当前位置: 首页 > 期刊 > 《动脉硬化血栓血管生物学》 > 2004年第11期 > 正文
编号:11330234
Reactive Oxygen Species
http://www.100md.com 《动脉硬化血栓血管生物学》
     From the Institute of Physiology (F.K., U.P.) and Cardiology Division (F.K., H.-Y.S.), Medizinische Poliklinik-Innenstadt, Ludwig-Maximilians-University, Munich, Germany.

    ABSTRACT

    Platelets participate not only in thrombus formation but also in the regulation of vessel tone, the development of atherosclerosis, angiogenesis, and in neointima formation after vessel wall injury. It is not surprising, therefore, that the platelet activation cascade (including receptor-mediated tethering to the endothelium, rolling, firm adhesion, aggregation, and thrombus formation) is tightly regulated. In addition to already well-defined platelet regulatory factors, such as nitric oxide (NO), prostacyclin (PGI2), and adenosine, reactive oxygen species (ROS) participate in the regulation of platelet activation. Although exogenously derived ROS are known to affect the regulation of platelet activation, recent data suggest that the platelets themselves generate ROS. Intracellular ROS signaling in activated platelets could be of significant relevance after transient platelet contact with the vessel wall, during the recruitment of additional platelets, and in thrombus formation. This review discusses the potential cellular and enzymatic sources of ROS in platelets, their molecular mechanisms of action in platelet activation, and summarizes in vitro and in vivo evidence for their physiological and potential therapeutic relevance.

    Platelets participate not only in thrombus formation but also in the regulation of vessel tone, the development of atherosclerosis, angiogenesis, and in neointima formation after vessel wall injury. It is not surprising, therefore, that the platelet activation cascade is tightly regulated. In addition to already well-defined platelet regulatory factors, such as nitric oxide (NO), prostacyclin (PGI2), and adenosine, reactive oxygen species (ROS) participate in the regulation of platelet activation. This review discusses the potential cellular and enzymatic sources of ROS in platelets, their molecular mechanisms of action in platelet activation, and summarizes in vitro and in vivo evidence for their physiological and potential therapeutic relevance.

    Key Words: platelets ? reactive oxygen species ? NAD(P)H-oxidase ? aggregation ? adhesion ? thrombus formation

    Introduction

    Platelet interaction with the vessel wall serves numerous physiological and pathophysiological functions. This is reflected by the fact that platelets release growth factors,1 lipid mediators,2,3 and cytokines.4 Consequently, the regulation of platelet activity plays a role not only for thrombus formation and the regulation of vascular tone5 but also for the vascular pathophysiology of angiogenesis and inflammation. Moreover, platelets participate in the development of atherothrombotic disease by promoting atherosclerotic lesion6 and neointima formation.7 Not surprisingly, the activation of platelets is regulated and modulated by numerous factors, blood-borne and cell-derived. Most of these factors are relatively well-characterized.8 However, in recent time, several publications have suggested that reactive oxygen species (ROS) represent a new modulator of platelet activity. It has been known for some time that ROS exert critical regulatory functions within the vascular wall and it is therefore plausible that platelets represent a relevant target for their action.

    Within the vessel wall (where endothelial cells, vascular smooth muscle cells, and fibroblasts express a variety of ROS-generating enzymes), there is a constant, low-quantity flux of ROS. It is already established that enhanced ROS release from the vascular wall can indirectly affect platelet activity by scavenging nitric oxide (NO), thereby decreasing the antiplatelet properties of the endothelium.9 In addition to their exposure to ROS derived from the vascular wall, platelets themselves also can generate ROS, and there is some evidence for a more direct role of ROS in the control of platelet activity. Finally, under inflammatory conditions, platelets are also exposed to phagocyte-dependent "burst-like" production of high quantities of ROS.10,11 With these circumstances in mind, it is surprising that our understanding of the influence of ROS on platelets lags far behind that of other soluble factors present in the vascular bed.

    This review focuses on current knowledge regarding ROS-dependent regulation of platelet activity and their role in platelet adhesion and aggregation in vitro and in vivo.

    ROS and Platelet Function

    Influence of Exogenous ROS on Platelet Function

    Although often referred to as a single entity, the term ROS actually comprises several factors that all potentially exert different vascular effects.12 The superoxide anion (O2–) is central to ROS chemistry, because it may be converted into other physiologically relevant ROS by enzymatic or nonenzymatic reactions. For example, O2– can react rapidly with NO to form peroxynitrite (ONOO–). It can also be converted to hydrogen peroxide (H2O2) by superoxide dismutase (SOD).12 Theoretically, both SOD and NO compete for O2– scavenging. Thus, it is worth mentioning that the rate of O2– conversion to H2O2 catalyzed by SOD is only one-third of the reaction rate of O2– with NO.13 H2O2 also serves as a substrate for the production of further detrimental ROS, such as hypochlorous acid (HOCl), generated by enzymatic conversion through neutrophil myeloperoxidase, and plasma levels of myeloperoxidase possess predictive value for coronary thrombotic events.14 Aside from enzymatic conversion, H2O2 reacts with ferrous iron (Fe2+), generating ferric iron (Fe3+) and the hydroxyl radical (OH–, "Fenton reaction").15 In the presence of catalase, H2O2 is degraded to water and oxygen. Aside from SOD or catalase, glutathione peroxidase (GPx) also exerts antioxidant enzymatic function, because it catalyzes a reaction that degrades H2O2 by oxidizing reduced glutathione (GSH) to its disulfide form (GSSG). Therefore, SOD, catalase, and GPx represent important antioxidant enzymes within vascular biology.

    Several in vitro experimental approaches have been used to investigate the specific roles for distinct ROS on platelets. A simple approach to identify the role of H2O2 has been to directly expose platelets to H2O2, which, unlike O2–, ONOO–, or OH–, is relatively stable, diffuses through membranes, and is available as a chemical reagent. Divergent results have been obtained from experiments using exogenous H2O2. It inhibited ADP-dependent platelet activation,16,17 whereas it enhanced collagen-dependent or arachidonic acid (AA)-dependent platelet activation.18 This discrepancy may be caused by the different antioxidant capacities of the buffers used.

    In addition to direct administration of H2O2, platelets have been exposed to O2–-generating enzymatic systems such as glucose/glucose oxidase, (hypo-)xanthine/xanthine oxidase (XO), or pyrogallol. Thereby, several independent studies have shown that O2– reduces the threshold for platelet activation to thrombin, collagen, ADP, or AA, and may even induce spontaneous aggregation.19–22 In all cases, it appeared that O2–, but not OH– or H2O2, mediated the enhanced platelet activity.21,22 In a more recent study, Pratico et al23 demonstrated that an alternative ROS, OH–, was generated during full blood platelet aggregation induced by collagen, and that the OH– scavenger mannitol partly prevented this aggregation, suggesting that >1 ROS may be involved in enhancing collagen-dependent platelet activation.

    In addition to directly activating platelets or decreasing the threshold for platelet activation, O2– reacts with platelet or endothelium-derived NO to ONOO–, which is of particular importance for vascular thrombosis. This is primarily caused by the decreased bioavailability of NO as a potent inhibitor of platelet activation. It is not surprising, therefore, that the antithrombotic effect of NO is lost either when ROS are exogenously added to the system or when they are scavenged by SOD.24,25 It is, however, likely that ONOO–, being a highly reactive molecule, exerts additional effects on platelets. Direct administration of peroxynitrite to platelets inhibits the aggregation of platelets induced by ADP, thrombin, collagen, and other platelet stimuli.26,27 However, ONOO– appears to possess a dual effect: whereas it activates platelets in normal buffer, this effect is reversed when plasma is added.27,28 In some cases, inhibitory effects of ONOO– are likely caused by the antiplatelet activity of S-nitrosothiols,29 which are formed when ONOO–-generating drugs like SIN-1 are applied.30 Unfortunately, a specific effect of ONOO– generated during the reaction of endothelial NO with O2– has not been investigated in mixed suspensions of these cells in vitro yet.

    Notably, results from in vitro exposure of platelets to exogenously applied H2O2, O2–, or ONOO– must be interpreted cautiously. First, the use of high, presumably nonphysiological concentrations of H2O2 or other ROS induces leakage of intracellular contents via membrane disruption (such effects may occur at H2O2 concentrations of 0.1 mmol/L and higher).31 Some reported inhibitory effects of H2O2 on platelet aggregation may have been caused by this effect.32,33 Second, the antioxidant capacity of the assay buffer must be considered when working with oxidants and platelets. Although studies with washed platelets allow for "unaltered" information about the function of a specific ROS, this scenario is not necessarily representative of physiological conditions, because plasma and full blood possess antioxidant capacities.34,35 However, studies of full blood may be biased by oxidants and antioxidants derived from cellular sources other than those from platelets. Finally, some data obtained in the presence of SOD, which is normally applied to identify the specific action of O2–, may also be explained by the effects of an increased concentration of H2O2, which is generated by the reaction of SOD with O2–. Thus, awareness of the assay conditions is of importance for the interpretation of data regarding the influence of ROS on platelets. If the influence of a specific ROS on platelets is investigated in vitro, then buffers that do not contain antioxidant substances may be appropriate. If consequences of such signaling should be assessed with relevance for the in vivo situation, however, then the effects of addition of defined amounts of plasma or full blood should be tested.

    Role of Platelet-Derived ROS

    In addition to ROS derived from exogenous sources, ROS are also generated by activated platelets. The release of O2– and other ROS by platelets was first observed by Marcus in 1977.36 In the following years, the release of several ROS, including O2–, OH–, and H2O2 from platelets was reported, both from unstimulated platelets and after platelet stimulation with agonists such as collagen or thrombin.37–39 This endogenous formation of ROS suggests they have autocrine or paracrine roles in platelet activation similar as described for exogenous ROS.

    Like endothelium-derived ROS, platelet-derived ROS potentially stem from different enzymatic sources (summarized in Figure 1), but one of them, the platelet isoform of NAD(P)H-oxidase, has gained most attention, because it can be activated on stimulation by agonists that also induce platelet activation.20,31,40–42 Evidence for the expression of an NAD(P)H-oxidase enzyme in platelets first arose from the observation that diphenylene iodonium chloride, a rather nonspecific inhibitor of flavoprotein-dependent enzymes (among them NAD(P)H-oxidase), inhibited platelet aggregation.43 Inhibition of platelet O2– formation by diphenylene iodonium chloride has been shown independently by several groups.44,45 The first direct evidence for the expression of a platelet NAD(P)H-oxidase was presented by Seno et al, who detected p22phox and p67phox, 2 subunits of the NAD(P)H-oxidase enzyme (there are at least 5 known subunits), in platelet lysates.46 The expression of a platelet isoform of NAD(P)H-oxidase was further underpinned by detection of the gp91phox and p47phox subunits.20,47 The specific inhibition of the gp91phox subunit could block platelet O2– production in a study that also demonstrated that collagen stimulation indeed increased platelet NAD(P)H-oxidase activity.20 Furthermore, NAD(P)H-oxidase–dependent platelet O2– production enhanced the recruitment of platelets to a growing thrombus, most likely by inactivating a platelet ectonucleotidase, thereby increasing bioavailability of ADP.20 Notably, the magnitude of NAD(P)H-oxidase derived O2– flux is in the nanomolar range,20 and thus it is similar to the flux that is present in endothelial cells48 but <1% of the amount of O2– released from activated neutrophils.49 When total release is measured in endothelial cell monolayers or defined platelet suspensions, again, levels of platelet O2– are similar to those produced by endothelial cells.20,48

    Figure 1. Enzymatic sources for reactive oxygen species (ROS) in the vascular endothelium and in platelets. Several enzymatic systems contribute to the production of ROS, thereby influencing platelet activity. In the endothelium (EC), NAD(P)H-oxidase, cyclooxygenase isoforms 1 and 2 (Cox), cytochrome P450 epoxygenase isoform 2C9 (CYP2C9), xanthine oxidase (XO), uncoupled endothelial NO synthase (unc. eNOS), and mitochondrial respiration contribute to the production of superoxide radicals (O2–), hydrogen peroxide (H2O2), and hydroxyl radicals (OH–). Platelets also contain NAD(P)H-oxidase activity, and many of its subunits have been found in platelets at the protein level (Nox2 or gp91phox, p22phox, p67phox, p47phox, and Rac). Platelet NAD(P)H-oxidase is stimulated on collagen (most likely mediated by GPVI) or thrombin exposure (most likely mediated by PAR1) and after membrane depolarization. It is unclear which of the Rac isoforms participates in platelet NAD(P)H-oxidase activation. Other potential sources for ROS from platelets are cyclooxygenase-1, xanthine oxidase, mitochondrial respiration, or uncoupled eNOS (when there is insufficient supply of cofactors such as tetrahydrobiopterin ). In addition, recent data indicate that there is redox potential-dependent regulation (GSH/GSSG ratio) of the platelet fibrinogen receptor (GPIIb/IIIa). Transparent symbols indicate that these enzymes have not been detected on a protein level to date, although there is pharmacological evidence for their existence. Note that macrophages, which are not shown in this Figure, are another important source for ROS with relevance to platelets.

    Recently, Clutton et al presented evidence that on stimulation with a thrombin-receptor activating peptide platelet phosphatidylinositol 3 kinase (PI3-kinase)-induced membrane translocation of platelet p67phox and thus regulated NAD(P)H-oxidase–dependent O2– release from platelets.50 Inhibition of phosphatidylinositol 3 kinase41,50 or protein kinase C20,39,41 have been found to decrease O2– release from platelets, whereas direct activation of protein kinase C resulted in enhanced platelet O2– production.20,39,41 Membrane depolarization also stimulated O2– production in platelets,51 similar as observed in endothelial cells, in which depolarization stimulated O2– production was found to be caused by a rac1-dependent activation of NAD(P)H-oxidase.48,52,53 Accordingly, we were able to show that platelet hyperpolarization prevented platelet-derived ROS release and platelet adhesion to endothelial cells in vitro.54

    Aside from NAD(P)H-oxidase, a variety of other enzymes is capable of producing ROS in platelets, as in most other cells of the vasculature. The endothelial isoform of NO synthase (eNOS) is a well-characterized source of O2– in endothelial cells when there is a shortage of cofactor supply ("uncoupling"),12 and there is evidence that eNOS may also be a source of platelet-derived O2– because platelets from eNOS-deficient mice show a markedly reduced O2– release.55 Furthermore, it has been shown that XO was not only present in platelets but also contributed to thrombin-induced platelet ROS production.39 Other observations imply an additional role for phospholipase A2 (PLA2)-dependent AA release contributing to ROS production.37 By analogy, this may be explained by the fact that the AA-converting prostaglandin H-synthase that also exists in platelets contains a cyclooxygenase and peroxidase moiety within its enzymatic complex,56 which can generate significant amounts of O2–.57 It is unknown, however, whether this really occurs in platelets. Other enzymes of AA metabolism, such as lipoxygenases, may also participate in platelet ROS release,58,59 but again, there is no evidence so far for a functional role of lipoxygenase or cyclooxygenase-dependent platelet ROS formation. Thus, in some cases, it remains to be shown that the release of ROS is not merely an epiphenomenon that does not necessarily play a functional role.

    Role of ROS in Platelet Signaling

    The hypothesis of an important role of platelet-derived ROS in platelet activation was initially supported by studies applying antioxidant enzymes to inhibit platelet activity. SOD, an O2–-degrading enzyme, has occasionally been observed to have an inhibitory effect on agonist-induced platelet activity when applied exogenously to platelet suspensions; however, the effect of SOD on platelets differed depending on the stimulus used.20,60

    Some of the conflicting data regarding the influence of platelet derived O2– on platelet aggregation can be explained by the kinetics of platelet ROS production after stimulus exposure. Platelet O2– production on collagen stimulation occurs with a delay of 3 to 5 minutes,20 so the involvement of platelet O2– in the initial steps of aggregation may be minimal. In the case of collagen, the aggregation starts only after a 1- to 2-minute lag phase after stimulus exposure. Although there was no effect of platelet-derived O2– in this initial phase of aggregation, late thrombus growth was significantly increased because O2– increased the bioavailability of ADP, leading to an increased recruitment of additional platelets.20 Similar results were reported after stimulating platelets with a thrombin receptor-activating peptide: the concomitant O2– release from platelets prevented late thrombus disaggregation that is normally caused by platelet-derived NO.50 There is accumulating evidence that platelet-derived O2– is a functionally relevant scavenger of platelet-derived NO, which inhibits the anti-aggregatory effects of platelet-derived NO.45,61 In accordance with this, GPIIb/IIIa inhibitors decrease platelet O2– release and enhance platelet NO release concomitantly.61 Some of these observations also indicate that an effect of platelet-derived ROS on platelet aggregation may play a prominent role in a late phase of aggregation. Although the mechanism by which O2– induces recruitment and enhanced platelet aggregation is not yet fully elucidated, it is likely that it involves scavenging of platelet NO by O2–,50 and/ or regulation of redox-sensitive ectonucleotidases located on platelet and endothelial cell membranes20,48,52,53 (Figure 2).

    Figure 2. Recruitment of platelets to a growing thrombus is enhanced by superoxide. When coming into close contact with subcellular matrix components (black curved lines), such as collagen, resting platelets (ovals) tether to the endothelium and are activated. Activated platelets (bizarre shapes) release substances, such as ADP (or thromboxane A2), that lead to recruitment of additional circulating resting platelets (ovals). On stimulation, platelets also release ROS such as superoxide (O2–), which increase thrombus formation either by preventing the degradation of ADP (thereby also lowering the bioavailability of the platelet inhibiting ultimate degradation product of ADP, adenosine) or by degrading bioavailability of nitric oxide (NO). NO, under physiological conditions, prevents platelet activation by increasing intraplatelet cGMP levels. Platelet release of O2– can be prevented by GPIIb/IIIa inhibitors.

    In studies using full blood, platelet activation on subthreshold exposure to collagen or AA was primed to result in irreversible aggregation when neutrophils were stimulated with N-formyl-Met-Leu-Phe.10 This effect could be prevented by the blockade of neutrophil NAD(P)H-oxidase and by catalase, supporting an additional role for H2O2.10,62 Of note, during the catalase reaction with H2O2, relevant amounts of the highly oxidized catalase intermediate "compound I" may be generated,63 which (similar to NO) is able to stimulate soluble guanylate cyclase.17,64 Such observations underpin the difficulties in identifying the specific effects of distinct ROS by using antioxidant enzymes.

    The fact that platelets express antioxidant enzymes65 also suggests a role of ROS in platelet signaling, because these antioxidant enzymes are likely not only to prevent cytotoxic effects of ROS but also to regulate oxidation sensitive signaling pathways in platelets. One of these enzymes, GPx, has gained notable attention with respect to arterial thrombus formation, because an innate deficiency of plasmatic GPx has been observed in 2 brothers with childhood cerebrovascular thrombosis.66 In addition to its antioxidant properties, GPx also enhances the bioavailability of NO by catalyzing its liberation from s-nitrosothiols and reducing lipid peroxides.67 By using GSH as a cosubstrate, GPx not only reduces H2O2 to water but also reduces lipid peroxides to their respective alcohols.68 Although it is not clear whether this reaction occurs in living cells, O2– can directly react with GSH, resulting in a chain reaction that ultimately produces oxidized glutathione (GSSG) and additional O2–.69,70 Thus, O2– has the potential to even directly decrease the equilibrium between reduced and oxidized glutathione (GSH/GSSG ratio), being of high importance for the redox regulation of protein thiol groups. The majority of these thiols exist in a reduced form when the GSH/GSSG ratio is high, which is the typical intracellular scenario, where the ratio amounts to 100/1.34 Shifting the intracellular redox balance to a lower GSH/GSSG ratio increases platelet sensitivity to activating agents, as demonstrated by increased intracellular calcium spiking.71 Consequently, platelet supplementation with GSH decreases platelet aggregation in vitro.72–74 Further, N-acetyl-L-cysteine, which is capable of restoring intracellular thiol stores by shifting the redox balance in favor of GSH, inhibits platelet aggregation and potentiates the antiplatelet effects of NO,75 underscoring the critical role for the platelets’ intracellular redox state.

    It is unclear at the moment which and how many proteins of importance for platelet function are potentially regulated by such redox-dependent mechanisms. There is, however, intriguing evidence that the IIb?3 integrin—a large transmembraneous heterodimeric glycoprotein receptor with a prominent role in platelet aggregation—is regulated by oxidants on several sites. The extracellular domains of IIb?3 are rich in disulfide bonds, which, on reduction, activate the receptor and induce a pro-aggregatory state, probably by stabilizing ligand–integrin interactions.76–78 The number of free thiols within IIb?3 is regulated by the extracellular GSH/GSSG ratio. Intriguingly, when the ratio is close to that in full blood (5/1), there is a maximum effect of agonist-induced aggregation.34 Recently, it has been described that platelets contain a glutathione reductase-like activity on their surface, which participates in regulating the extracellular GSH/GSSG ratio, and is capable of generating the appropriate redox potential for platelet activation.79 Similar mechanisms have been observed to regulate collagen-dependent platelet adhesion to the 2?1 integrin80 and are proposed to participate in inactivation of the platelet P2Y12 receptor.81

    Independent from this, another oxidative platelet activatory mechanism involves the short intracellular tail of ?3, which can be tyrosine phosphorylated on extracellular ligand binding.82 This has been suggested to represent an important step in platelet outside-in signaling, as it was associated with enhanced binding of tyrosine kinase adapter proteins such as GRB2.82 This tyrosine phosphorylation of ?3 could also be induced by H2O2 (in combination with a tyrosine phosphatase inhibitor), which resulted in spontaneous platelet aggregation.83

    Lastly, ex vivo studies indicate that antioxidant vitamins exert inhibitory actions on platelet function.84 Vitamin E (-tocopherol) inhibits platelet aggregation and the release reaction,85 suggesting that these effects were caused by a decreased O2– bioavailability in platelets.55 Vitamin C (ascorbic acid) and the red-wine phenolic antioxidant transresveratrol reportedly also reduced production of O2– in platelets.86 Recent reports also indicate that oral vitamin C administration can inhibit platelet aggregation ex vivo,87 and transresveratrol slightly reduces platelet aggregation and platelet lipid peroxidation.88 However, vitamins E and C can also exert pro-oxidant effects,86,89 and it remains unclear whether the administration of "potentially antioxidant" vitamins always exerts antithrombotic effects in vivo. Randomized clinical trials investigating a protective role for these compounds in cardiovascular disease have been disappointing thus far.90,91

    Thus, although redox regulation of platelet signaling is likely, a detailed understanding of redox-dependent platelet activatory signaling is at present limited to the data mentioned here.

    Role of ROS in Platelet Activation In Vivo

    Because of the concurrent presence of several distinct ROS, the actual influence of a specific ROS on platelet function in vivo may not necessarily match in vitro findings. Several lines of evidence, however, support the hypothesis that ROS are crucially involved in the physiology and pathophysiology of atherothrombosis in vivo.

    It is clearly evident that various risk factors for atherosclerosis and cardiovascular thrombosis, such as hypertension, hypercholesterinemia, hyperglycemia, cigarette smoking, or hyperhomocysteinemia, are associated with ROS-mediated development of endothelial dysfunction92,93 (a list is given in the Table). Because of their mass, platelets tend to flow outside the central high-velocity axis of laminar blood stream,94 so a large number of platelet interactions with the endothelium occur spontaneously, even under nonthrombotic conditions.95 Endothelial dysfunction or damage by oxidants96 are then associated with an enhanced risk for platelet activation and subsequent atherothrombotic complications.9,97 In addition, in vitro and in vivo evidence suggests that ROS, apart from directly activating platelets, increase their adhesion to the vascular endothelium by degrading the endothelial glycocalix,98 by acting on transcriptional mechanisms in the endothelium,13,99 by inactivating endothelial ectonucleotidases (enhancing the bioavailability of ADP52), or by an ischemia reperfusion-dependent endothelial cell activation.99

    Cardiovascular Risk Factors and Diseases Associated With Enhanced Platelet ROS Generation in Humans

    A direct involvement of ROS in thrombus formation in vivo has also been reported. It was first observed by Yao et al, who showed that recombinant CuZnSOD, recombinant MnSOD, and catalase infusion decreased cyclic flow variations (CFV), a correlate for in vivo platelet activation, in the coronary arteries of dogs.60 Corroborating an influence of ROS on platelet function in vivo, platelet recruitment to a growing thrombus was decreased after intravenous administration of SOD and catalase in rats, whereas it was enhanced by inhibition of NOS.100 Yao et al also showed that catalase and SOD prevented platelet-dependent CFV, suggesting that both O2– and H2O2 enhance thrombus formation in vivo. Strikingly, in animals in which catalase had no effect, SOD worsened CFV, underscoring a central role of high concentrations of H2O2.60 These results were later confirmed by Ikeda et al.101 In both studies, radical fluxes were generated by xanthine/XO-induced CFV; it was also observed that an inhibitor of endogenous XO, allopurinol, was able to prevent CFV induced by ADP.102 It was determined that O2– derived from uncoupled platelet eNOS was involved in platelet activation during CFV.103 Evidence for a significant role for other platelet sources of ROS, such as NAD(P)H-oxidase, in vivo is still limited: recently, hypercholesterinemic mice deficient of gp91phox, which is the large membrane-bound electron transferring complex of NAD(P)H-oxidase, were observed to have reduced P-selectin–dependent platelet rolling, compared with hypercholesterinemic controls.104 Whether an increase of platelet rolling was caused by endothelial, neutrophil, or platelet NAD(P)H-oxidase activity remained unclear in these experiments.

    Unfortunately, there are very few studies so far that give direct clues for the mechanisms underlying the prothrombotic effects of ROS in vivo. Several studies reported enhanced lipid peroxidation in association with platelet O2– formation and activation,42,105–108 but this is only a rather unspecific marker, similar as the reported decrease of plasmatic thiol levels associated with platelet O2– formation.109–111 Notably, plasma redox state is influenced by several plasmatic thiols such as homocysteine or cysteine; therefore, it is difficult to correlate the source of altered plasmatic redox state to thrombotic risk.

    Conclusions

    Overall, there is accumulating evidence supporting a net prothrombotic effect of vascular-derived and platelet-derived ROS in vitro. Several lines of evidence also suggest an oxidative enhancement of thrombus formation in vivo. Taken together, these data point to an important role of platelet-derived ROS and the intraplatelet redox state in the regulation of physiological platelet activation. In several cardiovascular disease states, augmented interactions of oxidants with platelets and the release of ROS from platelets have been observed, further supporting a causal role of enhanced ROS production in platelet pathophysiology.

    However, because of the complexity of ROS chemistry and because of the great number of regulatory enzymatic sources that are potentially involved, many questions about the role of ROS for platelet activation and the role of the different ROS-producing enzymatic systems remain unclear. In particular, the specific significance of the different types of ROS with respect to platelet activity has not definitely been determined. This is especially true for H2O2 and ONOO–, whereas the data concerning a stimulatory role for O2– or OH– are more consistent. The actual influence of ROS on biochemical redox levels also needs further attention because the GSH/GSSG ratio currently emerges as one potentially important regulator of platelet signaling. In addition, the specific role of oxidant-dependent platelet regulation and oxidant producing enzymes within platelets in comparison to other platelet regulatory mechanisms has not been evaluated. In some situations, it thus remains to be confirmed that platelet ROS production actually exerts a specific pathophysiological function, rather than simply existing as an epiphenomenon. However, because O2– fluxes from platelets are similar to those observed in endothelial cells, and because both cell types exhibit some similarity in terms of equipment with oxidant enzymes, it seems likely that the significance of platelet O2– is underestimated so far. This is even more so, because the role of platelets for atherogenic, inflammatory, or angiogenic processes, which are all likely to be influenced by platelet-derived ROS, is only gradually being defined.6,7,112–115 It will also be of significant interest to determine whether therapeutic intervention to reduce platelet ROS formation will be of use for antithrombotic treatment strategies, as suggested recently.61 Platelet ROS formation then could potentially be of use as a predictor for the progression of atherothrombotic disease.

    Although it is clearly evident that ROS participate in platelet activation and subsequent thrombus formation, it is not fully understood whether ROS serve as indispensable signaling molecules for specific platelet activation pathways. Possibly, the mere presence of ROS in the vicinity of platelets defines the conditions of the "physiological battleground" on which platelets encounter their physiological/pathophysiological destinations.

    Acknowledgments

    The authors thank Dr Darcy Lidington for critically reading this manuscript and for giving important suggestions for improvement.

    References

    English D, Garcia JG, Brindley DN. Platelet-released phospholipids link haemostasis and angiogenesis. Cardiovasc Res. 2001; 49: 588–599.

    Bolz SS, Pohl U. Highly effective non-viral gene transfer into vascular smooth muscle cells of cultured resistance arteries demonstrated by genetic inhibition of sphingosine-1-phosphate-induced vasoconstriction. J Vasc Res. 2003; 40: 399–405.

    Kaul S, Padgett RC, Heistad DD. Role of platelets and leukocytes in modulation of vascular tone. Ann N Y Acad Sci. 1994; 714: 122–135.

    Boehlen F, Clemetson KJ. Platelet chemokines and their receptors: what is their relevance to platelet storage and transfusion practice? Transfus Med. 2001; 11: 403–417.

    Agha A, Schluter H, Konig S, Biel K, Tepel M, Zidek W. A novel platelet-derived renal vasoconstrictor agent in normotensives and essential hypertensives. J Vasc Res. 1992; 29: 281–289.

    Huo Y, Schober A, Forlow SB, Smith DF, Hyman MC, Jung S, Littman DR, Weber C, Ley K. Circulating activated platelets exacerbate atherosclerosis in mice deficient in apolipoprotein E. Nat Med. 2003; 9: 61–67.

    Bienvenu JG, Tanguay JF, Chauvet P, Merhi Y. Relationship between platelets and neutrophil adhesion and neointimal growth after repeated arterial wall injury induced by angioplasty in pigs. J Vasc Res. 2001; 38: 153–162.

    Luscher TF. Platelet-vessel wall interaction: role of nitric oxide, prostaglandins and endothelins. Baillieres Clin Haematol. 1993; 6: 609–627.

    Loscalzo J. Oxidant stress: a key determinant of atherothrombosis. Biochem Soc Trans. 2003; 31: 1059–1061.

    Pratico D, Iuliano L, Alessandri C, Camastra C, Violi F. Polymorphonuclear leukocyte-derived O2-reactive species activate primed platelets in human whole blood. Am J Physiol. 1993; 264: H1582–H1587.

    Cerwinka WH, Cooper D, Krieglstein CF, Ross CR, McCord JM, Granger DN. Superoxide mediates endotoxin-induced platelet-endothelial cell adhesion in intestinal venules. Am J Physiol Heart Circ Physiol. 2003; 284: H535–H541.

    Wolin MS, Gupte SA, Oeckler RA. Superoxide in the vascular system. J Vasc Res. 2002; 39: 191–207.

    Wolin MS. Interactions of oxidants with vascular signaling systems. Arterioscler Thromb Vasc Biol. 2000; 20: 1430–1442.

    Brennan ML, Penn MS, Van Lente F, Nambi V, Shishehbor MH, Aviles RJ, Goormastic M, Pepoy ML, McErlean ES, Topol EJ, Nissen SE, Hazen SL. Prognostic value of myeloperoxidase in patients with chest pain. N Engl J Med. 2003; 349: 1595–1604.

    Winterbourn CC. Toxicity of iron and hydrogen peroxide: the Fenton reaction. Toxicol Lett. 1995; 82–83: 969–974.

    Stuart MJ, Holmsen H. Hydrogen peroxide, an inhibitor of platelet function: effect on adenine nucleotide metabolism, and the release reaction. Am J Hematol. 1977; 2: 53–63.

    Ambrosio G, Golino P, Pascucci I, Rosolowsky M, Campbell WB, DeClerck F, Tritto I, Chiariello M. Modulation of platelet function by reactive oxygen metabolites. Am J Physiol. 1994; 267: H308–H318.

    Pratico D, Iuliano L, Ghiselli A, Alessandri C, Violi F. Hydrogen peroxide as trigger of platelet aggregation. Haemostasis. 1991; 21: 169–174.

    Handin RI, Karabin R, Boxer GJ. Enhancement of platelet function by superoxide anion. J Clin Invest. 1977; 59: 959–965.

    Kr?tz F, Sohn HY, Gloe T, Zahler S, Riexinger T, Schiele TM, Becker BF, Theisen K, Klauss V, Pohl U. NAD(P)H-oxidase-dependent platelet superoxide anion release increases platelet recruitment. Blood. 2002; 100: 917–924.

    Salvemini D, de Nucci G, Sneddon JM, Vane JR. Superoxide anions enhance platelet adhesion and aggregation. Br J Pharmacol. 1989; 97: 1145–1150.

    De la Cruz JP, Garcia PJ, Sanchez DLC. Dipyridamole inhibits platelet aggregation induced by oxygen-derived free radicals. Thromb Res. 1992; 66: 277–285.

    Pratico D, Pasin M, Barry OP, Ghiselli A, Sabatino G, Iuliano L, Fitzgerald GA, Violi F. Iron-dependent human platelet activation and hydroxyl radical formation: involvement of protein kinase C. Circulation. 1999; 99: 3118–3124.

    Radomski MW, Palmer RM, Moncada S. Endogenous nitric oxide inhibits human platelet adhesion to vascular endothelium. Lancet. 1987; 2: 1057–1058.

    Sneddon JM, Vane JR. Endothelium-derived relaxing factor reduces platelet adhesion to bovine endothelial cells. Proc Natl Acad Sci U S A. 1988; 85: 2800–2804.

    Yin K, Lai PS, Rodriguez A, Spur BW, Wong PY. Antithrombotic effects of peroxynitrite: inhibition and reversal of aggregation in human platelets. Prostaglandins. 1995; 50: 169–178.

    Moro MA, Darley-Usmar VM, Goodwin DA, Read NG, Zamora-Pino R, Feelisch M, Radomski MW, Moncada S. Paradoxical fate and biological action of peroxynitrite on human platelets. Proc Natl Acad Sci U S A. 1994; 91: 6702–6706.

    Brown AS, Moro MA, Masse JM, Cramer EM, Radomski M, Darley-Usmar V. Nitric oxide-dependent and independent effects on human platelets treated with peroxynitrite. Cardiovasc Res. 1998; 40: 380–388.

    Maalej N, Albrecht R, Loscalzo J, Folts JD. The potent platelet inhibitory effects of S-nitrosated albumin coating of artificial surfaces. J Am Coll Cardiol. 1999; 33: 1408–1414.

    Mohr S, Stamler JS, Brune B. Mechanism of covalent modification of glyceraldehyde-3-phosphate dehydrogenase at its active site thiol by nitric oxide, peroxynitrite and related nitrosating agents. FEBS Lett. 1994; 348: 223–227.

    Iuliano L, Colavita AR, Leo R, Pratico D, Violi F. Oxygen free radicals and platelet activation. Free Radic Biol Med. 1997; 22: 999–1006.

    Belisario MA, Tafuri S, Di Domenico C, Squillacioti C, Della MR, Lucisano A, Staiano N. H(2)O(2) activity on platelet adhesion to fibrinogen and protein tyrosine phosphorylation. Biochim Biophys Acta. 2000; 1495: 183–193.

    Ohyashiki T, Kobayashi M, Matsui K. Oxygen-radical-mediated lipid peroxidation and inhibition of ADP-induced platelet aggregation. Arch Biochem Biophys. 1991; 288: 282–286.

    Essex DW, Li M. Redox control of platelet aggregation. Biochemistry. 2003; 42: 129–136.

    Anderson ME, Meister A. Dynamic state of glutathione in blood plasma. J Biol Chem. 1980; 255: 9530–9533.

    Marcus AJ, Silk ST, Safier LB, Ullman HL. Superoxide production and reducing activity in human platelets. J Clin Invest. 1977; 59: 149–158.

    Caccese D, Pratico D, Ghiselli A, Natoli S, Pignatelli P, Sanguigni V, Iuliano L, Violi F. Superoxide anion and hydroxyl radical release by collagen-induced platelet aggregation–role of arachidonic acid metabolism. Thromb Haemost. 2000; 83: 485–490.

    Finazzi-Agro A, Menichelli A, Persiani M, Biancini G, Del Principe D. Hydrogen peroxide release from human blood platelets. Biochim Biophys Acta. 1982; 718: 21–25.

    Wachowicz B, Olas B, Zbikowska HM, Buczynski A. Generation of reactive oxygen species in blood platelets. Platelets. 2002; 13: 175–182.

    Pignatelli P, Lenti L, Sanguigni V, Frati G, Simeoni I, Gazzaniga PP, Pulcinelli FM, Violi F. Carnitine inhibits arachidonic acid turnover, platelet function, and oxidative stress. Am J Physiol Heart Circ Physiol. 2003; 284: H41–H48.

    Zielinski T, Wachowicz B, Saluk-Juszczak J, Kaca W. The generation of superoxide anion in blood platelets in response to different forms of Proteus mirabilis lipopolysaccharide: effects of staurosporin, wortmannin, and indomethacin. Thromb Res. 2001; 103: 149–155.

    McVeigh GE, Hamilton P, Wilson M, Hanratty CG, Leahey WJ, Devine AB, Morgan DG, Dixon LJ, McGrath LT. Platelet nitric oxide and superoxide release during the development of nitrate tolerance: effect of supplemental ascorbate. Circulation. 2002; 106: 208–213.

    Salvemini D, Radziszewski W, Mollace V, Moore A, Willoughby D, Vane J. Diphenylene iodonium, an inhibitor of free radical formation, inhibits platelet aggregation. Eur J Pharmacol. 1991; 199: 15–18.

    Leo R, Pratico D, Iuliano L, Pulcinelli FM, Ghiselli A, Pignatelli P, Colavita AR, Fitzgerald GA, Violi F. Platelet activation by superoxide anion and hydroxyl radicals intrinsically generated by platelets that had undergone anoxia and then reoxygenated. Circulation. 1997; 95: 885–891.

    Tajima M, Sakagami H. Tetrahydrobiopterin impairs the action of endothelial nitric oxide via superoxide derived from platelets. Br J Pharmacol. 2000; 131: 958–964.

    Seno T, Inoue N, Gao D, Okuda M, Sumi Y, Matsui K, Yamada S, Hirata KI, Kawashima S, Tawa R, Imajoh-Ohmi S, Sakurai H, Yokoyama M. Involvement of NADH/NADPH oxidase in human platelet ROS production. Thromb Res. 2001; 103: 399–409.

    Pignatelli P, Sanguigni V, Lenti L, Ferro D, Finocchi A, Rossi P, Violi F. gp91phox-Dependent Expression of Platelet CD40 Ligand. Circulation. 2004;ePub ahead of Print.

    Sohn HY, Keller M, Gloe T, Morawietz H, Rueckschloss U, Pohl U. The small G-protein Rac mediates depolarization-induced superoxide formation in human endothelial cells. J Biol Chem. 2000; 275: 18745–18750.

    Lassegue B, Clempus RE. Vascular NAD(P)H-oxidases: specific features, expression, and regulation. Am J Physiol Regul Integr Comp Physiol. 2003; 285: R277–R297.

    Clutton P, Miermont A, Freedman JE. Regulation of endogenous reactive oxygen species in platelets can reverse aggregation. Arterioscler Thromb Vasc Biol. 2004; 24: 187–192.

    Kr?tz F, Riexinger T, Keller M, Sohn HY, Pohl U. 11,12-EETs hyperpolarize human platelets. In: Paul M. Vanhoutte, ed. EDHF 2002. London, UK: Taylor and Francis Publishing; 2003: 349–355.

    Kr?tz F, Sohn HY, Keller M, Gloe T, Bolz SS, Becker BF, Pohl U. Depolarization of endothelial cells enhances platelet aggregation through oxidative inactivation of endothelial NTPDase. Arterioscler Thromb Vasc Biol. 2002; 22: 2003–2009.

    Fisher AB, Al Mehdi AB, Wei Z, Song C, Manevich Y. Lung ischemia: endothelial cell signaling by reactive oxygen species. A progress report. Adv Exp Med Biol. 2003; 510: 343–347.

    Kr?tz F, Riexinger T, Buerkle MA, Nithipatikom K, Gloe T, Sohn HY, Campbell WB, Pohl U. Membrane Potential-Dependent Inhibition of Platelet Adhesion to Endothelial Cells by Epoxyeicosatrienoic Acids. Arterioscler Thromb Vasc Biol. 2004; 24: 595–600.

    Freedman JE, Li L, Sauter R, Keaney JF, Jr. -Tocopherol and protein kinase C inhibition enhance platelet-derived nitric oxide release. FASEB J. 2000; 14: 2377–2379.

    Fitzgerald GA. COX-2 and beyond: Approaches to prostaglandin inhibition in human disease. Nat Rev Drug Discov. 2003; 2: 879–890.

    Niwa K, Haensel C, Ross ME, Iadecola C. Cyclooxygenase-1 participates in selected vasodilator responses of the cerebral circulation. Circ Res. 2001; 88: 600–608.

    Singh D, Greenwald JE, Bianchine J, Metz EN, Sagone AL, Jr. Evidence for the generation of hydroxyl radical during arachidonic acid metabolism by human platelets. Am J Hematol. 1981; 11: 233–240.

    Jahn B, Hansch GM. Oxygen radical generation in human platelets: dependence on 12-lipoxygenase activity and on the glutathione cycle. Int Arch Allergy Appl Immunol. 1990; 93: 73–79.

    Yao SK, Ober JC, Gonenne A, Clubb FJ, Jr., Krishnaswami A, Ferguson JJ, Anderson HV, Gorecki M, Buja LM, Willerson JT. Active oxygen species play a role in mediating platelet aggregation and cyclic flow variations in severely stenosed and endothelium-injured coronary arteries. Circ Res. 1993; 73: 952–967.

    Chakrabarti S, Clutton P, Varghese S, Cox D, Mascelli MA, Freedman JE. Glycoprotein IIb/IIIa inhibition enhances platelet nitric oxide release. Thromb Res. 2004; 113: 225–233.

    Pignatelli P, Pulcinelli FM, Lenti L, Gazzaniga PP, Violi F. Hydrogen peroxide is involved in collagen-induced platelet activation. Blood. 1998; 91: 484–490.

    Burke TM, Wolin MS. Hydrogen peroxide elicits pulmonary arterial relaxation and guanylate cyclase activation. Am J Physiol. 1987; 252: H721–H732.

    Sabetkar M, Naseem KM, Tullett JM, Friebe A, Koesling D, Bruckdorfer KR. Synergism between nitric oxide and hydrogen peroxide in the inhibition of platelet function: the roles of soluble guanylyl cyclase and vasodilator-stimulated phosphoprotein. Nitric Oxide. 2001; 5: 233–242.

    Rey C, Vericel E, Nemoz G, Chen W, Chapuy P, Lagarde M. Purification and characterization of glutathione peroxidase from human blood platelets. Age-related changes in the enzyme. Biochim Biophys Acta. 1994; 1226: 219–224.

    Freedman JE, Loscalzo J, Benoit SE, Valeri CR, Barnard MR, Michelson AD. Decreased platelet inhibition by nitric oxide in two brothers with a history of arterial thrombosis. J Clin Invest. 1996; 97: 979–987.

    Freedman JE, Frei B, Welch GN, Loscalzo J. Glutathione peroxidase potentiates the inhibition of platelet function by S-nitrosothiols. J Clin Invest. 1995; 96: 394–400.

    Loscalzo J. Nitric oxide insufficiency, platelet activation, and arterial thrombosis. Circ Res. 2001; 88: 756–762.

    Winterbourn CC, Metodiewa D. The reaction of superoxide with reduced glutathione. Arch Biochem Biophys. 1994; 314: 284–290.

    Winterbourn CC, Metodiewa D. Reactivity of biologically important thiol compounds with superoxide and hydrogen peroxide. Free Radic Biol Med. 1999; 27: 322–328.

    van Gorp RM, Dam-Mieras MC, Hornstra G, Heemskerk JW. Effect of membrane-permeable sulfhydryl reagents and depletion of glutathione on calcium mobilisation in human platelets. Biochem Pharmacol. 1997; 53: 1533–1542.

    Pacchiarini L, Tua A, Grignani G. In vitro effect of reduced glutathione on platelet function. Haematologica. 1996; 81: 497–502.

    Thomas G, Skrinska V, Lucas FV, Schumacher OP. Platelet glutathione and thromboxane synthesis in diabetes. Diabetes. 1985; 34: 951–954.

    Matsuda S, Ikeda Y, Aoki M, Toyama K, Watanabe K, Ando Y. Role of reduced glutathione on platelet functions. Thromb Haemost. 1979; 42: 1324–1331.

    Loscalzo J. N-Acetylcysteine potentiates inhibition of platelet aggregation by nitroglycerin. J Clin Invest. 1985; 76: 703–708.

    Lahav J, Jurk K, Hess O, Barnes MJ, Farndale RW, Luboshitz J, Kehrel BE. Sustained integrin ligation involves extracellular free sulfhydryls and enzymatically catalyzed disulfide exchange. Blood. 2002; 100: 2472–2478.

    O’Neill S, Robinson A, Deering A, Ryan M, Fitzgerald DJ, Moran N. The platelet integrin IIb? 3 has an endogenous thiol isomerase activity. J Biol Chem. 2000; 275: 36984–36990.

    Walsh GM, Sheehan D, Kinsella A, Moran N, O’Neill S. Redox Modulation of Integrin (IIb)?(3) Involves a Novel Allosteric Regulation of Its Thiol Isomerase Activity. Biochemistry. 2004; 43: 2140.

    Essex DW, Li M, Feinman RD, Miller A. Platelet Surface Glutathione Reductase-Like Activity. Blood. 2004; 104: 1383–1385.

    Lahav J, Wijnen EM, Hess O, Hamaia SW, Griffiths D, Makris M, Knight CG, Essex DW, Farndale RW. Enzymatically catalyzed disulfide exchange is required for platelet adhesion to collagen via integrin 2?1. Blood. 2003; 102: 2085–2092.

    Ding Z, Kim S, Dorsam RT, Jin J, Kunapuli SP. Inactivation of the human P2Y12 receptor by thiol reagents requires interaction with both extracellular cysteine residues, Cys17 and Cys270. Blood. 2003; 101: 3908–3914.

    Law DA, Nannizzi-Alaimo L, Phillips DR. Outside-in integrin signal transduction. IIb ? 3-(GP IIb IIIa) tyrosine phosphorylation induced by platelet aggregation. J Biol Chem. 1996; 271: 10811–10815.

    Irani K, Pham Y, Coleman LD, Roos C, Cooke GE, Miodovnik A, Karim N, Wilhide CC, Bray PF, Goldschmidt-Clermont PJ. Priming of platelet IIb?3 by oxidants is associated with tyrosine phosphorylation of ?3. Arterioscler Thromb Vasc Biol. 1998; 18: 1698–1706.

    Keaney JF, Jr., Simon DI, Freedman JE. Vitamin E and vascular homeostasis: implications for atherosclerosis. FASEB J. 1999; 13: 965–975.

    Steiner M, Anastasi J. Vitamin E. An inhibitor of the platelet release reaction. J Clin Invest. 1976; 57: 732–737.

    Olas B, Wachowicz B. Resveratrol and vitamin C as antioxidants in blood platelets. Thromb Res. 2002; 106: 143–148.

    Wilkinson IB, Megson IL, MacCallum H, Sogo N, Cockcroft JR, Webb DJ. Oral vitamin C reduces arterial stiffness and platelet aggregation in humans. J Cardiovasc Pharmacol. 1999; 34: 690–693.

    Olas B, Wachowicz B, Stochmal A, Oleszek W. Anti-platelet effects of different phenolic compounds from Yucca schidigera Roezl. bark. Platelets. 2002; 13: 167–173.

    Abudu N, Miller JJ, Attaelmannan M, Levinson SS. Vitamins in human arteriosclerosis with emphasis on vitamin C and vitamin E. Clin Chim Acta. 2004; 339: 11–25.

    Clarke R, Armitage J. Antioxidant vitamins and risk of cardiovascular disease. Review of large-scale randomised trials. Cardiovasc Drugs Ther. 2002; 16: 411–415.

    Brown BG, Cheung MC, Lee AC, Zhao XQ, Chait A. Antioxidant vitamins and lipid therapy: end of a long romance? Arterioscler Thromb Vasc Biol. 2002; 22: 1535–1546.

    Cai H, Harrison DG. Endothelial dysfunction in cardiovascular diseases: the role of oxidant stress. Circ Res. 2000; 87: 840–844.

    Faraci FM, Lentz SR. Hyperhomocysteinemia, oxidative stress, and cerebral vascular dysfunction. Stroke. 2004; 35: 345–347.

    Tangelder GJ, Teirlinck HC, Slaaf DW, Reneman RS. Distribution of blood platelets flowing in arterioles. Am J Physiol. 1985; 248: H318–H323.

    Rosenblum WI. Platelet adhesion and aggregation without endothelial denudation or exposure of basal lamina and/or collagen. J Vasc Res. 1997; 34: 409–417.

    Shatos MA, Doherty JM, Hoak JC. Alterations in human vascular endothelial cell function by oxygen free radicals. Platelet adherence and prostacyclin release. Arterioscler Thromb. 1991; 11: 594–601.

    Bonetti PO, Lerman LO, Lerman A. Endothelial dysfunction: a marker of atherosclerotic risk. Arterioscler Thromb Vasc Biol. 2003; 23: 168–175.

    Vink H, Constantinescu AA, Spaan JA. Oxidized lipoproteins degrade the endothelial surface layer : implications for platelet-endothelial cell adhesion. Circulation. 2000; 101: 1500–1502.

    Cooper D, Stokes KY, Tailor A, Granger DN. Oxidative stress promotes blood cell-endothelial cell interactions in the microcirculation. Cardiovasc Toxicol. 2002; 2: 165–180.

    Peire MA, Puig-Parellada P. Oxygen-free radicals and nitric oxide are involved in the thrombus growth produced by iontophoresis of ADP. Pharmacol Res. 1998; 38: 353–356.

    Ikeda H, Koga Y, Oda T, Kuwano K, Nakayama H, Ueno T, Toshima H, Michael LH, Entman ML. Free oxygen radicals contribute to platelet aggregation and cyclic flow variations in stenosed and endothelium-injured canine coronary arteries. J Am Coll Cardiol. 1994; 24: 1749–1756.

    Kuwano K, Ikeda H, Oda T, Nakayama H, Koga Y, Toshima H, Imaizumi T. Xanthine oxidase mediates cyclic flow variations in a canine model of coronary arterial thrombosis. Am J Physiol. 1996; 270: H1993–H1999.

    Kanaya S, Ikeda H, Haramaki N, Murohara T, Imaizumi T. Intraplatelet tetrahydrobiopterin plays an important role in regulating canine coronary arterial thrombosis by modulating intraplatelet nitric oxide and superoxide generation. Circulation. 2001; 104: 2478–2484.

    Ishikawa M, Stokes KY, Zhang JH, Nanda A, Granger DN. Cerebral microvascular responses to hypercholesterolemia: roles of NADPH oxidase and P-selectin. Circ Res. 2004; 94: 239–244.

    McGrath LT, Dixon L, Morgan DR, McVeigh GE. Production of 8-epi prostaglandin F(2) in human platelets during administration of organic nitrates. J Am Coll Cardiol. 2002; 40: 820–825.

    Minuz P, Patrignani P, Gaino S, Seta F, Capone ML, Tacconelli S, Degan M, Faccini G, Fornasiero A, Talamini G, Tommasoli R, Arosio E, Santonastaso CL, Lechi A, Patrono C. Determinants of platelet activation in human essential hypertension. Hypertension. 2004; 43: 64–70.

    Minuz P, Patrignani P, Gaino S, Degan M, Menapace L, Tommasoli R, Seta F, Capone ML, Tacconelli S, Palatresi S, Bencini C, Del Vecchio C, Mansueto G, Arosio E, Santonastaso CL, Lechi A, Morganti A, Patrono C. Increased oxidative stress and platelet activation in patients with hypertension and renovascular disease. Circulation. 2002; 106: 2800–2805.

    Davi G, Ciabattoni G, Consoli A, Mezzetti A, Falco A, Santarone S, Pennese E, Vitacolonna E, Bucciarelli T, Costantini F, Capani F, Patrono C. In vivo formation of 8-iso-prostaglandin f2 and platelet activation in diabetes mellitus: effects of improved metabolic control and vitamin E supplementation. Circulation. 1999; 99: 224–229.

    Muscat JE, Kleinman W, Colosimo S, Muir A, Lazarus P, Park J, Richie JP, Jr. Enhanced protein glutathiolation and oxidative stress in cigarette smokers. Free Radic Biol Med. 2004; 36: 464–470.

    Haramaki N, Ikeda H, Takajo Y, Katoh A, Kanaya S, Shintani S, Haramaki R, Murohara T, Imaizumi T. Long-term smoking causes nitroglycerin resistance in platelets by depletion of intraplatelet glutathione. Arterioscler Thromb Vasc Biol. 2001; 21: 1852–1856.

    Takajo Y, Ikeda H, Haramaki N, Murohara T, Imaizumi T. Augmented oxidative stress of platelets in chronic smokers. Mechanisms of impaired platelet-derived nitric oxide bioactivity and augmented platelet aggregability. J Am Coll Cardiol. 2001; 38: 1320–1327.

    Wagner DD, Burger PC. Platelets in inflammation and thrombosis. Arterioscler Thromb Vasc Biol. 2003; 23: 2131–2137.

    Huo Y, Ley KF. Role of platelets in the development of atherosclerosis. Trends Cardiovasc Med. 2004; 14: 18–22.

    Brill A, Elinav H, Varon D. Differential role of platelet granular mediators in angiogenesis. Cardiovasc Res. 2004; 63: 226–235.

    Rhee JS, Black M, Schubert U, Fischer S, Morgenstern E, Hammes HP, Preissner KT. The functional role of blood platelet components in angiogenesis. Thromb Haemost. 2004; 92: 394–402.

    Schaeffer G, Wascher TC, Kostner GM, Graier WF. Alterations in platelet Ca2+ signalling in diabetic patients is due to increased formation of superoxide anions and reduced nitric oxide production. Diabetologia. 1999; 42: 167–176.

    Buczynski A, Wachowicz B, Kedziora-Kornatowska K, Tkaczewski W, Kedziora J. Changes in antioxidant enzymes activities, aggregability and malonyldialdehyde concentration in blood platelets from patients with coronary heart disease. Atherosclerosis. 1993; 100: 223–228.

    Kaur G, Pandey NR, Chandra M, Sanwal GG, Misra MK. Lipid peroxidation in human blood platelets during unstable angina and myocardial infarction. Boll Chim Farm. 2000; 139: 267–269.

    Pandey NR, Kaur G, Chandra M, Sanwal GG, Misra MK. Enzymatic oxidant and antioxidants of human blood platelets in unstable angina and myocardial infarction. Int J Cardiol. 2000; 76: 33–38.

    Chirkov YY, Holmes AS, Chirkova LP, Horowitz JD. Nitrate resistance in platelets from patients with stable angina pectoris. Circulation. 1999; 100: 129–134.

    Ellis GR, Anderson RA, Chirkov YY, Morris-Thurgood J, Jackson SK, Lewis MJ, Horowitz JD, Frenneaux MP. Acute effects of vitamin C on platelet responsiveness to nitric oxide donors and endothelial function in patients with chronic heart failure. J Cardiovasc Pharmacol. 2001; 37: 564–570.(Florian Kr?tz; Hae-Young )