Curr Hypertens Rep (2014) 16:431 DOI 10.1007/s11906-014-0431-2

MEDIATORS, MECHANISMS, AND PATHWAYS IN TISSUE INJURY (T FUJITA, SECTION EDITOR)

Angiotensin II and Vascular Injury Augusto C. Montezano & Aurelie Nguyen Dinh Cat & Francisco J. Rios & Rhian M. Touyz

# Springer Science+Business Media New York 2014

Abstract Vascular injury, characterized by endothelial dysfunction, structural remodelling, inflammation and fibrosis, plays an important role in cardiovascular diseases. Cellular processes underlying this include altered vascular smooth muscle cell (VSMC) growth/apoptosis, fibrosis, increased contractility and vascular calcification. Associated with these events is VSMC differentiation and phenotypic switching from a contractile to a proliferative/secretory phenotype. Inflammation, associated with macrophage infiltration and increased expression of redox-sensitive pro-inflammatory genes, also contributes to vascular remodelling. Among the many factors involved in vascular injury is Ang II. Ang II, previously thought to be the sole biologically active downstream peptide of the renin-angiotensin system (RAS), is converted to smaller peptides, [Ang III, Ang IV, Ang-(1-7)], that are functional and that modulate vascular tone and structure. The actions of Ang II are mediated via signalling pathways activated upon binding to AT1R and AT2R. AT1R activation induces effects through PLC-IP3-DAG, MAP kinases, tyrosine kinases, tyrosine phosphatases and RhoA/Rho kinase. Ang II elicits many of its (patho)physiological actions by stimulating reactive oxygen species (ROS) generation through activation of vascular NAD(P)H oxidase (Nox). ROS in turn influence redox-sensitive signalling molecules. Here we discuss the role of Ang II in vascular injury, focusing This article is part of the Topical Collection on Mediators, Mechanisms, and Pathways in Tissue Injury A. C. Montezano : A. Nguyen Dinh Cat : F. J. Rios : R. M. Touyz Institute of Cardiovascular and Medical Sciences, University of Glasgow, Glasgow, UK R. M. Touyz (*) Institute of Cardiovascular and Medical Sciences, BHF Glasgow Cardiovascular Research Centre, University of Glasgow, 126 University Place, Glasgow G12 8TA, UK e-mail: [email protected]

on molecular mechanisms and cellular processes. Implications in vascular remodelling, inflammation, calcification and atherosclerosis are highlighted. Keywords Angiotensin receptors . Inflammation . Signal transduction . Oxidative stress . Vascular remodelling . Calcification . Atherosclerosis

Introduction The renin angiotensin system (RAS) plays a major physiological role in regulating vascular function and a pathological role in vascular injury through its actions on endothelial dysfunction, vascular remodelling and vascular inflammation [1, 2]. As such, the RAS contributes to vascular damage associated with hypertension and atherosclerosis, which are important risk factors underlying cardiovascular, cerebrovascular, renal and metabolic diseases. Drugs that inhibit the RAS promote vascular health and reduce the risks of cardiovascular events [3]. Initially the RAS was identified as an endocrine system where circulating kidney-derived renin regulates cardiovascular function through Ang II binding to its AT1 and AT2 receptors on target tissues [4]. Until recently, this system was accepted as the conventional RAS. However, it is now evident that the kidney is not the only source of renin production and that angiotensin (Ang) peptides may be formed in extra-renal tissues and organs, such as the brain, adrenal gland, pituitary gland, reproductive tissues, gastrointestinal tract, haematopoietic tissue, heart and vessels [5, 6]. Functions of this local RAS remain unclear, but may be related to finetuning of Ang II at the tissue level. Vascular injury is associated with impaired endothelial cell function and alterations in vascular smooth muscle cell (VSMC) contraction/relaxation, migration, growth/apoptosis, senescence, calcification, production/degradation of

431, Page 2 of 11

extracellular matrix and inflammation [7–9]. Growing evidence indicates that activation of the immune system and adipokines derived from perivascular fat may also be important [10, 11]. Cellular mechanisms and signalling pathways implicated in vascular injury are complex and multifactorial, but among the numerous systems shown to be important is the RAS [12–14]. Here, we highlight molecular mechanisms whereby Ang II influences vascular cell function, and discuss the role of Ang II in vascular remodelling, inflammation, calcification and atherosclerosis.

The RAS—Old and New Concepts The RAS is considered an endocrine system where Ang II is the product of sequential enzymatic cleavage of angiotensinogen. Circulating renin converts hepaticderived angiotensinogen to Ang I, which is then cleaved by angiotensin converting enzyme (ACE) to form Ang II. Ang II mediates its physiological actions through two receptors, Ang II type I receptor (AT1R) and Ang II type 2 receptor (AT2R). Until recently, Ang II was considered the major downstream peptide of the RAS. However, over the past decade, the conventional view of the RAS has undergone major change. In particular, new RAS members (renin/prorenin receptor, ACE2 and Ang II-derived peptides) have been identified (Fig. 1), and a functionally active tissue-based RAS has been characterized [15, 16]. New Members of the RAS Renin, produced by the juxtaglomerular apparatus in the kidney, acts on

Fig. 1 New components of the renin-angiotensin system (RAS) and Ang receptors. In addition to the classical RAS, where Ang II is derived from angiotensinogen by renin and ACE, Ang II can be produced through nonACE dependent systems. Ang II itself can be metabolized to biologically active small peptides, such as Ang-(1-7), Ang-(1-9), Ang III and Ang IV. PEP, prolyl-endopeptidase; AMP, aminopeptidase; PCP, prolyl-carboxypeptidase; NEP, neutral-endopeptidase; (P)RR, prorenin receptor

Curr Hypertens Rep (2014) 16:431

angiotensinogen to form Ang I. It is the key enzyme of the RAS because of its rate-limiting hydrolytic activity [17]. Renin also acts as an agonist of the RAS by binding to the renin/ prorenin receptor [(P)RR], a new component of the RAS [18, 19]. Prorenin bound to (P)RR undergoes a conformational change and activates signalling pathways, e.g., ERK1/2 and p38MAP kinase, which promote cell growth and fibrosis independently of Ang II in cardiomyocytes, mesangial cells, podocytes, distal tubular cells, vascular endothelial cells and VSMCs [20, 21]. The exact (patho)physiological significance of this receptor remains unclear, but it may play a role in early development. Transgenic animals overexpressing (P)RR develop high BP or glomerulosclerosis, and increased expression of (P)RR has been shown in models of hypertension and kidney damage. However, definitive proof of a role for this receptor in human cardiovascular disease is still lacking [22]. Angiotensin Converting Enzyme (ACE) 2 ACE2 catalyses Ang I and Ang II to generate the Ang peptides Ang-(1-9) and Ang-(1-7), which mediate, in general, effects that are opposite to those of Ang II [23, 24]. Reduced ACE2, and consequent decreased Ang-(1-7) production, is associated with vascular injury in hypertension, diabetes, atherosclerosis and kidney disease [25, 26]. On the other hand, increased ACE2 activity promotes vasodilation, and as such is currently being tested as a potential therapeutic target [27]. Ang II-Derived Small Peptides Ang-(1-7) is formed from Ang II by prolyl endopeptidase, prolyl carboxipeptidase or ACE2, or directly from Ang I through hydrolysis by prolyl endopeptidase and endopeptidase 24.11 [28] (Fig. 1). Ang-(1-7) is present in the circulation and in tissues [29, 30], and it binds to its G protein-coupled receptor Mas [31]. Ang-(1-7) opposes Ang II effects by mediating vasodilation, growth-inhibition, anti-inflammatory responses, anti-arrhythmogenic and antithrombotic effects [32, 33]. These processes are mediated through nitric oxide synthase (NOS)-derived nitric oxide (NO) production, activation of protein tyrosine phosphatases, inhibition of MAP kinases, and inhibition of NADPH oxidase-derived generation of reactive oxygen species (ROS) [34, 35]. Ang-(1-7)-Mas can interact with AT1R, thereby inhibiting Ang II actions. The ACE2-Ang-(1-7)-Mas axis is now considered as a counter-regulatory system to the ACEAng II-AT1R axis [34]. Other Ang II-derived peptides include Ang III, which increases blood pressure, promotes vasoconstriction and stimulates aldosterone production [36]; Ang IV, which increases renal blood flow and improves cardiac function [37]; Ang-(3-7), implicated in the brain and kidney [38]; Ang-(1-9), which enhances bradykinin actions, NO production, arachidonic acid release and regulation of platelet function [39]; and Ang-(1-12), which has been implicated in cardiac function.

Curr Hypertens Rep (2014) 16:431

Tissue-Based RAS A tissue system is characterized by the presence of all RAS components, including angiotensinogen, renin, ACE, Ang I, Ang II, and Ang II receptors. Such a local system is present in the heart, vessels, kidney, adrenal gland, pancreas, central nervous system, reproductive system, lymphatic and adipose tissue [40]. Components of the RAS have also been identified in the eye, which may be important physiologically in maintaining ocular pressure [41], and pathologically in retinopathies associated with hypertension and diabetes [42]. Increased activation of tissue RAS has been demonstrated in cardiovascular diseases such as atherosclerosis, myocardial infarction, cardiac failure, diabetes and kidney disease [43], independently of blood pressure (BP) elevation.

Ang II in the Vascular System—Molecular Mechanisms Angiotensin Receptors and Signalling Ang II mediates effects via complex intracellular signalling pathways that are stimulated following binding of the peptide to its cell-surface receptors [44, 45] AT1R and AT2R [46, 47]. AT1R and AT2R are members of the seven transmembrane-domain, G proteincoupled receptor (GPCR) superfamily, and are distinguished pharmacologically by their affinities for antagonists. Ang II binding to the AT1R results in coupling of G proteins (Gq/G11 and/or Gi/Go) to the C-terminal of the receptor and to stimulation of many intracellular signalling pathways, including PLC, Ca2+ channels, PLD, PLA2, adneylate cyclase, MAP kinases, JAK-STAT pathway and NADPH oxidase, amongst others [48]. AT1R mediates most of the pathophysiological effects of Ang II, including vasoconstriction, inflammation growth, and fibrosis, whereas AT2R may counteract many of the AT1R-mediated actions [49, 50]. GPCR interact with G proteins as well as with accessory proteins, termed GPCR interacting proteins (GIP). GIP specifically associated with AT1R, called AT1R-associated protein (ATRAP), acts as a negative regulator of AT1R [51]. ATRAP-transgenic mice exhibit reduced neointimal formation, and decreased inflammatory responses and oxidative stress in response to Ang II, supporting the inhibitory role of ATRAP in cardiovascular remodelling [52]. Moreover, in ATRAP-/- mice, blood pressure and plasma volume are increased [53]. Another GIP has also been identified, ARAP1 (AT1R-associated protein 1), which is involved in the increase in membrane AT1R number following Ang II stimulation, and in the recycling of the receptor [54]. GPCR interacting proteins have also been associated with AT2R, including AT2R-interacting protein (ATIP), which is involved in transinactivation of receptor tyrosine kinases and growth inhibition [55]. ATIP is also known as mitochondrial tumor suppressor gene 1 (MTUS1) and AT2R binding protein of 50 kDa (ATBP50).

Page 3 of 11, 431

ATIP is expressed in tissues in which AT2R is abundant, including the uterus and adrenal tissue [56]. Both receptors play a role in regulating VSMC function, although they differ in their actions (Fig. 1). Whereas the AT1R is associated with growth, inflammation and vasoconstriction, the AT2R is generally associated with apoptosis and vasodilation [57, 58]. In pathological conditions, AT2R has also been shown to exert hypertrophic and pro-inflammatory actions. Ang II Signalling Through Mitogen-Activated Protein (MAP) Kinases, Tyrosine Kinases and RhoA/Rho Kinase Of the many growth-signalling pathways through which Ang II signals, is the MAP kinase family [59]. In VSMCs, Ang II activates all four of the major MAP kinases, including extracellular signal-regulated kinases (ERK1/2), p38MAP kinase, c-Jun N-terminal kinases (JNK) and ERK5 [60]. ERK1/2, phosphorylated by MEK1/2 (MAP/ERK kinase), is a key growth signalling kinase, whereas JNK and p38MAP kinase, phosphorylated by MEK4/7 and MEK3/6, respectively, influence cell survival, apoptosis, differentiation and inflammation. ERK5 is involved in protein synthesis, cell cycle progression and cell growth. Ang II induces phosphorylation of multiple tyrosine kinases, including c-Src, janus family kinases (JAK), focal adhesion kinase (FAK), Pyk2, p130Cas and phosphatidylinositol 3-kinase (PI3K) [61–63]. In human and rat VSMCs, c-Src is critically involved in trophic and contractile actions of Ang II [64, 65], effects that are augmented in vascular injury associated with hypertension [66, 67]. Ang II also activates receptor tyrosine kinases, even though it may not bind directly to these receptors. This process of transactivation has been demonstrated for epidermal growth factor (EGF) receptor, platelet-derived growth factor (PDGF) receptor, subtype β, and insulin-like 1 growth factor (IGF-1) receptor [68, 69]. Mechanisms whereby Ang II-induces receptor tyrosine kinase transactivation include activation of tyrosine kinases (Pyk2 and Src), redox-sensitive processes, and through a metalloprotease-dependent shedding of heparin-binding EGF-like growth factor (HB-EGF) [70]. The metalloprotease responsible for this process is ADAM17 [70]. Ang II also increases production of vasoactive hormones and growth factors, which could promote cell proliferation, protein synthesis and fibrosis, further contributing to growth processes in arterial remodelling. Activation of RhoA and its downstream target Rho-kinase play an important role in Ang II-mediated vascular contraction and growth [71, 72]. RhoA, a member of the Rho family of small GTPase binding proteins, is abundantly expressed in vascular smooth muscle cells and participates in vasoconstriction via phosphorylation of myosin light chain (MLC) and sensitization of contractile proteins to Ca2+. In VSMCs, Ang II/AT1R increases RhoA activity [73, 74] via the G12/13 family

431, Page 4 of 11

of G proteins as well as Gq [157]. RhoGEFs, which catalyse the exchange of GDP for GTP on RhoA, are sensitive to G12/ 13. In Ang II-infused rodents, vascular injury and hypertension are associated with increased vascular RhoA/Rho kinase activation, effects that are ameliorated by the Rho kinase inhibitors, fasudil or Y27632 [75, 76]. Ang II Signalling Through Reactive Oxygen Species in Vascular Cells Ang II mediates many of its molecular and cellular effects by stimulating production of ROS, which are highly reactive, bioactive, short-lived molecules derived from a reduction of molecular oxygen [77]. Of the many types of ROS generated, superoxide anion (•O2-), hydrogen peroxide (H2O2), nitric oxide (NO) and peroxynitrite (ONOO-) are particularly important in the cardiovascular system [78]. Vascular ROS are produced in endothelial, adventitial and VSMCs and generated mainly from NAD(P)H oxidase, a multi-subunit enzyme [79–81] that catalyses the production of •O2- by the one electron reduction of oxygen using NAD(P)H as the electron donor: 2O2 + NADPH→2O2- + NAD(P)H + H+. The prototypical NADPH oxidase (Nox) is phagocytic NADPH oxidase, which comprises five components: (phox for PHagocyte OXidase), p47phox, p67phox, p40phox, p22phox and gp91phox and the small G proteins Rac 1/2. The Nox family comprises seven isoforms, Nox1-5, Duox1 and Duox2. Unlike phagocytic NAD(P)H oxidase, which is activated only upon stimulation and which generates •O2- in a burst-like manner extracellularly, vascular oxidases (Nox1, Nox4, Nox5) are constitutively active, produce •O2- in a slow and sustained fashion and act as intracellular signalling molecules, influencing not only transcription factors, but other molecules involved in cell growth, contraction, migration, fibrosis and inflammation, such as MAP kinases, tyrosine kinases and protein phosphatases [82–84]. Nox1 is implicated in cell growth, inflammation and fibrosis and plays an important role in the development of atherosclerosis [85]. Nox4 has vasodilatory functions, is implicated in fibrosis and inflammation and may be both vasoprotective and vaso-injurious. [86, 87]. Nox5 is implicated in cell growth and inflammation and is upregulated in human atherosclerosis [88]. Increased Nox5 activity has recently been shown to be important in renal injury, proteinuria and hypertension [89]. All vascular Nox isoforms, including Nox1, Nox2, Nox4 and Nox5, are regulated by Ang II. The functional significance of different Nox isoforms still awaits clarification, but their differential tissue distribution, cellular localization and subcellular compartmentalization probably play a major role in Nox-specific actions. Ang II is a potent stimulator of vascular NAD(P)H oxidase [90, 91]. It induces activation of the enzyme, increases expression of NAD(P)H oxidase subunits and stimulates ROS production in cultured VSMC and intact arteries. Mechanisms linking Ang II to the enzyme and upstream signalling molecules modulating vascular Nox activity include PLD, PKC, c-

Curr Hypertens Rep (2014) 16:431

Src, PI3K and Rac [92•, 93, 94]. In hypertension, diabetes and atherosclerosis these processes are augmented, contributing to increased activation of the oxidase and consequent oxidative stress [95–97]. Superoxide and H2O2 activate multiple signalling molecules, including MAP kinases (p38MAP kinase, JNK, ERK-5 and ERK1/2), non-receptor tyrosine kinases (Src, JAK2, STAT, p21Ras, Pyk2 and Akt), receptor tyrosine kinases (EGFR, IGFR and PDGFR), protein tyrosine phosphatases, calcium channels and redox-sensitive transcription factors (NF-κB, AP-1 and HIF-1) [98–100]. Activation of these molecules participates in cell growth, migration, expression of pro-inflammatory genes, production of extracellular matrix proteins and contraction. All of these processes play important roles in vascular injury associated with cardiovascular diseases. Mechanisms whereby ROS modify signalling molecules are complex, but oxidative modification of proteins, e.g., protein tyrosine phosphatases (PTP), is important [101]. Inhibition of Nox is now being considered as a possible therapeutic target in patients. [102•]. Noxs may also be targets of conventional cardiovascular drugs such as ACE inhibitors, AT1 receptor blockers and calcium channel blockers, and effects of these agents may be mediated, in part, by decreasing Nox-derived ROS generation. Clinical studies support a direct antioxidant effect of AT1 receptor blockers, as demonstrated in hypertensive patients treated with candesartan, where oxidative stress and inflammation were reduced independently of blood pressure-lowering actions [103, 104].

Angiotensin II, Vascular Remodelling and Inflammation Vascular smooth muscle cells are critically involved in maintaining vascular integrity and tone. They are the main constitutive cells of the vascular wall, assuming various functional and structural functions. Vascular smooth muscle cells are dynamic, highly plastic, multifunctional cells that contribute to arterial remodelling by influencing cell growth (hyperplasia and hypertrophy), apoptosis, cell elongation, reorganization, altered production of extracellular matrix proteins, fibrosis, migration and inflammation [105, 106] (Fig. 2). Moreover, quiescent contractile VSMCs can be differentiated into a proliferative phenotype under pathological conditions. Such plasticity and phenotypic switching are major contributors to vascular injury in various cardiovascular diseases, and the RAS is critically involved in these processes. Ang II, through the AT1R, promotes a phenotypic switch from a contractile to a proliferative and synthetic phenotype of VSMCs, leading to changes in the contractile machinery, vascular hypertrophy/proliferation and secretion of contractile, mitogenic, inflammatory and pro-fibrotic mediators [8]. These processes are controlled by transcriptional factors

Curr Hypertens Rep (2014) 16:431

Page 5 of 11, 431

Fig. 2 Mechanisms whereby Ang II induces vascular injury. Binding of Ang II to its cell membrane receptor induces impairs NO synthesis by reducing eNOS activity and promoting NOS uncoupling. Ang II also activates NADPH oxidase (Nox), leading to increased generation of reactive oxygen species (ROS), that are proinflammatory, mitogenic, prothrombotic and pro-fibrotic. These events also contribute to

atherogenesis, through macrophage infiltration and foam cell formation. Ang II activates signaling pathways through calcium mobilization and MAP kinase activation and by inhibiting cyclic GMP (cGMP), leading to altered vascular reactivity, growth, calcification and fibrosis. Together, these processes contribute to vascular injury that underlies many cardiovascular diseases

regulated in large part by Ang II. Growing evidence also implicates microRNAs (miR) in VSMC differentiation and phenotypic switching. In particular, miR-143/145, miR-21, and miR-1 have been shown to promote the contractile phenotype, whereas miR-221, miR-146a, miR-24, and miR-26a may be important in the switch to the synthetic phenotype and cell proliferation associated with vascular injury [107]. Changes in intracellular calcium and magnesium, MAPK and Rho kinase signalling and generation of ROS, are all important in Ang II-mediated VSMC effects [8, 105–107]. Vascular fibrosis involves accumulation of extracellular matrix proteins, such as collagen, elastin, fibrillin, fibronectin and proteoglycans in the tunica media and is increased in vascular remodelling. Collagen accumulation may be a consequence of increased synthesis [108, 109], reduced degradation by matrix metalloproteinases (MMP) and/or enhanced tissue inhibitor of metalloproteinase-1 (TIMP-1) production [110]. All of these processes are influenced by Ang II. Apoptosis modulates fine-tuning of media growth and may be a growth-associated compensatory or adaptive process in Ang II-mediated vascular remodelling. An imbalance between growth and apoptosis may be important in the structural changes associated with vascular injury. Ang II-elicited growth/apoptosis and profibrotic effects are modulated, in part, by endogenous production of mitogenic factors, such as TGF-β, PDGF, EGF, IGF-1 and ET-1 [111, 112], and activation of mitogenic and pro-fibrotic pathways. We recently identified a new pathway for Ang II-mediated vascular cell growth and migration through Fat-1, an atypical cadherin, and Nox1 [113].

Vascular inflammation is characterized by recruitment of monocytes and lymphocytes into the subendothelial space, production of chemotactic cytokines, increased expression of adhesion molecules, reactive smooth muscle cell proliferation, platelet aggregation, altered extracellular matrix production and degradation. These processes, together with lipid oxidation, are pro-atherogenic, particularly in injured arteries in hypertension and diabetes. Ang II has significant proinflammatory actions inducing the production of ROS, cytokines, adhesion molecules and activation of redox-sensitive inflammatory genes in vascular cells [113–115]. Cytokines, chemokines and other pro-inflammatory mediators are also produced by circulating leukocytes and lymphocytes and by resident macrophages in the vessel wall in response to Ang II stimulation. These pro-inflammatory responses are associated with activation of the adaptive immune system. T lymphocytes, particularly effector T cells such as Thelper (Th)1 (interferon-gamma-producing) and Th2 lymphocytes [that produce interleukin (IL)-4], as well as Th17 (that produce IL-17), and T suppressor lymphocytes including regulatory T cells (Treg), which express the transcription factor forkhead box P3 (Foxp3), play critical roles in the vasculopathy associated with Ang II-dependent hypertension [116•, 117, 118]. Inflammation participates in vascular remodelling and atherosclerosis [119, 120], and may contribute to accelerated vascular damage in hypertension, diabetes, renal disease and in aging. Inflammation may activate the RAS, and thereby further contribute to vascular remodelling and hypertension. Activators of nuclear receptors, such as the peroxisome

431, Page 6 of 11

proliferator activated receptors (PPARs), which are hypolipidemic agents (the fibrates, PPARα agonists) or insulin sensitizers (glitazones, PPARγ agonists), downregulate the vascular inflammatory response in experimental animals and decrease serum markers of inflammation in humans [121, 122]. Thus, PPARs and vasoactive substances may be endogenous modulators of the inflammatory process involved in vascular injury. Ang II downregulates PPARs through activation of nuclear factor (NF)-κB [123].

Angiotensin II and Vascular Calcification Calcification of arteries predicts events in patients with cardiovascular disease [121, 122]. It is common in advanced age, chronic kidney disease, diabetes mellitus, atherosclerosis and hypertension, and is related to cardiovascular morbidity and mortality [124–126]. Vascular calcification is a complex and dynamic process involving many mechanisms that promote accumulation of calcium deposits in the vessel wall. This leads to increased arterial stiffness and decreased elasticity, which in turn impacts on hemodynamics and cardiovascular dysfunction. Vascular calcification is associated with mineralization of the internal elastic lamina and elastic fibres within the medial layer. It is a highly regulated process, not dissimilar to processes that control bone formation [127]. Factors that trigger and promote calcification include abnormalities in mineral metabolism, particularly hyperphosphatemia and hypercalcemia [129–132], which stimulate VSMC differentiation to an osteoblastic phenotype [128, 129]. This is driven by upregulation of transcription factors such as cbfa1 (core-binding factor 1α)/Runx2, MSX-2 and bone morphogenetic protein 2 (BMP-2) that are critically involved in normal bone development, and control the expression of osteogenic proteins, including osteocalcin, osteonectin, alkaline phosphatase, collagen-1 and bone sialoprotein [130, 131]. Another mechanism contributing to vascular mineralization is loss of calcification inhibitors, such as fetuin-A, matrix Gla protein, pyrophosphate and osteopontin [132, 133]. These events are influenced by Ang II, which induces upregulation of osteogenic proteins and signalling pathways. In particular, Ang II stimulates expression of bone morphogenetic protein 2 (BMP2) and the osteoblast transcription factor Runx2/Cbfa1 [134•], and it stimulates transcellular transport of cations such as Ca2+ and Mg2 [135, 136]. Ang II also inhibits matrix Gla, an important inhibitor of calcification [137•]. Targeting the RAS may have therapeutic potential in vascular calcification. AT1R blockade can inhibit arterial calcification by disrupting vascular osteogenesis, suggesting that patients with vascular calcification may benefit therapeutically from ARBs. The vasoprotective axis of the RAS Ang-(1-7)/Mas has also been shown to

Curr Hypertens Rep (2014) 16:431

prevent vascular calcification in experimental models, by decreasing the expression of osteogenic proteins [138].

Angiotensin II and Atherosclerosis Atherosclerosis is a chronic inflammatory disease of the vascular wall characterized by plaque formation in the intimal layer. The early hallmark of atherosclerosis, so-called fatty streaks, is already evident during the first decade of life. However, with time the small lesions progress to become fibrous plaques and complicated lesions, characterized by a decrease in lumen diameter and consequent obstruction of blood flow [139]. Although the exact mechanisms that trigger atherogenesis are unclear, increasing evidence indicates that oxidation of Low Density Lipoprotein (LDL) is one of the main early events of the process [140]. LDL diffuses from the plasma to the subendothelial space, where it is oxidatively modified by ROS or enzymes [141]. Oxidized LDL (oxLDL) has pro-inflammatory properties. oxLDL interacts with different membrane and nuclear receptors, such as CD36, PPARγ, and receptor for platelet-activating factor (PAFR), which stimulate pro-inflammatory signalling pathways [142–145]. oxLDL also activates endothelial cells and stimulates the expression of adhesion molecules such as E-selectin, vascular cell adhesion protein 1 (VCAM-1), and intercellular adhesion molecule-1 (ICAM-1). The production of chemokines attracts monocytes and T cells to the arterial intima, promoting an intense activation of the immune response. Circulating leukocytes are continuously recruited into the vessel wall, where monocytes differentiate into macrophages. Macrophages in turn take up oxLDL through the scavenger receptors, CD36 and SR-A, and transform into foam cells that are present in every stage of atherogenesis. Continuous recruitment of circulating leukocytes into the vessel wall further promotes inflammation [146]. oxLDL also interacts with CD36 and Toll-like receptor 4 in VSMCs, enhancing the production of inflammatory cytokines (IL-1β, IL-6, IL-8, and MCP-1) and growth factors such as granulocyte-macrophage colony-stimulating factor (GMCSF) and granulocyte-colony-stimulating factor (G-CSF) that induce recruitment and differentiation of macrophages [147•, 148•]. VSMCs migrate from the media to the arterial intima, where they proliferate, differentiate and produce extracellular matrix to form the fibrous cap present in advanced atherosclerosis lesions, with a very important role in the plaque stability. Lectin-like oxidized LDL receptor-1 (LOX-1) is a scavenger receptor present in endothelial cells and VSMCs, and promotes the interaction with oxLDL. The expression of LOX-1 is induced by inflammatory stimuli, ROS, oxLDL and Ang II [149]. Moreover downstream activation of LOX1 by oxLDL results in the generation of large amounts of ROS [150] and upregulates AT1R via NADPH oxidase, MAPK and

Curr Hypertens Rep (2014) 16:431

the NF-κB pathway [151], demonstrating the crosstalk between AT1R and LOX-1. LOX-1, through PKC activation, inhibits the eNOS pathway and eNOS-dependent NO production [152] and the inhibition of LOX-1 prevents endothelial dysfunction. The RAS is a key player in the pathogenesis of atherosclerosis. Expression of AT1R is increased in atherosclerosis [153], and experimental data have shown that inhibition of AT1R has protective effects on the development of atherosclerosis and metabolic disorders [154]. On the other hand, development of atherosclerosis is reduced in pro-atherogenic mice that have overexpression of AT2R [153]. Ang II stimulates expression of adhesion molecules, chemokines and cytokines, and promotes endothelial dysfunction, oxidation and uptake of LDL and proliferation of VSMCs. Many of these proatherogenic effects are mediated through Nox-derived ROS. We recently demonstrated that Nox1 plays a critical role in vascular injury associated with accelerated atherosclerosis in diabetes [85]. In advanced atherosclerosis, Ang II stimulates expression of matrix metalloproteinases (MMPs) and plasminogen activator inhibitor-1, with resultant destabilization of atherosclerotic plaques.

Page 7 of 11, 431 BHF Chair. ACM is supported by a Leadership fellowship from the University of Glasgow. Compliance with Ethics Guidelines Conflict of Interest Augusto C. Montezano, Aurelie Nguyen Dinh Cat, Francisco J. Rios, and Rhian M. Touyz declare that they have no conflict of interest. Human and Animal Rights and Informed Consent This article does not contain any studies with human or animal subjects performed by any of the authors.

References Papers of particular interest, published recently, have been highlighted as: • Of importance

1.

2.

Conclusions 3.

Vascular injury, characterized by endothelial dysfunction, structural remodelling and inflammation, underlies many cardiovascular disorders, including hypertension, atherosclerosis, diabetes and ischemic heart disease. At the cellular level, there is increased VSMC growth, migration, differentiation, calcification, production of extracellular matrix proteins and inflammation, stimulated in large part by activation of the RAS. Over the recent past, our views of Ang II have changed from being a simple vasoconstrictor to that of a complex growth factor that mediates effects through diverse signalling pathways involving PLC/PKC/Ca2+ mobilization, PLA2, PLD, MAP kinases, tyrosine kinases, proto-oncogene expression, RhoA/Rho kinase, and oxidative stress. Through increased Nox-derived ROS generation and activation of redoxsensitive transcription factors, Ang II promotes expression of cell adhesion molecules and induces synthesis of proinflammatory mediators and growth factors. These molecular and cellular processes facilitate increased vascular permeability, leukocyte recruitment, calcification and vascular fibrosis leading to vascular injury. Targeting some of these molecular events with RAS inhibitors may protect against vascular injury and promote vascular health. Acknowledgments Studies performed by RMT were supported by grants from the Canadian Institutes of Health Research (CIHR), JDRF and the British Heart Foundation (BHF). RMT is supported through a

4.

5.

6.

7. 8.

9.

10.

11. 12.

13.

Schiffrin EL, Touyz RM. From bedside to bench to bedside: role of renin-angiotensin-aldosterone system in remodeling of resistance arteries in hypertension. Am J Physiol Heart Circ Physiol. 2004;287(2):H435–46. Touyz RM. The role of angiotensin II in regulating vascular structural and functional changes in hypertension. Curr Hypertens Rep. 2003;5(2):155–64. Aroor AR, Demarco VG, Jia G, Sun Z, Nistala R, Meininger GA, et al. The role of tissue renin-angiotensin-aldosterone system in the development of endothelial dysfunction and arterial stiffness. Front Endocrinol. 2013;4:161–8. Ferrario CM. New physiological concepts of the renin-angiotensin system from the investigation of precursors and products of angiotensin I metabolism. Hypertension. 2010;55(2):445–52. Ferrario CM, Ahmad S, Nagata S, Simington SW, Varagic J, Kon N, et al. An evolving story of angiotensin-II-forming pathways in rodents and humans. Clin Sci (Lond). 2014;126(7):461–9. Bader M. Tissue renin-angiotensin-aldosterone systems: targets for pharmacological therapy. Annu Rev Pharmacol Toxicol. 2010;50:439–65. Gibbons GH, Dzau VJ. The emerging concept of vascular remodeling. N Engl J Med. 1994;330:1431–8. Intengan HD, Schiffrin EL. Vascular remodeling in hypertension: roles of apoptosis, inflammation, and fibrosis. Hypertension. 2001;38:581–7. Martinez-Lemus LA, Hill MA, Meininger GA. The plastic nature of the vascular wall: a continuum of remodeling events contributing to control of arteriolar diameter and structure. Physiology. 2009;24:45–57. Harrison DG, Vinh A, Lob H, Madhur MS. Role of the adaptive immune system in hypertension. Curr Opin Pharmacol. 2010;10(2):203–7. Thatcher S. The adipose renin-angiotensin system: role in cardiovascular disease. Mol Cell Endocrinol. 2009;302(2):111–7. Guzik TJ, Hoch NE, Brown KA, McCann LA, Rahman A, Dikalov S, et al. Role of the T cell in the genesis of angiotensin II induced hypertension and vascular dysfunction. J Exp Med. 2007;204(10):2449–60. Prieto MC, Gonzalez AA, Navar LG. Evolving concepts on regulation and function of renin in distal nephron. Pflugers Arch. 2013;465(1):121–32.

431, Page 8 of 11 14. 15.

16. 17.

18. 19. 20.

21.

22.

23.

24. 25.

26.

27. 28. 29.

30.

31. 32.

33.

34.

35.

Cassis LA. Local adipose tissue renin-angiotensin system. Curr Hypertens Rep. 2008;10(2):93–8. Miyazaki M, Takai S. Tissue angiotensin II generating system by angiotensin-converting enzyme and chymase. J Pharmacol Sci. 2006;100(5):391–7. Reudelhuber TL. Prorenin, Renin, and their receptor: moving targets. Hypertension. 2010;55(5):1071–107. Danser AJ, Nguyen G. The Renin Academy Summit: advancing the understanding of renin science. J Renin-AngiotensinAldosterone Syst. 2008;9(2):119–23. Nguyen G, Muller DN. The biology of the (pro)renin receptor. J Am Soc Nephrol. 2010;21(1):8–23. Nguyen G. The (pro)renin receptor in health and disease. Ann Med. 2010;42(1):13–8. Hitom H, Liu G, Nishiyama A. Role of (pro)renin receptor in cardiovascular cells from the aspect of signaling. Front Biosci. 2010;2:1246–9. Advani A, Kelly DJ, Cox AJ, White KE, Advani SL, Thai K, et al. The (pro)renin receptor: site-specific and functional linkage to the vacuolar H+ -ATPase in the kidney. Hypertension. 2009;54:261– 9. Cousin C, Bracquart D, Contrepas A, Nguyen G. Potential role of the (pro)renin receptor in cardiovascular and kidney diseases. J Nephrol. 2010;23:508–13. Xia H, Lazartigues E. Angiotensin-converting enzyme 2: central regulator for cardiovascular function. Curr Hypertens Rep. 2010;12(3):170–5. Bader M. ACE2, angiotensin-(1–7), and Mas: the other side of the coin. Pflugers Arch. 2013;465(1):79–85. Shaltout HA. Alterations in circulatory and renal angiotensin-converting enzyme and angiotensin-converting enzyme 2 in fetal programmed hypertension. Hypertension. 2009;53(2):404–8. Fraga-Silva RA, Savergnini SQ, Montecucco F, Nencioni A, Caffa I, Soncini D, et al. Treatment with Angiotensin-(1-7) reduces inflammation in carotid atherosclerotic plaques. Thromb Haemost. 2014;111(4):736–47. Imai Y. Angiotensin-converting enzyme 2 (ACE2) in disease pathogenesis. Circ J. 2010;74(3):405–10. Ferrario CM, Iyer SN. Angiotensin-(1-7): a bioactive fragment of the renin-angiotensin system. Regul Pept. 1998;78(1–3):13–8. Santiago NM. Lifetime overproduction of circulating Angiotensin-(1-7) attenuates deoxycorticosterone acetate-salt hypertension-induced cardiac dysfunction and remodeling. Hypertension. 2010;55(4):889–96. Cangussu LM. Angiotensin-(1-7) antagonist, A-779, microinjection into the caudal ventrolateral medulla of renovascular hypertensive rats restores baroreflex bradycardia. Peptides. 2009;30(10):1921–7. Silva DM. Evidence for a new angiotensin-(1-7) receptor subtype in the aorta of Sprague-Dawley rats. Peptides. 2007;28(3):702–7. Santos RA, Ferreira AJ, Simões E, Silva AC. Recent advances in the angiotensin-converting enzyme 2-angiotensin(1-7)-Mas axis. Exp Physiol. 2008;93(5):519–27. Sampaio WO, Souza dos Santos RA, Faria-Silva R, da Mata Machado LT, Schiffrin EL, Touyz RM. Angiotensin-(1-7) counter-regulates angiotensin II signaling in human endothelial cells. Hypertension. 2007;50(6):1093–8. Gwathmey TM, Pendergrass KD, Reid SD, Rose JC, Diz DI, Chappell MC. Angiotensin-(1-7)-angiotensin-converting enzyme 2 attenuates reactive oxygen species formation to angiotensin II within the cell nucleus. Hypertension. 2010;55(1):66–171. Santos RA, Frézard F, Ferreira AJ. Angiotensin-(1-7): blood, heart, and blood vessels. Curr Med Chem Cardiovasc Hematol Agents. 2005;3(4):383–91.

Curr Hypertens Rep (2014) 16:431 36.

37.

38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50. 51.

52.

53.

54.

55.

56.

Reaux A, Fournie-Zaluski MC, Llorens-Cortes C. Angiotensin III: a central regulator of vasopressin release and blood pressure. Trends Endocrinol Metab. 2001;12(4):157–62. Yang R, Walther T, Gembardt F, Smolders I, Vanderheyden P, Albiston AL, et al. Renal vasoconstrictor and pressor responses to angiotensin IV in mice are AT1a-receptor mediated. J Hypertens. 2010;28(3):487–94. Ferreira PM, Souza Dos Santos RA, Campagnole-Santos MJ. Angiotensin-(3-7) pressor effect at the rostral ventrolateral medulla. Regul Pept. 2007;141(1–3):168–74. Ocaranza MP, Jalil JE. Protective role of the ACE2/Ang-(1-9) axis in cardiovascular remodeling. Int J Hypertens. 2012;2012:594– 361. Ahmad S, Wei CC, Tallaj J, Dell'Italia LJ, Moniwa N, Varagic J, et al. Chymase mediates angiotensin-(1-12) metabolism in normal human hearts. J Am Soc Hypertens. 2013;7(2):128–36. Vaajanen A, Luhtala S, Oksala O, Vapaatalo H. Does the reninangiotensin system also regulate intra-ocular pressure? Ann Med. 2008;40(6):418–27. Wilkinson-Berka JL, Agrotis A, Deliyanti D. The retinal reninangiotensin system: roles of angiotensin II and aldosterone. Peptides. 2012;36(1):142–50. Kumar R, Singh VP, Baker KM. The intracellular reninangiotensin system in the heart. Curr Hypertens Rep. 2009;11(2):104–10. deGasparo M, Catt KJ, Inagami T, Wright JW, Unger T. International union of pharmacology XXII. The angiotension II receptors. Pharmacol Rev. 2000;52:415–72. Namsolleck P, Recarti C, Foulquier S, Steckelings UM, Unger T. AT(2) receptor and tissue injury: therapeutic implications. Curr Hypertens Rep. 2014;16(2):416. Carey RM, Padia SH. Role of angiotensin AT(2) receptors in natriuresis: Intrarenal mechanisms and therapeutic potential. Clin Exp Pharmacol Physiol. 2013;40(8):527–34. Levy BI. How to explain the differences between reninangiotensin system modulators. Am J Hypertens. 2005;18: 134S–41S. Higuchi S, Ohtsu H, Suzuki H, Shirai H, Frank GD, Eguchi S. Angiotensin II signal transduction through the AT1 receptor: novel insights into mechanisms and pathophysiology. Clin Sci (Lond). 2007;112(8):417–28. Matsubara H. Pathophysiological role of angiotensin II type 2 receptor in cardiovascular and renal disease. Circ Res. 1998;83: 1182–91. Steckelings UM, Kaschina E, Unger. The AT2 receptor—A matter of love or hate. Peptides. 2005;26:1401–9. Tamura K, Wakui H, Maeda A, Dejima T, Ohsawa M, Azushima K, et al. The physiology and pathophysiology of a novel angiotensin receptor-binding protein ATRAP/Agtrap. Curr Pharm Des. 2013;19(17):3043–8. Wakui H, Dejima T, Tamura K, Uneda K, Azuma K, Maeda A, et al. Activation of angiotensin II type 1 receptor-associated protein exerts an inhibitory effect on vascular hypertrophy and oxidative stress in angiotensin II-mediated hypertension. Cardiovasc Res. 2013;100(3):511–9. Oppermann M, Gess B, Schweda F, Castrop H. Atrap deficiency increases arterial blood pressure and plasma volume. J Am Soc Nephrol. 2010;21(3):468–77. Mogi M, Iwai M, Horiuchi M. Emerging concepts of regulation of angiotensin II receptors: new players and targets for traditional receptors. Arterioscler Thromb Vasc Biol. 2007;27(12):2532–9. Nouet S, Amzallag N. Trans-inactivation of receptor tyrosine kinases by novel angiotensin II AT2 receptor-interacting protein, ATIP. J Biol Chem. 2004;279:28989–97. Di Benedetto M. Structural organization and expression of human MTUS1, a candidate 8p22 tumor suppressor gene encoding a

Curr Hypertens Rep (2014) 16:431

57.

58.

59.

60.

61.

62.

63.

64.

65.

66.

67.

68.

69.

70.

71. 72.

73.

family of angiotensin II AT2 receptor-interacting proteins, ATIP. Gene. 2006;380(2):127–36. Henrion D, Kubis N, Lévy BI. Physiological and pathophysiological functions of the AT2 subtype receptor of angiotensin II - From large arteries to the microcirculation. Hypertension. 2001;38:1150–7. Akishita M. Inflammation influences vascular remodeling through AT2 receptor expression and signaling. Physiol Genomics. 2000;2:13–20. Touyz RM, He G, Deng LY, Schiffrin EL. Role of extracellular signal-regulated kinases in angiotensin II-Stimulated contraction of smooth muscle cells from human resistance arteries. Circulation. 1999;99:392–9. Xi XP. Central role of the MAPK pathway in Ang II-mediated DNA synthesis and migration in rat vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 1999;19:73–82. Eguchi S, Inagami T. Signal transduction of angiotensin II type 1 receptor through receptor tyrosine kinase. Regul Pept. 2000;91:13– 20. Montezano AC, Callera GE, Yogi A, He Y, Tostes RC, He G, et al. Aldosterone and angiotensin II synergistically stimulate migration in vascular smooth muscle cells through c-Src-regulated redoxsensitive RhoA pathways. Arterioscler Thromb Vasc Biol. 2008;28(8):1511–8. Touyz RM, Schiffrin EL. Role of calcium influx and intracellular calcium stores in angiotensin II-mediated calcium hyperresponsiveness in smooth muscle from spontaneously hypertensive rats. J Hypertens. 1997;15:1431–9. Touyz RM, Yao G, Schiffrin EL. c-Src induces phosphorylation and translocation of p47phox - Role in superoxide generation by angiotensin II in human vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 2003;23:981–7. Callera GE, Yogi A, Briones AM, Montezano AC, He Y, Tostes RC, et al. Vascular proinflammatory responses by aldosterone are mediated via c-Src trafficking to cholesterol-rich microdomains: role of PDGFR. Cardiovasc Res. 2011;91(4):720–31. Touyz RM. Increased angiotensin II-mediated Src signaling via epidermal growth factor receptor transactivation is associated with decreased c-terminal Src kinase activity in vascular smooth muscle cells from spontaneously hypertensive rats. Hypertension. 2002;39:479–85. Touyz RM. Src is an important mediator of extracellular signalregulated kinase 1/2-dependent growth signaling by angiotensin II in smooth muscle cells from resistance arteries of hypertensive patients. Hypertension. 2001;38:56–64. Saito Y, Berk BC. Transactivation: a novel signaling pathway from angiotensin II to tyrosine kinase receptors. J Mol Cell Cardiol. 2001;33:3–7. Burger D, Montezano AC, Nishigaki N, He Y, Carter A, Touyz RM. Endothelial microparticle formation by angiotensin II is mediated via Ang II receptor type I/NADPH oxidase/ Rho kinase pathways targeted to lipid rafts. Arterioscler Thromb Vasc Biol. 2011;31(8):1898–907. Ohtsu H, Dempsey PJ, Frank GD, Brailoiu E, Higuchi S, Suzuki H, et al. ADAM17 mediates epidermal growth factor receptor transactivation and vascular smooth muscle cell hypertrophy induced by angiotensin II. Arterioscler Thromb Vasc Biol. 2006;26(9):e133–7. Loirand G, Pacaud P. The role of Rho protein signaling in hypertension. Nat Rev Cardiol. 2010;7(11):637–47. Uehata M, Ishizaki T, Satoh H. Calcium sensitization of smooth muscle mediated by a Rho-associated protein kinase in hypertension. Nature. 1997;389(6654):990–4. Guilluy C, Bregeon J, Tourmaniantz G. The Rho exchange factor Arhgef1 mediates the effects of angiotensin II on vascular tone and blood pressure. Nat Med. 2010;16:183–90.

Page 9 of 11, 431 74.

75.

76.

77.

78.

79.

80.

81.

82.

83. 84. 85.

86.

87.

88.

89.

90.

91.

92.•

Bregeon J, Loirand G, Pacaud P. Angiotensin II induces RhoA activation through SHP2-dependent dephosphorylation of the RhoGAP p190A in vascular smooth muscle cells. Am J Physiol Cell Physiol. 2009;297(5):C1062–70. Cario-Toumaniantz C, Ferland-McCollough D, Chadeuf G, Toumaniantz G, Rodriguez M, Galizzi JP, et al. RhoA guanine exchange factor expression profile in arteries: evidence for a Rho kinase-dependent negative feedback in angiotensin II-dependent hypertension. Am J Physiol Cell Physiol. 2012;302:C1394–404. Kobayashi N, Nakano S, Mita S, Kobayashi T, Honda T, Tsubokou Y, et al. Involvement of Rho-kinase pathway for angiotensin II-induced plasminogen activator inhibitor-1 gene expression and cardiovascular remodeling in hypertensive rats. J Pharmacol Exp Ther. 2002;301:459–66. Touyz RM. Reactive oxygen species and angiotensin II signaling in vascular cells – implications in cardiovascular disease. Braz J Med Biol Res. 2004;37(8):1263–73. Taniyama Y, Griendling KK. Reactive oxygen species in the vasculature: molecular and cellular mechanisms. Hypertension. 2003;42:1075–81. Touyz RM, Briones AM, Sedeek M, Burger D, Montezano AC. NOX isoforms and reactive oxygen species in vascular health. Mol Interv. 2011;11(1):27–35. Lassegue B, Clempus RE. Vascular NAD(P)H oxidases: specific features, expression, and regulation. Am J Physiol Regul Integr Comp Physiol. 2003;285(2):R277–97. Rey FE, Pagano PJ. The reactive adventitia: fibroblast oxidase in vascular function. Arterioscler Thromb Vasc Biol. 2002;22(12): 1962–71. Montezano AC, Touyz RM. Molecular mechanisms of hypertension–reactive oxygen species and antioxidants: a basic science update for the clinician. Can J Cardiol. 2012;28(3):288–95. Griendling KK. Novel NAD(P)H oxidases in the cardiovascular system. Heart. 2004;90(5):491–3. Briones AM, Touyz RM. Oxidative stress and hypertension: current concepts. Curr Hypertens Rep. 2010;12(2):135–42. Gray SP, Di Marco E, Okabe J, Szyndralewiez C, Heitz F, Montezano AC, et al. NADPH oxidase 1 plays a key role in diabetes mellitus-accelerated atherosclerosis. Circulation. 2013;127(18):1888–902. Nguyen Dinh Cat A, Montezano AC, Burger D, Touyz RM. Angiotensin II, NADPH oxidase, and redox signaling in the vasculature. Antioxid Redox Signal. 2013;19(10):1110–20. Schramm A, Matusik P, Osmenda G, Guzik TJ. Targeting NADPH oxidases in vascular pharmacology. Vasc Pharmacol. 2012;56(5–6):216–31. Montezano AC, Burger D, Paravicini TM, Chignalia AZ, Yusuf H, Almasri M, et al. Nicotinamide adenine dinucleotide phosphate reduced oxidase 5 (Nox5) regulation by angiotensin II and endothelin-1 is mediated via calcium/calmodulin-dependent, rac1-independent pathways in human endothelial cells. Circ Res. 2010;106(8):1363–73. Holterman CE, Thibodeau JF, Towaij C, Gutsol A, Montezano AC, Parks RJ, et al. Nephropathy and elevated bp in mice with podocyte-specific NADPH oxidase 5 expression. J Am Soc Nephrol. 2013;25:784–97. Montezano AC, Touyz RM. Reactive oxygen species, vascular Noxs, and hypertension: focus on translational and clinical research. Antioxid Redox Signal. 2014;20(1):164–82. Touyz RM, Chen X, Tabet F, Yao G, He G, Quinn MT, et al. Expression of a gp91phox-containing leukocyte-type NADPH oxidase in human vascular smooth muscle cells—modulation by Ang II. Circ Res. 2006;90:1205–13. Paravicini TM, Montezano AC, Yusuf H. Touyz RM activation of vascular p38MAPK by mechanical stretch is independent of c-Src and NADPH oxidase: influence of hypertension and angiotensin

431, Page 10 of 11

93.

94.

95. 96.

97.

98.

99. 100. 101.

102.•

103.

104.

105.

106.

107.

108.

109.

110.

II. J Am Soc Hypertens. 2012;6(3):169–78. This study clearly identifies differential mechanisms whereby Ang II and blood pressure influence mechanosensitive signaling in vascular smooth muscle cells. Touyz RM, Schiffrin EL. Ang II-stimulated superoxide production is mediated via phospholipase D in human vascular smooth muscle cells. Hypertension. 1999;34(4):976–82. Touyz RM, Schiffrin EL. Increased generation of superoxide by angiotensin II in smooth muscle cells from resistance arteries of hypertensive patients: role of phospholipase D-dependent NAD(P)H oxidase-sensitive pathways. J Hypertens. 2001;19: 1245–54. Touyz RM. Oxidative stress and vascular damage in hypertension. Curr Hypertens Rep. 2001;2:98–105. Petrovič D. Association of the -262C/T polymorphism in the catalase gene promoter and the C242T polymorphism of the NADPH oxidase P22phox gene with essential arterial hypertension in patients with diabetes mellitus type 2. Clin Exp Hypertens. 2014;36(1):36–40. Virdis A, Neves MF, Amiri F, Touyz RM, Schiffrin EL. Role of NAD(P)H oxidase on vascular alterations in angiotensin IIinfused mice. J Hypertens. 2004;22(3):535–42. Chiarugi P, Cirri P. Redox regulation of protein tyrosine phosphatases during receptor tyrosine kinase signal transduction. Trends Biochem Sci. 2003;28:509–14. Nathan C. Specificity of a third kind: reactive oxygen and nitrogen intermediates in cell signaling. J Clin Invest. 2003;111:769–78. Torres M, Forman HJ. Redox signaling and the MAP kinase pathways. Biofactors. 2003;17:287–96. Tabet F, Schiffrin EL, Callera GE, He Y, Yao G, Ostman A, et al. Redox-sensitive signaling by angiotensin II involves oxidative inactivation and blunted phosphorylation of protein tyrosine phosphatase SHP-2 in vascular smooth muscle cells from SHR. Circ Res. 2008;103(2):149–58. Sedeek M, Gutsol A, Montezano AC, Burger D, Nguyen Dinh Cat A, Kennedy CR, et al. Renoprotective effects of a novel Nox1/4 inhibitor in a mouse model of Type 2 diabetes. Clin Sci (Lond). 2013;124(3):191–202. An in vivo study demonstrating that an orally active inhibitor of Nox1 and Nox4 ameliorates development of nephropathy in a model of diabetes, indicating a role for Nox1/4-derived reactive oxygen in diabetic nephropathy. Dohi Y. Candesartan reduces oxidative stress and inflammation in patients with essential hypertension. Hypertens Res. 2003;26: 691–7. Kamiyama M, Urushihara M, Morikawa T, Konishi Y, Imanishi M, Nishiyama A, et al. Oxidative stress/angiotensinogen/reninangiotensin system axis in patients with diabetic nephropathy. Int J Mol Sci. 2013;14(11):23045–62. Lacolley P, Regnault V, Nicoletti A, Li Z, Michel JB. The vascular smooth muscle cell in arterial pathology: a cell that can take on multiple roles. Cardiovasc Res. 2012;95(2):194–204. Savoia C, Burger D, Nishigaki N, Montezano A, Touyz RM. Angiotensin II and the vascular phenotype in hypertension. Expert Rev Mol Med. 2011;13:e11. Albinsson S, Sessa WC. Can microRNAs control vascular smooth muscle phenotypic modulation and the response to injury? Physiol Genomics. 2011;43(10):529–33. Tayebjee MH, MacFadyen RJ, Lip GY. Extracellular matrix biology: a new frontier in linking the pathology and therapy of hypertension? J Hypertens. 2003;21:2211–8. Delva P. Collagen I, and III mRNA gene expression and cell growth potential of skin fibroblasts in patients with essential hypertension. J Hypertens. 2002;20:1393–9. Castro MM, Rizzi E, Prado CM, Rossi MA, Tanus-Santos JE, Gerlach RF. Imbalance between matrix metalloproteinases and

Curr Hypertens Rep (2014) 16:431

111.

112. 113.

114.

115.

116.•

117.

118.

119.

120.

121.

122.

123.

124.

125.

126.

tissue inhibitor of metalloproteinases in hypertensive vascular remodeling. Matrix Biol. 2010;29(3):194–201. Satoh C. Role of endogenous angiotensin II in the increased expression of growth factors in vascular smooth muscle cells from spontaneously hypertensive rats. J Cardiovasc Pharmacol. 2001;37(1):108–18. Bobik A. Hypertension, transforming growth factor-β, angiotensin II and kidney disease. J Hypertens. 2004;22(7):1265–7. Bruder-Nascimento T, Chinnasamy P, Riascos-Bernal DF, Cau SB, Callera GE, Touyz RM, et al. Angiotensin II induces Fat1 expression/activation and vascular smooth muscle cell migration via Nox1-dependent reactive oxygen species generation. J Mol Cell Cardiol. 2014;66:18–26. Viel EC, Lemarié CA, Benkirane K, Paradis P, Schiffrin EL. Immune regulation and vascular inflammation in genetic hypertension. Am J Physiol Heart Circ Physiol. 2010;298(3): H938–44. Zhao Q, Ishibashi M, Hiasa K, Tan C, Takeshita A, Egashira K. Essential role of vascular endothelial growth factor in angiotensin II-induced vascular inflammation and remodeling. Hypertension. 2004;44(3):264–70. Kasal DA, Barhoumi T, Li MW, Yamamoto N, Zdanovich E, Rehman A, et al. T Regulatory Lymphocytes Prevent Aldosterone-Induced Vascular Injury. Hypertension. 2012;59: 324–30. An elegant study showing that adoptive transfer of T regulatory lymphocytes ameliorates vascular injury in mice infused with aldosterone, supporting a role for innate and adaptive immunity in vascular damage associated with activation of the renin angiotensin aldosterone system. Marvar PJ, Thabet SR, Guzik TJ, Lob HE, McCann LA, Weyand C, et al. Central and peripheral mechanisms of T-lymphocyte activation and vascular inflammation produced by angiotensin II-induced hypertension. Circ Res. 2010;107(2):263–70. Hoch NE, Guzik TJ, Chen W, Deans T, Maalouf SA, Gratze P, et al. Regulation of T-cell function by endogenously produced angiotensin II. Am J Physiol Regul Integr Comp Physiol. 2009;296(2):R208–16. Suematsu M. The inflammatory aspect of the microcirculation in hypertension: oxidative stress, leukocytes/endothelial interaction, apoptosis. Microcirculation. 2002;9:259–76. Wu L, Iwai M, Nakagami H. Roles of angiotensin II type 2 receptor stimulation associated with selective angiotensin II type 1 receptor blockade with valsartan in the improvement of inflammation-induced vascular injury. Circulation. 2001;104: 2716–21. Diep QN, Amiri F, Touyz RM, Cohn JS, Endemann D, Neves MF, et al. PPARa activator effects on Ang II-induced vascular oxidative stress and inflammation. Hypertension. 2002;40:866–71. Sigmund CD. Endothelial and vascular muscle PPARgamma in arterial pressure regulation: lessons from genetic interference and deficiency. Hypertension. 2010;55(2):437–44. Tham DM. Angiotensin II, is associated with activation of NFkappaB-mediated genes and downregulation of PPARs. Physiol Genomics. 2002;11:21–30. Schiffrin EL, Lipman ML, Mann JF. Chronic kidney disease: effects on the cardiovascular system. Circulation. 2007;116(1): 85–97. Anand DV, Lim E, Darko D, Bassett P, Hopkins D, Lipkin D, et al. Determinants of progression of coronary artery calcification in type 2 diabetes role of glycemic control and inflammatory/ vascular calcification markers. J Am Coll Cardiol. 2007;50(23): 2218–25. Ng K, Hildreth CM, Avolio AP, Phillips JK. Angiotensinconverting enzyme inhibitor limits pulse-wave velocity and aortic calcification in a rat model of cystic renal disease. Am J Physiol Renal Physiol. 2011;301(5):F959–66.

Curr Hypertens Rep (2014) 16:431 127.

128.

129.

130.

131.

132.

133.

134.•

135.

136.

137.•

138.

139.

140.

141.

Karwowski W, Naumnik B, Szczepański M. Myśliwiec The mechanism of vascular calcification - a systematic review. Med Sci Monit. 2012;18:1–11. Tyson KL, Reynolds JL, McNair R, Zhang Q, Weissberg PL, Shanahan CM. Osteo/chondrocytic transcription factors and their target genes exhibit distinct patterns of expression in human arterial calcification. Arterioscler Thromb Vasc Biol. 2003;23: 489–94. Huybers S, Bindels RJ. Vascular calcification in chronic kidney disease: new developments in drug therapy. Kidney Int. 2007;72: 663–5. Li X, Giachelli CM. Sodium-dependent phosphate cotransporters and vascular calcification. Curr Opin Nephrol Hypertens. 2007;16:325–8. Li X, Yang HY, Giachelli CM. BMP-2 promotes phosphate uptake, phenotypic modulation, and calcification of human vascular smooth muscle cells. Atherosclerosis. 2008;199:271–7. Freedman BI, Bowden DW, Ziegler JT, Langefeld CD, Lehtinen AB, Rudock ME, et al. Bone morphogenetic protein 7 (BMP7) gene polymorphisms are associated with inverse relationships between vascular calcification and BMD: the Diabetes Heart Study. J Bone Miner Res. 2009;24:1719–27. Scatena M, Liaw L, Giachelli CM. Osteopontin: a multifunctional molecule regulating chronic inflammation and vascular disease. Arterioscler Thromb Vasc Biol. 2007;27:2302–9. Osako MK, Nakagami H, Shimamura M, Koriyama H, Nakagami F, Shimizu H, et al. Cross-Talk of receptor activator of nuclear factor-κB ligand signaling with renin–angiotensin system in vascular calcification. Arterioscler Thromb Vasc Biol. 2013;33:1287– 96. In vitro and in vivo studies demonstrate that Ang II induces vascular calcification through pathways that involve receptor activator of nuclear factor-κB ligand (RANKL) and that RANKL itself influences the RAS. These data indicate a circuitous interaction between these systems that amplifies atherogenesis and vascular calcification. Montezano AC, Zimmerman D, Yusuf H, Burger D, Chignalia AZ, Wadhera V, et al. Vascular smooth muscle cell differentiation to an osteogenic phenotype involves TRPM7 modulation by magnesium. Hypertension. 2010;56(3):453–62. He Y, Yao G, Savoia C, Touyz RM. Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: role of angiotensin II. Circ Res. 2005;96(2):207–15. Jia G, Stormont RM, Gangahar DM, Agrawal DK. Role of matrix Gla protein in angiotensin II-induced exacerbation of vascular calcification. Am J Physiol Heart Circ Physiol. 2012;303(5):H523–32. Demonstration that Ang II influences vascular calcification through processes that interfere with the natural calcification inhibitor, matrix Gla protein and activation of the transcription factors, runt-related transcription factor 2 and NF-κB. Sui YB, Chang JR, Chen WJ, Zhao L, Zhang BH, Yu YR, et al. Angiotensin-(1-7) inhibits vascular calcification in rats. Peptides. 2013;42:25–34. Virmani R, Robinowitz M, Geer JC, Breslin PP, Beyer JC, McAllister HA. Coronary artery atherosclerosis revisited in Korean war combat casualties. Arch Pathol Lab Med. 1987;111(10):972–6. Steinbrecher UP, Parthasarathy S, Leake DS. Modification of low density lipoprotein by endothelial cells involves lipid peroxidation and degradation of lowdensity lipoprotein phospholipids. Proc Natl Acad Sci U S A. 1984;81:3883–7. Tabas I, Williams KJ, Borén J. Subendothelial lipoprotein retention as the initiating process in atherosclerosis: update and therapeutic implications. Circulation. 2007;16:1832–44.

Page 11 of 11, 431 142.

143.

144.

145.

146.

147.•

148.•

149.

150.

151.

152.

153.

154.

Nicholson AC, Frieda S, Pearce A, Silverstein RL. Oxidized LDL binds to CD36 on human monocyte-derived macrophages and transfected cell lines. Evidence implicating the lipid moiety of the lipoprotein as the binding site. Arterioscler Thromb Vasc Biol. 1995;15(2):269–75. Nagy L, Tontonoz P, Alvarez JG, Chen H, Evans RM. Oxidized LDL regulates macrophage gene expression through ligand activation of PPARgamma. Cell. 1998;93(2):229–40. Rios FJ, Ferracini M, Pecenin M, Koga MM, Wang Y, Ketelhuth DF, et al. Uptake of oxLDL and IL-10 production by macrophages requires PAFR and CD36 recruitment into the same lipid rafts. PLoS ONE. 2013;8(10):e76893. Rios FJ, Koga MM, Pecenin M, Ferracini M, Gidlund M, Jancar S. Oxidized LDL induces alternative macrophage phenotype through activation of CD36 and PAFR. Mediat Inflamm. 2013;2013:198–3. Libby P, Lichtman AH, Hansson GK. Immune effector mechanisms implicated in atherosclerosis: from mice to humans. Immunity. 2013;38(6):1092–104. Kiyan Y, Tkachuk S, Hilfiker-Kleiner D, Haller H, Fuhrman B, Dumler I. oxLDL induces inflammatory responses in vascular smooth muscle cells via urokinase receptor association with CD36 and TLR4. J Mol Cell Cardiol. 2014;66:72–82. This study elucidates a novel mechanism contributing to vascular inflammation through the urokinase receptor, which responds to endogenous atherogenic ligands, such as oxLDL, impacting on CD36 and TLR4 function. Wang YS, Wang HY, Liao YC, Tsai PC, Chen KC, Cheng HY, et al. MicroRNA-195 regulates vascular smooth muscle cell phenotype and prevents neointimal formation. Cardiovasc Res. 2012;95(4):517–26. This interesting study demonstrates that in vascular smooth muscle cells, microRNA-195 reduces proliferation, migration, and synthesis of IL-1β, IL-6, and IL-8. These findings were extended to a rat model where the microRNA-195 gene was introduced by adenovirus, and showed reduced neointimal formation in balloon-injured carotid arteries. Taken together, these data suggest that microRNA-195 may be vasculo-protective. Hu C, Dandapat A, Mehta JL. Angiotensin II induces capillary formation from endothelial cells via the LOX-1 dependent redoxsensitive pathway. Hypertension. 2007;50:952–7. Cominacini L, Pasini AF, Garbin U, Davoli A, Tosetti ML, Campagnola M, et al. Oxidized low density lipoprotein (oxLDL) binding to ox-LDL receptor-1 in endothelial cells induces the activation of NF-kappaB through an increased production of intracellular reactive oxygen species. J Biol Chem. 2000;275(17): 12633–8. Taye A, Saad AH, Kumar AH, Morawietz H. Effect of apocynin on NADPH oxidase-mediated oxidative stress-LOX-1-eNOS pathway in human endothelial cells exposed to high glucose. Eur J Pharmacol. 2010;627:42–8. Tsai KL, Chen LH, Chiou SH, Chiou GY, Chen YC, Chou HY, et al. Coenzyme Q10 suppresses oxLDL-induced endothelial oxidative injuries by the modulation of LOX-1mediated ROS generation via the AMPK/PKC/NADPH oxidase signaling pathway. Mol Nutr Food Res. 2011;55: S227–40. Hu C, Dandapat A, Chen J, Liu Y, Hermonat PL, Carey RM, et al. Over-expression of angiotensin II type 2 receptor (agtr2) reduces atherogenesis and modulates LOX-1, endothelial nitric oxide synthase and heme-oxygenase-1 expression. Atherosclerosis. 2008;199:284–94. Fukuda D, Enomoto S, Hirata Y, Nagai R, Sata M. The angiotensin receptor blocker, telmisartan, reduces and stabilizes atherosclerosis in ApoE and AT1aR double deficient mice. Biomed Pharmacother. 2010;64(10):712–7.

Angiotensin II and vascular injury.

Vascular injury, characterized by endothelial dysfunction, structural remodelling, inflammation and fibrosis, plays an important role in cardiovascula...
976KB Sizes 1 Downloads 3 Views