CVD and Oxidative Stress

Karla Cervantes Gracia, Daniel Llanas-Cornejo, Holger Husi, Karla Cervantes Gracia, Daniel Llanas-Cornejo, Holger Husi

Abstract

Nowadays, it is known that oxidative stress plays at least two roles within the cell, the generation of cellular damage and the involvement in several signaling pathways in its balanced normal state. So far, a substantial amount of time and effort has been expended in the search for a clear link between cardiovascular disease (CVD) and the effects of oxidative stress. Here, we present an overview of the different sources and types of reactive oxygen species in CVD, highlight the relationship between CVD and oxidative stress and discuss the most prominent molecules that play an important role in CVD pathophysiology. Details are given regarding common pharmacological treatments used for cardiovascular distress and how some of them are acting upon ROS-related pathways and molecules. Novel therapies, recently proposed ROS biomarkers, as well as future challenges in the field are addressed. It is apparent that the search for a better understanding of how ROS are contributing to the pathophysiology of CVD is far from over, and new approaches and more suitable biomarkers are needed for the latter to be accomplished.

Keywords: CVD; cardiovascular disease; oxidative stress; reactive oxygen species.

Conflict of interest statement

The authors declare no conflict of interest.

Figures

Figure 1
Figure 1
Atherosclerotic plaque formation process. LDLs at a high concentration inhibit endothelial cells’ (EC) endocytosis capacity, migrate and accumulate in the intima. LDLs get oxidized and induce VCAM (green square) expression in EC. Monocytes are recruited into the intima by VCAM interaction. Monocytes transform into macrophages, which take-up oxLDLs, forming foam cells. Macrophages and EC secrete chemokines and recruit T-cells. T-cells produce TNF-alpha and IFN-gamma, amplifying inflammation. Fibroblast growth factor (FGF) and platelet-derived growth factor (PDGF) stimulate smooth-muscle cells (SMC) migration and proliferation. SMCs can also accumulate lipids, migrate and proliferate. A lipid core is generated with necrotic foam cells surrounded by SMCs and a collagen fibrous cap, resulting in thrombus formation.
Figure 2
Figure 2
Metabolic pathways in ROS production and metabolism.
Figure 3
Figure 3
NADPH oxidase-modulated pathway in vascular smooth muscle cell differentiation. The TGF-β-mediated induction of NOX4 causes an inhibition of MKP-1 and activation of RhoA by ROS. Under physiological conditions, MKP-1 inhibits p38MAPK, thereby preventing the RhoA-mediated release and association of transcription elements SRF, which can be activated by p38MAPK, and MRTF, which drive SMA gene expression.
Figure 4
Figure 4
Illustration of neutrophil NOX-HOCl production. Neutrophils have phagosomes and granules in its cytoplasm. Granules contain myeloperoxidase (MPO) (orange). The NOX2 complex (blue) gets assembled once the neutrophils initiate bacterial endocytosis. NOX2 produces H2O2 by dismutation, and MPO reacts with H2O2 and chloride ions, thereby generating HOCl and water, leading to neutrophil bactericidal effects.
Figure 5
Figure 5
Molecular pathways associated with CVD etiology. Contributing CVD factors, such as diabetes, cause an imbalance of essential physiological pathways by impairing normal glycolysis (GL), which feeds into the tricarboxylic acid pathway (TCA) that is directly linked to the electron transport chain (ETC). Fatty acids, which can also contribute and feed into the TCA cycle through fatty acid oxidation (FAO), are also implicated in CVD manifestation. Multiple enzymes involved in this scheme have been shown to either modulate or directly generate ROS.

References

    1. Sies H. Oxidative stress: A concept in redox biology and medicine. Redox Biol. 2015;4:180–183. doi: 10.1016/j.redox.2015.01.002.
    1. Jones D.P. Redefining oxidative stress. Antioxid. Redox Signal. 2006;8:1865–1879. doi: 10.1089/ars.2006.8.1865.
    1. Mittal C.K., Murad F. Activation of guanylate cyclase by superoxide dismutase and hydroxyl radical: A physiological regulator of guanosine 3′,5′-monophosphate formation. Proc. Natl. Acad. Sci. USA. 1977;74:4360–4364. doi: 10.1073/pnas.74.10.4360.
    1. Olguin-Albuerne M., Moran J. Ros produced by NOX2 control in vitro development of cerebellar granule neurons development. ASN Neuro. 2015 doi: 10.1177/1759091415578712.
    1. Mandal D., Fu P., Levine A.D. REDOX regulation of IL-13 signaling in intestinal epithelial cells: Usage of alternate pathways mediates distinct gene expression patterns. Cell. Signal. 2010;22:1485–1494. doi: 10.1016/j.cellsig.2010.05.017.
    1. Abimannan T., Peroumal D., Parida J.R., Barik P.K., Padhan P., Devadas S. Oxidative stress modulates the cytokine response of differentiated Th17 and Th1 cells. Free Radic. Biol. Med. 2016;99:352–363. doi: 10.1016/j.freeradbiomed.2016.08.026.
    1. Fujino G., Noguchi T., Matsuzawa A., Yamauchi S., Saitoh M., Takeda K., Ichijo H. Thioredoxin and TRAF family proteins regulate reactive oxygen species-dependent activation of ASK1 through reciprocal modulation of the N-terminal homophilic interaction of ASK1. Mol. Cell. Biol. 2007;27:8152–8163. doi: 10.1128/MCB.00227-07.
    1. Heppner D.E., Hristova M., Dustin C.M., Danyal K., Habibovic A., van der Vliet A. The NADPH oxidases DUOX1 and NOX2 play distinct roles in redox regulation of epidermal growth factor receptor signaling. J. Biol. Chem. 2016;291:23282–23293. doi: 10.1074/jbc.M116.749028.
    1. Sart S., Song L., Li Y. Controlling redox status for stem cell survival, expansion, and differentiation. Oxid. Med. Cell. Longev. 2015;2015:105135. doi: 10.1155/2015/105135.
    1. Krumova K., Cosa G. Singlet Oxygen: Applications in Biosciences and Nanosciences. Volume 1. The Royal Society of Chemistry; London, UK: 2016. Overview of reactive oxygen species; pp. 1–21. Chapter 1.
    1. Holmstrom K.M., Finkel T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 2014;15:411–421. doi: 10.1038/nrm3801.
    1. Austin V., Crack P.J., Bozinovski S., Miller A.A., Vlahos R. Copd and stroke: Are systemic inflammation and oxidative stress the missing links? Clin. Sci. 2016;130:1039–1050. doi: 10.1042/CS20160043.
    1. Mendis S., Puska P., Norrving B. Global Atlas on Cardiovascular Disease Prevention and Control. World Health Organization; Geneva, Switzerland: 2011.
    1. Mozaffarian D., Benjamin E.J., Go A.S., Arnett D.K., Blaha M.J., Cushman M., Das S.R., de Ferranti S., Despres J.P., Fullerton H.J., et al. Heart disease and stroke statistics—2016 update: A report from the american heart association. Circulation. 2016;133:e38–e360. doi: 10.1161/CIR.0000000000000350.
    1. Go A.S., Mozaffarian D., Roger V.L., Benjamin E.J., Berry J.D., Blaha M.J., Dai S., Ford E.S., Fox C.S., Franco S., et al. Heart disease and stroke statistics—2014 update: A report from the american heart association. Circulation. 2014;129:e28–e292. doi: 10.1161/01.cir.0000441139.02102.80.
    1. Mehta J.L., Saldeen T.G., Rand K. Interactive role of infection, inflammation and traditional risk factors in atherosclerosis and coronary artery disease. J. Am. Coll. Cardiol. 1998;31:1217–1225. doi: 10.1016/S0735-1097(98)00093-X.
    1. Zwaka T.P., Hombach V., Torzewski J. C-reactive protein-mediated low density lipoprotein uptake by macrophages: Implications for atherosclerosis. Circulation. 2001;103:1194–1197. doi: 10.1161/01.CIR.103.9.1194.
    1. Schuh J., Fairclough G.F., Haschemeyer R.H. Oxygen-mediated heterogeneity of apo-low-density lipoprotein. Proc. Natl. Acad. Sci. USA. 1978;75:3173–3177. doi: 10.1073/pnas.75.7.3173.
    1. Vlassara H. The AGE-receptor in the pathogenesis of diabetic complications. Diabetes Metab. Res. Rev. 2001;17:436–443. doi: 10.1002/dmrr.233.
    1. Goldstein J.L., Ho Y.K., Basu S.K., Brown M.S. Binding site on macrophages that mediates uptake and degradation of acetylated low density lipoprotein, producing massive cholesterol deposition. Proc. Natl. Acad. Sci. USA. 1979;76:333–337. doi: 10.1073/pnas.76.1.333.
    1. Steinbrecher U.P., Fisher M., Witztum J.L., Curtiss L.K. Immunogenicity of homologous low density lipoprotein after methylation, ethylation, acetylation, or carbamylation: Generation of antibodies specific for derivatized lysine. J. Lipid Res. 1984;25:1109–1116.
    1. Ross R. Atherosclerosis—An inflammatory disease. N. Engl. J. Med. 1999;340:115–126. doi: 10.1016/S0002-8703(99)70266-8.
    1. Avogaro P., Cazzolato G., Bittolo-Bon G. Some questions concerning a small, more electronegative LDL circulating in human plasma. Atherosclerosis. 1991;91:163–171. doi: 10.1016/0021-9150(91)90198-C.
    1. Brownlee M., Cerami A., Vlassara H. Advanced glycosylation end products in tissue and the biochemical basis of diabetic complications. N. Engl. J. Med. 1988;318:1315–1321.
    1. Witztum J.L., Mahoney E.M., Branks M.J., Fisher M., Elam R., Steinberg D. Nonenzymatic glucosylation of low-density lipoprotein alters its biologic activity. Diabetes. 1982;31:283–291. doi: 10.2337/diab.31.4.283.
    1. Bucala R., Makita Z., Koschinsky T., Cerami A., Vlassara H. Lipid advanced glycosylation: Pathway for lipid oxidation in vivo. Proc. Natl. Acad. Sci. USA. 1993;90:6434–6438. doi: 10.1073/pnas.90.14.6434.
    1. Klein R.L., Laimins M., Lopes-Virella M.F. Isolation, characterization, and metabolism of the glycated and nonglycated subfractions of low-density lipoproteins isolated from type I diabetic patients and nondiabetic subjects. Diabetes. 1995;44:1093–1098. doi: 10.2337/diab.44.9.1093.
    1. Sobal G., Menzel J., Sinzinger H. Why is glycated LDL more sensitive to oxidation than native LDL? A comparative study. Prostaglandins Leukot. Essent. Fatty Acids. 2000;63:177–186. doi: 10.1054/plef.2000.0204.
    1. Pirillo A., Norata G.D., Catapano A.L. LOX-1, OxLDL, and atherosclerosis. Mediat. Inflam. 2013;2013:152786. doi: 10.1155/2013/152786.
    1. Raggi P. Inflammation, depression and atherosclerosis or depression, inflammation and atherosclerosis? Atherosclerosis. 2016;251:542–543. doi: 10.1016/j.atherosclerosis.2016.07.902.
    1. Assies J., Mocking R.J., Lok A., Ruhe H.G., Pouwer F., Schene A.H. Effects of oxidative stress on fatty acid- and one-carbon-metabolism in psychiatric and cardiovascular disease comorbidity. Acta Psychiatr. Scand. 2014;130:163–180. doi: 10.1111/acps.12265.
    1. Panth N., Paudel K.R., Parajuli K. Reactive oxygen species: A key hallmark of cardiovascular disease. Adv. Med. 2016;2016:12. doi: 10.1155/2016/9152732.
    1. Brown D.I., Griendling K.K. Regulation of signal transduction by reactive oxygen species in the cardiovascular system. Circ. Res. 2015;116:531–549. doi: 10.1161/CIRCRESAHA.116.303584.
    1. Mikhed Y., Daiber A., Steven S. Mitochondrial oxidative stress, mitochondrial DNA damage and their role in age-related vascular dysfunction. Int. J. Mol. Sci. 2015;16:15918–15953. doi: 10.3390/ijms160715918.
    1. Dikalov S. Cross talk between mitochondria and NADPH oxidases. Free Radic. Biol. Med. 2011;51:1289–1301. doi: 10.1016/j.freeradbiomed.2011.06.033.
    1. Jensen P.K. Antimycin-insensitive oxidation of succinate and reduced nicotinamide-adenine dinucleotide in electron-transport particles. I. Ph dependency and hydrogen peroxide formation. Biochim. Biophys. Acta. 1966;122:157–166. doi: 10.1016/0926-6593(66)90057-9.
    1. Chance B., Sies H., Boveris A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. 1979;59:527–605.
    1. Dan Dunn J., Alvarez L.A., Zhang X., Soldati T. Reactive oxygen species and mitochondria: A nexus of cellular homeostasis. Redox Biol. 2015;6:472–485. doi: 10.1016/j.redox.2015.09.005.
    1. Wenzel P., Mollnau H., Oelze M., Schulz E., Wickramanayake J.M., Muller J., Schuhmacher S., Hortmann M., Baldus S., Gori T., et al. First evidence for a crosstalk between mitochondrial and NADPH oxidase-derived reactive oxygen species in nitroglycerin-triggered vascular dysfunction. Antioxid. Redox Signal. 2008;10:1435–1447. doi: 10.1089/ars.2007.1969.
    1. Doughan A.K., Harrison D.G., Dikalov S.I. Molecular mechanisms of angiotensin ii-mediated mitochondrial dysfunction: Linking mitochondrial oxidative damage and vascular endothelial dysfunction. Circ. Res. 2008;102:488–496. doi: 10.1161/CIRCRESAHA.107.162800.
    1. Ago T., Kuroda J., Pain J., Fu C., Li H., Sadoshima J. Upregulation of NOX4 by hypertrophic stimuli promotes apoptosis and mitochondrial dysfunction in cardiac myocytes. Circ. Res. 2010;106:1253–1264. doi: 10.1161/CIRCRESAHA.109.213116.
    1. Nazarewicz R.R., Dikalova A.E., Bikineyeva A., Dikalov S.I. NOX2 as a potential target of mitochondrial superoxide and its role in endothelial oxidative stress. Am. J. Physiol. Heart Circ. Physiol. 2013;305:H1131–H1140. doi: 10.1152/ajpheart.00063.2013.
    1. Dikalov S.I., Nazarewicz R.R., Bikineyeva A., Hilenski L., Lassegue B., Griendling K.K., Harrison D.G., Dikalova A.E. NOX2-induced production of mitochondrial superoxide in angiotensin II-mediated endothelial oxidative stress and hypertension. Antioxid. Redox Signal. 2014;20:281–294. doi: 10.1089/ars.2012.4918.
    1. Landmesser U., Dikalov S., Price S.R., McCann L., Fukai T., Holland S.M., Mitch W.E., Harrison D.G. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J. Clin. Investig. 2003;111:1201–1209. doi: 10.1172/JCI200314172.
    1. Fleming I., Fisslthaler B., Dimmeler S., Kemp B.E., Busse R. Phosphorylation of Thr495 regulates Ca2+/calmodulin-dependent endothelial nitric oxide synthase activity. Circ. Res. 2001;88:E68–E75. doi: 10.1161/hh1101.092677.
    1. Lin M.I., Fulton D., Babbitt R., Fleming I., Busse R., Pritchard K.A., Jr., Sessa W.C. Phosphorylation of threonine 497 in endothelial nitric-oxide synthase coordinates the coupling of l-arginine metabolism to efficient nitric oxide production. J. Biol. Chem. 2003;278:44719–44726. doi: 10.1074/jbc.M302836200.
    1. Loot A.E., Schreiber J.G., Fisslthaler B., Fleming I. Angiotensin II impairs endothelial function via tyrosine phosphorylation of the endothelial nitric oxide synthase. J. Exp. Med. 2009;206:2889–2896. doi: 10.1084/jem.20090449.
    1. Gao L., Mann G.E. Vascular NAD(p)h oxidase activation in diabetes: A double-edged sword in redox signalling. Cardiovasc. Res. 2009;82:9–20. doi: 10.1093/cvr/cvp031.
    1. Bugger H., Abel E.D. Mitochondria in the diabetic heart. Cardiovasc. Res. 2010;88:229–240. doi: 10.1093/cvr/cvq239.
    1. Kayama Y., Raaz U., Jagger A., Adam M., Schellinger I.N., Sakamoto M., Suzuki H., Toyama K., Spin J.M., Tsao P.S. Diabetic cardiovascular disease induced by oxidative stress. Int. J. Mol. Sci. 2015;16:25234–25263. doi: 10.3390/ijms161025234.
    1. Quast U., Stephan D., Bieger S., Russ U. The impact of ATP-sensitive k+ channel subtype selectivity of insulin secretagogues for the coronary vasculature and the myocardium. Diabetes. 2004;53:S156–S164. doi: 10.2337/diabetes.53.suppl_3.S156.
    1. Madungwe N.B., Zilberstein N.F., Feng Y., Bopassa J.C. Critical role of mitochondrial ROS is dependent on their site of production on the electron transport chain in ischemic heart. Am. J. Cardiovasc. Dis. 2016;6:93–108.
    1. Mozaffari M.S., Baban B., Liu J.Y., Abebe W., Sullivan J.C., El-Marakby A. Mitochondrial complex I and NAD(P)H oxidase are major sources of exacerbated oxidative stress in pressure-overloaded ischemic-reperfused hearts. Basic Res. Cardiol. 2011;106:287–297. doi: 10.1007/s00395-011-0150-7.
    1. Baines C.P., Kaiser R.A., Purcell N.H., Blair N.S., Osinska H., Hambleton M.A., Brunskill E.W., Sayen M.R., Gottlieb R.A., Dorn G.W., et al. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature. 2005;434:658–662. doi: 10.1038/nature03434.
    1. Di Lisa F., Bernardi P. A capful of mechanisms regulating the mitochondrial permeability transition. J. Mol. Cell. Cardiol. 2009;46:775–780. doi: 10.1016/j.yjmcc.2009.03.006.
    1. Napoli C., Martin-Padura I., de Nigris F., Giorgio M., Mansueto G., Somma P., Condorelli M., Sica G., De Rosa G., Pelicci P. Deletion of the p66Shc longevity gene reduces systemic and tissue oxidative stress, vascular cell apoptosis, and early atherogenesis in mice fed a high-fat diet. Proc. Natl. Acad. Sci. USA. 2003;100:2112–2116. doi: 10.1073/pnas.0336359100.
    1. Bianchi P., Kunduzova O., Masini E., Cambon C., Bani D., Raimondi L., Seguelas M.H., Nistri S., Colucci W., Leducq N., et al. Oxidative stress by monoamine oxidase mediates receptor-independent cardiomyocyte apoptosis by serotonin and postischemic myocardial injury. Circulation. 2005;112:3297–3305. doi: 10.1161/CIRCULATIONAHA.104.528133.
    1. Carpi A., Menabo R., Kaludercic N., Pelicci P., Di Lisa F., Giorgio M. The cardioprotective effects elicited by p66Shc ablation demonstrate the crucial role of mitochondrial ROS formation in ischemia/reperfusion injury. Biochim. Biophys. Acta. 2009;1787:774–780. doi: 10.1016/j.bbabio.2009.04.001.
    1. Kaludercic N., Takimoto E., Nagayama T., Feng N., Lai E.W., Bedja D., Chen K., Gabrielson K.L., Blakely R.D., Shih J.C., et al. Monoamine oxidase a-mediated enhanced catabolism of norepinephrine contributes to adverse remodeling and pump failure in hearts with pressure overload. Circ. Res. 2010;106:193–202. doi: 10.1161/CIRCRESAHA.109.198366.
    1. Konior A., Schramm A., Czesnikiewicz-Guzik M., Guzik T.J. NADPH oxidases in vascular pathology. Antioxid. Redox Signal. 2014;20:2794–2814. doi: 10.1089/ars.2013.5607.
    1. Amanso A.M., Griendling K.K. Differential roles of NADPH oxidases in vascular physiology and pathophysiology. Front. Biosci. 2012;4:1044–1064.
    1. Li J., Stouffs M., Serrander L., Banfi B., Bettiol E., Charnay Y., Steger K., Krause K.H., Jaconi M.E. The NADPH oxidase NOX4 drives cardiac differentiation: Role in regulating cardiac transcription factors and map kinase activation. Mol. Biol. Cell. 2006;17:3978–3988. doi: 10.1091/mbc.E05-06-0532.
    1. Martin-Garrido A., Brown D.I., Lyle A.N., Dikalova A., Seidel-Rogol B., Lassegue B., San Martin A., Griendling K.K. NADPH oxidase 4 mediates TGF-β-induced smooth muscle α-actin via p38MAPK and serum response factor. Free Radic. Biol. Med. 2011;50:354–362. doi: 10.1016/j.freeradbiomed.2010.11.007.
    1. Szocs K., Lassegue B., Sorescu D., Hilenski L.L., Valppu L., Couse T.L., Wilcox J.N., Quinn M.T., Lambeth J.D., Griendling K.K. Upregulation of NOX-based NAD(P)H oxidases in restenosis after carotid injury. Arterioscler. Thromb. Vasc. Biol. 2002;22:21–27. doi: 10.1161/hq0102.102189.
    1. Brandes R.P., Takac I., Schröder K. No superoxide—No stress? Arterioscler. Thromb. Vasc. Biol. 2011;31:1255–1257. doi: 10.1161/ATVBAHA.111.226894.
    1. Schroder K., Zhang M., Benkhoff S., Mieth A., Pliquett R., Kosowski J., Kruse C., Luedike P., Michaelis U.R., Weissmann N., et al. NOX4 is a protective reactive oxygen species generating vascular NADPH oxidase. Circ. Res. 2012;110:1217–1225. doi: 10.1161/CIRCRESAHA.112.267054.
    1. Alvarez L.A., Kovacic L., Rodriguez J., Gosemann J.H., Kubica M., Pircalabioru G.G., Friedmacher F., Cean A., Ghise A., Sarandan M.B., et al. NADPH oxidase-derived H2O2 subverts pathogen signaling by oxidative phosphotyrosine conversion to PB-DOPA. Proc. Natl. Acad. Sci. USA. 2016;113:10406–10411. doi: 10.1073/pnas.1605443113.
    1. Winterbourn C.C., Hampton M.B., Livesey J.H., Kettle A.J. Modeling the reactions of superoxide and myeloperoxidase in the neutrophil phagosome: Implications for microbial killing. J. Biol. Chem. 2006;281:39860–39869. doi: 10.1074/jbc.M605898200.
    1. Li X.J., Tian W., Stull N.D., Grinstein S., Atkinson S., Dinauer M.C. A fluorescently tagged c-terminal fragment of p47 phox detects NADPH oxidase dynamics during phagocytosis. Mol. Biol. Cell. 2009;20:1520–1532. doi: 10.1091/mbc.E08-06-0620.
    1. Rosen H., Klebanoff S.J., Wang Y., Brot N., Heinecke J.W., Fu X. Methionine oxidation contributes to bacterial killing by the myeloperoxidase system of neutrophils. Proc. Natl. Acad. Sci. USA. 2009;106:18686–18691. doi: 10.1073/pnas.0909464106.
    1. Winterbourn C.C., Kettle A.J., Hampton M.B. Reactive oxygen species and neutrophil function. Ann. Rev. Biochem. 2016;85:765–792. doi: 10.1146/annurev-biochem-060815-014442.
    1. Giorgio M., Migliaccio E., Orsini F., Paolucci D., Moroni M., Contursi C., Pelliccia G., Luzi L., Minucci S., Marcaccio M., et al. Electron transfer between cytochrome c and p66 Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell. 2005;122:221–233. doi: 10.1016/j.cell.2005.05.011.
    1. Pagano P.J., Clark J.K., Cifuentes-Pagano M.E., Clark S.M., Callis G.M., Quinn M.T. Localization of a constitutively active, phagocyte-like NADPH oxidase in rabbit aortic adventitia: Enhancement by angiotensin II. Proc. Natl. Acad. Sci. USA. 1997;94:14483–14488. doi: 10.1073/pnas.94.26.14483.
    1. Bedard K., Jaquet V., Krause K.H. NOX5: From basic biology to signaling and disease. Free Radic. Biol. Med. 2012;52:725–734. doi: 10.1016/j.freeradbiomed.2011.11.023.
    1. Nisimoto Y., Diebold B.A., Cosentino-Gomes D., Lambeth J.D. NOX4: A hydrogen peroxide-generating oxygen sensor. Biochemistry. 2014;53:5111–5120. doi: 10.1021/bi500331y.
    1. Lassegue B., Sorescu D., Szocs K., Yin Q., Akers M., Zhang Y., Grant S.L., Lambeth J.D., Griendling K.K. Novel GP91(phox) homologues in vascular smooth muscle cells: NOX1 mediates angiotensin II-induced superoxide formation and redox-sensitive signaling pathways. Circ. Res. 2001;88:888–894. doi: 10.1161/hh0901.090299.
    1. San Jose G., Bidegain J., Robador P.A., Diez J., Fortuno A., Zalba G. Insulin-induced NADPH oxidase activation promotes proliferation and matrix metalloproteinase activation in monocytes/macrophages. Free Radic. Biol. Med. 2009;46:1058–1067. doi: 10.1016/j.freeradbiomed.2009.01.009.
    1. Ebrahimian T., Li M.W., Lemarie C.A., Simeone S.M., Pagano P.J., Gaestel M., Paradis P., Wassmann S., Schiffrin E.L. Mitogen-activated protein kinase-activated protein kinase 2 in angiotensin II-induced inflammation and hypertension: Regulation of oxidative stress. Hypertension. 2011;57:245–254. doi: 10.1161/HYPERTENSIONAHA.110.159889.
    1. Sahoo S., Meijles D.N., Pagano P.J. NADPH oxidases: Key modulators in aging and age-related cardiovascular diseases? Clin. Sci. 2016;130:317–335. doi: 10.1042/CS20150087.
    1. Guzik T.J., West N.E., Black E., McDonald D., Ratnatunga C., Pillai R., Channon K.M. Vascular superoxide production by NAD(P)H oxidase: Association with endothelial dysfunction and clinical risk factors. Circ. Res. 2000;86:E85–E90. doi: 10.1161/01.RES.86.9.e85.
    1. Sorescu D., Weiss D., Lassegue B., Clempus R.E., Szocs K., Sorescu G.P., Valppu L., Quinn M.T., Lambeth J.D., Vega J.D., et al. Superoxide production and expression of NOX family proteins in human atherosclerosis. Circulation. 2002;105:1429–1435. doi: 10.1161/01.CIR.0000012917.74432.66.
    1. Guzik T.J., Chen W., Gongora M.C., Guzik B., Lob H.E., Mangalat D., Hoch N., Dikalov S., Rudzinski P., Kapelak B., et al. Calcium-dependent NOX5 nicotinamide adenine dinucleotide phosphate oxidase contributes to vascular oxidative stress in human coronary artery disease. J. Am. Coll. Cardiol. 2008;52:1803–1809. doi: 10.1016/j.jacc.2008.07.063.
    1. Barry-Lane P.A., Patterson C., van der Merwe M., Hu Z., Holland S.M., Yeh E.T., Runge M.S. P47 phox is required for atherosclerotic lesion progression in ApoE(−/−) mice. J. Clin. Investig. 2001;108:1513–1522. doi: 10.1172/JCI200111927.
    1. Judkins C.P., Diep H., Broughton B.R., Mast A.E., Hooker E.U., Miller A.A., Selemidis S., Dusting G.J., Sobey C.G., Drummond G.R. Direct evidence of a role for Nox2 in superoxide production, reduced nitric oxide bioavailability, and early atherosclerotic plaque formation in ApoE−/− mice. Am. J. Physiol. Heart Circ. Physiol. 2010;298:H24–H32. doi: 10.1152/ajpheart.00799.2009.
    1. Sheehan A.L., Carrell S., Johnson B., Stanic B., Banfi B., Miller F.J. Role for Nox1 NADPH oxidase in atherosclerosis. Atherosclerosis. 2011;216:321–326. doi: 10.1016/j.atherosclerosis.2011.02.028.
    1. Vendrov A.E., Vendrov K.C., Smith A., Yuan J., Sumida A., Robidoux J., Runge M.S., Madamanchi N.R. NOX4 NADPH oxidase-dependent mitochondrial oxidative stress in aging-associated cardiovascular disease. Antioxid. Redox Signal. 2015;23:1389–1409. doi: 10.1089/ars.2014.6221.
    1. Gray S.P., Di Marco E., Okabe J., Szyndralewiez C., Heitz F., Montezano A.C., de Haan J.B., Koulis C., El-Osta A., Andrews K.L., et al. NADPH oxidase 1 plays a key role in diabetes mellitus-accelerated atherosclerosis. Circulation. 2013;127:1888–1902. doi: 10.1161/CIRCULATIONAHA.112.132159.
    1. Maalouf R.M., Eid A.A., Gorin Y.C., Block K., Escobar G.P., Bailey S., Abboud H.E. NOX4-derived reactive oxygen species mediate cardiomyocyte injury in early type 1 diabetes. Am. J. Physiol. Cell Physiol. 2012;302:C597–C604. doi: 10.1152/ajpcell.00331.2011.
    1. Green D.E., Murphy T.C., Kang B.Y., Kleinhenz J.M., Szyndralewiez C., Page P., Sutliff R.L., Hart C.M. The Nox4 inhibitor GKT137831 attenuates hypoxia-induced pulmonary vascular cell proliferation. Am. J. Respir. Cell Mol. Biol. 2012;47:718–726. doi: 10.1165/rcmb.2011-0418OC.
    1. Norton C.E., Broughton B.R., Jernigan N.L., Walker B.R., Resta T.C. Enhanced depolarization-induced pulmonary vasoconstriction following chronic hypoxia requires EGFR-dependent activation of NAD(P)H oxidase 2. Antioxid. Redox Signal. 2013;18:1777–1788. doi: 10.1089/ars.2012.4836.
    1. Reilly S.N., Jayaram R., Nahar K., Antoniades C., Verheule S., Channon K.M., Alp N.J., Schotten U., Casadei B. Atrial sources of reactive oxygen species vary with the duration and substrate of atrial fibrillation: Implications for the antiarrhythmic effect of statins. Circulation. 2011;124:1107–1117. doi: 10.1161/CIRCULATIONAHA.111.029223.
    1. Nishino T., Okamoto K., Eger B.T., Pai E.F., Nishino T. Mammalian xanthine oxidoreductase-mechanism of transition from xanthine dehydrogenase to xanthine oxidase. FEBS J. 2008;275:3278–3289. doi: 10.1111/j.1742-4658.2008.06489.x.
    1. McCord J.M., Fridovich I. The reduction of cytochrome c by milk xanthine oxidase. J. Biol. Chem. 1968;243:5753–5760.
    1. Maxwell A.J., Bruinsma K.A. Uric acid is closely linked to vascular nitric oxide activity. Evidence for mechanism of association with cardiovascular disease. J. Am. Coll. Cardiol. 2001;38:1850–1858. doi: 10.1016/S0735-1097(01)01643-6.
    1. Doehner W., Anker S.D. Xanthine oxidase inhibition for chronic heart failure: Is allopurinol the next therapeutic advance in heart failure? Heart. 2005;91:707–709. doi: 10.1136/hrt.2004.057190.
    1. Wei L., Fahey T., Struthers A.D., MacDonald T.M. Association between allopurinol and mortality in heart failure patients: A long-term follow-up study. Int. J. Clin. Pract. 2009;63:1327–1333. doi: 10.1111/j.1742-1241.2009.02118.x.
    1. Noman A., Ang D.S., Ogston S., Lang C.C., Struthers A.D. Effect of high-dose allopurinol on exercise in patients with chronic stable angina: A randomised, placebo controlled crossover trial. Lancet. 2010;375:2161–2167. doi: 10.1016/S0140-6736(10)60391-1.
    1. Higgins P., Dawson J., Lees K.R., McArthur K., Quinn T.J., Walters M.R. Xanthine oxidase inhibition for the treatment of cardiovascular disease: A systematic review and meta-analysis. Cardiovasc. Ther. 2012;30:217–226. doi: 10.1111/j.1755-5922.2011.00277.x.
    1. Jankov R.P., Kantores C., Pan J., Belik J. Contribution of xanthine oxidase-derived superoxide to chronic hypoxic pulmonary hypertension in neonatal rats. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008;294:L233–L245. doi: 10.1152/ajplung.00166.2007.
    1. Banks M.F., Gerasimovskaya E.V., Tucker D.A., Frid M.G., Carpenter T.C., Stenmark K.R. Egr-1 antisense oligonucleotides inhibit hypoxia-induced proliferation of pulmonary artery adventitial fibroblasts. J. Appl. Physiol. 2005;98:732–738. doi: 10.1152/japplphysiol.00821.2004.
    1. Li C.J., Ning W., Matthay M.A., Feghali-Bostwick C.A., Choi A.M. Mapk pathway mediates egr-1-hsp70-dependent cigarette smoke-induced chemokine production. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007;292:L1297–L1303. doi: 10.1152/ajplung.00194.2006.
    1. Givertz M.M., Mann D.L., Lee K.L., Ibarra J.C., Velazquez E.J., Hernandez A.F., Mascette A.M., Braunwald E. Xanthine oxidase inhibition for hyperuricemic heart failure patients. Circ. Heart Fail. 2013;6:862–868. doi: 10.1161/CIRCHEARTFAILURE.113.000394.
    1. Van Leyen K. Lipoxygenase: An emerging target for stroke therapy. CNS Neurol. Disord. Drug Targets. 2013;12:191–199. doi: 10.2174/18715273112119990053.
    1. Funk C.D., Cyrus T. 12/15-lipoxygenase, oxidative modification of LDL and atherogenesis. Trends Cardiovasc. Med. 2001;11:116–124.
    1. Lotzer K., Funk C.D., Habenicht A.J. The 5-lipoxygenase pathway in arterial wall biology and atherosclerosis. Biochim. Biophys. Acta. 2005;1736:30–37. doi: 10.1016/j.bbalip.2005.07.001.
    1. Back M., Hansson G.K. Leukotriene receptors in atherosclerosis. Ann. Med. 2006;38:493–502. doi: 10.1080/07853890600982737.
    1. Funk C.D. Leukotriene modifiers as potential therapeutics for cardiovascular disease. Nat. Rev. Drug Discov. 2005;4:664–672. doi: 10.1038/nrd1796.
    1. Flamand N., Mancuso P., Serezani C.H., Brock T.G. Leukotrienes: Mediators that have been typecast as villains. Cell. Mol. Life Sci. CMLS. 2007;64:2657–2670. doi: 10.1007/s00018-007-7228-2.
    1. Serhan C.N., Chiang N., Van Dyke T.E. Resolving inflammation: Dual anti-inflammatory and pro-resolution lipid mediators. Nat. Rev. Immunol. 2008;8:349–361. doi: 10.1038/nri2294.
    1. Poeckel D., Funk C.D. The 5-lipoxygenase/leukotriene pathway in preclinical models of cardiovascular disease. Cardiovasc. Res. 2010;86:243–253. doi: 10.1093/cvr/cvq016.
    1. Shekher A., Singh M. Role of eicosanoid inhibition of ischemia reperfusion injury: Intact and isolated rat heart studies. Methods Find. Exp. Clin. Pharmacol. 1997;19:223–229.
    1. Adamek A., Jung S., Dienesch C., Laser M., Ertl G., Bauersachs J., Frantz S. Role of 5-lipoxygenase in myocardial ischemia-reperfusion injury in mice. Eur. J. Pharmacol. 2007;571:51–54. doi: 10.1016/j.ejphar.2007.05.040.
    1. Chinetti-Gbaguidi G., Baron M., Bouhlel M.A., Vanhoutte J., Copin C., Sebti Y., Derudas B., Mayi T., Bories G., Tailleux A., et al. Human atherosclerotic plaque alternative macrophages display low cholesterol handling but high phagocytosis because of distinct activities of the PPARγ and LXRα pathways. Circ. Res. 2011;108:985–995. doi: 10.1161/CIRCRESAHA.110.233775.
    1. Levick S.P., Loch D.C., Taylor S.M., Janicki J.S. Arachidonic acid metabolism as a potential mediator of cardiac fibrosis associated with inflammation. J. Immunol. 2007;178:641–646. doi: 10.4049/jimmunol.178.2.641.
    1. Wen Y., Gu J., Chakrabarti S.K., Aylor K., Marshall J., Takahashi Y., Yoshimoto T., Nadler J.L. The role of 12/15-lipoxygenase in the expression of interleukin-6 and tumor necrosis factor-α in macrophages. Endocrinology. 2007;148:1313–1322. doi: 10.1210/en.2006-0665.
    1. Hansson G.K., Robertson A.K., Soderberg-Naucler C. Inflammation and atherosclerosis. Ann. Rev. Pathol. 2006;1:297–329. doi: 10.1146/annurev.pathol.1.110304.100100.
    1. Suzuki H., Kayama Y., Sakamoto M., Iuchi H., Shimizu I., Yoshino T., Katoh D., Nagoshi T., Tojo K., Minamino T., et al. Arachidonate 12/15-lipoxygenase-induced inflammation and oxidative stress are involved in the development of diabetic cardiomyopathy. Diabetes. 2015;64:618–630. doi: 10.2337/db13-1896.
    1. Baldus S., Rudolph V., Roiss M., Ito W.D., Rudolph T.K., Eiserich J.P., Sydow K., Lau D., Szocs K., Klinke A., et al. Heparins increase endothelial nitric oxide bioavailability by liberating vessel-immobilized myeloperoxidase. Circulation. 2006;113:1871–1878. doi: 10.1161/CIRCULATIONAHA.105.590083.
    1. Anatoliotakis N., Deftereos S., Bouras G., Giannopoulos G., Tsounis D., Angelidis C., Kaoukis A., Stefanadis C. Myeloperoxidase: Expressing inflammation and oxidative stress in cardiovascular disease. Curr. Top. Med. Chem. 2013;13:115–138. doi: 10.2174/1568026611313020004.
    1. Van der Veen B.S., de Winther M.P., Heeringa P. Myeloperoxidase: Molecular mechanisms of action and their relevance to human health and disease. Antioxid. Redox Signal. 2009;11:2899–2937. doi: 10.1089/ars.2009.2538.
    1. Fu X., Kassim S.Y., Parks W.C., Heinecke J.W. Hypochlorous acid oxygenates the cysteine switch domain of pro-matrilysin (MMP-7). A mechanism for matrix metalloproteinase activation and atherosclerotic plaque rupture by myeloperoxidase. J. Biol. Chem. 2001;276:41279–41287. doi: 10.1074/jbc.M106958200.
    1. Pullar J.M., Vissers M.C., Winterbourn C.C. Glutathione oxidation by hypochlorous acid in endothelial cells produces glutathione sulfonamide as a major product but not glutathione disulfide. J. Biol. Chem. 2001;276:22120–22125. doi: 10.1074/jbc.M102088200.
    1. Hawkins C.L. The role of hypothiocyanous acid (HOSCN) in biological systems. Free Radic. Res. 2009;43:1147–1158. doi: 10.3109/10715760903214462.
    1. Lane A.E., Tan J.T., Hawkins C.L., Heather A.K., Davies M.J. The myeloperoxidase-derived oxidant HOSCN inhibits protein tyrosine phosphatases and modulates cell signalling via the mitogen-activated protein kinase (MAPK) pathway in macrophages. Biochem. J. 2010;430:161–169. doi: 10.1042/BJ20100082.
    1. Podrez E.A., Schmitt D., Hoff H.F., Hazen S.L. Myeloperoxidase-generated reactive nitrogen species convert LDL into an atherogenic form in vitro. J. Clin. Investing. 1999;103:1547–1560. doi: 10.1172/JCI5549.
    1. Bergt C., Pennathur S., Fu X., Byun J., O’Brien K., McDonald T.O., Singh P., Anantharamaiah G.M., Chait A., Brunzell J., et al. The myeloperoxidase product hypochlorous acid oxidizes HDL in the human artery wall and impairs ABCA1-dependent cholesterol transport. Proc. Natl. Acad. Sci. USA. 2004;101:13032–13037. doi: 10.1073/pnas.0405292101.
    1. Exner M., Hermann M., Hofbauer R., Hartmann B., Kapiotis S., Gmeiner B. Thiocyanate catalyzes myeloperoxidase-initiated lipid oxidation in LDL. Free Radic. Biol. Med. 2004;37:146–155. doi: 10.1016/j.freeradbiomed.2004.04.039.
    1. Baldus S., Heitzer T., Eiserich J.P., Lau D., Mollnau H., Ortak M., Petri S., Goldmann B., Duchstein H.J., Berger J., et al. Myeloperoxidase enhances nitric oxide catabolism during myocardial ischemia and reperfusion. Free Radic. Biol. Med. 2004;37:902–911. doi: 10.1016/j.freeradbiomed.2004.06.003.
    1. Wang Z., Nicholls S.J., Rodriguez E.R., Kummu O., Horkko S., Barnard J., Reynolds W.F., Topol E.J., DiDonato J.A., Hazen S.L. Protein carbamylation links inflammation, smoking, uremia and atherogenesis. Nat. Med. 2007;13:1176–1184. doi: 10.1038/nm1637.
    1. Vozarova B., Weyer C., Lindsay R.S., Pratley R.E., Bogardus C., Tataranni P.A. High white blood cell count is associated with a worsening of insulin sensitivity and predicts the development of type 2 diabetes. Diabetes. 2002;51:455–461. doi: 10.2337/diabetes.51.2.455.
    1. Orio F., Jr., Palomba S., Cascella T., Di Biase S., Manguso F., Tauchmanova L., Nardo L.G., Labella D., Savastano S., Russo T., et al. The increase of leukocytes as a new putative marker of low-grade chronic inflammation and early cardiovascular risk in polycystic ovary syndrome. J. Clin. Endocrinol. Metab. 2005;90:2–5. doi: 10.1210/jc.2004-0628.
    1. Rovira-Llopis S., Rocha M., Falcon R., de Pablo C., Alvarez A., Jover A., Hernandez-Mijares A., Victor V.M. Is myeloperoxidase a key component in the ROS-induced vascular damage related to nephropathy in type 2 diabetes? Antioxid. Redox Signal. 2013;19:1452–1458. doi: 10.1089/ars.2013.5307.
    1. Victor V.M., Rovira-Llopis S., Banuls C., Diaz-Morales N., Martinez de Maranon A., Rios-Navarro C., Alvarez A., Gomez M., Rocha M., Hernandez-Mijares A. Insulin resistance in PCOS patients enhances oxidative stress and leukocyte adhesion: Role of myeloperoxidase. PLoS ONE. 2016;11:e0151960. doi: 10.1371/journal.pone.0151960.
    1. Ribeiro A.L., Scapinelli A., Tamanaha S., Oliveira R.M., Kowastch I., Mathias W., Jr., Aoki T., Aldrighi J.M. Myeloperoxidases and polycystic ovary syndrome. Gynecol. Endocrinol. 2012;28:3–6. doi: 10.3109/09513590.2011.579656.
    1. Shishehbor M.H., Brennan M.L., Aviles R.J., Fu X., Penn M.S., Sprecher D.L., Hazen S.L. Statins promote potent systemic antioxidant effects through specific inflammatory pathways. Circulation. 2003;108:426–431. doi: 10.1161/01.CIR.0000080895.05158.8B.
    1. Laursen J.B., Rajagopalan S., Galis Z., Tarpey M., Freeman B.A., Harrison D.G. Role of superoxide in angiotensin II-induced but not catecholamine-induced hypertension. Circulation. 1997;95:588–593. doi: 10.1161/01.CIR.95.3.588.
    1. Ono H., Minatoguchi S., Watanabe K., Yamada Y., Mizukusa T., Kawasaki H., Takahashi H., Uno T., Tsukamoto T., Hiei K., et al. Candesartan decreases carotid intima-media thickness by enhancing nitric oxide and decreasing oxidative stress in patients with hypertension. Hypertens. Res. 2008;31:271–279. doi: 10.1291/hypres.31.271.
    1. Hernandez R.H., Armas-Hernandez M.J., Velasco M., Israili Z.H., Armas-Padilla M.C. Calcium antagonists and atherosclerosis protection in hypertension. Am. J. Ther. 2003;10:409–414. doi: 10.1097/00045391-200311000-00006.
    1. Costanzo P., Perrone-Filardi P., Petretta M., Marciano C., Vassallo E., Gargiulo P., Paolillo S., Petretta A., Chiariello M. Calcium channel blockers and cardiovascular outcomes: A meta-analysis of 175,634 patients. J. Hypertens. 2009;27:1136–1151. doi: 10.1097/HJH.0b013e3283281254.
    1. Dewey C.M., Spitler K.M., Ponce J.M., Hall D.D., Grueter C.E. Cardiac-secreted factors as peripheral metabolic regulators and potential disease biomarkers. J. Am. Heart Assoc. 2016 doi: 10.1161/JAHA.115.003101.
    1. Donato A.J., Walker A.E., Magerko K.A., Bramwell R.C., Black A.D., Henson G.D., Lawson B.R., Lesniewski L.A., Seals D.R. Life-long caloric restriction reduces oxidative stress and preserves nitric oxide bioavailability and function in arteries of old mice. Aging Cell. 2013;12:772–783. doi: 10.1111/acel.12103.
    1. Gano L.B., Donato A.J., Pasha H.M., Hearon C.M., Jr., Sindler A.L., Seals D.R. The sirt1 activator srt1720 reverses vascular endothelial dysfunction, excessive superoxide production, and inflammation with aging in mice. Am. J. Physiol. Heart Circ. Physiol. 2014;307:H1754–H1763. doi: 10.1152/ajpheart.00377.2014.
    1. Koentges C., Bode C., Bugger H. SIRT3 in Cardiac Physiology and Disease. Front. Cardiovasc. Med. 2016;13:3–38. doi: 10.3389/fcvm.2016.00038.
    1. De Picciotto N.E., Gano L.B., Johnson L.C., Martens C.R., Sindler A.L., Mills K.F., Imai S., Seals D.R. Nicotinamide mononucleotide supplementation reverses vascular dysfunction and oxidative stress with aging in mice. Aging Cell. 2016;15:522–530. doi: 10.1111/acel.12461.
    1. LaRocca T.J., Hearon C.M., Jr., Henson G.D., Seals D.R. Mitochondrial quality control and age-associated arterial stiffening. Exp. Gerontol. 2014;58:78–82. doi: 10.1016/j.exger.2014.07.008.
    1. Kaplon R.E., Hill S.D., Bispham N.Z., Santos-Parker J.R., Nowlan M.J., Snyder L.L., Chonchol M., LaRocca T.J., McQueen M.B., Seals D.R. Oral trehalose supplementation improves resistance artery endothelial function in healthy middle-aged and older adults. Aging. 2016;8:1167–1183. doi: 10.18632/aging.100962.
    1. Gallego-Villar L., Perez B., Ugarte M., Desviat L.R., Richard E. Antioxidants successfully reduce ROS production in propionic acidemia fibroblasts. Biochem. Biophys. Res. Commun. 2014;452:457–461. doi: 10.1016/j.bbrc.2014.08.091.
    1. Escribano-Lopez I., Diaz-Morales N., Rovira-Llopis S., de Marañon A.M., Orden S., Alvarez A., Bañuls C., Rocha M., Murphy M.P., Hernandez-Mijares A., et al. The mitochondria-targeted antioxidant mitoq modulates oxidative stress, inflammation and leukocyte-endothelium interactions in leukocytes isolated from type 2 diabetic patients. Redox Biol. 2016;10:200–205. doi: 10.1016/j.redox.2016.10.017.
    1. Hassinen I., Hartikainen J., Hedman A., Kivelä A., Saraste A., Knuuti J., Husso M., Mussalo H., Hedman M., Toivanen P., et al. Abstract 11987: Adenoviral intramyocardial VEGF-D gene transfer increases myocardial perfusion in refractory angina patients. Circulation. 2015;132:A11987.
    1. Van Wanrooij E.J., de Vos P., Bixel M.G., Vestweber D., van Berkel T.J., Kuiper J. Vaccination against CD99 inhibits atherogenesis in low-density lipoprotein receptor-deficient mice. Cardiovasc. Res. 2008;78:590–596. doi: 10.1093/cvr/cvn025.
    1. Deshpande D., Kethireddy S., Janero D.R., Amiji M.M. Therapeutic efficacy of an ω-3-fatty acid-containing 17-β estradiol nano-delivery system against experimental atherosclerosis. PLoS ONE. 2016;11:e0147337. doi: 10.1371/journal.pone.0147337.
    1. Paul A., Hasan A., Kindi H.A., Gaharwar A.K., Rao V.T., Nikkhah M., Shin S.R., Krafft D., Dokmeci M.R., Shum-Tim D., et al. Injectable graphene oxide/hydrogel-based angiogenic gene delivery system for vasculogenesis and cardiac repair. ACS Nano. 2014;8:8050–8062. doi: 10.1021/nn5020787.
    1. Wang B., Qian H., Yang H., Xu L., Xu W., Yan J. Regression of atherosclerosis plaques in apolipoprotein e−/− mice after lentivirus-mediated RNA interference of CD40. Int. J. Cardiol. 2013;163:34–39. doi: 10.1016/j.ijcard.2011.05.053.
    1. Kowalski P.S., Lintermans L.L., Morselt H.W., Leus N.G., Ruiters M.H., Molema G., Kamps J.A. Anti-VCAM-1 and Anti-E-selectin SAINT-O-Somes for selective delivery of siRNA into inflammation-activated primary endothelial cells. Mol. Pharm. 2013;10:3033–3044. doi: 10.1021/mp4001124.
    1. Lee R.G., Crosby J., Baker B.F., Graham M.J., Crooke R.M. Antisense technology: An emerging platform for cardiovascular disease therapeutics. J. Cardiovasc. Transl. Res. 2013;6:969–980. doi: 10.1007/s12265-013-9495-7.
    1. Philippen L.E., Dirkx E., Wit J.B., Burggraaf K., de Windt L.J., da Costa Martins P.A. Antisense microrna therapeutics in cardiovascular disease: Quo vadis? Mol. Ther. 2015;23:1810–1818. doi: 10.1038/mt.2015.133.
    1. Wang P., Yang Y., Hong H., Zhang Y., Cai W., Fang D. Aptamers as therapeutics in cardiovascular diseases. Curr. Med. Chem. 2011;18:4169–4174. doi: 10.2174/092986711797189673.
    1. Du H., Li L., Bennett D., Guo Y., Key T.J., Bian Z., Sherliker P., Gao H., Chen Y., Yang L., et al. Fresh fruit consumption and major cardiovascular disease in china. N. Engl. J. Med. 2016;374:1332–1343. doi: 10.1056/NEJMoa1501451.
    1. Toh J.Y., Tan V.M., Lim P.C., Lim S.T., Chong M.F. Flavonoids from fruit and vegetables: A focus on cardiovascular risk factors. Curr. Atheroscler. Rep. 2013;15:368. doi: 10.1007/s11883-013-0368-y.
    1. Noratto G., Martino H.S., Simbo S., Byrne D., Mertens-Talcott S.U. Consumption of polyphenol-rich peach and plum juice prevents risk factors for obesity-related metabolic disorders and cardiovascular disease in zucker rats. J. Nutr. Biochem. 2015;26:633–641. doi: 10.1016/j.jnutbio.2014.12.014.
    1. Pereira B.L., Arruda F.C., Reis P.P., Felix T.F., Santos P.P., Rafacho B.P., Goncalves A.F., Claro R.T., Azevedo P.S., Polegato B.F., et al. Tomato (Lycopersicon esculentum) supplementation induces changes in cardiac mirna expression, reduces oxidative stress and left ventricular mass, and improves diastolic function. Nutrients. 2015;7:9640–9649. doi: 10.3390/nu7115493.
    1. Hooper L., Kay C., Abdelhamid A., Kroon P.A., Cohn J.S., Rimm E.B., Cassidy A. Effects of chocolate, cocoa, and flavan-3-ols on cardiovascular health: A systematic review and meta-analysis of randomized trials. Am. J. Clin. Nutr. 2012;95:740–751. doi: 10.3945/ajcn.111.023457.
    1. Ghayur M.N., Gilani A.H. Ginger lowers blood pressure through blockade of voltage-dependent calcium channels. J. Cardiovasc. Pharmacol. 2005;45:74–80. doi: 10.1097/00005344-200501000-00013.
    1. Hosseini B., Saedisomeolia A., Allman-Farinelli M. Association between antioxidant intake/status and obesity: A systematic review of observational studies. Biol. Trace Elem. Res. 2017;175:287–297. doi: 10.1007/s12011-016-0785-1.
    1. Srinivasan M., Sudheer A.R., Menon V.P. Ferulic acid: Therapeutic potential through its antioxidant property. J. Clin. Biochem. Nutr. 2007;40:92–100. doi: 10.3164/jcbn.40.92.
    1. Liu J.F., Liu Y.H., Chen C.M., Chang W.H., Chen C.Y. The effect of almonds on inflammation and oxidative stress in Chinese patients with type 2 diabetes mellitus: A randomized crossover controlled feeding trial. Eur. J. Nutr. 2012;52:927–935. doi: 10.1007/s00394-012-0400-y.
    1. Lopez-Uriarte P., Nogues R., Saez G., Bullo M., Romeu M., Masana L. Effect of nut consumption on oxidative stress and the endothelial function in metabolic syndrome. Clin. Nutr. 2010;29:373–380. doi: 10.1016/j.clnu.2009.12.008.
    1. Bjørklund G., Chirumbolo S. Role of oxidative stress and antioxidants in daily nutrition and human health. Nutrition. 2017;33:311–321. doi: 10.1016/j.nut.2016.07.018.
    1. Klotz L.O., Sánchez-Ramos C., Prieto-Arroyo I., Urbánek P., Steinbrenner H., Monsalve M. Redox regulation of FoxO transcription factors. Redox Biol. 2015;6:51–72. doi: 10.1016/j.redox.2015.06.019.
    1. Montague C.T., Farooqi I.S., Whitehead J.P., Soos M.A., Rau H., Wareham N.J. Congenital leptin deficiency is associated with severe early-onset obesity in humans. Nature. 1997;387:903–908.

Source: PubMed

3
Subscribe