Vous êtes sur la page 1sur 6

PERSPECTIVE

Nitric oxide, oxidants, and protein tyrosine nitration


Rafael Radi*
Departamento de Bioquı́mica and Center for Free Radical and Biomedical Research, Facultad de Medicina, Universidad de la República,
Avda. General Flores 2125, 11800 Montevideo, Uruguay

Communicated by Louis J. Ignarro, University of California School of Medicine, Los Angeles, CA, January 12, 2004 (received for review November 12, 2003)

The occurrence of protein tyrosine nitration under disease conditions is now firmly established and represents a shift from the signal
transducing physiological actions of •NO to oxidative and potentially pathogenic pathways. Tyrosine nitration is mediated by reactive ni-
trogen species such as peroxynitrite anion (ONOOⴚ) and nitrogen dioxide (•NO2), formed as secondary products of •NO metabolism in the
•ⴚ
presence of oxidants including superoxide radicals (O2 ), hydrogen peroxide (H2O2), and transition metal centers. The precise interplay be-

tween NO and oxidants and the identification of the proximal intermediate(s) responsible for nitration in vivo have been under contro-
versy. Despite the capacity of peroxynitrite to mediate tyrosine nitration in vitro, its role on nitration in vivo has been questioned, and al-
ternative pathways, including the nitrite兾H2O2兾hemeperoxidase and transition metal-dependent mechanisms, have been proposed. A
balanced analysis of existing evidence indicates that (i) different nitration pathways can contribute to tyrosine nitration in vivo, and (ii)
•ⴚ
most, if not all, nitration pathways involve free radical biochemistry with carbonate radicals (CO3 ) and兾or oxo–metal complexes oxidizing

tyrosine to tyrosyl radical followed by the diffusion-controlled reaction with NO2 to yield 3-nitrotyrosine. Although protein tyrosine nitra-
tion is a low-yield process in vivo, 3-nitrotyrosine has been revealed as a relevant biomarker of •NO-dependent oxidative stress; addition-
ally, site-specific nitration focused on particular protein tyrosines may result in modification of function and promote a biological effect.
Tissue distribution and quantitation of protein 3-nitrotyrosine, recognition of the predominant nitration pathways and individual identifi-
cation of nitrated proteins in disease states open new avenues for the understanding and treatment of human pathologies.

arly after the discovery of the sig- and chronic disease states, and can be a ute to protein tyrosine nitration. Interest-

E nal transducing physiological func-


tions of the free radical nitric ox-
ide (•NO) in the vasculature and
nervous system (e.g., vasodilation and
neurotransmission), it became evident that
predictor of disease risk.
Seminal work by Beckman et al. (6) and
Ischiropoulos et al. (7) demonstrated the
capacity of peroxynitrite to cause protein
tyrosine nitration in vitro and established
ingly, the alternative nitration pathways
share common characteristics, because
they involve free radical biochemistry with
the participation of transient tyrosyl,
•NO , and carbonate (CO•⫺) radicals
2 3
•NO could also participate as a cytotoxic the concept that biologically produced and兾or oxo–metal complexes.
effector molecule and兾or a pathogenic intermediates could promote nitration in
mediator when produced at high rates by vivo (8, 9), as was also suggested in an NO Reaction with Superoxide and the
either inflammatory stimuli-induced nitric earlier work by Ohshima et al. (10). How- Formation of Peroxynitrite
oxide synthase (iNOS) or overstimulation ever, the role of peroxynitrite as a central A relevant debate in the field has been
of the constitutive forms (eNOS and species in biological nitration has been whether peroxynitrite can be produced
nNOS) (1). Much of •NO-mediated more recently questioned (11, 12), and biologically at levels high enough to play a
pathogenicity depends on the formation alternative pathways, most notably mecha- significant role in •NO-dependent pathol-
of secondary intermediates such as per- nisms that depend on the formation of ogy (16, 17). Because peroxynitrite is a
oxynitrite anion (ONOO⫺) and nitrogen •NO by the action of hemeperoxidases
2 transient species with a biological half-life
dioxide (•NO2) that are typically more and兾or transition metal complexes on a [10–20 ms (18)] even shorter than that of
•NO [1–30 s (1)], it cannot be directly
reactive and toxic than •NO per se (2). main product of •NO metabolism, nitrite
The formation of reactive nitrogen species (NO⫺ 2 ), have been presented as contribu- measured, and its presence must be in-
from •NO requires the presence of oxi- tors (12–15). ferred from a combination of analytical,
dants such as superoxide radicals (O•⫺ 2 ), In addition to nitration, inflammatory pharmacological, and兾or genetic ap-
hydrogen peroxide (H2O2), and transition conditions promote other oxidative modi- proaches (19).† The biological reactions of
metal centers, the concentration of which fications in tyrosine such as chlorination, NO with O•⫺ 2 were initially proposed dur-
can be increased either by •NO itself or bromination, and hydroxylation to ing studies that characterized the chemical
by the same mediators that up-regulate 3-chloro-, 3-bromo-, or 3-hydroxytyrosine, nature of the endothelial-derived relaxing
•NO production. Nitrogen dioxide can the detection of which may assist in iden- factor as •NO (20–22). Indeed, the bio-
also be formed in hydrophobic environ- tifying preferential nitration pathways. logical half-life and actions of •NO in the
ments from the reactions of •NO with The extent of these oxidative modifica- vasculature were prolonged by superoxide
molecular oxygen, where these species tions in vivo can reach similar values per dismutase (SOD), the enzyme that elimi-
concentrate (3, 4). One of the molecular milligram of tissue protein (e.g., 10–100 nates O•⫺2 at diffusion-controlled rates,
footprints left by the reactions of reactive pmol兾mg) and are comparable to the lev-
nitrogen species with biomolecules is the els of protein S-nitrosation. It is also im-
nitration (i.e., addition of nitro group, portant to appreciate that other oxidative Abbreviations: NOS, NO synthase; MPO, myeloperoxidase;
EPO, eosinophil peroxidase; SOD, superoxide dismutase;
⫺NO2) of protein tyrosine residues to processes triggered by reactive nitrogen LMW, low molecular weight.
3-nitrotyrosine. The formation of protein species such as thiol and methionine oxi- *E-mail: rradi@fmed.edu.uy.
3-nitrotyrosine was originally addressed in dation, disruption of iron–sulfur clusters, †The debate on the biological production and measure-
early protein chemistry studies with tetra- and oxidation of transition metal centers ment of peroxynitrite is analogous to a previous one on
nitromethane aimed at establishing the (2) can, in many cases, be more relevant •⫺
O2 that arose after the discovery of SOD by J. M. McCord
function of tyrosines in proteins (5). This than nitration in the promotion cell dys- and I. Fridovich. The recurring discussions during the first
decade led I. Fridovich and associates to invest large re-
now-established posttranslational modifi- function兾death. search efforts into fully (and redundantly) establishing the
cation attracts considerable interest to A balanced analysis of all of the exist- concept that O2•⫺ was a biologically relevant reactive in-
biomedical research, because it can alter ing in vitro and in vivo evidence indicates termediate.
protein function, is associated to acute that more than one pathway can contrib- © 2004 by The National Academy of Sciences of the USA

www.pnas.org兾cgi兾doi兾10.1073兾pnas.0307446101 PNAS 兩 March 23, 2004 兩 vol. 101 兩 no. 12 兩 4003– 4008
and were decreased by enhanced O•⫺ 2 for- The Transition Metal-Dependent M⫺1s⫺1) oxo–iron intermediates (14, 15)
mation by redox-cycling molecules or high Formation of Strong Oxidants and (Eqs. 3–5):
oxygen tensions. Thus, early data firmly Nitrogen Dioxide
established the existence of •NO and O•⫺ MPO 共Fe3⫹兲 ⫹ H2O2 3
2 Elevations of redox active metals systems
interactions, although at the time the re- such as iron (Fe), copper (Cu), and man- MPO •␲⫹ 共Fe4⫹兲 共compound I兲 [3]
action was perceived as one to just limit ganese (Mn) contribute to oxidative dam-
the biological half-life of •NO to yield age in disease states (12, 14). Also, in- MPO •␲⫹ 4⫹
共Fe 兲 ⫹ NO⫺
2 3
relatively inert nitrate (NO⫺ 3 ). flammatory conditions trigger the release
The rate constant of the reaction of of leukocyte peroxidases, including myelo- MPO 共Fe4⫹兲 ⫹ •NO2 共compound II兲
O•⫺ •
2 with NO is larger (⬃10
10 M⫺1s⫺1)
peroxidase (MPO) and eosinophil peroxi-
[4]
than with SOD (1–2 ⫻ 109 M⫺1s⫺1), and dase (EPO), which play central roles in
therefore •NO sometimes outcompetes tissue oxidant formation (13, 15, 41). Oxi- MPO 共Fe 兲 ⫹ 4⫹
NO⫺
2 3
SOD for O•⫺ •
2 . The reaction of NO with dative stress promotes metal mobilization
O•⫺
2 leads to the diffusion-controlled for- from proteins by mechanisms that include MPO 共Fe3⫹兲 ⫹ •NO2 [5]
mation of peroxynitrite anion (ONOO⫺) O•⫺2 -mediated oxidation of labile iron–
(Eq. 1; for a review, see ref. 2): sulfur clusters [e.g., aconitase and other Other hemeproteins, free hemin, and
dehydratases (42)], redox-dependent LMW complexes can undergo similar re-

NO ⫹ O•⫺ ⫺ actions to produce •NO2, although pre-
2 3 ONOO [1] metal release from storage or transport
proteins (transferrin, ferritin, and cerulo- sumably at lower rates (12).
The influence of SOD on the half-life of plasmin), oxidative modifications of heme The reactions of peroxynitrite with oxi-
•NO and the formation of peroxynitrite proteins that result in heme release dized transition metal centers including
under various fluxes of •NO and O•⫺ 2 (23) and兾or degradation (myoglobin and cyto- MPO (45) can sometimes concomitantly
and in different vascular tissue layers (24) chrome c), and histidine oxidation with yield oxo–metal complexes and •NO2. In-
have been recently analyzed. Being O•⫺ deed, the OOO bond in ONOO⫺ is sig-
2 , disruption of metal coordination sites (i.e.,
one of the factors controlling the half-life, Cu-Zn SOD), among others. Metals re- nificantly weakened by association with
for instance, of vascular- (20–22) and in- leased from proteins are trapped by low Lewis acids such as H⫹, CO2, or transition
flammatory cell- (25) derived •NO, the metal centers (MenX), which favor its ho-
molecular-weight (LMW) chelators such
molysis (ref. 46; Eq. 6):
subsequent formation of peroxynitrite is as adenine nucleotides (ATP3-and ADP2-),
obligatory.‡ However, •NO and O•⫺ 2 reac- tri- and dicarboxylic acids (citrate, isocit- ONOO⫺ ⫹ MenX 3
tions do not necessarily result in tissue rate, and piruvate), or phosphate or bind
oxidative injury and in some cases can to macromolecules and in conjunction ONOO–MenX 3 •NO2 ⫹ •O–MenX 3
even be cytoprotective (30, 31). Indeed, with free hemin (12, 14) become more

low levels of peroxynitrite could be detox- readily accessible to undesired redox bio- NO2 ⫹ O⫽Men⫹1X [6]
ified by enzymatic and nonenzymatic sys- chemistry (43). Typically, in the reductive
Although the H⫹- or CO2-catalyzed ho-
tems (32–35) and then, direct toxic effects environment of tissues, these metal com-
molysis results in radical yields of ⬇30%
of either •NO or O•⫺ 2 neutralized. Forma-
plexes become reduced, and in particular
and 35%, respectively, oxidant yield from
tion of peroxynitrite might also play subtle ferrous iron can react with hydrogen per-
metals (Eq. 6) can, in some cases, reach
roles in signal transduction processes (36– oxide to form variable amounts of free
•OH and oxoferryl species [Fenton reac- up to 100% (6, 46). Thus, there is positive
39). However, identification of oxidation feedback among redox active metals and
and nitration products in conjunction with tion (43); Eq. 2]:
oxidants and •NO to facilitate the produc-
pharmacological and兾or genetic ap- tion of the mediators of nitration.
proaches has substantiated direct toxicity X-Fe3⫹ ⫹ •OH [2a]
of peroxynitrite in pathophysiologically 2⫹ m Iron Regulatory Proteins, •NO, and
X-Fe ⫹ H 2O 2
relevant conditions (see ref. 19 and refer- Oxidant Stress
ences therein and below). Peroxynitrite is n NO and oxidants influence the biochemi-
a strong one- and two-electron oxidant X-Fe4⫹⫽O [2b] cal mechanisms that regulate cell iron ho-
and can also evolve to •OH, •NO2, and meostasis and that involve iron regulatory
CO•⫺ The large reactivity of free •OH with proteins 1 and 2 (IRP-1兾2). IRPs play
3 radicals. Peroxynitrite does not re-
most biomolecules and its minimal diffu- pivotal roles in assuring the maintenance
act directly with tyrosine (2) but oxidizes
sion distance (three to four molecular di- of iron in the LMW pool at levels suffi-
and nitrates it through its radical products.
ameters) makes it unselective for protein cient for the biosynthesis of heme- and
Revisions on the biological chemistry of
modification unless site-specifically formed nonheme proteins but not exceeding
peroxynitrite and its derived radicals have
within metal-binding sites; oxo–iron com- amounts that may impose a risk of metal-
been published recently (2, 19, 40).
plexes live longer and are more selective dependent oxidative damage (47). IRP-1
in substrate兾target reactions. Oxidized in its inactive state contains a cubane
transition metals also react with H2O2 to 4Fe-4S cluster with a labile iron (Fe␣)
‡Under some circumstances, SOD could augment the bio-
form secondary oxidizing species, e.g., and displays aconitase activity. Total dis-
availability of •NO by O2•⫺-independent mechanisms such ferric heme reaction with H2O2 yields an assembly of the cluster to the apoprotein
as (i) the SOD-mediated oxidation of nitroxyl anion (NO⫺) oxoferryl intermediate plus a porphyrin yields a fully active IRP-1, which in turn
to •NO (26);. and (ii) the SOD-mediated reduction of S- radical cation (14). promotes down-regulation of ferritin and
nitrosothiols (27). These alternative mechanisms for the
actions of SOD could mainly operate intracellularly (28).
Oxo–metal complexes, in turn, favor up-regulation of the transferrin receptor
The immediate vascular responses to extracellularly added formation of •NO2. Indeed, NO⫺ 2 evolves with an overall increase of LMW iron.
SOD are due to the prevention of the oxidative inactiva- to •NO2 via a one-electron oxidation. For Mild oxidative stress and兾or •NO levels
tion of readily diffusible •NO. NO兾O2•⫺ interactions in the
instance, MPO (and EPO) can oxidize lead to IRP-1 activation, whereas over-
vasculature are viewed as contributions to the develop-
ment of vascular disease states, including atherogenesis, NO⫺ 2 by compounds I (k ⫽ 2 ⫻ 10
6
production of reactive oxygen species or
⫺1 ⫺1
hypertension, and hyperglycemia (29). M s ) (44) and II (k ⫽ 5.5 ⫻ 102 peroxynitrite extends their action beyond

4004 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0307446101 Radi


Then, the polarized carrier or NO⫹ 2 itself
could attack tyrosine as a two-electron
acceptor to yield a nitroarenium ion inter-
mediate, which then evolves to 3-nitroty-
rosine and a proton (50) (Eqs. 8–10):

NO2–O–MenX ⫹ TyrH 3

关TyrH-NO2–O–MenX兴 3

NO2–Tyr ⫹ O⫽MenX ⫹ H⫹ [8]

NO⫹
2 ⫹ TyrH 3

关NO2–TyrH⫹兴 3 NO2–Tyr ⫹ H⫹ [9]

O⫽MnnX ⫹ 2H⫹ 3 MenX ⫹ H2O


[10]
The half-life of NO⫹
is too short in aque-
2
Fig. 1. The free radical pathways of tyrosine nitration. ous systems because of its fast reaction
with H2O and subsequent decay to NO⫺ 3,
and therefore the metal-bound NO⫹ 2 com-
reversible cluster disassembly and pro- thiols. The dimerization of tyrosyl radicals plex would be the proximal reactant (Eq.
duces irreversible inactivation via oxida- to 3,3⬘-dityrosine competes with the for- 8). This mechanism of nitration§ may op-
tion of critical thiols. IRP-2 does not con- mation of 3-nitrotyrosine. However, pro- erate for some metal complexes such as in
tain the iron–sulfur cluster, and its activity tein tyrosyl radicals can be stabilized, with the peroxynitrite-dependent nitrations of
is controlled by level of expression and intra- and intermolecular dimerization Tyr-108 of bovine CuZn SOD (6) or
oxidant-stimulated proteolytic limited due to spatial and diffusional con- Tyr-34 of human MnSOD (51), although
degradation. straints, both in aqueous and hydrophobic current data cannot rule out a free radical
The oxidant-dependent inactivation of compartments, in which case their reac- mechanism.
IRPs may be viewed as a negative feed- tion with •NO2 is favored. Another com- Electrophilic aromatic nitration can also
back to inhibit the amplification of oxida- peting pathway is the formation of 3-hy- occur in vitro by the action of nitryl chlo-
tive damage but, in fact, it disrupts the droxytyrosine, which can be mediated by ride (NO2Cl), formed from the relatively
reversible mechanisms for controlling iron •OH and oxo–metal complexes (15). An
slow reaction of MPO-derived hypochlor-
homeostasis and may render the cell more alternative radical mechanism for tyrosine ous acid with NO⫺ 2 . However, the role of
vulnerable. Also, an increase of the labile nitration involves the reaction of a tyrosyl NO2Cl as a nitrating molecule in vivo is
iron pool by •NO may favor, on initiation radical with •NO to form 3-nitrosotyro- highly unlikely (52). In summary, present
of oxidant stress, nitration reactions in sine followed by a sequential two-electron evidence cannot yet confirm that electro-
target cells before the homeostatic re- oxidation to 3-nitrotyrosine via tyrosine philic aromatic nitration is a biologically
sponses for iron sequestration at the pro- iminoxyl radical; this mechanism may relevant mechanism.
tein expression level are achieved. Inter- operate in transition metal-containing
estingly, during inflammatory conditions, proteins that can readily oxidize The Limited Efficiency of the Nitration
IRP-1 is found to be not only inactivated 3-nitrosotyrosine before the latter can Reactions in Biology
but also nitrated (48). The role of IRPs in reversibly generate tyrosyl radicals and Biological nitration yields are low, and,
the regulation of iron-dependent nitration •NO, such as in the case of prostaglandin
under inflammatory conditions, one to
reactions remains to be elucidated. H synthase-2 (49). five 3-nitrotyrosine residues per 10,000
The Free Radical Pathways to tyrosine residues (100–500 ␮mol兾mol) are
Is Electrophilic Aromatic Nitration a detected (15, 53). This is due to a multi-
Tyrosine Nitration Biological Mechanism? plicity of processes (reactions, diffusion)
The very existence of 3-nitrotyrosine in The reactions of peroxynitrite with transi- that the precursors of nitrating species
vivo supports free radical biochemistry (O•⫺ •
tion metal centers may promote tyrosine 2 , H2O2, NO) and the proximal in-
(Fig. 1). In fact, the two most widely in- termediates of nitration (•NO2, CO•⫺
nitration by a free radical-independent 3 ,
voked mechanisms of biological nitration, oxo–metal complexes) can undergo,
chemistry, namely electrophilic aromatic
namely the peroxynitrite and the heme- to repair mechanisms of the tyrosyl
nitration (6, 7). In this mechanism, per-
peroxidase pathways, lead to the concomi- radical such as reduction by glutathione
oxynitrite would first form a complex with
tant formation of tyrosyl radicals and and, presumably, by metabolism of
•NO , which combine at diffusion- the transition metal to yield a polarized
2
carrier of nitronium cation (NO⫹ 2 ), which 3-nitrotyrosine.¶
controlled rates to form 3-nitrotyrosine
may then decompose to free NO⫹ 2 by het-
(Fig. 1).
erolysis (Eq. 7):
The oxidants leading to tyrosyl radical §In the electrophilic aromatic nitration mechanism, there is

are either CO•⫺3 or oxo–metal complexes ONOO⫺ ⫹ MenX 3 no net redox change in the metal center (Eqs. 7–10). How-
and, to a lesser extent, •OH. Importantly, ever, direct and unambiguous spectroscopic observation is
•NO alone is inefficient in promoting sometimes difficult due to low steady-state concentrations
2 ONOO–MenX 3 and兾or the absorption of peroxynitrite–metal complexes;
nitration, because it must react first with additionally, there is no formation of •NO2.
tyrosine to yield tyrosyl radical, a reaction NO␦2⫹–O␦⫺–MenX 3 ¶Nitrated proteins can also undergo enhanced proteolytic
that is slow compared to other processes degradation (54), and reports have suggested that a deni-
that •NO2 undergoes, for example, with NO⫹
2 ⫹ O⫽Me X n
[7] trase activity may be present in vivo (55).

Radi PNAS 兩 March 23, 2004 兩 vol. 101 兩 no. 12 兩 4005


The idea of peroxynitrite as a mediator dases may serve as catalytic sinks of •NO that are more superficial than the metal-
of biological nitration was challenged by and influence nitration yields (64). adjacent Tyr-34, and that the association
observations demonstrating that the oxi- of nitration plus inactivation of MnSOD
dative biochemistry of peroxynitrite could The Biological Significance of Protein may serve to unravel the chemical nature
be strongly modulated by the relative Tyrosine Nitration of the nitrating species. A recent quantita-
fluxes of •NO and O•⫺ 2 (56–58). Indeed, The small fraction of nitrated protein in tion of the extent of MnSOD nitration in
in homogenous systems, reasonable nitra- the context of total tissue protein ques- rat kidneys during angiotensin II infusion
tion yields are obtained only at an optimal tions its possible biological relevance. It is provided support for the hypothesis of
•NO兾O•⫺ flux ratio ⬃ one (11, 58), be-
2 noteworthy to indicate, however, that (i) a nitration as a direct cause of enzyme inac-
cause excess of either •NO or O•⫺ 2 reacts relatively limited number of proteins are tivation (71). Similarly, the vascular en-
with tyrosyl radicals and •NO2 and deacti- preferential targets of nitration, and (ii) zyme prostacyclin synthase (PGI2) syn-
vates the process. Also, under low radical within these proteins, only one or a few thase becomes site-specifically nitrated by
fluxes, the direct reaction of tyrosine with specific tyrosines can be nitrated (65); peroxynitrite at heme-adjacent Tyr-430 in
•NO competes with the recombination
2 because the species participating in nitra- a process catalyzed by the active site
reaction of •NO2 with the tyrosyl radical. tion have short diffusion distances and heme-thiolate involving transient ferryl-
The observations demonstrated that free site-specific nitration occurs at metal- or species (72, 73). Note that the subcellular
radical processes triggered by pure per- hemeperoxidase-binding sites, nitration distribution of PGI2 synthase in vascular
oxynitrite may result in outcomes different reactions can be concentrated on proteins, endothelium, colocalized with eNOS at
than those initiated by the simultaneous cell, or tissue compartments as revealed
production of •NO and O•⫺ the caveola, may further contribute for it
2 (11, 58) do by immunochemical (with antibodies
not reconcile well with solid biological to be a sensitive and critical target of per-
against protein 3-nitrotyrosine) and pro- oxynitrite.储 Recent experiments indicate
evidence (see below), supporting per- teomic-based methods. Thus, although
oxynitrite-mediated nitration in vivo. This that inflammation of artery walls results
there is a dilutional effect in quantitating in rapid PGI2 synthase nitration by a per-
apparent conflict can be solved by appre- total nitrated protein by analytical tech-
ciating the role of two key factors present oxynitrite-dependent mechanism (73).
niques in tissues, nitration can be focused Examples of MnSOD and PGI2 synthase
in cells and tissues: first, SOD, which sig- on specific tyrosines and potentially result
nificantly lessens increases in the steady- affirm the concept that transition metal
in modification, loss, or gain of function. catalysis provides selectivity and efficiency
state levels of O•⫺
2 , despite large varia- Addition of a ⫺NO2 group to tyrosine
tions of O•⫺2 production rates; and
to peroxynitrite as a biological nitrating
lowers the pKa of its phenolic ⫺OH by agent. Finally, it is relevant to note that
second, the rapid transmembrane diffu- 2–3 units and adds a bulky substituent; if
sion and兾or cell consumption of •NO that actin, which can constitute 5% or more of
placed in relevant tyrosines, nitration can
operates as a drain of excess •NO; thus, cell protein, is heavily nitrated in sickle
alter protein function and conformation,
augmentation of either •NO or O•⫺ 2 cell disease and that the extents of nitra-
impose steric restrictions, and also inhibit
fluxes in vivo will serve to trap more of tion observed in tissues are sufficient to
tyrosine phosphorylation. However, to
the partner radical, which otherwise is induce defective cytoskeletal polymeriza-
have biological significance, a loss-of-func-
being consumed by other processes in- tion (53).
tion modification requires a large fraction
stead of accumulating and interfering with The clinical relevance of protein ty-
the nitration process. Experimental con- of protein to become nitrated at specific
rosine nitration has been recently under-
founding factors have also obscured this critical tyrosines and result in 3-nitroty-
scored by the observation of a strong as-
discussion, including, in some works, the rosine to tyrosine ratios in the range of
0.1–1.0; documentation of these large ra- sociation between protein 3-nitrotyrosine
presence of uric acid (see below) and the levels and coronary artery disease risk
lack of consideration of the diffusional tios for a given protein in vivo is scanty,
and it is doubtful that many proteins will (75). Circulating levels of protein 3-nitro-
properties of extracellular peroxynitrite tyrosine may serve as a biomarker to as-
and CO•⫺ undergo such an extent of nitration. An
3 . Thus, a modest nitration but sess atherosclerosis risk as well as to mon-
one that is responsive to the increase on alternative scenario is a gain-of-function
modification, in which case a small frac- itor the vasculoprotective action of drugs.
the absolute •NO and兾or O•⫺ 2 production Interestingly, circulating MPO levels may
rates can be generated by peroxynitrite in tion of nitrated protein can elicit a sub-
stantive biological signal. This attractive also serve as predictor of cardiovascular
vivo. As examples at the cell level, the role risk (76). Thus, enhanced •NO consump-
of endogenous peroxynitrite in protein concept has been shown in a few proteins
such as cytochrome c, which acquires a tion by either vascular wall兾inflammatory
nitration has been substantiated by thor-
strong peroxidase activity after nitration cell-derived O•⫺2 and兾or MPO provides a
ough studies on motoneurons (59) and
(66), in nitrated fibrinogen, which acceler- common link between decreases in •NO
smooth muscle cells (60) undergoing bio-
ates clot formation (67) and in protein bioavailability and increases in protein
logically relevant challenges.
In inflammatory cells, the hemeperoxi- kinase C␧, which becomes activated and tyrosine nitration.
dase nitration pathway (MPO, EPO) ap- translocates on nitration (68).
A notable example of loss of enzyme Mitochondrial Nitration and Its
pears to more important extra- than intra- Relevance to Cell Death
cellularly, because intracellular activity linked to nitration in vivo is the
peroxidases may have limited supply of mitochondrial enzyme MnSOD. This pro- Mitochondria have been recognized as
NO⫺ tein is nitrated by peroxynitrite in Tyr-34 critical sources and targets of nitrating
2 (61, 62). However, after polymor-
phonuclear cell degranulation, MPO can by a Mn-catalyzed process, which leads to species (reviewed in ref. 77). Although the
be taken up by nonphagocytic cells, and a enzyme inactivation. Nitrated and inacti- formation of •NO mainly occurs at extra-
strong codistribution of MPO and 3-nitro- vated MnSOD is found in acute and mitochondrial sites, the facile diffusion of
•NO to mitochondria and its combination
tyrosine is observed in different inflam- chronic inflammatory processes both in
matory processes (63). Excess •NO can animal models and human diseases (69).
also modulate the hemeperoxidase-medi- Interestingly, MnSOD nitration via the
储Some
ated nitration; indeed, •NO readily reacts hemeperoxidase兾•NO2 pathway does not metabolic conditions favor the role of NOS as a
significant source of peroxynitrite (74). NOS, in turn, can
with resting state MPO as well as with lead to significant inactivation (70), sug- be nitrated and potentially inactivated both in vitro and
compounds I and II, and therefore peroxi- gesting that nitration occurs at tyrosines in vivo.

4006 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0307446101 Radi


with mitochondrial-derived O•⫺ 2 results in respiratory chain uncouplers, among jury and nitration in neurotoxicity and
the formation of peroxynitrite, which ac- others. If the mechanism is peroxynitrite- vascular injury models (92, 93).** On the
counts for much of the disruption of mito- dependent, it should be inhibited SOD contrary, heterozygous MnSOD knock-
chondrial metabolism initially attributed (native, modified, or liposome-entrapped) outs resulted in larger extents of brain
to direct actions of •NO and to the nitra- or SOD-mimics, as long as the enzyme is nitration in a model of neurotoxicity, con-
tion of mitochondrial proteins. A cyto- located in the same compartment where sistent with enhanced mitochondrial for-
chrome c-dependent peroxidase-like O⫺ 2 is formed, and is unaffected by MPO mation and reactions of peroxynitrite (94).
mechanism of nitration may also operate inhibitors and catalase (60, 85). In cases MPO and EPO knockouts have evi-
(70). Mitochondrial proteins are nitrated where SOD is not close enough to O⫺ 2 denced the peroxidase-dependent nitra-
in vitro and in vivo, including MnSOD, sources, it can enhance peroxynitrite- tion pathway in models of inflammation
aconitase, cytochrome c, voltage-depen- dependent nitration (7, 8). On the other (15), but the relative contribution was
dent anion channel, ATPase, and succi- hand, hemeperoxidase-dependent path- highly model-dependent. The inability of
nyl-CoA oxoacid-CoA transferase (77, 78). ways can be stimulated by SOD, inhibited MPO from human neutrophils to cause
The nitration兾inactivation of MnSOD op- by catalase and MPO inhibitors, as ob- nitration in some models may result from
erates as a positive loop for enhanced in- served in H2O2-producing inflammatory lack of H2O2 substrate, because all O•⫺ 2
tramitochondrial peroxynitrite formation, cells in the presence of NO⫺ 2 (13). The can react with •NO to form peroxynitrite
which in turn triggers apoptotic signaling participation of LMW iron in the nitration (25) and兾or because the nitration capabili-
of cell death, in part by the thiol oxida- reactions can be assessed by metal chela-
tion-dependent assembly of the perme- ties of hemeperoxidases may vary accord-
tors, which block Fenton chemistry, al- ing to their access to NO⫺
ability transition pore (77). Detection of 2 and peroxyni-
though experiments in this line are scarce. trite (61, 62, 95). In this context, release
nitrated mitochondrial proteins and even Importantly, desferrioxamine can also di-
nitrated cytochrome c in the cytosol may and sequestration of neutrophil-derived
rectly trap •NO2 and CO•⫺ 3 and prevent MPO in specific tissue compartments dur-
serve to reflect extents of mitochondrial tyrosine nitration (86).
•NO-dependent oxidative stress (66, 77, ing inflammation are important aspects of
Uric acid inhibits nitration reactions MPO distribution and nitration capacity in
78). Importantly, the enhanced peroxi- and has been used as a scavenger of per-
datic activity of nitrated cytochrome c (66, vivo (63). However, results involving MPO
oxynitrite in vitro and in vivo (87). Al-
77) may further contribute to oxidative in mice should be cautiously extrapolated
though uric acid does not react at appre-
damage. to human inflammatory disease; indeed,
ciable rates with peroxynitrite (2), it
surprisingly, whereas MPO levels posi-
readily reacts with peroxynitrite-derived
Red Blood Cell Hemoglobin as a Sink of tively correlate with cardiovascular risk in
Nitrating Species radicals and oxo–metal complexes. Unfor-
humans (76), MPO knockout mice de-
tunately, the presence of uric acid has
Not all hemeproteins promote peroxyni- velop atherosclerosis more rapidly than
confounded the analysis of nitration yields
trite-dependent nitration under biologi- WT (96). MPO or EPO overexpresses are
by peroxynitrite in various reports when
cally relevant conditions, because some not available yet, but in the future they
xanthine兾xanthine oxidase has been used
reduce it by two electrons to NO⫺ 2 or may also provide clues to the role of
as a source of O•⫺ 2 and can be easily over-
isomerize it to NO⫺ 3 (79–82). In this con- MPO and EPO in pathology. Enhanced
text, oxyhemoglobin (oxyHb), present at 5 come with the use of alternative sub-
strates (88). LMW porphyrin complexes removal of H2O2 by stable transfection of
mM in red blood cells, serves as a sink to cells with catalase can also be useful to
capture intravascularly formed nitrating with Mn and Fe have SOD-mimic and
peroxynitrite-reductase activities and react assess the contribution of the hemeperoxi-
species. Indeed, a fraction of intravascu- dase pathway (60). The hemeperoxidase-
larly formed peroxynitrite could diffuse quickly with CO•⫺ 3 (89). They have been
successfully used to inhibit •NO-depen- independent nitration observed in MPO
into red blood cells before enacting target and EPO knockouts (15) may be due, in
molecule reactions with plasma compo- dent injury and nitration in cell and ani-
mal models of inflammation (90, 91). The addition to peroxynitrite, to the LMW
nents (83) and could undergo a fast reac- metal-dependent pathway (12). The recent
tion with oxyHb, which results in its redox and kinetic properties of these
SOD-mimics and peroxynitrite decompo- development of IRPs knockout mice (97)
isomerization to NO⫺ 3 , without significant
sition catalysts can be modulated by offers opportunities to study inflammatory
formation of nitrating species (82). Hemo-
changes in the porphyrin substituents, models and nitration in the context of
globin nitration is observed only under
excess peroxynitrite over hemoglobin. Be- which can also affect its cellular兾tissue altered iron homeostasis.
cause •NO and NO⫺ distribution, aspects that require further Circulating immune system cells (i.e.,
2 are also oxidized by
oxyHb to rather unreactive NO⫺ research. monocytes and neutrophils) from patients
3 , oxyHb
may ultimately serve to inhibit the intra- suffering from genetic deficiencies of
vascular formation of •NO2. A similar Lessons from Genetics: Knockouts, NADPH oxidase (chronic granulomatous
inhibitory role on the formation of Transgenics, and Human Deficiency disease) or MPO (MPO deficiency) have
nitrating species may be played by deoxy- Syndromes altered interactions of •NO with oxidants
hemoglobin, which reduces NO⫺ 2 to
Knockout and transgenic animals and ge- (25, 62). Future studies on cells and tis-
•NO (84) and, presumably, peroxynitrite netically engineered cells of the enzymes sues from these patients may offer impor-
to NO⫺ 2.
that participate in the formation or elimi- tant opportunities to understand the
nation of nitrating species are available. mechanisms of biological protein nitration
Lessons from Pharmacology The preferential nitration pathways under in humans.
Pharmacological experiments have been various pathologically relevant conditions
instrumental in defining the biological have been assessed by genetic approaches
mechanisms of nitration. Inhibition of that involve SODs, namely overexpression
NOS consistently leads to inhibition of of MnSOD or CuZnSOD or heterozygous **On the contrary, expression of amyotrophic lateral scle-
nitration, and the sources of O•⫺
2 兾H2O2 MnSOD knockouts and MPO and EPO rosis-associated mutants of CuZnSOD leads to increased
motoneuron nitration and apoptosis (59), possibly due to
can be discriminated with inhibitors of knockouts. MnSOD or CuZnSOD over- a toxic gain of function of SOD that promotes peroxyni-
NAD(P)H oxidases, xanthine oxidase, or expressors decreased •NO-dependent in- trite formation and nitration reactions.

Radi PNAS 兩 March 23, 2004 兩 vol. 101 兩 no. 12 兩 4007


Conclusion and Perspectives contribution of the peroxynitrite and matory, vascular, and neurodegenerative
The analysis presented herein shows that hemeperoxidase兾transition metal-depen- diseases.
biological protein tyrosine nitration†† dent pathways to biological nitration, the
detection of nitrated proteins with altered I thank Drs. Stanley L. Hazen, Harry Ischiro-
mainly occurs through free radical path- poulos, Bruce A. Freeman, Joe S. Beckman,
ways, and that there is an interplay involv- function in vivo, and the data arising from
Irwin Fridovich, Jonathan S. Stamler, Jean
ing excess •NO, oxidants, and transition pharmacological and genetic approaches Claude Drapier, and Ronald P. Mason for
metal centers. The individual or combined support the significance of nitration as a valuable comments, and my coinvestigators in
biochemical process linked to •NO-depen- Montevideo for contributions to the elucida-
dent pathophysiology. Approaches di- tion of various aspects presented in the
rected at inhibiting the oxidative modifica- manuscript. This work was supported
††Biological nitration is not restricted only to tyrosine nitra- by the Howard Hughes Medical Institute,
tions caused by reactive nitrogen species, Fogarty–National Institutes of Health, the
tion but may also result in nitration of tryptophan resi-
dues, as well as DNA bases, sugars, and lipids that can including protein tyrosine nitration, open John Simon Guggenheim Memorial Founda-
result in altered or new biological activities. new avenues for the treatment of inflam- tion, and Universidad de la República.

1. Ignarro, L. J. (1990) Annu. Rev. Pharmacol. Toxicol. 30, 535–560. 37. Go, Y. M., Patel, R. P., Maland, M. C., Park, H., Beckman, J. S., 69. MacMillan-Crow, L. A., Crow, J. P., Kerby, J. D., Beckman, J. S.
2. Radi, R., Denicola, A., Alvarez, B., Ferrer-Sueta, G. & Rubbo, Darley-Usmar, V. M. & Jo, H. (1999) Am. J. Physiol. 277, & Thompson, J. A. (1996) Proc. Natl. Acad. Sci. USA 93,
H. (2000) in Nitric Oxide, ed. Ignarro, L. (Academic, San Diego), H1647–H1653. 11853–11858.
pp. 57–82. 38. Mihm, M. J., Wattanapitayakul, S. K., Piao, S. F., Hoyt, D. G. & 70. Castro, L., Eiserich, J. P., Sweeney, S., Radi, R. & Freeman, B. A.
3. Liu, X., Miller, M. J., Joshi, M. S., Thomas, D. D. & Lancaster, Bauer, J. A. (2003) Biochem. Pharmacol. 65, 1189–1197. (2004) Arch. Biochem. Biophys. 421, 99–107.
J. R., Jr. (1998) Proc. Natl. Acad. Sci. USA 95, 2175–2179. 39. Cosentino, F., Eto, M., De Paolis, P., van der Loo, B., Bach- 71. Guo, W., Adachi, T., Matsui, R., Xu, S., Jiang, B., Zou, M. H.,
4. Denicola, A., Batthyany, C., Lissi, E., Freeman, B. A., Rubbo, H. schmid, M., Ullrich, V., Kouroedov, A., Delli Gatti, C., Joch, H., Kirber, M., Lieberthal, W. & Cohen, R. A. (2003) Am. J. Physiol.
& Radi, R. (2002) J. Biol. Chem. 277, 932–936. Volpe, M., et al. (2003) Circulation 107, 1017–1023. 285, H1396–H1403.
5. Sokolovsky, M., Riordan, J. F. & Vallee, B. L. (1966) Biochem- 40. Augusto, O., Bonini, M. G., Amanso, A. M., Linares, E., Santos,
72. Schmidt, P., Youhnovski, N., Daiber, A., Balan, A., Arsic, M.,
istry 5, 3582–3589. C. C. & De Menezes, S. L. (2002) Free Radical Biol. Med. 32,
Bachschmid, M., Przybylski, M. & Ullrich, V. (2003) J. Biol.
6. Beckman, J. S., Ischiropoulos, H., Zhu, L., van der Woerd, M., 841–859.
Chem. 278, 12813–12819.
Smith, C., Chen, J., Harrison, J., Martin, J. C. & Tsai, M. (1992) 41. Hazen, S. L., Gaut, J. P., Hsu, F. F., Crowley, J. R., d’Avignon,
A. & Heinecke, J. W. (1997) J. Biol. Chem. 272, 16990–16998. 73. Bachschmid, M., Thurau, S., Zou, M. H. & Ullrich, V. (2003)
Arch. Biochem. Biophys. 298, 438–445. FASEB J. 17, 914–916.
7. Ischiropoulos, H., Zhu, L., Chen, J., Tsai, M., Martin, J. C., 42. Fridovich, I. (1997) J. Biol. Chem. 272, 18515–18517.
43. Yamazaki, I. & Piette, L. H. (1990) J. Biol. Chem. 265, 13589– 74. Stuehr, D., Pou, S. & Rosen, G. M. (2001) J. Biol. Chem. 276,
Smith, C. D. & Beckman, J. S. (1992) Arch. Biochem. Biophys. 14533–14536.
13594.
298, 431–437. 75. Shishehbor, M. H., Aviles, R. J., Brennan, M. L., Fu, X.,
44. Burner, U., Furtmuller, P. G., Kettle, A. J., Koppenol, W. H. &
8. Ischiropoulos, H., Zhu, L. & Beckman, J. S. (1992) Arch.
Obinger, C. (2000) J. Biol. Chem. 275, 20597–20601. Goormastic, M., Pearce, G. L., Gokce, N., Keaney, J. F., Jr.,
Biochem. Biophys. 298, 446–451.
45. Sampson, J. B., Rosen, H. & Beckman, J. S. (1996) Methods Penn, M. S., Sprecher, D. L., et al. (2003) J. Am. Med. Assoc. 289,
9. Beckmann, J. S., Ye, Y. Z., Anderson, P. G., Chen, J., Accavitti,
Enzymol. 269, 210–218. 1675–1680.
M. A., Tarpey, M. M. & White, C. R. (1994) Biol. Chem.
46. Ferrer-Sueta, G., Quijano, C., Alvarez, B. & Radi, R. (2002) 76. Brennan, M. L., Penn, M. S., Van Lente, F., Nambi, V.,
Hoppe–Seyler 375, 81–88.
Methods Enzymol. 349, 23–37. Shishehbor, M. H., Aviles, R. J., Goormastic, M., Pepoy, M. L.,
10. Ohshima, H., Friesen, M., Brouet, I. & Bartsch, H. (1990) Food 47. Bouton, C. & Drapier, J. C. (2003) Sci. STKE 182, pe17. McErlean, E. S., Topol, E. J., et al. (2003) N. Engl. J. Med. 349,
Chem. Toxicol. 28, 647–652. 48. Soum, E., Brazzolotto, X., Goussias, C., Bouton, C., Moulis,
11. Pfeiffer, S., Schmidt, K. & Mayer, B. (2000) J. Biol. Chem. 275, 1595–1604.
J. M., Mattioli, T. A. & Drapier, J. C. (2003) Biochemistry 42, 77. Radi, R., Cassina, A., Hodara, R., Quijano, C. & Castro, L.
6346–6352. 7648–7654.
12. Thomas, D. D., Espey, M. G., Vitek, M. P., Miranda, K. M. & (2002) Free Radical Biol. Med. 33, 1451–1464.
49. Gunther, M. R., Hsi, L. C., Curtis, J. F., Gierse, J. K., Marnett, 78. Turko, I. V., Li, L., Aulak, K. S., Stuehr, D. J., Chang, J. Y. &
Wink, D. A. (2002) Proc. Natl. Acad. Sci. USA 99, 12691–12696. L. J., Eling, T. E. & Mason, R. P. (1997) J. Biol. Chem. 272,
13. Eiserich, J. P., Hristova, M., Cross, C. E., Jones, A. D., Freeman, Murad, F. (2003) J. Biol. Chem. 278, 33972–33977.
17086–17090.
B. A., Halliwell, B. & van der Vliet, A. (1998) Nature 391, 79. Pearce, L. L., Pitt, B. R. & Peterson, J. (1999) J. Biol. Chem. 274,
50. Esteves, P. M., De M Carneiro, J. W., Cardoso, S. P., Barbosa,
393–397. 35763–35767.
A. G., Laali, K. K., Rasul, G., Prakash, G. K. & Olah, G. A.
14. Bian, K., Gao, Z., Weisbrodt, N. & Murad, F. (2003) Proc. Natl. (2003) J. Am. Chem. Soc. 125, 4836–4849. 80. Pietraforte, D., Salzano, A. M., Scorza, G., Marino, G. &
Acad. Sci. USA 100, 5712–5717. 51. Quijano, C., Hernandez-Saavedra, D., Castro, L., McCord, J. M., Minetti, M. (2001) Biochemistry 40, 15300–15309.
15. Brennan, M. L., Wu, W., Fu, X., Shen, Z., Song, W., Frost, H., Freeman, B. A. & Radi, R. (2001) J. Biol. Chem. 276, 11631– 81. Herold, S., Shivashankar, K. & Mehl, M. (2002) Biochemistry 41,
Vadseth, C., Narine, L., Lenkiewicz, E., Borchers, M. T., et al. 11638. 13460–13472.
(2002) J. Biol. Chem. 277, 17415–17427. 52. Whiteman, M., Siau, J. L. & Halliwell, B. (2003) J. Biol. Chem. 82. Romero, N., Radi, R., Linares, E., Augusto, O., Detweiler, C. D.,
16. Fukuto, J. & Ignarro, L. (1997) Acc. Chem. Res. 30, 149–152. 278, 8380–8384. Mason, R. P. & Denicola, A. (2003) J. Biol. Chem. 278, 44049–
17. Eu, J. P., Liu, L., Zeng, M. & Stamler, J. S. (2000) Biochemistry 53. Aslan, M., Ryan, T. M., Townes, T. M., Coward, L., Kirk, M. C., 44057.
39, 1040–1047. Barnes, S., Alexander, C. B., Rosenfeld, S. S. & Freeman, B. A. 83. Romero, N., Denicola, A., Souza, J. M. & Radi, R. (1999) Arch.
18. Denicola, A., Souza, J. M. & Radi, R. (1998) Proc. Natl. Acad. (2003) J. Biol. Chem. 278, 4194–4204. Biochem. Biophys. 368, 23–30.
Sci. USA 95, 3566–3571. 54 Grune, T., Blasig, I. E., Sitte, N., Roloff, B., Haseloff, R. & 84. Cosby, K., Partovi, K. S., Crawford, J. H., Patel, R. P., Reiter,
19. Radi, R., Peluffo, G., Alvarez, M. N., Naviliat, M. & Cayota, A. Davies, K. J. (1998) J. Biol. Chem. 273, 10857–10862. C. D., Martyr, S., Yang, B. K., Waclawiw, M. A., Zalos, G., Xu,
(2001) Free Radical Biol. Med. 30, 463–488. 55 Irie, Y., Saeki, M., Kamisaki, Y., Martin, E. & Murad, F. (2003) X., et al. (2003) Nat. Med. 9, 1498–1505.
20. Gryglewski, R. J., Palmer, R. M. & Moncada, S. (1986) Nature Proc. Natl. Acad. Sci. USA 100, 5634–5639. 85. Estevez, A. G., Sampson, J. B., Zhuang, Y., Spear, N., Richard-
320, 454–456. 56. Rubbo, H., Radi, R., Trujillo, M., Telleri, R., Kalyanaraman, B., son, G. J., Crow, J. P., Tarpey, M. M., Barbeito, L. & Beckman,
21. Rubanyi, G. M. & Vanhoutte, P. M. (1986) Am. J. Physiol. 250, Barnes, S., Kirk, M. & Freeman, B. A. (1994) J. Biol. Chem. 269, J. S. (2000) Free Radical Biol. Med. 28, 437–446.
H822–H827. 26066–26075. 86. Bartesaghi, S., Trujillo, M., Denicola, A., Folkes, L., Wardman,
22. Ignarro, L. J., Byrns, R. E., Buga, G. M., Wood, K. S. & 57. Miles, A. M., Bohle, D. S., Glassbrenner, P. A., Hansert, B.,
P. & Radi, R. (2004) Free Radical Biol. Med. 36, 471–483.
Chaudhuri, G. (1988) J. Pharmacol. Exp. Ther. 244, 181–189. Wink, D. A. & Grisham, M. B. (1996) J. Biol. Chem. 271, 40–47.
87. Scott, G. S. & Hooper, D. C. (2001) Med. Hypotheses 56, 95–100.
23. Liochev, S. I. & Fridovich, I. (2002) Free Radical Biol. Med. 33, 58. Goldstein, S., Czapski, G., Lind, J. & Merenyi, G. (2000) J. Biol.
Chem. 275, 3031–3036. 88. Trujillo, M., Alvarez, M. N., Peluffo, G., Freeman, B. A. & Radi,
137–141. R. (1998) J. Biol. Chem. 273, 7828–7834.
24. Buerk, D. G., Lamkin-Kennard, K. & Jaron, D. (2003) Free 59. Estevez, A. G., Crow, J. P., Sampson, J. B., Reiter, C., Zhuang,
Y., Richardson, G. J., Tarpey, M. M., Barbeito, L. & Beckman, 89. Ferrer-Sueta, G., Vitturi, D., Batinic-Haberle, I., Fridovich, I.,
Radical Biol. Med. 34, 1488–1503. Goldstein, S., Czapski, G. & Radi, R. (2003) J. Biol. Chem. 278,
25. Clark, S. R., Coffey, M. J., Maclean, R. M., Collins, P. W., Lewis, J. S. (1999) Science 286, 2498–2500.
60. Fries, D. M., Paxinou, E., Themistocleous, M., Swanberg, E., 27432–27438.
M. J., Cross, A. R. & O’Donnell, V. B. (2002) J. Immunol. 169,
Griendling, K. K., Salvemini, D., Slot, J. W., Heijnen, H. F., 90. Muscoli, C., Cuzzocrea, S., Riley, D. P., Zweier, J. L., Thiemer-
5889–5896.
Hazen, S. L. & Ischiropoulos, H. (2003) J. Biol. Chem. 278, mann, C., Wang, Z. Q. & Salvemini, D. (2003) Br. J. Pharmacol.
26 Murphy, M. E. & Sies, H. (1991) Proc. Natl. Acad. Sci. USA 88,
22901–22907. 140, 445–460.
10860–10864.
61. Jiang, Q. & Hurst, J. K. (1997) J. Biol. Chem. 272, 32767–32772. 91. Pacher, P., Liaudet, L., Bai, P., Mabley, J. G., Kaminski, P. M.,
27 Jourd’heuil D., Laroux, F. S., Miles, A. M., Wink, D. A. &
62. Rosen, H., Crowley, J. R. & Heinecke, J. W. (2002) J. Biol. Chem. Virag, L., Deb, A., Szabo, E., Ungvari, Z., Wolin, M. S., Groves,
Grisham, M. B. (1999) Arch. Biochem. Biophys. 361, 323–330.
277, 30463–30468. J. T. & Szabo, C. (2003) Circulation 107, 896–904.
28. Hobbs, A. J., Fukuto, J. M. & Ignarro, L. (1994) Proc. Natl. Acad.
63. Baldus, S., Eiserich, J. P., Brennan, M. L., Jackson, R. M., 92. Keller, J. N., Kindy, M. S., Holtsberg, F. W., St Clair, D. K., Yen,
Sci. USA 91, 10992–10996. Alexander, C. B. & Freeman, B. A. (2002) Free Radical Biol.
29. Sowers, J. R. (2002) N. Engl. J. Med. 346, 1999–2001. H. C., Germeyer, A., Steiner, S. M., Bruce-Keller, A. J.,
Med. 33, 1010. Hutchins, J. B. & Mattson, M. P. (1998) J. Neurosci. 18, 687–697.
30. Gutierrez, H. H., Nieves, B., Chumley, P., Rivera, A. & Freeman, 64. Hazen, S. L., Zhang, R., Shen, Z., Wu, W., Podrez, E. A.,
B. A. (1996) Free Radical Biol. Med. 21, 43–52. 93. Wang, H. D., Johns, D. G., Xu, S. & Cohen, R. A. (2002) Am. J.
MacPherson, J. C., Schmitt, D., Mitra, S. N., Mukhopadhyay, C., Physiol. 282, H1697–H1702.
31. Sarti, P., Avigliano, L., Gorlach, A. & Brune, B. (2002) Cell Chen, Y., et al. (1999) Circ. Res. 85, 950–958.
Death Differ. 9, 1160–1162. 94. Kim, G. W. & Chan, P. H. J (2002) Cereb. Blood Flow Metab. 22,
65. Souza, J. M., Daikhin, E., Yudkoff, M., Raman, C. S. &
32. Radi, R., Beckman, J. S., Bush, K. M. & Freeman, B. A. (1991) 798–809.
Ischiropoulos, H. (1999) Arch. Biochem. Biophys. 371, 169–178.
J. Biol. Chem. 266, 4244–4250. 95. Gaut, J. P., Byun, J., Tran, H. D., Lauber, W. M., Carroll, J. A.,
66. Cassina, A., Hodara, R., Souza, J., Thomson, L., Castro, L.,
33. Trujillo, M. & Radi, R. (2002) Arch. Biochem. Biophys. 397, Ischiropoulos, H., Freeman, B. A. & Radi, R. (2000) J. Biol. Hotchkiss, R. S., Belaaouaj, A. & Heinecke, J. W. (2002) J. Clin.
91–98. Chem. 275, 21409–21415. Invest. 109, 1311–1319.
34. Arteel, G. E., Briviba, K. & Sies, H. (1999) FEBS Lett. 445, 67. Vadseth, C., Souza, J. M., Thomson, L., Seagraves, A., Na- 96. Brennan, M. L. Anderson, M. M., Shih, D. M., Qu, X. D., Wang,
226–230. gaswami, C., Scheiner, T., Torbet, J., Vilaire, G., Bennett, J. S., X., Mehta, A. C., Lim, L. L., Shi, W., Hazen, S. L., Jacob, J. S.,
35. Bryk, R., Griffin, P. & Nathan, C. (2000) Nature 407, 211–215. Murciano, J. C., et al. (2003) J. Biol. Chem., jbc兾M306101200v1. et al. (2001) J. Clin. Invest. 107, 419–430.
36. Brito, C., Naviliat, M., Tiscornia, A. C., Vuillier, F., Gualco, G., 68. Balafanova, Z., Bolli, R., Zhang, J., Zheng, Y., Pass, J. M., 97. LaVaute, T., Smith, S., Cooperman, S., Iwai, K., Land, W.,
Dighiero, G., Radi, R. & Cayota, A. M. (1999) J. Immunol. 162, Bhatnagar, A., Tang, X.-L., Wang, O., Cardwell, E. & Ping, P. Meyron-Holtz, E., Drake, S. K., Miller, G., Abu-Asab, M.,
3356–3366. (2002) J. Biol. Chem. 277, 15021–15027. Tsokos, M., et al. (2001) Nat. Genet. 27, 209–214.

4008 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0307446101 Radi

Vous aimerez peut-être aussi