Abstract

HomeHypertensionVol. 53, No. 2Mitochondrial Dysfunction and Mitochondrial-Produced Reactive Oxygen Species Free AccessEditorialPDF/EPUBAboutView PDFView EPUBSections ToolsAdd to favoritesDownload citationsTrack citationsPermissions ShareShare onFacebookTwitterLinked InMendeleyReddit Jump toFree AccessEditorialPDF/EPUBMitochondrial Dysfunction and Mitochondrial-Produced Reactive Oxygen SpeciesNew Targets for Neurogenic Hypertension? Matthew C. Zimmerman and Irving H. Zucker Matthew C. ZimmermanMatthew C. Zimmerman From the Department of Cellular and Integrative Physiology (M.C.Z., I.H.Z.), University of Nebraska Medical Center, Omaha; and the Redox Biology Center (M.C.Z.), University of Nebraska, Lincoln. Search for more papers by this author and Irving H. ZuckerIrving H. Zucker From the Department of Cellular and Integrative Physiology (M.C.Z., I.H.Z.), University of Nebraska Medical Center, Omaha; and the Redox Biology Center (M.C.Z.), University of Nebraska, Lincoln. Search for more papers by this author Originally published29 Dec 2008https://doi.org/10.1161/HYPERTENSIONAHA.108.125567Hypertension. 2009;53:112–114Other version(s) of this articleYou are viewing the most recent version of this article. Previous versions: December 29, 2008: Previous Version 1 Over the past 10 to 15 years, a vast collection of studies have provided evidence indicating that reactive oxygen species (ROS), particularly superoxide (O2·−) and hydrogen peroxide (H2O2), contribute to the pathogenesis of cardiovascular diseases, such as heart failure and hypertension. Griendling et al1 first demonstrated that NADPH oxidase present in the vasculature is a primary source of the elevated ROS levels. Since these initial studies, NADPH oxidase-derived ROS in the kidney,2 heart,3 and brain4 have been linked to the development and progression of numerous cardiovascular-related diseases. More recently, however, mitochondria have also been identified as important sources of ROS in controlling cardiovascular function. Considering that mitochondria are the primary source of ROS in most cells during normal respiration because of the leaking of electrons from the electron transport chain (ETC), perhaps it should not be all that surprising that mitochondrial-produced ROS are involved in pathophysiological conditions of the cardiovascular system.To date, most of the evidence linking mitochondrial dysfunction and mitochondrial-produced ROS to the pathogenesis of cardiovascular diseases comes from studies on the peripheral renin-angiotensin system.5 For example, using a model of cardiac ischemic reperfusion injury, Kimura et al6 reported that angiotensin II (Ang II)-induced preconditioning is mediated by mitochondrial-produced ROS. The authors further demonstrated that Ang II-induced NADPH oxidase-derived ROS lie upstream of mitochondrial-produced ROS, thus, implicating a ROS-induced ROS mechanism. Similarly, it was demonstrated recently that, in aortic endothelial cells, Ang II-induced NADPH oxidase activation leads to an increase in mitochondrial ROS production, as well as mitochondrial dysfunction, as determined by a decrease in mitochondrial membrane potential and mitochondrial respiration.7 Together, these studies and others (detailed elsewhere5) clearly illustrate a role for mitochondrial-produced ROS and mitochondrial dysfunction in peripheral tissues in the pathogenesis of cardiovascular diseases, primarily those associated with increased Ang II signaling. However, in the central nervous system, the contribution of defective mitochondria and mitochondrial-produced ROS in cardiovascular diseases has been mostly overlooked.In the present issue of Hypertension, Chan et al,8 studying neurogenic hypertension, provide new evidence that, indeed, mitochondrial dysfunction and the subsequent production of mitochondrial-localized ROS in the central nervous system, particularly the rostral ventrolateral medulla (RVLM), play a critical role in cardiovascular function. More specifically, they report a decrease in ETC complex I and complex III activity accompanied by an increase in mitochondrial-produced ROS, particularly O2·− and H2O2, in the RLVM of spontaneously hypertensive rats (SHRs). Direct RVLM administration of coenzyme Q10, a mitochondrial electron transporter and antioxidant, restored electron transport capacity, decreased mitochondrial-localized ROS production, and significantly reduced mean systemic arterial pressure and sympathetic neurogenic vasomotor tone in SHRs. Using mitochondrial ETC inhibitors rotenone and antimycin A, Chan et al8 provide further evidence that mitochondrial dysfunction in the RVLM results in mitochondrial-produced ROS, which, in turn, induce changes in cardiovascular function. In normotensive Wistar-Kyoto rats or prehypertensive SHRs, RVLM administration of rotenone or antimycin A significantly elevated mitochondrial H2O2 production, as well as mean systemic arterial pressure and power density of vasomotor activity. Similar to its action in hypertensive SHRs, coenzyme Q10 markedly attenuated the augmented cardiovascular responses induced by ETC inhibition in normotensive animals. These data indicate that diminished mitochondrial ETC activity and the subsequent production of ROS in the RVLM contribute to the SHR hypertensive phenotype.To address the hypothesis that a cytoplasm-to-mitochondria ROS-induced ROS mechanism in the RLVM is involved in neurogenic hypertension, Chan et al8 used an intracerebroventricular (ICV) Ang II infusion model. ICV Ang II infusion results in an increase in sympathetic tone and blood pressure, at least in part, via NADPH oxidase activation in the RVLM.9 Compared with ICV infusion of artificial cerebrospinal fluid, Ang II infusion decreased ETC complex I through III activity while increasing mitochondrial-produced H2O2 in the RVLM. Inhibition of NADPH oxidase via p22phox antisense significantly attenuated the Ang II-induced increase in mitochondrial H2O2 levels, thus implicating a ROS-induced ROS mechanism initiated by NADPH oxidase-derived ROS. Similar to the SHR experiments, RVLM administration of coenzyme Q10 inhibited the Ang II-induced increase in mitochondrial-produced ROS, mean systemic arterial pressure, and sympathetic vasomotor tone. Together, these studies strongly suggest that increased sympathetic tone and the pathogenesis of neurogenic hypertension are mediated by mitochondrial ETC dysfunction and an ensuing increase in mitochondrial-produced ROS in the RVLM.In an attempt to further support the ROS-induced ROS hypothesis, Chan et al8 used adenoviral-mediated gene transfer to overexpress 3 different antioxidants, copper/zinc superoxide dismutase (SOD1), manganese superoxide dismutase (SOD2), and catalase in the RVLM of SHRs. Previously, this group demonstrated that SOD1, SOD2, or catalase overexpression in the RVLM significantly reduces the elevated arterial pressure of SHRs.10 In the current study, they report that all 3 of the antioxidants restored complex I and III activity and reduced the elevated levels of mitochondrial-localized O2·− and H2O2. Considering that SOD1, a O2·− scavenging enzyme, and catalase, an H2O2 scavenger, are primarily localized in the cytoplasm and peroxisomes, respectively, the authors conclude that ROS generated in the cytosolic compartment induce ETC damage and an increase in mitochondrial-produced ROS. However, alternative interpretations should also be considered. For example, SOD1 is also present in mitochondria,11 and, thus, it is possible that overexpression of mitochondrial-localized SOD1 protects ETC complexes and scavenges mitochondrial-produced ROS. Regarding the catalase overexpression experiments, it remains unclear how the H2O2-scavenging, peroxisome-targeted enzyme reduces both O2·− and H2O2 in RVLM mitochondria of SHRs. Future studies designed to measure cytoplasmic- and mitochondrial-localized ROS simultaneously will help substantiate the ROS-induced ROS mechanism postulated by Chan et al.8Perhaps the most exciting aspect of the study by Chan et al8 is that it provides evidence that damaged ETC complexes are a source for mitochondrial-produced ROS in central nervous system-dependent cardiovascular responses. Until this report, a potential source of mitochondrial-produced ROS in Ang II-stimulated neurons, as suggested in previous studies, remained speculative. For example, Zimmerman et al12 first proposed a role for mitochondrial-produced ROS in brain angiotensinergic signaling by reporting that overexpressing SOD2, the mitochondrial-targeted isoform of SOD, in the brain significantly attenuates the cardiovascular responses induced by ICV administration of Ang II. However, this earlier study failed to identify a potential source of Ang II-induced mitochondrial-produced ROS in central neurons. More recently, Nozoe et al13 also showed that SOD2 overexpression in the brain, notably, the RVLM, attenuates the acute pressor response of Ang II microinjected into the RVLM. Similar to Chan et al,8 Nozoe et al13 suggest that Ang II signaling in neurons involves a ROS-induced ROS mechanism, which starts with NADPH oxidase-derived ROS and ends with mitochondrial-produced ROS. However, in contrast, Nozoe et al13 report that Ang II does not alter the activity of ETC complexes. This discrepancy is likely attributable to the fact that Nozoe et al13 measured the acute effect of Ang II (1-hour stimulation) on ETC complex activity in cultured PC-12 cells, whereas Chan et al8 measured complex activity in vivo in RVLM tissue after 5 days of ICV Ang II infusion. As discussed earlier, the fact that rotenone or antimycin A, 2 ETC inhibitors, microinjected into the RVLM increased mitochondrial-localized ROS, mean systemic arterial pressure, and sympathetic tone strengthens the conclusion by Chan et al8 that, in neurons, damaged ETC complexes are a source of mitochondrial-produced ROS. Nevertheless, further experiments, perhaps using genetic strategies to inhibit ETC activity in central neurons, are required to corroborate this conclusion.In summary, Chan et al8 report a role for mitochondrial dysfunction and mitochondrial-produced ROS in the central nervous system in the pathogenesis of neurogenic hypertension. The data indicate that impaired ETC complexes are a source of mitochondrial-localized ROS and that NADPH oxidase-derived ROS may mediate the impairment of the ETC (Figure). Additional studies are required to examine the downstream mechanism(s) by which mitochondrial-produced ROS increase sympathetic tone and drive the development of hypertension. Such studies should use mitochondrial-targeted antioxidants, including SOD2, and focus on the redox sensitivity of neuronal ion channels, as well as redox control of transcription factors (Figure). The results of these future experiments may strengthen the conclusions by Chan et al8 and may help distinguish damaged ETC complexes and mitochondrial-produced ROS as novel therapeutic targets in neurogenic hypertension. Download figureDownload PowerPointFigure. Proposed Ang II signaling pathway in RVLM neurons involving mitochondrial dysfunction and mitochondrial-produced ROS. Chan et al10 provide evidence indicating that Ang II stimulation of neurons in the RVLM increases NADPH oxidase-derived ROS, which, in turn, damage mitochondrial ETC complexes leading to an increase in mitochondrial-produced ROS (solid-line arrows). Additional experiments using mitochondrial-targeted antioxidants, such as SOD2, are needed to determine the downstream signaling events mediated by mitochondrial-produced ROS. Possible targets of mitochondrial-produced ROS include redox-sensitive transcription factors (TF) and/or ion channels (broken-line arrows).The opinions expressed in this editorial are not necessarily those of the editors or of the American Heart Association.Sources of FundingM.C.Z.’s research is supported by a National Institutes of Health Centers of Biomedical Research Excellence grant awarded to the Redox Biology Center at the University of Nebraska-Lincoln. I.H.Z.’s research is supported by National Institutes of Health grant PO-1 HL62222.DisclosuresNone.FootnotesCorrespondence to Irving H. Zucker, University of Nebraska Medical Center, Department of Cellular and Integrative Physiology, 985850 Nebraska Medical Center, Omaha, NE 68198-5850. E-mail [email protected] References 1 Griendling KK, Minieri CA, Ollerenshaw JD, Alexander RW. Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circ Res. 1994; 74: 1141–1148.CrossrefMedlineGoogle Scholar2 Wilcox CS. Reactive oxygen species: roles in blood pressure and kidney function. Curr Hyperten Rep. 2002; 4: 160–166.CrossrefMedlineGoogle Scholar3 Hingtgen SD, Tian X, Yang J, Dunlay SM, Peek AS, Wu Y, Sharma RV, Engelhardt JF, Davisson RL. Nox2-containing NADPH oxidase and Akt activation play a key role in angiotensin II-induced cardiomyocyte hypertrophy. Physiol Genomics. 2006; 26: 180–191.CrossrefMedlineGoogle Scholar4 Zimmerman MC, Davisson RL. Redox signaling in central neural regulation of cardiovascular function. Prog Biophys Mol Biol. 2004; 84: 125–149.CrossrefMedlineGoogle Scholar5 de Cavanagh EM, Inserra F, Ferder M, Ferder L. From mitochondria to disease: role of the renin-angiotensin system. Am J Nephrol. 2007; 27: 545–553.CrossrefMedlineGoogle Scholar6 Kimura S, Zhang GX, Nishiyama A, Shokoji T, Yao L, Fan YY, Rahman M, Suzuki T, Maeta H, Abe Y. Role of NAD(P)H oxidase- and mitochondria-derived reactive oxygen species in cardioprotection of ischemic reperfusion injury by angiotensin II. Hypertension. 2005; 45: 860–866.LinkGoogle Scholar7 Doughan AK, Harrison DG, Dikalov SI. Molecular mechanisms of angiotensin II-mediated mitochondrial dysfunction: linking mitochondrial oxidative damage and vascular endothelial dysfunction. Circ Res. 2008; 102: 488–496.LinkGoogle Scholar8 Chan SHH, Wu KLH, Chang AYW, Tai M-H, Chan JYH. Oxidative impairment of mitochondrial electron transport chain complexes in rostral ventrolateral medulla contributes to neurogenic hypertension. Hypertension. 2009; 53: 217–227.LinkGoogle Scholar9 Gao L, Wang W, Li YL, Schultz HD, Liu D, Cornish KG, Zucker IH. Sympathoexcitation by central ANG II: roles for AT1 receptor upregulation and NAD(P)H oxidase in RVLM. Am J Physiol Heart Circ Physiol. 2005; 288: H2271–H2279.CrossrefMedlineGoogle Scholar10 Chan SH, Tai MH, Li CY, Chan JY. Reduction in molecular synthesis or enzyme activity of superoxide dismutases and catalase contributes to oxidative stress and neurogenic hypertension in spontaneously hypertensive rats. Free Radic Biol Med. 2006; 40: 2028–2039.CrossrefMedlineGoogle Scholar11 Okado-Matsumoto A, Fridovich I. Subcellular distribution of superoxide dismutases (SOD) in rat liver: Cu,Zn-SOD in mitochondria. J Biol Chem. 2001; 276: 38388–38393.CrossrefMedlineGoogle Scholar12 Zimmerman MC, Lazartigues E, Lang JA, Sinnayah P, Ahmad IM, Spitz DR, Davisson RL. Superoxide mediates the actions of angiotensin II in the central nervous system. Circ Res. 2002; 91: 1038–1045.LinkGoogle Scholar13 Nozoe M, Hirooka Y, Koga Y, Araki S, Konno S, Kishi T, Ide T, Sunagawa K. Mitochondria-derived reactive oxygen species mediate sympathoexcitation induced by angiotensin II in the rostral ventrolateral medulla. J Hypertens. 2008; 26: 2176–2184.CrossrefMedlineGoogle Scholar Previous Back to top Next FiguresReferencesRelatedDetailsCited By Morales P, Arias-Durán C, Ávalos-Guajardo Y, Aedo G, Verdejo H, Parra V and Lavandero S (2020) Emerging role of mitophagy in cardiovascular physiology and pathology, Molecular Aspects of Medicine, 10.1016/j.mam.2019.09.006, 71, (100822), Online publication date: 1-Feb-2020. Mohammed M, Berdasco C and Lazartigues E (2020) Brain angiotensin converting enzyme-2 in central cardiovascular regulation, Clinical Science, 10.1042/CS20200483, 134:19, (2535-2547), Online publication date: 16-Oct-2020. Chan J and Chan S (2019) Differential impacts of brain stem oxidative stress and nitrosative stress on sympathetic vasomotor tone, Pharmacology & Therapeutics, 10.1016/j.pharmthera.2019.05.015, 201, (120-136), Online publication date: 1-Sep-2019. Souders C, Liang X, Wang X, Ector N, Zhao Y and Martyniuk C (2018) High-throughput assessment of oxidative respiration in fish embryos: Advancing adverse outcome pathways for mitochondrial dysfunction, Aquatic Toxicology, 10.1016/j.aquatox.2018.03.031, 199, (162-173), Online publication date: 1-Jun-2018. Yang T, Rodriguez V, Malphurs W, Schmidt J, Ahmari N, Sumners C, Martyniuk C and Zubcevic J (2018) Butyrate regulates inflammatory cytokine expression without affecting oxidative respiration in primary astrocytes from spontaneously hypertensive rats, Physiological Reports, 10.14814/phy2.13732, 6:14, (e13732), Online publication date: 1-Jul-2018. Zubcevic J, Baker A and Martyniuk C (2017) Transcriptional networks in rodent models support a role for gut-brain communication in neurogenic hypertension: a review of the evidence, Physiological Genomics, 10.1152/physiolgenomics.00010.2017, 49:7, (327-338), Online publication date: 1-Jul-2017. Kishi T (2016) Heart Failure as a Disruption of Dynamic Circulatory Homeostasis Mediated by the Brain, International Heart Journal, 10.1536/ihj.15-517, 57:2, (145-149), . Jouett N, Moralez G, White D, Eubank W, Chen S, Tian J, Smith M, Zimmerman M and Raven P (2016) N -Acetylcysteine reduces hyperacute intermittent hypoxia-induced sympathoexcitation in human subjects , Experimental Physiology, 10.1113/EP085546, 101:3, (387-396), Online publication date: 1-Mar-2016. Chan S and Chan J (2014) Brain Stem NOS and ROS in Neural Mechanisms of Hypertension, Antioxidants & Redox Signaling, 10.1089/ars.2013.5230, 20:1, (146-163), Online publication date: 1-Jan-2014. Warowicka A, Kwasniewska A and Gozdzicka-Jozefiak A (2013) Alterations in mtDNA: A qualitative and quantitative study associated with cervical cancer development, Gynecologic Oncology, 10.1016/j.ygyno.2013.01.001, 129:1, (193-198), Online publication date: 1-Apr-2013. Kishi T (2013) Regulation of the sympathetic nervous system by nitric oxide and oxidative stress in the rostral ventrolateral medulla: 2012 Academic Conference Award from the Japanese Society of Hypertension, Hypertension Research, 10.1038/hr.2013.73, 36:10, (845-851), Online publication date: 1-Oct-2013. Mahalwar R and Khanna D (2013) Pleiotropic antioxidant potential of rosuvastatin in preventing cardiovascular disorders, European Journal of Pharmacology, 10.1016/j.ejphar.2013.04.025, 711:1-3, (57-62), Online publication date: 1-Jul-2013. Zheng H and Yu Y (2012) Chronic hydrogen-rich saline treatment attenuates vascular dysfunction in spontaneous hypertensive rats, Biochemical Pharmacology, 10.1016/j.bcp.2012.01.031, 83:9, (1269-1277), Online publication date: 1-May-2012. Cardoso L, Colombari E and Toney G (2012) Endogenous hydrogen peroxide in the hypothalamic paraventricular nucleus regulates sympathetic nerve activity responses to l -glutamate , Journal of Applied Physiology, 10.1152/japplphysiol.00912.2012, 113:9, (1423-1431), Online publication date: 1-Nov-2012. Yu Y and Zheng H (2012) Chronic hydrogen-rich saline treatment reduces oxidative stress and attenuates left ventricular hypertrophy in spontaneous hypertensive rats, Molecular and Cellular Biochemistry, 10.1007/s11010-012-1264-4, 365:1-2, (233-242), Online publication date: 1-Jun-2012. Pal C, Bindu S, Dey S, Alam A, Goyal M, Iqbal M, Sarkar S, Kumar R, Halder K, Debnath M, Adhikari S and Bandyopadhyay U (2012) Tryptamine-Gallic Acid Hybrid Prevents Non-steroidal Anti-inflammatory Drug-induced Gastropathy, Journal of Biological Chemistry, 10.1074/jbc.M111.307199, 287:5, (3495-3509), Online publication date: 1-Jan-2012. Lam M, Scruggs S, Kim T, Zong C, Lau E, Wang D, Ryan C, Faull K and Ping P (2012) An MRM-based workflow for quantifying cardiac mitochondrial protein phosphorylation in murine and human tissue, Journal of Proteomics, 10.1016/j.jprot.2012.02.014, 75:15, (4602-4609), Online publication date: 1-Aug-2012. Kishi T and Hirooka Y (2012) Central Mechanisms of Abnormal Sympathoexcitation in Chronic Heart Failure, Cardiology Research and Practice, 10.1155/2012/847172, 2012, (1-7), . Jin H, Hong K, Kim B, Kim J, Yoo Y, Oh B and Jeong S (2011) Replicated association between genetic variation in the PARK2 gene and blood pressure, Clinica Chimica Acta, 10.1016/j.cca.2011.05.026, 412:17-18, (1673-1677), Online publication date: 1-Aug-2011. Kim J, Minkler P, Salomon R, Anderson V and Hoppel C (2011) Cardiolipin: characterization of distinct oxidized molecular species, Journal of Lipid Research, 10.1194/jlr.M010520, 52:1, (125-135), Online publication date: 1-Jan-2011. Hirooka Y, Kishi T, Sakai K, Takeshita A and Sunagawa K (2011) Imbalance of central nitric oxide and reactive oxygen species in the regulation of sympathetic activity and neural mechanisms of hypertension, American Journal of Physiology-Regulatory, Integrative and Comparative Physiology, 10.1152/ajpregu.00426.2010, 300:4, (R818-R826), Online publication date: 1-Apr-2011. Jin H, Sõber S, Hong K, Org E, Kim B, Laan M, Oh B and Jeong S (2011) Age-Dependent Association of the Polymorphisms in the Mitochondria-Shaping Gene, OPA1, With Blood Pressure and Hypertension in Korean Population, American Journal of Hypertension, 10.1038/ajh.2011.131, 24:10, (1127-1135), Online publication date: 1-Oct-2011. Mariappan N, Elks C, Sriramula S, Guggilam A, Liu Z, Borkhsenious O and Francis J (2009) NF-κB-induced oxidative stress contributes to mitochondrial and cardiac dysfunction in type II diabetes, Cardiovascular Research, 10.1093/cvr/cvp305, 85:3, (473-483), Online publication date: 1-Feb-2010., Online publication date: 1-Feb-2010. Kumarasamy S, Gopalakrishnan K, Shafton A, Nixon J, Thangavel J, Farms P and Joe B (2010) Mitochondrial polymorphisms in rat genetic models of hypertension, Mammalian Genome, 10.1007/s00335-010-9259-5, 21:5-6, (299-306), Online publication date: 1-Jun-2010. Addabbo F, Montagnani M and Goligorsky M (2009) Mitochondria and Reactive Oxygen Species, Hypertension, 53:6, (885-892), Online publication date: 1-Jun-2009. Kim J, Kim H, Ko J, Bang H, Lee D and Vina J (2013) The Relationship between Leukocyte Mitochondrial DNA Copy Number and Telomere Length in Community-Dwelling Elderly Women, PLoS ONE, 10.1371/journal.pone.0067227, 8:6, (e67227) February 2009Vol 53, Issue 2 Advertisement Article InformationMetrics https://doi.org/10.1161/HYPERTENSIONAHA.108.125567PMID: 19114641 Originally publishedDecember 29, 2008 PDF download Advertisement SubjectsHypertension

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call