite: Implications for Endothelial Injury from Nitric Oxide and Superoxide. Proc. Natl. Acad. Sci. U. S. A. 1990, 87, 1620. (3) Botti, H.; Möller, M. N.; Steinmann, D.; Nauser, T.; Koppenol, W. H.; Denicola, A.; Radi, R. Distance-Dependent DiffusionControlled Reaction of *NO and O2*- at Chemical Equilibrium with ONOO-. J. Phys. Chem. B 2010, 114, 16584. (4) Radi, R. Peroxynitrite, a Stealthy Biological Oxidant. J. Biol. Chem. 2013, 288, 26464. (5) Kissner, R.; Nauser, T.; Bugnon, P.; Lye, P. G.; Koppenol, W. H. Formation and Properties of Peroxynitrite as Studied by Laser Flash Photolysis, High-Pressure Stopped-Flow Technique, and Pulse Radiolysis. Chem. Res. Toxicol. 1997, 10, 1285. (6) Goldstein, S.; Czapski, G. The Reaction of NO. With O2.- and HO2.: A Pulse Radiolysis Study. Free Radical Biol. Med. 1995, 19, 505. (7) Radi, R.; Beckman, J. S.; Bush, K. M.; Freeman, B. A. Peroxynitrite-Induced Membrane Lipid Peroxidation: The Cytotoxic Potential of Superoxide and Nitric Oxide. Arch. Biochem. Biophys. 1991, 288, 481. (8) Radi, R.; Beckman, J. S.; Bush, K. M.; Freeman, B. A. Peroxynitrite Oxidation of Sulfhydryls. The Cytotoxic Potential of Superoxide and Nitric Oxide. J. Biol. Chem. 1991, 266, 4244. (9) Beckman, J. S.; Ischiropoulos, H.; Zhu, L.; van der Woerd, M.; Smith, C.; Chen, J.; Harrison, J.; Martin, J. C.; Tsai, M. Kinetics of Superoxide Dismutase- and Iron-Catalyzed Nitration of Phenolics by Peroxynitrite. Arch. Biochem. Biophys. 1992, 298, 438. (10) Ischiropoulos, H.; Zhu, L.; Beckman, J. S. Peroxynitrite Formation from Macrophage-Derived Nitric Oxide. Arch. Biochem. Biophys. 1992, 298, 446. (11) Ischiropoulos, H.; Zhu, L.; Chen, J.; Tsai, M.; Martin, J. C.; Smith, C. D.; Beckman, J. S. Peroxynitrite-Mediated Tyrosine Nitration Catalyzed by Superoxide Dismutase. Arch. Biochem. Biophys. 1992, 298, 431. (12) Koppenol, W. H.; Moreno, J. J.; Pryor, W. A.; Ischiropoulos, H.; Beckman, J. S. Peroxynitrite, a Cloaked Oxidant Formed by Nitric Oxide and Superoxide. Chem. Res. Toxicol. 1992, 5, 834. (13) Haddad, I. Y.; Ischiropoulos, H.; Holm, B. A.; Beckman, J. S.; Baker, J. R.; Matalon, S. Mechanisms of Peroxynitrite-Induced Injury to Pulmonary Surfactants. Am. J. Physiol. 1993, 265, L555. (14) Graham, A.; Hogg, N.; Kalyanaraman, B.; O’Leary, V.; DarleyUsmar, V.; Moncada, S. Peroxynitrite Modification of Low-Density Lipoprotein Leads to Recognition by the Macrophage Scavenger Receptor. FEBS Lett. 1993, 330, 181. (15) Hogg, N.; Darley-Usmar, V. M.; Graham, A.; Moncada, S. Peroxynitrite and Atherosclerosis. Biochem. Soc. Trans. 1993, 21, 358. (16) Pacher, P.; Beckman, J. S.; Liaudet, L. Nitric Oxide and Peroxynitrite in Health and Disease. Physiol. Rev. 2007, 87, 315. (17) Szabo, C.; Ischiropoulos, H.; Radi, R. Peroxynitrite: Biochemistry, Pathophysiology and Development of Therapeutics. Nat. Rev. Drug Discovery 2007, 6, 662. (18) Halfpenny, E.; Robinson, P. L. The Nitration and Hydroxylation of Aromatic Compounds by Pernitrous Acid. J. Chem. Soc. 1952, 939. 1387

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

and by Proximity to Nitric Oxide Synthase. Free Radical Biol. Med. 2006, 40, 1903. (41) Crow, J. P.; Beckman, J. S. Quantitation of Protein Tyrosine, 3Nitrotyrosine, and 3-Aminotyrosine Utilizing Hplc and Intrinsic Ultrviolet Absorbance. Methods 1995, 7, 116. (42) Sacksteder, C. A.; Qian, W. J.; Knyushko, T. V.; Wang, H.; Chin, M. H.; Lacan, G.; Melega, W. P.; Camp, D. G., 2nd; Smith, R. D.; Smith, D. J.; et al. Endogenously Nitrated Proteins in Mouse Brain: Links to Neurodegenerative Disease. Biochemistry 2006, 45, 8009. (43) Xu, S.; Ying, J.; Jiang, B.; Guo, W.; Adachi, T.; Sharov, V.; Lazar, H.; Menzoian, J.; Knyushko, T. V.; Bigelow, D.; et al. Detection of Sequence-Specific Tyrosine Nitration of Manganese Sod and Serca in Cardiovascular Disease and Aging. Am. J. Physiol.: Heart Circ. Physiol. 2006, 290, H2220. (44) Brennan, M.-L.; Wu, W.; Fu, X.; Shen, Z.; Song, W.; Frost, H.; Vadseth, C.; Narine, L.; Lenkiewicz, E.; Borchers, M. T.; et al. A Tale of Two Controversies: Defining Both the Role of Peroxidases in Nitrotyrosine Formation in Vivo Using Eosinophil Peroxidase and Myeloperoxidase-Deficient Mice, and the Nature of PeroxidaseGenerated Reactive Nitrogen Species. J. Biol. Chem. 2002, 277, 17415. (45) Pennathur, S.; Jackson-Lewis, V.; Przedborski, S.; Heinecke, J. W. Mass Spectrometric Quantification of 3-Nitrotyrosine, OrthoTyrosine, and O,O′-Dityrosine in Brain Tissue of 1-Methyl-4-Phenyl1,2,3, 6-Tetrahydropyridine-Treated Mice, a Model of Oxidative Stress in Parkinson’s Disease. J. Biol. Chem. 1999, 274, 34621. (46) van der Vliet, A.; Eiserich, J. P.; Halliwell, B.; Cross, C. E. Formation of Reactive Nitrogen Species During Peroxidase-Catalyzed Oxidation of Nitrite. A Potential Additional Mechanism of Nitric Oxide-Dependent Toxicity. J. Biol. Chem. 1997, 272, 7617. (47) Wu, W.; Chen, Y.; Hazen, S. L. Eosinophil Peroxidase Nitrates Protein Tyrosyl Residues. Implications for Oxidative Damage by Nitrating Intermediates in Eosinophilic Inflammatory Disorders. J. Biol. Chem. 1999, 274, 25933. (48) Trujillo, M.; Budde, H.; Pineyro, M. D.; Stehr, M.; Robello, C.; Flohe, L.; Radi, R. Trypanosoma Brucei and Trypanosoma Cruzi Tryparedoxin Peroxidases Catalytically Detoxify Peroxynitrite Via Oxidation of Fast Reacting Thiols. J. Biol. Chem. 2004, 279, 34175. (49) Ghosh, S.; Janocha, A. J.; Aronica, M. A.; Swaidani, S.; Comhair, S. A.; Xu, W.; Zheng, L.; Kaveti, S.; Kinter, M.; Hazen, S. L.; et al. Nitrotyrosine Proteome Survey in Asthma Identifies Oxidative Mechanism of Catalase Inactivation. J. Immunol. 2006, 176, 5587. (50) Keith, W. G.; Powell, R. E. Kinetics of Decomposition of Peroxynitrous Acid. J. Chem. Soc. A 1969, 90. (51) Pryor, W. A.; Squadrito, G. L. The Chemistry of Peroxynitrite: A Product from the Reaction of Nitric Oxide with Superoxide. Am. J. Physiol. 1995, 268, L699. (52) Marnett, L. J. Forum: Reactive Species of Peroxynitrite. Chem. Res. Toxicol. 1998, 11, 709. (53) Bartberger, M. D.; Olson, L. P.; Houk, K. N. Mechanisms of Peroxynitrite Oxidations and Rearrangements: The Theoretical Perspective. Chem. Res. Toxicol. 1998, 11, 710. (54) Merenyi, G.; Lind, J.; Goldstein, S.; Czapski, G. Peroxynitrous Acid Homolyzes into *OH and *NO2 Radicals. Chem. Res. Toxicol. 1998, 11, 712. (55) Lymar, S. V.; Hurst, J. K. Radical Nature of Peroxynitrite Reactivity. Chem. Res. Toxicol. 1998, 11, 714. (56) Koppenol, W. H. Peroxynitrite Uncloaked? Chem. Res. Toxicol. 1998, 11, 716. (57) Squadrito, G. L.; Pryor, W. A. The Nature of Reactive Species in Systems That Produce Peroxynitrite. Chem. Res. Toxicol. 1998, 11, 718. (58) Radi, R. Peroxynitrite Reactions and Diffusion in Biology. Chem. Res. Toxicol. 1998, 11, 720. (59) Goldstein, S.; Lind, J.; Merenyi, G. Chemistry of Peroxynitrites as Compared to Peroxynitrates. Chem. Rev. 2005, 105, 2457. (60) Koppenol, W. H.; Bounds, P. L.; Nauser, T.; Kissner, R.; Rüegger, H. Peroxynitrous Acid: Controversy and Consensus Surrounding an Enigmatic Oxidant. Dalton Trans. 2012, 41, 13779.

(61) Gatti, R. M.; Alvarez, B.; Vasquez-Vivar, J.; Radi, R.; Augusto, O. Formation of Spin Trap Adducts During the Decomposition of Peroxynitrite. Arch. Biochem. Biophys. 1998, 349, 36. (62) Augusto, O.; Radi, R.; Gatti, R. M.; Vasquez-Vivar, J. Detection of Secondary Radicals from Peroxynitrite-Medicated Oxidations by Electron Spin Resonance. Methods Enzymol. 1996, 269, 346. (63) Denicola, A.; Souza, J. M.; Gatti, R. M.; Augusto, O.; Radi, R. Desferrioxamine Inhibition of the Hydroxyl Radical-Like Reactivity of Peroxynitrite: Role of the Hydroxamic Groups. Free Radical Biol. Med. 1995, 19, 11. (64) Augusto, O.; Gatti, R. M.; Radi, R. Spin-Trapping Studies of Peroxynitrite Decomposition and of 3-Morpholinosydnonimine NEthylcarbamide Autooxidation: Direct Evidence for Metal-Independent Formation of Free Radical Intermediates. Arch. Biochem. Biophys. 1994, 310, 118. (65) Mahoney, L. R. Evidence for the Formation of Hydroxyl Radicals in the Isomerization of Pernitrous Acid to Nitric Acid in Aqueous Solution [33]. J. Am. Chem. Soc. 1970, 92, 5262. (66) Ferrer-Sueta, G.; Radi, R. Chemical Biology of Peroxynitrite: Kinetics, Diffusion, and Radicals. ACS Chem. Biol. 2009, 4, 161. (67) Bonini, M. G.; Radi, R.; Ferrer-Sueta, G.; Ferreira, A. M.; Augusto, O. Direct EPR Detection of the Carbonate Radical Anion Produced from Peroxynitrite and Carbon Dioxide. J. Biol. Chem. 1999, 274, 10802. (68) Bryk, R.; Griffin, P.; Nathan, C. Peroxynitrite Reductase Activity of Bacterial Peroxiredoxins. Nature 2000, 407, 211. (69) Briviba, K.; Kissner, R.; Koppenol, W. H.; Sies, H. Kinetic Study of the Reaction of Glutathione Peroxidase with Peroxynitrite. Chem. Res. Toxicol. 1998, 11, 1398. (70) Stadler, K.; Bonini, M. G.; Dallas, S.; Jiang, J.; Radi, R.; Mason, R. P.; Kadiiska, M. B. Involvement of Inducible Nitric Oxide Synthase in Hydroxyl Radical-Mediated Lipid Peroxidation in StreptozotocinInduced Diabetes. Free Radical Biol. Med. 2008, 45, 866. (71) Bartesaghi, S.; Wenzel, J.; Trujillo, M.; Lopez, M.; Joseph, J.; Kalyanaraman, B.; Radi, R. Lipid Peroxyl Radicals Mediate Tyrosine Dimerization and Nitration in Membranes. Chem. Res. Toxicol. 2010, 23, 821. (72) Bartesaghi, S.; Herrera, D.; Martinez, D. M.; Petruk, A.; Demicheli, V.; Trujillo, M.; Marti, M. A.; Estrin, D. A.; Radi, R. Tyrosine Oxidation and Nitration in Transmembrane Peptides Is Connected to Lipid Peroxidation. Arch. Biochem. Biophys. 2017, 622, 9. (73) Huie, R. E.; Padmaja, S. The Reaction of NO with Superoxide. Free Radical Res. Commun. 1993, 18, 195. (74) Lancaster, J. R., Jr. Simulation of the Diffusion and Reaction of Endogenously Produced Nitric Oxide. Proc. Natl. Acad. Sci. U. S. A. 1994, 91, 8137. (75) Takahashi, M. A.; Asada, K. Superoxide Anion Permeability of Phospholipid Membranes and Chloroplast Thylakoids. Arch. Biochem. Biophys. 1983, 226, 558. (76) Chang, L. Y.; Slot, J. W.; Geuze, H. J.; Crapo, J. D. Molecular Immunocytochemistry of the Cuzn Superoxide Dismutase in Rat Hepatocytes. J. Cell Biol. 1988, 107, 2169. (77) Halliwell, B.; Gutteridge, J. M. C. In Free Radical Biology & Medicine; Oxford University Press: Oxford, 1999. (78) Quijano, C.; Hernandez-Saavedra, D.; Castro, L.; McCord, J. M.; Freeman, B. A.; Radi, R. Reaction of Peroxynitrite with MnSuperoxide Dismutase. Role of the Metal Center in Decomposition Kinetics and Nitration. J. Biol. Chem. 2001, 276, 11631. (79) Fridovich, I. Superoxide Anion Radical (O2-.), Superoxide Dismutases, and Related Matters. J. Biol. Chem. 1997, 272, 18515. (80) McCord, J. M.; Keele, B. B., Jr.; Fridovich, I. An Enzyme-Based Theory of Obligate Anaerobiosis: The Physiological Function of Superoxide Dismutase. Proc. Natl. Acad. Sci. U. S. A. 1971, 68, 1024. (81) Martinez, A.; Peluffo, G.; Petruk, A. A.; Hugo, M.; Piñeyro, D.; Demicheli, V.; Moreno, D. M.; Lima, A.; Batthyány, C.; Durán, R.; et al. Structural and Molecular Basis of the Peroxynitrite-Mediated Nitration and Inactivation of Trypanosoma Cruzi Iron-Superoxide Dismutases (Fe-SODs) A and B: Disparate Susceptibilities Due to the 1388

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Repair of Tyr35 Radical by Cys83 in Fe-SODB through Intramolec. J. Biol. Chem. 2014, 289, 12760. (82) Fielden, E. M.; Roberts, P. B.; Bray, R. C.; Lowe, D. J.; Mautner, G. N.; Rotilio, G.; Calabrese, L. Mechanism of Action of Superoxide Dismutase from Pulse Radiolysis and Electron Paramagnetic Resonance. Evidence That Only Half the Active Sites Function in Catalysis. Biochem. J. 1974, 139, 49. (83) Hsu, J. L.; Hsieh, Y.; Tu, C.; O’Connor, D.; Nick, H. S.; Silverman, D. N. Catalytic Properties of Human Manganese Superoxide Dismutase. J. Biol. Chem. 1996, 271, 17687. (84) Klug-Roth, D.; Fridovich, I.; Rabani, J. Pulse Radiolytic Investigations of Superoxide Catalyzed Disproportionation. Mechanism for Bovine Superoxide Dismutase. J. Am. Chem. Soc. 1973, 95, 2786. (85) Gardner, P. R.; Raineri, I.; Epstein, L. B.; White, C. W. Superoxide Radical and Iron Modulate Aconitase Activity in Mammalian Cells. J. Biol. Chem. 1995, 270, 13399. (86) Denicola, A.; Souza, J. M.; Radi, R.; Lissi, E. Nitric Oxide Diffusion in Membranes Determined by Fluorescence Quenching. Arch. Biochem. Biophys. 1996, 328, 208. (87) Subczynski, W. K.; Lomnicka, M.; Hyde, J. S. Permeability of Nitric Oxide through Lipid Bilayer Membranes. Free Radical Res. 1996, 24, 343. (88) Thomas, D. D.; Liu, X.; Kantrow, S. P.; Lancaster, J. R., Jr. The Biological Lifetime of Nitric Oxide: Implications for the Perivascular Dynamics of NO and O2. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 355. (89) Herold, S.; Exner, M.; Nauser, T. Kinetic and Mechanistic Studies of the NO*-Mediated Oxidation of Oxymyoglobin and Oxyhemoglobin. Biochemistry 2001, 40, 3385. (90) Liu, X.; Miller, M. J.; Joshi, M. S.; Sadowska-Krowicka, H.; Clark, D. A.; Lancaster, J. R., Jr. Diffusion-Limited Reaction of Free Nitric Oxide with Erythrocytes. J. Biol. Chem. 1998, 273, 18709. (91) Liochev, S. I.; Fridovich, I. Superoxide and Nitric Oxide: Consequences of Varying Rates of Production and Consumption: A Theoretical Treatment. Free Radical Biol. Med. 2002, 33, 137. (92) Beckman, J. S.; Koppenol, W. H. Nitric Oxide, Superoxide, and Peroxynitrite: The Good, the Bad, and Ugly. Am. J. Physiol. 1996, 271, C1424. (93) Buerk, D. G.; Lamkin-Kennard, K.; Jaron, D. Modeling the Influence of Superoxide Dismutase on Superoxide and Nitric Oxide Interactions, Including Reversible Inhibition of Oxygen Consumption. Free Radical Biol. Med. 2003, 34, 1488. (94) Nalwaya, N.; Deen, W. M. Analysis of Cellular Exposure to Peroxynitrite in Suspension Cultures. Chem. Res. Toxicol. 2003, 16, 920. (95) Nalwaya, N.; Deen, W. M. Peroxynitrite Exposure of Cells Cocultured with Macrophages. Ann. Biomed. Eng. 2004, 32, 664. (96) Quijano, C.; Romero, N.; Radi, R. Tyrosine Nitration by Superoxide and Nitric Oxide Fluxes in Biological Systems: Modeling the Impact of Superoxide Dismutase and Nitric Oxide Diffusion. Free Radical Biol. Med. 2005, 39, 728. (97) Pfeiffer, S.; Mayer, B. Lack of Tyrosine Nitration by Peroxynitrite Generated at Physiological pH. J. Biol. Chem. 1998, 273, 27280. (98) Rubbo, H.; Radi, R.; Trujillo, M.; Telleri, R.; Kalyanaraman, B.; Barnes, S.; Kirk, M.; Freeman, B. A. Nitric Oxide Regulation of Superoxide and Peroxynitrite-Dependent Lipid Peroxidation. Formation of Novel Nitrogen-Containing Oxidized Lipid Derivatives. J. Biol. Chem. 1994, 269, 26066. (99) Jourd’heuil, D.; Jourd’heuil, F. L.; Kutchukian, P. S.; Musah, R. A.; Wink, D. A.; Grisham, M. B. Reaction of Superoxide and Nitric Oxide with Peroxynitrite. Implications for Peroxynitrite-Mediated Oxidation Reactions in Vivo. J. Biol. Chem. 2001, 276, 28799. (100) Jourd’heuil, D.; Miranda, K. M.; Kim, S. M.; Espey, M. G.; Vodovotz, Y.; Laroux, S.; Mai, C. T.; Miles, A. M.; Grisham, M. B.; Wink, D. A. The Oxidative and Nitrosative Chemistry of the Nitric Oxide/Superoxide Reaction in the Presence of Bicarbonate. Arch. Biochem. Biophys. 1999, 365, 92.

(101) Wardman, P. Fluorescent and Luminescent Probes for Measurement of Oxidative and Nitrosative Species in Cells and Tissues: Progress, Pitfalls, and Prospects. Free Radical Biol. Med. 2007, 43, 995. (102) Zielonka, J.; Sikora, A.; Joseph, J.; Kalyanaraman, B. Peroxynitrite Is the Major Species Formed from Different Flux Ratios of Co-Generated Nitric Oxide and Superoxide: Direct Reaction with Boronate-Based Fluorescent Probe. J. Biol. Chem. 2010, 285, 14210. (103) Rios, N.; Piacenza, L.; Trujillo, M.; Martinez, A.; Demicheli, V.; Prolo, C.; Alvarez, M. N.; Lopez, G. V.; Radi, R. Sensitive Detection and Estimation of Cell-Derived Peroxynitrite Fluxes Using Fluorescein-Boronate. Free Radical Biol. Med. 2016, 101, 284. (104) Prolo, C.; Alvarez, M. N.; Rios, N.; Peluffo, G.; Radi, R.; Romero, N. Nitric Oxide Diffusion to Red Blood Cells Limits Extracellular, but Not Intraphagosomal, Peroxynitrite Formation by Macrophages. Free Radical Biol. Med. 2015, 87, 346. (105) Chen, Z. J.; Ren, W.; Wright, Q. E.; Ai, H. W. Genetically Encoded Fluorescent Probe for the Selective Detection of Peroxynitrite. J. Am. Chem. Soc. 2013, 135, 14940. (106) Chen, Z. J.; Tian, Z.; Kallio, K.; Oleson, A. L.; Ji, A.; Borchardt, D.; Jiang, D. E.; Remington, S. J.; Ai, H. W. The N-B Interaction through a Water Bridge: Understanding the Chemoselectivity of a Fluorescent Protein Based Probe for Peroxynitrite. J. Am. Chem. Soc. 2016, 138, 4900. (107) Hogg, N.; Zielonka, J.; Kalyanaraman, B. In Nitric Oxide Biology and Pathobiology; Ignarro, L. J., Freeman, B. A., Eds.; Academic Press: London, 2017. (108) Ríos, N.; Prolo, C.; Alvarez, M. N.; Piacenza, L.; Radi, R. In Nitric Oxide Biology and Pathobiology; Ignarro, L. J., Freeman, B. A., Eds.; Academic Press: London, 2017. (109) Fukuto, J. M.; Bartberger, M. D.; Dutton, A. S.; Paolocci, N.; Wink, D. A.; Houk, K. N. The Physiological Chemistry and Biological Activity of Nitroxyl (HNO): The Neglected, Misunderstood, and Enigmatic Nitrogen Oxide. Chem. Res. Toxicol. 2005, 18, 790. (110) Miranda, K. M. The Chemistry of Nitroxyl (HNO) and Implications in Biology. Coord. Chem. Rev. 2005, 249, 433. (111) Doctorovich, F.; Bikiel, D. E.; Pellegrino, J.; Suarez, S. A.; Marti, M. A. Reactions of HNO with Metal Porphyrins: Underscoring the Biological Relevance of HNO. Acc. Chem. Res. 2014, 47, 2907. (112) Bianco, C. L.; Toscano, J. P.; Bartberger, M. D.; Fukuto, J. M. The Chemical Biology of HNO Signaling. Arch. Biochem. Biophys. 2017, 617, 129. (113) Doctorovich, F.; Farmer, P. J.; Marti, M. A. The Chemistry and Biology of Nitroxyl (HNO); Elsevier: Boston, 2017. (114) Bartberger, M. D.; Liu, W.; Ford, E.; Miranda, K. M.; Switzer, C.; Fukuto, J. M.; Farmer, P. J.; Wink, D. A.; Houk, K. N. The Reduction Potential of Nitric Oxide (NO) and Its Importance to NO Biochemistry. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 10958. (115) Shafirovich, V.; Lymar, S. V. Nitroxyl and Its Anion in Aqueous Solutions: Spin States, Protic Equilibria, and Reactivities toward Oxygen and Nitric Oxide. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 7340. (116) Marti, M. A.; Á lvarez, L.; Suarez, S. A.; Doctorovich, F. In The Chemistry and Biology of Nitroxyl (HNO); Elsevier: Boston, 2017. (117) Shafirovich, V.; Lymar, S. V. Spin-Forbidden Deprotonation of Aqueous Nitroxyl (HNO). J. Am. Chem. Soc. 2003, 125, 6547. (118) Liochev, S. I.; Fridovich, I. The Mode of Decomposition of Angeli’s Salt (Na2N2O3) and the Effects Thereon of Oxygen, Nitrite, Superoxide Dismutase, and Glutathione. Free Radical Biol. Med. 2003, 34, 1399. (119) Miranda, K. M.; Paolocci, N.; Katori, T.; Thomas, D. D.; Ford, E.; Bartberger, M. D.; Espey, M. G.; Kass, D. A.; Feelisch, M.; Fukuto, J. M.; et al. A Biochemical Rationale for the Discrete Behavior of Nitroxyl and Nitric Oxide in the Cardiovascular System. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 9196. (120) Smulik, R.; Debski, D.; Zielonka, J.; Michalowski, B.; Adamus, J.; Marcinek, A.; Kalyanaraman, B.; Sikora, A. Nitroxyl (HNO) Reacts with Molecular Oxygen and Forms Peroxynitrite at Physiological pH. Biological Implications. J. Biol. Chem. 2014, 289, 35570. 1389

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(121) Jorolan, J. H.; Buttitta, L. A.; Cheah, C.; Miranda, K. M. Comparison of the Chemical Reactivity of Synthetic Peroxynitrite with That of the Autoxidation Products of Nitroxyl or Its Anion. Nitric Oxide 2015, 44, 39. (122) Doyle, M. P.; Hoekstra, J. W. Oxidation of Nitrogen Oxides by Bound Dioxygen in Hemoproteins. J. Inorg. Biochem. 1981, 14, 351. (123) Wade, R. S.; Castro, C. E. Reactions of Oxymyoglobin with NO, NO2, and NO2- under Argon and in Air. Chem. Res. Toxicol. 1996, 9, 1382. (124) Herold, S. Mechanistic Studies of the Oxidation of Pyridoxalated Hemoglobin Polyoxyethylene Conjugate by Nitrogen Monoxide. Arch. Biochem. Biophys. 1999, 372, 393. (125) Herold, S. Kinetic and Spectroscopic Characterization of an Intermediate Peroxynitrite Complex in the Nitrogen Monoxide Induced Oxidation of Oxyhemoglobin. FEBS Lett. 1998, 439, 85. (126) Eich, R. F.; Li, T.; Lemon, D. D.; Doherty, D. H.; Curry, S. R.; Aitken, J. F.; Mathews, A. J.; Johnson, K. A.; Smith, R. D.; Phillips, G. N., Jr; et al. Mechanism of NO-Induced Oxidation of Myoglobin and Hemoglobin. Biochemistry 1996, 35, 6976. (127) Gardner, P. R. Nitric Oxide Dioxygenase Function and Mechanism of Flavohemoglobin, Hemoglobin, Myoglobin and Their Associated Reductases. J. Inorg. Biochem. 2005, 99, 247. (128) Arnold, E. V.; Bohle, D. S. In Methods in Enzymology; Academic Press, 1996; Vol. 269. (129) Subedi, H.; Brasch, N. E. Mechanistic Studies on the Reaction of Nitroxylcobalamin with Dioxygen: Evidence for Formation of a Peroxynitritocob(III)Alamin Intermediate. Inorg. Chem. 2013, 52, 11608. (130) Carballal, S.; Cuevasanta, E.; Yadav, P. K.; Gherasim, C.; Ballou, D. P.; Alvarez, B.; Banerjee, R. Kinetics of Nitrite Reduction and Peroxynitrite Formation by Ferrous Heme in Human Cystathionine B-Synthase. J. Biol. Chem. 2016, 291, 8004. (131) Herold, S.; Koppenol, W. H. Peroxynitritometal Complexes. Coord. Chem. Rev. 2005, 249, 499. (132) Kim, S.; Siegler, M. A.; Karlin, K. D. Peroxynitrite Chemistry Derived from Nitric Oxide Reaction with a Cu(II)-OOH Species and a Copper Mediated NO Reductive Coupling Reaction. Chem. Commun. (Cambridge, U. K.) 2014, 50, 2844. (133) Barry, S. M.; Kers, J. A.; Johnson, E. G.; Song, L.; Aston, P. R.; Patel, B.; Krasnoff, S. B.; Crane, B. R.; Gibson, D. M.; Loria, R.; et al. Cytochrome P450-Catalyzed L-Tryptophan Nitration in Thaxtomin Phytotoxin Biosynthesis. Nat. Chem. Biol. 2012, 8, 814. (134) Baeyer, A.; Villiger, V. Uber Die Salpetrige Saüre. Ber. Dtsch. Chem. Ges. 1901, 34, 755. (135) Reed, J. W.; Ho, H. H.; Jolly, W. L. Chemical Syntheses with a Quenched Flow Reactor. Hydroxytrihydroborate and Peroxynitrite. J. Am. Chem. Soc. 1974, 96, 1248. (136) Saha, A.; Goldstein, S.; Cabelli, D.; Czapski, G. Determination of Optimal Conditions for Synthesis of Peroxynitrite by Mixing Acidified Hydrogen Peroxide with Nitrite. Free Radical Biol. Med. 1998, 24, 653. (137) Pryor, W. A.; Cueto, R.; Jin, X.; Koppenol, W. H.; NguSchwemlein, M.; Squadrito, G. L.; Uppu, P. L.; Uppu, R. M. A Practical Method for Preparing Peroxynitrite Solutions of Low Ionic Strength and Free of Hydrogen Peroxide. Free Radical Biol. Med. 1995, 18, 75. (138) Hughes, M. N.; Nicklin, H. G. Autoxidation of Hydroxylamine in Alkaline Solution. J. Chem. Soc. A 1971, 164. (139) Uppu, R. M. Synthesis of Peroxynitrite Using Isoamyl Nitrite and Hydrogen Peroxide in a Homogeneous Solvent System. Anal. Biochem. 2006, 354, 165. (140) Uppu, R. M.; Pryor, W. A. Synthesis of Peroxynitrite in a TwoPhase System Using Isoamyl Nitrite and Hydrogen Peroxide. Anal. Biochem. 1996, 236, 242. (141) Koppenol, W. H.; Kissner, R.; Beckman, J. S. Syntheses of Peroxynitrite: To Go with the Flow or on Solid Grounds? Methods Enzymol. 1996, 269, 296.

(142) Bohle, D. S.; Glassbrenner, P. A.; Hansert, B. Synthesis of Pure Tetramethylammonium Peroxynitrite. Methods Enzymol. 1996, 269, 302. (143) Bohle, D. S.; Hansert, B.; Paulson, S. C.; Smith, B. D. Biomimetic Synthesis of the Putative Cytotoxin Peroxynitrite, ONOO−, and Its Characterization as a Tetramethylammonium Salt. J. Am. Chem. Soc. 1994, 116, 7423. (144) Bartesaghi, S.; Trujillo, M.; Denicola, A.; Folkes, L.; Wardman, P.; Radi, R. Reactions of Desferrioxamine with Peroxynitrite-Derived Carbonate and Nitrogen Dioxide Radicals. Free Radical Biol. Med. 2004, 36, 471. (145) Rocha, B. S.; Gago, B.; Barbosa, R. M.; Lundberg, J. O.; Radi, R.; Laranjinha, J. Intragastric Nitration by Dietary Nitrite: Implications for Modulation of Protein and Lipid Signaling. Free Radical Biol. Med. 2012, 52, 693. (146) Rocha, S.; Gago, B.; Barbosa, R. M.; Lundberg, J. O.; Mann, G. E.; Radi, R.; Laranjinha, J. Pepsin Is Nitrated in the Rat Stomach, Acquiring Antiulcerogenic Activity: A Novel Interaction between Dietary Nitrate and Gut Proteins. Free Radical Biol. Med. 2013, 58, 26. (147) Rocha, B. S.; Lundberg, J. O.; Radi, R.; Laranjinha, J. Role of Nitrite, Urate and Pepsin in the Gastroprotective Effects of Saliva. Redox Biol. 2016, 8, 407. (148) Lundberg, J. O.; Gladwin, M. T.; Ahluwalia, A.; Benjamin, N.; Bryan, N. S.; Butler, A.; Cabrales, P.; Fago, A.; Feelisch, M.; Ford, P. C.; et al. Nitrate and Nitrite in Biology, Nutrition and Therapeutics. Nat. Chem. Biol. 2009, 5, 865. (149) Shiva, S.; Gladwin, M. T. Nitrite Mediates Cytoprotection after Ischemia/Reperfusion by Modulating Mitochondrial Function. Basic Res. Cardiol. 2009, 104, 113. (150) van Faassen, E. E.; Bahrami, S.; Feelisch, M.; Hogg, N.; Kelm, M.; Kim-Shapiro, D. B.; Kozlov, A. V.; Li, H.; Lundberg, J. O.; Mason, R.; et al. Nitrite as Regulator of Hypoxic Signaling in Mammalian Physiology. Med. Res. Rev. 2009, 29, 683. (151) Gladwin, M. T.; Raat, N. J.; Shiva, S.; Dezfulian, C.; Hogg, N.; Kim-Shapiro, D. B.; Patel, R. P. Nitrite as a Vascular Endocrine Nitric Oxide Reservoir That Contributes to Hypoxic Signaling, Cytoprotection, and Vasodilation. Am. J. Physiol. Heart Circ Physiol 2006, 291, H2026. (152) Shiva, S.; Gladwin, M. T. Nitrite Therapeutics: Back to the Future. Crit. Care Med. 2005, 33, 1865. (153) Schmidt, K.; Pfeiffer, S.; Mayer, B. Reaction of Peroxynitrite with Hepes or Mops Results in the Formation of Nitric Oxide Donors. Free Radical Biol. Med. 1998, 24, 859. (154) Lomonosova, E. E.; Kirsch, M.; Rauen, U.; de Groot, H. The Critical Role of Hepes in Sin-1 Cytotoxicity, Peroxynitrite Versus Hydrogen Peroxide. Free Radical Biol. Med. 1998, 24, 522. (155) Kirsch, M.; Lomonosova, E. E.; Korth, H. G.; Sustmann, R.; de Groot, H. Hydrogen Peroxide Formation by Reaction of Peroxynitrite with Hepes and Related Tertiary Amines. Implications for a General Mechanism. J. Biol. Chem. 1998, 273, 12716. (156) Gadelha, F. R.; Thomson, L.; Fagian, M. M.; Costa, A. D.; Radi, R.; Vercesi, A. E. Ca2+-Independent Permeabilization of the Inner Mitochondrial Membrane by Peroxynitrite Is Mediated by Membrane Protein Thiol Cross-Linking and Lipid Peroxidation. Arch. Biochem. Biophys. 1997, 345, 243. (157) Hughes, M. N.; Nicklin, H. G. The Chemistry of Pernitrites. Part I. Kinetics of Decomposition of Pernitrous Acid. J. Chem. Soc. A 1968, A, 450. (158) Denicola, A.; Rubbo, H.; Rodriguez, D.; Radi, R. PeroxynitriteMediated Cytotoxicity to Trypanosoma Cruzi. Arch. Biochem. Biophys. 1993, 304, 279. (159) Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill, 1995. (160) Khairutdinov, R. F.; Coddington, J. W.; Hurst, J. K. Permeation of Phospholipid Membranes by Peroxynitrite. Biochemistry 2000, 39, 14238. (161) Marla, S. S.; Lee, J.; Groves, J. T. Peroxynitrite Rapidly Permeates Phospholipid Membranes. Proc. Natl. Acad. Sci. U. S. A. 1997, 94, 14243. 1390

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(162) Möller, M. N.; Lancaster, J. R.; Denicola, A. Chapter 2 the Interaction of Reactive Oxygen and Nitrogen Species with Membranes. Curr. Top. Membr. 2008, 61, 23. (163) Carballal, S.; Bartesaghi, S.; Radi, R. Kinetic and Mechanistic Considerations to Assess the Biological Fate of Peroxynitrite. Biochim. Biophys. Acta, Gen. Subj. 2014, 1840, 768. (164) Romero, N.; Denicola, A.; Souza, J. M.; Radi, R. Diffusion of Peroxynitrite in the Presence of Carbon Dioxide. Arch. Biochem. Biophys. 1999, 368, 23. (165) Alvarez, M. N.; Piacenza, L.; Irigoin, F.; Peluffo, G.; Radi, R. Macrophage-Derived Peroxynitrite Diffusion and Toxicity to Trypanosoma Cruzi. Arch. Biochem. Biophys. 2004, 432, 222. (166) Lymar, S. V.; Hurst, J. K. Carbon Dioxide: Physiological Catalyst for Peroxynitrite-Mediated Cellular Damage or Cellular Protectant? Chem. Res. Toxicol. 1996, 9, 845. (167) Hurst, J. K.; Lymar, S. V. Toxicity of Peroxynitrite and Related Reactive Nitrogen Species toward Escherichia Coli. Chem. Res. Toxicol. 1997, 10, 802. (168) Kuwahara, H.; Miyamoto, Y.; Akaike, T.; Kubota, T.; Sawa, T.; Okamoto, S.; Maeda, H. Helicobacter Pylori Urease Suppresses Bactericidal Activity of Peroxynitrite Via Carbon Dioxide Production. Infect. Immun. 2000, 68, 4378. (169) Gobert, A. P.; Wilson, K. T. The Immune Battle against Helicobacter Pylori Infection: NO Offense. Trends Microbiol. 2016, 24, 366. (170) Signorelli, S.; Möller, M. N.; Coitino, E. L.; Denicola, A. Nitrogen Dioxide Solubility and Permeation in Lipid Membranes. Arch. Biochem. Biophys. 2011, 512, 190. (171) Denicola, A.; Souza, J. M.; Radi, R. Diffusion of Peroxynitrite across Erythrocyte Membranes. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 3566. (172) Macfadyen, A. J.; Reiter, C.; Zhuang, Y.; Beckman, J. S. A Novel Superoxide Dismutase-Based Trap for Peroxynitrite Used to Detect Entry of Peroxynitrite into Erythrocyte Ghosts. Chem. Res. Toxicol. 1999, 12, 223. (173) Alvarez, M. N.; Peluffo, G.; Piacenza, L.; Radi, R. Intraphagosomal Peroxynitrite as a Macrophage-Derived Cytotoxin against Internalized Trypanosoma Cruzi: Consequences for Oxidative Killing and Role of Microbial Peroxiredoxins in Infectivity. J. Biol. Chem. 2011, 286, 6627. (174) Hogg, N.; Darley-Usmar, V. M.; Wilson, M. T.; Moncada, S. The Oxidation of Alpha-Tocopherol in Human Low-Density Lipoprotein by the Simultaneous Generation of Superoxide and Nitric Oxide. FEBS Lett. 1993, 326, 199. (175) Pryor, W. A.; Jin, X.; Squadrito, G. L. One- and Two-Electron Oxidations of Methionine by Peroxynitrite. Proc. Natl. Acad. Sci. U. S. A. 1994, 91, 11173. (176) Shi, X.; Rojanasakul, Y.; Gannett, P.; Liu, K.; Mao, Y.; Daniel, L. N.; Ahmed, N.; Saffiotti, U. Generation of Thiyl and Ascorbyl Radicals in the Reaction of Peroxynitrite with Thiols and Ascorbate at Physiological Ph. J. Inorg. Biochem. 1994, 56, 77. (177) van der Vliet, A.; O'Neill, C. A.; Halliwell, B.; Cross, C. E.; Kaur, H. Aromatic Hydroxylation and Nitration of Phenylalanine and Tyrosine by Peroxynitrite. Evidence for Hydroxyl Radical Production from Peroxynitrite. FEBS Lett. 1994, 339, 89. (178) Lymar, S. V.; Hurst, J. K. Rapid Reaction between Peroxynitrite Ion and Carbon Dioxide: Implications for Biological Activity. J. Am. Chem. Soc. 1995, 117, 8867. (179) Roussyn, I.; Briviba, K.; Masumoto, H.; Sies, H. SeleniumContaining Compounds Protect DNA from Single-Strand Breaks Caused by Peroxynitrite. Arch. Biochem. Biophys. 1996, 330, 216. (180) Padmaja, S.; Ramazenian, M. S.; Bounds, P. L.; Koppenol, W. H. Reaction of Peroxynitrite with L-Tryptophan. Redox Rep. 1996, 2, 173. (181) Alvarez, B.; Rubbo, H.; Kirk, M.; Barnes, S.; Freeman, B. A.; Radi, R. Peroxynitrite-Dependent Tryptophan Nitration. Chem. Res. Toxicol. 1996, 9, 390. (182) Moreno, J. J.; Pryor, W. A. Inactivation of Alpha 1-Proteinase Inhibitor by Peroxynitrite. Chem. Res. Toxicol. 1992, 5, 425.

(183) Floris, R.; Piersma, S. R.; Yang, G.; Jones, P.; Wever, R. Interaction of Myeloperoxidase with Peroxynitrite. A Comparison with Lactoperoxidase, Horseradish Peroxidase and Catalase. Eur. J. Biochem. 1993, 215, 767. (184) King, P. A.; Jamison, E.; Strahs, D.; Anderson, V. E.; Brenowitz, M. ’Footprinting’ Proteins on DNA with Peroxonitrous Acid. Nucleic Acids Res. 1993, 21, 2473. (185) Castro, L.; Rodriguez, M.; Radi, R. Aconitase Is Readily Inactivated by Peroxynitrite, but Not by Its Precursor, Nitric Oxide. J. Biol. Chem. 1994, 269, 29409. (186) Deliconstantinos, G.; Villiotou, V.; Stavrides, J. C. Scavenging Effects of Hemoglobin and Related Heme Containing Compounds on Nitric Oxide, Reactive Oxidants and Carcinogenic Volatile Nitrosocompounds of Cigarette Smoke. A New Method for Protection against the Dangerous Cigarette Constituents. Anticancer Res. 1994, 14, 2717. (187) Swain, J. A.; Darley-Usmar, V.; Gutteridge, J. M. Peroxynitrite Releases Copper from Caeruloplasmin: Implications for Atherosclerosis. FEBS Lett. 1994, 342, 49. (188) Rubbo, H.; Denicola, A.; Radi, R. Peroxynitrite Inactivates Thiol-Containing Enzymes of Trypanosoma Cruzi Energetic Metabolism and Inhibits Cell Respiration. Arch. Biochem. Biophys. 1994, 308, 96. (189) Mohr, S.; Stamler, J. S.; Brune, B. Mechanism of Covalent Modification of Glyceraldehyde-3-Phosphate Dehydrogenase at Its Active Site Thiol by Nitric Oxide, Peroxynitrite and Related Nitrosating Agents. FEBS Lett. 1994, 348, 223. (190) Radi, R.; Rodriguez, M.; Castro, L.; Telleri, R. Inhibition of Mitochondrial Electron Transport by Peroxynitrite. Arch. Biochem. Biophys. 1994, 308, 89. (191) Haddad, I. Y.; Crow, J. P.; Hu, P.; Ye, Y.; Beckman, J.; Matalon, S. Concurrent Generation of Nitric Oxide and Superoxide Damages Surfactant Protein A. Am. J. Physiol. 1994, 267, L242. (192) Crow, J. P.; Beckman, J. S.; McCord, J. M. Sensitivity of the Essential Zinc-Thiolate Moiety of Yeast Alcohol Dehydrogenase to Hypochlorite and Peroxynitrite. Biochemistry 1995, 34, 3544. (193) Thomson, L.; Trujillo, M.; Telleri, R.; Radi, R. Kinetics of Cytochrome C2+ Oxidation by Peroxynitrite: Implications for Superoxide Measurements in Nitric Oxide-Producing Biological Systems. Arch. Biochem. Biophys. 1995, 319, 491. (194) Balazy, M. Peroxynitrite and Arachidonic Acid. Identification of Arachidonate Epoxides. Pol. J. Pharmacol. 1994, 46, 593. (195) Moore, K. P.; Darley-Usmar, V.; Morrow, J.; Roberts, L. J., 2nd Formation of F2-Isoprostanes During Oxidation of Human LowDensity Lipoprotein and Plasma by Peroxynitrite. Circ. Res. 1995, 77, 335. (196) Moro, M. A.; Darley-Usmar, V. M.; Lizasoain, I.; Su, Y.; Knowles, R. G.; Radomski, M. W.; Moncada, S. The Formation of Nitric Oxide Donors from Peroxynitrite. Br. J. Pharmacol. 1995, 116, 1999. (197) Salgo, M. G.; Stone, K.; Squadrito, G. L.; Battista, J. R.; Pryor, W. A. Peroxynitrite Causes DNA Nicks in Plasmid Pbr322. Biochem. Biophys. Res. Commun. 1995, 210, 1025. (198) Yermilov, V.; Rubio, J.; Ohshima, H. Formation of 8Nitroguanine in DNA Treated with Peroxynitrite in Vitro and Its Rapid Removal from DNA by Depurination. FEBS Lett. 1995, 376, 207. (199) Epe, B.; Ballmaier, D.; Roussyn, I.; Briviba, K.; Sies, H. DNA Damage by Peroxynitrite Characterized with DNA Repair Enzymes. Nucleic Acids Res. 1996, 24, 4105. (200) Radi, R. Kinetic Analysis of Reactivity of Peroxynitrite with Biomolecules. Methods Enzymol. 1996, 269, 354. (201) Arteel, G. E.; Briviba, K.; Sies, H. Protection against Peroxynitrite. FEBS Lett. 1999, 445, 226. (202) Radi, R.; Cosgrove, T. P.; Beckman, J. S.; Freeman, B. A. Peroxynitrite-Induced Luminol Chemiluminescence. Biochem. J. 1993, 290, 51. (203) Denicola, A.; Freeman, B. A.; Trujillo, M.; Radi, R. Peroxynitrite Reaction with Carbon Dioxide/Bicarbonate: Kinetics 1391

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

and Influence on Peroxynitrite-Mediated Oxidations. Arch. Biochem. Biophys. 1996, 333, 49. (204) Zhang, H.; Squadrito, G. L.; Pryor, W. A. The Mechanism of the Peroxynitrite-Carbon Dioxide Reaction Probed Using Tyrosine. Nitric Oxide 1997, 1, 301. (205) Goldstein, S.; Czapski, G. The Effect of Bicarbonate on Oxidation by Peroxynitrite: Implication for Its Biological Activity. Inorg. Chem. 1997, 36, 5113. (206) Lymar, S. V.; Jiang, Q.; Hurst, J. K. Mechanism of Carbon Dioxide-Catalyzed Oxidation of Tyrosine by Peroxynitrite. Biochemistry 1996, 35, 7855. (207) Lymar, S. V.; Hurst, J. K. Co2-Catalyzed One-Electron Oxidations by Peroxynitrite: Properties of the Reactive Intermediate. Inorg. Chem. 1998, 37, 294. (208) Meli, R.; Nauser, T.; Koppenol, W. H. Direct Observation of Intermediates in the Reaction of Peroxynitrite with Carbon Dioxide. Helv. Chim. Acta 1999, 82, 722. (209) Pryor, W. A.; Lemercier, J. N.; Zhang, H.; Uppu, R. M.; Squadrito, G. L. The Catalytic Role of Carbon Dioxide in the Decomposition of Peroxynitrite. Free Radical Biol. Med. 1997, 23, 331. (210) Harned, H. S.; Bonner, F. T. The First Ionization of Carbonic Acid in Aqueous Solutions of Sodium Chloride. J. Am. Chem. Soc. 1945, 67, 1026. (211) Kern, D. M. The Hydration of Carbon Dioxide. J. Chem. Educ. 1960, 37, 14. (212) Sly, W. S.; Hu, P. Y. Human Carbonic Anhydrases and Carbonic Anhydrase Deficiencies. Annu. Rev. Biochem. 1995, 64, 375. (213) Balboni, E.; Lehninger, A. L. Entry and Exit Pathways of CO2 in Rat Liver Mitochondria Respiring in a Bicarbonate Buffer System. J. Biol. Chem. 1986, 261, 3563. (214) Llopis, J.; McCaffery, J. M.; Miyawaki, A.; Farquhar, M. G.; Tsien, R. Y. Measurement of Cytosolic, Mitochondrial, and Golgi Ph in Single Living Cells with Green Fluorescent Proteins. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 6803. (215) Alayash, A. I.; Ryan, B. A.; Cashon, R. E. PeroxynitriteMediated Heme Oxidation and Protein Modification of Native and Chemically Modified Hemoglobins. Arch. Biochem. Biophys. 1998, 349, 65. (216) Exner, M.; Herold, S. Kinetic and Mechanistic Studies of the Peroxynitrite-Mediated Oxidation of Oxymyoglobin and Oxyhemoglobin. Chem. Res. Toxicol. 2000, 13, 287. (217) Furtmuller, P. G.; Jantschko, W.; Zederbauer, M.; Schwanninger, M.; Jakopitsch, C.; Herold, S.; Koppenol, W. H.; Obinger, C. Peroxynitrite Efficiently Mediates the Interconversion of Redox Intermediates of Myeloperoxidase. Biochem. Biophys. Res. Commun. 2005, 337, 944. (218) Gebicka, L.; Didik, J.; Gebicki, J. L. Catalytic Scavenging of Peroxynitrite by Lactoperoxidase in the Absence and Presence of Bicarbonate. Free Radical Res. 2010, 44, 217. (219) Edens, W. A.; Sharling, L.; Cheng, G.; Shapira, R.; Kinkade, J. M.; Lee, T.; Edens, H. A.; Tang, X.; Sullards, C.; Flaherty, D. B.; et al. Tyrosine Cross-Linking of Extracellular Matrix Is Catalyzed by Duox, a Multidomain Oxidase/Peroxidase with Homology to the Phagocyte Oxidase Subunit Gp91phox. J. Cell Biol. 2001, 154, 879. (220) Gebicka, L.; Didik, J. Catalytic Scavenging of Peroxynitrite by Catalase. J. Inorg. Biochem. 2009, 103, 1375. (221) Trostchansky, A.; O’Donnell, V. B.; Goodwin, D. C.; Landino, L. M.; Marnett, L. J.; Radi, R.; Rubbo, H. Interactions between Nitric Oxide and Peroxynitrite During Prostaglandin Endoperoxide H Synthase-1 Catalysis: A Free Radical Mechanism of Inactivation. Free Radical Biol. Med. 2007, 42, 1029. (222) Zou, M. H.; Daiber, A.; Peterson, J. A.; Shoun, H.; Ullrich, V. Rapid Reactions of Peroxynitrite with Heme-Thiolate Proteins as the Basis for Protection of Prostacyclin Synthase from Inactivation by Nitration. Arch. Biochem. Biophys. 2000, 376, 149. (223) Gebicka, L.; Didik, J. Kinetic Studies of the Reaction of HemeThiolate Enzyme Chloroperoxidase with Peroxynitrite. J. Inorg. Biochem. 2007, 101, 159.

(224) Marechal, A.; Mattioli, T. A.; Stuehr, D. J.; Santolini, J. Activation of Peroxynitrite by Inducible Nitric-Oxide Synthase: A Direct Source of Nitrative Stress. J. Biol. Chem. 2007, 282, 14101. (225) Ogusucu, R.; Rettori, D.; Munhoz, D. C.; Soares Netto, L. E.; Augusto, O. Reactions of Yeast Thioredoxin Peroxidases I and II with Hydrogen Peroxide and Peroxynitrite: Rate Constants by Competitive Kinetics. Free Radical Biol. Med. 2007, 42, 326. (226) Wengenack, N. L.; Jensen, M. P.; Rusnak, F.; Stern, M. K. Mycobacterium Tuberculosis Katg Is a Peroxynitritase. Biochem. Biophys. Res. Commun. 1999, 256, 485. (227) McLean, S.; Bowman, L. A.; Poole, R. K. Katg from Salmonella Typhimurium Is a Peroxynitritase. FEBS Lett. 2010, 584, 1628. (228) Marechal, A.; Mattioli, T. A.; Stuehr, D. J.; Santolini, J. NO Synthase Isoforms Specifically Modify Peroxynitrite Reactivity. FEBS J. 2010, 277, 3963. (229) Daiber, A.; Herold, S.; Schoneich, C.; Namgaladze, D.; Peterson, J. A.; Ullrich, V. Nitration and Inactivation of Cytochrome P450bm-3 by Peroxynitrite. Stopped-Flow Measurements Prove Ferryl Intermediates. Eur. J. Biochem. 2000, 267, 6729. (230) Ascenzi, P.; Pesce, A. Peroxynitrite Scavenging by Campylobacter Jejuni Truncated Hemoglobin P. J. Biol. Inorg. Chem. 2017, 22, 1141. (231) Celano, L.; Gil, M.; Carballal, S.; Durán, R.; Denicola, A.; Banerjee, R.; Alvarez, B. Inactivation of Cystathionine B-Synthase with Peroxynitrite. Arch. Biochem. Biophys. 2009, 491, 96. (232) Herold, S.; Shivashankar, K. Metmyoglobin and Methemoglobin Catalyze the Isomerization of Peroxynitrite to Nitrate. Biochemistry 2003, 42, 14036. (233) Demicheli, V.; Quijano, C.; Alvarez, B.; Radi, R. Inactivation and Nitration of Human Superoxide Dismutase (SOD) by Fluxes of Nitric Oxide and Superoxide. Free Radical Biol. Med. 2007, 42, 1359. (234) Sharpe, M. A.; Cooper, C. E. Interaction of Peroxynitrite with Mitochondrial Cytochrome Oxidase. Catalytic Production of Nitric Oxide and Irreversible Inhibition of Enzyme Activity. J. Biol. Chem. 1998, 273, 30961. (235) Pearce, L. L.; Pitt, B. R.; Peterson, J. The Peroxynitrite Reductase Activity of Cytochrome C Oxidase Involves a Two-Electron Redox Reaction at the Heme a(3)-Cu(B) Site. J. Biol. Chem. 1999, 274, 35763. (236) Cassina, A.; Radi, R. Differential Inhibitory Action of Nitric Oxide and Peroxynitrite on Mitochondrial Electron Transport. Arch. Biochem. Biophys. 1996, 328, 309. (237) Radi, R.; Cassina, A.; Hodara, R.; Quijano, C.; Castro, L. Peroxynitrite Reactions and Formation in Mitochondria. Free Radical Biol. Med. 2002, 33, 1451. (238) Pearce, L. L.; Kanai, A. J.; Birder, L. A.; Pitt, B. R.; Peterson, J. The Catabolic Fate of Nitric Oxide: The Nitric Oxide Oxidase and Peroxynitrite Reductase Activities of Cytochrome Oxidase. J. Biol. Chem. 2002, 277, 13556. (239) Cassina, A. M.; Hodara, R.; Souza, J. M.; Thomson, L.; Castro, L.; Ischiropoulos, H.; Freeman, B. A.; Radi, R. Cytochrome C Nitration by Peroxynitrite. J. Biol. Chem. 2000, 275, 21409. (240) Castro, L. A.; Robalinho, R. L.; Cayota, A.; Meneghini, R.; Radi, R. Nitric Oxide and Peroxynitrite-Dependent Aconitase Inactivation and Iron-Regulatory Protein-1 Activation in Mammalian Fibroblasts. Arch. Biochem. Biophys. 1998, 359, 215. (241) Hausladen, A.; Fridovich, I. Superoxide and Peroxynitrite Inactivate Aconitases, but Nitric Oxide Does Not. J. Biol. Chem. 1994, 269, 29405. (242) Keyer, K.; Imlay, J. A. Inactivation of Dehydratase [4Fe-4S] Clusters and Disruption of Iron Homeostasis Upon Cell Exposure to Peroxynitrite. J. Biol. Chem. 1997, 272, 27652. (243) Tortora, V.; Quijano, C.; Freeman, B.; Radi, R.; Castro, L. Mitochondrial Aconitase Reaction with Nitric Oxide, S-Nitrosoglutathione, and Peroxynitrite: Mechanisms and Relative Contributions to Aconitase Inactivation. Free Radical Biol. Med. 2007, 42, 1075. (244) Kennedy, M. C.; Antholine, W. E.; Beinert, H. An Epr Investigation of the Products of the Reaction of Cytosolic and 1392

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Mitochondrial Aconitases with Nitric Oxide. J. Biol. Chem. 1997, 272, 20340. (245) Drapier, J. C.; Hibbs, J. B., Jr. Differentiation of Murine Macrophages to Express Nonspecific Cytotoxicity for Tumor Cells Results in L-Arginine-Dependent Inhibition of Mitochondrial IronSulfur Enzymes in the Macrophage Effector Cells. J. Immunol. 1988, 140, 2829. (246) Drapier, J. C.; Hibbs, J. B., Jr. Murine Cytotoxic Activated Macrophages Inhibit Aconitase in Tumor Cells. Inhibition Involves the Iron-Sulfur Prosthetic Group and Is Reversible. J. Clin. Invest. 1986, 78, 790. (247) Drapier, J. C.; Wietzerbin, J.; Hibbs, J. B., Jr. InterferonGamma and Tumor Necrosis Factor Induce the L-Arginine-Dependent Cytotoxic Effector Mechanism in Murine Macrophages. Eur. J. Immunol. 1988, 18, 1587. (248) Hibbs, J. B., Jr.; Vavrin, Z.; Taintor, R. R. L-Arginine Is Required for Expression of the Activated Macrophage Effector Mechanism Causing Selective Metabolic Inhibition in Target Cells. J. Immunol. 1987, 138, 550. (249) Amber, I. J.; Hibbs, J. B., Jr.; Parker, C. J.; Johnson, B. B.; Taintor, R. R.; Vavrin, Z. Activated Macrophage Conditioned Medium: Identification of the Soluble Factors Inducing Cytotoxicity and the L-Arginine Dependent Effector Mechanism. J. Leukoc. Biol. 1991, 49, 610. (250) Soum, E.; Drapier, J. C. Nitric Oxide and Peroxynitrite Promote Complete Disruption of the [4Fe-4S] Cluster of Recombinant Human Iron Regulatory Protein 1. JBIC, J. Biol. Inorg. Chem. 2003, 8, 226. (251) Gonzalez, D.; Drapier, J. C.; Bouton, C. Endogenous Nitration of Iron Regulatory Protein-1 (Irp-1) in Nitric Oxide-Producing Murine Macrophages: Further Insight into the Mechanism of Nitration in Vivo and Its Impact on Irp-1 Functions. J. Biol. Chem. 2004, 279, 43345. (252) Grune, T.; Blasig, I. E.; Sitte, N.; Roloff, B.; Haseloff, R.; Davies, K. J. Peroxynitrite Increases the Degradation of Aconitase and Other Cellular Proteins by Proteasome. J. Biol. Chem. 1998, 273, 10857. (253) Han, D.; Canali, R.; Garcia, J.; Aguilera, R.; Gallaher, T. K.; Cadenas, E. Sites and Mechanisms of Aconitase Inactivation by Peroxynitrite: Modulation by Citrate and Glutathione. Biochemistry 2005, 44, 11986. (254) Kanski, J.; Behring, A.; Pelling, J.; Schoneich, C. Proteomic Identification of 3-Nitrotyrosine-Containing Rat Cardiac Proteins: Effects of Biological Aging. Am. J. Physiol. Heart. Circ. Physiol. 2004, 288, H371. (255) Basso, M.; Samengo, G.; Nardo, G.; Massignan, T.; D’Alessandro, G.; Tartari, S.; Cantoni, L.; Marino, M.; Cheroni, C.; De Biasi, S.; et al. Characterization of Detergent-Insoluble Proteins in Als Indicates a Causal Link between Nitrative Stress and Aggregation in Pathogenesis. PLoS One 2009, 4, e8130. (256) Pearce, L. L.; Kanai, A. J.; Epperly, M. W.; Peterson, J. Nitrosative Stress Results in Irreversible Inhibition of Purified Mitochondrial Complexes I and III without Modification of Cofactors. Nitric Oxide 2005, 13, 254. (257) Pearce, L. L.; Martinez-Bosch, S.; Manzano, E. L.; Winnica, D. E.; Epperly, M. W.; Peterson, J. The Resistance of Electron-Transport Chain Fe-S Clusters to Oxidative Damage During the Reaction of Peroxynitrite with Mitochondrial Complex II and Rat-Heart Pericardium. Nitric Oxide 2009, 20, 135. (258) Ohnishi, T.; Lim, J.; Winter, D. B.; King, T. E. Thermodynamic and Epr Characteristics of a Hipip-Type Iron-Sulfur Center in the Succinate Dehydrogenase of the Respiratory Chain. J. Biol. Chem. 1976, 251, 2105. (259) Scandroglio, F.; Tortora, V.; Radi, R.; Castro, L. Metabolic Control Analysis of Mitochondrial Aconitase: Influence over Respiration and Mitochondrial Superoxide and Hydrogen Peroxide Production. Free Radical Res. 2014, 48, 684. (260) Trujillo, M.; Radi, R. Peroxynitrite Reaction with the Reduced and the Oxidized Forms of Lipoic Acid: New Insights into the

Reaction of Peroxynitrite with Thiols. Arch. Biochem. Biophys. 2002, 397, 91. (261) Bordwell, F. G.; Hughes, D. L. Thiol Acidities and Thiolate Ion Reactivities toward Butyl Chloride in Dimethyl Sulfoxide Solution. The Question of Curvature in Broensted Plots. J. Org. Chem. 1982, 47, 3224. (262) Bulaj, G.; Kortemme, T.; Goldenberg, D. P. IonizationReactivity Relationships for Cysteine Thiols in Polypeptides. Biochemistry 1998, 37, 8965. (263) Roberts, D. D.; Lewis, S. D.; Ballou, D. P.; Olson, S. T.; Shafer, J. A. Reactivity of Small Thiolate Anions and Cysteine-25 in Papain toward Methyl Methanethiosulfonate. Biochemistry 1986, 25, 5595. (264) Danehy, J. P.; Noel, C. J. The Relative Nucleophilic Character of Several Mercaptans toward Ethylene Oxide1. J. Am. Chem. Soc. 1960, 82, 2511. (265) Barany, G.; Merrifield, R. B. Kinetics and Mechanism of the Thiolytic Removal of the Dithiasuccinoyl (Dts) Amino Protecting Group. J. Am. Chem. Soc. 1980, 102, 3084. (266) Whitesides, G. M.; Lilburn, J. E.; Szajewski, R. P. Rates of Thiol-Disulfide Interchange Reactions between Mono- and Dithiols and Ellman’s Reagent. J. Org. Chem. 1977, 42, 332. (267) Trujillo, M.; Clippe, A.; Manta, B.; Ferrer-Sueta, G.; Smeets, A.; Declercq, J. P.; Knoops, B.; Radi, R. Pre-Steady State Kinetic Characterization of Human Peroxiredoxin 5: Taking Advantage of Trp84 Fluorescence Increase Upon Oxidation. Arch. Biochem. Biophys. 2007, 467, 95. (268) Trujillo, M.; Alvarez, B.; Souza, J. M.; Romero, N.; Castro, L.; Thomson, L.; Radi, R. In Nitric Oxide. Biology and Pathobiology; Ignarro, L. J., Ed.; Elsevier: Los Angeles, CA, 2010. (269) Trindade, D. F.; Cerchiaro, G.; Augusto, O. A Role for Peroxymonocarbonate in the Stimulation of Biothiol Peroxidation by the Bicarbonate/Carbon Dioxide Pair. Chem. Res. Toxicol. 2006, 19, 1475. (270) Ferrer-Sueta, G.; Manta, B.; Botti, H.; Radi, R.; Trujillo, M.; Denicola, A. Factors Affecting Protein Thiol Reactivity and Specificity in Peroxide Reduction. Chem. Res. Toxicol. 2011, 24, 434. (271) Trujillo, M.; Ferrer-Sueta, G.; Thomson, L.; Flohe, L.; Radi, R. Kinetics of Peroxiredoxins and Their Role in the Decomposition of Peroxynitrite. Subcell. Biochem. 2007, 44, 83. (272) Selles, B.; Hugo, M.; Trujillo, M.; Srivastava, V.; Wingsle, G.; Jacquot, J. P.; Radi, R.; Rouhier, N. Hydroperoxide and Peroxynitrite Reductase Activity of Poplar Thioredoxin-Dependent Glutathione Peroxidase 5: Kinetics, Catalytic Mechanism and Oxidative Inactivation. Biochem. J. 2012, 442, 369. (273) Alegria, T. G.; Meireles, D. A.; Cussiol, J. R.; Hugo, M.; Trujillo, M.; de Oliveira, M. A.; Miyamoto, S.; Queiroz, R. F.; Valadares, N. F.; Garratt, R. C.; et al. Ohr Plays a Central Role in Bacterial Responses against Fatty Acid Hydroperoxides and Peroxynitrite. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, E132. (274) Trujillo, M.; Alvarez, B.; Radi, R. One- and Two-Electron Oxidation of Thiols: Mechanisms, Kinetics and Biological Fates. Free Radical Res. 2016, 50, 150. (275) Dubuisson, M.; Vander Stricht, D.; Clippe, A.; Etienne, F.; Nauser, T.; Kissner, R.; Koppenol, W. H.; Rees, J. F.; Knoops, B. Human Peroxiredoxin 5 Is a Peroxynitrite Reductase. FEBS Lett. 2004, 571, 161. (276) Zeida, A.; González Lebrero, M. C.; Radi, R.; Trujillo, M.; Estrin, D. A. Mechanism of Cysteine Oxidation by Peroxynitrite: An Integrated Experimental and Theoretical Study. Arch. Biochem. Biophys. 2013, 539, 81. (277) Brigelius-Flohe, R.; Flohe, L. Basic Principles and Emerging Concepts in the Redox Control of Transcription Factors. Antioxid. Redox Signaling 2011, 15, 2335. (278) Ashby, M. T.; Nagy, P. Revisiting a Proposed Kinetic Model for the Reaction of Cysteine and Hydrogen Peroxide Via Cysteine Sulfenic Acid. Int. J. Chem. Kinet. 2007, 39, 32. (279) Flohé, L.; Budde, H.; Bruns, K.; Castro, H.; Clos, J.; Hofmann, B.; Kansal-Kalavar, S.; Krumme, D.; Menge, U.; Plank-Schumacher, K.; et al. Tryparedoxin Peroxidase of Leishmania Donovani: Molecular 1393

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Cloning, Heterologous Expression, Specificity, and Catalytic Mechanism. Arch. Biochem. Biophys. 2002, 397, 324. (280) Rouhier, N.; Gelhaye, E.; Corbier, C.; Jacquot, J. P. Active Site Mutagenesis and Phospholipid Hydroperoxide Reductase Activity of Poplar Type II Peroxiredoxin. Physiol. Plant. 2004, 120, 57. (281) Portillo-Ledesma, S.; Sardi, F.; Manta, B.; Tourn, M. V.; Clippe, A.; Knoops, B.; Alvarez, B.; Coitiño, E. L.; Ferrer-Sueta, G. Deconstructing the Catalytic Efficiency of Peroxiredoxin-5 Peroxidatic Cysteine. Biochemistry 2014, 53, 6113. (282) Pedre, B.; van Bergen, L. A.; Pallo, A.; Rosado, L. A.; Dufe, V. T.; Molle, I. V.; Wahni, K.; Erdogan, H.; Alonso, M.; Proft, F. D.; et al. The Active Site Architecture in Peroxiredoxins: A Case Study on Mycobacterium Tuberculosis Ahpe. Chem. Commun. (Cambridge, U. K.) 2016, 52, 10293. (283) Tairum, C. A., Jr.; de Oliveira, M. A.; Horta, B. B.; Zara, F. J.; Netto, L. E. Disulfide Biochemistry in 2-Cys Peroxiredoxin: Requirement of Glu50 and Arg146 for the Reduction of Yeast Tsa1 by Thioredoxin. J. Mol. Biol. 2012, 424, 28. (284) Hall, A.; Parsonage, D.; Poole, L. B.; Karplus, P. A. Structural Evidence That Peroxiredoxin Catalytic Power Is Based on TransitionState Stabilization. J. Mol. Biol. 2010, 402, 194. (285) Hugo, M.; Turell, L.; Manta, B.; Botti, H.; Monteiro, G.; Netto, L. E.; Alvarez, B.; Radi, R.; Trujillo, M. Thiol and Sulfenic Acid Oxidation of Ahpe, the One-Cysteine Peroxiredoxin from Mycobacterium Tuberculosis: Kinetics, Acidity Constants, and Conformational Dynamics. Biochemistry 2009, 48, 9416. (286) Pineyro, M. D.; Arcari, T.; Robello, C.; Radi, R.; Trujillo, M. Tryparedoxin Peroxidases from Trypanosoma Cruzi: High Efficiency in the Catalytic Elimination of Hydrogen Peroxide and Peroxynitrite. Arch. Biochem. Biophys. 2011, 507, 287. (287) Jaeger, T.; Budde, H.; Flohe, L.; Menge, U.; Singh, M.; Trujillo, M.; Radi, R. Multiple Thioredoxin-Mediated Routes to Detoxify Hydroperoxides in Mycobacterium Tuberculosis. Arch. Biochem. Biophys. 2004, 423, 182. (288) Staudacher, V.; Trujillo, M.; Diederichs, T.; Dick, T. P.; Radi, R.; Morgan, B.; Deponte, M. Redox-Sensitive GFP Fusions for Monitoring the Catalytic Mechanism and Inactivation of Peroxiredoxins in Living Cells. Redox Biol. 2018, 14, 549. (289) Manta, B.; Hugo, M.; Ortiz, C.; Ferrer-Sueta, G.; Trujillo, M.; Denicola, A. The Peroxidase and Peroxynitrite Reductase Activity of Human Erythrocyte Peroxiredoxin 2. Arch. Biochem. Biophys. 2009, 484, 146. (290) Peskin, A. V.; Low, F. M.; Paton, L. N.; Maghzal, G. J.; Hampton, M. B.; Winterbourn, C. C. The High Reactivity of Peroxiredoxin 2 with H2O2 Is Not Reflected in Its Reaction with Other Oxidants and Thiol Reagents. J. Biol. Chem. 2007, 282, 11885. (291) Loumaye, E.; Ferrer-Sueta, G.; Alvarez, B.; Rees, J. F.; Clippe, A.; Knoops, B.; Radi, R.; Trujillo, M. Kinetic Studies of Peroxiredoxin 6 from Arenicola Marina: Rapid Oxidation by Hydrogen Peroxide and Peroxynitrite but Lack of Reduction by Hydrogen Sulfide. Arch. Biochem. Biophys. 2011, 514, 1. (292) Reyes, A. M.; Vazquez, D. S.; Zeida, A.; Hugo, M.; Piñeyro, M. D.; De Armas, M. I.; Estrin, D.; Radi, R.; Santos, J.; Trujillo, M. Prxq B from Mycobacterium Tuberculosis Is a Monomeric, ThioredoxinDependent and Highly Efficient Fatty Acid Hydroperoxide Reductase. Free Radical Biol. Med. 2016, 101, 249. (293) Horta, B. B.; de Oliveira, M. A.; Discola, K. F.; Cussiol, J. R.; Netto, L. E. Structural and Biochemical Characterization of Peroxiredoxin Qbeta from Xylella Fastidiosa: Catalytic Mechanism and High Reactivity. J. Biol. Chem. 2010, 285, 16051. (294) Yuan, Y.; Knaggs, M.; Poole, L.; Fetrow, J.; Salsbury, F., Jr. Conformational and Oligomeric Effects on the Cysteine Pk(a) of Tryparedoxin Peroxidase. J. Biomol. Struct. Dyn. 2010, 28, 51. (295) Toledo, J. C., Jr.; Audi, R.; Ogusucu, R.; Monteiro, G.; Netto, L. E.; Augusto, O. Horseradish Peroxidase Compound I as a Tool to Investigate Reactive Protein-Cysteine Residues: From Quantification to Kinetics. Free Radical Biol. Med. 2011, 50, 1032. (296) Mastronicola, D.; Falabella, M.; Testa, F.; Pucillo, L. P.; Teixeira, M.; Sarti, P.; Saraiva, L. M.; Giuffre, A. Functional

Characterization of Peroxiredoxins from the Human Protozoan Parasite Giardia Intestinalis. PLoS Neglected Trop. Dis. 2014, 8, e2631. (297) Takakura, K.; Beckman, J. S.; MacMillan-Crow, L. A.; Crow, J. P. Rapid and Irreversible Inactivation of Protein Tyrosine Phosphatases Ptp1b, Cd45, and Lar by Peroxynitrite. Arch. Biochem. Biophys. 1999, 369, 197. (298) Wang, P. F.; McLeish, M. J.; Kneen, M. M.; Lee, G.; Kenyon, G. L. An Unusually Low Pk(a) for Cys282 in the Active Site of Human Muscle Creatine Kinase. Biochemistry 2001, 40, 11698. (299) Konorev, E. A.; Hogg, N.; Kalyanaraman, B. Rapid and Irreversible Inhibition of Creatine Kinase by Peroxynitrite. FEBS Lett. 1998, 427, 171. (300) Andres-Mateos, E.; Perier, C.; Zhang, L.; Blanchard-Fillion, B.; Greco, T. M.; Thomas, B.; Ko, H. S.; Sasaki, M.; Ischiropoulos, H.; Przedborski, S.; et al. Dj-1 Gene Deletion Reveals That Dj-1 Is an Atypical Peroxiredoxin-Like Peroxidase. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 14807. (301) Witt, A. C.; Lakshminarasimhan, M.; Remington, B. C.; Hasim, S.; Pozharski, E.; Wilson, M. A. Cysteine Pka Depression by a Protonated Glutamic Acid in Human Dj-1. Biochemistry 2008, 47, 7430. (302) Peixoto, A. S.; Geyer, R. R.; Iqbal, A.; Truzzi, D. R.; Soares Moretti, A. I.; Laurindo, F. R. M.; Augusto, O. Peroxynitrite Preferentially Oxidizes the Dithiol Redox Motifs of Protein Disulfide Isomerase. J. Biol. Chem. 2017, jbc.M117.807016. (303) Souza, J. M.; Radi, R. Glyceraldehyde-3-Phosphate Dehydrogenase Inactivation by Peroxynitrite. Arch. Biochem. Biophys. 1998, 360, 187. (304) Polgar, L. Ion-Pair Formation as a Source of Enhanced Reactivity of the Essential Thiol Group of D-Glyceraldehyde-3Phosphate Dehydrogenase. Eur. J. Biochem. 1975, 51, 63. (305) Dairou, J.; Atmane, N.; Rodrigues-Lima, F.; Dupret, J. M. Peroxynitrite Irreversibly Inactivates the Human Xenobiotic-Metabolizing Enzyme Arylamine N-Acetyltransferase 1 (Nat1) in Human Breast Cancer Cells: A Cellular and Mechanistic Study. J. Biol. Chem. 2004, 279, 7708. (306) Kuhn, D. M.; Geddes, T. J. Peroxynitrite Inactivates Tryptophan Hydroxylase Via Sulfhydryl Oxidation. Coincident Nitration of Enzyme Tyrosyl Residues Has Minimal Impact on Catalytic Activity. J. Biol. Chem. 1999, 274, 29726. (307) Trujillo, M.; Mauri, P.; Benazzi, L.; Comini, M.; De Palma, A.; Flohe, L.; Radi, R.; Stehr, M.; Singh, M.; Ursini, F.; et al. The Mycobacterial Thioredoxin Peroxidase Can Act as a One-Cysteine Peroxiredoxin. J. Biol. Chem. 2006, 281, 20555. (308) Dormeyer, M.; Reckenfelderbaumer, N.; Ludemann, H.; Krauth-Siegel, R. L. Trypanothione-Dependent Synthesis of Deoxyribonucleotides by Trypanosoma Brucei Ribonucleotide Reductase. J. Biol. Chem. 2001, 276, 10602. (309) Alvarez, B.; Ferrer-Sueta, G.; Freeman, B. A.; Radi, R. Kinetics of Peroxynitrite Reaction with Amino Acids and Human Serum Albumin. J. Biol. Chem. 1999, 274, 842. (310) Torres, M. J.; Turell, L.; Botti, H.; Antmann, L.; Carballal, S.; Ferrer-Sueta, G.; Radi, R.; Alvarez, B. Modulation of the Reactivity of the Thiol of Human Serum Albumin and Its Sulfenic Derivative by Fatty Acids. Arch. Biochem. Biophys. 2012, 521, 102. (311) Crawford, M. A.; Tapscott, T.; Fitzsimmons, L. F.; Liu, L.; Reyes, A. M.; Libby, S. J.; Trujillo, M.; Fang, F. C.; Radi, R.; VázquezTorres, A. Redox-Active Sensing by Bacterial DksA Transcription Factors Is Determined by Cysteine and Zinc Content, mBio 2016, 7, e02161-1510.1128/mBio.02161-15. (312) Luo, D.; Smith, S. W.; Anderson, B. D. Kinetics and Mechanism of the Reaction of Cysteine and Hydrogen Peroxide in Aqueous Solution. J. Pharm. Sci. 2005, 94, 304. (313) Turell, L.; Botti, H.; Carballal, S.; Ferrer-Sueta, G.; Souza, J. M.; Duran, R.; Freeman, B. A.; Radi, R.; Alvarez, B. Reactivity of Sulfenic Acid in Human Serum Albumin. Biochemistry 2008, 47, 358. (314) Sohn, J.; Rudolph, J. Catalytic and Chemical Competence of Regulation of Cdc25 Phosphatase by Oxidation/Reduction. Biochemistry 2003, 42, 10060. 1394

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(315) Peskin, A. V.; Winterbourn, C. C. Kinetics of the Reactions of Hypochlorous Acid and Amino Acid Chloramines with Thiols, Methionine, and Ascorbate. Free Radical Biol. Med. 2001, 30, 572. (316) Augusto, O.; Bonini, M. G.; Trindade, D. Spin Trapping of Glutathiyl and Protein Radicals Produced from Nitric Oxide-Derived Oxidants. Free Radical Biol. Med. 2004, 36, 1224. (317) Carballal, S.; Radi, R.; Kirk, M. C.; Barnes, S.; Freeman, B. A.; Alvarez, B. Sulfenic Acid Formation in Human Serum Albumin by Hydrogen Peroxide and Peroxynitrite. Biochemistry 2003, 42, 9906. (318) Rhee, S. G. Redox Signaling: Hydrogen Peroxide as Intracellular Messenger. Exp. Mol. Med. 1999, 31, 53. (319) Flohé, L. Changing Paradigms in Thiology from Antioxidant Defense toward Redox Regulation. Methods Enzymol. 2010, 473, 1. (320) Winterbourn, C. C.; Hampton, M. B. Thiol Chemistry and Specificity in Redox Signaling. Free Radical Biol. Med. 2008, 45, 549. (321) Randall, L. M.; Ferrer-Sueta, G.; Denicola, A. Peroxiredoxins as Preferential Targets in H2O2-Induced Signaling. Methods Enzymol. 2013, 527, 41. (322) Soito, L.; Williamson, C.; Knutson, S. T.; Fetrow, J. S.; Poole, L. B.; Nelson, K. J. Prex: Peroxiredoxin Classification Index, a Database of Subfamily Assignments across the Diverse Peroxiredoxin Family. Nucleic Acids Res. 2011, 39, D332. (323) Daiber, A.; Daub, S.; Bachschmid, M.; Schildknecht, S.; Oelze, M.; Steven, S.; Schmidt, P.; Megner, A.; Wada, M.; Tanabe, T.; et al. Protein Tyrosine Nitration and Thiol Oxidation by PeroxynitriteStrategies to Prevent These Oxidative Modifications. Int. J. Mol. Sci. 2013, 14, 7542. (324) Requejo, R.; Hurd, T. R.; Costa, N. J.; Murphy, M. P. Cysteine Residues Exposed on Protein Surfaces Are the Dominant Intramitochondrial Thiol and May Protect against Oxidative Damage. FEBS J. 2010, 277, 1465. (325) Piacenza, L.; Peluffo, G.; Alvarez, M. N.; Kelly, J. M.; Wilkinson, S. R.; Radi, R. Peroxiredoxins Play a Major Role in Protecting Trypanosoma Cruzi against Macrophage- and Endogenously-Derived Peroxynitrite. Biochem. J. 2008, 410, 359. (326) Trujillo, M.; Ferrer-Sueta, G.; Radi, R. Kinetic Studies on Peroxynitrite Reduction by Peroxiredoxins. Methods Enzymol. 2008, 441, 173. (327) Romero-Puertas, M. C.; Laxa, M.; Matte, A.; Zaninotto, F.; Finkemeier, I.; Jones, A. M.; Perazzolli, M.; Vandelle, E.; Dietz, K. J.; Delledonne, M. S-Nitrosylation of Peroxiredoxin II E Promotes Peroxynitrite-Mediated Tyrosine Nitration. Plant Cell 2007, 19, 4120. (328) Pedrajas, J. R.; Carreras, A.; Valderrama, R.; Barroso, J. B. Mitochondrial 1-Cys-Peroxiredoxin/Thioredoxin System Protects Manganese-Containing Superoxide Dismutase (Mn-Sod) against Inactivation by Peroxynitrite in Saccharomyces Cerevisiae. Nitric Oxide 2010, 23, 206. (329) Master, S. S.; Springer, B.; Sander, P.; Boettger, E. C.; Deretic, V.; Timmins, G. S. Oxidative Stress Response Genes in Mycobacterium Tuberculosis: Role of Ahpc in Resistance to Peroxynitrite and Stage-Specific Survival in Macrophages. Microbiology 2002, 148, 3139. (330) Hu, Y.; Coates, A. R. Acute and Persistent Mycobacterium Tuberculosis Infections Depend on the Thiol Peroxidase Tpx. PLoS One 2009, 4, e5150. (331) Hattori, F.; Murayama, N.; Noshita, T.; Oikawa, S. Mitochondrial Peroxiredoxin-3 Protects Hippocampal Neurons from Excitotoxic Injury in Vivo. J. Neurochem. 2003, 86, 860. (332) Dons, L. E.; Mosa, A.; Rottenberg, M. E.; Rosenkrantz, J. T.; Kristensson, K.; Olsen, J. E. Role of the Listeria Monocytogenes 2-Cys Peroxiredoxin Homologue in Protection against Oxidative and Nitrosative Stress and in Virulence. Pathog. Dis. 2014, 70, 70. (333) Binesse, J.; Lindgren, H.; Lindgren, L.; Conlan, W.; Sjostedt, A. Roles of Reactive Oxygen Species-Degrading Enzymes of Francisella Tularensis Schu S4. Infect. Immun. 2015, 83, 2255. (334) Barr, S. D.; Gedamu, L. Role of Peroxidoxins in Leishmania Chagasi Survival. Evidence of an Enzymatic Defense against Nitrosative Stress. J. Biol. Chem. 2003, 278, 10816. (335) Yang, K. S.; Kang, S. W.; Woo, H. A.; Hwang, S. C.; Chae, H. Z.; Kim, K.; Rhee, S. G. Inactivation of Human Peroxiredoxin I During

Catalysis as the Result of the Oxidation of the Catalytic Site Cysteine to Cysteine-Sulfinic Acid. J. Biol. Chem. 2002, 277, 38029. (336) Wood, Z. A.; Poole, L. B.; Karplus, P. A. Peroxiredoxin Evolution and the Regulation of Hydrogen Peroxide Signaling. Science 2003, 300, 650. (337) Randall, L. M.; Manta, B.; Hugo, M.; Gil, M.; Batthyany, C.; Trujillo, M.; Poole, L. B.; Denicola, A. Nitration Transforms a Sensitive Peroxiredoxin 2 into a More Active and Robust Peroxidase. J. Biol. Chem. 2014, 289, 15536. (338) Randall, L.; Manta, B.; Nelson, K. J.; Santos, J.; Poole, L. B.; Denicola, A. Structural Changes Upon Peroxynitrite-Mediated Nitration of Peroxiredoxin 2; Nitrated Prx2 Resembles Its DisulfideOxidized Form. Arch. Biochem. Biophys. 2016, 590, 101. (339) Peshenko, I. V.; Shichi, H. Oxidation of Active Center Cysteine of Bovine 1-Cys Peroxiredoxin to the Cysteine Sulfenic Acid Form by Peroxide and Peroxynitrite. Free Radical Biol. Med. 2001, 31, 292. (340) de Bem, A. F.; Fiuza, B.; Calcerrada, P.; Brito, P. M.; Peluffo, G.; Dinis, T. C.; Trujillo, M.; Rocha, J. B.; Radi, R.; Almeida, L. M. Protective Effect of Diphenyl Diselenide against PeroxynitriteMediated Endothelial Cell Death: A Comparison with Ebselen. Nitric Oxide 2013, 31, 20. (341) Fiuza, B.; Subelzu, N.; Calcerrada, P.; Straliotto, M. R.; Piacenza, L.; Cassina, A.; Rocha, J. B.; Radi, R.; de Bem, A. F.; Peluffo, G. Impact of Sin-1-Derived Peroxynitrite Flux on Endothelial Cell Redox Homeostasis and Bioenergetics: Protective Role of Diphenyl Diselenide Via Induction of Peroxiredoxins. Free Radical Res. 2015, 49, 122. (342) Storkey, C.; Pattison, D. I.; Ignasiak, M. T.; Schiesser, C. H.; Davies, M. J. Kinetics of Reaction of Peroxynitrite with Selenium- and Sulfur-Containing Compounds: Absolute Rate Constants and Assessment of Biological Significance. Free Radical Biol. Med. 2015, 89, 1049. (343) Stadtman, T. C. Selenocysteine. Annu. Rev. Biochem. 1996, 65, 83. (344) Wessjohann, L. A.; Schneider, A.; Abbas, M.; Brandt, W. Selenium in Chemistry and Biochemistry in Comparison to Sulfur. Biol. Chem. 2007, 388, 997. (345) Padmaja, S.; Squadrito, G. L.; Pryor, W. A. Inactivation of Glutathione Peroxidase by Peroxynitrite. Arch. Biochem. Biophys. 1998, 349, 1. (346) Flohe, L.; Loschen, G.; Gunzler, W. A.; Eichele, E. Glutathione Peroxidase, V. The Kinetic Mechanism. Hoppe-Seyler's Z. Physiol. Chem. 1972, 353, 987. (347) Prabhakar, R.; Morokuma, K.; Musaev, D. G. Peroxynitrite Reductase Activity of Selenoprotein Glutathione Peroxidase: A Computational Study. Biochemistry 2006, 45, 6967. (348) Fu, Y.; Sies, H.; Lei, X. G. Opposite Roles of SeleniumDependent Glutathione Peroxidase-1 in Superoxide Generator Diquatand Peroxynitrite-Induced Apoptosis and Signaling. J. Biol. Chem. 2001, 276, 43004. (349) Knight, T. R.; Kurtz, A.; Bajt, M. L.; Hinson, J. A.; Jaeschke, H. Vascular and Hepatocellular Peroxynitrite Formation During Acetaminophen Toxicity: Role of Mitochondrial Oxidant Stress. Toxicol. Sci. 2001, 62, 212. (350) Knight, T. R.; Ho, Y. S.; Farhood, A.; Jaeschke, H. Peroxynitrite Is a Critical Mediator of Acetaminophen Hepatotoxicity in Murine Livers: Protection by Glutathione. J. Pharmacol. Exp. Ther. 2002, 303, 468. (351) Yun, J. W.; Lum, K.; Lei, X. G. A Novel Upregulation of Glutathione Peroxidase 1 by Knockout of Liver-Regenerating Protein Reg3beta Aggravates Acetaminophen-Induced Hepatic Protein Nitration. Free Radical Biol. Med. 2013, 65, 291. (352) Jozsef, L.; Filep, J. G. Selenium-Containing Compounds Attenuate Peroxynitrite-Mediated NF-kappaB and AP-1 Activation and Interleukin-8 Gene and Protein Expression in Human Leukocytes. Free Radical Biol. Med. 2003, 35, 1018. (353) Arteel, G. E.; Mostert, V.; Oubrahim, H.; Briviba, K.; Abel, J.; Sies, H. Protection by Selenoprotein P in Human Plasma against Peroxynitrite-Mediated Oxidation and Nitration. Biol. Chem. 1998, 379, 1201. 1395

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(354) Sies, H.; Arteel, G. E. Interaction of Peroxynitrite with Selenoproteins and Glutathione Peroxidase Mimics. Free Radical Biol. Med. 2000, 28, 1451. (355) Moore, R. B.; Mankad, M. V.; Shriver, S. K.; Mankad, V. N.; Plishker, G. A. Reconstitution of Ca(2+)-Dependent K+ Transport in Erythrocyte Membrane Vesicles Requires a Cytoplasmic Protein. J. Biol. Chem. 1991, 266, 18964. (356) Cho, C. S.; Kato, G. J.; Yang, S. H.; Bae, S. W.; Lee, J. S.; Gladwin, M. T.; Rhee, S. G. Hydroxyurea-Induced Expression of Glutathione Peroxidase 1 in Red Blood Cells of Individuals with Sickle Cell Anemia. Antioxid. Redox Signaling 2010, 13, 1. (357) Cox, A. G.; Winterbourn, C. C.; Hampton, M. B. Mitochondrial Peroxiredoxin Involvement in Antioxidant Defence and Redox Signalling. Biochem. J. 2010, 425, 313. (358) Peskin, A. V.; Dickerhof, N.; Poynton, R. A.; Paton, L. N.; Pace, P. E.; Hampton, M. B.; Winterbourn, C. C. Hyperoxidation of Peroxiredoxins 2 and 3: Rate Constants for the Reactions of the Sulfenic Acid of the Peroxidatic Cysteine. J. Biol. Chem. 2013, 288, 14170. (359) Perrin, D.; Koppenol, W. H. The Quantitative Oxidation of Methionine to Methionine Sulfoxide by Peroxynitrite. Arch. Biochem. Biophys. 2000, 377, 266. (360) Jensen, J. L.; Miller, B. L.; Zhang, X.; Hug, G. L.; Schöneich, C. Oxidation of Threonylmethionine by Peroxynitrite. Quantification of the One-Electron Transfer Pathway by Comparison to One-Electron Photooxidation. J. Am. Chem. Soc. 1997, 119, 4749. (361) Nakao, L. S.; Iwai, L. K.; Kalil, J.; Augusto, O. Radical Production from Free and Peptide-Bound Methionine Sulfoxide Oxidation by Peroxynitrite and Hydrogen Peroxide/Iron(II). FEBS Lett. 2003, 547, 87. (362) Yashiro, H.; White, R. C.; Yurkovskaya, A. V.; Forbes, M. D. E. Methionine Radical Cation: Structural Studies as a Function of Ph Using X- and Q-Band Time-Resolved Electron Paramagnetic Resonance Spectroscopy. J. Phys. Chem. A 2005, 109, 5855. (363) Spasojević, I.; Bogdanović Pristov, J.; Vujisić, L.; Spasić, M. The Reaction of Methionine with Hydroxyl Radical: Reactive Intermediates and Methanethiol Production. Amino Acids 2012, 42, 2439. (364) Huhmer, A. F.; Gerber, N. C.; Ortiz de Montellano, P. R.; Schoneich, C. Peroxynitrite Reduction of Calmodulin Stimulation of Neuronal Nitric Oxide Synthase. Chem. Res. Toxicol. 1996, 9, 484. (365) Berlett, B. S.; Levine, R. L.; Stadtman, E. R. Carbon Dioxide Stimulates Peroxynitrite-Mediated Nitration of Tyrosine Residues and Inhibits Oxidation of Methionine Residues of Glutamine Synthetase: Both Modifications Mimic Effects of Adenylylation. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 2784. (366) Chen, H. J.; Chen, Y. C. Reactive Nitrogen Oxide SpeciesInduced Post-Translational Modifications in Human Hemoglobin and the Association with Cigarette Smoking. Anal. Chem. 2012, 84, 7881. (367) Khor, H. K.; Fisher, M. T.; Schoneich, C. Potential Role of Methionine Sulfoxide in the Inactivation of the Chaperone Groel by Hypochlorous Acid (Hocl) and Peroxynitrite (ONOO-). J. Biol. Chem. 2004, 279, 19486. (368) Lancellotti, S.; De Filippis, V.; Pozzi, N.; Peyvandi, F.; Palla, R.; Rocca, B.; Rutella, S.; Pitocco, D.; Mannucci, P. M.; De Cristofaro, R. Formation of Methionine Sulfoxide by Peroxynitrite at Position 1606 of Von Willebrand Factor Inhibits Its Cleavage by Adamts-13: A New Prothrombotic Mechanism in Diseases Associated with Oxidative Stress. Free Radical Biol. Med. 2010, 48, 446. (369) Zhang, H.; Zielonka, J.; Sikora, A.; Joseph, J.; Xu, Y.; Kalyanaraman, B. The Effect of Neighboring Methionine Residue on Tyrosine Nitration and Oxidation in Peptides Treated with MPO, H2O2, and NO2(−) or Peroxynitrite and Bicarbonate: Role of Intramolecular Electron Transfer Mechanism? Arch. Biochem. Biophys. 2009, 484, 134. (370) Schoneich, C. Methionine Oxidation by Reactive Oxygen Species: Reaction Mechanisms and Relevance to Alzheimer’s Disease. Biochim. Biophys. Acta, Proteins Proteomics 2005, 1703, 111.

(371) Levine, R. L.; Moskovitz, J.; Stadtman, E. R. Oxidation of Methionine in Proteins: Roles in Antioxidant Defense and Cellular Regulation. IUBMB Life 2000, 50, 301. (372) Weissbach, H.; Etienne, F.; Hoshi, T.; Heinemann, S. H.; Lowther, W. T.; Matthews, B., St; St. John, G.; Nathan, C.; Brot, N. Peptide Methionine Sulfoxide Reductase: Structure, Mechanism of Action, and Biological Function. Arch. Biochem. Biophys. 2002, 397, 172. (373) Fomenko, D. E.; Novoselov, S. V.; Natarajan, S. K.; Lee, B. C.; Koc, A.; Carlson, B. A.; Lee, T. H.; Kim, H. Y.; Hatfield, D. L.; Gladyshev, V. N. Msrb1 (Methionine-R-Sulfoxide Reductase 1) Knock-out Mice: Roles of Msrb1 in Redox Regulation and Identification of a Novel Selenoprotein Form. J. Biol. Chem. 2009, 284, 5986. (374) St. John, G.; Brot, N.; Ruan, J.; Erdjument-Bromage, H.; Tempst, P.; Weissbach, H.; Nathan, C. Peptide Methionine Sulfoxide Reductase from Escherichia Coli and Mycobacterium Tuberculosis Protects Bacteria against Oxidative Damage from Reactive Nitrogen Intermediates. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 9901. (375) Padmaja, S.; Squadrito, G. L.; Lemercier, J. N.; Cueto, R.; Pryor, W. A. Peroxynitrite-Mediated Oxidation of D,L-Selenomethionine: Kinetics, Mechanism and the Role of Carbon Dioxide. Free Radical Biol. Med. 1997, 23, 917. (376) Pietraforte, D.; Minetti, M. Direct Esr Detection or Peroxynitrite-Induced Tyrosine-Centred Protein Radicals in Human Blood Plasma. Biochem. J. 1997, 325, 675. (377) Mahmoudi, L.; Kissner, R.; Nauser, T.; Koppenol, W. H. Electrode Potentials of L-Tryptophan, L-Tyrosine, 3-Nitro-L-Tyrosine, 2,3-Difluoro-L-Tyrosine, and 2,3,5-Trifluoro-L-Tyrosine. Biochemistry 2016, 55, 2849. (378) Chen, S. N.; Hoffman, M. Z. Rate Constants for the Reaction of the Carbonate Radical with Compounds of Biochemical Interest in Neutral Aqueous Solution. Radiat. Res. 1973, 56, 40. (379) Koppenol, W. H.; Bounds, P. L. Signaling by Sulfur-Containing Molecules. Quantitative Aspects. Arch. Biochem. Biophys. 2017, 617, 3. (380) Sala, A.; Nicolis, S.; Roncone, R.; Casella, L.; Monzani, E. Peroxidase Catalyzed Nitration of Tryptophan Derivatives. Mechanism, Products and Comparison with Chemical Nitrating Agents. Eur. J. Biochem. 2004, 271, 2841. (381) Suzuki, T.; Mower, H. F.; Friesen, M. D.; Gilibert, I.; Sawa, T.; Ohshima, H. Nitration and Nitrosation of N-Acetyl-L-Tryptophan and Tryptophan Residues in Proteins by Various Reactive Nitrogen Species. Free Radical Biol. Med. 2004, 37, 671. (382) Yamakura, F.; Ikeda, K. Modification of Tryptophan and Tryptophan Residues in Proteins by Reactive Nitrogen Species. Nitric Oxide 2006, 14, 152. (383) Jantschko, W.; Furtmuller, P. G.; Allegra, M.; Livrea, M. A.; Jakopitsch, C.; Regelsberger, G.; Obinger, C. Redox Intermediates of Plant and Mammalian Peroxidases: A Comparative Transient-Kinetic Study of Their Reactivity toward Indole Derivatives. Arch. Biochem. Biophys. 2002, 398, 12. (384) Nuriel, T.; Hansler, A.; Gross, S. S. Protein Nitrotryptophan: Formation, Significance and Identification. J. Proteomics 2011, 74, 2300. (385) Yamakura, F.; Matsumoto, T.; Fujimura, T.; Taka, H.; Murayama, K.; Imai, T.; Uchida, K. Modification of a Single Tryptophan Residue in Human Cu,Zn-Superoxide Dismutase by Peroxynitrite in the Presence of Bicarbonate. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol. 2001, 1548, 38. (386) Yamakura, F.; Matsumoto, T.; Ikeda, K.; Taka, H.; Fujimura, T.; Murayama, K.; Watanabe, E.; Tamaki, M.; Imai, T.; Takamori, K. Nitrated and Oxidized Products of a Single Tryptophan Residue in Human Cu,Zn-Superoxide Dismutase Treated with Either Peroxynitrite-Carbon Dioxide or Myeloperoxidase-Hydrogen Peroxide-Nitrite. J. Biochem. 2005, 138, 57. (387) Ikeda, K.; Yukihiro Hiraoka, B.; Iwai, H.; Matsumoto, T.; Mineki, R.; Taka, H.; Takamori, K.; Ogawa, H.; Yamakura, F. Detection of 6-Nitrotryptophan in Proteins by Western Blot Analysis 1396

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

and Its Application for Peroxynitrite-Treated Pc12 Cells. Nitric Oxide 2007, 16, 18. (388) Degendorfer, G.; Chuang, C. Y.; Kawasaki, H.; Hammer, A.; Malle, E.; Yamakura, F.; Davies, M. J. Peroxynitrite-Mediated Oxidation of Plasma Fibronectin. Free Radical Biol. Med. 2016, 97, 602. (389) Nuriel, T.; Whitehouse, J.; Ma, Y.; Mercer, E. J.; Brown, N.; Gross, S. S. Ansid: A Solid-Phase Proteomic Approach for Identification and Relative Quantification of Aromatic Nitration Sites. Front. Chem. 2016, 3, 70. (390) Ishii, Y.; Ogara, A.; Katsumata, T.; Umemura, T.; Nishikawa, A.; Iwasaki, Y.; Ito, R.; Saito, K.; Hirose, M.; Nakazawa, H. Quantification of Nitrated Tryptophan in Proteins and Tissues by High-Performance Liquid Chromatography with Electrospray Ionization Tandem Mass Spectrometry. J. Pharm. Biomed. Anal. 2007, 44, 150. (391) Rebrin, I.; Bregere, C.; Kamzalov, S.; Gallaher, T. K.; Sohal, R. S. Nitration of Tryptophan 372 in Succinyl-Coa:3-Ketoacid Coa Transferase During Aging in Rat Heart Mitochondria. Biochemistry 2007, 46, 10130. (392) Bregere, C.; Rebrin, I.; Gallaher, T. K.; Sohal, R. S. Effects of Age and Calorie Restriction on Tryptophan Nitration, Protein Content, and Activity of Succinyl-Coa:3-Ketoacid Coa Transferase in Rat Kidney Mitochondria. Free Radical Biol. Med. 2010, 48, 609. (393) Cong, W.; Ma, W.; Zhao, T.; Zhu, Z.; Wang, Y.; Tan, Y.; Li, X.; Jin, L.; Cai, L. Metallothionein Prevents Diabetes-Induced Cardiac Pathological Changes, Likely Via the Inhibition of Succinyl-Coa:3Ketoacid Coenzyme a Transferase-1 Nitration at Trp(374). Am. J. Phys. Endocr. Metab. 2013, 304, E826. (394) Uda, M.; Kawasaki, H.; Shigenaga, A.; Baba, T.; Yamakura, F. Proteomic Analysis of Endogenous Nitrotryptophan-Containing Proteins in Rat Hippocampus and Cerebellum. Biosci. Rep. 2012, 32, 521. (395) Galetskiy, D.; Lohscheider, J. N.; Kononikhin, A. S.; Popov, I. A.; Nikolaev, E. N.; Adamska, I. Mass Spectrometric Characterization of Photooxidative Protein Modifications in Arabidopsis Thaliana Thylakoid Membranes. Rapid Commun. Mass Spectrom. 2011, 25, 184. (396) Kawasaki, H.; Tominaga, M.; Shigenaga, A.; Kamo, A.; Kamata, Y.; Iizumi, K.; Kimura, U.; Ogawa, H.; Takamori, K.; Yamakura, F. Importance of Tryptophan Nitration of Carbonic Anhydrase III for the Morbidity of Atopic Dermatitis. Free Radical Biol. Med. 2014, 73, 75. (397) Buddha, M. R.; Tao, T.; Parry, R. J.; Crane, B. R. Regioselective Nitration of Tryptophan by a Complex between Bacterial Nitric-Oxide Synthase and Tryptophanyl-tRNA Synthetase. J. Biol. Chem. 2004, 279, 49567. (398) Santos, C. X.; Anjos, E. I.; Augusto, O. Uric Acid Oxidation by Peroxynitrite: Multiple Reactions, Free Radical Formation, and Amplification of Lipid Oxidation. Arch. Biochem. Biophys. 1999, 372, 285. (399) Squadrito, G. L.; Cueto, R.; Splenser, A. E.; Valavanidis, A.; Zhang, H.; Uppu, R. M.; Pryor, W. A. Reaction of Uric Acid with Peroxynitrite and Implications for the Mechanism of Neuroprotection by Uric Acid. Arch. Biochem. Biophys. 2000, 376, 333. (400) Whiteman, M.; Ketsawatsakul, U.; Halliwell, B. A Reassessment of the Peroxynitrite Scavenging Activity of Uric Acid. Ann. N. Y. Acad. Sci. 2002, 962, 242. (401) García-Ruiz, I.; Rodríguez-Juan, C.; Díaz-Sanjuan, T.; del Hoyo, P.; Colina, F.; Muñoz-Yagüe, T.; Solís-Herruzo, J. A. Uric Acid and Anti-Tnf Antibody Improve Mitochondrial Dysfunction in Ob/Ob Mice. Hepatology 2006, 44, 581. (402) Masuda, T.; Shinohara, H.; Kondo, M. J. Radiat. Res. 1975, 16, 153. (403) Ford, E.; Hughes, M. N.; Wardman, P. Kinetics of the Reactions of Nitrogen Dioxide with Glutathione, Cysteine, and Uric Acid at Physiological pH. Free Radical Biol. Med. 2002, 32, 1314. (404) Augusto, O.; Bonini, M. G.; Amanso, A. M.; Linares, E.; Santos, C. C.; De Menezes, S. L. Nitrogen Dioxide and Carbonate Radical Anion: Two Emerging Radicals in Biology. Free Radical Biol. Med. 2002, 32, 841.

(405) Imaram, W.; Gersch, C.; Kim, K. M.; Johnson, R. J.; Henderson, G. N.; Angerhofer, A. Radicals in the Reaction between Peroxynitrite and Uric Acid Identified by Electron Spin Resonance Spectroscopy and Liquid Chromatography Mass Spectrometry. Free Radical Biol. Med. 2010, 49, 275. (406) Levine, M. New Concepts in the Biology and Biochemistry of Ascorbic Acid. N. Engl. J. Med. 1986, 314, 892. (407) Otero, P.; Viana, M.; Herrera, E.; Bonet, B. Antioxidant and Prooxidant Effects of Ascorbic Acid, Dehydroascorbic Acid and Flavonoids on LDL Submitted to Different Degrees of Oxidation. Free Radical Res. 1997, 27, 619. (408) Du, J.; Cullen, J. J.; Buettner, G. R. Ascorbic Acid: Chemistry, Biology and the Treatment of Cancer. Biochim. Biophys. Acta, Rev. Cancer 2012, 1826, 443. (409) Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-Flow Kinetic Study of the Reaction of Ascorbic Acid with Peroxynitrite. Arch. Biochem. Biophys. 1995, 322, 53. (410) Kurz, C. R.; Kissner, R.; Nauser, T.; Perrin, D.; Koppenol, W. H. Rapid Scavenging of Peroxynitrous Acid by Monohydroascorbate. Free Radical Biol. Med. 2003, 35, 1529. (411) Huie, R. E.; Shoute, L. C. T.; Neta, P. Temperature Dependence of the Rate Constants for Reactions of the Carbonate Radical with Organic and Inorganic Reductants. Int. J. Chem. Kinet. 1991, 23, 541. (412) Alfassi, Z. B.; Huie, R. E.; Neta, P.; Shoute, L. C. T. Temperature Dependence of the Rate Constants for Reaction of Inorganic Radicals with Organic Reductants. J. Phys. Chem. 1990, 94, 8800. (413) Vásquez-Vivar, J.; Santos, A. M.; Junqueira, V. B.; Augusto, O. Peroxynitrite-Mediated Formation of Free Radicals in Human Plasma: Epr Detection of Ascorbyl, Albumin-Thiyl and Uric Acid-Derived Free Radicals. Biochem. J. 1996, 314, 869. (414) Scorza, G.; Minetti, M. One-Electron Oxidation Pathway of Thiols by Peroxynitrite in Biological Fluids: Bicarbonate and Ascorbate Promote the Formation of Albumin Disulphide Dimers in Human Blood Plasma. Biochem. J. 1998, 329, 405. (415) Guidarelli, A.; De Sanctis, R.; Cellini, B.; Fiorani, M.; Dacha, M.; Cantoni, O. Intracellular Ascorbic Acid Enhances the DNA SingleStrand Breakage and Toxicity Induced by Peroxynitrite in U937 Cells. Biochem. J. 2001, 356, 509. (416) Fiorani, M.; Azzolini, C.; Cerioni, L.; Guidarelli, A.; Cantoni, O. Superoxide Dictates the Mode of U937 Cell Ascorbic Acid Uptake and Prevents the Enhancing Effects of the Vitamin to Otherwise Nontoxic Levels of Reactive Oxygen/Nitrogen Species. J. Nutr. Biochem. 2013, 24, 467. (417) Kuzkaya, N.; Weissmann, N.; Harrison, D. G.; Dikalov, S. Interactions of Peroxynitrite with Uric Acid in the Presence of Ascorbate and Thiols: Implications for Uncoupling Endothelial Nitric Oxide Synthase. Biochem. Pharmacol. 2005, 70, 343. (418) Monteiro, G.; Horta, B. B.; Pimenta, D. C.; Augusto, O.; Netto, L. E. Reduction of 1-Cys Peroxiredoxins by Ascorbate Changes the Thiol-Specific Antioxidant Paradigm, Revealing Another Function of Vitamin C. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 4886. (419) Goldstein, S.; Czapski, G. Reactivity of Peroxynitrite Versus Simultaneous Generation of (*)NOoand O(2)(*)(−) toward NADH. Chem. Res. Toxicol. 2000, 13, 736. (420) Kirsch, M.; de Groot, H. Reaction of Peroxynitrite with Reduced Nicotinamide Nucleotides, the Formation of Hydrogen Peroxide. J. Biol. Chem. 1999, 274, 24664. (421) Land, E. J.; Swallow, A. J. One-Electron Reactions in Biochemical Systems as Studied by Pulse Radiolysis. I. Nicotinamide-Adenine Dinucleotide and Related Compounds. Biochim. Biophys. Acta, Bioenerg. 1968, 162, 327. (422) Petersen Shay, K.; Moreau, R. F.; Smith, E. J.; Hagen, T. M. Is Alpha-Lipoic Acid a Scavenger of Reactive Oxygen Species in Vivo? Evidence for Its Initiation of Stress Signaling Pathways That Promote Endogenous Antioxidant Capacity. IUBMB Life 2008, 60, 362. 1397

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(423) Handelman, G. J.; Han, D.; Tritschler, H.; Packer, L. AlphaLipoic Acid Reduction by Mammalian Cells to the Dithiol Form, and Release into the Culture Medium. Biochem. Pharmacol. 1994, 47, 1725. (424) Richards, E. M.; Rosenthal, R. E.; Kristian, T.; Fiskum, G. Postischemic Hyperoxia Reduces Hippocampal Pyruvate Dehydrogenase Activity. Free Radical Biol. Med. 2006, 40, 1960. (425) Chen, W.; Druhan, L. J.; Chen, C. A.; Hemann, C.; Chen, Y. R.; Berka, V.; Tsai, A. L.; Zweier, J. L. Peroxynitrite Induces Destruction of the Tetrahydrobiopterin and Heme in Endothelial Nitric Oxide Synthase: Transition from Reversible to Irreversible Enzyme Inhibition. Biochemistry 2010, 49, 3129. (426) Kuzkaya, N.; Weissmann, N.; Harrison, D. G.; Dikalov, S. Interactions of Peroxynitrite, Tetrahydrobiopterin, Ascorbic Acid, and Thiols: Implications for Uncoupling Endothelial Nitric-Oxide Synthase. J. Biol. Chem. 2003, 278, 22546. (427) Milstien, S.; Katusic, Z. Oxidation of Tetrahydrobiopterin by Peroxynitrite: Implications for Vascular Endothelial Function. Biochem. Biophys. Res. Commun. 1999, 263, 681. (428) Franco, M. do Carmo P.; Fortes, Z. B.; Akamine, E. H.; Kawamoto, E. M.; Scavone, C.; de Britto, L. R.; Muscara, M. N.; Teixeira, S. A.; Tostes, R. C.; Carvalho, M. H.; et al. Tetrahydrobiopterin Improves Endothelial Dysfunction and Vascular Oxidative Stress in Microvessels of Intrauterine Undernourished Rats. J. Physiol. 2004, 558, 239. (429) Nakamura, M.; Nagayoshi, R.; Ijiri, K.; Nakashima-Matsushita, N.; Takeuchi, T.; Matsuyama, T. Nitration and Chlorination of Folic Acid by Peroxynitrite and Hypochlorous Acid, and the Selective Binding of 10-Nitro-Folate to Folate Receptor Beta. Biochem. Biophys. Res. Commun. 2002, 297, 1238. (430) Uppu, R. M.; Winston, G. W.; Pryor, W. A. Reactions of Peroxynitrite with Aldehydes as Probes for the Reactive Intermediates Responsible for Biological Nitration. Chem. Res. Toxicol. 1997, 10, 1331. (431) Nakao, L. S.; Ouchi, D.; Augusto, O. Oxidation of Acetaldehyde by Peroxynitrite and Hydrogen Peroxide/Iron(II). Production of Acetate, Formate, and Methyl Radicals. Chem. Res. Toxicol. 1999, 12, 1010. (432) Merenyi, G.; Lind, J.; Goldstein, S. The Rate of Homolysis of Adducts of Peroxynitrite to the CO Double Bond. J. Am. Chem. Soc. 2002, 124, 40. (433) Vasquez-Vivar, J.; Denicola, A.; Radi, R.; Augusto, O. Peroxynitrite-Mediated Decarboxylation of Pyruvate to Both Carbon Dioxide and Carbon Dioxide Radical Anion. Chem. Res. Toxicol. 1997, 10, 786. (434) Schöpfer, F.; Riobó, N.; Carreras, M. C.; Alvarez, B.; Radi, R.; Boveris, A.; Cadenas, E.; Poderoso, J. J. Oxidation of Ubiquinol by Peroxynitrite: Implications for Protection of Mitochondria against Nitrosative Damage. Biochem. J. 2000, 349, 35. (435) Romero, N.; Radi, R.; Linares, E.; Augusto, O.; Detweiler, C. D.; Mason, R. P.; Denicola, A. Reaction of Human Hemoglobin with Peroxynitrite. Isomerization to Nitrate and Secondary Formation of Protein Radicals. J. Biol. Chem. 2003, 278, 44049. (436) Stern, M. K.; Jensen, M. J.; Kramer, K. Peroxynitrite Decomposition Catalysts. J. Am. Chem. Soc. 1996, 118, 8735. (437) Lee, J.; Hunt, J. A.; Groves, J. T. Mechanism of Iron Porphyrin Reactions with Peroxynitrite. J. Am. Chem. Soc. 1998, 120, 7493. (438) Jensen, M. P.; Riley, D. P. Peroxynitrite Decomposition Activity of Iron Porphyrin Complexes. Inorg. Chem. 2002, 41, 4788. (439) Cuzzocrea, S.; Mazzon, E.; Di Paola, R.; Esposito, E.; Macarthur, H.; Matuschak, G. M.; Salvemini, D. A Role for Nitric Oxide-Mediated Peroxynitrite Formation in a Model of EndotoxinInduced Shock. J. Pharmacol. Exp. Ther. 2006, 319, 73. (440) Muscoli, C.; Cuzzocrea, S.; Ndengele, M. M.; Mollace, V.; Porreca, F.; Fabrizi, F.; Esposito, E.; Masini, E.; Matuschak, G. M.; Salvemini, D. Therapeutic Manipulation of Peroxynitrite Attenuates the Development of Opiate-Induced Antinociceptive Tolerance in Mice. J. Clin. Invest. 2007, 117, 3530. (441) Ding, R.; Feng, L.; He, L.; Chen, Y.; Wen, P.; Fu, Z.; Lin, C.; Yang, S.; Deng, X.; Zeng, J.; et al. Peroxynitrite Decomposition

Catalyst Prevents Matrix Metalloproteinase-9 Activation and Neurovascular Injury after Hemoglobin Injection into the Caudate Nucleus of Rats. Neuroscience 2015, 297, 182. (442) Trujillo, M.; Ferrer-Sueta, G.; Radi, R. Peroxynitrite Detoxification and Its Biologic Implications. Antioxid. Redox Signaling 2008, 10, 1607. (443) Parnham, M. J.; Kindt, S. A Novel Biologically Active SelenoOrganic Compound–III. Effects of Pz 51 (Ebselen) on Glutathione Peroxidase and Secretory Activities of Mouse Macrophages. Biochem. Pharmacol. 1984, 33, 3247. (444) Bhabak, K. P.; Vernekar, A. A.; Jakka, S. R.; Roy, G.; Mugesh, G. Mechanistic Investigations on the Efficient Catalytic Decomposition of Peroxynitrite by Ebselen Analogues. Org. Biomol. Chem. 2011, 9, 5193. (445) Arteel, G. E.; Briviba, K.; Sies, H. Function of Thioredoxin Reductase as a Peroxynitrite Reductase Using Selenocystine or Ebselen. Chem. Res. Toxicol. 1999, 12, 264. (446) Wagner, G.; Schuch, G.; Akerboom, T. P.; Sies, H. Transport of Ebselen in Plasma and Its Transfer to Binding Sites in the Hepatocyte. Biochem. Pharmacol. 1994, 48, 1137. (447) Daiber, A.; Zou, M. H.; Bachschmid, M.; Ullrich, V. Ebselen as a Peroxynitrite Scavenger in Vitro and Ex Vivo. Biochem. Pharmacol. 2000, 59, 153. (448) Choi, M. H.; Sajed, D.; Poole, L.; Hirata, K.; Herdman, S.; Torian, B. E.; Reed, S. L. An Unusual Surface Peroxiredoxin Protects Invasive Entamoeba Histolytica from Oxidant Attack. Mol. Biochem. Parasitol. 2005, 143, 80. (449) Puntel, R. L.; Roos, D. H.; Seeger, R. L.; Aschner, M.; Rocha, J. B. Organochalcogens Inhibit Mitochondrial Complexes I and II in Rat Brain: Possible Implications for Neurotoxicity. Neurotoxic. Res. 2013, 24, 109. (450) Bhabak, K. P.; Satheeshkumar, K.; Jayavelu, S.; Mugesh, G. Inhibition of Peroxynitrite- and Peroxidase-Mediated Protein Tyrosine Nitration by Imidazole-Based Thiourea and Selenourea Derivatives. Org. Biomol. Chem. 2011, 9, 7343. (451) Bhabak, K. P.; Mugesh, G. Synthesis, Characterization, and Antioxidant Activity of Some Ebselen Analogues. Chem. - Eur. J. 2007, 13, 4594. (452) Bem, A. F.; Fiuza, B.; Calcerrada, P.; Brito, P. M.; Peluffo, G.; Dinis, T. C.; Trujillo, M.; Rocha, J. B.; Radi, R.; Almeida, L. M. Protective Effect of Diphenyl Diselenide against PeroxynitriteMediated Endothelial Cell Death: A Comparison with Ebselen. Nitric Oxide 2013, 31, 20. (453) Kachadourian, R.; Johnson, C. A.; Min, E.; Spasojevic, I.; Day, B. J. Flavin-Dependent Antioxidant Properties of a New Series of Meso-N,N′-Dialkyl-Imidazolium Substituted Manganese(III) Porphyrins. Biochem. Pharmacol. 2004, 67, 77. (454) Ferrer-Sueta, G.; Hannibal, L.; Batinic-Haberle, I.; Radi, R. Reduction of Manganese Porphyrins by Flavoenzymes and Submitochondrial Particles: A Catalytic Cycle for the Reduction of Peroxynitrite. Free Radical Biol. Med. 2006, 41, 503. (455) Valez, V.; Cassina, A.; Batinic-Haberle, I.; Kalyanaraman, B.; Ferrer-Sueta, G.; Radi, R. Peroxynitrite Formation in Nitric OxideExposed Submitochondrial Particles: Detection, Oxidative Damage and Catalytic Removal by Mn-Porphyrins. Arch. Biochem. Biophys. 2013, 529, 45. (456) Weitner, T.; Kos, I.; Mandic, Z.; Batinic-Haberle, I.; Birus, M. Acid-Base and Electrochemical Properties of Manganese Meso(Orthoand Meta-N-Ethylpyridyl)Porphyrins: Voltammetric and Chronocoulometric Study of Protolytic and Redox Equilibria. Dalton Trans. 2013, 42, 14757. (457) Spasojevic, I.; Chen, Y.; Noel, T. J.; Yu, Y.; Cole, M. P.; Zhang, L.; Zhao, Y.; St. Clair, D. K.; Batinic-Haberle, I. Mn Porphyrin-Based Superoxide Dismutase (Sod) Mimic, MnIIITE-2-PyP5+, Targets Mouse Heart Mitochondria. Free Radical Biol. Med. 2007, 42, 1193. (458) Groves, J. T.; Marla, S. S. Peroxynitrite -Induced DNA Strand Scission Mediated by a Manganese Porphyrin. J. Am. Chem. Soc. 1995, 117, 9578. 1398

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(459) Ferrer-Sueta, G.; Ruiz-Ramirez, L.; Radi, R. Ternary Copper Complexes and Manganese (III) Tetrakis(4-Benzoic Acid) Porphyrin Catalyze Peroxynitrite-Dependent Nitration of Aromatics. Chem. Res. Toxicol. 1997, 10, 1338. (460) Lee, J.; Hunt, J.; Groves, J. T. Rapid Decomposition of Peroxynitrite by Manganese Porphyrin-Antioxidant Redox Couples. Bioorg. Med. Chem. Lett. 1997, 7, 2913. (461) Ferrer-Sueta, G.; Batinic-Haberle, I.; Spasojevic, I.; Fridovich, I.; Radi, R. Catalytic Scavenging of Peroxynitrite by Isomeric Mn(III) N- Methylpyridylporphyrins in the Presence of Reductants. Chem. Res. Toxicol. 1999, 12, 442. (462) Ferrer-Sueta, G.; Vitturi, D.; Batinic-Haberle, I.; Fridovich, I.; Goldstein, S.; Czapski, G.; Radi, R. Reactions of Manganese Porphyrins with Peroxynitrite and Carbonate Radical Anion. J. Biol. Chem. 2003, 278, 27432. (463) Trostchansky, A.; Ferrer-Sueta, G.; Batthyany, C.; Botti, H.; Batinic-Haberle, I.; Radi, R.; Rubbo, H. Peroxynitrite Flux-Mediated LDL Oxidation Is Inhibited by Manganese Porphyrins in the Presence of Uric Acid. Free Radical Biol. Med. 2003, 35, 1293. (464) Spasojevic, I.; Chen, Y.; Noel, T. J.; Fan, P.; Zhang, L.; Reboucas, J. S., St; St. Clair, D. K.; Batinic-Haberle, I. Pharmacokinetics of the Potent Redox-Modulating Manganese Porphyrin, MnTE2-PyP(5+), in Plasma and Major Organs of B6c3f1Mice. Free Radical Biol. Med. 2008, 45, 943. (465) Czapski, G.; Lymar, S. V.; Schwarz, H. A. Acidity of the Carbonate Radical. J. Phys. Chem. A 1999, 103, 3447. (466) Huie, R. E.; Clifton, C. L.; Neta, P. Electron-Transfer Reaction Rates and Equilibria of the Carbonate and Sulfate Radical Anions. Radiat. Phys. Chem. 1991, 38, 477. (467) Lymar, S. V.; Schwarz, H. A.; Czapski, G. Medium Effects on Reactions of the Carbonate Radical with Thiocyanide, Iodide and Ferrocyanide Ions. Radiat. Phys. Chem. 2000, 59, 387. (468) Bonini, M. G.; Augusto, O. Carbon Dioxide Stimulates the Production of Thiyl, Sulfinyl, and Disulfide Radical Anion from Thiol Oxidation by Peroxynitrite. J. Biol. Chem. 2001, 276, 9749. (469) Goss, S. P.; Singh, R. J.; Kalyanaraman, B. Bicarbonate Enhances the Peroxidase Activity of Cu,Zn-Superoxide Dismutase. Role of Carbonate Anion Radical. J. Biol. Chem. 1999, 274, 28233. (470) Liochev, S. I.; Fridovich, I. On the Role of Bicarbonate in Peroxidations Catalyzed by Cu,Zn Superoxide Dismutase. Free Radical Biol. Med. 1999, 27, 1444. (471) Liochev, S. I.; Fridovich, I. Co2, Not Hco3-, Facilitates Oxidations by Cu,Zn Superoxide Dismutase Plus H2O2. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 743. (472) Bonini, M. G.; Miyamoto, S.; Di Mascio, P.; Augusto, O. Production of the Carbonate Radical Anion During Xanthine Oxidase Turnover in the Presence of Bicarbonate. J. Biol. Chem. 2004, 279, 51836. (473) Lilie, J.; Hanrahan, R. J.; Henglein, A. O-Transfer Reactions of the Carbonate Radical Anion. Radiat. Phys. Chem. 1978, 11, 225. (474) Baseggio Conrado, A.; D’Angelantonio, M.; Torreggiani, A.; Pecci, L.; Fontana, M. Reactivity of Hypotaurine and Cysteine Sulfinic Acid toward Carbonate Radical Anion and Nitrogen Dioxide as Explored by the Peroxidase Activity of Cu,Zn Superoxide Dismutase and by Pulse Radiolysis. Free Radical Res. 2014, 48, 1300. (475) Shafirovich, V.; Dourandin, A.; Huang, W.; Geacintov, N. E. The Carbonate Radical Is a Site-Selective Oxidizing Agent of Guanine in Double-Stranded Oligonucleotides. J. Biol. Chem. 2001, 276, 24621. (476) Dassanayake, R. S.; Shelley, J. T.; Cabelli, D. E.; Brasch, N. E. Pulse Radiolysis and Ultra-High-Performance Liquid Chromatography/High-Resolution Mass Spectrometry Studies on the Reactions of the Carbonate Radical with Vitamin B12 Derivatives. Chem. - Eur. J. 2015, 21, 6409. (477) Domazou, A. S.; Koppenol, W. H. Oxidation-State-Dependent Reactions of Cytochrome C with the Trioxidocarbonate(*1-) Radical: A Pulse Radiolysis Study. JBIC, J. Biol. Inorg. Chem. 2006, 12, 118. (478) Boccini, F.; Domazou, A. S.; Herold, S. Pulse Radiolysis Studies of the Reactions of Carbonate Radical Anion with Myoglobin and Hemoglobin. J. Phys. Chem. A 2004, 108, 5800.

(479) Trujillo, M.; Folkes, L.; Bartesaghi, S.; Kalyanaraman, B.; Wardman, P.; Radi, R. Peroxynitrite-Derived Carbonate and Nitrogen Dioxide Radicals Readily React with Lipoic and Dihydrolipoic Acid. Free Radical Biol. Med. 2005, 39, 279. (480) Zhao, C. Y.; Shi, Y. M.; Yao, S. D.; Jia, Z. J.; Fan, B. T.; Wang, W. F.; Lin, W. Z.; Lin, N. Y.; Zheng, R. L. Scavenging Effects of Natural Phenols on Oxidizing Intermediates of Peroxynitrite. Die Pharmazie 2003, 58, 742. (481) Simic, M. G.; Hunter, E. P. L. The Reactivities of Organic Oxygen (Oxy) Radicals; Walter de Gruyter: Berlin, 1984. (482) Goldstein, S.; Samuni, A.; Merenyi, G. Reactions of Nitric Oxide, Peroxynitrite, and Carbonate Radicals with Nitroxides and Their Corresponding Oxoammonium Cations. Chem. Res. Toxicol. 2004, 17, 250. (483) Krishna, M. C.; Samuni, A. Nitroxides as Antioxidants. Methods Enzymol. 1994, 234, 580. (484) Linares, E.; Seixas, L. V.; dos Prazeres, J. N.; Ladd, F. V.; Ladd, A. A.; Coppi, A. A.; Augusto, O. Tempol Moderately Extends Survival in a HSOD1(G93a) Als Rat Model by Inhibiting Neuronal Cell Loss, Oxidative Damage and Levels of Non-Native HSOD1(G93a) Forms. PLoS One 2013, 8, e55868. (485) Offer, T.; Russo, A.; Samuni, A. The Pro-Oxidative Activity of SOD and Nitroxide SOD Mimics. FASEB J. 2000, 14, 1215. (486) Thiemermann, C. Membrane-Permeable Radical Scavengers (Tempol) for Shock, Ischemia-Reperfusion Injury, and Inflammation. Crit. Care Med. 2003, 31, S76. (487) Bonini, M. G.; Mason, R. P.; Augusto, O. The Mechanism by Which 4-Hydroxy-2,2,6,6-Tetramethylpiperidene-1-Oxyl (Tempol) Diverts Peroxynitrite Decomposition from Nitrating to Nitrosating Species. Chem. Res. Toxicol. 2002, 15, 506. (488) Fernandes, D. C.; Medinas, D. B.; Alves, M. J.; Augusto, O. Tempol Diverts Peroxynitrite/Carbon Dioxide Reactivity toward Albumin and Cells from Protein-Tyrosine Nitration to ProteinCysteine Nitrosation. Free Radical Biol. Med. 2005, 38, 189. (489) Behar, D.; Czapski, G.; Duchovny, I. Carbonate Radical in Flash Photolysis and Pulse Radiolysis of Aqueous Carbonate Solutions. J. Phys. Chem. 1970, 74, 2206. (490) Hayon, E.; McGarvey, J. J. Flash Photolysis in the Vacuum Ultraviolet Region of Sulfate, Carbonate, and Hydroxyl Ions in Aqueous Solutions. J. Phys. Chem. 1967, 71, 1472. (491) Alvarez, M. N.; Peluffo, G.; Folkes, L.; Wardman, P.; Radi, R. Reaction of the Carbonate Radical with the Spin-Trap 5,5-Dimethyl-1Pyrroline-N-Oxide in Chemical and Cellular Systems: Pulse Radiolysis, Electron Paramagnetic Resonance, and Kinetic-Competition Studies. Free Radical Biol. Med. 2007, 43, 1523. (492) Wrona, M.; Patel, K.; Wardman, P. Reactivity of 2′,7′Dichlorodihydrofluorescein and Dihydrorhodamine 123 and Their Oxidized Forms toward Carbonate, Nitrogen Dioxide, and Hydroxyl Radicals. Free Radical Biol. Med. 2005, 38, 262. (493) Zhang, H.; Joseph, J.; Felix, C.; Kalyanaraman, B. Bicarbonate Enhances the Hydroxylation, Nitration, and Peroxidation Reactions Catalyzed by Copper, Zinc Superoxide Dismutase. Intermediacy of Carbonate Anion Radical. J. Biol. Chem. 2000, 275, 14038. (494) Crean, C.; Lee, Y. A.; Yun, B. H.; Geacintov, N. E.; Shafirovich, V. Oxidation of Guanine by Carbonate Radicals Derived from Photolysis of Carbonatotetramminecobalt(III) Complexes and the pH Dependence of Intrastrand DNA Cross-Links Mediated by Guanine Radical Reactions. ChemBioChem 2008, 9, 1985. (495) Medinas, D. B.; Gozzo, F. C.; Santos, L. F.; Iglesias, A. H.; Augusto, O. A Ditryptophan Cross-Link Is Responsible for the Covalent Dimerization of Human Superoxide Dismutase 1 During Its Bicarbonate-Dependent Peroxidase Activity. Free Radical Biol. Med. 2010, 49, 1046. (496) Zhang, H.; Andrekopoulos, C.; Joseph, J.; Chandran, K.; Karoui, H.; Crow, J. P.; Kalyanaraman, B. Bicarbonate-Dependent Peroxidase Activity of Human Cu,Zn-Superoxide Dismutase Induces Covalent Aggregation of Protein: Intermediacy of TryptophanDerived Oxidation Products. J. Biol. Chem. 2003, 278, 24078. 1399

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(517) Bartesaghi, S.; Romero, N.; Radi, R. Nitric Oxide and Derived Oxidants; Nova Science Publishers, Inc., 2011. (518) Nottingham, W. C.; Sutter, J. R. Kinetics of the Oxidation of Nitric Oxide by Chlorine and Oxygen in Nonaqueous Media. Int. J. Chem. Kinet. 1986, 18, 1289. (519) Burner, U.; Furtmuller, P. G.; Kettle, A. J.; Koppenol, W. H.; Obinger, C. Mechanism of Reaction of Myeloperoxidase with Nitrite. J. Biol. Chem. 2000, 275, 20597. (520) Marquez, L. A.; Dunford, H. B. Kinetics of Oxidation of Tyrosine and Dityrosine by Myeloperoxidase Compounds I and II. Implications for Lipoprotein Peroxidation Studies. J. Biol. Chem. 1995, 270, 30434. (521) Dunford, H. B. One-Electron Oxidations by Peroxidases. Xenobiotica 1995, 25, 725. (522) Logager, T.; Sehested, K. Formation and Decay of Peroxynitric Acid: A Pulse Radiolysis Study. J. Phys. Chem. 1993, 97, 10047. (523) Huie, R. E. The Reaction Kinetics of NO2(.). Toxicology 1994, 89, 193. (524) Prutz, W. A.; Monig, H.; Butler, J.; Land, E. J. Reactions of Nitrogen Dioxide in Aqueous Model Systems: Oxidation of Tyrosine Units in Peptides and Proteins. Arch. Biochem. Biophys. 1985, 243, 125. (525) Jiang, H.; Kruger, N.; Lahiri, D. R.; Wang, D.; Vatele, J. M.; Balazy, M. Nitrogen Dioxide Induces Cis-Trans-Isomerization of Arachidonic Acid within Cellular Phospholipids. Detection of TransArachidonic Acids in Vivo. J. Biol. Chem. 1999, 274, 16235. (526) Titov, A. I. Free Radical Mechanism of Nitration. Tetrahedron 1963, 19, 557. (527) Zghibeh, C. M.; Raj Gopal, V.; Poff, C. D.; Falck, J. R.; Balazy, M. Determination of Trans-Arachidonic Acid Isomers in Human Blood Plasma. Anal. Biochem. 2004, 332, 137. (528) Cui, T.; Schopfer, F. J.; Zhang, J.; Chen, K.; Ichikawa, T.; Baker, P. R.; Batthyany, C.; Chacko, B. K.; Feng, X.; Patel, R. P.; et al. Nitrated Fatty Acids: Endogenous Anti-Inflammatory Signaling Mediators. J. Biol. Chem. 2006, 281, 35686. (529) Ferreira, A. M.; Ferrari, M. I.; Trostchansky, A.; Batthyany, C.; Souza, J. M.; Alvarez, M. N.; Lopez, G. V.; Baker, P. R.; Schopfer, F. J.; O’Donnell, V.; et al. Macrophage Activation Induces Formation of the Anti-Inflammatory Lipid Cholesteryl-Nitrolinoleate. Biochem. J. 2009, 417, 223. (530) Villacorta, L.; Gao, Z.; Schopfer, F. J.; Freeman, B. A.; Chen, Y. E. Nitro-Fatty Acids in Cardiovascular Regulation and Diseases: Characteristics and Molecular Mechanisms. Front. Biosci. (Landmark Ed) 2016, 21, 873. (531) Forni, L. G.; Mora-Arellano, V. O.; Packer, J. E.; Willson, R. L. Nitrogen Dioxide and Related Free Radicals: Electron-Transfer Reactions with Organic Compounds in Solutions Containing Nitrite or Nitrate. J. Chem. Soc., Perkin Trans. 2 1986, 1. (532) Pryor, W. A.; Lightsey, J. W. Mechanisms of Nitrogen Dioxide Reactions: Initiation of Lipid Peroxidation and the Production of Nitrous Acid. Science 1981, 214, 435. (533) Pryor, W. A.; Lightsey, J. W.; Church, D. F. Reaction of Nitrogen Dioxide with Alkenes and Polyunsaturated Fatty Acids: Addition and Hydrogen-Abstraction Mechanisms. J. Am. Chem. Soc. 1982, 104, 6685. (534) Folkes, L. K.; Patel, K. B.; Wardman, P.; Wrona, M. Kinetics of Reaction of Nitrogen Dioxide with Dihydrorhodamine and the Reaction of the Dihydrorhodamine Radical with Oxygen: Implications for Quantifying Peroxynitrite Formation in Cells. Arch. Biochem. Biophys. 2009, 484, 122. (535) Health Effects of Outdoor Air Pollution. Part 2. Committee of the Environmental and Occupational Health Assembly of the American Thoracic Society. Am. J. Respir. Crit. Care Med. 1996, 153, 477. (536) Sauerbeck, G.; Parkin, C. S.; Houston, S.; Whall, C.; Newton, A.; Carlyle, J. Nitrogen Dioxide and Particle Pollution near Trunk Roads and in Towns of the South Midlands in England. J. R. Soc. Promot. Health 2000, 120, 183.

(497) Zhang, H.; Joseph, J.; Crow, J.; Kalyanaraman, B. Mass Spectral Evidence for Carbonate-Anion-Radical-Induced Posttranslational Modification of Tryptophan to Kynurenine in Human Cu, Zn Superoxide Dismutase. Free Radical Biol. Med. 2004, 37, 2018. (498) Queiroz, R. F.; Paviani, V.; Coelho, F. R.; Marques, E. F.; Di Mascio, P.; Augusto, O. The Carbonylation and Covalent Dimerization of Human Superoxide Dismutase 1 Caused by Its BicarbonateDependent Peroxidase Activity Is Inhibited by the Radical Scavenger Tempol. Biochem. J. 2013, 455, 37. (499) Paviani, V.; Queiroz, R. F.; Marques, E. F.; Di Mascio, P.; Augusto, O. Production of Lysozyme and Lysozyme-Superoxide Dismutase Dimers Bound by a Ditryptophan Cross-Link in Carbonate Radical-Treated Lysozyme. Free Radical Biol. Med. 2015, 89, 72. (500) Radi, R.; Peluffo, G.; Alvarez, M. N.; Naviliat, M.; Cayota, A. Unraveling Peroxynitrite Formation in Biological Systems. Free Radical Biol. Med. 2001, 30, 463. (501) Santos, C. X.; Bonini, M. G.; Augusto, O. Role of the Carbonate Radical Anion in Tyrosine Nitration and Hydroxylation by Peroxynitrite. Arch. Biochem. Biophys. 2000, 377, 146. (502) Surmeli, N. B.; Litterman, N. K.; Miller, A. F.; Groves, J. T. Peroxynitrite Mediates Active Site Tyrosine Nitration in Manganese Superoxide Dismutase. Evidence of a Role for the Carbonate Radical Anion. J. Am. Chem. Soc. 2010, 132, 17174. (503) Berlett, B. S.; Stadtman, E. R. Protein Oxidation in Aging, Disease, and Oxidative Stress. J. Biol. Chem. 1997, 272, 20313. (504) Davies, M. J.; Fu, S.; Wang, H.; Dean, R. T. Stable Markers of Oxidant Damage to Proteins and Their Application in the Study of Human Disease. Free Radical Biol. Med. 1999, 27, 1151. (505) Grune, T.; Jung, T.; Merker, K.; Davies, K. J. Decreased Proteolysis Caused by Protein Aggregates, Inclusion Bodies, Plaques, Lipofuscin, Ceroid, and ’Aggresomes’ During Oxidative Stress, Aging, and Disease. Int. J. Biochem. Cell Biol. 2004, 36, 2519. (506) Stadtman, E. R.; Berlett, B. S. Reactive Oxygen-Mediated Protein Oxidation in Aging and Disease. Chem. Res. Toxicol. 1997, 10, 485. (507) Andrekopoulos, C.; Zhang, H.; Joseph, J.; Kalivendi, S.; Kalyanaraman, B. Bicarbonate Enhances Alpha-Synuclein Oligomerization and Nitration: Intermediacy of Carbonate Radical Anion and Nitrogen Dioxide Radical. Biochem. J. 2004, 378, 435. (508) Coelho, F. R.; Iqbal, A.; Linares, E.; Silva, D. F.; Lima, F. S.; Cuccovia, I. M.; Augusto, O. Oxidation of the Tryptophan 32 Residue of Human Superoxide Dismutase 1 Caused by Its BicarbonateDependent Peroxidase Activity Triggers the Non-Amyloid Aggregation of the Enzyme. J. Biol. Chem. 2014, 289, 30690. (509) Iqbal, A.; Paviani, V.; Moretti, A. I.; Laurindo, F. R.; Augusto, O. Oxidation, Inactivation and Aggregation of Protein Disulfide Isomerase Promoted by the Bicarbonate-Dependent Peroxidase Activity of Human Superoxide Dismutase. Arch. Biochem. Biophys. 2014, 557, 72. (510) Queliconi, B. B.; Marazzi, T. B.; Vaz, S. M.; Brookes, P. S.; Nehrke, K.; Augusto, O.; Kowaltowski, A. J. Bicarbonate Modulates Oxidative and Functional Damage in Ischemia-Reperfusion. Free Radical Biol. Med. 2013, 55, 46. (511) Gamon, L. F.; Wille, U. Oxidative Damage of Biomolecules by the Environmental Pollutants NO2* and NO3*. Acc. Chem. Res. 2016, 49, 2136. (512) Beckman, J. S. In Nitric Oxide Principles and Actions; Lancaster, J., Ed.; Academic Press: New York, 1996. (513) Stanbury, D. M. Reduction Potentials Involving Inorganic Free Radicals in Aqueous Solutions. Adv. Inorg. Chem. 1989, 33, 69. (514) Schwartz, S. E.; White, W. H. Kinetics of Reactive Dissolution of Nitrogen Dioxide in Aqueous Solution. Adv. Environ. Sci. Technol. 1983, 12, 1. (515) Ross, A. B.; Mallard, W. G.; Helman, W. P.; Buxton, G. V.; Huie, R. T.; Neta, P., NDRL-NIST Solution Kinetics Database, version 3; Notre Dame Radiation Laboratory and NIST Standard Reference Data: Notre Dame, IN and Gaithersburg, MD, 1998. (516) Kirsch, M.; Korth, H. G.; Sustmann, R.; de Groot, H. The Pathobiochemistry of Nitrogen Dioxide. Biol. Chem. 2002, 383, 389. 1400

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(537) Bokhoven, C.; Niessen, H. J. Amounts of Oxides of Nitrogen and Carbon Monoxide in Cigarette Smoke, with and without Inhalation. Nature 1961, 192, 458. (538) Bayram, H.; Sapsford, R. J.; Abdelaziz, M. M.; Khair, O. A. Effect of Ozone and Nitrogen Dioxide on the Release of Proinflammatory Mediators from Bronchial Epithelial Cells of Nonatopic Nonasthmatic Subjects and Atopic Asthmatic Patients in Vitro. J. Allergy Clin. Immunol. 2001, 107, 287. (539) Postlethwait, E. M.; Langford, S. D.; Bidani, A. Reactive Absorption of Nitrogen Dioxide by Pulmonary Epithelial Lining Fluid. J. Appl. Physiol. 1990, 69, 523. (540) Postlethwait, E. M.; Langford, S. D.; Jacobson, L. M.; Bidani, A. NO2 Reactive Absorption Substrates in Rat Pulmonary Surface Lining Fluids. Free Radical Biol. Med. 1995, 19, 553. (541) Sagai, M.; Ichinose, T.; Kubota, K. Studies on the Biochemical Effects of Nitrogen Dioxide. Iv. Relation between the Change of Lipid Peroxidation and the Antioxidative Protective System in Rat Lungs Upon Life Span Exposure to Low Levels of NO2. Toxicol. Appl. Pharmacol. 1984, 73, 444. (542) Persinger, R. L.; Blay, W. M.; Heintz, N. H.; Hemenway, D. R.; Janssen-Heininger, Y. M. Nitrogen Dioxide Induces Death in Lung Epithelial Cells in a Density-Dependent Manner. Am. J. Respir. Cell Mol. Biol. 2001, 24, 583. (543) Velsor, L. W.; Ballinger, C. A.; Patel, J.; Postlethwait, E. M. Influence of Epithelial Lining Fluid Lipids on NO(2)-Induced Membrane Oxidation and Nitration. Free Radical Biol. Med. 2003, 34, 720. (544) Velsor, L. W.; Postlethwait, E. M. NO2-Induced Generation of Extracellular Reactive Oxygen Is Mediated by Epithelial Lining Layer Antioxidants. Am. J. Physiol. 1997, 273, L1265. (545) Carballal, S.; Trujillo, M.; Cuevasanta, E.; Bartesaghi, S.; Möller, M. N.; Folkes, L. K.; García-Bereguiaín, M. A.; GutiérrezMerino, C.; Wardman, P.; Denicola, A.; et al. Reactivity of Hydrogen Sulfide with Peroxynitrite and Other Oxidants of Biological Interest. Free Radical Biol. Med. 2011, 50, 196. (546) Das, T. N.; Huie, R. E.; Neta, P.; Padmaja, S. Reduction Potential of the Sulfhydryl Radical: Pulse Radiolysis and Laser Flash Photolysis Studies of the Formation and Reactions of •SH and HSSH•− in Aqueous Solutions. J. Phys. Chem. A 1999, 103, 5221. (547) Prütz, W. A.; Mönig, H.; Butler, J.; Land, E. J. Reactions of Nitrogen Dioxide in Aqueous Model Systems: Oxidation of Tyrosine Units in Peptides and Proteins. Arch. Biochem. Biophys. 1985, 243, 125. (548) Patel, K. B.; Stratford, M. R.; Wardman, P.; Everett, S. A. Oxidation of Tetrahydrobiopterin by Biological Radicals and Scavenging of the Trihydrobiopterin Radical by Ascorbate. Free Radical Biol. Med. 2002, 32, 203. (549) Yermilov, V.; Rubio, J.; Becchi, M.; Friesen, M. D.; Pignatelli, B.; Ohshima, H. Formation of 8-Nitroguanine by the Reaction of Guanine with Peroxynitrite in Vitro. Carcinogenesis 1995, 16, 2045. (550) Sodum, R. S.; Fiala, E. S. Analysis of Peroxynitrite Reactions with Guanine, Xanthine, and Adenine Nucleosides by High-Pressure Liquid Chromatography with Electrochemical Detection: C8-Nitration and -Oxidation. Chem. Res. Toxicol. 2001, 14, 438. (551) Sawa, T.; Zaki, M. H.; Okamoto, T.; Akuta, T.; Tokutomi, Y.; Kim-Mitsuyama, S.; Ihara, H.; Kobayashi, A.; Yamamoto, M.; Fujii, S.; et al. Protein S-Guanylation by the Biological Signal 8-Nitroguanosine 3′,5′-Cyclic Monophosphate. Nat. Chem. Biol. 2007, 3, 727. (552) Fujii, S.; Sawa, T.; Ihara, H.; Tong, K. I.; Ida, T.; Okamoto, T.; Ahtesham, A. K.; Ishima, Y.; Motohashi, H.; Yamamoto, M.; et al. The Critical Role of Nitric Oxide Signaling, Via Protein S-Guanylation and Nitrated Cyclic Gmp, in the Antioxidant Adaptive Response. J. Biol. Chem. 2010, 285, 23970. (553) Ingold, C. K.; Millen, D. J.; Poole, H. G. Vibrational Spectra of Ionic Forms of Oxides and Oxy-Acids of Nitrogen. Part I. RamanSpectral Evidence of the Ionisation of Nitric Acid by Perchloric, Sulphuric, and Selenic Acids. Spectroscopic Identification of the Nitronium Ion, NO2+. J. Chem. Soc. 1950, 2576.

(554) Hughes, E. D.; Ingold, C. K.; Reed, R. I. Kinetic and Mechanism of Aromatic Nitration. Part II. Nitration by the Nitronium Ion, NO2+, Derived from Nitric Acid. J. Chem. Soc. 1950, 2400. (555) Olah, G. A.; Narang, S. C.; Olah, J. A.; Lammertsma, K. Recent Aspects of Nitration: New Preparative Methods and Mechanistic Studies (a Review). Proc. Natl. Acad. Sci. U. S. A. 1982, 79, 4487. (556) Halfpenny, E.; Robinson, P. L. Pernitrous Acid. The Reaction between Hydrogen Peroxide and Nitrous Acid, and the Properties of an Intermediate Product. J. Chem. Soc. 1952, 928. (557) Wormall, A. The Immunological Specificity of Chemically Altered Proteins: Halogenated and Nitrated Proteins. J. Exp. Med. 1930, 51, 295. (558) Ehrenberg, L.; Fischer, I.; Löfgren, N. Inhibitory Effect of Tetranitromethane on the Diphteria Toxin-Antitoxin Precipitin Reaction. Nature 1946, 157, 730. (559) Astrup, T.; Sjögren, B. Inactivation of Thrombin by Means of Tetranitromethane. Acta Chem. Scand. 1947, 1, 744. (560) Riordan, J. F.; Sokolovsky, M.; Vallee, B. L. Tetranitromethane. A Reagent for the Nitration of Tyrosine and Tyrosyl Residues of Proteins. J. Am. Chem. Soc. 1966, 88, 4104. (561) Sokolovsky, M.; Riordan, J. F.; Vallee, B. L. Tetranitromethane. A Reagent for the Nitration of Tyrosyl Residues in Proteins. Biochemistry 1966, 5, 3582. (562) Riordan, J. F.; Sokolovsky, M.; Vallee, B. L. The Functional Tyrosyl Residues of Carboxypeptidase A. Nitration with Tetranitromethane. Biochemistry 1967, 6, 3609. (563) Zhu, L.; Gunn, C.; Beckman, J. S. Bactericidal Activity of Peroxynitrite. Arch. Biochem. Biophys. 1992, 298, 452. (564) Smith, C. D.; Carson, M.; van der Woerd, M.; Chen, J.; Ischiropoulos, H.; Beckman, J. S. Crystal Structure of PeroxynitriteModified Bovine Cu,Zn Superoxide Dismutase. Arch. Biochem. Biophys. 1992, 299, 350. (565) Eiserich, J. P.; Cross, C. E.; Jones, A. D.; Halliwell, B.; van der Vliet, A. Formation of Nitrating and Chlorinating Species by Reaction of Nitrite with Hypochlorous Acid. J. Biol. Chem. 1996, 271, 19199. (566) Eiserich, J. P.; Hristova, M.; Cross, C. E.; Jones, A. D.; Freeman, B. A.; Halliwell, B.; van der Vliet, A. Formation of Nitric Oxide-Derived Inflammatory Oxidants by Myeloperoxidase in Neutrophils. Nature 1998, 391, 393. (567) Thomas, D. D.; Espey, M. G.; Vitek, M. P.; Miranda, K. M.; Wink, D. a. Protein Nitration Is Mediated by Heme and Free Metals through Fenton-Type Chemistry: An Alternative to the NO/O2Reaction. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12691. (568) Pietraforte, D.; Minetti, M. One-Electron Oxidation Pathway of Peroxynitrite Decomposition in Human Blood Plasma: Evidence for the Formation of Protein Tryptophan-Centred Radicals. Biochem. J. 1997, 321, 743. (569) Herold, S.; Shivashankar, K.; Mehl, M. Myoglobin Scavenges Peroxynitrite without Being Significantly Nitrated. Biochemistry 2002, 41, 13460. (570) Kawasaki, H.; Ikeda, K.; Shigenaga, A.; Baba, T.; Takamori, K.; Ogawa, H.; Yamakura, F. Mass Spectrometric Identification of Tryptophan Nitration Sites on Proteins in Peroxynitrite-Treated Lysates from Pc12 Cells. Free Radical Biol. Med. 2011, 50, 419. (571) Solar, S.; Solar, W.; Getoff, N. Reactivity of Oh with Tyrosine in Aqueous Solution Studied by Pulse Radiolysis. J. Phys. Chem. 1984, 88, 2091. (572) Gow, A.; Duran, D.; Thom, S. R.; Ischiropoulos, H. Carbon Dioxide Enhancement of Peroxynitrite-Mediated Protein Tyrosine Nitration. Arch. Biochem. Biophys. 1996, 333, 42. (573) Crow, J. P. Manganese and Iron Porphyrins Catalyze Peroxynitrite Decomposition and Simultaneously Increase Nitration and Oxidant Yield: Implications for Their Use as Peroxynitrite Scavengers in Vivo. Arch. Biochem. Biophys. 1999, 371, 41. (574) Ramezanian, M. S.; Padmaja, S.; Koppenol, W. H. Nitration and Hydroxylation of Phenolic Compounds by Peroxynitrite. Chem. Res. Toxicol. 1996, 9, 232. 1401

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(575) Folkes, L. K.; Bartesaghi, S.; Trujillo, M.; Radi, R.; Wardman, P. Kinetics of Oxidation of Tyrosine by a Model Alkoxyl Radical. Free Radical Res. 2012, 46, 1150. (576) Heinecke, J. W.; Li, W.; Daehnke, H. L.; Goldstein, J. A. Dityrosine, a Specific Marker of Oxidation, Is Synthesized by the Myeloperoxidase-Hydrogen Peroxide System of Human Neutrophils and Macrophages. J. Biol. Chem. 1993, 268, 4069. (577) Pfeiffer, S.; Schmidt, K.; Mayer, B. Dityrosine Formation Outcompetes Tyrosine Nitration at Low Steady-State Concentrations of Peroxynitrite. J. Biol. Chem. 2000, 275, 6346. (578) van der Vliet, A.; O’Neill, C. A.; Halliwell, B.; Cross, C. E.; Kaur, H. Aromatic Hydroxylation and Nitration of Phenylalanine and Tyrosine by Peroxynitrite. Evidence for Hydroxyl Radical Production from Peroxynitrite. FEBS Lett. 1994, 339, 89. (579) Folkes, L. K.; Trujillo, M.; Bartesaghi, S.; Radi, R.; Wardman, P. Kinetics of Reduction of Tyrosine Phenoxyl Radicals by Glutathione. Arch. Biochem. Biophys. 2011, 506, 242. (580) Hunter, E. P.; Desrosiers, M. F.; Simic, M. G. The Effect of Oxygen, Antioxidants, and Superoxide Radical on Tyrosine Phenoxyl Radical Dimerization. Free Radical Biol. Med. 1989, 6, 581. (581) Eiserich, J. P.; Butler, J.; van der Vliet, A.; Cross, C. E.; Halliwell, B. Nitric Oxide Rapidly Scavenges Tyrosine and Tryptophan Radicals. Biochem. J. 1995, 310, 745. (582) Jin, F.; Leitich, J.; von Sonntag, C. The Superoxide Radical Reacts with Tyrosine-Derived Phenoxyl Radicals by Addition Rather Than by Electron Transfer. J. Chem. Soc., Perkin Trans. 2 1993, 1583. (583) Radi, R. Protein Tyrosine Nitration: Biochemical Mechanisms and Structural Basis of Functional Effects. Acc. Chem. Res. 2013, 46, 550. (584) Sampson, J. B.; Rosen, H.; Beckman, J. S. PeroxynitriteDependent Tyrosine Nitration Catalyzed by Superoxide Dismutase, Myeloperoxidase, and Horseradish Peroxidase. Methods Enzymol. 1996, 269, 210. (585) Campolo, N.; Bartesaghi, S.; Radi, R. Metal-Catalyzed Protein Tyrosine Nitration in Biological Systems. Redox Rep. 2014, 19, 221. (586) Frost, M. T.; Halliwell, B.; Moore, K. P. Analysis of Free and Protein-Bound Nitrotyrosine in Human Plasma by a Gas Chromatography/Mass Spectrometry Method That Avoids Nitration Artifacts. Biochem. J. 2000, 345, 453. (587) Schwedhelm, E.; Tsikas, D.; Gutzki, F. M.; Frölich, J. C. Gas Chromatographic-Tandem Mass Spectrometric Quantification of Free 3-Nitrotyrosine in Human Plasma at the Basal State. Anal. Biochem. 1999, 276, 195. (588) Baldus, S.; Eiserich, J. P.; Brennan, M.-L.; Jackson, R. M.; Alexander, C. B.; Freeman, B. A. Spatial Mapping of Pulmonary and Vascular Nitrotyrosine Reveals the Pivotal Role of Myeloperoxidase as a Catalyst for Tyrosine Nitration in Inflammatory Diseases. Free Radical Biol. Med. 2002, 33, 1010. (589) Castro, L.; Eiserich, J. P.; Sweeney, S.; Radi, R.; Freeman, B. A. Cytochrome C: A Catalyst and Target of Nitrite-Hydrogen PeroxideDependent Protein Nitration. Arch. Biochem. Biophys. 2004, 421, 99. (590) Sainz, M.; Calvo-Begueria, L.; Perez-Rontome, C.; Wienkoop, S.; Abian, J.; Staudinger, C.; Bartesaghi, S.; Radi, R.; Becana, M. Leghemoglobin Is Nitrated in Functional Legume Nodules in a Tyrosine Residue within the Heme Cavity by a Nitrite/PeroxideDependent Mechanism. Plant J. 2015, 81, 723. (591) Monzani, E.; Nicolis, S.; Roncone, R.; Barbieri, M.; Granata, A.; Casella, L. Protein Self-Modification by Heme-Generated Reactive Species. IUBMB Life 2008, 60, 41. (592) Nicolis, S.; Monzani, E.; Ciaccio, C.; Ascenzi, P.; Moens, L.; Casella, L. Reactivity and Endogenous Modification by Nitrite and Hydrogen Peroxide: Does Human Neuroglobin Act Only as a Scavenger? Biochem. J. 2007, 407, 89. (593) Herold, S. Nitrotyrosine, Dityrosine, and Nitrotryptophan Formation from Metmyoglobin, Hydrogen Peroxide, and Nitrite. Free Radical Biol. Med. 2004, 36, 565. (594) Gunther, M. R.; Hsi, L. C.; Curtis, J. F.; Gierse, J. K.; Marnett, L. J.; Eling, T. E.; Mason, R. P. Nitric Oxide Trapping of the Tyrosyl

Radical of Prostaglandin H Synthase-2 Leads to Tyrosine Iminoxyl Radical and Nitrotyrosine Formation. J. Biol. Chem. 1997, 272, 17086. (595) Sturgeon, B. E.; Glover, R. E.; Chen, Y. R.; Burka, L. T.; Mason, R. P. Tyrosine Iminoxyl Radical Formation from Tyrosyl Radical/Nitric Oxide and Nitrosotyrosine. J. Biol. Chem. 2001, 276, 45516. (596) Sanakis, Y.; Goussias, C.; Mason, R. P.; Petrouleas, V. NO Interacts with the Tyrosine Radical Y(D). Of Photosystem II to Form an Iminoxyl Radical. Biochemistry 1997, 36, 1411. (597) Oldreive, C.; Zhao, K.; Paganga, G.; Halliwell, B.; Rice-Evans, C. Inhibition of Nitrous Acid-Dependent Tyrosine Nitration and DNA Base Deamination by Flavonoids and Other Phenolic Compounds. Chem. Res. Toxicol. 1998, 11, 1574. (598) Pannala, A. S.; Mani, A. R.; Rice-Evans, C. A.; Moore, K. P. pH-Dependent Nitration of Para -Hydroxyphenylacetic Acid in the Stomach. Free Radical Biol. Med. 2006, 41, 896. (599) Kosaka, H.; Imaizumi, K.; Imai, K.; Tyuma, I. Stoichiometry of the Reaction of Oxyhemoglobin with Nitrite. Biochim. Biophys. Acta, Protein Struct. 1979, 581, 184. (600) Kosaka, H.; Tyuma, I. Mechanism of Autocatalytic Oxidation of Oxyhemoglobin by Nitrite. Environ. Health Perspect. 1987, 73, 147. (601) Lissi, E. Autocatalytic Oxidation of Hemoglobin by Nitrite: A Possible Mechanism. Free Radical Biol. Med. 1998, 24, 1535. (602) Keszler, A.; Piknova, B.; Schechter, A. N.; Hogg, N. The Reaction between Nitrite and Oxyhemoglobin: A Mechanistic Study. J. Biol. Chem. 2008, 283, 9615. (603) Tsikas, D.; Boger, R. H.; Bode-Boger, S. M.; Gutzki, F. M.; Frolich, J. C. Quantification of Nitrite and Nitrate in Human Urine and Plasma as Pentafluorobenzyl Derivatives by Gas ChromatographyMass Spectrometry Using Their 15n-Labelled Analogs. J. Chromatogr., Biomed. Appl. 1994, 661, 185. (604) Moshage, H.; Kok, B.; Huizenga, J. R.; Jansen, P. L. Nitrite and Nitrate Determinations in Plasma: A Critical Evaluation. Clin. Chem. 1995, 41, 892. (605) Dejam, A.; Hunter, C. J.; Pelletier, M. M.; Hsu, L. L.; Machado, R. F.; Shiva, S.; Power, G. G.; Kelm, M.; Gladwin, M. T.; Schechter, A. N. Erythrocytes Are the Major Intravascular Storage Sites of Nitrite in Human Blood. Blood 2005, 106, 734. (606) Schwarz, A.; Modun, D.; Heusser, K.; Tank, J.; Gutzki, F. M.; Mitschke, A.; Jordan, J.; Tsikas, D. Stable-Isotope Dilution Gc-Ms Approach for Nitrite Quantification in Human Whole Blood, Erythrocytes, and Plasma Using Pentafluorobenzyl Bromide Derivatization: Nitrite Distribution in Human Blood. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2011, 879, 1485. (607) Eiserich, J. P.; Estévez, A. G.; Bamberg, T. V.; Ye, Y. Z.; Chumley, P. H.; Beckman, J. S.; Freeman, B. A. Microtubule Dysfunction by Posttranslational Nitrotyrosination of Alpha-Tubulin: A Nitric Oxide-Dependent Mechanism of Cellular Injury. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 6365. (608) Blanchard-Fillion, B.; Prou, D.; Polydoro, M.; Spielberg, D.; Tsika, E.; Wang, Z.; Hazen, S. L.; Koval, M.; Przedborski, S.; Ischiropoulos, H. Metabolism of 3-Nitrotyrosine Induces Apoptotic Death in Dopaminergic Cells. J. Neurosci. 2006, 26, 6124. (609) Kalisz, H. M.; Erck, C.; Plessmann, U.; Wehland, J. Incorporation of Nitrotyrosine into A-Tubulin by Recombinant Mammalian Tubulin-Tyrosine Ligase. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol. 2000, 1481, 131. (610) Neumann, H.; Hazen, J. L.; Weinstein, J.; Mehl, R. A.; Chin, J. W. Genetically Encoding Protein Oxidative Damage. J. Am. Chem. Soc. 2008, 130, 4028. (611) Yokoyama, K.; Uhlin, U.; Stubbe, J. Site-Specific Incorporation of 3-Nitrotyrosine as a Probe of Pka Perturbation of Redox-Active Tyrosines in Ribonucleotide Reductase. J. Am. Chem. Soc. 2010, 132, 8385. (612) DiDonato, J. A.; Aulak, K.; Huang, Y.; Wagner, M.; Gerstenecker, G.; Topbas, C.; Gogonea, V.; DiDonato, A. J.; Tang, W. H. W.; Mehl, R. A.; et al. Site-Specific Nitration of Apolipoprotein a-I at Tyrosine 166 Is Both Abundant within Human Atherosclerotic Plaque and Dysfunctional. J. Biol. Chem. 2014, 289, 10276. 1402

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(631) Zheng, L.; Nukuna, B.; Brennan, M.-L.; Sun, M.; Goormastic, M.; Settle, M.; Schmitt, D.; Fu, X.; Thomson, L.; Fox, P. L.; et al. Apolipoprotein a-I Is a Selective Target for MyeloperoxidaseCatalyzed Oxidation and Functional Impairment in Subjects with Cardiovascular Disease. J. Clin. Invest. 2004, 114, 529. (632) Zheng, L.; Settle, M.; Brubaker, G.; Schmitt, D.; Hazen, S. L.; Smith, J. D.; Kinter, M. Localization of Nitration and Chlorination Sites on Apolipoprotein a-I Catalyzed by Myeloperoxidase in Human Atheroma and Associated Oxidative Impairment in Abca1-Dependent Cholesterol Efflux from Macrophages. J. Biol. Chem. 2005, 280, 38. (633) Yamakura, F.; Taka, H.; Fujimura, T.; Murayama, K. Inactivation of Human Manganese-Superoxide Dismutase by Peroxynitrite Is Caused by Exclusive Nitration of Tyrosine 34 to 3Nitrotyrosine. J. Biol. Chem. 1998, 273, 14085. (634) Moreno, D. M.; Martí, M. A.; De Biase, P. M.; Estrin, D. A.; Demicheli, V.; Radi, R.; Boechi, L. Exploring the Molecular Basis of Human Manganese Superoxide Dismutase Inactivation Mediated by Tyrosine 34 Nitration. Arch. Biochem. Biophys. 2011, 507, 304. (635) Deeb, R. S.; Hao, G.; Gross, S. S.; Lainé, M.; Qiu, J. H.; Resnick, B.; Barbar, E. J.; Hajjar, D. P.; Upmacis, R. K. Heme Catalyzes Tyrosine 385 Nitration and Inactivation of Prostaglandin H2 Synthase-1 by Peroxynitrite. J. Lipid Res. 2006, 47, 898. (636) Schmidt, P.; Youhnovski, N.; Daiber, A.; Balan, A.; Arsic, M.; Bachschmid, M.; Przybylski, M.; Ullrich, V. Specific Nitration at Tyrosine 430 Revealed by High Resolution Mass Spectrometry as Basis for Redox Regulation of Bovine Prostacyclin Synthase. J. Biol. Chem. 2003, 278, 12813. (637) Goodwin, D. C.; Gunther, M. R.; Hsi, L. C.; Crews, B. C.; Eling, T. E.; Mason, R. P.; Marnett, L. J. Nitric Oxide Trapping of Tyrosyl Radicals Generated During Prostaglandin Endoperoxide Synthase Turnover: Detection of the Radical Derivative of Tyrosine 385. J. Biol. Chem. 1998, 273, 8903. (638) Lepoivre, M.; Houée-Levin, C.; Coeytaux, K.; Decottignies, P.; Auger, G.; Lemaire, G. Nitration of the Tyrosyl Radical in Ribonucleotide Reductase by Nitrogen Dioxide: A Gamma Radiolysis Study. Free Radical Biol. Med. 2005, 38, 1511. (639) Guittet, O.; Decottignies, P.; Serani, L.; Henry, Y.; Le Maréchal, P.; Laprévote, O.; Lepoivre, M. Peroxynitrite-Mediated Nitration of the Stable Free Radical Tyrosine Residue of the Ribonucleotide Reductase Small Subunit. Biochemistry 2000, 39, 4640. (640) van der Vliet, A.; Eiserich, J. P.; O’Neill, C. A.; Halliwell, B.; Cross, C. E. Tyrosine Modification by Reactive Nitrogen Species - a Closer Look. Arch. Biochem. Biophys. 1995, 319, 341. (641) Linares, E.; Giorgio, S.; Mortara, R. A.; Santos, C. X. C.; Yamada, A. T.; Augusto, O. Role of Peroxynitrite in Macrophage Microbicidal Mechanisms in Vivo Revealed by Protein Nitration and Hydroxylation. Free Radical Biol. Med. 2001, 30, 1234. (642) Hensley, K.; Maidt, M. L.; Yu, Z.; Sang, H.; Markesbery, W. R.; Floyd, R. A. Electrochemical Analysis of Protein Nitrotyrosine and Dityrosine in the Alzheimer Brain Indicates Region-Specific Accumulation. J. Neurosci. 1998, 18, 8126. (643) Gross, A. J.; Sizer, I. W. The Oxidation of Tyramine, Tyrosine, and Related Compounds by Peroxidase. J. Biol. Chem. 1959, 234, 1611. (644) Andersen, S. O. The Cross-Links in Resilin Identified as Dityrosine and Trityrosine. Biochim. Biophys. Acta, Gen. Subj. 1964, 93, 213. (645) LaBella, F.; Keeley, F.; Vivian, S.; Thornhill, D. Evidence for Dityrosine in Elastin. Biochem. Biophys. Res. Commun. 1967, 26, 748. (646) Raven, D. J.; Earland, C.; Little, M. Occurrence of Dityrosine in Tussah Silk Fibroin and Keratin. Biochim. Biophys. Acta, Protein Struct. 1971, 251, 96. (647) LaBella, F.; Waykole, P.; Queen, G. Formation of Insoluble Gels and Dityrosine by the Action of Peroxidase on Soluble Collagens. Biochem. Biophys. Res. Commun. 1968, 30, 333. (648) Aeschbach, R.; Amadoò, R.; Neukom, H. Formation of Dityrosine Cross-Links in Proteins by Oxidation of Tyrosine Residues. Biochim. Biophys. Acta, Protein Struct. 1976, 439, 292.

(613) Franco, M. C.; Ricart, K. C.; Gonzalez, A. S.; Dennys, C. N.; Nelson, P. A.; Janes, M. S.; Mehl, R. A.; Landar, A.; Estévez, A. G. Nitration of Hsp90 on Tyrosine 33 Regulates Mitochondrial Metabolism. J. Biol. Chem. 2015, 290, 19055. (614) Tomita, H.; Katsuyama, Y.; Minami, H.; Ohnishi, Y. Identification and Characterization of a Bacterial Cytochrome P450 Monooxygenase Catalyzing the 3-Nitration of Tyrosine in Rufomycin Biosynthesis. J. Biol. Chem. 2017, 292, 15859. (615) Ara, J.; Przedborski, S.; Naini, A. B.; Jackson-Lewis, V.; Trifiletti, R. R.; Horwitz, J.; Ischiropoulos, H. Inactivation of Tyrosine Hydroxylase by Nitration Following Exposure to Peroxynitrite and 1Methyl-4-Phenyl-1, 2, 3, 6- Tetrahydropyridine (Mptp). Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 7659. (616) Crow, J. P.; Ye, Y. Z.; Strong, M.; Kirk, M.; Barnes, S.; Beckman, J. Superoxide Dismutase Catalyzes Nitration of Tyrosines by Peroxynitrite in the Rod and Head Domains of Neurofilament-L. J. Neurochem. 1997, 69, 1945. (617) Ischiropoulos, H. Biological Tyrosine Nitration: A Pathophysiological Function of Nitric Oxide and Reactive Oxygen Species. Arch. Biochem. Biophys. 1998, 356, 1. (618) Franco, M. C.; Ye, Y.; Refakis, C. A.; Feldman, J. L.; Stokes, A. L.; Basso, M.; Melero Fernández de Mera, R. M.; Sparrow, N. A.; Calingasan, N. Y.; Kiaei, M.; et al. Nitration of Hsp90 Induces Cell Death. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, E1102. (619) Peinado, M. Á .; Hernández, R.; Peragón, J.; Ovelleiro, D.; Pedrosa, J. Á .; Blanco, S. Proteomic Characterization of Nitrated Cell Targets after Hypobaric Hypoxia and Reoxygenation in Rat Brain. J. Proteomics 2014, 109, 309. (620) Souza, J. M.; Daikhin, E.; Yudkoff, M.; Raman, C. S.; Ischiropoulos, H. Factors Determining the Selectivity of Protein Tyrosine Nitration. Arch. Biochem. Biophys. 1999, 371, 169. (621) Ischiropoulos, H. Biological Selectivity and Functional Aspects of Protein Tyrosine Nitration. Biochem. Biophys. Res. Commun. 2003, 305, 776. (622) Castro, L.; Demicheli, V.; Tórtora, V.; Radi, R. Mitochondrial Protein Tyrosine Nitration. Free Radical Res. 2011, 45, 37. (623) Liu, Z.; Cao, J.; Ma, Q.; Gao, X.; Xue, Y.; Ren, J. GPS-YNO2: Computational Prediction of Tyrosine Nitration Sites in Proteins. Mol. BioSyst. 2011, 7, 1197. (624) Ng, J. Y.; Boelen, L.; Wong, J. W. H. Bioinformatics Analysis Reveals Biophysical and Evolutionary Insights into the 3-Nitrotyrosine Post-Translational Modification in the Human Proteome. Open Biol. 2013, 3, 120148. (625) Bayden, A. S.; Yakovlev, V. A.; Graves, P. R.; Mikkelsen, R. B.; Kellogg, G. E. Factors Influencing Protein Tyrosine Nitration– Structure-Based Predictive Models. Free Radical Biol. Med. 2011, 50, 749. (626) Cheng, S.; Lian, B.; Liang, J.; Shi, T.; Xie, L.; Zhao, Y.-l. Site Selectivity for Protein Tyrosine Nitration: Insights from Features of Structure and Topological Networks. Mol. BioSyst. 2013, 9, 2860. (627) Zhang, H.; Xu, Y.; Joseph, J.; Kalyanaraman, B. Intramolecular Electron Transfer between Tyrosyl Radical and Cysteine Residue Inhibits Tyrosine Nitration and Induces Thiyl Radical Formation in Model Peptides Treated with Myeloperoxidase, H2O2, and NO2-: Epr Spin Trapping Studies. J. Biol. Chem. 2005, 280, 40684. (628) Petruk, A. a.; Bartesaghi, S.; Trujillo, M.; Estrin, D. A.; Murgida, D.; Kalyanaraman, B.; Marti, M. A.; Radi, R. Molecular Basis of Intramolecular Electron Transfer in Proteins During RadicalMediated Oxidations: Computer Simulation Studies in Model Tyrosine-Cysteine Peptides in Solution. Arch. Biochem. Biophys. 2012, 525, 82. (629) Levine, R. L.; Mosoni, L.; Berlett, B. S.; Stadtman, E. R. Methionine Residues as Endogenous Antioxidants in Proteins. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 15036. (630) Mozziconacci, O.; Mirkowski, J.; Rusconi, F.; Kciuk, G.; Wisniowski, P. B.; Bobrowski, K.; Houée-Levin, C. Methionine Residue Acts as a Prooxidant in the •OH-Induced Oxidation of Enkephalins. J. Phys. Chem. B 2012, 116, 12460. 1403

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

(649) Sawa, T.; Akaike, T.; Maeda, H. Tyrosine Nitration by Peroxynitrite Formed from Nitric Oxide and Superoxide Generated by Xanthine Oxidase. J. Biol. Chem. 2000, 275, 32467. (650) Zhang, H.; Joseph, J.; Feix, J.; Hogg, N.; Kalyanaraman, B. Nitration and Oxidation of a Hydrophobic Tyrosine Probe by Peroxynitrite in Membranes: Comparison with Nitration and Oxidation of Tyrosine by Peroxynitrite in Aqueous Solution. Biochemistry 2001, 40, 7675. (651) Bartesaghi, S.; Valez, V.; Trujillo, M.; Peluffo, G.; Romero, N.; Zhang, H.; Kalyanaraman, B.; Radi, R. Mechanistic Studies of Peroxynitrite-Mediated Tyrosine Nitration in Membranes Using the Hydrophobic Probe N-T-Boc-L-Tyrosine Tert-Butyl Ester. Biochemistry 2006, 45, 6813. (652) Ischiropoulos, H.; Al-Mehdi, A. B. Peroxynitrite-Mediated Oxidative Protein Modifications. FEBS Lett. 1995, 364, 279. (653) MacMillan-Crow, L. A.; Crow, J. P.; Thompson, J. A. Peroxynitrite-Mediated Inactivation of Manganese Superoxide Dismutase Involves Nitration and Oxidation of Critical Tyrosine Residues. Biochemistry 1998, 37, 1613. (654) Nowak, P.; Wachowicz, B. Peroxynitrite-Mediated Modification of Fibrinogen Affects Platelet Aggregation and Adhesion. Platelets 2002, 13, 293. (655) Bagnasco, P.; MacMillan-Crow, L. A.; Greendorfer, J. S.; Young, C. J.; Andrews, L.; Thompson, J. A. Peroxynitrite Modulates Acidic Fibroblast Growth Factor (Fgf-1) Activity. Arch. Biochem. Biophys. 2003, 419, 178. (656) van der Vliet, A.; Hristova, M.; Cross, C. E.; Eiserich, J. P.; Goldkorn, T. Peroxynitrite Induces Covalent Dimerization of Epidermal Growth Factor Receptors in A431 Epidermoid Carcinoma Cells. J. Biol. Chem. 1998, 273, 31860. (657) Giasson, B. I.; Duda, J. E.; Murray, I. V.; Chen, Q.; Souza, J. M.; Hurtig, H. I.; Ischiropoulos, H.; Trojanowski, J. Q.; Lee, V. M. Oxidative Damage Linked to Neurodegeneration by Selective AlphaSynuclein Nitration in Synucleinopathy Lesions. Science 2000, 290, 985. (658) Souza, J. M.; Giasson, B. I.; Chen, Q.; Lee, V. M.; Ischiropoulos, H. Dityrosine Cross-Linking Promotes Formation of Stable Alpha -Synuclein Polymers. Implication of Nitrative and Oxidative Stress in the Pathogenesis of Neurodegenerative Synucleinopathies. J. Biol. Chem. 2000, 275, 18344. (659) Reynolds, M. R.; Berry, R. W.; Binder, L. I. Site-Specific Nitration and Oxidative Dityrosine Bridging of the T Protein by Peroxynitrite: Implications for Alzheimer’ S Disease. Biochemistry 2005, 44, 1690. (660) Jacob, J. S.; Cistola, D. P.; Hsu, F. F.; Muzaffar, S.; Mueller, D. M.; Hazen, S. L.; Heinecke, J. W. Human Phagocytes Employ the Myeloperoxidase-Hydrogen Peroxide System to Synthesize Dityrosine, Trityrosine, Pulcherosine, and Isodityrosine by a Tyrosyl RadicalDependent Pathway. J. Biol. Chem. 1996, 271, 19950. (661) Leeuwenburgh, C.; Hansen, P. A.; Holloszy, J. O.; Heinecke, J. W. Oxidized Amino Acids in the Urine of Aging Rats: Potential Markers for Assessing Oxidative Stress in Vivo. Am. J. Physiol. 1999, 276, R128. (662) Bhattacharjee, S.; Pennathur, S.; Byun, J.; Crowley, J.; Mueller, D.; Gischler, J.; Hotchkiss, R. S.; Heinecke, J. W. Nadph Oxidase of Neutrophils Elevates O,O′-Dityrosine Cross-Links in Proteins and Urine During Inflammation. Arch. Biochem. Biophys. 2001, 395, 69. (663) Kato, Y.; Wu, X.; Naito, M.; Nomura, H.; Kitamoto, N.; Osawa, T. Immunochemical Detection of Protein Dityrosine in Atherosclerotic Lesion of Apo-E-Deficient Mice Using a Novel Monoclonal Antibody. Biochem. Biophys. Res. Commun. 2000, 275, 11. (664) Orhan, H.; Vermeulen, N. P. E.; Tump, C.; Zappey, H.; Meerman, J. H. N. Simultaneous Determination of Tyrosine, Phenylalanine and Deoxyguanosine Oxidation Products by Liquid Chromatography-Tandem Mass Spectrometry as Non-Invasive Biomarkers for Oxidative Damage. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2004, 799, 245. (665) Colombo, G.; Reggiani, F.; Cucchiari, D.; Portinaro, N. M.; Giustarini, D.; Rossi, R.; Garavaglia, M. L.; Saino, N.; Milzani, A.;

Badalamenti, S.; et al. Plasma Protein-Bound Di-Tyrosines as Biomarkers of Oxidative Stress in End Stage Renal Disease Patients on Maintenance Haemodialysis. BBA Clinical 2017, 7, 55. (666) DiMarco, T.; Giulivi, C. Current Analytical Methods for the Detection of Dityrosine, a Biomarker of Oxidative Stress, in Biological Samples. Mass Spectrom. Rev. 2007, 26, 108. (667) Chen, Y.-R.; Chen, C.-L.; Chen, W.; Zweier, J. L.; Augusto, O.; Radi, R.; Mason, R. P. Formation of Protein Tyrosine OrthoSemiquinone Radical and Nitrotyrosine from Cytochrome C-Derived Tyrosyl Radical. J. Biol. Chem. 2004, 279, 18054. (668) Riordan, J. F.; Sokolovsky, M.; Vallee, B. L. Environmentally Sensitive Tyrosyl Residues. Nitration with Tetranitromethane. Biochemistry 1967, 6, 358. (669) Sokolovsky, M.; Riordan, J. F.; Vallee, B. L. Conversion of 3Nitrotyrosine to 3-Aminotyrosine in Peptides and Proteins. Biochem. Biophys. Res. Commun. 1967, 27, 20. (670) Abello, N.; Kerstjens, H. A. M.; Postma, D. S.; Bischoff, R. Protein Tyrosine Nitration: Selectivity, Physicochemical and Biological Consequences, Denitration, and Proteomics Methods for the Identification of Tyrosine-Nitrated Proteins. J. Proteome Res. 2009, 8, 3222. (671) Souza, J. M.; Peluffo, G.; Radi, R. Protein Tyrosine Nitration– Functional Alteration or Just a Biomarker? Free Radical Biol. Med. 2008, 45, 357. (672) Chopineaux-Courtois, V.; Reymond, F.; Bouchard, G.; Carrupt, P.-A.; Testa, B.; Girault, H. H. Effects of Charge and Intramolecular Structure on the Lipophilicity of Nitrophenols. J. Am. Chem. Soc. 1999, 121, 1743. (673) Abraham, M. H.; Du, C. M.; Platts, J. A. Lipophilicity of the Nitrophenols. J. Org. Chem. 2000, 65, 7114. (674) De Filippis, V.; Frasson, R.; Fontana, A. 3-Nitrotyrosine as a Spectroscopic Probe for Investigating Protein − Protein Interactions. Protein Sci. 2006, 15, 976. (675) Koide, S.; Sidhu, S. S. The Importance of Being Tyrosine: Lessons in Molecular Recognition from Minimalist Synthetic Binding Proteins. ACS Chem. Biol. 2009, 4, 325. (676) Souza, J. M.; Castro, L.; Cassina, A. M.; Batthyany, C.; Radi, R. Nitrocytochrome C: Synthesis, Purification, and Functional Studies. Methods Enzymol. 2008, 441, 197. (677) Savvides, S. N.; Scheiwein, M.; Bohme, C. C.; Arteel, G. E.; Karplus, P. A.; Becker, K.; Schirmer, R. H. Crystal Structure of the Antioxidant Enzyme Glutathione Reductase Inactivated by Peroxynitrite. J. Biol. Chem. 2002, 277, 2779. (678) Mostad, A.; Natarajan, S. Crystal and Molecular Structure of 3Nitro-4-Hydroxy-Phenylalanine Nitrate. Z. Kristallogr. 1990, 193, 127. (679) Yee, C. S.; Seyedsayamdost, M. R.; Chang, M. C. Y.; Nocera, D. G.; Stubbe, J. Generation of the R2 Subunit of Ribonucleotide Reductase by Intein Chemistry: Insertion of 3-Nitrotyrosine at Residue 356 as a Probe of the Radical Initiation Process. Biochemistry 2003, 42, 14541. (680) Gole, M. D.; Souza, J. M.; Choi, I.; Hertkorn, C.; Malcolm, S.; Foust, R. F.; Finkel, B.; Lanken, P. N.; Ischiropoulos, H. Plasma Proteins Modified by Tyrosine Nitration in Acute Respiratory Distress Syndrome. Am. J. Phys. Lung. Cell Mol. Phys. 2000, 278, L961. (681) Tien, M.; Berlett, B. S.; Levine, R. L.; Chock, P. B.; Stadtman, E. R. Peroxynitrite-Mediated Modification of Proteins at Physiological Carbon Dioxide Concentration: pH Dependence of Carbonyl Formation, Tyrosine Nitration, and Methionine Oxidation. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 7809. (682) MacMillan-Crow, L. A.; Crow, J. P.; Kerby, J. D.; Beckman, J. S.; Thompson, J. A. Nitration and Inactivation of Manganese Superoxide Dismutase in Chronic Rejection of Human Renal Allografts. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 11853. (683) Quint, P.; Reutzel, R.; Mikulski, R.; McKenna, R.; Silverman, D. N. Crystal Structure of Nitrated Human Manganese Superoxide Dismutase: Mechanism of Inactivation. Free Radical Biol. Med. 2006, 40, 453. (684) MacMillan-Crow, L. A.; Cruthirds, D. L.; Ahki, K. M.; Sanders, P. W.; Thompson, J. A. Mitochondrial Tyrosine Nitration Precedes 1404

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Chronic Allograft Nephropathy. Free Radical Biol. Med. 2001, 31, 1603. (685) Guo, W.; Adachi, T.; Matsui, R.; Xu, S.; Jiang, B.; Zou, M.-H.; Kirber, M.; Lieberthal, W.; Cohen, R. A. Quantitative Assessment of Tyrosine Nitration of Manganese Superoxide Dismutase in Angiotensin II-Infused Rat Kidney. Am. J. Phys. Heart Circ. Phys. 2003, 285, H1396. (686) Gray, K. D.; MacMillan-Crow, L.-A.; Simovic, M. O.; Stain, S. C.; May, A. K. Pulmonary MnSOD Is Nitrated Following Hepatic Ischemia-Reperfusion. Surg. Infect. 2004, 5, 166. (687) Batthyány, C.; Souza, J. M.; Durán, R.; Cassina, A.; Cerveñansky, C.; Radi, R. Time Course and Site (S) of Cytochrome C Tyrosine Nitration by Peroxynitrite. Biochemistry 2005, 44, 8038. (688) Abriata, L. A.; Cassina, A.; Tórtora, V.; Marín, M.; Souza, J. M.; Castro, L.; Vila, A. J.; Radi, R. Nitration of Solvent-Exposed Tyrosine 74 on Cytochrome C Triggers Heme Iron-Methionine 80 Bond Disruption. Nuclear Magnetic Resonance and Optical Spectroscopy Studies. J. Biol. Chem. 2009, 284, 17. (689) Capdevila, D. A.; Á lvarez-Paggi, D.; Castro, M. A.; Tórtora, V.; Demicheli, V.; Estrín, D. A.; Radi, R.; Murgida, D. H. Coupling of Tyrosine Deprotonation and Axial Ligand Exchange in Nitrocytochrome C. Chem. Commun. (Cambridge, U. K.) 2014, 50, 2592. (690) Radi, R.; Turrens, J. F.; Freeman, B. A. Cytochrome CCatalyzed Membrane Lipid Peroxidation by Hydrogen Peroxide. Arch. Biochem. Biophys. 1991, 288, 118. (691) Kagan, V. E.; Tyurin, V. A.; Jiang, J.; Tyurina, Y. Y.; Ritov, V. B.; Amoscato, A. A.; Osipov, A. N.; Belikova, N. A.; Kapralov, A. A.; Kini, V.; et al. Cytochrome C Acts as a Cardiolipin Oxygenase Required for Release of Proapoptotic Factors. Nat. Chem. Biol. 2005, 1, 223. (692) Shao, B.; Pennathur, S.; Heinecke, J. W. Myeloperoxidase Targets Apolipoprotein a-I, the Major High Density Lipoprotein Protein, for Site-Specific Oxidation in Human Atherosclerotic Lesions. J. Biol. Chem. 2012, 287, 6375. (693) Cimino, F.; Anderson, W. B.; Stadtman, E. R. Ability of Nonenzymic Nitration or Acetylation of E. Coli Glutamine Synthetase to Produce Effects Analogous to Enzymic Adenylylation. Proc. Natl. Acad. Sci. U. S. A. 1970, 66, 564. (694) Berlett, B. S.; Friguet, B.; Yim, M. B.; Chock, P. B.; Stadtman, E. R. Peroxynitrite-Mediated Nitration of Tyrosine Residues in Escherichia Coli Glutamine Synthetase Mimics Adenylylation: Relevance to Signal Transduction. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 1776. (695) Markland, F. S.; Bacharach, A. D.; Weber, B. H.; O’Grady, T. C.; Saunders, G. C.; Umemura, N. Chemical Modification of Yeast 3Phosphoglycerate Kinase. J. Biol. Chem. 1975, 250, 1301. (696) Lauritzen, A. M.; Landfear, S. M.; Lipscomb, W. N. Inactivation of the Catalytic Subunit of Aspartate Transcarbamylase by Nitration with Tetranitromethane. J. Biol. Chem. 1980, 255, 602. (697) Kochhar, S.; Dua, R. D. Chemical Modification of Liquefying Alpha-Amylase: Role of Tyrosine Residues at Its Active Center. Arch. Biochem. Biophys. 1985, 240, 757. (698) Shimokawa, T.; Kulmacz, R. J.; DeWitt, D. L.; Smith, W. L. Tyrosine 385 of Prostaglandin Endoperoxide Synthase Is Required for Cyclooxygenase Catalysis. J. Biol. Chem. 1990, 265, 20073. (699) Francescutti, D.; Baldwin, J.; Lee, L.; Mutus, B. Peroxynitrite Modification of Glutathione Reductase: Modeling Studies and Kinetic Evidence Suggest the Modification of Tyrosines at the Glutathione Disulfide Binding Site. Protein Eng., Des. Sel. 1996, 9, 189. (700) Haddad, I. Y.; Zhu, S.; Ischiropoulos, H.; Matalon, S. Nitration of Surfactant Protein a Results in Decreased Ability to Aggregate Lipids. Am. J. Physiol. 1996, 270, L281. (701) Greis, K. D.; Zhu, S.; Matalon, S. Identification of Nitration Sites on Surfactant Protein a by Tandem Electrospray Mass Spectrometry. Arch. Biochem. Biophys. 1996, 335, 396. (702) Roberts, E. S.; Lin, H.; Crowley, J. R.; Vuletich, J. L.; Osawa, Y.; Hollenberg, P. F. Peroxynitrite-Mediated Nitration of Tyrosine and Inactivation of the Catalytic Activity of Cytochrome P450 2b1. Chem. Res. Toxicol. 1998, 11, 1067.

(703) Blanchard-Fillion, B.; Souza, J. M.; Friel, T.; Jiang, G. C.; Vrana, K.; Sharov, V.; Barron, L.; Schoneich, C.; Quijano, C.; Alvarez, B.; et al. Nitration and Inactivation of Tyrosine Hydroxylase by Peroxynitrite. J. Biol. Chem. 2001, 276, 46017. (704) Kuhn, D. M.; Aretha, C. W.; Geddes, T. J. Peroxynitrite Inactivation of Tyrosine Hydroxylase: Mediation by Sulfhydryl Oxidation, Not Tyrosine Nitration. J. Neurosci. 1999, 19, 10289. (705) Nakagawa, H.; Komai, N.; Takusagawa, M.; Miura, Y.; Toda, T.; Miyata, N.; Ozawa, T.; Ikota, N. Nitration of Specific Tyrosine Residues of Cytochrome C Is Associated with Caspase-Cascade Inactivation. Biol. Pharm. Bull. 2007, 30, 15. (706) Balafanova, Z.; Bolli, R.; Zhang, J.; Zheng, Y.; Pass, J. M.; Bhatnagar, A.; Tang, X. L.; Wang, O.; Cardwell, E.; Ping, P. Nitric Oxide (NO) Induces Nitration of Protein Kinase Cepsilon (Pkcepsilon), Facilitating Pkcepsilon Translocation Via Enhanced Pkcepsilon -Rack2 Interactions: A Novel Mechanism of NO-Triggered Activation of Pkcepsilon. J. Biol. Chem. 2002, 277, 15021. (707) Vadseth, C.; Souza, J. M.; Thomson, L.; Seagraves, A.; Nagaswami, C.; Scheiner, T.; Torbet, J.; Vilaire, G.; Bennett, J. S.; Murciano, J. C.; et al. Pro-Thrombotic State Induced by PostTranslational Modification of Fibrinogen by Reactive Nitrogen Species. J. Biol. Chem. 2004, 279, 8820. (708) Ji, Y.; Bennett, B. M. Activation of Microsomal Glutathione STransferase by Peroxynitrite. Mol. Pharmacol. 2003, 63, 136. (709) Ji, Y.; Neverova, I.; Van Eyk, J. E.; Bennett, B. M. Nitration of Tyrosine 92 Mediates the Activation of Rat Microsomal Glutathione S-Transferase by Peroxynitrite. J. Biol. Chem. 2006, 281, 1986. (710) Hodara, R.; Norris, E. H.; Giasson, B. I.; Mishizen-Eberz, A. J.; Lynch, D. R.; Lee, V. M.; Ischiropoulos, H. Functional Consequences of Alpha-Synuclein Tyrosine Nitration: Diminished Binding to Lipid Vesicles and Increased Fibril Formation. J. Biol. Chem. 2004, 279, 47746. (711) Aggarwal, S.; Gross, C. M.; Kumar, S.; Datar, S.; Oishi, P.; Kalkan, G.; Schreiber, C.; Fratz, S.; Fineman, J. R.; Black, S. M. Attenuated Vasodilatation in Lambs with Endogenous and Exogenous Activation of Cgmp Signaling: Role of Protein Kinase G Nitration. J. Cell. Physiol. 2011, 226, 3104. (712) Aggarwal, S.; Gross, C. M.; Rafikov, R.; Kumar, S.; Fineman, J. R.; Ludewig, B.; Jonigk, D.; Black, S. M. Nitration of Tyrosine 247 Inhibits Protein Kinase G-1alpha Activity by Attenuating Cyclic Guanosine Monophosphate Binding. J. Biol. Chem. 2014, 289, 7948. (713) Shao, B.; Bergt, C.; Fu, X.; Green, P.; Voss, J. C.; Oda, M. N.; Oram, J. F.; Heinecke, J. W. Tyrosine 192 in Apolipoprotein A-I Is the Major Site of Nitration and Chlorination by Myeloperoxidase, but Only Chlorination Markedly Impairs Abca1-Dependent Cholesterol Transport. J. Biol. Chem. 2005, 280, 5983. (714) Monteiro, H. P.; Arai, R. J.; Travassos, L. R. Protein Tyrosine Phosphorylation and Protein Tyrosine Nitration in Redox Signaling. Antioxid. Redox Signaling 2008, 10, 843. (715) Martin, B. L.; Wu, D.; Jakes, S.; Graves, D. J. Chemical Influences on the Specificity of Tyrosine Phosphorylation. J. Biol. Chem. 1990, 265, 7108. (716) Gow, A. J.; Duran, D.; Malcolm, S.; Ischiropoulos, H. Effects of Peroxynitrite-Induced Protein Modifications on Tyrosine Phosphorylation and Degradation. FEBS Lett. 1996, 385, 63. (717) Kong, S. K.; Yim, M. B.; Stadtman, E. R.; Chock, P. B. Peroxynitrite Disables the Tyrosine Phosphorylation Regulatory Mechanism: Lymphocyte-Specific Tyrosine Kinase Fails to Phosphorylate Nitrated Cdc2(6−20)Nh2 Peptide. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 3377. (718) Saeki, M.; Maeda, S. P130Cas Is a Cellular Target Protein for Tyrosine Nitration Induced by Peroxynitrite. Neurosci. Res. 1999, 33, 325. (719) Newman, D. K.; Hoffman, S.; Kotamraju, S.; Zhao, T.; Wakim, B.; Kalyanaraman, B.; Newman, P. J. Nitration of Pecam-1 Itim Tyrosines Abrogates Phosphorylation and Shp-2 Binding. Biochem. Biophys. Res. Commun. 2002, 296, 1171. 1405

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Cerevisiae Involved in Mating Signal Transduction. Proteomics 2015, 15, 580. (738) Viera, L.; Radmilovich, M.; Vargas, M. R.; Dennys, C. N.; Wilson, L.; Barnes, S.; Franco, M. C.; Beckman, J. S.; Estévez, A. G. Temporal Patterns of Tyrosine Nitration in Embryo Heart Development. Free Radical Biol. Med. 2013, 55, 101. (739) Yakovlev, V. A.; Barani, I. J.; Rabender, C. S.; Black, S. M.; Leach, J. K.; Graves, P. R.; Kellogg, G. E.; Mikkelsen, R. B. Tyrosine Nitration of Iκbalpha: A Novel Mechanism for Nf-Kb Activation. Biochemistry 2007, 46, 11671. (740) Yakovlev, V. A.; Bayden, A. S.; Graves, P. R.; Kellogg, G. E.; Mikkelsen, R. B. Nitration of the Tumor Suppressor Protein P53 at Tyrosine 327 Promotes P53 Oligomerization and Activation. Biochemistry 2010, 49, 5331. (741) Shibuya, A.; Wada, K.; Nakajima, A.; Saeki, M.; Katayama, K.; Mayumi, T.; Kadowaki, T.; Niwa, H.; Kamisaki, Y. Nitration of Pparγ Inhibits Ligand-Dependent Translocation into the Nucleus in a Macrophage-Like Cell Line, Raw 264. FEBS Lett. 2002, 525, 43. (742) Wagner, A. H.; Hildebrandt, A.; Baumgarten, S.; Jungmann, A.; Müller, O. J.; Sharov, V. S.; Schöneich, C.; Hecker, M. Tyrosine Nitration Limits Stretch-Induced CD40 Expression and Disconnects CD40 Signaling in Human Endothelial Cells. Blood 2011, 118, 3734. (743) Estévez, A. G.; Radi, R.; Barbeito, L.; Shin, J. T.; Thompson, J. A.; Beckman, J. S. Peroxynitrite-Induced Cytotoxicity in Pc12 Cells: Evidence for an Apoptotic Mechanism Differentially Modulated by Neurotrophic Factors. J. Neurochem. 1995, 65, 1543. (744) Lin, K.-T.; Xue, J.-Y.; Nomen, M.; Spur, B.; Wong, P. Y.-K. Peroxynitrite-Induced Apoptosis in HL-60 Cells. J. Biol. Chem. 1995, 270, 16487. (745) Salgo, M. G.; Squadrito, G. L.; Pryor, W. A. Peroxynitrite Causes Apoptosis in Rat Thymocytes. Biochem. Biophys. Res. Commun. 1995, 215, 1111. (746) Moulian, N.; Truffault, F. r.; Gaudry-Talarmain, Y. M.; Serraf, A.; Berrih-Aknin, S. In Vivo and in Vitro Apoptosis of Human Thymocytes Are Associated with Nitrotyrosine Formation. Blood 2001, 97, 3521. (747) Aslan, M.; Yücel, I.; Akar, Y.; Yücel, G.; Ciftçioğlu, M. A.; Sanlioglu, S. Nitrotyrosine Formation and Apoptosis in Rat Models of Ocular Injury. Free Radical Res. 2006, 40, 147. (748) Manabe, Y.; Wang, J. M.; Murakami, T.; Warita, H.; Hayashi, T.; Shoji, M.; Abe, K. Expressions of Nitrotyrosine and Tunel Immunoreactivities in Cultured Rat Spinal Cord Neurons after Exposure to Glutamate, Nitric Oxide, or Peroxynitrite. J. Neurosci. Res. 2001, 65, 371. (749) Rodríguez-Ariza, A.; López-Sánchez, L. M.; González, R.; Corrales, F. J.; López, P.; Bernardos, A.; Muntané, J. Altered Protein Expression and Protein Nitration Pattern During D-GalactosamineInduced Cell Death in Human Hepatocytes: A Proteomic Analysis. Liver Int. 2005, 25, 1259. (750) Natal, C.; Modol, T.; Osés-Prieto, J. A.; López-Moratalla, N.; Iraburu, M. J.; López-Zabalza, M. J. Specific Protein Nitration in Nitric Oxide-Induced Apoptosis of Human Monocytes. Apoptosis 2008, 13, 1356. (751) Nazarewicz, R. R.; Zenebe, W. J.; Parihar, A.; Larson, S. K.; Alidema, E.; Choi, J.; Ghafourifar, P. Tamoxifen Induces Oxidative Stress and Mitochondrial Apoptosis Via Stimulating Mitochondrial Nitric Oxide Synthase. Cancer Res. 2007, 67, 1282. (752) Ye, Y.; Quijano, C.; Robinson, K. M.; Ricart, K. C.; Strayer, A. L.; Sahawneh, M. A.; Shacka, J. J.; Kirk, M.; Barnes, S.; Accavitti-Loper, M. A.; et al. Prevention of Peroxynitrite-Induced Apoptosis of Motor Neurons and Pc12 Cells by Tyrosine-Containing Peptides. J. Biol. Chem. 2007, 282, 6324. (753) Beckmann, J. S.; Ye, Y. Z.; Anderson, G.; Chen, J.; Accavitti, M. A.; Tarpey, M. M.; White, C. R. Extensive Nitration of Protein Tyrosines in Human Atherosclerosis Detected by Immunohistochemistry. Biol. Chem. Hoppe-Seyler 1994, 375, 81. (754) Ye, Y. Z.; Strong, M.; Huang, Z. Q.; Beckman, J. S. Antibodies That Recognize Nitrotyrosine. Methods Enzymol. 1996, 269, 201.

(720) Mondoro, T. H.; Shafer, B. C.; Vostal, J. G. PeroxynitriteInduced Tyrosine Nitration and Phosphorylation in Human Platelets. Free Radical Biol. Med. 1997, 22, 1055. (721) Li, X.; De Sarno, P.; Song, L.; Beckman, J. S.; Jope, R. S. Peroxynitrite Modulates Tyrosine Phosphorylation and Phosphoinositide Signalling in Human Neuroblastoma Sh-Sy5y Cells: Attenuated Effects in Human 1321n1 Astrocytoma Cells. Biochem. J. 1998, 331, 599. (722) Brito, C.; Naviliat, M.; Tiscornia, A. C.; Vuillier, F.; Gualco, G.; Dighiero, G.; Radi, R.; Cayota, A. M. Peroxynitrite Inhibits T Lymphocyte Activation and Proliferation by Promoting Impairment of Tyrosine Phosphorylation and Peroxynitrite-Driven Apoptotic Death. J. Immunol. 1999, 162, 3356. (723) Mallozzi, C.; Di Stasi, A.; Minetti, M. Peroxynitrite Modulates Tyrosine-Dependent Signal Transduction Pathway of Human Erythrocyte Band 3. FASEB J. 1997, 11, 1281. (724) Di Stasi, A. M. M.; Mallozzi, C.; Macchia, G.; Petrucci, T. C.; Minetti, M. Peroxynitrite Induces Tyrosine Nitration and Modulates Tyrosine Phosphorylation of Synaptic Proteins. J. Neurochem. 1999, 73, 727. (725) Mallozzi, C.; Di Stasi, A. M.; Minetti, M. Activation of Src Tyrosine Kinases by Peroxynitrite. FEBS Lett. 1999, 456, 201. (726) MacMillan-Crow, L. a.; Greendorfer, J. S.; Vickers, S. M.; Thompson, J. A. Tyrosine Nitration of C-Src Tyrosine Kinase in Human Pancreatic Ductal Adenocarcinoma. Arch. Biochem. Biophys. 2000, 377, 350. (727) Mallozzi, C.; Di Stasi, A. M.; Minetti, M. Nitrotyrosine Mimics Phosphotyrosine Binding to the Sh2 Domain of the Src Family Tyrosine Kinase Lyn. FEBS Lett. 2001, 503, 189. (728) Zhou, J.; Li, H.; Zeng, J.; Huang, K. Effects of PeroxynitriteInduced Protein Tyrosine Nitration on Insulin-Stimulated Tyrosine Phosphorylation in Hepg2 Cells. Mol. Cell. Biochem. 2009, 331, 49. (729) Joshi, M. S.; Mihm, M. J.; Cook, A. C.; Schanbacher, B. L.; Bauer, J. A. Alterations in Connexin 43 During Diabetic Cardiomyopathy: Competition of Tyrosine Nitration Versus Phosphorylation. J. Diabetes 2015, 7, 250. (730) Mallozzi, C.; D’Amore, C.; Camerini, S.; Macchia, G.; Crescenzi, M.; Petrucci, T. C.; Di Stasi, A. M. M. Phosphorylation and Nitration of Tyrosine Residues Affect Functional Properties of Synaptophysin and Dynamin I, Two Proteins Involved in ExoEndocytosis of Synaptic Vesicles. Biochim. Biophys. Acta, Mol. Cell Res. 2013, 1833, 110. (731) Elsasser, T. H.; Li, C.-J.; Caperna, T. J.; Kahl, S.; Schmidt, W. F. Growth Hormone (Gh)-Associated Nitration of Janus Kinase-2 at the 1007y-1008y Epitope Impedes Phosphorylation at This Site: Mechanism for and Impact of a Gh, Akt, and Nitric Oxide Synthase Axis on Gh Signal Transduction. Endocrinology 2007, 148, 3792. (732) Marcondes, S.; Cardoso, M. H. M.; Morganti, R. P.; Thomazzi, S. M.; Lilla, S.; Murad, F.; De Nucci, G.; Antunes, E. Cyclic GmpIndependent Mechanisms Contribute to the Inhibition of Platelet Adhesion by Nitric Oxide Donor: A Role for Alpha-Actinin Nitration. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 3434. (733) Reinehr, R.; Görg, B.; Höngen, A.; Häussinger, D. Cd95Tyrosine Nitration Inhibits Hyperosmotic and Cd95 Ligand-Induced Cd95 Activation in Rat Hepatocytes. J. Biol. Chem. 2004, 279, 10364. (734) Pilon, G.; Charbonneau, A.; White, P. J.; Dallaire, P.; Perreault, M.; Kapur, S.; Marette, A. Endotoxin Mediated-Inos Induction Causes Insulin Resistance Via ONOO- Induced Tyrosine Nitration of Irs-1 in Skeletal Muscle. PLoS One 2010, 5, No. e15912. (735) Speckmann, B.; Steinbrenner, H.; Grune, T.; Klotz, L. O. Peroxynitrite: From Interception to Signaling. Arch. Biochem. Biophys. 2016, 595, 153. (736) Naseem, K. M.; Low, S. Y.; Sabetkar, M.; Bradley, N. J.; Khan, J.; Jacobs, M.; Bruckdorfer, K. R. The Nitration of Platelet Cytosolic Proteins During Agonist-Induced Activation of Platelets. FEBS Lett. 2000, 473, 119. (737) Kang, J. W.; Lee, N. Y.; Cho, K. C.; Lee, M. Y.; Choi, D. Y.; Park, S. H.; Kim, K. P. Analysis of Nitrated Proteins in Saccharomyces 1406

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Proteinase Complex, Proteasome. Arch. Biochem. Biophys. 1994, 311, 329. (773) Grune, T.; Merker, K.; Sandig, G.; Davies, K. J. A. Selective Degradation of Oxidatively Modified Protein Substrates by the Proteasome. Biochem. Biophys. Res. Commun. 2003, 305, 709. (774) Grune, T.; Reinheckel, T.; Davies, K. J. Degradation of Oxidized Proteins in Mammalian Cells. FASEB J. 1997, 11, 526. (775) Pacifici, R. E.; Kono, Y.; Davies, K. J. A. Hydrophobicity as the Signal for Selective Degradation of Hydroxyl Radical-Modified Hemoglobin by the Multicatalytic Proteinase Complex, Proteasome. J. Biol. Chem. 1993, 268, 15405. (776) Pacifici, R. E.; Salo, D. C.; Davies, K. J. A. Macroxyproteinase (M.O.P.): A 670 Kda Proteinase Complex That Degrades Oxidatively DeNatured Proteins in Red Blood Cells. Free Radical Biol. Med. 1989, 7, 521. (777) Salo, D. C.; Pacifici, R. E.; Lin, S. W.; Giulivi, C.; Davies, K. J. A. Superoxide Dismutase Undergoes Proteolysis and Fragmentation Following Oxidative Modification and Inactivation. J. Biol. Chem. 1990, 265, 11919. (778) Shringarpure, R.; Grune, T.; Mehlhase, J.; Davies, K. J. A. Ubiquitin Conjugation Is Not Required for the Degradation of Oxidized Proteins by Proteasome. J. Biol. Chem. 2003, 278, 311. (779) Souza, J. M.; Choi, I.; Chen, Q.; Weisse, M.; Daikhin, E.; Yudkoff, M.; Obin, M.; Ara, J.; Horwitz, J.; Ischiropoulos, H. Proteolytic Degradation of Tyrosine Nitrated Proteins. Arch. Biochem. Biophys. 2000, 380, 360. (780) Curry-McCoy, T. V.; Osna, N. A.; Donohue, T. M., Jr. Modulation of Lysozyme Function and Degradation after Nitration with Peroxynitrite. Biochim. Biophys. Acta, Gen. Subj. 2009, 1790, 778. (781) Wang, Y. Y.; Lin, S. Y.; Chuang, Y. H.; Mao, C. H.; Tung, K. C.; Sheu, W. H. H. Protein Nitration Is Associated with Increased Proteolysis in Skeletal Muscle of Bile Duct Ligation-Induced Cirrhotic Rats. Metab., Clin. Exp. 2010, 59, 468. (782) Bulteau, A. L.; Mena, N. P.; Auchere, F.; Lee, I.; Prigent, A.; Lobsiger, C. S.; Camadro, J. M.; Hirsch, E. C. Dysfunction of Mitochondrial Lon Protease and Identification of Oxidized Protein in Mouse Brain Following Exposure to MPTP: Implications for Parkinson Disease. Free Radical Biol. Med. 2017, 108, 236. (783) Ohshima, H.; Friesen, M.; Brouet, I.; Bartsch, H. Nitrotyrosine as a New Marker for Endogenous Nitrosation and Nitration of Proteins. Food Chem. Toxicol. 1990, 28, 647. (784) Fell, V.; Hoskins, J. A.; Pollitt, R. J. The Labelling of Urinary Acids after Oral Doses of Deuterated L-Phenylalanine and L-Tyrosine in Normal Subjects. Quantitative Studies with Implications for the Deuterated Phenylalanine Load Test in Phenylketonuria. Clin. Chim. Acta 1978, 83, 259. (785) Tabrizi-Fard, M. A.; Maurer, T. S.; Fung, H. L. In Vivo Disposition of 3-Nitro-L-Tyrosine in Rats: Implications on Tracking Systemic Peroxynitrite Exposure. Drug Metab. Dispos. 1999, 27, 429. (786) Mani, A. R.; Pannala, A. S.; Orie, N. N.; Ollosson, R.; Harry, D.; Rice-Evans, C. A.; Moore, K. P. Nitration of Endogenous ParaHydroxyphenylacetic Acid and the Metabolism of Nitrotyrosine. Biochem. J. 2003, 374, 521. (787) Liu, M. C.; Yasuda, S.; Idell, S. Sulfation of Nitrotyrosine: Biochemistry and Functional Implications. IUBMB Life 2007, 59, 622. (788) Lipmann, F. Biological Sulfate Activation and Transfer. Science 1958, 128, 575. (789) Yasuda, S.; Idell, S.; Liu, M. C. Generation and Release of Nitrotyrosine O-Sulfate by Hepg2 Human Hepatoma Cells Upon Sin1 Stimulation: Identification of Sult1a3 as the Enzyme Responsible. Biochem. J. 2007, 401, 497. (790) Yasuda, S.; Yasuda, T.; Liu, M. Y.; Shetty, S.; Idell, S.; Boggaram, V.; Suiko, M.; Sakakibara, Y.; Fu, J.; Liu, M. C. Sulfation of Chlorotyrosine and Nitrotyrosine by Human Lung Endothelial and Epithelial Cells: Role of the Human Sult1a3. Toxicol. Appl. Pharmacol. 2011, 251, 104. (791) Kamisaki, Y.; Wada, K.; Bian, K.; Balabanli, B.; Davis, K.; Martin, E.; Behbod, F.; Lee, Y.-C.; Murad, F. An Activity in Rat

(755) Birnboim, H. C.; Lemay, A.-M.; Lam, D. K. Y.; Goldstein, R.; Webb, J. R. Cutting Edge: Mhc Class II-Restricted Peptides Containing the Inflammation-Associated Marker 3-Nitrotyrosine Evade Central Tolerance and Elicit a Robust Cell-Mediated Immune Response. J. Immunol. 2003, 171, 528. (756) Herzog, J.; Maekawa, Y.; Cirrito, T. P.; Illian, B. S.; Unanue, E. R. Activated Antigen-Presenting Cells Select and Present Chemically Modified Peptides Recognized by Unique Cd4 T Cells. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 7928. (757) Thomson, L.; Christie, J.; Vadseth, C.; Lanken, P. N.; Fu, X.; Hazen, S. L.; Ischiropoulos, H. Identification of Immunoglobulins That Recognize 3-Nitrotyrosine in Patients with Acute Lung Injury after Major Trauma. Am. J. Respir. Cell Mol. Biol. 2007, 36, 152. (758) Thomson, L.; Tenopoulou, M.; Lightfoot, R.; Tsika, E.; Parastatidis, I.; Martinez, M.; Greco, T. M.; Doulias, P. T.; Wu, Y.; Tang, W. H. W.; et al. Immunoglobulins against Tyrosine Nitrated Epitopes in Coronary Artery Disease. Circulation 2012, 126, 2392. (759) Khan, F.; Siddiqui, A. A. Prevalence of Anti-3-Nitrotyrosine Antibodies in the Joint Synovial Fluid of Patients with Rheumatoid Arthritis, Osteoarthritis and Systemic Lupus Erythematosus. Clin. Chim. Acta 2006, 370, 100. (760) Gauba, V.; Grünewald, J.; Gorney, V.; Deaton, L. M.; Kang, M.; Bursulaya, B.; Ou, W.; Lerner, R. A.; Schmedt, C.; Geierstanger, B. H.; et al. Loss of CD4 T-Cell-Dependent Tolerance to Proteins with Modified Amino Acids. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 12821. (761) Hardy, L. L.; Wick, D. A.; Webb, J. R. Conversion of Tyrosine to the Inflammation-Associated Analog 3′-Nitrotyrosine at Either TCR- or MHC-Contact Positions Can Profoundly Affect Recognition of the MHC Class I-Restricted Epitope of Lymphocytic Choriomeningitis Virus Glycoprotein 33 by CD8 T Ce. J. Immunol. 2008, 180, 5956. (762) Madhurantakam, C.; Duru, A. D.; Sandalova, T.; Webb, J. R.; Achour, A. Inflammation-Associated Nitrotyrosination Affects Tcr Recognition through Reduced Stability and Alteration of the Molecular Surface of the Mhc Complex. PLoS One 2012, 7, No. e32805. (763) Nagaraj, S.; Gupta, K.; Pisarev, V.; Kinarsky, L.; Sherman, S.; Kang, L.; Herber, D. L.; Schneck, J.; Gabrilovich, D. I. Altered Recognition of Antigen Is a Mechanism of Cd8+ T Cell Tolerance in Cancer. Nat. Med. 2007, 13, 828. (764) Molon, B.; Ugel, S.; Del Pozzo, F.; Soldani, C.; Zilio, S.; Avella, D.; De Palma, A.; Mauri, P.; Monegal, A.; Rescigno, M.; et al. Chemokine Nitration Prevents Intratumoral Infiltration of AntigenSpecific T Cells. J. Exp. Med. 2011, 208, 1949. (765) Rivett, A. J. Preferential Degradation of the Oxidatively Modified Form of Glutamine Synthetase by Intracellular Mammalian Proteases. J. Biol. Chem. 1985, 260, 300. (766) Levine, R. L.; Oliver, C. N.; Fulks, R. M.; Stadtman, E. R. Turnover of Bacterial Glutamine Synthetase: Oxidative Inactivation Precedes Proteolysis. Proc. Natl. Acad. Sci. U. S. A. 1981, 78, 2120. (767) Davies, K. J. A.; Lin, S. W.; Pacifici, R. E. Protein Damage and Degradation by Oxygen Radicals. Iv. Degradation of Denatured Protein. J. Biol. Chem. 1987, 262, 9914. (768) Grune, T.; Reinheckel, T.; Joshi, M.; Davies, K. J. A. Proteolysis in Cultured Liver Epithelial Cells During Oxidative Stress. J. Biol. Chem. 1995, 270, 2344. (769) Davies, K. J. Intracellular Proteolytic Systems May Function as Secondary Antioxidant Defenses: An Hypothesis. J. Free Radicals Biol. Med. 1986, 2, 155. (770) Davies, K. J. A.; Lin, S. W. Degradation of Oxidatively Denatured Proteins in Escherichia Coli. Free Radical Biol. Med. 1988, 5, 215. (771) Davies, K. J. A.; Lin, S. W. Oxidatively Denatured Proteins Are Degraded by an Atp-Independent Proteolytic Pathway in Escherichia Coli. Free Radical Biol. Med. 1988, 5, 225. (772) Giulivi, C.; Pacifici, R. E.; Davies, K. J. A. Exposure of Hydrophobic Moieties Promotes the Selective Degradation of Hydrogen Peroxide-Modified Hemoglobin by the Multicatalytic 1407

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408

Chemical Reviews

Review

Tissues That Modifies Nitrotyrosine-Containing Proteins. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 11584. (792) Irie, Y.; Saeki, M.; Kamisaki, Y.; Martin, E.; Murad, F. Histone H1.2 Is a Substrate for Denitrase, an Activity That Reduces Nitrotyrosine Immunoreactivity in Proteins. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 5634. (793) Koeck, T.; Fu, X.; Hazen, S. L.; Crabb, J. W.; Stuehr, D. J.; Aulak, K. S. Rapid and Selective Oxygen-Regulated Protein Tyrosine Denitration and Nitration in Mitochondria. J. Biol. Chem. 2004, 279, 27257. (794) Balabanli, B.; Kamisaki, Y.; Martin, E.; Murad, F. Requirements for Heme and Thiols for the Nonenzymatic Modification of Nitrotyrosine. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 13136. (795) Chen, H.-J. C.; Chang, C.-M.; Lin, W.-P.; Cheng, D.-L.; Leong, M.-I. H2O2/Nitrite-Induced Post-Translational Modifications of Human Hemoglobin Determined by Mass Spectrometry: Redox Regulation of Tyrosine Nitration and 3-Nitrotyrosine Reduction by Antioxidants. ChemBioChem 2008, 9, 312. (796) Chen, H. J. C.; Chen, Y. M.; Chang, C. M. Lipoyl Dehydrogenase Catalyzes Reduction of Nitrated DNA and Protein Adducts Using Dihydrolipoic Acid or Ubiquinol as the Cofactor. Chem.-Biol. Interact. 2002, 140, 199. (797) Smyth, G. E.; Orsi, B. A. Nitroreductase Activity of Nadh Dehydrogenase of the Respiratory Redox Chain. Biochem. J. 1989, 257, 859. (798) Smallwood, H. S.; Lourette, N. M.; Boschek, C. B.; Bigelow, D. J.; Smith, R. D.; Pasa-Tolić, L.; Squier, T. C. Identification of a Denitrase Activity against Calmodulin in Activated Macrophages Using High-Field Liquid Chromatography–Fticr Mass Spectrometry. Biochemistry 2007, 46, 10498. (799) Görg, B.; Qvartskhava, N.; Voss, P.; Grune, T.; Häussinger, D.; Schliess, F. Reversible Inhibition of Mammalian Glutamine Synthetase by Tyrosine Nitration. FEBS Lett. 2007, 581, 84. (800) Kang, M.; Akbarali, H. I. Denitration of L-Type Calcium Channel. FEBS Lett. 2008, 582, 3033. (801) Léger, C. L.; Torres-Rasgado, E.; Fouret, G.; Carbonneau, M.A. First Evidence for an LDL- and HDL-Associated Nitratase Activity That Denitrates Albumin-Bound Nitrotyrosine–Physiological Consequences. IUBMB Life 2008, 60, 73. (802) Deeb, R. S.; Nuriel, T.; Cheung, C.; Summers, B.; Lamon, B. D.; Gross, S. S.; Hajjar, D. P. Characterization of a Cellular Denitrase Activity That Reverses Nitration of Cyclooxygenase. Am. J. Phys. Heart. Circ. Phys. 2013, 305, H687. (803) Ackaert, C.; Kofler, S.; Horejs-Hoeck, J.; Zulehner, N.; Asam, C.; Von Grafenstein, S.; Fuchs, J. E.; Briza, P.; Liedl, K. R.; Bohle, B.; et al. The Impact of Nitration on the Structure and Immunogenicity of the Major Birch Pollen Allergen Bet V 1.0101. PLoS One 2014, 9, No. e104520. (804) Yokoyama, K.; Uhlin, U.; Stubbe, J. A Hot Oxidant, 3NO2Y122 Radical, Unmasks Conformational Gating in Ribonucleotide Reductase. J. Am. Chem. Soc. 2010, 132, 15368. (805) Jarmuła, A.; Rode, W. Computational Study of the Effects of Protein Tyrosine Nitrations on the Catalytic Activity of Human Thymidylate Synthase. J. Comput.-Aided Mol. Des. 2013, 27, 45. (806) Lin, Y. W.; Shu, X. G.; Du, K. J.; Nie, C. M.; Wen, G. B. Computational Insight into Nitration of Human Myoglobin. Comput. Biol. Chem. 2014, 52, 60. (807) Quaroni, L.; Smith, W. E. Nitration of Internal Tyrosine of Cytochrome C Probed by Resonance Raman Scattering. Biospectroscopy 1999, 5, S71. (808) Gómez-Mingot, M.; Alcaraz, L. A.; Heptinstall, J.; Donaire, A.; Piccioli, M.; Montiel, V.; Iniesta, J. Electrochemical Nitration of Myoglobin at Tyrosine 103: Structure and Stability. Arch. Biochem. Biophys. 2013, 529, 26. (809) Hamilton, R. T.; Asatryan, L.; Nilsen, J. T.; Isas, J. M.; Gallaher, T. K.; Sawamura, T.; Hsiai, T. K. LDL Protein Nitration: Implication for LDL Protein Unfolding. Arch. Biochem. Biophys. 2008, 479, 1.

(810) Uversky, V. N.; Yamin, G.; Munishkina, L. A.; Karymov, M. A.; Millett, I. S.; Doniach, S.; Lyubchenko, Y. L.; Fink, A. L. Effects of Nitration on the Structure and Aggregation of Alpha-Synuclein. Mol. Brain Res. 2005, 134, 84.

1408

DOI: 10.1021/acs.chemrev.7b00568 Chem. Rev. 2018, 118, 1338−1408