当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 动脉硬化血栓血管生物学 > 2005年 > 第9期 > 正文
编号:11257734
Heme Oxygenase and Atherosclerosis
     From the Department of Laboratory Medicine, Toho University School of Medicine, Tokyo, Japan.

    Correspondence to Toshisuke Morita, Department of Laboratory Medicine, Toho University School of Medicine, 5-21-16 Omori-nishi, Ota-ku, Tokyo, 143-8540, Japan. E-mail toshimrt@med.toho-u.ac.jp

    Abstract

    Overproduction of reactive oxygen species under pathophysiological conditions, including dyslipidemia, hypertension, diabetes, and smoking, is integral in the development of cardiovascular diseases (CVD). The reactive oxygen species released from all types of vascular cells regulate various signaling pathways that mediate not only vascular inflammation in atherogenesis but also antioxidative and antiinflammatory responses. One such protective and stress-induced protein is heme oxygenase (HO). HO is the first rate-limiting enzyme in heme breakdown to generate equimolar quantities of carbon monoxide, biliverdin, and free ferrous iron. Accumulating evidence has shown that inducible HO (HO-1) and its products function as adaptive molecules against oxidative insults. The proposed mechanisms by which HO-1 exerts its cytoprotective effects include its abilities to degrade the pro-oxidative heme, to release biliverdin and subsequently convert it bilirubin, both of which have antioxidant properties, and to generate carbon monoxide, which has antiproliferative and antiinflammatory as well as vasodilatory properties. Herein, I highlight the relationship of HO and cardiovascular disease, especially atherosclerosis, gene-targeting approaches in animal models, and the potential for and concern about HO-1 as a novel therapeutic target for cardiovascular diseases.

    Heme oxygenase (HO) is the rate-limiting enzyme which catalyzes heme to carbon monoxide, biliverdin and free ferrous iron. Inducible HO (HO-1) functions as adaptive molecules against oxidative insults. Herein, I highlight the relationship of HO and atherosclerosis and the potential for HO-1 as a novel therapeutic target for cardiovascular diseases.

    Key Words: heme ? reactive oxygen species ? inflammation ? cytoprotection

    Introduction

    Well-established cardiovascular risk factors, such as dyslipidemia, hypertension, diabetes, and smoking, can initiate endothelial dysfunction by altering the cell redox state in the vessel wall and thus induce atherosclerosis. In this paradigm, a common mechanism by which cardiovascular risk factors initiate the disease process is oxidative stress, which leads to vascular inflammation. Excessive production of reactive oxygen species (ROS) has been implicated in many pathophysiological conditions such as oxidative stress.

    The oxidative modification of plasma low-density lipoproteins (LDL), mainly formed in the arterial wall, markedly increases their atherogenicity,1 and deposition of oxidized LDL in arterial walls is central to the development of an atherosclerotic lesion.2 LDL, which is initially accumulated in the extracellular subendothelial space in arterial walls, is mildly oxidized to minimally modified LDL, which in turn induces monocyte chemotactic protein 1 and macrophage colony-stimulating factor, both of which stimulate monocyte recruitment and differentiation to macrophages in arterial walls.3 The accumulation of monocytes/macrophages in arterial walls generates ROS, resulting in the conversion of minimally modified LDL into oxidized LDL. Because oxidized LDL-elicited dysfunction of endothelial cells is one of the earliest steps in the development of atherosclerosis, the adaptive and/or protective responses of the vasculature to oxidative stress are important in preventing atherogenesis.

    One such protective and stress-induced protein is heme oxygenase (HO). HO is the rate-limiting enzyme in heme degradation.4 For much of the time, HO was studied within the confines of heme metabolism. After HO-1, an inducible form of HO, was found to be readily upregulated by diverse factors associated with various cardiovascular diseases, however, many studies have been reported and accumulating evidence suggests that HO-1 induction confers protection against injuries associated with ROS.

    In this article, I review the enzymatic aspects of HO, the biological properties of HO reaction products, the possible involvement of HO in atherosclerosis, and the therapeutic application of HO for atherosclerosis in humans by directly delivering or pharmacologically activating the HO-1 gene.

    Heme and Heme Oxygenase

    Heme is a crucial substance that is absolutely required for aerobic life. Free heme can be quite cytotoxic, however, particularly in the presence of oxidants or activated phagocytes. Of all sites in the body, the vasculature may be at greatest risk of exposure to free heme, as erythrocytes contain heme at a concentration of 20 mmol/L and are prone to undergoing unexpected lysis. Extracellular hemoglobin is easily oxidized and readily releases heme. Given the hydrophobic nature of heme, it can rapidly intercalate with membranes and cause severe damage. Exposure of microsomal membrane fractions to free heme in the presence of H2O2 leads to a dose-dependent peroxidation of unsaturated lipids in vitro.5 Heme also catalyzes the protein peroxidation of susceptible groups in membranes, serum, or cytosolic proteins, leading to the formation of nonreducible covalent cross-linkages or protein fragmentation. Heme has been shown to be involved in the pathogenesis of atherosclerosis, which is considered to be an inflammatory disease characterized by the accumulation of lipids and fibrous material in arterial walls. Hemoglobin-derived heme has been demonstrated to act as a catalyst for the oxidation of LDL, which in turn causes atherosclerosis.6 Wagener and colleagues7 observed heme-induced expression of adhesion molecules, such as intercellular adhesion molecule-1, vascular adhesion molecule-1, and E-selectin, in cultured endothelial cells.

    HO is a rate-limiting and microsomal enzyme that participates in heme breakdown to generate equimolar amounts of biliverdin IXa, free ferrous iron, and carbon monoxide (CO). Subsequently, biliverdin is rapidly converted to bilirubin by biliverdin reductase, and free iron is sequestered by ferritin4,8,9 (Figure 1). This reaction requires oxygen and occurs in association with NADPH-ferrihemoprotein reductase (cytochrome reductase), which supplies electrons to reduce heme iron or to activate molecular oxygen.10 Three isoforms of HO that are the products of separate genes have been identified in mammals.11–13 HO-1 is transcriptionally upregulated as a sensitive antiinflammatory protein by various types of oxidative stress, such as oxidized LDL,14 ultra-violet radiation,15 thiol scavengers,16 and hypoxia,17,18 as well as substrate heme19 in the cardiovascular system. A common feature of these HO-1 inducers is their ability to regulate the intracellular redox state. HO-1 is also transcriptionally activated through several regulatory mechanisms. Studies on the promoter region of HO-1 have revealed transcriptionally responsive elements, including activator protein I, activator protein II, nuclear factor-B, interleukin-6–responsive elements, and an antioxidant response element (ARE).20–22 HO-2 is constitutively expressed in many organs throughout the body, although it is particularly high in the brain and testes, but is unresponsive to any of the inducers of HO-1.11 The third isoform, HO-3, is nearly devoid of catalytic activity;13 however, Hayashi and colleagues23 recently reported that HO-3 is a pseudo gene derived from HO-2 transcripts.

    Figure 1. The heme oxygenase enzyme reaction. Schema of the catalytic conversion of heme into bilirubin, carbon monoxide (CO) and iron.

    Potential Roles of HO Reaction Products on Vascular Function

    HO and the subsequent metabolites of heme catabolism seem to play vital roles in regulating important biological responses, including inflammation, oxidative stress, cell survival, and cell proliferation. The proposed mechanisms by which HO-1 exerts its biological effects include its ability to release biliverdin with its subsequent conversion to bilirubin and to generate ferritin and CO, as well as to decompose the pro-oxidative heme.

    Biliverdin and Bilirubin

    Biliverdin and bilirubin have been considered to be waste products of heme degradation. These products can be toxic at high concentrations, especially in neonates.24 Little attention had been paid to the physiological roles of these products until the observation that bilirubin suppresses the oxidation of chemically induced peroxyl radicals to a greater extent than -tocopherol, which provides powerful protection against lipid peroxidation in vitro.25 After bilirubin was recognized as an antioxidant, many epidemiological studies demonstrated that higher serum levels of bilirubin are inversely associated with the incidence of coronary arterial disease.26–29 The antiatherogenic properties of bilirubin may include inhibitory effects against LDL oxidation and scavenging of oxygen radicals, which are generated from phospholipids, triglycerides, and cholesterol esters.30 Indeed, Ishikawa and colleagues31 demonstrated that HO inhibition resulted in enhanced formation of lipid peroxides in aortic tissue in Watanabe heritable hyperlipidemic (WHHL) rabbits. They speculated that HO reaction products like biliverdin and bilirubin might reduce the formation of plasma and tissue lipid peroxidation products.

    The mechanism by which bilirubin reacts with ROS is not completely understood, although its hydrophobic tetrapyrrole structure has been reported to inhibit the activation of superoxide-producing NADPH oxidase.32 Bilirubin has also been shown to have an inhibitory effect on protein phosphorylation33 and protein kinase C activity,34 both of which lead to the inactivation of various proatherogenic factors, such as lipid oxidation,35 immune reactions,36 and inflammatory processes.37 Oxidative metabolites of bilirubin have recently been detected in atherosclerotic lesions.38 The physiological role of neonatal hyperbilirubinemia has been assumed to be the result of the inhibitory action of bilirubin on lipid and protein peroxidation.39 Oxidized metabolites of bilirubin (biopyrrins) have been reported to be excreted in urine after exposure to oxidative stress.40 The antioxidant actions of bilirubin have been demonstrated in vivo in a variety of diseases, including Alzheimer disease, ischemia-reperfusion injury, and coronary artery diseases.41,42 These observations suggest that increased bilirubin production is an adaptive response against oxidative insults.

    Carbon Monoxide

    CO is another product of the HO reaction in vivo, and the similarities between nitric oxide (NO) and CO have suggested that CO may have a physiological role.43 Earlier studies have suggested that CO might bind to and activate guanylate cyclase, thereby increasing intracellular levels of cyclic guanine monophosphate (cGMP), as has been demonstrated for NO.4,18,44 However, the physiological relevance of CO as a vasodilator is controversial. NO is a potent activator of guanylate cyclase and increases cGMP production in vitro 130 fold, whereas the increase in cGMP induced by CO is 4.4 fold.45,46 Because CO is more chemically stable than NO and there appears to be no enzymatic pathway catalyzing CO degradation in vivo, the biological availability of NO and CO may differ.4 CO has been suggested to act as a partial antagonist for soluble guanylate cyclase (sGC) activation.47–49 It has been posited that overproduction of CO might impair NO-elicited activation of sGC, resulting in suppression of the cGMP increase in the aortas of transgenic mice that overexpressed HO-1 in vascular smooth muscle cells.49 HO activity, as measured by CO production, has been detected in arteries of rodents and humans in vivo.50 Some reports have suggested that CO may act as a physiological regulator of vascular tone through cGMP-mediated responses in large vessels (eg, aorta),46,51 whereas CO may dilate the smaller renal arteries or pial arterioles via activation of Ca2+-activated potassium channels.52,53 Other reports have hypothesized that CO may play a role in blood pressure regulation in acute hypertension.54 In addition to relaxing vascular smooth muscle, CO has been shown to have a regulatory effect on smooth muscle cell proliferation and death,55–58 which may play an important negative feedback role in limiting the development of vascular diseases.

    Recently, CO has been reported to exert cGMP-independent actions on Ca-dependent potassium channels53,59 and on the mitogen-activated protein (MAP) kinase.60,61 CO directly activates Ca-dependent potassium channels, leading to vascular relaxation through Ca desensitization. Another cGMP-independent action of CO is the activation of the p38 MAP kinase pathway. The inhibitory effect of CO against lipopolysaccharide-induced proinflammatory cytokine production in macrophages and its antiproliferative effect on vascular smooth muscle cells after balloon injury are mediated by the p38 MAP kinase pathway, including MAP kinases 3 and 6. The effect of CO on p38 MAP kinase and caspase 3 has been reported in pulmonary artery endothelial cells, indicating that CO has antiapototic effects and reduces ischemia-reperfusion lung injury.62 In addition, p38 MAP kinase pathway activation by CO has been observed to protect endothelial cells from tumor necrosis factor-–mediated apoptosis.63 Thus, the actions of CO via cGMP-dependent and cGMP-independent pathways may explain a number of the potential actions of CO regarding the pathogenesis of cardiovascular diseases.

    Ferritin and Iron

    Ferritin is a ubiquitous and highly conserved iron-binding protein. In vertebrates, the cytosolic form consists of 2 subunits, termed H and L. Twenty-four ferritin subunits assemble to form the apoferritin shell.64 In the ferritin shell, the proportion of H and L depends on the iron status in the cell or tissue and varies among organs and species. Ferritin H catalyzes the oxidation of ferrous iron to ferric iron under aerobic conditions to allow intracellular iron storage in biological systems. The content of cytoplasmic ferritin is regulated by the translation of ferritin H and L mRNAs in response to the content of intracellular free iron.65,66 When iron levels are low, ferritin synthesis decreases; conversely, when iron levels are high, ferritin synthesis increases. Although under certain circumstances there is an increase in ferritin mRNA in response to iron,67 the regulatory response of ferritin to iron is largely posttranscriptional68 and is due to the recruitment of stored mRNA from monosomes to polysomes in the presence of iron.66 The induction of ferritin has been shown to provide marked antioxidant cellular protection by rapidly sequestering free cytosolic iron, the crucial catalyst of oxygen-centered radical formation via the Fenton reaction in biological systems.

    HO-1 releases Fe2+ from the core of the heme molecule, leading to the rapid expression of the protein ferritin in tandem with HO-1.69 Ferritin H is induced after HO-1 induction in vascular smooth muscle cells.70 Furthermore, the enhancement of intracellular ferritin protein through HO-1 has been reported to reduce the cytotoxic effects of hemin and hydrogen peroxide in vascular endothelial cells.71 Additionally, increased expression of ferritin mRNA has been shown in a rat model of brain ischemia/reperfusion injury,72 in pericardial fluid of patients with coronary disease,73 and in human atherosclerotic aortas with abdominal aneurysm.74 Hoekstra and colleagues75,76 revealed that aortic endothelial cells in a strain of atherosclerosis-susceptible Japanese quail were more susceptible to oxidative stress than those from an atherosclerosis-resistant strain, which coincided with higher HO activity, HO-1 expression, ferritin, and glutathione levels. These findings suggest that the cytoprotective effect of ferritin production after HO-1 induction on ROS-induced cellular injury may act to limit intracellular-free iron.

    HO-1 Induction by ROS

    Well-established cardiovascular risk factors, such as dyslipidemia, hypertension, diabetes, and smoking, increase ROS production in arterial wall cells. In endothelial cells as well as in vascular smooth muscle cells and monocytes/macrophages, all of which constitute atherosclerotic lesions, ROS generated via several enzymatic reactions initiate and promote the atherogenic process. Xanthine oxidase,77 NAD(P)H oxidase,78 uncoupled NO synthase,79 and myeloperoxidase80 are well-known ROS-generating enzymes expressed in atherosclerotic lesions. Among these enzymes, NAD(P)H oxidase is a key player in generating ROS in the vasculature, and it produces superoxide anion, one of the important reactive oxygen species, by catalyzing a 1-electron reduction of oxygen using NAD(P)H as the electron donor. ROS affect many proatherogenic processes such as lipoprotein oxidation, endothelial apoptosis, and matrix metalloproteinase expression and cause atherosclerosis, restenosis, and diabetes mellitus vasculopathy. ROS production activates several transcription factors such as nuclear factor-B, activator protein 1, and early growth response gene 1 (Egr-1) in vascular walls, leading to the initiation of proinflammatory processes.81,82 By contrast, cells exposed to ROS also induce a group of antioxidative genes via the ARE and Nrf2, a transcription factor for ARE.83,84

    Stimulation of the HO-1 gene is primarily controlled at the transcriptional level, which is governed by responsive elements localized in the promoter 5'-flanking region of the HO-1 gene. Studies of the promoter region of HO-1 have revealed transcriptionally responsive elements, including nuclear factor-B, activator protein 1, activator protein 2, interleukin-6–responsive elements and ARE.20,21,22,85 Although considerable information is available on the inducible responsive elements of the HO-1 gene, much less is known about the transcriptional factors that actually mediate this induction. The transcriptional factor Nrf2, which belongs to the family of Cap‘n’Collar basic leucine zipper proteins, has previously been shown to be essential for gene induction mediated by ARE86 and serves as a key transcription factor in the cytoprotection of tissues against oxidative stress. Nrf2 acts in a wide-ranging metabolic response to oxidative stress and constitutes a cellular sensor for oxidative stress through a nuclear shuttling mechanism with the cytosolic regulator protein Keap I.87 Alam and colleagues88 have demonstrated that a DNA sequence containing an activator protein 1 consensus site is regulated by Nrf2 and that Nrf2 mediates the induction of the HO-1 gene. The important role of Nrf2 in the stress-dependent induction of HO-1 has been confirmed by the finding that HO-1 is not inducible in Nrf2 null mice.89 More recently, Ishii and colleagues90,91 have demonstrated that oxidized LDL and 4-hydroxynonenal increased HO-1 expression in murine macrophages, whereas they failed to increase it in Nrf2-deficient macrophages.

    ??Oxidized LDL Accumulation and HO-1 Expression in Atherosclerotic

    ??

    As noted earlier, numerous studies have shown that oxidized LDL plays a pivotal role in the development of atherosclerosis. Development of fatty-streak lesions is initiated with the production of adhesion molecules, monocyte chemotactic factors, and vascular smooth muscle growth factors. As the atherosclerosis lesions progress, migration and proliferation of smooth muscle cells and deposition of fibrous tissue lead to an advanced, complicated lesion. HO-1 in vascular endothelial cells, vascular smooth muscle cells, and macrophages is markedly upregulated by oxidized LDL, whereas HO-1 is not increased in vascular endothelial cells or smooth muscle cells when exposed to native LDL.14,92,93 The component in oxidized LDL responsible for HO-1 induction appears to be oxidized arachidonic acid-containing phospholipids, such as 1-palmitoyl-2-isoprostanoyl-sn-glycero-3-phosphorylcholine92 and linoleyl hydroperoxide,94 but not lysophosphatidylcholine.14 HO-1 expression was observed throughout the development of the lesions, from an early fatty streak to an advanced complex atherosclerotic lesion14,31; that is, HO-1 is expressed in vascular endothelial cells and macrophages in the early stages of atherosclerotic lesion formation and in foam cells and smooth muscle cells residing in the necrotic core of advanced lesions. In all these atherosclerotic lesions, HO-1 has been found to be colocalized with oxidized phospholipids,31,95 strongly suggesting that HO-1 is induced by oxidized phospholipids in vivo. The fact that HO-1 was detected in early atherosclerotic lesions suggests that the changes may be the cause rather than a consequence.

    Antiatherogenic Properties of HO-1

    In Vitro Study

    Atherosclerosis is thought to be accelerated when the antioxidative defense is not sufficiently induced in the vascular wall under oxidative and inflammatory stress. Accumulating evidence suggests that HO-1 contributes to the balance of pro-oxidant and antioxidant factors in the vascular wall through multiple mechanisms. Taille et al96 explored the role of HO-1 in the expression and activity of NADP(H) oxidase, the key player in generating ROS in the vasculature. They revealed that HO-1 expression by HO-1–encoded plasmid transfection in macrophages resulted in a decrease of NAD(P)H oxidase activity and expression of gp91(phox), its heme-containing catalytic component, because reduction of the heme content due to HO-1 activation limited heme availability for maturation of the gp91(phox) subunit and assembly of the functional NAD(P)H oxidase.96 Endothelial dysfunction caused by ROS is a key step in the initiation of lesion formation. Abraham et al97 revealed that increased expression of HO-1 attenuated ROS-mediated cell growth and apoptosis of endothelial cells. The protective effect of the HO-1 signaling mechanism, which attenuated glucose (ROS)-mediated endothelial cell apoptosis, was maintained through suppression of p21 and p27 cyclin kinase inhibitors or, perhaps, through its powerful antioxidant effect. Monocyte transmigration into the vessel wall is an essential event for the development of atherosclerosis. Ishikawa et al92 demonstrated that HO-1 induction significantly inhibited monocyte transmigration induced by oxidized LDL, whereas HO inhibition had the opposite effect. The inhibitory effect of HO-1 on monocyte transmigration could be mediated by biliverdin and bilirubin, as these compounds can directly inhibit monocyte transmigration. We confirmed the role of bilirubin in monocyte transmigration.98 My colleagues and I also revealed that angiotensin II increased CC chemokine receptor 2 (CCR2) expression in monocytes and increased ROS production, leading to enhanced chemotactic activity. However, HO-1–induced monocytes with increased accumulation of bilirubin exhibited enhanced CCR2 expression, as well as ROS production and chemotactic activity induced by angiotensin II. Similar inhibitory action of bilirubin on leukocyte adhesion to activated vascular endothelial cells has been reported.99

    Smooth muscle cell proliferation and monocytes recruitment are essential steps for the development of atherosclerosis. My colleagues and I have previously investigated the role of HO-1 in hypoxia-induced growth factor expression in vitro.55 My colleagues and I demonstrated that HO-1 was induced by hypoxia in vascular smooth muscle cells, and that the smooth muscle cells–derived CO inhibited the endothelial cells production of endothelin 1, platelet-derived growth factor-B, and vascular endothelial growth factor. The inhibition of these factors by CO led to a decrease in vascular smooth muscle cells proliferation. In addition, smooth muscle cell–derived CO directly decreased vascular smooth muscle cells growth by inhibiting E2F-1, a transcription factor that participates in the control of cell cycle progression from G1 to S phase.56 These reports suggest that CO, by limiting vascular smooth muscle cell growth, may be a key mediator in the body’s compensatory response to vascular remodeling associated with hypoxia.

    In Vivo Animal Study

    Several investigators have also demonstrated that endogenous HO-1 plays a role in atherosclerotic lesion formation based on the fact that chronic inhibition of the HO system using metalloporphyrins promoted exaggerated lesion formation in a rat model of balloon injury, LDL-receptor knockout mice, and Watanabe heritable hyperlipidemic rabbits. My colleagues and I previously demonstrated that HO-1 was upregulated in rat carotid artery after endothelial denudation using a balloon catheter and that inhibition of endogenous HO-1 by metalloporphyrins promoted neointimal formation, whereas HO induction by hemin resulted in a reduction of neointimal formation.100 In LDL receptor knockout mice, HO-1 is abundantly expressed in atherosclerotic lesions after dietary cholesterol feeding.85 In this mouse model, HO overexpression resulted in a reduction of plaque formation, whereas HO inhibition by Sn-protoporphyrin IX promoted lesion development. In the WHHL rabbit model, inhibition of HO-1 by Sn-protoporphyrin IX increased plasma and tissue lipid peroxide levels atherosclerotic plaque formation.95 These studies provided insight regarding the role of endogenous HO enzymes and their therapeutic application against atherosclerotic lesion formation; however, confirmation of this response in HO-1–deficient mice is critical because the effects of metalloporphyrins are not selective for HO-1 or even the HO system.101 Metalloporphyrins have been reported to inhibit NOS and sGC in addition to heme oxygenase.101,102 Other nonspecific effects of metalloporphyrins include the inhibition of muscle relaxation and suppression of cyclic adenosine monophosphate and cGMP.103 These non-HO activity-related inhibitory effects may be consequential to the interaction of metalloporphyrins with membrane receptors or their downstream signal transduction pathways. Heme oxygenase inducers are also known to exert an adverse effect when used especially for chronic upregulation of HO-1. The disadvantage of using SnCl2, a synthetic substance with unique selectivity for renal heme oxygenase, is its nephrotoxicity. The application of CoCl2 was limited by its excessive toxicity and high mortality found in the animal study. Hemin, the most prominent and frequently used HO inducer, increases oxidative stress. Heme arginate increases not only the production of CO from HO-1, but also that of NO, as arginate is a derivative of L-arginine.

    Recent studies, therefore, have explored the strategy of administering the HO gene with a vector and generating transgenic mice to investigate the long-term and specific protective effects of HO-1 against atherosclerosis and vascular injury. Abraham et al104 demonstrated that retrovirus-mediated HO-1 overexpression reduced the number of detached endothelial cells in a rat model of diabetes, indicating that endothelial HO-1 protects against ROS-mediated endothelial damage. Overexpression and underexpression of human HO in endothelial cells can be modulated on a long-term basis by the introduction of the HO-1 gene with a retroviral vector.105 The findings that overexpression of the HO-1 gene in the rat carotid artery reduced neointimal formation and inhibited medial-wall DNA synthesis after balloon injury further suggested that adenovirus-mediated HO-1 gene delivery has a potential therapeutic application.106 Juan and colleagues107 demonstrated that selective overexpression of HO-1 was also able to decrease lesion formation in apolipoprotein-E (ApoE)–deficient mice using adenovirus-mediated gene transfer of HO-1. Duckers et al108 showed that adenovirus-mediated HO-1 introduction to arterial walls directly reduced vasoconstriction and inhibited vascular cell proliferation in a pig model of vascular injury. They revealed that the vasorelaxation response was mediated by guanylate cyclase and cGMP, whereas growth inhibition and cell-cycle arrest in cells expressing the HO-1 transgene were associated with the induction of p21cip1. Yet et al generated mice deficient in both HO-1 and ApoE. Despite similarly elevated total plasma cholesterol levels in response to a Western diet, mice deficient in both HO-1 and ApoE developed larger and more advanced atherosclerotic lesions than mice deficient in ApoE alone.109 In a mouse model of vein graft stenosis, vein grafts from HO-1 knockout mice showed a robust neointima formed by -actin–positive vascular smooth muscle cells as compared with that in vein grafts from wild-type mice. Further studies revealed that vascular smooth muscle cells from HO-1 knockout mice were more susceptible to oxidant H2O2-induced cell death than those isolated from wild type mice, suggesting that increased susceptibility to oxidative stress may lead to cell death in HO-1 knockout mice.109

    To confine HO-1 gene delivery to vascular smooth muscle cells, My colleagues and I generated transgenic mice which overexpress HO-1 under the control of the SM22 promoter, which is specifically expressed in vascular smooth muscle cells.49 Using these transgenic mice, my colleagues and I investigated the pathophysiological role of HO-1 in vascular smooth muscle cells after arterial injury induced by angiotensin II combined with high salt.110 We demonstrated that wild-type mice exhibited increased production of ROS and inflammatory changes such as expression of inflammatory cytokines and macrophage infiltration in and around the coronary artery, whereas HO-1 transgenic mice did not show any changes in ROS production or inflammatory responses. Interestingly, in vascular smooth muscle cells, HO-1 reduced ROS production not only in vascular smooth muscle cells but also in adjacent endothelial cells and cardiomyocytes.

    Human Cases

    The importance of HO-1 in vascular biology was highlighted by the discovery of a child with HO-1 deficiency.111,112 In this HO-1–deficient child, both intravascular hemolysis and endothelial cell injury were prominent. Importantly, oxidation of hemoglobin to methemoglobin occurred in the plasma, and iron was accumulating in his LDL. Iron-induced oxidative modification of lipoproteins is cytotoxic and causes endothelial damage, leading to the development of fatty streaks and fibrous plaques in the aorta. Endothelial cells of this child were susceptible to oxidative insults because of heme-mediated oxidation of LDL and an associated lack of adaptive responses, suggesting that HO-1 plays a crucial role in protecting vessels from oxidative insults. To support this concept in humans, Exner et al113 carried out a cohort study patients to evaluate HO-1 gene promoter polymorphisms and their risk for restenosis. They found that a dinucleotide repeat in the promoter region of the HO-1 gene showed a length polymorphism that modulate the level of gene transcription. In their study, patients with short (<25 GT) dinucleotide repeats in the HO-1 gene promoter on either allele had restenosis significantly less often that patients with longer (25 GT) dinucleotide repeats. These data imply that upregulation of HO-1, associated with shorter dinucleotide repeats, may be a protective factor after balloon angioplasty. More recently, a study that also assessed microsatellite polymorphisms in the HO-1 gene promoter showed that type 2 diabetic patients carrying longer (2 GT) repeats had higher oxidative stress and increased susceptibility to the development of atherosclerosis and coronary artery disease.114 Although the results of these studies on genetic polymorphism of the HO-1 gene may indicate the potential importance of HO-1 in the pathogenesis of cardiovascular diseases, direct measurement of HO-1 activity in the patients is necessary, as varying oxidative stress may elicit the HO-1 response in endothelial cells, smooth muscle cells, and inflammatory cells to different degrees. Development of a serial monitoring system of HO-1 activity for use in clinical setting is desirable.

    HO and Pharmacological Interventions

    HO may have potential in the treatment of atherosclerosis and vascular injury in humans by being directly delivered or by pharmacological activation of the HO-1 gene. However, most of the pharmacological inducers of HO-1 used in experimental studies, such as hemin and heavy metals, exhibit cellular and tissue toxicity, and the long-term and adverse effects of HO-1 gene delivery must be elucidated before its application in humans. Recent studies, however, revealed that some well-used drugs modulate HO-1 expression in vascular cells. Aspirin is known to reduce the incidence of thrombotic occlusive events such as myocardial infarction and stroke by inhibiting platelet activity. Recently, more direct effects of aspirin on the integrity of the vascular wall, free radical scavenging, and the capacity to protect endothelial cells from the deleterious effects of hydrogen peroxide have been reported. Grosser et al115 reported that aspirin increased HO-1 protein levels in a concentration-dependent fashion via NO dependent pathways in cultured endothelial cells derived from the human umbilical vein. They concluded that induction of HO-1 activity is a novel mechanism by which aspirin prevents cellular injury under inflammatory conditions and in cardiovascular disease. Statins, widely used lipid-lowering agents, substantially decrease cardiovascular morbidity and mortality in patients with and without coronary disease via not only a cholesterol-lowering effect but also their pleiotropic cholesterol-independent effects. The precise mechanisms of those pleiotropic effects, however, remain unclear. Grosser and colleagues116,117 explored the role of HO-1 as a target and mediator of statins. They revealed that statins, including simvastatin, lovastatin, and rosuvastatin, increased the transcription of HO-1 mRNA in human umbilical vein endothelial cells, which coincided with a reduction of free radicals formation. These results may explain the pleiotropic antioxidant, antiinflammatory, and antiatherogenic actions of statins. Rapamycin, a macrolide antibiotic, blocks cell cycle progression at the G1 phase and rapamycin-coated coronary stents have been shown to reduce coronary restenosis.118 Visner and colleagues reported that rapamycin induced HO-1 and suppressed platelet-derived growth factor-dependent vascular smooth muscle cell growth.119 It has been suggested that HO-1 induced by rapamycin in smooth muscle cells shows an antiproliferative effect, resulting in the reduction of restenosis rate.

    Conclusion and Perspectives

    Ample evidence has been provided to support HO-1 as a key player in various cardiovascular diseases, especially those in which ROS have been implicated (Figure 2). With regard to the possible relevance of HO-1 in atherosclerosis, it should be noted that HO-1 is induced by most of the well-established cardiovascular risk factors in vascular cells and circulating macrophages. These HO-1 responses appear to have a protective role in the vascular wall against atherogenesis through multiple pathways. Interventions aimed at modulating the levels of HO in the vascular wall, therefore, might be a novel target to treat or prevent atherosclerotic diseases. Some studies, however, have clearly demonstrated the surprising role of HO-1 in promoting endothelial cell apoptosis, endothelial dysfunction, and vasoconstriction. Akin et al120 reported that bilirubin (10 to 100 mol/L) induced apoptosis in bovine brain microvascular endothelial cells. Thom and colleagues121 showed that biochemical effects of CO occur at environmentally relevant concentrations and that exposure to relatively high concentrations of CO causes apoptotic cell death in bovine pulmonary artery endothelial cells. Johnson et al revealed122 that exogenous as well as endogenous CO induces vasoconstriction in isolated gracilis muscle arterioles, most likely through the inhibition of endothelial NO synthesis, and that elevated HO-1 levels and activity in arteries contribute to endothelial dysfunction, leading to hypertension in deoxycorticosterone acetate rats and Dahl-S rats.123–125 Furthermore, reduction of heme and heme proteins caused by HO-1 may impair the function of various crucial enzymes, such as mitochondrial cytochromes, microsomal cytocromes, cyclooxygenases, NO synthase, and NAD(P)H oxidase.96,126–130 Taken together, these observations suggest that the protective effect of HO-1 may vary depending on the level of and the specific site of expression. Most HO inducers that are heme analogues or heme derivatives are administered intraperitoneally, whereas SnCl2 is generally given subcutaneously. In contrast, the HO-1 gene is mainly delivered by intracardiac injection. Whether these different routes of administration influence the effectiveness and specificity of the therapy remains unclear. No matter what the route of administration, it seems that HO-1 inducers and HO-1 gene delivery do not produce an organ-specific overexpression of HO-1; the intracardiac injection of the HO-1 gene alters other organs, including lungs, kidneys, liver, and brain. Nevertheless, gene delivery has great potential for becoming more specific with future development, in the light of reports that some genes have been successfully overexpressed in specific targeted tissues, including extracellular matrix protein Del-1. Subcutaneously introduced Del-1 exerted endothelial cell-specific activities by binding to integrin v?5 localized only on endothelial cells, resulting in the promotion of neovascularization of ischemic tissues.131 Thus, a key to future therapeutic application of HO-1 for humans is to properly target and regulate HO-1 expression at a specific site to allow maximal cytoprotection and minimum adverse effects at the site of the lesion. Further advances are necessary to achieve these goals.

    Figure 2. Antiatherogenic properties of HO-1. Increased ROS production by well-established cardiovascular risk factors induces HO-1 expression in endothelial cells, vascular smooth muscle cells and macrophages. The enhanced HO-1/CO/bilirubin system exerts its cytoprotective properties in the atherosclerotic lesion through multi steps indicated as dotted arrows.

    Acknowledgments

    I thank all of my colleagues, past and present, for their efforts in our studies. I also thank my past mentors, Dr Stella Kourembanas, Dr Shigehiro Katayama, and Dr Gen Yoshino.

    References

    Wiztum JL, Steinberg D. Role of oxidized low density lipoprotein in atherogenesis. J Clin Invest. 1991; 88: 1785–1792.

    Berliner JA, Heinecke JW. The role of oxidized lipoproteins in atherosclerosis. Free Radic Biol Med. 1996; 20: 707–727.

    Parhami F, Fang ZT, Fogelman AM, Andalibi A, Territo MC, Berliner JA. Minimally modified low density lipoprotein-induced inflammatory responses in endothelial cells are mediated by cyclic adenosine monophosphate. J Clin Invest. 1993; 92: 471–478.

    Maines MD. The heme oxygenase system: a regulator of second messenger gases. Annu Rev Pharmacol Toxicol. 1997; 37: 517–554.

    Vincent SH, Grady RW, Shaklai N, Snider JM, Muller-Eberhard U. The influence of heme-binding proteins in heme-catalyzed oxidations. Arch Biochem Biophys. 1988; 265: 539–550.

    Jeney V, Balla J, Yachie A, Varga Z, Vercellotti GM, Eaton J, Balla G. Pro-oxidant and cytotoxic effects of circulating heme. Blood. 2002; 100: 879–887.

    Wagener, Feldman E, de Witte T, Abraham NG. Heme induces the expression of adhesion molecules ICAM-1, VCAM-1, and E selectin in vascular endothelial cells. Proc Soc Exp Biol Med. 1997; 216: 456–463.

    Tenhunen R, Marver HS, Schmid R. Microsomal heme oxygenase: characterization of the enzyme. J Biol Chem. 1969; 244: 6388–6394.

    Yoshida T, Kikuchi G. Purification and properties of heme oxygenase from pig spleen microsomes. J Biol Chem. 1978; 253: 4224–4229.

    Maines MD, Kappas A. Cobalt induction of hepatic heme oxygenase; with evidence that cytochrome P-450 is not essential for this enzyme activity. Proc Natl Acad Sci U S A. 1974; 71: 4293–4297.

    McCoubrey WKJ, Maines MD. The structure, organization and differential expression of the gene encoding rat heme oxygenase-2. Gene. 1994; 139: 155–161.

    Shibahara S, Yoshizawa M, Suzuki H, Takeda K, Meguro K, Endo K. Functional analysis of cDNAs for two types of human heme oxygenase and evidence for their separate regulation. J Biochem (Tokyo). 1993; 113: 214–218.

    McCoubrey WKJ, Huang TJ, Maines MD. Isolation and characterization of a cDNA from the rat brain that encodes hemoprotein heme oxygenase-3. Eur J Biochem. 1997; 247: 725–732.

    Wang LJ, Lee TS, Lee FY, Pai RC, Chau LY. Expression of heme oxygenase-1 in atherosclerotic lesions. Am J Pathol. 1998; 152: 711–720.

    Keyse SM, Tyrrell RM. Heme oxygenase is the major 32-kDa stress protein induced in human skin fibroblasts by UVA radiation, hydrogen peroxide, and sodium arsenite. Proc Nat Acad Sci U S A. 1989; 86: 99–103.

    Maines MD, Kappas A. Regulation of heme pathway enzymes and cellular glutathione content by metals that do not chelate with tetrapyrroles: blockade of metal effects by thiols. Proc Natl Acad Sci U S A. 1977; 74: 1875–1878.

    Lee PJ, Jiang BH, Chin BY, Iyer NV, Alam J, Semenza GL, Choi AM. Hypoxia-inducible fantor-1 mediates transcriptional activation of the heme oxygenase-1 gene in response to hypoxia. J Biol Chem. 1997; 272: 5375–5381.

    Morita T, Perrella MA, Lee ME, Kourembanas S. Smooth muscle cell-derived carbon monoxide is a regulator of vascular cGMP. Proc Natl Acad Sci U S A. 1995; 92: 1475–1479.

    Shibahara S, Yoshida T, Kikuchi G. Mechanism of increase of heme oxygenase activity induced by hemin in cultured pig alveolar macrophages. Arch Biochem Biophys. 1979; 197: 607–617.

    Alam J, Den Z. Distal AP-1 binding sites mediate basal level enhancement and TPA induction of the mouse heme oxygenase-1 gene. J Biol Chem. 1992; 267: 21894–21900.

    Lavrovsky Y, Schwartzman ML, Levere RD, Kappas A, Abraham NG. Identification of finding sites for transcription factors NF-kappa B and AP-2 in the promoter region of the human heme oxygenase 1 gene. Proc Natl Acad Sci U S A. 1994; 91: 5987–5991.

    Dalton T, Palmiter RD, Andrews GK. Transcriptional induction of the mouse metallothionein-I gene in hydrogen peroxide-treated Hepa cells involves a composite major late transcription factor/antioxidant response element and metal response promoter elements. Nucleic Acids Res. 1994; 22: 5016–5023.

    Hayashi S, Omata Y, Sakamoto H, Higashimoto Y, Hara T, Sagara Y, Noguchi M. Characterization of rat heme oxygenase-3 gene: implication of processed pseudogenes derived from heme oxygenase-2 gene. Gene. 2004; 336: 241–250.

    Mireles LC, Lum MA, Dennery PA. Antioxidant and cytotoxic effects of bilirubin on neonatal erythrocytes. Pediatr Res. 1999; 45: 355–362.

    Stocker R, Yamamoto Y, McDonagh AF, Glazer AN, Ames BN. Bilirubin is an antioxidant of possible physiological importance. Science. 1987; 235: 1043–1046.

    Schwertner HA, Jackson WG, Tolan G. Association of low serum concentration of bilirubin with increased risk of coronary artery disease. Clin Chem. 1994; 40: 18–23.

    Hopkins PN, Wu LL, Hunt SC, James BC, Vincent GM, Williams RR. Higher serum bilirubin is associated with decreased risk for early familial coronary artery disease. Arterioscler Thromb Vasc Biol. 1996; 16: 250–255.

    Breimer LH, Wannamethee G, Ebrahim S, Shaper AG. Serum bilirubin and risk of ischemic heart disease in middle-aged British men. Clin Chem. 1995; 41: 1504–1508.

    Hunt SC, Wu LL, Hopkins PN, Williams RR. Evidence for a major gene elevating serum bilirubin concentration in Utah pedigrees. Arterioscler Thromb Vasc Biol. 1996; 16: 912–917.

    Neuzil J, Stocker R. Free and albumin-bound bilirubin are density lipoprotein lipid peroxidation. J Biol Chem. 1994; 269: 16712–16719.

    Ishikawa K, Sugawara D, Wang XP, Suzuki K, Itabe H, Maruyama Y, Lusis AJ. Heme oxygenase-1 inhibits atherosclerotic lesion formation in LDL-receptor knockout mice. Circ Rec. 2001; 88: 506–512.

    Kwak JY, Takeshige K, Cheung BS, Minakami S. Bilirubin inhibits the activation of superoxide-producing NADPH oxidase in a neutrophil cell-free system. Biochem Biophys Acta. 1991; 1076: 369–373.

    Hansen TW, Mathiesen SB, Walaas SI. Bilirubin has widespread inhibitory effects on protein phosphorylation. Pediatr Res. 1996; 39: 1072–1077.

    Amit Y, Boneh A. Bilirubin inhibits protein kinase C activity and protein kinase C-mediated phosphorylation of endogenous substrates in human skin fibroblasts. Clin Chem Acta. 1993; 223: 103–111.

    Mayer M. Association of serum bilirubin concentration with risk of coronary artery disease. Clin Chem. 2000; 46: 1723–1727.

    Nakagami T, Toyomura K, Kinoshita T, Morisawa S. A beneficial role of bile pigments as an endogenous tissue protector: anti-complement effects of biliverdin and conjugated bilirubin. Biochim Biophys Acta. 1993; 1158: 189–193.

    Willis D, Moore AR, Frederick R, Willoughby DA. Heme oxygenase: a novel target for the modulation of the inflammatory response. Nat Med. 1996; 2: 87–90.

    Nakayama M, Takahashi K, Komaru T, Fukuchi M, Shioiri H, Sato Ki, Kitamuro T, Shirato K, Yamaguchi T, Suematsu M, Shibahara S. Increased expression of heme oxygenase-1 and bilirubin accumulation in foam cells of rabbit atherosclerotic lesions. Arterioscler Thromb Vasc Biol. 2001; 21: 1373–1377.

    Dennery PA, Seidman DS, Stevenson DK. Neonatal hyperbilirubinemia. N Engl J Med. 2001; 344: 581–590.

    Yamaguchi T, Shioji I, Sugimoto A, Yamaoka M. Psychological stress increases bilirubin metabolites in human urine. Biochem Biophys Res Commun. 2002; 293: 517–520.

    Shimomura H, Ogawa H, Takazoe K, Soejima H, Miyamoto S, Sakamoto T, Kawano H, Suefuji H, Nishikawa H, Arai H, Hokamaki J, Kajiwara I, Kugiyama K, Yoshimura M. Comparison of urinary of biopyrrin levels in acute myocardial infarction (after reperfusion therapy) versus stable angina pectoris and their usefulness in predicting subsequent cardiac events. Am J Cardiol. 2002; 90: 108–111.

    Dore S. Decreased activity of the antioxidant heme oxygenase enzyme: implications in ischemia and in Alzheimer’s disease. Free Radic Biol Med. 2002; 32: 1276–1282.

    Marks GS, Brien JF, Nakatsu K, McLaughlin BE. Does carbon monoxide have a physiological function? Trends Pharmacol Sci. 1991; 12: 185–188.

    Verma A, Hirsch DJ, Glatt CE, Ronnett GV, Snyder SH. Carbon monoxide: a putative neural messenger. Science. 1993; 259: 381–384.

    Furchgott RF, Jothianandan D. Endothelium-dependent and –independent vasodilation involving cyclic GMP: relaxation induced by nitric oxide, carbon monoxide and light. Blood Vessels. 1991; 28: 52–61.

    Kozma F, Johnson RA, Zhang F, Yu C, Tong X, Nasjletti A. Contribution of endogenous carbon monoxide to regulation of diameter in resistance vessels. Am J Physiol. 1999; 276: R1087–R1094.

    Kajimura M, Goda N, Suematsu M. Organ design for generation and reception of CO: lessons from the liver. Antioxid Redox Signal. 2002; 4: 633–637.

    Kajimura M, Shimoyama M, Tsuyama S, Suzuki T, Kozaki S, Takenaka S, Tsubota K, Oguchi Y, Suematsu M. Visualization of gaseous monoxide reception by soluble guanylate cyclase in the rat retina. FASEB J. 2003; 17 (3): 506–508.

    Imai T, Morita T, Shindo T, Nagai R, Yazaki Y, Kurihara H, Suematsu M, Katayama S. Vascular smooth muscle cell-directed overexpression of heme oxygenase-1 elevates blood pressure through attenuation of nitric oxide-induced vasodilation in mice. Circ Res. 2001; 89: 55–62.

    Grundemar L, Johansson MB, Ekelund M, Hogestatt ED. Haem oxygenase activity in blood vessel homogenates as measured by carbon monoxide production. Acta Physiol Scand. 1995; 153: 203–204.

    Sammut IA, Foresti R, Clark JE, Exon DJ, Vesely MJ, Sarathchandra P, Green CJ, Motterlini R. Carbon monoxide is a major contributor to the regulation of vascular tone in aortas expressing high levels of haeme oxygenase-1. Br J Pharmacol. 1998; 125: 1437–1444.

    Leffler C, Nasjletti A, Yu C, Johnson RA, Fedinec AL, walker N. Carbon monoxide and cerebral microvascular tone in newborn pigs. Am J Physiol. 1999; 276: H1641–H1646.

    Kaide JI, Zhang F, Wei Y, Jiang H, Yu C, Wang WH, Balazy M, Abraham NG, Nasjletti A. Carbon monoxide of vascular origin attenuates the sensitivity of renal arterial vessels to vasoconstrictors. J Clin Invest. 2001; 107: 1163–1171.

    Motterlini R, Gonzales A, Foresti R, Clark JE, Green CJ, Winslow RM. Heme oxygenase-1-derived carbon monoxide contributes to the suppression of acute hypertensive response in vivo. Circ Res. 1998; 83: 568–577.

    Morita T, Kourembanas S. Endothelial cell expression of vasoconstrictors and growth factors is regulated by smooth muscle cell-derived carbon monoxide. J Clin Invest. 1995; 96: 2676–2682.

    Morita T, Mitsialis SA, Koike H, Liu Y, Kourembanas S. Carbon monoxide controls the proliferation of hypoxic vascular smooth muscle cells. J Biol Chem. 1997; 272: 32804–32809.

    Peyton KJ, Reyna SV, Chapman GB, Ensenat D, Liu XM, Wang H, Schafer AI, Durante W. Heme oxygenase-1 derived carbon monoxide is an autocrine inhibitor of vascular smooth muscle cell growth. Blood. 2002; 99: 4443–4448.

    Liu XM, Chapman GB, Peyton KJ, Schafer AI, Durante W. Carbon monoxide inhibits apoptosis in vascular smooth muscle cells. Cardiovasc Res. 2002; 55: 396–405.

    Wang R, Wu L. The chemical modification of KCa channels by carbon monoxide in vascular smooth muscle cells. J Biol Chem. 1997; 272: 8222–8226.

    Otterbein LE, Bach FH, Alam J, Soares M, Tao LuH, Wysk M, Davis RJ, Flavell RA, Choi AM. Carbon monoxide has anti-inflammatory effects involving the mitogen-activated protein kinase pathway. Nat Med. 2000; 6: 422–428.

    Otterbein LE, Zuckerbraun BS, Haga M, Liu F, Song R, Usheva A, Stachulak C, Bodyak N, Smith RN, Csizmadia E, Tyagi S, Akamatsu Y, Flavell RJ, Billiar TR, Tzeng E, Bach FH, Choi AM, Soares MP. Carbon monoxide suppresses arteriosclerotic lesions associated with chronic graft rejection and with balloon injury. Nat Med. 2003; 9: 183–190.

    Amersi F, Shen XD, Anselmo D, Mclinek J, Iyer S, Southard DJ, Katori M, Volk HD, Busuttil RW, Kupiec-Weglinski JW. Ex vivo exposure to carbon monoxide presents hepatic ischemia/reperfusion injury through p38 MAP kinase pathway. Hepatology. 2002; 35: 815–823.

    Brouard S, Berberat PO, Tobiasch E, Seldon MP, Bach FH, Soares MP. Heme oxygenase-1-derived carbon monoxide requires the activation of transcription factor NF-kappa B to protect endothelial cells from tumor necrosis factor-alpha-mediated apoptosis. J Biol Chem. 2002; 277: 17950–17961.

    Harrison PM, Arosio P. The ferritins: molecular properties, iron storage function and cellular regulation. Biochim Biophys Acta. 1996; 1275: 161–203.

    Aziz N, Munro HN. Both subunits of rat liver ferritin are regulated at a translational level by iron induction. Nucleic Acids Res. 1986; 14: 915–927.

    Rogers J, Munro H. Translation of ferritin light and heavy subunit mRNAs is regulated by intracellular chelatable iron levels in rat hepatoma cells. Proc Natl Acad Sci U S A. 1987; 84: 2277–2281.

    Cairo G, Bardella L, Schiaffonati L, Arosio P, Levi S, Bernelli-Zazzera A. Multiple mechanisms of iron-induced ferritin synthesis in HeLa cells. Biochem Biophys Res Commun. 1985; 133: 314–321.

    Zahringer J, Baliga BS, Munro HN. Novel mechanism for translational control in regulation of ferritin synthesis by iron. Proc Natl Acad Sci U S A. 1976; 73: 857–861

    Fogg S, Agarwal A, Nick HS, Visner GA. Iron regulates hyperoxia-dependent human heme oxygenase 1 gene expression in pulmonary endothelial cells. Am J Respir Cell Mol Biol. 1999; 20: 797–804.

    Marton LS, Wang X, Kowalczuk A, Zhang ZD, Windmeyer E, Macdonald RL. Effects of hemoglobin on heme oxygenase gene expression and viability of cultured smooth muscle cells. Am J Physiol Heart Circ Physiol. 2000; 279: H2405–H2413.

    Balla G, Jacob HS, Balla J, Rosenberg M, Nath K, Apple F, Eaton JW, Vercellotti GM. Ferritin: a cytoprotective antioxidant strategem of endothelium. J Biol Chem. 1992; 267: 18148–18153.

    Chi SI, Wang CK, Chen JJ, Chau LY, Lin TN. Differential regulation of H- and L-ferritin messenger RNA subunits, ferritin protein and iron following focal cerebral ischemia-reperfusion. Neuroscience. 2000; 100: 475–484.

    Selmeci L, Antal M, Horkay F, Merkely B, Szokodi I, Biro L, Szekely M, Jobbagt J, Szepvolgyi J, Toth M. Enhanced accumulation of pericardial fluid ferritin in patients with coronary artery disease. Coron Artery Dis. 2000; 11: 53–56.

    Pang JH, Jiang MJ, Chen YL, Wang FW, Wang DL, Chu SH, Chau LY. Increased ferritin gene expression in atherosclerotic lesions. J Clin Invest. 1996; 97: 2204–2212.

    Hoekstra KA, Godin DV, Kurtu J, Cheng KM. Heme oxygenase and antioxidant status in cultured aortic endothelial cells isolated from atherosclerosis-susceptible and -resistant Japanese quail. Mol Cell Biochem. 2003; 252: 253–262.

    Hoekstra KA, Godin DV, Kurtu J, Cheng KM. Effect of oxidant-induced injury on heme oxygenase and glutathione in cultured aortic endothelial cells isolated from atherosclerosis-susceptible and -resistant Japanese quail. Mol Cell Biochem. 2003; 254: 61–71.

    White CR, Darley-Usmar V, Berrington WR, McAdams M, Gore JZ, Thompson JA, Parks DA, Tarpey MM, Freeman BA. Circulating plasma xanthine oxidase contributes to vascular dysfunction in hypercholesterolemic rabbits. Proc Natl Acad Sci U S A. 1996; 93: 8745–8749.

    Griendling KK, Sorescu D, Ushio-Fukai M. NAD(P)H oxidase: role in cardiovascular biology and disease. Circ Res. 2000; 86: 494–501.

    Landmesser U, Dikalov S, Price SR, McCann L, Fukai T, Holland SM, Mitch WE, Harrison DG. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J Clin Invest. 2003; 111: 1201–1209.

    Gaut JP, Byun J, Tran HD, Lauber WM, Carroll JA, Hotchkiss RS, Belaaouaj A, Heinecke JW. Myeloperoxidase produces nitrating oxidants in vivo. J Clin Invest. 2002; 109: 1311–1319.

    Lusis AJ. Atherosclerosis. Nature. 2000; 407: 233–221.

    Libby P. Inflammation in atherosclerosis. Nature. 2002; 420: 868–874.

    Wasserman WW, Fahl WE. Functional antioxidant responsive elements. Proc Natl Acad Sci U S A. 1997; 94: 5361–5366.

    Nguyen T, Sherratt PJ, Pickett CB. Regulatory mechanisms controlling gene expression mediated by the antioxidant response element. Annu Rev Pharmacol Toxicol. 2003; 43: 233–260.

    Alam A. Multiple elements within the 5' distal enhancer of the mouse heme oxigenase-1 gene mediate induction by heavy metals. J Biol Chem. 1994; 269: 25049–25056.

    Venugopal R, Jaiswal AK. Nrf1 abd Nrf2 positively and c-FOS and Fra1 negatively regulate the human antioxidant response element-mediated expression of NAD(P)H: quinone oxidoreductase 1 gene. Proc Natl Acad Sci U S A. 1996; 93: 14960–14965.

    Itoh K, Wakabayashi N, Katoh Y, Ishii T, Igarashi K, Engel JD, Yamamoto M. Keap I represses nuclear activation of antioxidant response elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 1999; 13: 76–86.

    Alam J, Stewart D, Touchars C, Boinapally S, Choi AM, Cook JL. Nrf2, a Cap‘n’Collar transcription factor, regulates induction of the heme oxygenase-1 gene. J Biol Chem. 1999; 274: 26071–26078.

    Chan K, Kan YW. Nrf2 is essential for protection against acute pulmonary injury in mice. Proc Natl Acad Sci U S A. 1999; 96: 12731–12736.

    Ishii T, Itoh K, Takahashi S, Sato H, Yanagawa T, Katoh Y, Bannai S, Yamanoto M. Transcription factor Nrf2 coordinately regulates a group of oxidative stress-inducible genes in macrophages. J Biol Chem. 2000; 275: 16023–16029.

    Ishii T, Itoh K, Ruiz E, Leake DS, Unoki H, Yamanoto M, Mann GE. Role of Nrf2 in the regulation of CD36 and stress protein expression in murine macrophages: activation by oxidatively modified LDL and 4-hydroxynonenal. Circ Res. 2004; 94: 609–616.

    Ishikawa K, Navab M, Leitinger N, Fogelman AM, Lusis AJ. Induction of heme oxygenase-1 inhibits the monocyte transmigration induced by mildly oxidized LDL. J Clin Invest. 1997; 100: 1209–1216.

    Yamaguchi M, Sato H, Bannai S. Induction of stress proteins in mouse peritoneal macrophages by oxidized low-density lipoprotein. Biochem Biophys Res Commun. 1993; 193: 1198–1201.

    Agarwal A, Shiraishi F, Visner GA, Nick HS. Linoleyl hydroperoxide transcriptionally upregulates heme oxygenase-1 gene expression in human renal epithelial and aortic endothelial cells. J Am Soc Nephrol. 1998; 9: 1990–1997.

    Ishikawa K, Sugawara D, Goto J, Watanabe Y, Kawamura K, Shiomi M, Itabe H, Maruyama Y. Heme oxygenase-1 inhibits atherogenesis in Watanabe heritable hyperlipidemic rabbits. Circulation. 2001; 104: 1831–1836.

    Taille C, El-Benna J, Lanone S, Dang MC, Ogier-Denis E, Aubler M, Boczkowski J. Induction of heme oxygenase-1 inhibits NAD(P)H oxidase activity by down-regulating cytochrome b558 expression via the reduction of heme availability. J Biol Chem. 2004; 279: 28681–286818.

    Abraham NG, Kushida T, McClung J, Weiss M, Quan S, Lafaro R, Darzynkiewicz Z, Wolin M. Heme oxygenase-1 attenuates glucose-mediated cell growth arrest and apoptosis in human microvessel endothelial cells. Circ Res. 2003; 93: 507–514.

    Morita T, Imai T, Yamaguchi T, Sugiyama T, Katayama S, Yoshino G. Induction of heme oxygenase-1 in monocytes suppresses angiotensin II-elicited chemotactic activity through inhibition of CCR2: role of bilirubin and carbon monoxide generated by the enzyme. Antioxid Redox Signal. 2003; 5: 439–447.

    Hayashi S, Takamiya R, Yamaguchi T, Matsumoto K, Tojo SJ, Tamatani T, Kitajima M, Makino N, Ishimura Y, Suematsu M. Induction of heme oxygenase-1 suppresses venular leukocyte adhesion elicited by oxidative stress: role of bilirubin generated by the enzyme. Circ Res. 1999; 85: 663–671.

    Togane Y, Morita T, Suematsu M, Ishimura Y, Yamazaki J, Katayama S. Protective roles of endogenous carbon monoxide in neointimal development elicited by arterial injury. Am J Physiol Heart Circ Physiol. 2000; 278: H623–H632.

    Grundemar L, Ny L. Pitfalls using metalloporphyrins in carbon monoxide research. Trends Pharmacol Sci. 1997; 18: 193–195.

    Luo D, Vincent SR. Metalloporphyrins inhibit nitric oxide-dependent cGMP formation in vivo. Eur J Pharmacol. 1994; 267: 263–267.

    Ny L, Andersson KE, Grundemar L. Inhibition by zinc protoporphyrin-IX of receptor-mediated relaxation of the rat aorta in a manner distinct from inhibition of haeme oxygenase. Br J Pharmacol. 1995; 115: 186–190.

    Abraham NG, Rezzani R, Rodella L, Kruger A, Taller D, Li Volti G, Goodman Al, Kappas A. Overexpression of human heme oxygenase-1 attenuates endothelial cell sloughing in experimental diabetes. Am J Physiol Heart Circ Physiol. 2004; 287: H2468–H2477.

    Quan S, Yang L, Abraham N, Kappas A. Regulation of human heme oxygenase in endothelial cells by using sense and antisense retroviral constructs. Proc Natl Acad Sci U S A. 2001; 98: 12203–12208.

    Tulis DA, Durante W, Liu X, Evans AJ, Peyton KJ, Schafer AI. Adenovirus-mediated heme oxygenase-1 gene delivery inhibits injury-induced vascular neointima formation. Circulation. 2001; 104: 2710–2715.

    Juan SH, Lee TS, Tseng KW, Liou JY, Shyue SK, Wu KK, Chau LY. Adenovirus-mediated heme oxygenase-1 gene transfer inhibits the development of atherosclerosis in apolipoprotein E-deficient mice. Circulation. 2001; 104: 1519–1525.

    Duckers HJ, Boehm M, True AL, Yet SF, San H, Park JL, Clinton Webb R, Lee ME, Nabel GJ, Nabel EG. Heme oxygenase-1 protects against vascular constriction and proliferation. Nat Med. 2001; 7: 693–698.

    Yet SF, Layne MD, Liu X, Chen YH, Ith B, Sibinga NE, Perrella MA. Absence of heme oxygenase-1 exacerbates atherosclerotic lesion formation and vascular remodeling. FASEB J. 2003; 17: 1759–1761.

    Morita T, Imai T, Sugiyama T, Katayama S, Yoshino G. Heme oxygenase-1 in vascular smooth muscle cells counteracts cardiovascular damage induced by angiotensin II. Current Neurovascular Research. 2005; 2: 113–120.

    Yachie A, Niida Y, Wada T, Igarashi N, Kaneda H, Toma T, Ohta K, Kasahara Y, Koizumi S. Oxidative stress causes enhanced endothelial cell injury in human heme oxygenase-1 deficiency. J Clin Invest. 1999; 103: 129–135.

    Kawashima A, Oda Y, Yachie A, Koizumi S, Nakanishi I. Heme oxigenase-1 deficiency: the first autopsy case. Hum Pathol. 2002; 33: 125–130.

    Exner M, Schillinger M, Minar E, Mlekusch W, Schlerka G, Haumer M, Mannhalter C, Wagner O. Heme oxygenase-1 gene promoter microsatellite polymorphism is associated with restenosis after percutaneous transluminal angioplasty. J Endovasc Ther. 2001; 8: 433–440.

    Chen YH, Lin SJ, Lin MW, Tsai HL, Kuo SS, Chen JW, Charng MJ, Wu TC, Chen LC, Ding YA, Pan WH, Jou YS, Chau LY. Microsatellite polymorphism in promoter of heme oxygenase-1 gene is associated with susceptibility to coronary artery disease in type 2 diabetic patients. Hum Genet. 2002; 111: 1–8.

    Grosser N, Abate A, Oberle S, Vreman HJ, Dennery PA, Becker JC, Pohle T, Seidman DS, Schroder H. Heme oxygenase-1 induction may explain the antioxidant profile of aspirin. Biochem Biophys Res Commun. 2003; 308: 956–960.

    Grosser N, Erdmann K, Hemmerle A, Berndt G, Hinkelmann U, Smith G, Schroder H. Rosuvastatin upregulates the antioxidant defense protein heme oxygenase-1. Biochem Biophys Res Commun. 2004; 325: 871–876.

    Grosser N, Hemmerle A, Berndt G, Erdmann K, Hinkelmann U, Schurgerc S, Wijayanti N, Immenschuh S, Schroder H. The antioxidant defense protein heme oxygenase 1 is a novel target for statins in endothelial cells. Free Radic Biol Med. 2004; 37: 2064–2071.

    Morice MC, Serruys PW, Sousa JE, Fajadet J, Ban Hayashi E, Perin M, Colombo A, Schuler G, Barragan P, Guagliumi G, Molnar F, Falotico R; RAVEL Study Group. Randomized study with the sirolimus-coated Bx velocity balloon-expandable stent in the treatment of patients with de novo native coronary artery lesions. A randomized comparison of a sirolimus-eluting stent with a standard stent for coronary revascularization. N Engl J Med. 2002; 346: 1773–1780.

    Visner GA, Lu F, Zhou H, Liu J, Kazemfar K, Agarwal A. Rapamycin induces heme oxygenase-1 in human pulmonary vascular cells: implications in the antiproliferative response to rapamycin. Circulation. 2003; 107: 911–916.

    Akin E, Clower B, Tibbs R, Tang J, Zhang J. Bilirubin produces apoptosis in cultured bovine brain endothelial cells. Brain Res. 2002; 931: 168–175.

    Thom SR, Fisher D, Xu YA, Notar FK, Ischiropoulos H. Adaptive responses and apoptosis in endothelial cells exposed to carbon monoxide. Proc Natl Acad Sci U S A. 2000; 97: 1305–1310.

    Johnson FK, Johnson RA. Carbon monoxide promotes endothelium-dependent constriction of isolated gracilis muscle arterioles. Am J Physiol Regul Integr Comp Physiol. 2003; 285: R536–R541.

    Johnson FK, Durante W, Peyton KJ, Johnson RA. Heme oxygenase-mediated endothelial dysfunction in DOCA-salt, but not in spontaneously hypertensive, rat arterioles. Am J Physiol Heart Circ Physiol. 2004; 286: H1681–H1687.

    Johnson FK, Durante W, Peyton KJ, Johnson RA. Heme oxygenase inhibitor restores arteriolar nitric oxide function in Dahl rats. Hypertension. 2003; 41: 149–155.

    Teran FJ, Johnson RA, Stevenson BK, Peyton KJ, Jackson KE, Appleton SD, Durante W, Johnson FK. Heme oxygenase-derived carbon monoxide promotes arteriolar endothelial dysfunction and contributes to salt-induced hypertension in Dahl salt-sensitive rats. Am J Physiol Regul Integr Comp Physiol. 2005; 288: R615–R622.

    Kutty RK, Maines MD. Hepatic heme metabolism: possible role of biliverdin in the regulation of heme oxygenase activity. Biochem Biophys Res Commun. 1984; 122: 40–46.

    Stocker R. Induction of haem oxygenase as a defense against oxidative stress. Free Radic Res Commun. 1990; 9: 101–112.

    Muller-Eberhard U, Fraig M. Bioactivity of heme and its containment. Am J Hematol. 1993; 42: 59–62.

    Willis D, Tomlinson A, Frederick R, Paul-Clark MJ., Willoughby DA. Modulation of heme oxygenase activity in rat brain and spleen by inhibitors and donors of nitric oxide. Biochem Biophys Res Commun. 1995; 214: 1152–1156.

    Balla J, Jacob HS, Balla G, Nath K, Eaton JW, Vercellotti GM. Endothelial-cell heme uptake from heme proteins: induction of sensitization and desensitization to oxidant damage. Proc Natl Acad Sci U S A. 1993; 90: 9285–9289.

    Zhong J, Eliceiri B, Stupack D, Penta K, Sakamoto G, Quertermous T, Coleman M, Bodreau N, Varner JA. Neovascularization of ischemic tissues by gene delivery of the extracellular matrix protein Del-1. J Clin Invest. 2003; 112: 30–41.(Toshisuke Morita)