Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2391 2023-08-08 22:51:53 |
2 format change + 1 word(s) 2392 2023-08-09 03:03:47 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Koumallos, N.; Sigala, E.; Milas, T.; Baikoussis, N.G.; Aragiannis, D.; Sideris, S.; Tsioufis, K. Angiotensin Regulation of Vascular Homeostasis. Encyclopedia. Available online: https://encyclopedia.pub/entry/47807 (accessed on 27 July 2024).
Koumallos N, Sigala E, Milas T, Baikoussis NG, Aragiannis D, Sideris S, et al. Angiotensin Regulation of Vascular Homeostasis. Encyclopedia. Available at: https://encyclopedia.pub/entry/47807. Accessed July 27, 2024.
Koumallos, Nikolaos, Evangelia Sigala, Theodoros Milas, Nikolaos G. Baikoussis, Dimitrios Aragiannis, Skevos Sideris, Konstantinos Tsioufis. "Angiotensin Regulation of Vascular Homeostasis" Encyclopedia, https://encyclopedia.pub/entry/47807 (accessed July 27, 2024).
Koumallos, N., Sigala, E., Milas, T., Baikoussis, N.G., Aragiannis, D., Sideris, S., & Tsioufis, K. (2023, August 08). Angiotensin Regulation of Vascular Homeostasis. In Encyclopedia. https://encyclopedia.pub/entry/47807
Koumallos, Nikolaos, et al. "Angiotensin Regulation of Vascular Homeostasis." Encyclopedia. Web. 08 August, 2023.
Angiotensin Regulation of Vascular Homeostasis
Edit

Extensive research has been conducted to elucidate and substantiate the crucial role of the Renin-Angiotensin System (RAS) in the pathogenesis of hypertension, cardiovascular disorders, and renal diseases. Furthermore, the role of oxidative stress in maintaining vascular balance has been well established. It has been observed that many of the cellular effects induced by Angiotensin II (Ang II) are facilitated by reactive oxygen species (ROS) produced by nicotinamide adenine dinucleotide phosphate (NADPH) oxidase.

redox signaling oxidative stress RAS ACE inhibitors ARBs

1. Introduction

Reactive oxygen species (ROS) are considered highly reactive molecules that, depending on the amount produced, can affect cells and vascular functions [1]. Their precise regulation of production in the endothelium is crucial for controlling various cell functions in the vascular system [2][3][4]. Otherwise, excessive production of ROS (including superoxide O2_, hydroxyl radical _OH, and peroxynitrite anion ONOO_) can cause the disruption of redox homeostasis, damage the immune system response, and favour the manifestation of vascular diseases [5][6]. This situation can occur in various circumstances where there is an imbalance between ROS generation and antioxidant defence mechanisms, such as chronic inflammation, ischemia-reperfusion injury, aging, and metabolic disorders. Another important factor is impaired endothelial function (since endothelium is the main source of ROS), which disrupts the production of protective blood agents and contributes to cytotoxic effects, cell death, and, inevitably, oxidative stress [7][8]. Interestingly, it is worth noting that ROS have beneficial effects due to their cytotoxic properties and can serve as a defence mechanism against infections [9].
As for the overproduction of ROS in the development of cardiovascular disease, damage to the cardiovascular system caused by oxidative stress favours diseases such as hypertension, atherosclerosis, and heart failure [10][11]. The renin-angiotensin system (RAS) interacts with ROS and contributes to these diseases. Angiotensin II (Ang II), as a key component of the RAS, plays an important role in this interaction as it can stimulate ROS production through various mechanisms, such as activating nicotinamide adenine dinucleotide phosphate (NADPH) oxidase through its binding to the AT1 receptor (AT1R) [10]. This overproduction of ROS over Ang II can in turn affect the activity of components of the RAS, such as increase in the expression of the Angiotensin I to ACE and the production of renin.

2. Physiology of the RAS in Vascular Regulation

Understanding the physiology of RAS is vital for comprehending the interplay between ROS and RAS in vascular regulation. Increased levels of Ang II, the main effector of the RAS molecule, play a crucial role in cardiovascular disease progression when its production or signaling becomes dysregulated [10][11]. The physiological intricacies of the RAS have been thoroughly examined and comprehensively expounded upon in previous scientific investigations [10]. However, in the context of the present research on the influence of ROS in the progression of cardiovascular diseases, it is imperative to revisit and restate the fundamental principles of RAS physiology. By revisiting this well-established foundation, the researchers' objective is to establish a cohesive link between the RAS and the role of ROS in the development and manifestation of cardiovascular disorders.
The release of renin (an enzyme produced by conditions of reduced perfusion such as low blood pressure, low blood volume, or sympathetic stimulation), will activate a sequential processing of glycoprotein angiotensinogen (AGT), leading to the production of the decapeptide Ang I [12]. Subsequently, Ang I is further cleaved by ACE, a membrane-bound metalloprotease, and converts to Ang II. Once Ang II is formed it binds G-protein-coupled receptors (AT1-4R) [13][14][15]. AT1Rs are the prime mediators of Ang II leading to vasoconstriction, aldosterone and vasopressin secretion, and sodium and water retention. Because of this, AT1Rs play a crucial role in cardiovascular regulation, inflammation, fibrosis, endothelial dysfunction, and organ damage like nephrosclerosis [16][17].
On the other hand, the effects of AT2-4Rs are not fully understood, but they act differently than AT1Rs. It is believed that AT2Rs, when activated, counteract the short- and long-term effects of AT1Rs leading to beneficial antiproliferative and vasodilatory effects [18]. Other potential effects of AT2Rs include the regulation of cell growth and processing of the neuronal tissue [19]. AT3Rs were recently discovered and their exact role has not been elucidated. AT4Rs play a protective role in thrombotic effects as a fibrinolysis buffer by controlling the production of plasminogen activator inhibitor-1 (PAI-1). The nature of AT4Rs has been extensively investigated in the past with insulin-regulated aminopeptidase (IRAP) being the prominent candidate [20]. Angiotensin IV not only inhibits the peptidase activity of this enzyme but also facilitates IRAP translocation to the cell surface and enhances insulin-mediated glucose uptake [20]. In addition to the impacts mediated by G-protein-coupled receptors, recent discoveries of alternative enzyme systems and novel effector peptides have broadened our conventional understanding [21]. Among these systems are the receptors for prorenin and the G-protein-coupled receptor MAS [22]. The identification of these systems enhances our understanding of the intricate nature of the RAS by revealing that prorenin can activate renin and subsequently trigger protein kinases ERK1 and ERK2 [23][24]. The activation of these receptors introduces additional proliferative and metabolic effects that are not dependent on Ang II [25]. For blood pressure regulation, Ang II fast response to vascular changes is important. G protein-depended pathways enhance smooth muscle cell contraction [26] and trigger molecules that will contract smooth muscle cells. Moreover, when Ang II activates the G protein, it triggers the activation of phospholipase C, which leads to the production of inositol-1,4,5-trisphosphate (IP3) and diacylglycerol. These molecules then initiate the release of calcium ions (Ca2+) into the cytoplasm. The binding of Ca2+ to calmodulin activates myosin, a protein involved in muscle contraction, and enhances its interaction with actin. As a result, smooth muscle cells contract, contributing to the modulation of blood pressure.
The provided section of physiology elucidates the mechanism behind blood pressure regulation, which are mainly regulated by the rapid action of Ang II. By activating smooth muscle cell contraction through the G protein-dependent pathway, Ang II triggers a cascade of events involving phospholipase C, inositol-1,4,5-trisphosphate (IP3), diacylglycerol, and calcium ions (Ca2+). This series of molecular interactions ultimately leads to the contraction of smooth muscle cells. Understanding these intricate processes helps us grasp the physiological basis of blood pressure adjustments in response to changes in posture.

3. Pathological Implications of the RAS and the Involvement of ROS

As established thus far, the pathological manifestations of the RAS (abnormal cellular proliferation, inflammation, disrupted vascular balance) are closely related to the excessive presence and prolonged exposure of Ang II [27]. ROS, such as O2•_ (superoxide) and H2O2 (hydrogen peroxide), contribute to these pathological processes through Ang II signaling [28][29][30][31]. In the cardiovascular system, Ang II contributes significantly to hypertension through its central, vascular, or renal effects. The multimeric enzyme of NADPH oxidase-derived ROS production is a deleterious equation in the development of Ang II-induced hypertension. NAPDH oxidase induces Ang II-induced oxidative stress, due to increased enzymatic activity. This occurs because of the rapid translocation and phosphorylation of cytosolic subunits of the small GTPase rac1 and p47phox to the cytochrome complex, via protein kinase C (PKC) [32][33]. Thereafter, PKC activates Janus kinase (JAK), which transduces and activates the JAK/STAT signaling pathway. This sequence promotes Ang II multiplication processes. Other components in revealing Ang II adverse effects are the early growth responsive genes and redox-sensitive proteins (c-Src, epidermal growth factor EGFR) [34][35][36]. Specifically, EGFR activates the Ras/Raf/ERK cascade, which subsequently upregulates c-Fos. Together with c-Jun, activated by c-Src via JNK, c-Fos forms the transcription factor known as activator protein-1 (AP-q1) [33].
Elevated levels of Ang II can have additional consequences on cellular viability and potentially induce DNA damage. The increased presence of O2•_ and H2O2 activates additional redox-sensitive proteins, including p38/MAPK, which in turn stimulates the pro-survival factor Akt [37][38]. In a cascade of subsequent reactions, Akt inhibits various pro-apoptotic proteins. Another essential mediator of the RAS is aldosterone, which plays a crucial role in maintaining sodium and potassium balance, thus influencing extracellular volume. Aldosterone also exhibits potent pro-fibrotic effects [39]. The release of aldosterone is triggered by Ang II, and emerging evidence suggests that aldosterone may contribute to and exacerbate the detrimental effects of Ang II. Through activation of mineralocorticoid receptors, aldosterone promotes endothelial dysfunction and thrombosis, reduces vascular compliance and baroreceptor function, and induces fibrosis in both myocardial and vascular tissues [40]. The resulting increase in blood pressure and circulating volume, caused by the effects of Ang II and aldosterone on their target organs, establishes a negative feedback loop that suppresses renin release. The maintenance of this feedback inhibition critically relies on the Ang II-mediated activation of the AT1R [41].
Understanding the pathological effects of the RAS and the involvement of ROS is crucial for several reasons. It provides insights into the molecular mechanisms underlying cardiovascular diseases associated with RAS dysregulation, aiding in the advancement of precision therapeutic interventions. The identification of specific proteins and pathways involved opens opportunities for drug development, enabling the design of medications that selectively target these molecules or their associated receptors. This knowledge also contributes to personalized medicine approaches, considering individual variability in RAS responses. Moreover, it drives research advancements by uncovering novel signaling pathways and potential biomarkers, enhancing our understanding of cardiovascular disease pathophysiology. Overall, comprehending the pathological effects of RAS and ROS has significant implications for improving patient outcomes and advancing cardiovascular research.

4. The Application of ACEIs and ARBs in the Treatment of Cardiovascular Disorders

For a considerable period, the ACE/Ang II/AT1R pathway was recognized as the primary mechanism through which the RAS influences cardiovascular processes. Various categories of antihypertensive medications, including ACEIs, ARBs, β blockers (BBs), direct renin inhibitors (DRIs), and mineralocorticoid receptor antagonists (MRAs), have been employed to provide cardiorenal protection. ACEIs and ARBs, along with BBs, DRIs, and MRAs, act by interfering with the signaling pathways within the RAS. They are considered the first-line treatment options for managing hypertension and other cardiovascular disorders such as heart failure [42][43][44]. Furthermore, ACEIs and ARBs have a significant impact on the cardiovascular system and offer a protective effect against the occurrence and progression of kidney disease [45].
In addition to the well-established protection that ACEIs offer to the cardiovascular and renal systems by effectively controlling arterial pressure thresholds, they have also been associated with additional beneficial effects on the endothelium [46]. This has been confirmed in research studies conducted on patients with coronary artery disease or hypercholesterolemia, regardless of the effect on blood pressure reduction. These studies demonstrate that ACEIs and ARBs can additionally improve endothelial and vascular function in these patients. Furthermore, large studies have shown that ACEIs and ARBs are effective as monotherapy in managing other conditions such as left ventricular hypertrophy, systolic dysfunction, heart failure, and myocardial infarction [47]. Additionally, trials like the HOPE study, ONTARGET, and TRANSCEND have reported the extended benefits of ACEIs and ARBs to patients with an increased susceptibility to adverse outcomes but without left ventricular dysfunction [48]. For instance, the HOPE trial showed that ramipril significantly reduced the incidence of death, myocardial infarction, and stroke in high-risk patients [49]. Similarly, the EUROPA trial revealed a relative risk reduction in cardiovascular events with perindopril treatment in patients with stable coronary heart disease. However, the PEACE trial, which focused on stable coronary artery disease patients with preserved ventricular function, did not find a therapeutic benefit of ACEIs when added to conventional therapy, potentially because of the minimal occurrence of significant outcome events in that patient group [47][50].

5. Enhancing Endothelial Function and Mitigating Oxidative Stress: Exploring the Effects of ACEIs and ARBs

Several studies have demonstrated the favourable effects of ACEIs and ARBs on endothelial function and their ability to control oxidative stress levels in populations with cardiovascular diseases. As it has already been cleared, Ang II excessive presence is thought to play a pivotal role in the increased generation of O2•_ and the impairment of endothelial function in blood vessels. This effect is achieved through the activation of NADPH oxidases, triggered by the stimulation of AT1 receptors [51]. Additionally, studies conducted in laboratory settings have revealed that AT1 receptors can be upregulated by low-density lipoproteins [52][53]. By augmenting the activity of SOD3 and blocking NADPH oxidase activation, ACEIs and ARBs provide an indirect mitigation of oxidative stress [46][54].
In patients with coronary artery disease, it has been observed that ramipril (ACEI) and losartan (ARB) improve endothelial function by increasing the availability of nitric oxide (NO) through the mitigation of oxidative stress within the arterial wall [55]. Additionally, ARBs offer vascular protective effects [56]. Losartan promotes the phosphorylation of endothelial NO synthase (eNOS) and suppresses endothelial apoptosis induced by Tissue Necrosis Factor-α (TNF-α) through the activation of the VEGFR2/PI3K/Akt pathway [57]. Additionally, in diabetic rats, losartan restored glomerular NO production by increasing GCH1 protein expression and tetrahydrobiopterin (BH4) levels [58]. Furthermore, valsartan and irbesartan (ARBs) exert effects that counteract the development of atherosclerosis. The promotion of eNOS Ser117-phosphorylation increases eNOS mRNA stability and this leads to the reduced NADPH oxidase expression and also to the augmented vascular BH4 and restored eNOS uncoupling [59][60].
Some of the positive effects of ACEIs may be attributed to their impact on an ACE signaling cascade, resulting in improved endothelial function that appears to be independent of their effects on vasoactive substances [61]. Another mechanism suggested to beneficially affect the endothelium is the introduction of ACEIs as a treatment, with potential ACE signaling cascade involvement. This effect seems unrelated to the impact of ACEIs on vasoactive substances [62]. Moreover, when ACEI binds to the cell surface ectoenzyme ACE, it triggers a cascade that ultimately increases ACE and cyclooxygenase-2 expression. Firstly, ACEI binding activates casein kinase 2, leading to serine residue phosphorylation at the molecule’s C-terminal end [62]. Secondly, ACE-associated c-Jun N-terminal kinase is activated, possibly through mitogen-activated protein kinase 7 activation [63]. This cascade eventually leads to an increase in the expression of ACE and cyclooxygenase-2 (through the accumulation of phosphorylated c-Jun in the nucleus, enhancing the DNA-binding activity of activator protein-1) [64]. Additionally, the elevated expression of cyclooxygenase-2 benefits endothelial function by promoting the production of prostacyclin, a vasodilator and antiplatelet agent, by endothelial cells [65].
To conclude this section, it is evident that ACEIs and ARBs have shown potential in modulating ADMA metabolism and improving endothelial function. These drugs may reduce ADMA levels through various mechanisms, including activation of DDAH and attenuation of oxidative stress. However, further research is required to fully elucidate the precise mechanisms involved. Understanding the impact of ACEIs and ARBs on ADMA metabolism contributes to our knowledge of their potential benefits in cardiovascular health and warrants continued investigation in this field.

References

  1. Antoniades, C.; Tousoulis, D.; Tentolouris, C.; Toutouzas, P.; Stefanadis, C. Oxidative stress, antioxidant vitamins, and atherosclerosis. From basic research to clinical practice. Herz 2003, 28, 628–638.
  2. Fujiwara, N.; Yamashita, S.; Okamoto, M.; Cooley, M.A.; Ozaki, K.; Everett, E.T.; Suzuki, M. Perfluorooctanoic acid-induced cell death via the dual roles of ROS-MAPK/ERK signaling in ameloblast-lineage cells. Ecotoxicol. Environ. Saf. 2023, 260, 115089.
  3. Lee, Y.-P.; Lin, C.-R.; Chen, S.-S.; Chen, R.-J.; Wu, Y.-H.; Chen, Y.-H.; Huang, B.-M. Combination treatment of cordycepin and radiation induces MA-10 mouse Leydig tumor cell death via ROS accumulation and DNA damage. Am. J. Cancer Res. 2023, 13, 1329–1346.
  4. Wang, B.; Wang, Y.; Zhang, J.; Hu, C.; Jiang, J.; Li, Y.; Peng, Z. ROS-induced lipid peroxidation modulates cell death outcome: Mechanisms behind apoptosis, autophagy, and ferroptosis. Arch. Toxicol. 2023, 97, 1439–1451.
  5. Cheng, C.K.; Ding, H.; Jiang, M.; Yin, H.; Gollasch, M.; Huang, Y. Perivascular adipose tissue: Fine-tuner of vascular redox status and inflammation. Redox Biol. 2023, 62, 102683.
  6. Hernández-Reséndiz, S.; Muñoz-Vega, M.; Contreras, W.E.; Crespo-Avilan, G.E.; Rodriguez-Montesinos, J.; Arias-Carrión, O.; Pérez-Méndez, O.; Boisvert, W.A.; Preissner, K.T.; Cabrera-Fuentes, H.A. Responses of Endothelial Cells Towards Ischemic Conditioning Following Acute Myocardial Infarction. Cond. Med. 2018, 1, 247–258.
  7. Da Silva, F.C.; de Araújo, B.J.; Cordeiro, C.S.; Arruda, V.M.; Faria, B.Q.; Guerra, J.F.D.C.; De Araújo, T.G.; Fürstenau, C.R. Endothelial dysfunction due to the inhibition of the synthesis of nitric oxide: Proposal and characterization of an in vitro cellular model. Front. Physiol. 2022, 13, 978378.
  8. Shaito, A.; Aramouni, K.; Assaf, R.; Parenti, A.; Orekhov, A.; Yazbi, A.E.; Pintus, G.; Eid, A.H. Oxidative Stress-Induced Endothelial Dysfunction in Cardiovascular Diseases. Front. Biosci. 2022, 27, 105.
  9. Hänze, J.; Weissmann, N.; Grimminger, F.; Seeger, W.; Rose, F. Cellular and molecular mechanisms of hypoxia-inducible factor driven vascular remodeling. Thromb. Haemost. 2007, 97, 774–787.
  10. Paul, M.; Poyan Mehr, A.; Kreutz, R. Physiology of local renin-angiotensin systems. Physiol. Rev. 2006, 86, 747–803.
  11. Kasiakogias, A.; Rosei, E.A.; Camafort, M.; Ehret, G.; Faconti, L.; Ferreira, J.P.; Brguljan, J.; Januszewicz, A.; Kahan, T.; Manolis, A.; et al. Hypertension and heart failure with preserved ejection fraction: Position paper by the European Society of Hypertension. J. Hypertens. 2021, 39, 1522–1545.
  12. McFarlane, S.I.; Kumar, A.; Sowers, J.R. Mechanisms by which angiotensin-converting enzyme inhibitors prevent diabetes and cardiovascular disease. Am. J. Cardiol. 2003, 91, 30H–37H.
  13. Lévy, B.I. Can angiotensin II type 2 receptors have deleterious effects in cardiovascular disease? Implications for therapeutic blockade of the renin-angiotensin system. Circulation 2004, 109, 8–13.
  14. Zhang, H.; Luginina, A.; Mishin, A.; Baidya, M.; Shukla, A.K.; Cherezov, V. Structural insights into ligand recognition and activation of angiotensin receptors. Trends Pharmacol. Sci. 2021, 42, 577–587.
  15. Ziaja, M.; Urbanek, K.A.; Kowalska, K.; Piastowska-Ciesielska, A.W. Angiotensin II and Angiotensin Receptors 1 and 2—Multifunctional System in Cells Biology, What Do We Know? Cells 2021, 10, 381.
  16. Duprez, D.A. Role of the renin–angiotensin–aldosterone system in vascular remodeling and inflammation: A clinical review. J. Hypertens. 2006, 24, 983–991.
  17. Ferrario, C.M.; Strawn, W.B. Role of the renin-angiotensin-aldosterone system and proinflammatory mediators in cardiovascular disease. Am. J. Cardiol. 2006, 98, 121–128.
  18. Stoll, M.; Steckelings, U.M.; Paul, M.; Bottari, S.P.; Metzger, R.; Unger, T. The angiotensin AT2-receptor mediates inhibition of cell proliferation in coronary endothelial cells. J. Clin. Investig. 1995, 95, 651–657.
  19. Steckelings, U.M.; Kaschina, E.; Unger, T. The AT2 receptor—A matter of love and hate. Peptides 2005, 26, 1401–1409.
  20. Albiston, A.L.; McDowall, S.G.; Matsacos, D.; Sim, P.; Clune, E.; Mustafa, T.; Lee, J.; Mendelsohn, F.A.O.; Simpson, R.J.; Connolly, L.M.; et al. Evidence that the angiotensin IV (AT4) receptor is the enzyme insulin-regulated aminopeptidase. J. Biol. Chem. 2001, 276, 48623–48626.
  21. Silva, I.M.S.; Assersen, K.B.; Willadsen, N.N.; Jepsen, J.; Artuc, M.; Steckelings, U.M. The role of the renin-angiotensin system in skin physiology and pathophysiology. Exp. Dermatol. 2020, 29, 891–901.
  22. Nguyen, G.; Danser, A.H.J. Prorenin and (pro)renin receptor: A review of available data from in vitro studies and experimental models in rodents. Exp. Physiol. 2008, 93, 557–563.
  23. Santos, R.A.S.; Simoes e Silva, A.C.; Maric, C.; Silva, D.M.R.; Machado, R.P.; de Buhr, I.; Heringer-Walther, S.; Pinheiro, S.V.B.; Lopes, M.T.; Bader, M.; et al. Angiotensin-(1–7) is an endogenous ligand for the G protein-coupled receptor Mas. Proc. Natl. Acad. Sci. USA 2003, 100, 8258–8263.
  24. Pimenta, E.; Oparil, S. Role of aliskiren in cardio-renal protection and use in hypertensives with multiple risk factors. Ther. Clin. Risk Manag. 2009, 5, 459–464.
  25. Zhang, L.; Ju, Q.; Sun, J.; Huang, L.; Wu, S.; Wang, S.; Li, Y.; Guan, Z.; Zhu, Q.; Xu, Y. Discovery of Novel Dual Extracellular Regulated Protein Kinases (ERK) and Phosphoinositide 3-Kinase (PI3K) Inhibitors as a Promising Strategy for Cancer Therapy. Molecules 2020, 25, 5693.
  26. Nakamura, F.; Tsukamoto, I.; Inoue, S.; Hashimoto, K.; Akagi, M. Cyclic compressive loading activates angiotensin II type 1 receptor in articular chondrocytes and stimulates hypertrophic differentiation through a G-protein-dependent pathway. FEBS Open Bio 2018, 8, 962–973.
  27. Touyz, R.M. Reactive oxygen species and angiotensin II signaling in vascular cells: Implications in cardiovascular disease. Braz. J. Med. Biol. Res. 2004, 37, 1263–1273.
  28. Branicky, R.; Wang, Y.; Khaki, A.; Liu, J.-L.; Kramer-Drauberg, M.; Hekimi, S. Stimulation of RAS-dependent ROS signaling extends longevity by modulating a developmental program of global gene expression. Sci. Adv. 2022, 8, eadc9851.
  29. Mohamed, H.R.H. Acute Oral Administration of Cerium Oxide Nanoparticles Suppresses Lead Acetate–Induced Genotoxicity, Inflammation, and ROS Generation in Mice Renal and Cardiac Tissues. Biol. Trace Elem. Res. 2022, 200, 3284–3293.
  30. Barp, C.G.; Bonaventura, D.; Assreuy, J. NO, ROS, RAS, and PVAT: More Than a Soup of Letters. Front. Physiol. 2021, 12, 640021.
  31. Griendling, K.K.; Minieri, C.A.; Ollerenshaw, J.D.; Alexander, R.W. Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circ. Res. 1994, 74, 1141–1148.
  32. Wassmann, S.; Laufs, U.; Bäumer, A.T.; Müller, K.; Konkol, C.; Sauer, H.; Böhm, M.; Nickenig, G. Inhibition of geranylgeranylation reduces angiotensin II-mediated free radical production in vascular smooth muscle cells: Involvement of angiotensin AT1 receptor expression and Rac1 GTPase. Mol. Pharmacol. 2001, 59, 646–654.
  33. Mehta, P.K.; Griendling, K.K. Angiotensin II cell signaling: Physiological and pathological effects in the cardiovascular system. Am. J. Physiol. Cell Physiol. 2007, 292, C82–C97.
  34. Berk, B.C.; Corson, M.A. Angiotensin II signal transduction in vascular smooth muscle: Role of tyrosine kinases. Circ. Res. 1997, 80, 607–616.
  35. Ren, Y.; Xie, W.; Yang, S.; Jiang, Y.; Wu, D.; Zhang, H.; Sheng, S. Angiotensin-converting enzyme 2 inhibits inflammation and apoptosis in high glucose-stimulated microvascular endothelial cell damage by regulating the JAK2/STAT3 signaling pathway. Bioengineered 2022, 13, 10802–10810.
  36. Jung, J.-H.; Choi, J.-W.; Lee, M.-K.; Choi, Y.-H.; Nam, T.-J. Effect of Cyclophilin from Pyropia Yezoensis on the Proliferation of Intestinal Epithelial Cells by Epidermal Growth Factor Receptor/Ras Signaling Pathway. Mar. Drugs 2019, 17, 297.
  37. Kim, D.-H.; Chung, J.-K. Akt: Versatile mediator of cell survival and beyond. Korean Soc. Biochem. Mol. Biol.-BMB Rep. 2002, 35, 106–115.
  38. Azouz, A.A.; Abdel-Rahman, D.M.; Messiha, B.A.S. Balancing renal Ang-II/Ang-(1-7) by xanthenone; an ACE2 activator; contributes to the attenuation of Ang-II/p38 MAPK/NF-κB p65 and Bax/caspase-3 pathways in amphotericin B-induced nephrotoxicity in rats. Toxicol. Mech. Methods 2023, 33, 452–462.
  39. Johar, S.; Cave, A.C.; Narayanapanicker, A.; Grieve, D.J.; Shah, A.M. Aldosterone mediates angiotensin II-induced interstitial cardiac fibrosis via a Nox2-containing NADPH oxidase. FASEB J. 2006, 20, 1546–1548.
  40. Struthers, A.D.; MacDonald, T.M. Review of aldosterone- and angiotensin II-induced target organ damage and prevention. Cardiovasc. Res. 2004, 61, 663–670.
  41. Dehe, L.; Mousa, S.A.; Aboryag, N.; Shaqura, M.; Beyer, A.; Schäfer, M.; Treskatsch, S. Identification of Mineralocorticoid Receptors, Aldosterone, and Its Processing Enzyme CYP11B2 on Parasympathetic and Sympathetic Neurons in Rat Intracardiac Ganglia. Front. Neuroanat. 2021, 15, 802359.
  42. Jia, Q.; Han, W.; Shi, S.; Hu, Y. The effects of ACEI/ARB, aldosterone receptor antagonists and statins on preventing recurrence of atrial fibrillation: A protocol for systematic review and network meta-analysis. Medicine 2021, 100, e24280.
  43. Eberhardt, R.T.; Kevak, R.M.; Kang, P.M.; Frishman, W.H. Angiotensin II receptor blockade: An innovative approach to cardiovascular pharmacotherapy. J. Clin. Pharmacol. 1993, 33, 1023–1038.
  44. Papadopoulos, D.P.; Votteas, V. Role of perindopril in the prevention of stroke. Recent Pat. Cardiovasc. Drug Discov. 2006, 1, 283–289.
  45. Mogensen, U.M.; Gong, J.; Jhund, P.S.; Shen, L.; Køber, L.; Desai, A.S.; Lefkowitz, M.P.; Packer, M.; Rouleau, J.L.; Solomon, S.D.; et al. Effect of sacubitril/valsartan on recurrent events in the Prospective comparison of ARNI with ACEI to Determine Impact on Global Mortality and morbidity in Heart Failure trial (PARADIGM-HF). Eur. J. Heart Fail. 2018, 20, 760–768.
  46. Hornig, B.; Landmesser, U.; Kohler, C.; Ahlersmann, D.; Spiekermann, S.; Christoph, A.; Tatge, H.; Drexler, H. Comparative effect of ace inhibition and angiotensin II type 1 receptor antagonism on bioavailability of nitric oxide in patients with coronary artery disease: Role of superoxide dismutase. Circulation 2001, 103, 799–805.
  47. Yusuf, S.; Sleight, P.; Pogue, J.; Bosch, J.; Davies, R.; Dagenais, G.; Heart Outcomes Prevention Evaluation Study Investigators. Effects of an angiotensin-converting-enzyme inhibitor, ramipril, on cardiovascular events in high-risk patients. N. Engl. J. Med. 2000, 342, 145–153.
  48. Schmieder, R.E.; Martus, P.; Klingbeil, A. Reversal of left ventricular hypertrophy in essential hypertension. A meta-analysis of randomized double-blind studies. JAMA 1996, 275, 1507–1513.
  49. Fox, K.M. The EURopean Trial on Reduction of Cardiac Events with Perindopril in Stable Coronary Artery Disease Investigators. Efficacy of perindopril in reduction of cardiovascular events among patients with stable coronary artery disease: Randomised, double-blind, placebo-controlled, multicentre trial (the EUROPA study). Lancet 2003, 362, 782–788.
  50. Braunwald, E.; Domanski, M.J.; Fowler, S.E.; Geller, N.L.; Gersh, B.J.; Hsia, J.; Pfeffer, M.A.; Rice, M.M.; Rosenberg, Y.D.; Rouleau, J.L.; et al. Angiotensin-converting-enzyme inhibition in stable coronary artery disease. N. Engl. J. Med. 2004, 351, 2058–2068.
  51. Griendling, K.K.; Sorescu, D.; Ushio-Fukai, M. NAD(P)H oxidase: Role in cardiovascular biology and disease. Circ. Res. 2000, 86, 494–501.
  52. Warnholtz, A.; Nickenig, G.; Schulz, E.; Macharzina, R.; Bräsen, J.H.; Skatchkov, M.; Heitzer, T.; Stasch, J.P.; Griendling, K.K.; Harrison, D.G.; et al. Increased NADH-oxidase-mediated superoxide production in the early stages of atherosclerosis: Evidence for involvement of the renin-angiotensin system. Circulation 1999, 99, 2027–2033.
  53. Cernecka, H.; Doka, G.; Srankova, J.; Pivackova, L.; Malikova, E.; Galkova, K.; Kyselovic, J.; Krenek, P.; Klimas, J. Ramipril restores PPARβ/δ and PPARγ expressions and reduces cardiac NADPH oxidase but fails to restore cardiac function and accompanied myosin heavy chain ratio shift in severe anthracycline-induced cardiomyopathy in rat. Eur. J. Pharmacol. 2016, 791, 244–253.
  54. Petersen, S.V.; Poulsen, N.B.; Linneberg Matthiesen, C.; Vilhardt, F. Novel and Converging Ways of NOX2 and SOD3 in Trafficking and Redox Signaling in Macrophages. Antioxidants 2021, 10, 172.
  55. Chao, Y.; Ye, P.; Zhu, L.; Kong, X.; Qu, X.; Zhang, J.; Luo, J.; Yang, H.; Chen, S. Low shear stress induces endothelial reactive oxygen species via the AT1R/eNOS/NO pathway. J. Cell. Physiol. 2018, 233, 1384–1395.
  56. Watanabe, T.; Suzuki, J.; Yamawaki, H.; Sharma, V.K.; Sheu, S.-S.; Berk, B.C. Losartan metabolite EXP3179 activates Akt and endothelial nitric oxide synthase via vascular endothelial growth factor receptor-2 in endothelial cells: Angiotensin II type 1 receptor-independent effects of EXP3179. Circulation 2005, 112, 1798–1805.
  57. Goh, K.L.; Bhaskaran, K.; Minassian, C.; Evans, S.J.; Smeeth, L.; Douglas, I.J. Angiotensin receptor blockers and risk of dementia: Cohort study in UK Clinical Practice Research Datalink. Br. J. Clin. Pharmacol. 2015, 79, 337–350.
  58. Satoh, M.; Fujimoto, S.; Arakawa, S.; Yada, T.; Namikoshi, T.; Haruna, Y.; Horike, H.; Sasaki, T.; Kashihara, N. Angiotensin II type 1 receptor blocker ameliorates uncoupled endothelial nitric oxide synthase in rats with experimental diabetic nephropathy. Nephrol. Dial. Transplant. 2008, 23, 3806–3813.
  59. Imanishi, T.; Tsujioka, H.; Ikejima, H.; Kuroi, A.; Takarada, S.; Kitabata, H.; Tanimoto, T.; Muragaki, Y.; Mochizuki, S.; Goto, M.; et al. Renin inhibitor aliskiren improves impaired nitric oxide bioavailability and protects against atherosclerotic changes. Hypertension 2008, 52, 563–572.
  60. Nussberger, J.; Aubert, J.-F.; Bouzourene, K.; Pellegrin, M.; Hayoz, D.; Mazzolai, L. Renin inhibition by aliskiren prevents atherosclerosis progression: Comparison with irbesartan, atenolol, and amlodipine. Hypertension 2008, 51, 1306–1311.
  61. Fleming, I. Signaling by the angiotensin-converting enzyme. Circ. Res. 2006, 98, 887–896.
  62. Horne, M.C.; Hell, J.W. Angiotensin II signalling kicks out p27Kip1: Casein kinase 2 augmentation of Cav1.2 L-type Ca2+ channel activity in immature ventricular cardiomyocytes. J. Physiol. 2017, 595, 4131–4132.
  63. Kashihara, T.; Nakada, T.; Kojima, K.; Takeshita, T.; Yamada, M. Angiotensin II activates CaV1.2 Ca2+ channels through β-arrestin2 and casein kinase 2 in mouse immature cardiomyocytes. J. Physiol. 2017, 595, 4207–4225.
  64. Fleming, I.; Kohlstedt, K.; Busse, R. New fACEs to the renin-angiotensin system. Physiology 2005, 20, 91–95.
  65. Li, X.; Hu, H.; Wang, Y.; Xue, M.; Li, X.; Cheng, W.; Xuan, Y.; Yin, J.; Yang, N.; Yan, S. Valsartan Upregulates Kir2.1 in Rats Suffering from Myocardial Infarction via Casein Kinase 2. Cardiovasc. Drugs Ther. 2015, 29, 209–218.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , ,
View Times: 184
Revisions: 2 times (View History)
Update Date: 09 Aug 2023
1000/1000
Video Production Service