Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Mar 5.
Published in final edited form as: Circ Res. 2010 Mar 5;106(4):633–646. doi: 10.1161/CIRCRESAHA.109.207381

S-Nitrosylation in Cardiovascular Signaling

Brian Lima a, Michael T Forrester b, Douglas T Hess c,d, Jonathan S Stamler b,c,d,*
PMCID: PMC2891248  NIHMSID: NIHMS166939  PMID: 20203313

Abstract

Well over two decades has passed since the endothelium-derived relaxing factor (EDRF) was reported to be the gaseous molecule nitric oxide (NO). Although soluble guanylyl cyclase (which generates cyclic guanosine monophosphate, cGMP) was the first identified receptor for NO, it has become increasingly clear that NO exerts a ubiquitous influence in a cGMP-independent manner. In particular, many if not most effects of NO are mediated by S-nitrosylation, the covalent modification of a protein cysteine thiol by an NO group to generate an S-nitrosothiol (SNO). Moreover, within the current framework of NO biology, EDRF activity—i.e. G-protein coupled receptor-mediated, or shear-induced endothelium-derived NO bioactivity—is understood to involve a central role for SNOs, acting both as second messengers and signal effectors. Further, essential roles for S-nitrosylation have been implicated in virtually all major functions of NO in the cardiovascular system. Here we review the basic biochemistry of S-nitrosylation (and denitrosylation), discuss the role of S-nitrosylation in the vascular and cardiac functions of NO and identify current and potential clinical applications.

Keywords: Nitric oxide, S-nitrosylation, denitrosylation, cysteine, redox, angiogenesis, apoptosis, inflammation, atherogenesis, electrophysiology, excitation-contraction coupling, heart failure, β-adrenergic receptor, hemoglobin

NO as the Third Respiratory Gas

Molecular oxygen (O2) and carbon dioxide (CO2) are critical components of cardiovascular physiology (and pathophysiology). These gaseous molecules are central to tissue physiology and cellular respiration, and it has long been understood that disturbances in O2 or CO2 processing are both causative and indicative of pathophysiology1. Over time, however, it has become increasingly clear that nitric oxide (NO) is also an endogenous regulator in cardiovascular physiology and cellular respiration, operating at considerably lower concentrations than O2 or CO2. These observations have lead to the proposal that NO is the “third gas” of the respiratory cycle24.

The major sources of NO in vivo are the NO synthase (NOS) isoforms. These include predominantly the neuronal (nNOS/NOS1), inducible (iNOS/NOS2) and endothelial (eNOS/NOS3) enzymes. It is worth noting that this naming system is primarily of historical significance; NOS tissue expression is far less strict than implied by this nomenclature, and all three isoforms may be constitutive or inducible. NOS’s are heme- and flavin-containing enzymes that employ NAPDH, tetrahydrobiopterin and O2 to convert L-arginine to L-citrulline with concomitant release of NO 5.

NO-Based Signaling: The Roles of cGMP and S-Nitrosylation

One of the earliest described intracellular receptors for NO is the soluble guanylyl cyclase (sGC)6, 7. Binding of NO to the heme group of sGC leads to increased conversion of GTP to cGMP, which in turn activates protein kinase G (PKG). Despite the wealth of studies that have focused on sGC, it has become increasingly clear that NO exerts most of its cellular influence in a cGMP-independent manner. More generally, it is now appreciated that hemes in proteins do not generally mediate NO-based signaling that involves posttranslational protein modification, but rather serve to promote the requisite redox chemistry of NO. These observations led to the exploration of alternative molecular mechanisms through which NO might regulate cellular function, which culminated in the discovery of NO-mediated modification of protein cysteine (Cys) residues (to generate an S-nitrosothiol, SNO), designated S-nitrosylation (Figure 1A).

Figure 1.

Figure 1

The roles of cGMP and S-nitrosylation in NO-based signaling (A) and enzymatic protein denitrosylation mediated by the S-nitrosoglutathione reductase (GSNOR) and thioredoxin (Trx) systems (B). (A) NO synthase (NOS) synthesizes NO, which may activate soluble guanylyl cyclase and thereby enhance production of cGMP (left) or subserve protein S-nitrosylation (right). The cGMP-dependent pathway is deactivated by cGMP-phosphodiesterase (PDE), which hydrolyzes cGMP to GMP (PDE may also be activated allosterically by cGMP). The SNO-based mechanisms are dynamically regulated via S-nitrosylation and denitrosylation of a multitude of cysteine-containing proteins. In contrast to the multiple elements regulated by S-nitrosylation, the cGMP-based signaling system relies primarily on the cGMP-dependent protein kinase, PKG. (B) Proteins undergo reversible S-nitrosylation and denitrosylation (center). Denitrosylation mediated by GSNOR is depicted on the left. Transnitrosylation of glutathione (GSH) by S-nitrosylated proteins generates GSNO and native protein. GSNO undergoes NADH-dependent reduction by GSNOR to generate glutathione S-hydroxysulfenamide (GSNHOH), which can undergo further reaction with GSH to generate oxidized glutathione (GSSG). The redox cycle is completed by reduction of GSSG to GSH via GSSG reductase. Denitrosylation mediated by the thioredoxin (Trx) system is depicted on the right. The active site dithiol motif (CXXC) of Trx1 (cytoplasmic) or Trx2 (mitochondrial) undergoes oxidation coupled to denitrosylation of SNO substrate. Oxidized Trx is reduced by the selenoprotein thioredoxin reductase (TrxR), which employs the reducing power of NADPH to regenerate active Trx.

Cys is a unique amino acid due to its thiol side chain. This functional group is nucleophilic, acidic (pKa ~ 8) and redox active due to its hybridized p- and d-orbitals, which together underlie the large range of reactivities for Cys residues within proteins. Within the realm of redox chemistry (i.e., transfer of electrons and consequent change in atomic oxidation state), numerous reactions are known to occur on Cys thiol side chains that affect protein structure and function. Of particular physiological significance is the redox reaction between NO and a Cys thiol leading to S-nitrosylation (forming a protein SNO). In contrast to the cGMP axis that employs a single principal molecular effector (i.e., PKG) to carry out the downstream functions of NO, S-nitrosylation allows for a wide range of NO-mediated functions, inasmuch as a plethora of proteins may undergo this modification 8, 9. S-nitrosylation therefore helps to explain the wide range of cellular effects of NO in the cardiovascular system, some of which are listed in Table 1.

Table 1.

Exemplary SNO-proteins in the Cardiovascular System

SNO-protein Cell/Tissue Effects of S-nitrosylation
Albumin Serum NO bioactivity reservoir104
β-arrestin 2 Endothelium Enhanced binding of β-Arr 2 to clathrin
and internalization of β-adrenergic
receptor28
Caspase 3 Endothelium,
lymphocytes
Anti-apoptosis and preservation of
endothelial function6467, 163
Dimethylarginine
dimethylaminohydrolase
Endothelium Accumulation of dimethylarginine and
NOS inhibition16
G-protein-coupled
receptor kinase 2
Endothelium,
Myocardium
Attenuation of β-adrenergic receptor
desensitization29
Hemoglobin Erythrocyte Hypoxic vasodilation and regulation of
vessel tone4, 17
Hypoxia-inducible factor
1 α
Myocardium Increased VEGF production and
myocardial capillary density31
MAP kinase
phosphatase 7
Endothelium Promotes endothelial cell migration and
angiogenesis54
N-ethylmaleimide-
sensitive factor
Platelets Prevention of platelet activation73, 74
Ryanodine receptor 2 Cardiac
muscle
Enhanced cardiac Ca2+ release and
contractility136, 137
Tissue
transglutaminase
Endothelial
surface
Inhibition of platelet aggregation164
Metallothionein Vascular smooth muscle Myogenic reflex, pulmonary vasoconstriction170,171

Examples of S-nitrosylated protein of interest, and the general location and overall effect of S-nitrosylation. See Table 2 for dysregulated SNO-proteins in cardiovascular disease. MPK7 indicates mitogen-activated protein kinase phosphatase 7.

In addition, the ongoing delineation of cellular SNO-proteins has revealed multiple loci through which S-nitrosylation might influence levels of cGMP. It has been reported that S-nitrosylation inhibits sGC 10 and cGMP phosphodiesterase 11, as well as eNOS itself 12 and eNOS-regulating proteins including HSP9013 and Akt/protein kinase B 14. PKG has regulatory thiols as well, which may be susceptible to S-nitrosylation. Further, activating S-nitrosylation of arginase 15 and inhibitory S-nitrosylation of dimethylarginine dimethylaminohydrolase 16 would decrease NOS substrate levels and increase levels of endogenous, methylarginine NOS inhibitors, respectively.

Protein Denitrosylation: A Critical Regulator of SNO Biology

Numerous studies have focused on the mechanistic aspects of S-nitrosylation, leading to the identification of proteins that may either catalyze S-nitrosylation (e.g. hemoglobin)17, or participate in protein trans-nitrosylation (i.e., NO group transfer between proteins)18. More recently, however, protein denitrosylation has been shown to play a major role in controlling cellular S-nitrosylation1921 (similar to the role of phosphatases in protein phosphorylation), and has been shown to operate on hundreds of proteins in intact cells21, 22. To date, two major enzymatic systems mediating denitrosylation have been described (Figure 1B), and are discussed in greater detail below.

A number of enzymes have been reported to catalyze the reduction of SNOs, and thus may be viewed as candidate denitrosylases. One of the first described is known as S-nitrosoglutathione (GSNO) reductase (GSNOR)23, 24. This enzyme employs the reducing power of NADH to convert GSNO to glutathione S-hydroxysulfenamide (GSNHOH), which in turn is converted to oxidized glutathione (GSSG); reduction of GSSG by glutathione reductase completes the denitrosylation cycle (GSSG-reductase activity is therefore required for physiological denitrosylation of GSNO25). Although GSNOR acts only on GSNO, i.e. SNO-proteins are not substrates, it governs protein S-nitrosylation by influencing the cellular equilibrium between SNO-proteins and GSNO26, 27 (Figure 1B). Importantly, GSNOR has been shown to play a role in regulating SNO signaling downstream of the β-adrenergic receptor28, 29, and is therefore operative in cellular signal transduction (discussed further below). Pharmacological inhibition or genetic deletion of GSNOR leads to enhanced vasodilation19, 30, 31, consistent with a role for GSNO in conveying the systemic activity of NO derived from eNOS.

GSNOR is an atypical member of the alcohol dehydrogenase family, inasmuch as it has no known alcohol-based substrate. In methylotropic bacteria, GSNOR also metabolizes formaldehyde. A recent report indicates that another NADPH-oxidoreductase (carbonyl reductase 1) possesses GSNOR activity32. In addition, xanthine oxidase metabolizes GSNO, but the Km is high and its physiological relevance is therefore not clear33. Nonetheless, these studies, taken together, raise the idea that multiple enzymes may modulate GSNO levels in vivo.

A new line of investigation has recently revealed that the ubiquitous thioredoxin enzyme family—originally described as protein disulfide reductases34, 35—are also bona fide intracellular denitrosylases36, 37. In contrast to the strict substrate specificity of GSNOR for GSNO, a small-molecular-weight SNO, the cytoplasmic and mitochondrial thioredoxins (Trx1 and Trx2, respectively) directly mediate the denitrosylation of multiple substrate SNO-proteins. As illustrated in Figure 1B, the Trx system employs a thioredoxin reductase (TrxR) and NADPH to regenerate reduced Trx following denitrosylation. Recent examples demonstrate that, in the context of signal transduction, denitrosylation by Trx/TrxR can be stimulus-coupled, substrate specific and spatially restricted (compartmentalized) 20, 21, 30.

Accumulating evidence indicates that protein S-nitrosylation status in vivo is not determined simply by rates of NO synthesis (i.e. NOS activities), but rather involves a precisely regulated equilibrium between S-nitrosylation and denitrosylation pathways, in particular involving transnitrosylation reactions between a variety of peptides and proteins, and that consequently protein denitrosylation is critical in SNO-based signal transduction8,21. Enzymatic control of both S-nitrosylation and denitrosylation, established by stringent genetic criteria, underlies the spatiotemporal specificity necessary for cellular signaling. In addition, elucidation of the mechanisms of denitrosylation may provide novel genetic and pharmacological tools for manipulating SNO-based signaling in vivo (e.g., as revealed in studies of GSNOR−/− mice, discussed further below) and help identify potential targets for therapeutic intervention in dysregulated SNO processing in the cardiovascular system38.

Roles of S-Nitrosylation in Vascular Signaling

EDRF and systemic vascular resistance

NO derived constitutively from eNOS (which mediates endothelium-dependent relaxation) is thought to account for the increase in blood pressure that is produced by both NOS inhibitors and genetic deletion of eNOS. It is worth considering in this light the implications of the findings that inhibitors of GSNOR elicit vasodilation30 and deletion of GSNOR results in lowering of systemic vascular resistance 19, 31; GSNOR null mice are in fact highly susceptible to hypotension19. Thus, to the extent that peripheral vasodilation by eNOS is identified with EDRF, analysis in GSNOR mutant mice indicates that GSNO is a major effector of EDRF action.

Vasodilation by EDRF may be mediated by cGMP or may be cGMP-independent, depending on the animal species and vessel type. In the classic Furchgott bioassay of rabbit thoracic aorta, the EDRF response is equally dependent on cGMP-and non-cGMP-regulated pathways39, 40. Further, increases in cGMP in and of themselves provide little insight into the nature of the NO-based effector, because both NO and GSNO can increase cGMP levels. In addition, cGMP elevations may either result not only from NO binding to heme in sGC but also from inhibitory S-nitrosylation of phosphodiesterase 511. GSNO-based EDRF activity would be fully consistent with these data. GSNO is in equilibrium with protein SNO, and it has recently been reported that shear-induced activation of endothelial cells is associated with S-nitrosylation of over one hundred proteins41. Furchgott’s EDRF was not generated by shear but rather by acetylcholine, a G protein-coupled receptor (GPCR) that activates both activates eNOS and releases NO from SNO-protein reservoirs 172. As discussed in more detail below, it has been shown recently that GPCR-mediated vasodilation via a different, eNOS-coupled GPCR, the β2-adrenergic receptor, is regulated critically by and very likely dependent in large part upon GSNO-mediated S-nitrosylation of a set of proteins that includes G protein-coupled receptor kinase GRK228, 29, 42. Inhibition of GRK2 by S-nitrosylation prevents receptor densensitization. Thus, the available evidence is consistent with a prominent role for SNO in conveying NO-based vasodilatory signals4345.

Substantial evidence supports a model in which endothelial dysfunction contributes to evolution of vascular disease, including hypertension, diabetes and atherosclerosis. A signature feature of this model is the excess production of superoxide, which eliminates NO43. Endothelial sources of superoxide include NADPH oxidases and eNOS itself, through a process known as “uncoupling” that results from co-factor or substrate deficiency. It is therefore of interest that S-nitrosylation of NADPH oxidase may preserve NO bioavailabilty in healthy endothelium by inhibiting production of superoxide44. By contrast, excessive S-nitrosylation of arginase has been implicated in the uncoupling eNOS that is characteristic of arteriosclerotic vessels45. It has been recently suggested that impaired S-nitrosylation of endothelial clock-related proteins may be linked to hypertension46, emphasizing the emerging theme that hypo- and hyper-nitrosylation of specific protein targets, rather than a general NO/superoxide imbalance that may be interpreted principally in terms of altered levels of NO, may correlate best with pathophysiology.

Angiogenesis

Endogenously synthesized NO is an established facilitator of endothelial function and survival: eNOS is both induced and activated following endothelial stimulation with vascular endothelial growth factor (VEGF)47, 48, a major promoter of vessel growth in vivo, as well as by shear stress49. Importantly, eNOS−/− mice are deficient in VEGF responsiveness, thus establishing that NO is indeed a critical element in angiogenesis50. It has also been demonstrated that both shear stress41, 51 and VEGF52, 53 regulate protein S-nitrosylation in the vascular endothelium. In a recent specific example, VEGF production downstream of the chemokine-type GPCR CXCR4 was shown to be intimately coupled to S-nitrosylation of MAP kinase phosphatase 7 (MKP7), which facilitates endothelial cell migration via activation of c-Jun N-terminal kinase 3 (JNK3), thus promoting angiogenesis54. Promotion of endothelial cell survival and angiogenesis also appears to be mediated via S-nitrosylation and activation of dynamin55, a regulator of endothelial cell endocytosis. Finally, endothelial S-nitrosylation is perturbed by known pathophysiological stimuli including aging56 and hyperglycemic states57, clearly linking defective S-nitrosylation to vascular disease.

It is well recognized that hypoxia stimulates angiogenesis primarily via the transcription factor hypoxia-inducible factor (HIF), which augments VEGF expression58. Under normoxic conditions, HIF is typically undetectable due to rapid proteolytic degradation that is initiated by prolyl hyroxylation. Interestingly, exogenously administered SNO donors exert a hypoxia-mimetic effect59, 60, leading to nuclear accumulation of HIF. HIF stablilization by SNO under conditions of normoxia, observed both in vitro61, 62 and in vivo31, is mediated by S-nitrosylation of HIF itself. Specifically, HIF is constitutively S-nitrosylated in normoxic GSNOR−/− mice, with increased binding of S-nitrosylated HIF to the gene for VEGF31. These mice also exhibited increased myocardial capillary density, lending further support for an integral role of S-nitrosylation in promoting angiogenesis.

Apoptosis

Some of the earliest studies examining the functions of S-nitrosylation focused on the anti-apoptotic/protective effects of endogenous NO63. These efforts demonstrated that NO S-nitrosylates and inhibits the active site Cys residue of the pro-apoptotic effector caspase-36466. It was shown subsequently that caspase-3 undergoes stimulus-coupled activation, driven by proapoptotic Fas stimulation, via thioredoxin-mediated denitrosylation20. Importantly, this mechanism has been shown to operate in endothelial cells67, suggesting that the S-nitrosylation/denitrosylation equilibrium of caspase-3 may be a critical determinant of endothelial cell survival and vessel function. Furthermore, the oxidoreductase function of thioredoxin, a vital element in preserving endothelial redox equilibrium and protecting against the deleterious effects of oxidative and/or nitrosative stress, is itself stimulated by S-nitrosylation68.

Inflammation

The robust anti-inflammatory attributes of NO were first appreciated in experimental observations of diminished leukocyte adherence to vascular endothelium in the presence of exogenous NO donors69. Administration of NOS inhibitors predictably results in increased leukocyte rolling along the endothelium. Studies in knockout mice lacking a specific NOS isoform also underscore the contribution of endogenous NO sources in mitigating leukocyte adherence: compared to wild-type, eNOS−/−, nNOS−/− and iNOS−/− mice exhibit increased leukocyte adherence to endothelium70, 71.

The molecular bases of these findings have been elucidated in part, and encompass two key areas of SNO-mediated regulation: control of endothelial protein trafficking and suppression of nuclear factor-κB (NF-κB)-dependent expression of pro-inflammatory cytokines and adhesion molecules72. During the initial phase of an inflammatory response, leukocyte rolling requires interactions between P-selectins on the endothelial cell surface with the cognate P-selectin glycoprotein ligand-1 (PSGL-1) on the leukocyte surface. P-selectins are transmembrane proteins that reside within resting endothelial cells in granules designated Weibel-Palade bodies (WPB). Upon endothelial cell activation by an inflammatory stimulus, these granules translocate to the cell surface, resulting in exposure of P-selectin to the vessel lumen. N-ethylmaleimide-sensitive factor (NSF), a principal component of this exocytic trafficking machinery, is subject to direct inhibition by S-nitrosylation of critical Cys residues73. The resultant interruption of NSF-mediated disassembly of soluble NSF-attachment protein receptor (SNARE) complexes prevents WPB exocytosis from endothelial cells. Thus, S-nitrosylation of NSF, consequent, for example, upon stimulation with the GPCR agonist thrombin, is identified with the anti-inflammatory activity of eNOS (Figure 2). Similarly, inhibitory S-nitrosylation of NSF in platelets is anti-thrombotic through a similar mechanism74 (see below).

Figure 2.

Figure 2

S-nitrosylation of NSF is anti-inflammatory and anti-thrombotic73, 74. In endothelial cells, inhibitory S-nitrosylation of NSF suppresses exocytosis of Weibel-Palade bodies and thereby externalization of P-selectin, which inhibits leukocyte rolling and thus vascular inflammation. Similarly, in platelets (labeling in parentheses), inhibitory S-nitrosylation of NSF suppresses exocytosis of secretory granules and thereby externalization of P-selectin (and other adhesive molecules), which reduce platelet activation, adhesion, aggregation and rolling on the endothelium. These effects are anti-thrombotic.

Other phases of the inflammatory response and leukocyte trafficking are impacted by S-nitrosylation. Specifically, NO has been shown to limit the expression of integrins and intracellular adhesion molecules (ICAMs) required for leukocyte adherence7577. These and other pro-inflammatory effectors, including cytokines and cytokine receptors, are under direct transcriptional control by NF-κB78. Inhibitory S-nitrosylation of both NK-κB79 and its upstream activating enzyme complex, inhibitory κB kinase (IKK)80, has been demonstrated in multiple studies. Taken together, these demonstrations of multiple loci of S-nitrosylation in the inflammatory signaling cascade support a comprehensive and multifaceted regulatory scheme akin to that subserved by phosphorylation/dephosphorylation. It may be anticipated that the anti-inflammatory actions of NO via S-nitrosylation will be relevant across a range of vascular pathologies from atherosclerosis to vasculitis and septic shock.

Reperfusion Injury

Following a period of transient tissue ischemia, re-establishment of vascular blood flow and O2 delivery causes paradoxical tissue damage referred to as “reperfusion injury”81. Elucidating the biochemical and molecular mechanisms of reperfusion injury has been an active area of investigation, inasmuch as amelioration would be of significant benefit during both percutaneous and pharmacological reperfusion techniques. Altered S-nitrosylation is intimately linked with reperfusion injury, helping to explain the salutary actions of statins, estrogen and mitochondrial respiratory chain inhibitors. In particular, atorvastatin stimulates iNOS-mediated S-nitrosylation of COX282, thereby generating cytoprotective prostaglandins. In addition, estrogen appears to exert its cardioprotective effect, at least in part, by augmenting S-nitrosylation of mitochondrial proteins 8385. Whereas dysregulated S-nitrosylation appears to facilitate injury via irreversible inhibition of mitochondrial complex I (necessary for converting electrons from NADH to an ATP-producing proton gradient)8688, targeted delivery of nitrosylating agents to mitochondria is protective in ischemia/reperfusion (I/R), and inhibition of reactive oxygen generation by complex I may be involved89. iNOS also contributes to the protective effects of pre-conditioning (exposure to moderate hypoxia prior to I/R, which attenuates reperfusion injury) that are partly recapitulated by nitroglycerin90, 91. Inasmuch as nitroglycerin bioactivation occurs predominantly in mitochondria and results in accumulation of protein S-nitrosothiols92, it may be suggested that S-nitrosylation plays a protective role85, 89, 93, 94. In support of this idea, SNO-proteins that have been shown to increase in pre-conditioned hearts are also identified following I/R95.

Atherogenesis, risk factors and circulating SNO

Defective S-nitrosylation may contribute significantly to the pathophysiology of atherosclerosis. This condition reflects a contribution from myriad factors, including disruptions in the NO/redox equilibrium, immunologic/inflammatory stresses, platelet activation, aberrant vessel tone, and age-associated endothelial dysfunction. A major role in atherosclerosis for oxidative stress with resultant NO/redox disequilibrium is well-characterized43. As described above, thioredoxin functions as a critical regulator of cellular redox status and as an important modulator of S-nitrosylation. Recent studies suggest that statins, frequently employed anti-dyslipidemic and vasculoprotective agents, may exert their effects, at least in part, by inducing both S-nitrosylation and activation of thioredoxin96.

Platelet activation is highly relevant in atherosclerosis. The role of NO in platelet biology has been complicated in recent years by findings that both eNOS and iNOS may contribute to platelet cGMP production, and that cGMP may exert both inhibitory and stimulatory effects97. In addition, accumulating evidence also supports the idea that NO may inhibit platelet aggregation via a cGMP-independent pathway98102. Platelet aggregation is the third and final stage of the platelet activation process, preceded by platelet adherence and granule secretion (exocytosis). The contents of the various platelet granules (i.e., dense, α-, and lysosomal granules) figure prominently in platelet recruitment, rolling, adherence, and aggregation. As in the case of endothelial exocytosis discussed above, S-nitrosylation of NSF also exerts an inhibitory effect on exocytosis in platelets, thereby suppressing thrombosis and vascular inflammation 74(Figure 2). These effects are mediated by endogenously generated NO, inasmuch as platelets from eNOS−/− mice exhibit increased rolling on venules, increased arteriolar thrombosis, and increased exocytosis in vivo74.

Hypertension is a primary risk factor for progression of atherosclerotic disease and cardiovascular morbidity and mortality. Endogenous SNOs are implicated as key mediators of vasodilation and blood pressure control19, 46, 103, and in plasma, SNO-albumin provides a major reservoir of NO bioactivity104. However, albumin can also serve as a deleterious NO sink, whereby excessive sequestration of endogenous NO as S-nitrosoalbumin (SNO-albumin) negatively impacts vascular homeostasis in a variety of pathophysiological states103. Notably, albumin infusions may precipitate elevations in blood pressure by limiting the pool of bioavailable SNO for basal vessel relaxation105. Similarly, increased plasma SNO levels, suggestive of impaired NO delivery or excessive NO sequestration, are associated with adverse cardiovascular outcomes and hypertension in end-stage renal disease patients106, and misappropriation of NO as SNO-albumin is also directly implicated in the pathogenesis of hypertension in preeclampsia107, 108. It is important to emphasize, however, that exogenously administered SNO-albumin has been shown to serve as an effective therapeutic agent in a number of animal models including ischaemia/reperfusion-associated heart damage109, 110, lung injury in sickle cell disease111 and cardiopulmonary dysfunction in endotoxemia112, 113.

A number of studies have confirmed that the aging process is accompanied by a progressive decrease in bioavailable NO and concomitant endothelial dysfunction15, 114116. Explanations include increased superoxide production and elevated levels of naturally occurring NOS inhibitors. Upregulation of arginase activity in aging vasculature has also been espoused as a predominant mechanism for age-related endothelial dysfunction. Elevated levels of arginase, which competes directly with NOS for the common substrate L-arginine, would theoretically limit the amount of NO synthesized. Indeed, it was reported that in vitro inhibition of arginase activity restores (NO-based) vasodilation in aortic rings derived from aged rats114, 116. A subsequent study revealed that arginase is activated by S-nitrosylation of a single Cys residue, leading to its stabilization and to substantially enhanced substrate affinity (6-fold reduction in Km)15, which might enhance its ability to compete with NOS. Moreover, S-nitrosylation of arginase was increased in blood vessels from aging rats and was mediated by iNOS, previously shown to be expressed in aging vasculature115.

SNO-Hemoglobin and Hypoxic Vasodilation

Hypoxic vasodilation is an autoregulatory physiological response that maximizes blood flow to regions in the arterial periphery with low hemoglobin (Hb) O2 saturation, thereby matching perfusion with tissue O2 demand4, 17, 117. The progressive diminution in blood O2 content accompanying the decline in arteriolar diameter within the microcirculation results in graded vasodilation. Autoregulation of blood flow occurs within seconds or less (A–V transit times) and is recapitulated by direct intra-arterial infusion of variably deoxygenated but not of oxygenated RBCs118, and RBCs added to aortic ring bioassays at varying PO2 actuate graded vasodilation (Figure 3A)17. Moreover, these RBC-induced responses can be replicated by S-nitrosohemoglobin (SNO-Hb), which has a well-documented role in mediating hypoxic vasodilation4, 119122. Infusion of SNO-Hb (but not unmodified Hb) augments blood flow in vivo under normoxic conditions (Figure 3B)122, and vasodilation is blunted in the setting of supraphysiological PO2 (e.g., ambient O2 at 3 atmospheres absolute)122 (Figure 3C). By contrast, hypoxemia augments vasodilation by SNO-Hb. Changes in peripheral blood flow are predictably correlated with circulating SNO-Hb concentrations123.

Figure 3.

Figure 3

SNO-Hb subserves hypoxic vasodilation. (A) PO2 determines the ability of RBCs to constrict or relax aortic ring preparations on a second-by-second time scale. PO2 is indicated for each curve, which illustrate a graded response. (B) and (C) O2-dependent effects of SNO-Hb and Hb on local cerebral blood flow are shown in normoxia and hyperoxia. SNO-Hb infusion in vivo (1µmol/kg over 3min, beginning at time 0) immediately increases local cerebral blood flow in the caudate-putamen nucleus of rats breathing 21% O2 at 1 atmosphere absolute (ATA), where tissue PO2 ranges from 19 to 37 mmHg. Thus, SNO-Hb appropriately increases blood flow in relatively hypoxic tissue; however, non-nitrosylated Hb decreases perfusion. In 100% O2 at 3 ATA, where tissue PO2 ranges from 365 to 538 mmHg, vasodilation is abrogated because SNO-Hb cannot allosterically dispense NO bioactivity (Adapted from Allen et al.)158. (D) Allosteric transitions of circulating hemoglobin (Hb) regulate delivery of NO bioactivity to preserve vascular O2 homeostasis. Hb in RBCs senses [O2] and responds through allosterically controlled NO binding, SNO formation, and NO group release. At high O2 in the pulmonary venous system, Hb is in the R-state, Cys β93 is reactive and Cys93-SNO is shielded in a hydrophobic pocket. On partial RBC deoxygenation in the periphery, Hb adopts the T configuration and Cys β93-SNO is exposed to solvent. Further, in venous blood a population of deoxygenated (T-state) Hb reacts with NO to produce nitrosylated heme in the β-chain (bottom left). Transition to R-state draws Cys β93 close to the nitrosylated heme (top left) with a subsequent transfer of NO from heme to Cys β93, forming a SNO (top right). Deoxygenation of Hb favors the T conformation (bottom right), allowing SNO-Cys β93 to react with other cellular thiols, and thereby facilitating release of NO/SNO from the RBC. (Adapted from 126).

Hemoglobin exists predominantly in one of two structural states: R (relaxed, high O2 affinity) and T (tense, low O2 affinity)120. S-nitrosylation of hemoglobin (to generate SNO-Hb) occurs at Cys93 of the β subunit (Cys β93)120. The allosteric conformation of the Hb molecule governs reactivity of the Cys β93 residue, and thus the propensity for NO binding (Figure 3D). SNO-Hb formation is favored in the oxygenated (R) structure, whereas in the T configuration (e.g., hypoxia, low pH), NO groups are released to the surrounding tissues with resultant vasodilation. The duality of SNO-Hb as simultaneous O2 carrier and NO donor may plausibly be harnessed in creating viable blood substitutes, with some applicability already reported in ischemic myocardium124. Additionally, it appears that the coronary vasodilator nitroglycerin improves perfusion of ischemic myocardium by utilizing SNO-Hb-mediated O2 delivery in concert with NO unloading125. Inasmuch as RBCs play a central role in autoregulation of blood flow4, perturbations in the delivery of SNO by RBCs may underlie a variety of pathophysiological states characterized by microvascular dysfunction126. For example, pulmonary hypertension, a clinical entity often triggered by sustained hypoxemia, leads to depletion of RBC SNO-Hb stores and consequently to defective PO2-coupled vasoregulation and ventilation/perfusion mismatching127. Moreover, in vivo repletion of SNO-Hb can correct these physiologic deficits. Defective production of SNO-Hb by sickle RBCs has been implicated in impaired vasoregulation in sickle cell disease; the severity of symptoms is correlated with the degree of impairment of SNO-Hb processing and of RBC-mediated vasodilation, and these deficits can be ameliorated by repletion of SNO-Hb 128. In diabetes, derangements of SNO delivery by RBCs, resulting from glycosylation of Hb, which promotes the R configuration and thereby limits NO delivery, may exacerbate the vasculopathy associated with this disease129, 130.

Roles of S-Nitrosylation in Cardiac Signaling

Electrophysiology

Within the heart, S-nitrosylation has emerged as a ubiquitous signaling modality, impacting virtually every facet of cardiac function and dysfunction. The elaborate cascade of Ca2+ cycling that underlies excitation-contraction coupling (ECC) is no exception131. ECC spans an ordered sequence from electrical excitation of the individual myocyte to heart contraction, subserved by the tightly regulated trafficking of Ca2+ flux from one cellular compartment to another132. Upon membrane depolarization of the cardiac myocyte generated by voltage-gated Na+ channels, a cytosolic influx of Ca2+ occurs via the plasmalemmal L-type Ca2+ channel. Through a process known as Ca2+-induced Ca2+ release, this initial Ca2+ current triggers a more pronounced Ca2+ release from the sarcoplasmic reticulum (SR) through the ryanodine receptor/Ca2+ release channel (RyR2). Myocyte contraction proceeds when Ca2+ binds to troponin C in myofilaments, activating myosin ATPase. Relaxation of the myocyte entails diastolic reuptake and extrusion of cytosolic Ca2+ by way of the SR Ca2+ ATPase (SERCA2a) and the sarcolemmal Na+/Ca2+ exchanger132, respectively.

The ion channels participating in ECC, as well as those that determine the shape and duration of the action potential (see below) are modulated by S-nitrosylation85, 131, 133138, which thereby exerts effects on both contractility and arrhythmogenesis. The specific effects exerted by NO in the myocyte are dictated in part by the subcellular compartmentalization of NOSs (NOS1and NOS3), which reside in close proximity to substrates for S-nitrosylation139(Figure 4). NOS3 is spatially confined to sarcolemmal membrane caveolae and is thus adjacent to the L-type Ca2+ channel, whose S-nitrosylation inhibits ion influx135, 140. In a similar fashion, NOS1 resides in the SR where it is complexed with RyR2, and S-nitrosylation activates RyR2 (increases channel opening probability, Po)136, 141. Co-localization and targeted S-nitrosylation may also hold true for SERCA2a142, 143. Collectively, these observations demonstrate the precise spatiotemporal regulation of S-nitrosylation that underlies control by NO of cardiac ECC.

Figure 4.

Figure 4

S-nitrosylation regulates myocardial Ca2+ handling and thereby excitation-contraction coupling. RyR2, the cardiac form of the tetrameric ryanodine receptor/Ca2+ release channel, is localized to the SR membrane in proximity with the plasma membrane L-type calcium channel (LTCC), which provides the substrate for calcium-mediated calcium release from SR to cytosol. The SR-localized Ca2+-ATPase (SERCA) replenishes SR Ca2+. RyR2 co-localizes with nNOS in the SR, and S-nitrosylation of RyR2 (mediated by GSNO) potentiates Ca2+ release. As in skeletal muscle RyR1, physiological S-nitrosylation of one or a few Cys within each RyR2 monomer is likely to be the case. S-nitrosylation of the LTCC (α1C subunit; resulting in, for example, attenuated β-AR-dependent contractility) and of SERCA is inhibitory. Hypo-S-nitrosylation of RyR2 is associated with diastolic Ca2+ leakage and arrhythmia characteristic of sudden cardiac death. S-nitrosylation of the LTCC has been associated with ischemic preconditioning that reduces reperfusion injury, whereas hyper-S-nitrosylation of the LTCC has been associated with atrial fibrillation. Note in addition that aberrant S-nitrosylation can result from the translocation of nNOS to the plasma membrane that is seen in association with myocardial infarction and cardiomyopathy. Further, aberrant and in particular hyper-S-nitrosylation can result in irreversible oxidative modification of S-nitrosylated proteins in concert with reactive oxygen species produced by endogenous enzymes including xanthine oxidase.

It has been reported that nNOS re-distributes to the sarcolemma in heart failure, where it may regulate both β-adrenergic responsiveness and Ca2+ flux143145, and the deleterious consequences of myocardial infarction in mice (ventricular arrhythmia and mortality) are significantly more severe in nNOS−/− animals than in wild-type animals, in association with decreased S-nitrosylation of L-type Ca2+ channels143. Thus, inhibitory S-nitrosylation of L-type Ca2+ channels by nNOS is likely anti-arrhythmogenic.

More generally, disruption of the NO/redox equilibrium in myocytes, through alteration of either levels or spatiotemporal distribution of NO/ROS, is widely regarded as a sine qua non of heart failure146. Up-regulation of oxidant (ROS) production, notably by xanthine oxidase, can overwhelm endogenous, NO-based signaling and promote the mechano-energetic uncoupling characteristic of cardiac dysfunction147. Therapies directed against xanthine oxidase enable reverse remodeling in rats with dilated cardiomyopathy148. Thus, restoration of NO/redox homeostasis provides a potentially fruitful approach to restoring cardiac contractile function149

β-Adrenergic Receptor System

In cardiac myocytes, eNOS is activated following β-adrenergic receptor stimulation150, and important roles have been demonstrated for S-nitrosylation in transducing adrenergic signals. For example, S-nitrosylation of the L-type Ca2+ channel increases following isoproterenol stimulation in an eNOS-dependent manner135. Interestingly, a difference in protein S-nitrosylation appears to explain, at least in part, the gender disparity in I/R injury: females exhibit higher SNO levels and improved protection83.

Densensitization of GPCRs is a characteristic feature of disease, as is a deficiency of NO bioactivity. Recent studies have helped to connect these phenomena by demonstrating that GRK2 undergoes agonist-coupled, inhibitory S-nitrosylation (Figure 5B). GRK2 activity is a molecular correlate of receptor densensitization. Thus, S-nitrosylation leads to decreased β-adrenergic receptor (β-AR) phosphorylation and desensitization29, and absent S-nitrosylation, cardiac contractility declines rapidly during maintained adrenergic stimulation (Figure 5B). β-arrestin 2, a scaffolding protein that targets receptors for stimulus-coupled internalization, has also been shown to undergo S-nitrosylation28, leading to enhanced eNOS-dependent receptor trafficking (Figure 5A). Importantly, these studies reported increased protein S-nitrosylation (SNO-GRK2 and SNO-β-arrestin) in GSNOR−/− mice. Finally, S-nitrosylation of dynamin facilitates clathrin-dependent endocytosis of membrane receptors including the β-AR and thereby receptor downregulation 151 (Figure 5C). S-nitrosylation is thus under enzymatic control and GSNO is a central player in β-adrenergic receptor signaling. Other studies have demonstrated that VEGF may regulate GSNOR expression152; crosstalk between VEGF (and other pro-vascular signals) and β-adrenergic receptors may be mediated via the GSNOR/S-nitrosylation axis.

Figure 5.

Figure 5

A Schematic summary of the regulation of agonist-induced β2-adrenergic receptor trafficking by S-nitrosylation/denitrosylation of β-arrestin 2 (β-Arr 2), G protein-coupled receptor kinase 2 (GRK2), and dynamin. (A) β-Arr 2 serves as a scaffold that functionally colocalizes eNOS and β-ARs (as well as other G protein-coupled receptors [GPCRs]). Ligand (isoproterenol) stimulation results in activation of eNOS and S-nitrosylation of β-Arr 2. S-nitrosylation of β-Arr 2 promotes its dissociation from eNOS and its association with clathrin heavy chain/β-adaptin, which facilitates routing of the β2-AR into the clathrin-based endocytotic pathway, and β-Arr 2 is subsequently denitrosylated. (B) Inhibition of GRK2 by ligand-coupled S-nitrosylation suppresses agonist-stimulated β-AR phosphorylation, β-Arr 2 recruitment, and receptor desensitization and downregulation (schematic at top) At bottom: desensitization (decline in cardiac contractility in the continued presence of ISO) is enhanced by inhibiting NO production. (C) After GPCR activation, eNOS-mediated S-nitrosylation of dynamin promotes multimerization and GTPase activity, as well as relocation to the plasma membrane, which facilitates scission of endocytotic vesicles and receptor internalization. (Adapted from28, 29).

Arrhythmogenesis

As indicated above, the shape and duration of the cardiac action potential are regulated by multiple ion channels that are subject to regulatory S-nitrosylation138 (Figure 6 summarizes findings in the case of the ventricular action potential). For example, the inward-rectifying K current (IK1) shapes phase 3, and S-nitrosylation of a single cysteine in the relevant channel protein, Kir2.1, shortens the action potential153. Chronic atrial fibrillation was associated with decreased Kir2.1 S-nitrosylation, as assessed in human atrial samples.

Figure 6.

Figure 6

S-nitrosylation of channel proteins regulates all phases of the ventricular action potential. S-nitrosylation of SCN5A channels enhances the Na+ current (INa)157, 159, whereas S-nitrosylation of the α1C subunit of the L-type Ca2+ channel inhibits the L-type Ca2+ current (ICaL)135, 140. Among voltage-gated potassium channels, S-nitrosylation of the KCNQ1 subunit facilitates the slowly activating component of the delayed rectifier K+ current (IKs)133, whereas S-nitrosylation exerts an inhibitory influence on Kv4.3 and thus the transient outward potassium current (ITo)160, as well as Kir2.1153, and thus IK1 (phase 4). In addition, heterologously expressed human ether-a-go-go-related gene 1 (hERG1) potassium channels, which mediate the rapidly activating delayed rectifier K+ channel (IKr) in their native environment, are inhibited by NO in a cGMP-independent fashion161. Note that, in the atrium, S-nitrosylation inhibits hKv1.5 and thus the ultra-rapid delayed rectifier current 162 (Adapted from138).

A role for dysregulated S-nitrosylation in the development of a number of cardiac arrhythmias is supported by additional studies. Gonzalez et al.141 demonstrated that nNOS-mediated S-nitrosylation of RyR2 is critical for maintaining intracellular Ca2+ homeostasis (Figure 4). Mice deficient in nNOS exhibit a diastolic calcium leak141, which creates contractile dysfunction and a pro-arrythmogenic state154, 155. nNOS mutant mice also exhibit a pro-arrhythmic state following myocardial infarction, which is associated with diminished S-nitrosylation of RyR2, SERCA2a and L-type calcium channel143.

Mutations in α-syntrophin, a dystrophin-associated protein that acts as a scaffold between nNOS and the plasma membrane Ca-ATPase156, have been shown to contribute to long-QT syndrome157. The Ala390→Val mutation in α-syntrophin alters the inhibitory interaction between nNOS and the plasma membrane Ca-ATPase, and the resultant S-nitrosylation of the Na+ channel SCN5A enhances Na+ influx157, a pro-arrhythmic event recapitulating long-QT syndrome.

Conclusions

NO plays an important role in virtually all aspects of cardiac and vascular physiology. However, the molecular details are understood in only very few instances. The emergence of SNOs as second messengers and of S-nitrosylation as the preeminent NO-based signal presages a new era in cardiovascular biology. Unraveling the molecular underpinnings of SNO-based cardiovascular function and pathophysiology will undoubtedly yield novel therapeutic targets with great potential to improve clinical outcomes38.

Table 2.

Aberrant S-nitrosylation in Cardiovascular Disease

SNO-protein Disease State
Serum albumin Pre-eclampsia107, 108
Ischemic coronary syndromes106
Hemoglobin Congestive heart failure165
Pulmonary arterial hypertension127
Sickle cell disease128
Diabetes (Type 1) 129, 130
Septic shock19
Hif-1α Pulmonary arterial hypertension60
Matrix metalloproteinase 9 Stroke166
Ryanodine receptor 2 Arrhythmogenesis, Heart failure 141
L-type Ca2+ channel (α1C subunit) Atrial fibrillation/arrhythmia135, 143, 167
Cardiac Na+ channel SNC5a Long Q/T syndrome157
Slowly activating delayed-rectifier K+ channel Atrial fibrillation/arrhythmia 133, 134
Insulin receptor β Diabetes (Type 2)168, 169
Insulin receptor substrate 1 Diabetes (Type 2)168, 169
Akt (protein kinase B) Diabetes (Type 2)14, 168, 169

Examples of proteins for which hypo- or hyper-S-nitrosylation has been implicated in the mechanism of disease. Note in addition that S-nitrosylation of multiple substrates including Cox2, Hif-1α, the L-type Ca2+ channel, RyR2 and SERCA2 is implicated in the cardio-protective effects of both statins and ischemia-or drug-induced preconditioning, and in amelioration of the effects of myocardial infarction (see text).

Acknowledgments

Sources of Funding

This work was supported by NIH grants R01-HL091876, R01-HL095463, P01-HL075443, R01-HL096673, and R01-HL059130.

Non-standard abbreviations and acronyms

EDRF

endothelium-derived relaxation factor

GSNO

S-nitrosoglutathione

GSNOR

S-nitrosoglutathione reductase

SNO

S-nitrosothiol

Footnotes

Disclosures: None.

This Review is in a thematic series on Novel Post-translational Modifications of Proteins and Their Cardiovascular Significance, which includes the following articles:

The Emerging Characterization of Lysine Residue Deacetylation on The Modulation of Mitochondrial Function and Cardiovascular Biology [2009;105:830–841]

Protein Acetylation in the Cardiorenal Axis: The Promise of Histone Deacetylase Inhibitors [please fill in pub. info from 2/5/10 issue]

Protein S-Nitrosylation and Cardioprotection [please fill in pub. info from 2/5/10 issue]

Sent to Destroy: The Ubiquitin Proteasome System Regulates Cell Signaling and Protein Quality Control in Cardiovascular Development and Disease [please fill in pub. info from 2/19/10 issue]

S-Nitrosylation in Cardiovascular Signaling

S-Nitrosylation and Cardiac Ischemia

Glycosylation and Cardiovascular Signaling

Ubiquitination and Cardiovascular Signaling

Elizabeth Murphy, Guest Editor

References

  • 1.Guyton AC, Coleman TG, Granger HJ. Circulation: overall regulation. Annu Rev Physiol. 1972;34:13–46. doi: 10.1146/annurev.ph.34.030172.000305. [DOI] [PubMed] [Google Scholar]
  • 2.McMahon TJ, Moon RE, Luschinger BP, Carraway MS, Stone AE, Stolp BW, Gow AJ, Pawloski JR, Watke P, Singel DJ, Piantadosi CA, Stamler JS. Nitric oxide in the human respiratory cycle. Nat Med. 2002;8:711–717. doi: 10.1038/nm718. [DOI] [PubMed] [Google Scholar]
  • 3.Dzik S. Nitric oxide: nature's third respiratory gas. Transfusion. 2002;42:1532–1533. doi: 10.1046/j.1537-2995.2002.00288.x. [DOI] [PubMed] [Google Scholar]
  • 4.Allen BW, Stamler JS, Piantadosi CA. Hemoglobin, nitric oxide and molecular mechanisms of hypoxic vasodilation. Trends Mol Med. 2009;15:452–460. doi: 10.1016/j.molmed.2009.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Bredt DS, Snyder SH. Nitric oxide: a physiologic messenger molecule. Annu Rev Biochem. 1994;63:175–195. doi: 10.1146/annurev.bi.63.070194.001135. [DOI] [PubMed] [Google Scholar]
  • 6.Kots AY, Martin E, Sharina IG, Murad F. A short history of cGMP, guanylyl cyclases, and cGMP-dependent protein kinases. Handb Exp Pharmacol. 2009;191:1–14. doi: 10.1007/978-3-540-68964-5_1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Murad F. Cyclic guanosine monophosphate as a mediator of vasodilation. J Clin Invest. 1986;78:1–5. doi: 10.1172/JCI112536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Foster MW, McMahon TJ, Stamler JS. S-nitrosylation in health and disease. Trends Mol Med. 2003;9:160–168. doi: 10.1016/s1471-4914(03)00028-5. [DOI] [PubMed] [Google Scholar]
  • 9.Hess DT, Matsumoto A, Kim SO, Marshall HE, Stamler JS. Protein S-nitrosylation: purview and parameters. Nat Rev Mol Cell Biol. 2005;6:150–166. doi: 10.1038/nrm1569. [DOI] [PubMed] [Google Scholar]
  • 10.Sayed N, Kim DD, Fioramonti X, Iwahashi T, Duran WN, Beuve A. Nitroglycerin-induced S-nitrosylation and desensitization of soluble guanylyl cyclase contribute to nitrate tolerance. Circ Res. 2008;103:606–614. doi: 10.1161/CIRCRESAHA.108.175133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Murray CI, Gebska MA, Haile A, Zhang M, Kass DA, Champion HC, Van Eyk JE. cGMP Specific phosphodiesterase type 5A activity is regulated by S-nitrosylation at Cys 181. Circulation. 2008;118:S415. [Google Scholar]
  • 12.Ravi K, Brennan LA, Levic S, Ross PA, Black SM. S-nitrosylation of endothelial nitric oxide synthase is associated with monomerization and decreased enzyme activity. Proc Natl Acad Sci U S A. 2004;101:2619–2624. doi: 10.1073/pnas.0300464101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Martinez-Ruiz A, Villanueva L, Gonzalez de Orduna C, Lopez-Ferrer D, Higueras MA, Tarin C, Rodriguez-Crespo I, Vazquez J, Lamas S. S-nitrosylation of Hsp90 promotes the inhibition of its ATPase and endothelial nitric oxide synthase regulatory activities. Proc Natl Acad Sci U S A. 2005;102:8525–8530. doi: 10.1073/pnas.0407294102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Yasukawa T, Tokunaga E, Ota H, Sugita H, Martyn JA, Kaneki M. S-nitrosylation- dependent inactivation of Akt/protein kinase B in insulin resistance. J Biol Chem. 2005;280:7511–7518. doi: 10.1074/jbc.M411871200. [DOI] [PubMed] [Google Scholar]
  • 15.Santhanam L, Lim HK, Lim HK, Miriel V, Brown T, Patel M, Balanson S, Ryoo S, Anderson M, Irani K, Khanday F, Di Costanzo L, Nyhan D, Hare JM, Christianson DW, Rivers R, Shoukas A, Berkowitz DE. Inducible NO synthase dependent S-nitrosylation and activation of arginase1 contribute to age-related endothelial dysfunction. Circ Res. 2007;101:692–702. doi: 10.1161/CIRCRESAHA.107.157727. [DOI] [PubMed] [Google Scholar]
  • 16.Leiper J, Murray-Rust J, McDonald N, Vallance P. S-nitrosylation of dimethylarginine dimethylaminohydrolase regulates enzyme activity: further interactions between nitric oxide synthase and dimethylarginine dimethylaminohydrolase. Proc Natl Acad Sci U S A. 2002;99:13527–13532. doi: 10.1073/pnas.212269799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Singel DJ, Stamler JS. Chemical physiology of blood flow regulation by red blood cells: the role of nitric oxide and S-nitrosohemoglobin. Annu Rev Physiol. 2005;67:99–145. doi: 10.1146/annurev.physiol.67.060603.090918. [DOI] [PubMed] [Google Scholar]
  • 18.Mitchell DA, Marletta MA. Thioredoxin catalyzes the S-nitrosation of the caspase-3 active site cysteine. Nat Chem Biol. 2005;1:154–158. doi: 10.1038/nchembio720. [DOI] [PubMed] [Google Scholar]
  • 19.Liu L, Yan Y, Zeng M, Zhang J, Hanes MA, Ahearn G, McMahon TJ, Dickfeld T, Marshall HE, Que LG, Stamler JS. Essential roles of S-nitrosothiols in vascular homeostasis and endotoxic shock. Cell. 2004;116:617–628. doi: 10.1016/s0092-8674(04)00131-x. [DOI] [PubMed] [Google Scholar]
  • 20.Benhar M, Forrester MT, Hess DT, Stamler JS. Regulated protein denitrosylation by cytosolic and mitochondrial thioredoxins. Science. 2008;320:1050–1054. doi: 10.1126/science.1158265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Benhar M, Forrester MT, Stamler JS. Protein denitrosylation: enzymatic mechanisms and cellular functions. Nat Rev Mol Cell Biol. 2009;10:721–732. doi: 10.1038/nrm2764. [DOI] [PubMed] [Google Scholar]
  • 22.Forrester MT, Thompson JW, Foster MW, Nogueira L, Moseley MA, Stamler JS. Proteomic analysis of S-nitrosylation and denitrosylation by resin-assisted capture. Nat Biotechnol. 2009;27:557–559. doi: 10.1038/nbt.1545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Jensen DE, Belka GK, Du Bois GC. S-Nitrosoglutathione is a substrate for rat alcohol dehydrogenase class III isoenzyme. Biochem J. 1998;331:659–668. doi: 10.1042/bj3310659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Liu L, Hausladen A, Zeng M, Que L, Heitman J, Stamler JS. A metabolic enzyme for S-nitrosothiol conserved from bacteria to humans. Nature. 2001;410:490–494. doi: 10.1038/35068596. [DOI] [PubMed] [Google Scholar]
  • 25.Forrester MT, Foster MW, Stamler JS. Assessment and application of the biotin switch technique for examining protein S-nitrosylation under conditions of pharmacologically induced oxidative stress. J Biol Chem. 2007;282:13977–13983. doi: 10.1074/jbc.M609684200. [DOI] [PubMed] [Google Scholar]
  • 26.Foster MW, Liu L, Zeng M, Hess DT, Stamler JS. A genetic analysis of nitrosative stress. Biochemistry. 2009;48:792–799. doi: 10.1021/bi801813n. [DOI] [PubMed] [Google Scholar]
  • 27.Davis ME, Grumbach IM, Fukai T, Cutchins A, Harrison DG. Shear stress regulates endothelial nitric-oxide synthase promoter activity through nuclear factor κB binding. J Biol Chem. 2004;279:163–168. doi: 10.1074/jbc.M307528200. [DOI] [PubMed] [Google Scholar]
  • 28.Ozawa K, Whalen EJ, Nelson CD, Mu Y, Hess DT, Lefkowitz RJ, Stamler JS. S-nitrosylation of β-arrestin regulates β-adrenergic receptor trafficking. Mol Cell. 2008;31:395–405. doi: 10.1016/j.molcel.2008.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Whalen EJ, Foster MW, Matsumoto A, Ozawa K, Violin JD, Que LG, Nelson CD, Benhar M, Keys JR, Rockman HA, Koch WJ, Daaka Y, Lefkowitz RJ, Stamler JS. Regulation of β-adrenergic receptor signaling by S-nitrosylation of G-protein-coupled receptor kinase 2. Cell. 2007;129:511–522. doi: 10.1016/j.cell.2007.02.046. [DOI] [PubMed] [Google Scholar]
  • 30.Sanghani PC, Davis WI, Fears SL, Green SL, Zhai L, Tang Y, Martin E, Bryan NS, Sanghani SP. Kinetic and cellular characterization of novel inhibitors of S-nitrosoglutathione reductase. J Biol Chem. 2009;284:24354–24362. doi: 10.1074/jbc.M109.019919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Lima B, Lam GK, Xie L, Diesen DL, Villamizar N, Nienaber J, Messina E, Bowles D, Kontos CD, Hare JM, Stamler JS, Rockman HA. Endogenous S-nitrosothiols protect against myocardial injury. Proc Natl Acad Sci U S A. 2009;106:6297–6302. doi: 10.1073/pnas.0901043106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Bateman RL, Rauh D, Tavshanjian B, Shokat KM. Human carbonyl reductase 1 is an S-nitrosoglutathione reductase. J Biol Chem. 2008;283:35756–35762. doi: 10.1074/jbc.M807125200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Trujillo M, Alvarez MN, Peluffo G, Freeman BA, Radi R. Xanthine oxidase-mediated decomposition of S-nitrosothiols. J Biol Chem. 1998;273:7828–7834. doi: 10.1074/jbc.273.14.7828. [DOI] [PubMed] [Google Scholar]
  • 34.Arner ES, Holmgren A. Physiological functions of thioredoxin and thioredoxin reductase. Eur J Biochem. 2000;267:6102–6109. doi: 10.1046/j.1432-1327.2000.01701.x. [DOI] [PubMed] [Google Scholar]
  • 35.Holmgren A. Thioredoxin and glutaredoxin systems. J Biol Chem. 1989;264:13963–13966. [PubMed] [Google Scholar]
  • 36.Sengupta R, Ryter SW, Zuckerbraun BS, Tzeng E, Billiar TR, Stoyanovsky DA. Thioredoxin catalyzes the denitrosation of low-molecular mass and protein S-nitrosothiols. Biochemistry. 2007;46:8472–8483. doi: 10.1021/bi700449x. [DOI] [PubMed] [Google Scholar]
  • 37.Stoyanovsky DA, Tyurina YY, Tyurin VA, Anand D, Mandavia DN, Gius D, Ivanova J, Pitt B, Billiar TR, Kagan VE. Thioredoxin and lipoic acid catalyze the denitrosation of low molecular weight and protein S-nitrosothiols. J Am Chem Soc. 2005;127:15815–15823. doi: 10.1021/ja0529135. [DOI] [PubMed] [Google Scholar]
  • 38.Foster MW, Hess DT, Stamler JS. Protein S-nitrosylation in health and disease: a current perspective. Trends Mol Med. 2009;15:391–404. doi: 10.1016/j.molmed.2009.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Bolotina VM, Najibi S, Palacino JJ, Pagano PJ, Cohen RA. Nitric oxide directly activates calcium-dependent potassium channels in vascular smooth muscle. Nature. 1994;368:850–853. doi: 10.1038/368850a0. [DOI] [PubMed] [Google Scholar]
  • 40.Moro MA, Russel RJ, Cellek S, Lizasoain I, Su Y, Darley-Usmar VM, Radomski MW, Moncada S. cGMP mediates the vascular and platelet actions of nitric oxide: confirmation using an inhibitor of the soluble guanylyl cyclase. Proc Natl Acad Sci U S A. 1996;93:1480–1485. doi: 10.1073/pnas.93.4.1480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Huang B, Chen SC, Wang DL. Shear flow increases S-nitrosylation of proteins in endothelial cells. Cardiovasc Res. 2009;83:536–546. doi: 10.1093/cvr/cvp154. [DOI] [PubMed] [Google Scholar]
  • 42.Whalen EJ, Johnson AK, Lewis SJ. β-adrenoceptor dysfunction after inhibition of NO synthesis. Hypertension. 2000;36:376–382. doi: 10.1161/01.hyp.36.3.376. [DOI] [PubMed] [Google Scholar]
  • 43.Harrison D, Griendling KK, Landmesser U, Hornig B, Drexler H. Role of oxidative stress in atherosclerosis. Am J Cardiol. 2003;91:7A–11A. doi: 10.1016/s0002-9149(02)03144-2. [DOI] [PubMed] [Google Scholar]
  • 44.Selemidis S, Dusting GJ, Peshavariya H, Kemp-Harper BK, Drummond GR. Nitric oxide suppresses NADPH oxidase-dependent superoxide production by S-nitrosylation in human endothelial cells. Cardiovasc Res. 2007;75:349–358. doi: 10.1016/j.cardiores.2007.03.030. [DOI] [PubMed] [Google Scholar]
  • 45.Kim JH, Bugaj LJ, Oh YJ, Bivalacqua TJ, Ryoo S, Soucy KG, Santhanam L, Webb A, Camara A, Sikka G, Nyhan D, Shoukas AA, Ilies M, Christianson DW, Champion HC, Berkowitz DE. Arginase inhibition restores NOS coupling and reverses endothelial dysfunction and vascular stiffness in old rats. J Appl Physiol. 2009;107:1249–1257. doi: 10.1152/japplphysiol.91393.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Kunieda T, Minamino T, Miura K, Katsuno T, Tateno K, Miyauchi H, Kaneko S, Bradfield CA, FitzGerald GA, Komuro I. Reduced nitric oxide causes age-associated impairment of circadian rhythmicity. Circ Res. 2008;102:607–614. doi: 10.1161/CIRCRESAHA.107.162230. [DOI] [PubMed] [Google Scholar]
  • 47.Kroll J, Waltenberger J. VEGF-A induces expression of eNOS and iNOS in endothelial cells via VEGF receptor-2 (KDR) Biochem Biophys Res Commun. 1998;252:743–746. doi: 10.1006/bbrc.1998.9719. [DOI] [PubMed] [Google Scholar]
  • 48.Papapetropoulos A, Garcia-Cardena G, Madri JA, Sessa WC. Nitric oxide production contributes to the angiogenic properties of vascular endothelial growth factor in human endothelial cells. J Clin Invest. 1997;100:3131–3139. doi: 10.1172/JCI119868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Dimmeler S, Fleming I, Fisslthaler B, Hermann C, Busse R, Zeiher AM. Activation of nitric oxide synthase in endothelial cells by Akt-dependent phosphorylation. Nature. 1999;399:601–605. doi: 10.1038/21224. [DOI] [PubMed] [Google Scholar]
  • 50.Ahmad S, Hewett PW, Wang P, Al-Ani B, Cudmore M, Fujisawa T, Haigh JJ, le Noble F, Wang L, Mukhopadhyay D, Ahmed A. Direct evidence for endothelial vascular endothelial growth factor receptor-1 function in nitric oxide-mediated angiogenesis. Circ Res. 2006;99:715–722. doi: 10.1161/01.RES.0000243989.46006.b9. [DOI] [PubMed] [Google Scholar]
  • 51.Hoffmann J, Dimmeler S, Haendeler J. Shear stress increases the amount of S-nitrosylated molecules in endothelial cells: important role for signal transduction. FEBS Lett. 2003;551:153–158. doi: 10.1016/s0014-5793(03)00917-7. [DOI] [PubMed] [Google Scholar]
  • 52.Erwin PA, Lin AJ, Golan DE, Michel T. Receptor-regulated dynamic S-nitrosylation of endothelial nitric-oxide synthase in vascular endothelial cells. J Biol Chem. 2005;280:19888–19894. doi: 10.1074/jbc.M413058200. [DOI] [PubMed] [Google Scholar]
  • 53.Sayed N, Baskaran P, Ma X, van den Akker F, Beuve A. Desensitization of soluble guanylyl cyclase, the NO receptor, by S-nitrosylation. Proc Natl Acad Sci U S A. 2007;104:12312–12317. doi: 10.1073/pnas.0703944104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Pi X, Wu Y, Ferguson JE, 3rd, Portbury AL, Patterson C. SDF-1α stimulates JNK3 activity via eNOS-dependent nitrosylation of MKP7 to enhance endothelial migration. Proc Natl Acad Sci U S A. 2009;106:5675–5680. doi: 10.1073/pnas.0809568106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Kang-Decker N, Cao S, Chatterjee S, Yao J, Egan LJ, Semela D, Mukhopadhyay D, Shah V. Nitric oxide promotes endothelial cell survival signaling through S-nitrosylation and activation of dynamin-2. J Cell Sci. 2007;120:492–501. doi: 10.1242/jcs.03361. [DOI] [PubMed] [Google Scholar]
  • 56.Hoffmann J, Haendeler J, Aicher A, Rossig L, Vasa M, Zeiher AM, Dimmeler S. Aging enhances the sensitivity of endothelial cells toward apoptotic stimuli: important role of nitric oxide. Circ Res. 2001;89:709–715. doi: 10.1161/hh2001.097796. [DOI] [PubMed] [Google Scholar]
  • 57.Wadham C, Parker A, Wang L, Xia P. High glucose attenuates protein S-nitrosylation in endothelial cells: role of oxidative stress. Diabetes. 2007;56:2715–2721. doi: 10.2337/db06-1294. [DOI] [PubMed] [Google Scholar]
  • 58.Pugh CW, Ratcliffe PJ. Regulation of angiogenesis by hypoxia: role of the HIF system. Nat Med. 2003;9:677–684. doi: 10.1038/nm0603-677. [DOI] [PubMed] [Google Scholar]
  • 59.Metzen E, Zhou J, Jelkmann W, Fandrey J, Brune B. Nitric oxide impairs normoxic degradation of HIF-1α by inhibition of prolyl hydroxylases. Mol Biol Cell. 2003;14:3470–3481. doi: 10.1091/mbc.E02-12-0791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Palmer LA, Doctor A, Chhabra P, Sheram ML, Laubach VE, Karlinsey MZ, Forbes MS, Macdonald T, Gaston B. S-nitrosothiols signal hypoxia-mimetic vascular pathology. J Clin Invest. 2007;117:2592–2601. doi: 10.1172/JCI29444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Li F, Sonveaux P, Rabbani ZN, Liu S, Yan B, Huang Q, Vujaskovic Z, Dewhirst MW, Li CY. Regulation of HIF-1α stability through S-nitrosylation. Mol Cell. 2007;26:63–74. doi: 10.1016/j.molcel.2007.02.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Sumbayev VV, Budde A, Zhou J, Brune B. HIF-1α protein as a target for S-nitrosation. FEBS Lett. 2003;535:106–112. doi: 10.1016/s0014-5793(02)03887-5. [DOI] [PubMed] [Google Scholar]
  • 63.Mannick JB, Asano K, Izumi K, Kieff E, Stamler JS. Nitric oxide produced by human B lymphocytes inhibits apoptosis and Epstein-Barr virus reactivation. Cell. 1994;79:1137–1146. doi: 10.1016/0092-8674(94)90005-1. [DOI] [PubMed] [Google Scholar]
  • 64.Dimmeler S, Haendeler J, Nehls M, Zeiher AM. Suppression of apoptosis by nitric oxide via inhibition of interleukin-1β-converting enzyme (ICE)-like and cysteine protease protein (CPP)-32-like proteases. J Exp Med. 1997;185:601–607. doi: 10.1084/jem.185.4.601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Kim YM, Talanian RV, Billiar TR. Nitric oxide inhibits apoptosis by preventing increases in caspase-3-like activity via two distinct mechanisms. J Biol Chem. 1997;272:31138–31148. doi: 10.1074/jbc.272.49.31138. [DOI] [PubMed] [Google Scholar]
  • 66.Mannick JB, Hausladen A, Liu L, Hess DT, Zeng M, Miao QX, Kane LS, Gow AJ, Stamler JS. Fas-induced caspase denitrosylation. Science. 1999;284:651–654. doi: 10.1126/science.284.5414.651. [DOI] [PubMed] [Google Scholar]
  • 67.Hoffmann J, Haendeler J, Zeiher AM, Dimmeler S. TNFα and oxLDL reduce protein S-nitrosylation in endothelial cells. J Biol Chem. 2001;276:41383–41387. doi: 10.1074/jbc.M107566200. [DOI] [PubMed] [Google Scholar]
  • 68.Haendeler J, Hoffmann J, Tischler V, Berk BC, Zeiher AM, Dimmeler S. Redox regulatory and anti-apoptotic functions of thioredoxin depend on S-nitrosylation at cysteine 69. Nat Cell Biol. 2002;4:743–749. doi: 10.1038/ncb851. [DOI] [PubMed] [Google Scholar]
  • 69.Kubes P, Suzuki M, Granger DN. Nitric oxide: an endogenous modulator of leukocyte adhesion. Proc Natl Acad Sci U S A. 1991;88:4651–4655. doi: 10.1073/pnas.88.11.4651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lefer DJ, Jones SP, Girod WG, Baines A, Grisham MB, Cockrell AS, Huang PL, Scalia R. Leukocyte-endothelial cell interactions in nitric oxide synthase-deficient mice. Am J Physiol. 1999;276:H1943–H1950. doi: 10.1152/ajpheart.1999.276.6.H1943. [DOI] [PubMed] [Google Scholar]
  • 71.Hickey MJ, Sharkey KA, Sihota EG, Reinhardt PH, Macmicking JD, Nathan C, Kubes P. Inducible nitric oxide synthase-deficient mice have enhanced leukocyte-endothelium interactions in endotoxemia. FASEB J. 1997;11:955–964. doi: 10.1096/fasebj.11.12.9337148. [DOI] [PubMed] [Google Scholar]
  • 72.Lowenstein CJ. Nitric oxide regulation of protein trafficking in the cardiovascular system. Cardiovasc Res. 2007;75:240–246. doi: 10.1016/j.cardiores.2007.03.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Matsushita K, Morrell CN, Cambien B, Yang SX, Yamakuchi M, Bao C, Hara MR, Quick RA, Cao W, O'Rourke B, Lowenstein JM, Pevsner J, Wagner DD, Lowenstein CJ. Nitric oxide regulates exocytosis by S-nitrosylation of N-ethylmaleimide-sensitive factor. Cell. 2003;115:139–150. doi: 10.1016/s0092-8674(03)00803-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Morrell CN, Matsushita K, Chiles K, Scharpf RB, Yamakuchi M, Mason RJ, Bergmeier W, Mankowski JL, Baldwin WM, 3rd, Faraday N, Lowenstein CJ. Regulation of platelet granule exocytosis by S-nitrosylation. Proc Natl Acad Sci U S A. 2005;102:3782–3787. doi: 10.1073/pnas.0408310102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.De Caterina R, Libby P, Peng HB, Thannickal VJ, Rajavashisth TB, Gimbrone MA, Jr, Shin WS, Liao JK. Nitric oxide decreases cytokine-induced endothelial activation. Nitric oxide selectively reduces endothelial expression of adhesion molecules and proinflammatory cytokines. J Clin Invest. 1995;96:60–68. doi: 10.1172/JCI118074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Khan BV, Harrison DG, Olbrych MT, Alexander RW, Medford RM. Nitric oxide regulates vascular cell adhesion molecule 1 gene expression and redox-sensitive transcriptional events in human vascular endothelial cells. Proc Natl Acad Sci U S A. 1996;93:9114–9119. doi: 10.1073/pnas.93.17.9114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Tsao PS, Buitrago R, Chan JR, Cooke JP. Fluid flow inhibits endothelial adhesiveness. Nitric oxide and transcriptional regulation of VCAM-1. Circulation. 1996;94:1682–1689. doi: 10.1161/01.cir.94.7.1682. [DOI] [PubMed] [Google Scholar]
  • 78.Ghosh S, Karin M. Missing pieces in the NF-κB puzzle. Cell. 2002;109 Suppl:S81–S96. doi: 10.1016/s0092-8674(02)00703-1. [DOI] [PubMed] [Google Scholar]
  • 79.Marshall HE, Stamler JS. Inhibition of NF-κB by S-nitrosylation. Biochemistry. 2001;40:1688–1693. doi: 10.1021/bi002239y. [DOI] [PubMed] [Google Scholar]
  • 80.Reynaert NL, Ckless K, Korn SH, Vos N, Guala AS, Wouters EF, van der Vliet A, Janssen-Heininger YM. Nitric oxide represses inhibitory κB kinase through S-nitrosylation. Proc Natl Acad Sci U S A. 2004;101:8945–8950. doi: 10.1073/pnas.0400588101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Murphy E, Steenbergen C. Mechanisms underlying acute protection from cardiac ischemia-reperfusion injury. Physiol Rev. 2008;88:581–609. doi: 10.1152/physrev.00024.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Atar S, Ye Y, Lin Y, Freeberg SY, Nishi SP, Rosanio S, Huang MH, Uretsky BF, Perez-Polo JR, Birnbaum Y. Atorvastatin-induced cardioprotection is mediated by increasing inducible nitric oxide synthase and consequent S-nitrosylation of cyclooxygenase-2. Am J Physiol Heart Circ Physiol. 2006;290:H1960–H1968. doi: 10.1152/ajpheart.01137.2005. [DOI] [PubMed] [Google Scholar]
  • 83.Murphy E, Steenbergen C. Gender-based differences in mechanisms of protection in myocardial ischemia-reperfusion injury. Cardiovasc Res. 2007;75:478–486. doi: 10.1016/j.cardiores.2007.03.025. [DOI] [PubMed] [Google Scholar]
  • 84.Murphy E, Steenbergen C. Cardioprotection in females: a role for nitric oxide and altered gene expression. Heart Fail Rev. 2007;12:293–300. doi: 10.1007/s10741-007-9035-0. [DOI] [PubMed] [Google Scholar]
  • 85.Sun J, Morgan M, Shen RF, Steenbergen C, Murphy E. Preconditioning results in S-nitrosylation of proteins involved in regulation of mitochondrial energetics and calcium transport. Circ Res. 2007;101:1155–1163. doi: 10.1161/CIRCRESAHA.107.155879. [DOI] [PubMed] [Google Scholar]
  • 86.Beltran B, Orsi A, Clementi E, Moncada S. Oxidative stress and S-nitrosylation of proteins in cells. Br J Pharmacol. 2000;129:953–960. doi: 10.1038/sj.bjp.0703147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Clementi E, Brown GC, Feelisch M, Moncada S. Persistent inhibition of cell respiration by nitric oxide: crucial role of S-nitrosylation of mitochondrial complex I and protective action of glutathione. Proc Natl Acad Sci U S A. 1998;95:7631–7636. doi: 10.1073/pnas.95.13.7631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Moncada S, Erusalimsky JD. Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat Rev Mol Cell Biol. 2002;3:214–220. doi: 10.1038/nrm762. [DOI] [PubMed] [Google Scholar]
  • 89.Prime TA, Blaikie FH, Evans C, Nadtochiy SM, James AM, Dahm CC, Vitturi DA, Patel RP, Hiley CR, Abakumova I, Requejo R, Chouchani ET, Hurd TR, Garvey JF, Taylor CT, Brookes PS, Smith RA, Murphy MP. A mitochondria-targeted S-nitrosothiol modulates respiration, nitrosates thiols, and protects against ischemia-reperfusion injury. Proc Natl Acad Sci U S A. 2009;106:10764–10769. doi: 10.1073/pnas.0903250106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Bolli R. Cardioprotective function of inducible nitric oxide synthase and role of nitric oxide in myocardial ischemia and preconditioning: an overview of a decade of research. J Mol Cell Cardiol. 2001;33:1897–1918. doi: 10.1006/jmcc.2001.1462. [DOI] [PubMed] [Google Scholar]
  • 91.Jneid H, Chandra M, Alshaher M, Hornung CA, Tang XL, Leesar M, Bolli R. Delayed preconditioning-mimetic actions of nitroglycerin in patients undergoing exercise tolerance tests. Circulation. 2005;111:2565–2571. doi: 10.1161/CIRCULATIONAHA.104.515445. [DOI] [PubMed] [Google Scholar]
  • 92.Stamler JS. Nitroglycerin-mediated S-nitrosylation of proteins: a field comes full cycle. Circ Res. 2008;103:557–559. doi: 10.1161/CIRCRESAHA.108.184341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Konorev EA, Joseph J, Tarpey MM, Kalyanaraman B. The mechanism of cardioprotection by S-nitrosoglutathione monoethyl ester in rat isolated heart during cardioplegic ischaemic arrest. Br J Pharmacol. 1996;119:511–518. doi: 10.1111/j.1476-5381.1996.tb15701.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Tao L, Gao E, Bryan NS, Qu Y, Liu HR, Hu A, Christopher TA, Lopez BL, Yodoi J, Koch WJ, Feelisch M, Ma XL. Cardioprotective effects of thioredoxin in myocardial ischemia and reperfusion: role of S-nitrosation [corrected] Proc Natl Acad Sci U S A. 2004;101:11471–11476. doi: 10.1073/pnas.0402941101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Lin J, Steenbergen C, Murphy E, Sun J. Estrogen receptor-β activation results in S-nitrosylation of proteins involved in cardioprotection. Circulation. 2009;120:245–254. doi: 10.1161/CIRCULATIONAHA.109.868729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Haendeler J, Hoffmann J, Zeiher AM, Dimmeler S. Antioxidant effects of statins via S-nitrosylation and activation of thioredoxin in endothelial cells: a novel vasculoprotective function of statins. Circulation. 2004;110:856–861. doi: 10.1161/01.CIR.0000138743.09012.93. [DOI] [PubMed] [Google Scholar]
  • 97.Li Z, Zhang G, Marjanovic JA, Ruan C, Du X. A platelet secretion pathway mediated by cGMP-dependent protein kinase. J Biol Chem. 2004;279:42469–42475. doi: 10.1074/jbc.M401532200. [DOI] [PubMed] [Google Scholar]
  • 98.Pawloski JR, Swaminathan RV, Stamler JS. Cell-free and erythrocytic S-nitrosohemoglobin inhibits human platelet aggregation. Circulation. 1998;97:263–267. doi: 10.1161/01.cir.97.3.263. [DOI] [PubMed] [Google Scholar]
  • 99.Beghetti M, Sparling C, Cox PN, Stephens D, Adatia I. Inhaled NO inhibits platelet aggregation and elevates plasma but not intraplatelet cGMP in healthy human volunteers. Am J Physiol Heart Circ Physiol. 2003;285:H637–H642. doi: 10.1152/ajpheart.00622.2002. [DOI] [PubMed] [Google Scholar]
  • 100.Crane MS, Rossi AG, Megson IL. A potential role for extracellular nitric oxide generation in cGMP-independent inhibition of human platelet aggregation: biochemical and pharmacological considerations. Br J Pharmacol. 2005;144:849–859. doi: 10.1038/sj.bjp.0706110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Sogo N, Magid KS, Shaw CA, Webb DJ, Megson IL. Inhibition of human platelet aggregation by nitric oxide donor drugs: relative contribution of cGMP-independent mechanisms. Biochem Biophys Res Commun. 2000;279:412–419. doi: 10.1006/bbrc.2000.3976. [DOI] [PubMed] [Google Scholar]
  • 102.Tsikas D, Ikic M, Tewes KS, Raida M, Frolich JC. Inhibition of platelet aggregation by S-nitroso-cysteine via cGMP-independent mechanisms: evidence of inhibition of thromboxane A2 synthesis in human blood platelets. FEBS Lett. 1999;442:162–166. doi: 10.1016/s0014-5793(98)01633-0. [DOI] [PubMed] [Google Scholar]
  • 103.Foster MW, Pawloski JR, Singel DJ, Stamler JS. Role of circulating S-nitrosothiols in control of blood pressure. Hypertension. 2005;45:15–17. doi: 10.1161/01.HYP.0000150160.41992.71. [DOI] [PubMed] [Google Scholar]
  • 104.Stamler JS, Jaraki O, Osborne J, Simon DI, Keaney J, Vita J, Singel D, Valeri CR, Loscalzo J. Nitric oxide circulates in mammalian plasma primarily as an S-nitroso adduct of serum albumin. Proc Natl Acad Sci U S A. 1992;89:7674–7677. doi: 10.1073/pnas.89.16.7674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Rafikova O, Rafikov R, Nudler E. Catalysis of S-nitrosothiols formation by serum albumin: the mechanism and implication in vascular control. Proc Natl Acad Sci U S A. 2002;99:5913–5918. doi: 10.1073/pnas.092048999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Massy ZA, Fumeron C, Borderie D, Tuppin P, Nguyen-Khoa T, Benoit MO, Jacquot C, Buisson C, Drueke TB, Ekindjian OG, Lacour B, Iliou MC. Increased plasma S-nitrosothiol concentrations predict cardiovascular outcomes among patients with end-stage renal disease: a prospective study. J Am Soc Nephrol. 2004;15:470–476. doi: 10.1097/01.asn.0000106716.22153.bb. [DOI] [PubMed] [Google Scholar]
  • 107.Gandley RE, Tyurin VA, Huang W, Arroyo A, Daftary A, Harger G, Jiang J, Pitt B, Taylor RN, Hubel CA, Kagan VE. S-nitrosoalbumin-mediated relaxation is enhanced by ascorbate and copper: effects in pregnancy and preeclampsia plasma. Hypertension. 2005;45:21–27. doi: 10.1161/01.HYP.0000150158.42620.3e. [DOI] [PubMed] [Google Scholar]
  • 108.Tyurin VA, Liu SX, Tyurina YY, Sussman NB, Hubel CA, Roberts JM, Taylor RN, Kagan VE. Elevated levels of S-nitrosoalbumin in preeclampsia plasma. Circ Res. 2001;88:1210–1215. doi: 10.1161/hh1101.092179. [DOI] [PubMed] [Google Scholar]
  • 109.Hallstrom S, Franz M, Gasser H, Vodrazka M, Semsroth S, Losert UM, Haisjackl M, Podesser BK, Malinski T. S-nitroso human serum albumin reduces ischaemia/reperfusion injury in the pig heart after unprotected warm ischaemia. Cardiovasc Res. 2008;77:506–514. doi: 10.1093/cvr/cvm052. [DOI] [PubMed] [Google Scholar]
  • 110.Semsroth S, Fellner B, Trescher K, Bernecker OY, Kalinowski L, Gasser H, Hallstrom S, Malinski T, Podesser BK. S-nitroso human serum albumin attenuates ischemia/reperfusion injury after cardioplegic arrest in isolated rabbit hearts. J Heart Lung Transplant. 2005;24:2226–2234. doi: 10.1016/j.healun.2005.08.004. [DOI] [PubMed] [Google Scholar]
  • 111.de Franceschi L, Malpeli G, Scarpa A, Janin A, Muchitsch EM, Roncada P, Leboeuf C, Corrocher R, Beuzard Y, Brugnara C. Protective effects of S-nitrosoalbumin on lung injury induced by hypoxia-reoxygenation in mouse model of sickle cell disease. Am J Physiol Lung Cell Mol Physiol. 2006;291:L457–L465. doi: 10.1152/ajplung.00462.2005. [DOI] [PubMed] [Google Scholar]
  • 112.Gluckman TL, Grossman JE, Folts JD, Kruse-Elliott KT. Modulation of endotoxin-induced cardiopulmonary dysfunction by S-nitroso-albumin. J Endotoxin Res. 2002;8:17–26. [PubMed] [Google Scholar]
  • 113.Jakubowski A, Maksimovich N, Olszanecki R, Gebska A, Gasser H, Podesser BK, Hallstrom S, Chlopicki S. S-nitroso human serum albumin given after LPS challenge reduces acute lung injury and prolongs survival in a rat model of endotoxemia. Naunyn Schmiedebergs Arch Pharmacol. 2009;379:281–290. doi: 10.1007/s00210-008-0351-2. [DOI] [PubMed] [Google Scholar]
  • 114.Berkowitz DE, White R, Li D, Minhas KM, Cernetich A, Kim S, Burke S, Shoukas AA, Nyhan D, Champion HC, Hare JM. Arginase reciprocally regulates nitric oxide synthase activity and contributes to endothelial dysfunction in aging blood vessels. Circulation. 2003;108:2000–2006. doi: 10.1161/01.CIR.0000092948.04444.C7. [DOI] [PubMed] [Google Scholar]
  • 115.Cernadas MR, Sanchez de Miguel L, Garcia-Duran M, Gonzalez-Fernandez F, Millas I, Monton M, Rodrigo J, Rico L, Fernandez P, de Frutos T, Rodriguez-Feo JA, Guerra J, Caramelo C, Casado S, Lopez F. Expression of constitutive and inducible nitric oxide synthases in the vascular wall of young and aging rats. Circ Res. 1998;83:279–286. doi: 10.1161/01.res.83.3.279. [DOI] [PubMed] [Google Scholar]
  • 116.White AR, Ryoo S, Li D, Champion HC, Steppan J, Wang D, Nyhan D, Shoukas AA, Hare JM, Berkowitz DE. Knockdown of arginase I restores NO signaling in the vasculature of old rats. Hypertension. 2006;47:245–251. doi: 10.1161/01.HYP.0000198543.34502.d7. [DOI] [PubMed] [Google Scholar]
  • 117.Singel DJ, Stamler JS. Blood traffic control. Nature. 2004;430:297. doi: 10.1038/430297a. [DOI] [PubMed] [Google Scholar]
  • 118.Ross JM, Fairchild HM, Weldy J, Guyton AC. Autoregulation of blood flow by oxygen lack. Am J Physiol. 1962;202:21–24. doi: 10.1152/ajplegacy.1962.202.1.21. [DOI] [PubMed] [Google Scholar]
  • 119.Diesen DL, Hess DT, Stamler JS. Hypoxic vasodilation by red blood cells: evidence for an S-nitrosothiol-based signal. Circ Res. 2008;103:545–553. doi: 10.1161/CIRCRESAHA.108.176867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Jia L, Bonaventura C, Bonaventura J, Stamler JS. S-nitrosohaemoglobin: a dynamic activity of blood involved in vascular control. Nature. 1996;380:221–226. doi: 10.1038/380221a0. [DOI] [PubMed] [Google Scholar]
  • 121.Pawloski JR, Hess DT, Stamler JS. Export by red blood cells of nitric oxide bioactivity. Nature. 2001;409:622–626. doi: 10.1038/35054560. [DOI] [PubMed] [Google Scholar]
  • 122.Stamler JS, Jia L, Eu JP, McMahon TJ, Demchenko IT, Bonaventura J, Gernert K, Piantadosi CA. Blood flow regulation by S-nitrosohemoglobin in the physiological oxygen gradient. Science. 1997;276:2034–2037. doi: 10.1126/science.276.5321.2034. [DOI] [PubMed] [Google Scholar]
  • 123.Sonveaux P, Kaz AM, Snyder SA, Richardson RA, Cardenas-Navia LI, Braun RD, Pawloski JR, Tozer GM, Bonaventura J, McMahon TJ, Stamler JS, Dewhirst MW. Oxygen regulation of tumor perfusion by S-nitrosohemoglobin reveals a pressor activity of nitric oxide. Circ Res. 2005;96:1119–1126. doi: 10.1161/01.RES.0000168740.04986.a7. [DOI] [PubMed] [Google Scholar]
  • 124.Asanuma H, Nakai K, Sanada S, Minamino T, Takashima S, Ogita H, Fujita M, Hirata A, Wakeno M, Takahama H, Kim J, Asakura M, Sakuma I, Kitabatake A, Hori M, Komamura K, Kitakaze M. S-nitrosylated and pegylated hemoglobin, a newly developed artificial oxygen carrier, exerts cardioprotection against ischemic hearts. J Mol Cell Cardiol. 2007;42:924–930. doi: 10.1016/j.yjmcc.2006.12.001. [DOI] [PubMed] [Google Scholar]
  • 125.Bin JP, Doctor A, Lindner J, Hendersen EM, Le DE, Leong-Poi H, Fisher NG, Christiansen J, Kaul S. Effects of nitroglycerin on erythrocyte rheology and oxygen unloading: novel role of S-nitrosohemoglobin in relieving myocardial ischemia. Circulation. 2006;113:2502–2508. doi: 10.1161/CIRCULATIONAHA.106.627091. [DOI] [PubMed] [Google Scholar]
  • 126.Sonveaux P, Lobysheva II, Feron O, McMahon TJ. Transport and peripheral bioactivities of nitrogen oxides carried by red blood cell hemoglobin: role in oxygen delivery. Physiology (Bethesda) 2007;22:97–112. doi: 10.1152/physiol.00042.2006. [DOI] [PubMed] [Google Scholar]
  • 127.McMahon TJ, Ahearn GS, Moya MP, Gow AJ, Huang YC, Luchsinger BP, Nudelman R, Yan Y, Krichman AD, Bashore TM, Califf RM, Singel DJ, Piantadosi CA, Tapson VF, Stamler JS. A nitric oxide processing defect of red blood cells created by hypoxia: deficiency of S-nitrosohemoglobin in pulmonary hypertension. Proc Natl Acad Sci U S A. 2005;102:14801–14806. doi: 10.1073/pnas.0506957102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Pawloski JR, Hess DT, Stamler JS. Impaired vasodilation by red blood cells in sickle cell disease. Proc Natl Acad Sci U S A. 2005;102:2531–2536. doi: 10.1073/pnas.0409876102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Milsom AB, Jones CJ, Goodfellow J, Frenneaux MP, Peters JR, James PE. Abnormal metabolic fate of nitric oxide in Type I diabetes mellitus. Diabetologia. 2002;45:1515–1522. doi: 10.1007/s00125-002-0956-9. [DOI] [PubMed] [Google Scholar]
  • 130.Padron J, Peiro C, Cercas E, Llergo JL, Sanchez-Ferrer CF. Enhancement of S-nitrosylation in glycosylated hemoglobin. Biochem Biophys Res Commun. 2000;271:217–221. doi: 10.1006/bbrc.2000.2617. [DOI] [PubMed] [Google Scholar]
  • 131.Hare JM. Nitric oxide and excitation-contraction coupling. J Mol Cell Cardiol. 2003;35:719–729. doi: 10.1016/s0022-2828(03)00143-3. [DOI] [PubMed] [Google Scholar]
  • 132.Bers DM. Cardiac excitation-contraction coupling. Nature. 2002;415:198–205. doi: 10.1038/415198a. [DOI] [PubMed] [Google Scholar]
  • 133.Asada K, Kurokawa J, Furukawa T. Redox- and calmodulin-dependent S-nitrosylation of the KCNQ1 channel. J Biol Chem. 2009;284:6014–6020. doi: 10.1074/jbc.M807158200. [DOI] [PubMed] [Google Scholar]
  • 134.Bai CX, Namekata I, Kurokawa J, Tanaka H, Shigenobu K, Furukawa T. Role of nitric oxide in Ca2+ sensitivity of the slowly activating delayed rectifier K+ current in cardiac myocytes. Circ Res. 2005;96:64–72. doi: 10.1161/01.RES.0000151846.19788.E0. [DOI] [PubMed] [Google Scholar]
  • 135.Sun J, Picht E, Ginsburg KS, Bers DM, Steenbergen C, Murphy E. Hypercontractile female hearts exhibit increased S-nitrosylation of the L-type Ca2+ channel α1 subunit and reduced ischemia/reperfusion injury. Circ Res. 2006;98:403–411. doi: 10.1161/01.RES.0000202707.79018.0a. [DOI] [PubMed] [Google Scholar]
  • 136.Sun J, Yamaguchi N, Xu L, Eu JP, Stamler JS, Meissner G. Regulation of the cardiac muscle ryanodine receptor by O2 tension and S-nitrosoglutathione. Biochemistry. 2008;47:13985–13990. doi: 10.1021/bi8012627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Xu L, Eu JP, Meissner G, Stamler JS. Activation of the cardiac calcium release channel (ryanodine receptor) by poly-S-nitrosylation. Science. 1998;279:234–237. doi: 10.1126/science.279.5348.234. [DOI] [PubMed] [Google Scholar]
  • 138.Gonzalez DR, AT MS, Sun QA, Stamler JS, Hare JM. S-nitrosylation of cardiac ion channels. J Cardiovasc Pharmacol. 2009;54:188–195. doi: 10.1097/FJC.0b013e3181b72c9f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Barouch LA, Harrison RW, Skaf MW, Rosas GO, Cappola TP, Kobeissi ZA, Hobai IA, Lemmon CA, Burnett AL, O'Rourke B, Rodriguez ER, Huang PL, Lima JA, Berkowitz DE, Hare JM. Nitric oxide regulates the heart by spatial confinement of nitric oxide synthase isoforms. Nature. 2002;416:337–339. doi: 10.1038/416337a. [DOI] [PubMed] [Google Scholar]
  • 140.Poteser M, Romanin C, Schreibmayer W, Mayer B, Groschner K. S-nitrosation controls gating and conductance of the alpha 1 subunit of class C L-type Ca2+ channels. J Biol Chem. 2001;276:14797–14803. doi: 10.1074/jbc.M008244200. [DOI] [PubMed] [Google Scholar]
  • 141.Gonzalez DR, Beigi F, Treuer AV, Hare JM. Deficient ryanodine receptor S-nitrosylation increases sarcoplasmic reticulum calcium leak and arrhythmogenesis in cardiomyocytes. Proc Natl Acad Sci U S A. 2007;104:20612–20617. doi: 10.1073/pnas.0706796104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Adachi T, Weisbrod RM, Pimentel DR, Ying J, Sharov VS, Schoneich C, Cohen RA. S-Glutathiolation by peroxynitrite activates SERCA during arterial relaxation by nitric oxide. Nat Med. 2004;10:1200–1207. doi: 10.1038/nm1119. [DOI] [PubMed] [Google Scholar]
  • 143.Burger DE, Lu X, Lei M, Xiang FL, Hammoud L, Jiang M, Wang H, Jones DL, Sims SM, Feng Q. Neuronal nitric oxide synthase protects against myocardial infarction-induced ventricular arrhythmia and mortality in mice. Circulation. 2009;120:1345–1354. doi: 10.1161/CIRCULATIONAHA.108.846402. [DOI] [PubMed] [Google Scholar]
  • 144.Bendall JK, Damy T, Ratajczak P, Loyer X, Monceau V, Marty I, Milliez P, Robidel E, Marotte F, Samuel JL, Heymes C. Role of myocardial neuronal nitric oxide synthase-derived nitric oxide in β-adrenergic hyporesponsiveness after myocardial infarction-induced heart failure in rat. Circulation. 2004;110:2368–2375. doi: 10.1161/01.CIR.0000145160.04084.AC. [DOI] [PubMed] [Google Scholar]
  • 145.Damy T, Ratajczak P, Shah AM, Camors E, Marty I, Hasenfuss G, Marotte F, Samuel JL, Heymes C. Increased neuronal nitric oxide synthase-derived NO production in the failing human heart. Lancet. 2004;363:1365–1367. doi: 10.1016/S0140-6736(04)16048-0. [DOI] [PubMed] [Google Scholar]
  • 146.Hare JM, Stamler JS. NO/redox disequilibrium in the failing heart and cardiovascular system. J Clin Invest. 2005;115:509–517. doi: 10.1172/JCI200524459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Saavedra WF, Paolocci N, St John ME, Skaf MW, Stewart GC, Xie JS, Harrison RW, Zeichner J, Mudrick D, Marban E, Kass DA, Hare JM. Imbalance between xanthine oxidase and nitric oxide synthase signaling pathways underlies mechanoenergetic uncoupling in the failing heart. Circ Res. 2002;90:297–304. doi: 10.1161/hh0302.104531. [DOI] [PubMed] [Google Scholar]
  • 148.Minhas KM, Saraiva RM, Schuleri KH, Lehrke S, Zheng M, Saliaris AP, Berry CE, Barouch LA, Vandegaer KM, Li D, Hare JM. Xanthine oxidoreductase inhibition causes reverse remodeling in rats with dilated cardiomyopathy. Circ Res. 2006;98:271–279. doi: 10.1161/01.RES.0000200181.59551.71. [DOI] [PubMed] [Google Scholar]
  • 149.Yano M, Okuda S, Oda T, Tokuhisa T, Tateishi H, Mochizuki M, Noma T, Doi M, Kobayashi S, Yamamoto T, Ikeda Y, Ohkusa T, Ikemoto N, Matsuzaki M. Correction of defective interdomain interaction within ryanodine receptor by antioxidant is a new therapeutic strategy against heart failure. Circulation. 2005;112:3633–3643. doi: 10.1161/CIRCULATIONAHA.105.555623. [DOI] [PubMed] [Google Scholar]
  • 150.Balligand JL, Kelly RA, Marsden PA, Smith TW, Michel T. Control of cardiac muscle cell function by an endogenous nitric oxide signaling system. Proc Natl Acad Sci U S A. 1993;90:347–351. doi: 10.1073/pnas.90.1.347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Wang G, Moniri NH, Ozawa K, Stamler JS, Daaka Y. Nitric oxide regulates endocytosis by S-nitrosylation of dynamin. Proc Natl Acad Sci U S A. 2006;103:1295–1300. doi: 10.1073/pnas.0508354103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Bhandari V, Choo-Wing R, Chapoval SP, Lee CG, Tang C, Kim YK, Ma B, Baluk P, Lin MI, McDonald DM, Homer RJ, Sessa WC, Elias JA. Essential role of nitric oxide in VEGF-induced, asthma-like angiogenic, inflammatory, mucus, and physiologic responses in the lung. Proc Natl Acad Sci U S A. 2006;103:11021–11026. doi: 10.1073/pnas.0601057103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Gomez R, Caballero R, Barana A, Amoros I, Calvo E, Lopez JA, Klein H, Vaquero M, Osuna L, Atienza F, Almendral J, Pinto A, Tamargo J, Delpon E. Nitric oxide increases cardiac IK1 by nitrosylation of cysteine 76 of Kir2.1 channels. Circ Res. 2009;105:383–392. doi: 10.1161/CIRCRESAHA.109.197558. [DOI] [PubMed] [Google Scholar]
  • 154.Wehrens XH, Lehnart SE, Huang F, Vest JA, Reiken SR, Mohler PJ, Sun J, Guatimosim S, Song LS, Rosemblit N, D'Armiento JM, Napolitano C, Memmi M, Priori SG, Lederer WJ, Marks AR. FKBP12.6 deficiency and defective calcium release channel (ryanodine receptor) function linked to exercise-induced sudden cardiac death. Cell. 2003;113:829–840. doi: 10.1016/s0092-8674(03)00434-3. [DOI] [PubMed] [Google Scholar]
  • 155.Wehrens XH, Lehnart SE, Marks AR. Intracellular calcium release and cardiac disease. Annu Rev Physiol. 2005;67:69–98. doi: 10.1146/annurev.physiol.67.040403.114521. [DOI] [PubMed] [Google Scholar]
  • 156.Williams JC, Armesilla AL, Mohamed TM, Hagarty CL, McIntyre FH, Schomburg S, Zaki AO, Oceandy D, Cartwright EJ, Buch MH, Emerson M, Neyses L. The sarcolemmal calcium pump, α-1 syntrophin, and neuronal nitric-oxide synthase are parts of a macromolecular protein complex. J Biol Chem. 2006;281(33):23341–23348. doi: 10.1074/jbc.M513341200. [DOI] [PubMed] [Google Scholar]
  • 157.Ueda K, Valdivia C, Medeiros-Domingo A, Tester DJ, Vatta M, Farrugia G, Ackerman MJ, Makielski JC. Syntrophin mutation associated with long QT syndrome through activation of the nNOS-SCN5A macromolecular complex. Proc Natl Acad Sci U S A. 2008;105:9355–9360. doi: 10.1073/pnas.0801294105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Allen BW, Piantadosi CA. How do red blood cells cause hypoxic vasodilation? The SNO-hemoglobin paradigm. Am J Physiol Heart Circ Physiol. 2006;291:H1507–H1512. doi: 10.1152/ajpheart.00310.2006. [DOI] [PubMed] [Google Scholar]
  • 159.Ahern GP, Hsu SF, Klyachko VA, Jackson MB. Induction of persistent sodium current by exogenous and endogenous nitric oxide. J Biol Chem. 2000;275:28810–28815. doi: 10.1074/jbc.M003090200. [DOI] [PubMed] [Google Scholar]
  • 160.Gomez R, Nunez L, Vaquero M, Amoros I, Barana A, de Prada T, Macaya C, Maroto L, Rodriguez E, Caballero R, Lopez-Farre A, Tamargo J, Delpon E. Nitric oxide inhibits Kv4.3 and human cardiac transient outward potassium current (Ito1) Cardiovasc Res. 2008;80:375–384. doi: 10.1093/cvr/cvn205. [DOI] [PubMed] [Google Scholar]
  • 161.Taglialatela M, Pannaccione A, Iossa S, Castaldo P, Annunziato L. Modulation of the K+ channels encoded by the human ether-a-gogo-related gene-1 (hERG1) by nitric oxide. Mol Pharmacol. 1999;56:1298–1308. doi: 10.1124/mol.56.6.1298. [DOI] [PubMed] [Google Scholar]
  • 162.Nunez L, Vaquero M, Gomez R, Caballero R, Mateos-Caceres P, Macaya C, Iriepa I, Galvez E, Lopez-Farre A, Tamargo J, Delpon E. Nitric oxide blocks hKv1.5 channels by S-nitrosylation and by a cyclic GMP-dependent mechanism. Cardiovasc Res. 2006;72:80–89. doi: 10.1016/j.cardiores.2006.06.021. [DOI] [PubMed] [Google Scholar]
  • 163.Maejima Y, Adachi S, Morikawa K, Ito H, Isobe M. Nitric oxide inhibits myocardial apoptosis by preventing caspase-3 activity via S-nitrosylation. J Mol Cell Cardiol. 2005;38:163–174. doi: 10.1016/j.yjmcc.2004.10.012. [DOI] [PubMed] [Google Scholar]
  • 164.Lai TS, Hausladen A, Slaughter TF, Eu JP, Stamler JS, Greenberg CS. Calcium regulates S-nitrosylation, denitrosylation, and activity of tissue transglutaminase. Biochemistry. 2001;40:4904–4910. doi: 10.1021/bi002321t. [DOI] [PubMed] [Google Scholar]
  • 165.Datta B, Tufnell-Barrett T, Bleasdale RA, Jones CJ, Beeton I, Paul V, Frenneaux M, James P. Red blood cell nitric oxide as an endocrine vasoregulator: a potential role in congestive heart failure. Circulation. 2004;109:1339–1342. doi: 10.1161/01.CIR.0000124450.07016.1D. [DOI] [PubMed] [Google Scholar]
  • 166.Gu Z, Kaul M, Yan B, Kridel SJ, Cui J, Strongin A, Smith JW, Liddington RC, Lipton SA. S-nitrosylation of matrix metalloproteinases: signaling pathway to neuronal cell death. Science. 2002;297:1186–1190. doi: 10.1126/science.1073634. [DOI] [PubMed] [Google Scholar]
  • 167.Carnes CA, Janssen PM, Ruehr ML, Nakayama H, Nakayama T, Haase H, Bauer JA, Chung MK, Fearon IM, Gillinov AM, Hamlin RL, Van Wagoner DR. Atrial glutathione content, calcium current, and contractility. J Biol Chem. 2007;282:28063–28073. doi: 10.1074/jbc.M704893200. [DOI] [PubMed] [Google Scholar]
  • 168.Carvalho-Filho MA, Ueno M, Hirabara SM, Seabra AB, Carvalheira JB, de Oliveira MG, Velloso LA, Curi R, Saad MJ. S-nitrosation of the insulin receptor, insulin receptor substrate 1, and protein kinase B/Akt: a novel mechanism of insulin resistance. Diabetes. 2005;54:959–967. doi: 10.2337/diabetes.54.4.959. [DOI] [PubMed] [Google Scholar]
  • 169.Pauli JR, Ropelle ER, Cintra DE, Carvalho-Filho MA, Moraes JC, De Souza CT, Velloso LA, Carvalheira JB, Saad MJ. Acute physical exercise reverses S-nitrosation of the insulin receptor, insulin receptor substrate 1 and protein kinase B/Akt in diet-induced obese Wistar rats. J Physiol. 2008;586:659–671. doi: 10.1113/jphysiol.2007.142414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Pearce LL, Gandley RE, Wasserloos K, Stitt M, Kanai AJ, McLaughlin MK, Pitt BR, Levitan ES. Role of metallothionein in nitric oxide signaling as revealed by a green fluorescent fusion protein. Proc Natl Acad Sci U S A. 2000;97:477–482. doi: 10.1073/pnas.97.1.477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Bernal PJ, Leelavanichkul K, Bauer E, Cao R, Wilson A, Wasserloos KJ, Watkins SC, Pitt BR, St Croix CM. Nitric-oxide-mediated zinc release contributes to hypoxic regulation of pulmonary vascular tone. Circ Res. 2008;102:1451–1454. doi: 10.1161/CIRCRESAHA.108.171264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Chvanov M, Gerasimenko OV, Peterson OH, Tepikin AV. Calcium-dependent release of NO from intracellular S-nitrosothiols. EMBO J. 2006;25:3024–3032. doi: 10.1038/sj.emboj.7601207. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES