the renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

13
REVIEWS: CURRECT TOPICS The reninangiotensin system in adipose tissue and its metabolic consequences during obesity Maria E. Frigolet a , Nimbe Torres b , Armando R. Tovar b, a Mount Sinai Hospital, Toronto, ON, Canada M5G1X b Instituto Nacional de Ciencias Médicas y Nutrición Salvador Zubirán, 1Dept. Fisiología de la Nutrición, México, D.F. 14000, México Received 19 February 2013; received in revised form 24 May 2013; accepted 22 July 2013 Abstract Obesity is a worldwide disease that is accompanied by several metabolic abnormalities such as hypertension, hyperglycemia and dyslipidemia. The accelerated adipose tissue growth and fat cell hypertrophy during the onset of obesity precedes adipocyte dysfunction. One of the features of adipocyte dysfunction is dysregulated adipokine secretion, which leads to an imbalance of pro-inflammatory, pro-atherogenic versus anti-inflammatory, insulin-sensitizing adipokines. The production of reninangiotensin system (RAS) components by adipocytes is exacerbated during obesity, contributing to the systemic RAS and its consequences. Increased adipose tissue RAS has been described in various models of diet-induced obesity (DIO) including fructose and high-fat feeding. Up- regulation of the adipose RAS by DIO promotes inflammation, lipogenesis and reactive oxygen species generation and impairs insulin signaling, all of which worsen the adipose environment. Consequently, the increase of circulating RAS, for which adipose tissue is partially responsible, represents a link between hypertension, insulin resistance in diabetes and inflammation during obesity. However, other nutrients and food components such as soy protein attenuate adipose RAS, decrease adiposity, and improve adipocyte functionality. Here, we review the molecular mechanisms by which adipose RAS modulates systemic RAS and how it is enhanced in obesity, which will explain the simultaneous development of metabolic syndrome alterations. Finally, dietary interventions that prevent obesity and adipocyte dysfunction will maintain normal RAS concentrations and effects, thus preventing metabolic diseases that are associated with RAS enhancement. © 2013 Elsevier Inc. All rights reserved. Keywords: Adipose tissue; Reninangiotensin system; Diet; Adipocyte; Hypertension; Diabetes 1. The classical reninangiotensin system (RAS) The enzymatic function of renin was established in 1898 by Tigerstedt and Bergman. Since then, the RAS has been extensively studied, and various hormones and receptors are now recognized as players in this system [1]. Therefore, multiple effects of the system and its regulation at cellular and systemic levels have been described. Because regulation of the RAS influences several levels, every level should be independently regulated in harmony with other systems to maintain homeostasis. The RAS system produces angiotensin II (Ang II) from angiotensin I (Ang I) and angiotensinogen (AGT) via renin and angiotensin-converting enzyme (ACE). In the classical pathway, AGT is produced mainly by the liver and is drained to the circulation. Subsequently, AGT is converted into the biologically inactive peptide Ang I through the action of renin, which is also released into the bloodstream. Renin is an aspartyl protease that is primarily produced by juxtaglomerular cells in the renal afferent arteriole as preprorenin, which is converted to prorenin and then to active renin [2,3]. This enzyme cleaves AGT to release the N-terminal decapeptide Ang I [4]. Then ACE, a peptidyldipeptide hydrolase located in vascular endothelial cells, separates the C-terminal dipeptide from angioten- sin I to produce the octapeptide Ang II [5]. Additionally, ACE metabolizes the vasodilator bradykinin and inactivates it [6]. For these reasons, ACE has a dual role of increasing vasoconstriction and inactivating vasodilation. The main effector hormone of the system is Ang II, which binds to the Ang II type 1 (AT1), and Ang II type 2 (AT2) G-protein coupled receptors to exert its biological functions. AT1 interacts with multiple heterotrimeric G proteins to produce second messengers such as inositol trisphosphate, diacylglycerol and reactive oxygen species [7]. Although other roles have been described for Ang II, most of the pathophysiological effects of the hormone are induced when bound to AT1 such as vasoconstriction, increased thirst, aldosterone produc- tion, Na+ reabsorption, nervous sympathetic system activation, hypertrophy and fibrosis [4]. AT2 is the predominant fetal isoform, and its expression is low in most tissues. AT2 mediates its actions through protein tyrosine phosphatase activation, nitric oxide gener- ation, and sphingolipid signaling to stimulate vasodilation, natriure- sis, anti-inflammatory and anti-fibrotic actions [8,9]. Thus, AT1 and AT2 receptors have opposite functions in several cell types [10]. The explanation of AT1 and AT2 functions in adipocytes will be described in this review. Available online at www.sciencedirect.com ScienceDirect Journal of Nutritional Biochemistry 24 (2013) 2003 2015 Corresponding author. Fax: +52 5556553038. E-mail address: [email protected] (A.R. Tovar). 0955-2863/$ - see front matter © 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.jnutbio.2013.07.002

Upload: armando-r

Post on 18-Dec-2016

217 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

Available online at www.sciencedirect.com

ScienceDirect

Journal of Nutritional Biochemistry 24 (2013) 2003–2015

REVIEWS: CURRECT TOPICS

The renin–angiotensin system in adipose tissue and its metabolic consequencesduring obesity

Maria E. Frigoleta, Nimbe Torresb, Armando R. Tovarb,⁎

aMount Sinai Hospital, Toronto, ON, Canada M5G1XbInstituto Nacional de Ciencias Médicas y Nutrición Salvador Zubirán, 1Dept. Fisiología de la Nutrición, México, D.F. 14000, México

Received 19 February 2013; received in revised form 24 May 2013; accepted 22 July 2013

Abstract

Obesity is a worldwide disease that is accompanied by several metabolic abnormalities such as hypertension, hyperglycemia and dyslipidemia. Theaccelerated adipose tissue growth and fat cell hypertrophy during the onset of obesity precedes adipocyte dysfunction. One of the features of adipocytedysfunction is dysregulated adipokine secretion, which leads to an imbalance of pro-inflammatory, pro-atherogenic versus anti-inflammatory, insulin-sensitizingadipokines. The production of renin–angiotensin system (RAS) components by adipocytes is exacerbated during obesity, contributing to the systemic RAS and itsconsequences. Increased adipose tissue RAS has been described in various models of diet-induced obesity (DIO) including fructose and high-fat feeding. Up-regulation of the adipose RAS by DIO promotes inflammation, lipogenesis and reactive oxygen species generation and impairs insulin signaling, all of whichworsen the adipose environment. Consequently, the increase of circulating RAS, for which adipose tissue is partially responsible, represents a link betweenhypertension, insulin resistance in diabetes and inflammation during obesity. However, other nutrients and food components such as soy protein attenuateadipose RAS, decrease adiposity, and improve adipocyte functionality. Here, we review the molecular mechanisms by which adipose RAS modulates systemicRAS and how it is enhanced in obesity, which will explain the simultaneous development of metabolic syndrome alterations. Finally, dietary interventionsthat prevent obesity and adipocyte dysfunction will maintain normal RAS concentrations and effects, thus preventing metabolic diseases that are associated withRAS enhancement.© 2013 Elsevier Inc. All rights reserved.

Keywords: Adipose tissue; Renin–angiotensin system; Diet; Adipocyte; Hypertension; Diabetes

1. The classical renin–angiotensin system (RAS)

The enzymatic function of renin was established in 1898 byTigerstedt and Bergman. Since then, the RAS has been extensivelystudied, and various hormones and receptors are now recognized asplayers in this system [1]. Therefore, multiple effects of the systemand its regulation at cellular and systemic levels have been described.Because regulation of the RAS influences several levels, every levelshould be independently regulated in harmony with other systems tomaintain homeostasis. The RAS system produces angiotensin II (AngII) from angiotensin I (Ang I) and angiotensinogen (AGT) via reninand angiotensin-converting enzyme (ACE). In the classical pathway,AGT is produced mainly by the liver and is drained to the circulation.Subsequently, AGT is converted into the biologically inactive peptideAng I through the action of renin, which is also released into thebloodstream. Renin is an aspartyl protease that is primarily producedby juxtaglomerular cells in the renal afferent arteriole as preprorenin,which is converted to prorenin and then to active renin [2,3]. Thisenzyme cleaves AGT to release the N-terminal decapeptide Ang I [4].

⁎ Corresponding author. Fax: +52 5556553038.E-mail address: [email protected] (A.R. Tovar).

0955-2863/$ - see front matter © 2013 Elsevier Inc. All rights reserved.http://dx.doi.org/10.1016/j.jnutbio.2013.07.002

Then ACE, a peptidyldipeptide hydrolase located in vascularendothelial cells, separates the C-terminal dipeptide from angioten-sin I to produce the octapeptide Ang II [5]. Additionally, ACEmetabolizes the vasodilator bradykinin and inactivates it [6]. Forthese reasons, ACE has a dual role of increasing vasoconstriction andinactivating vasodilation.

The main effector hormone of the system is Ang II, which binds tothe Ang II type 1 (AT1), and Ang II type 2 (AT2) G-protein coupledreceptors to exert its biological functions. AT1 interacts with multipleheterotrimeric G proteins to produce second messengers such asinositol trisphosphate, diacylglycerol and reactive oxygen species [7].Although other roles have been described for Ang II, most of thepathophysiological effects of the hormone are inducedwhen bound toAT1 such as vasoconstriction, increased thirst, aldosterone produc-tion, Na+ reabsorption, nervous sympathetic system activation,hypertrophy and fibrosis [4]. AT2 is the predominant fetal isoform,and its expression is low in most tissues. AT2 mediates its actionsthrough protein tyrosine phosphatase activation, nitric oxide gener-ation, and sphingolipid signaling to stimulate vasodilation, natriure-sis, anti-inflammatory and anti-fibrotic actions [8,9]. Thus, AT1 andAT2 receptors have opposite functions in several cell types [10]. Theexplanation of AT1 and AT2 functions in adipocytes will be describedin this review.

Page 2: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2004 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

Other angiotensin peptides hydrolyzed from AGT also havebiological effects. These are angiotensin 2–8 (Ang III), angiotensin3–8 (Ang IV), and angiotensin-(1–7) [Ang (1–7)]. The peptides Ang IIIand IV can be generated from Ang II degradation at the N-terminal byaminopeptidase (A and M), whereas Ang (1–7) is formed from Ang IIthrough ACE2 [11]. Ang (1–7) acts via the Mas receptor (MasR), andAng IV binds to the AT4 receptor, which is associated with insulin-regulated aminopeptidase (IRAP) (Fig. 1).

It is in the endocrine pathway that Ang II is produced in thecirculation and exerts its biological function in the vasculature or indistinct tissues. However, the RAS is significantly more complex. Todeal with such complexity, the RAS can be divided into the endocrineand local systems. The local RAS is defined by Ang II production fromAGT and locally synthesized enzymes [12]. Because AGT is a secretoryprotein, Ang II could be produced inside the cell or in the interstitialspace; consequently, the hormone could be bound to its receptor inneighboring cells, thus acting in an autocrine or paracrine manner[13]. Local Ang II production relies on additional enzymes such astonin, D and G cathepsin and chymase. For example, chymaseconverts 90% of Ang I to Ang II in myocardial extracts [14,15]. Inaddition, cultured skeletal muscle fibers and myoblasts lack reninmRNA, while cathepsin D is expressed during stretching when RAScomponents are up-regulated [16]. Other tissues that have beendescribed to contain the local RAS are the kidney, pancreas, heart,adipose tissue, muscle, liver, adrenal glands and bone marrow[17–20]. Thus, the local RAS in distinct tissues acts via autocrineand paracrine mechanisms to exacerbate the effects of circulatingRAS and/or works independently to induce a response within atissue or cell type.

Fig. 1. The RAS. The classical RAS comprises a series of steps that take place in the circulation. Tpotent enzyme. Ang I is cleaved into Ang II by the ACE. This same enzyme inactivates the vasoblood pressure acting as a vasoconstrictor. The actions of Ang II are mediated by two receptors:peptides derived from Ang II, which include Ang III and Ang IV. These are synthesized by aminoaction of ACE2. The receptor for Ang IV is AT4, which is bound to IRAP in the plasma membbloodstream and tissues, cathepsins and chymase catalyze the production of Ang II locally in vred, and the RAS local enzymes in orange. Angiotensin II and other angiotensin peptides caadipocyte. PRR: prorenin receptor; AT4: angiotensin receptor type IV.

2. The local RAS in adipocytes and its metabolic role

Adipose tissue has recently been described as an endocrineorgan with a great impact on energy metabolism. Thus, adipocytehomeostasis is required to maintain the equilibrium betweennutrient utilization and storage. Adipocyte functions includebalanced hormonal secretion, sensing hormones and other signals,performing lipogenesis and lipolysis at equilibrium, and normaladipocyte growth and development. Other general adipocytefunctions are metabolic processes such as angiogenesis, extracellu-lar matrix reformation, steroid metabolism, immune responses andhemostasis [21].

Additionally, recent findings demonstrate that human and rodentadipose tissues also contain all of the RAS components. AGTmessenger RNA (mRNA) expression has been demonstrated in ratwhite and brown adipose tissue, although greater gene expression isfound in white adipose tissue. Because adipose tissue contains severalcell types aside from adipocytes such as preadipocytes, monocytes,macrophages, vascular stromal cells, nerve cells and fibroblasts, itbecame essential to determine whether adipocytes were responsiblefor adipose tissue AGT or other RAS component production. Severalreports demonstrate that isolated adipocytes from humans [22] androdents [23] as well as the white adipocyte cell line 3T3-L1 allsynthesize AGT [24]. Adipose tissue is a major contributor ofextrahepatic AGT, particularly in obesity [25,26], and AGT mRNAexpression is higher in human visceral adipose depots thansubcutaneous adipose depots [27]. Accordingly, production of AGTin adipose tissue (especially in visceral adipose tissue) in obeseindividuals increases circulating AGT levels as a precursor that can be

he precursor AGT is converted to Ang I through renin, which bound to the PRR is a moredilator bradykinin. Ang II is typically the effector hormone of the system and regulatesAT1 and AT2, which usually have opposing effects. The novel RAS comprises angiotensinpeptidase A andM, respectively. Ang II is also a precursor of Ang(1–7), produced by therane, and the receptor for Ang (1–7) is Mas. Apart from renin and ACE found in thearious tissues. Note that the classical RAS is represented in black text, the novel RAS inn bind their receptors (AT1, AT2, MasR, and AT4) in several cell types including the

Page 3: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2005M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

cleaved by systemic RAS enzymes to produce Ang II and otherangiotensin peptides (Fig. 2).

Other components of the RAS needed for Ang II synthesis andresponsiveness have also been found in adipocytes [24]. For instance,ACE mRNA was present in both subcutaneous and visceral adiposetissue, and in situ hybridization demonstrated gene expression inadipocytes [28]. Although adipocyte renin mRNA expression was notdetectable according to some studies, it has been found ininterscapular brown adipose tissue, and renin activity has beendemonstrated in white and brown adipose tissue as well as 3T3-L1adipocytes [29–31]. However, adipocyte renin activity is relativelylow; therefore, ACE inhibitors do not have direct effects on adipocytes[32]. Determination of low renin activity levels in adipocytes could bedue to decreased renin receptor production because renin and pro-renin bind to their receptors and enhance renin enzymatic activity,which is accompanied by increased Ang II generation [4]. Reninreceptors have been found in human adipose tissue with greaterexpression in visceral compared with subcutaneous adipose tissue[33]. Although renin receptors aremainly located in the adipose tissuestromal vascular fraction, adipocytes also express these receptors[34,35]. Therefore, renin-dependent formation of angiotensin pep-tides is locally modulated by renin receptors in adipocytes and couldpossibly contribute to systemic Ang II concentrations.

As a local system, Ang II production from adipocyte AGT can alsobe achieved through the action of cathepsin isoforms D and G [36,37]and chymase [37], thereby bypassing renin and ACE [36]. Thus,AGT synthesis is a key regulatory step of the RAS in adipose tissuebecause it is present in the vasculature [38]. Additionally, cathepsins

Fig. 2. Nutrients and food components influence the adipocyte local and circulating RAS. HFD,and obesity. AGT and Ang II act locally or are released from the adipocyte into the bloodstreamhypertension and insulin resistance, important components of the metabolic syndrome. In cfunction and preventing metabolic abnormalities.

affect adipocyte biology because they have been associated withpreadipocyte differentiation and adipogenesis and are up-regulatedin obesity [39–41].

Responsiveness to Ang II in adipocytes occurs because of AT1 andAT2 receptor presence. A first study identified Ang receptors as aresult of Ang II binding in rat epidydymal adipocyte membranes [42],and we currently know that rodent and human adipocytes containthese receptors. AT1 receptors have been well described in 3T3-L1and human adipocytes by Western blotting, polymerase chainreaction (PCR), confocal microscopy and binding assays [43–47].Less is known about the biological effects of AT2 receptors, and theexpression of these receptors increases with preadipocyte differen-tiation [47]. Real-time PCR has also confirmed that both rat andhuman adipocytes contain AT2 [48,49]. AT2 is stimulated in certainconditions or when AT1 is blocked because Ang II preferentially actsvia AT2 [50].

Adipocytes and preadipocytes contain the Ang IV and Ang (1–7)receptors, which are AT4-IRAP and MasR, respectively [51]. MasRis a G protein-coupled receptor that is encoded by the Mas proto-oncogene and is related to increased glucose uptake and insulinsensitivity [52–54]. The AT4 receptor associates with membrane-bound IRAP [55]. Interestingly, AT4 could be implicated in insulinsensitivity by binding to angiotensin peptides because IRAPis linked to the glucose transporter glucose transporter 4(GLUT4) [56,57].

With the discovery of the Ang receptor, the autocrine andparacrine activity of the adipocyte RAS was confirmed. Local AGTand Ang II production contribute to circulating concentrations of

sucrose and fructose have shown to induce adipocyte dysfunction, increased local RAS. As a consequence, circulating RAS along with increased leptin and FFA contribute to

ontrast, dietary soy protein and PUFAS decrease adipocyte RAS maintaining adipocyte

Page 4: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2006 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

these peptides and hence participate in vascular function and bloodpressure regulation in humans. However, AGT could also be releasedand used by adipose tissue as a precursor for Ang II and otherangiotensin peptide production, which can affect the adipocytes andother neighboring cells. It needs to be determined whether the localadipose tissue RAS works independently of the circulating RAS.Adipose tissue renin does not correlate with plasma enzymeconcentrations, as has been shown for cardiac renin [58]. However,adipocyte-generated AGT represents approximately 30% of circulat-ing levels in rodents [59–61]. Moreover, adipose tissue RAS isincreased during the development of metabolic abnormalities suchas obesity, and adipose tissue-secreted AGT is increased in obesehumans and correlates with AGT plasma levels in mice [62]. Excessbody fat, particularly visceral fat, is associated with hypertension andother abnormalities; therefore, adipose tissue RAS could contribute tosystemic RAS in pathophysiological but not regular conditions.Likewise, nutrients can modulate RAS component expression andactivity in adipocytes, adding to whole body RAS and its conse-quences. In further studies, proper techniques should be used todissect local from systemic RAS in physiological, pathological anddiverse nutritional conditions to demonstrate the importance of eachtissue in the total system (Fig. 2).

Conversely, adipose tissue-derived RAS impacts not only circula-tion but also neighboring tissues. Adipocytes within the perivascularadipose tissue secrete RAS components and produce adipokines inresponse to locally generated Ang II, all of which influence vascularfunction [63–65]. The perivascular tissue in aorta and mesentericarteries is mainly brown and white adipose tissue, respectively.Interestingly, Ang II levels were higher in rat mesenteric than in aorticperivascular tissue [66], suggesting differences in local brown andwhite adipose tissue RAS as well as different expression levels of RASmolecules in distinct anatomical regions. In this case, crosstalkbetween adipocytes and vascular cells could explain the mechanisticlink between obesity, insulin resistance, endothelial dysfunction, andhypertension. Furthermore, visceral adipose tissue is the main playerin the development of obesity-associated metabolic abnormalities, incontrast to subcutaneous adipose tissue. Human visceral adiposetissue consistently exhibits significantly elevated AGT mRNA com-pared with subcutaneous adipose tissue [67]. The latter fact explainswhy visceral adiposity is clinically related to hypertension and insulinresistance [68,69].

Although the concept of adipocyte dysfunction is still beingdebated, scientists have thoroughly explored adipocyte physiology inrecent years. In obesity, adipocyte dysfunction has been associatedwith increased pro-atherogenic, pro-diabetic and pro-inflammatoryhormone secretion, which is accompanied by reduced adiponectinrelease [70]. Resistance to hormones such as insulin and leptininfluence adipocyte development. The inability to recruit preadipo-cytes when more mature adipocytes are needed to buffer excess fuelcauses adipocyte hyperthrophy. Larger fat cells release more fattyacids, which in turn increases macrophages in adipose tissue andworsens adipose tissue inflammation [71]. Consequences of adipocytedysfunction comprise obesity-associated abnormalities such asdiabetes, hypertension, cardiovascular diseases, sexual dysfunctionand cancers.

In addition, because adipocyte dysfunction includes imbalancedhormone secretion, increased AGT and other RAS componentproduction could be considered to be a feature of impaired adipocytefunction and is associated with several metabolic abnormalities suchas insulin resistance and hypertension [72]. In this review, we willdiscuss how RAS is associated with glucose and lipid metabolism inthe adipocytes, its involvement in adipose tissue inflammation andthe relationship between the exacerbation of local adipocyte RAS andconcomitant metabolic consequences during obesity as well as itsregulation by nutrients.

3. RAS and glucose metabolism in adipocytes

Increased RAS activity has been associated with insulin resistanceand decreased glucose uptake in various tissues, including adiposetissue. Ang II infusion in the subcutaneous abdominal adipose tissueof healthymen decreased glucose uptake in a dose-dependent fashion[73]. These Ang II effects are possibly mediated by receptor AT1because AT1a-deficient mice maintained insulin sensitivity andincreased energy expenditure after being fed a high-fat diet (HFD)[74]. Ang II receptor blockers (ARB) such as candesartan increaseadipose tissue and skeletal muscle glucose uptake in type 2 diabeticKK-A(y) mice [75]. Additionally, whole body and adipose tissueinsulin resistance in obese Zucker rats is improved by acute andchronic AT1 antagonism with irbesartan [76,77]. Modulation of otherRAS components has also been associated with insulin sensitivity. Forinstance, pharmacological or genetic ACE inhibition improves insulinaction on peripheral glucose disposal in animals [78,79] and clinicalinvestigations [80–84]. In addition, renin inhibition by aliskirenincreases glucose tolerance and whole body insulin sensitivity inobese Zucker rats and Ren2 rats, which manifest increased RASactivity [85,86].

Several mechanisms have been addressed to explain the effects ofAng II on insulin resistance; cross talk between Ang II and insulinsignaling is one of these mechanisms. In responsive tissues, Ang IIinduces serine phosphorylation of proteins that are implicated in theinsulin signaling pathway such as the insulin receptor (IR) and insulinreceptor substrate (IRS-1) and reduces insulin-mediated interactionsbetween IRS-1 and the regulatory p85 subunit of phosphatidylinositol3-kinase [87,88]. While tyrosine phosphorylation of these insulinsignaling proteins is necessary for insulin-mediated effects, serinephosphorylation impedes tyrosine phosphorylation of these mole-cules and therefore decreases insulin action [89,90]. Insulin signalingis also impaired by Ang II signaling through its AT1 receptor via JAK/STAT (Janus kinase/signal transducer and activator of transcription)pathway activation, which induces suppressor of cytokine signaling-3(SOCS-3). Also, c-Jun N-terminal kinase (JNK), a negative feedbackregulator of insulin signaling at the level of IRS-1 [91], is up-regulatedby Ang II [92] and increases SOCS-3 [93]. SOCS-3 can bind to tyrosineswithin the IR and hinder its ability to induce IRS isoform tyrosinephosphorylation [94–96]. Moreover, SOCS-3 binds directly to IRS anddrives its proteosomal degradation in an ubiquitin-dependentmechanism [97,98]. Subsequently, Ang II-mediated impairment ofpost-receptor insulin signaling via insulin signaling protein serinephosphorylation and/or SOCS activation decreases GLUT4 transloca-tion to the plasma membrane (Fig. 3).

Ang II also increases aldosterone secretion, which is anotherpromoter of insulin resistance that is secreted in response toadipocyte-derived compounds including aldosterone stimulatingfactors [58]. More recent evidence indicates that aldosterone causesoxidative stress in cultured adipocytes, contributing to adipose tissuedysfunction [99] and systemic insulin resistance.

As previously mentioned, ACE inhibition increases insulin sensi-tivity, which is particularly relevant in adipocytes, where ACEinhibition causes bradykinin accumulation [100,101]. In turn, brady-kinin increases GLUT4 translocation to the plasma membrane incanine adipocytes [102]. Other reports demonstrate that bradykininincreases glucose uptake in rat adipocytes via endothelial nitric oxidesynthase-mediated JNK inhibition, which is a negative regulator ofinsulin signaling [103]. Therefore, ACE inhibition not only decreasesangiotensin peptide production but also increases bradykinin pres-ence in adipocytes and therefore reduces insulin resistance.

Moreover, the insulin-sensitizing nuclear receptor peroxisomeproliferator activated receptor gamma (PPAR-g) is activated bytelmisartan in adipocytes [104], which is an ARB and PPAR-g agonist[105,106]. PPAR-g enhances insulin action by increasing the secretion

Page 5: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

Fig. 3. The adipocyte RAS and insulin signaling. Circulating or locally produced Ang II via its AT1 promotes the activation of JNK through increased proinflammatory cytokines, oxidativestress, or Bk cleavage by ACE. Signaling by the JAK/STAT pathway and JNK induction leads to increased SOCS-3 mediated degradation of IRS-1. Also, JNK can directly phosphorylate IRS-1 in serine residues inhibiting tyrosine phosphorylation of IRS-1 and impairing insulin signaling and GLUT4 translocation to the plasma membrane. The antagonism of AT1 improvesinsulin signaling and glucose uptake while increasing PPAR-gamma activity and adiponectin transcription.

2007M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

of the anti-inflammatory adipokine adiponectin and preadipocytedifferentiation, which increases lipid storage capacity by preventinglipid overload in non-adipose tissues such as skeletal muscle,pancreas and liver [107,108].

Other mechanisms by which Ang II and angiotensin peptidesinfluence insulin action have been recently described. Ang II increasesinflammatory markers [109,110], which have been associated withinsulin signaling impairment. In preadipocytes, Ang II is an additionalstimulus for producingmonocyte chemoattractant protein 1 (MCP-1),which is induced by the proinflammatory cytokine tumor necrosisfactor alpha (TNF-α) through extracellular signal-regulated kinase(ERK) and p38 MAPK (p38 mitogen activated protein kinase)-dependent pathways [111]. Ang II alters adipokine secretion, andAT1 blockers have favorable effects in terms of improving adiponectinproduction in diet-induced obese (DIO) mouse adipose tissue [112].These blockers also increase apelin levels in adipocyte culture, whichis another beneficial adipokine [113]. In contrast with Ang II, Ang(1–7) acts through MasR to prevent diabetes-induced cardiovascu-lar disease [114], and chronic elevation of plasma Ang (1–7) levelsimproves insulin sensitivity and increases adipocyte glucose uptake[115]. It is interesting that angiotensin peptides have divergenteffects; this could even explain some observations related Ang II toinsulin sensitivity versus insulin resistance in adipocytes. Onereport demonstrated that acute Ang II treatment increasedinsulin-stimulated glucose uptake in adipocytes [48] possiblybecause of increased production of Ang (1–7) or other angiotensinpeptides that utilize Ang II as a substrate and have insulin-sensitizing actions. Also, the time of exposure to Ang II could

explain the contrasting effects of the hormone, since AT1 binding ismodified by time and various stimuli. However, whether Ang (1–7)production from Ang II or exposure time are explanation fordifferent conclusions will need to be assessed in the future. Insummary, Ang II alters glucose metabolism and is implicated intype 2 diabetes-associated insulin resistance via distinctmechanisms.

4. RAS, adipogenesis, and lipid metabolism in adipocytes

Adipogenesis refers to the generation of mature adipocytes frompreadipocytes. This process occurs mainly during the late embryonicstage, although adipogenic capacity is maintained throughout thelifetime in all animal species [116]. AGT, renin, cathepsin D, ACE, Ang IIand AT1 expression are all increased during adipogenesis [22,117]. AngII stimulates adipocyte differentiation, which releases arachidonic acidfrom plasmamembrane phospholipids. Arachidonic acid is used by thecyclooxygenase enzyme to synthesize prostacyclin (prostaglandin I2).The latter is a potent and specific paracrine/autocrine effector ofpreadipocyte differentiation, which also reverses Ang II-mediatedvasoconstriction in adipose tissue [118]. In vivo, Ang II-inducedprostacyclin liberation has been demonstrated by microdialysis inepididymal fat pads, which was attributed to adipocytes [119,120]. Inkeeping with this finding, another study demonstrated that Ang IIinduces the adipogenic marker glycerol-3-phosphate dehydrogenase(GPDH) expression. Expression of this marker is abolished by acyclooxygenase inhibitor, indicating the role of prostacyclin in Ang II-induced adipogenesis [121]. This mechanism appears to be dependent

Page 6: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2008 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

on PPAR-g because prostacyclin and its stable analogue carbaprostacy-clin bind to this nuclear receptor to regulate fatty acid binding proteinand uncoupling protein-2 expression. However, recent evidencesuggests that rather than altering PPAR-g expression, Ang II regulatesadipogenesis through increased ERK1/2-mediated PPAR-g phosphory-lation in human preadipocytes [122]. Additionally, prostacyclin up-regulates the expression of transcription factors that are involved inadipocyte differentiation, such as C/EBP and , via its cell surfacereceptor [123–125].

In co-culture experiments, evidence suggests that Ang II promotespreadipocyte differentiation by AT2 receptor binding and thuscontrolling prostacyclin-mediated adipogenesis [126]. Moreover, ina mouse model of atherosclerosis, AT2 deficiency was associated withdecreased adipocyte number as well as C/EBP and PPAR-g expres-sion [127]. In contrast, AT1 stimulation by Ang II reduced adipocyteconversion, whereas the ARBs telmisartan and irbestartan inducedadipogenesis and simultaneously activated PPAR-g target genes [22].These results indicate that Ang II is involved in adipocyte differen-tiation through AT2 activation but not through AT1 stimulation inadipose tissue. Therefore, the reported controversies regarding theadipogenic or anti-adipogenic [122,128] effects of Ang II in pre-adipocytes could be because of divergent expression of the two Ang IIreceptor isoforms in rodents versus humans, receptor content indistinct fat pads or variable hormone affinity to these receptors incertain conditions. Another possible mechanism explaining theobserved anti-adipogenic effect of Ang II is down-regulation ofinsulin signaling, whichmediates cell proliferation apart from glucosehomeostasis [129]. Finally, because AT2 expression is regulated bygrowth hormones, use of different conditioned media could alsoaccount for the discrepancies among other in vitro studies.

Ang II also influences lipid metabolism and promotes lipogenesis,which is the production of fatty acids and its esterification intotriglycerides. Ang II exposure increases fatty acid synthase (FAS) andGPDH expression and triglyceride accumulation in adipocytes inrodents and humans [46,130]. In addition, perfusion of subcutaneousadipose tissue with Ang II by microdialysis inhibited lipolysis inhealthy volunteers [73]. This antilipolytic effect could be partlydependent on the vasoconstricting ability of Ang II [129]. Ang II-mediated increase in lipogenesis and decrease in lipolysis contributesto adipocyte enlargement. Generation of transgenic mice in whichadipose AGT is overexpressed or restricted to adipose tissuedemonstrated that targeted AGT expression in adipose tissue in-creases body weight, fat mass, FAS activity and fat cell hypertrophycompared with controls. Hypertrophy was accompanied by hypopla-sia because the cell number of adipocytes was reduced in transgenicmice [61]. Thus, adipose-specific AGT as an Ang II precursor influenceslipid metabolism and subsequently increases adipocyte size.

Thus, Ang II increases lipogenesis, adipocyte hypertrophy, andadipocyte differentiation through its distinct receptors AT1 and AT2.While AT1 decreases adipocyte differentiation [131,132], AT2induces adipocyte maturation and adipose tissue hyperplasia.Preadipocyte conversion into mature adipocytes has been associatedwith increased insulin sensitivity because of efficient lipid storage inadipose tissue. Conversely, adipocyte enlargement precedes macro-phage infiltration and adipose tissue inflammation, which impairsinsulin action. Thus, the effects of Ang II on lipid metabolism andadipocyte morphology via AT2 oppose those of AT1, which affectsglucose metabolism. Moreover, AT1 expression in both visceral andsubcutaneous adipose tissue of obese individuals was found to besignificantly increased, although there were no changes in AT1density in adipose tissues of obese Zucker rats versus lean controls[133]. Therefore, the balance between these two receptors variesamong species and during the development of metabolic abnormal-ities and therefore determines adipose tissue developmental fateand homeostasis.

In summary, the exacerbated secretion of Ang II in obesity,contributes to increased fat mass because of fatty acid accumulationand reesterification with inhibition of lipolysis.

5. Angiotensin and adipocyte inflammation

Adipose tissue undergoes inflammation during obesity onset.Recent studies demonstrated increased angiogenic factor and proin-flammatory cytokine production in murine adipose tissue over-expressing adipose AGT [134,135]. Furthermore, silencing of AGT incultured adipocytes decreases the expression of pro-inflammatorymarkers such as interleukin-6, TNF-α, and MCP-1 [136]. Ang IIinduces proinflammatory adipokine release from both pre- andmature adipocytes and possibly from adipose tissue stromal vascularcells through nuclear factor-kappa B (NF-κB) pathway activation[137–139]. The role of AT1 and AT2 receptors on the inflammatoryprocess is still controversial. Telmisartan administration, whichantagonizes AT1, was found to reduce TNF-α expression andmodulate adipose tissue macrophage polarization to an anti-inflam-matory state in high fat-fed mice [140]. The ARB olmesartan alsoreduced proinflammatory cytokine expression in obese KK-Ay mouseadipose tissue [112]. In isolated adipocytes, valsartan decreasedinflammatory signals in mice that were fed a Western (high fat) diet[141]. This evidence suggests that signaling through AT1mediates theinflammatory consequences of Ang II in adipose tissue and adipo-cytes. However, AT2, which regularly exerts contrasting effects toAT1, has also been implicated in adipose tissue inflammation whenAGT is overexpressed [135]. Further studies should focus on clarifyingwhether Ang II signals via AT1 or AT2 or other receptors such as toll-like receptors to promote inflammation.

Inflammatory reactions induce reactive oxygen species, and thereverse is also true [142]. Ang II increases oxidative stress, which isregulated by AT1 receptor activation [143]. Accordingly, data suggesta beneficial effect of ARBs on oxidative stress. In transgenic ratsexpressing human renin and AGT genes as well as in hypertensivesubjects, oxidative stress was normalized by valsartan treatment[144,145]. The production of reactive oxygen species (ROS) and theexpression of nicotinamide adenine dinucleotide phosphate dehy-drogenase oxidase (NADPH ox) subunits from adipose tissue werereduced by olmesartan treatment in rodents. Simultaneously, incultured adipocytes, the ARB olmesartan acted as an antioxidant andimproved adipokine dysregulation [112].

Inflammation causes glucose intolerance, alters lipid metabolismand contributes to the development of metabolic syndrome. Both AngII-mediated increases in inflammation and ROS production play a rolein insulin resistance and hypertension. Therefore, enhancement ofadipose RAS increases the numerous causes of obesity-associatedmetabolic abnormalities through up-regulating inflammation.

6. Nutrients that modulate the adipose tissue RAS

A new era in nutrition science has evolved in which theinteractions among metabolism, the genome and nutrients are usedto develop new therapeutics for prevention and treatment of chronicdiseases. Nutritional status, nutrient distribution, and nutrients per semodulate the RAS expression and activity in various tissues includingadipose tissue. Fasting and feeding affect AGT production. AGTexpression in epididymal white adipose tissue is regulated by foodintake; AGT decreases during fasting and increases upon re-feeding[24]. Modifying normal nutrient distribution with a HFD activates theadipose and systemic RAS components in humans and rodents, andtargeted inhibition of the RAS protects against DIO in animals [38].Finally, nutrients or food components, such as fructose, lipids, and soyprotein, among others, influence tissue and systemic RAS activity. The

Page 7: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2009M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

next section will describe the mechanisms by which the RAS isaffected by these nutrients.

6.1. Fructose

Along with an increase in total energy, there has been a shift in thetypes of nutrients that are consumed in the Western diet. Fructoseconsumption has increased significantly, largely because of the softdrinks, baked goods, condiments and desserts that are prepared withhigh-fructose corn syrup [146]. Dietary fructose treatment results inhypertension, glucose intolerance and hypertriglyceridemia in ani-mals [146]. This is because fructose feeding exacerbates presence andactivity of the RAS. In rats, the development of hypertension andmetabolic syndrome after 8 weeks of treatment with 60% dietaryfructose was prevented by renin inhibition with aliskiren treatment[147], thus suggesting a role for Ang II production in fructose-inducedmetabolic abnormalities. In a preliminary study, the nonproteolyticactivation of prorenin, which increases Ang II synthesis andcontributes to insulin resistance, was increased in skeletal muscleand adipose tissue of fructose-fed rats [148]. Dietary fructose alsoincreased AT1 expression and density as well as ACE mRNAabundance in the vasculature and cardiac AGT expression [149–151]. Interestingly, NADPH ox activity and contraction effects Ang II inthe aorta were enhanced in fructose-fed rats compared with controls.These responses to fructose feeding were normalized by administra-tion of the ARB losartan, implicating participation of AT1 receptor[152]. AT1 receptor involvement in vascular dysfunction and insulinresistance was confirmed by a study that utilized gene therapy todemonstrate that fructose consumption-induced hypertension andglucose intolerance were prevented by a single neonatal treatmentwith AT1 antisense DNA [153]. In contrast to Ang II actions throughAT1, the effects of Ang (1–7) are beneficial as the peptide reverseshepatic, skeletal muscle and adipose tissue insulin resistance [154]after high-fructose diet consumption.

In adipose tissue, dietary fructose also increases RAS activation.Sprague–Dawley (SD) rats fed a fructose-enriched diet had increasedAT1 but not AGT expression in adipose tissue [155] along withaugmented fat mass and adipocyte size in epididymal fat pads [156].Both temocapril, an ACE inhibitor, and olmesartan treatment restoredadipocyte size and improved insulin signaling. These findings suggestthat adipose RAS plays a role in fructose-induced metabolicabnormalities apart from systemic RAS.

Other sugars such as sucrose also represent a model of DIO;increases in fat pad weight and adipocyte diameter have beenobserved after sucrose consumption [157]. Sucrose-enriched dietsincrease blood pressure (BP) through sympathetic nervous systemactivation and by increasing renin activity and Ang II plasmaconcentrations. AT1 and ACE inhibition decreased the high-sucrosediet-induced BP elevation. Therefore, carbohydrates modify vascu-lar function through different mechanism including activation ofthe RAS.

6.2. Soy protein

The effects of dietary protein have not been as extensivelyexplored as the effects of lipids and carbohydrates. However, ourgroup has demonstrated that a soy protein-enriched diet exertsbeneficial actions in preventing metabolic abnormalities such aswhole body insulin resistance, impaired pancreatic insulin secretion[158], hepatic and skeletal muscle lipotoxicity [159,160] and adiposetissue dysfunction [161] in a rat model of DIO. In addition, evidenceindicates that soy protein prevents metabolic diseases such ashypertension because of the high arginine content and antioxidantactivity that is exhibited by isoflavones [162].

In retroperitoneal adipose tissue (rAT), dietary soy proteinsignificantly reduced AT1 and cathepsin expression with simulta-neous decreases in adipocyte area and in circulating concentrations ofleptin and insulin. Furthermore, isolated adipocytes from rAT releasedlower Ang II concentrations and were more sensitive to insulin-stimulated Akt activation [163].

The mechanism by which soy protein decreases BP and Ang II-derived functions could be explained by the fact that soy protein-derived peptides and amino acids can alter RAS activity. A soy proteinisolate digest produced in vitro by sequential digestion with pepsinand pancreatin effectively suppresses ACE. Peptides from the latterdigestion process with lower molecular masses and higher hydro-phobicities were more potent ACE inhibitors [164]. In vivo studiesthat have investigated the ingestion of soy protein hydrolysate areconsistent with these findings. Inflammation and ACE activity in thehearts of rats that were fed with the hydrolysate were significantlylower than in control animals [165].

Human and animal studies have documented that specific aminoacids such as cysteine, glutathione, glutamate and arginine attenuatehypertension-associated alterations including insulin resistance,altered RAS functions, and oxidative stress [162]. The isoflavonegenistein, which is also present in soy protein isolates, inhibits ACEexpression in serum and the vasculature [166]. These resultsindicate that soy protein gastrointestinal digestion-derived aminoacids, peptides and isoflavones explain modulation of the systemicand tissue RAS and its beneficial consequences to amelioratemetabolic syndrome.

6.3. Lipids

Dietary fat content and consumption of specific lipids modulatethe RAS components. High-fat feeding has been broadly used toinvestigate obesity-induced metabolic consequences as well aspharmacological, genetic, and dietary factors that potentially improvethese consequences. The relationship between HFD and RAS has beenestablished, and certain RAS components that increase blood pressureare increased with HFD consumption.

AGT-deficient rodents gain less weight and AT1 knockout (KO)mice have less body fat than their littermates after a HFD [167,168].HFD ingestion increases BP in conjunction with AGT mRNAabundance in rAT [169,170]. Along with increased BP, a HFDaugments circulating free fatty acids (FFA) and triglycerides[171,172]. Elevations in BP occur subsequent to obesity in HFD-fedrats. The ARB losartan reverses the HFD-induced increase in BP,supporting a role of RAS activation in obesity-associated hypertension[173]. Moreover, aliskiren, irbesartan and captopril, which are renin,AT1 and ACE inhibitors, respectively, reversed HFD-induced weightgain and increased adiposity. The mechanism by which the RASblockade opposes the HFD-mediated increase in adiposity includeslower leptin and higher adiponectin secretion from adipose tissue[174–176]. Other RAS components that are also modified by high-fat feeding include the pro-renin receptor. Post-natal HFD in ratsresults in increases in the number of adipose tissue pro-renin receptorexpressing cells. In addition, a positive relationship was determinedbetween plasma renin activity and pro-renin receptor positivecell density [35]. Taken together, these data suggest that theadipose RAS is exacerbated during HFD, which affects systemicRAS, thus increasing the risk of developing hypertension and relatedco-morbidities.

In non-adipose tissues, a HFD also enhances the RAS, whichcontributes to metabolic alterations. Myocardial AT1, renin, andNF-κB expression were increased but AT2 expression was reduced inHFD-fed rats [171,172]. Simultaneously, inflammatory markers areinduced in the kidneys, pancreas, aorta and liver by a HFD. RASinhibition attenuated this inflammatory process and improved cell

Page 8: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2010 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

function in normal rats or models of spontaneous hypertension andstreptozotocin-induced diabetes [177–179]. Angiotensin peptidessuch as Ang (1–7) and Ang IV have contrasting effects with thoseof Ang II regarding insulin sensitivity and ROS generation [180]. AHFD and Ang II infusion [181] impair insulin sensitivity more severelyin ACE2 KOmice than inWTmice, whichwas eradicated by Ang (1–7)infusion [182]. Thus, ACE2, which is the enzyme that cleavesAng II into Ang (1–7), protects against HFD-induced insulin resistancein mice.

In addition to the content of fat in the diet, fatty acids can also alterBP and RAS activity. Increased circulating FFAs as a consequence of aHFD increase vascular sympathetic tone and BP through increasedadrenergic stimulation, oxidative stress, and endothelial dysfunction[183]. However, dietary polyunsaturated fatty acids (PUFAs) havebeneficial effects on BP and other pathologies. Gamma-linolenic acid,an omega-6 PUFA, attenuates the development of hypertension inspontaneously hypertensive rats [184]. This effect was accompaniedby lower plasma aldosterone levels and decreased adrenal AT1density and affinity [185]. An omega-3 PUFA-sufficient diet versus anomega-3-deficient diet in Ren-2 rats reduced hypertension bymodulating the RAS [186]. An omega-3 PUFA-enriched post-nataldiet prevented hypertension partly by regulating the intrarenal RAS inrats [187]. Interestingly, Zucker fa/fa rats fed with the conjugatedlinoleic acid (CLA) trans-10, cis-12 maintained decreased adipocytesize, higher circulating adiponectin levels and lower BP. In this model,AGT expression was higher in larger adipocytes, possibly resulting indecreased AGT expression in adipose tissue of obese rats that were fedwith this CLA [188]. Finally, human studies demonstrate thatconsuming 3 g PUFA/day enhances the favorable effects of blockingthe RAS on kidney injury [189].

Other nutritional factors that modulate the RAS include mono-sodium glutamate (MG), sodium, and food-derived peptides. It hasbeen hypothesized that Katsuobushi oligopeptides from the bonitofish (Sarda sarda), which inhibit ACE, could prevent hypertension andinsulin resistance by boosting Bk action on adipocytes [100]. It is notclear whether MG-DIO regulates AT1 activity and signaling becauseMG elevated AT1 protein levels but reduced Ang II affinity to theadipocyte plasma membrane [190]. Nevertheless, high sodium intaketogether with dietary fructose increases BP and AT1 receptormRNA inrat adipose tissue [155]. High sodium intake also increases renal Ang IIand reduces ACE2 and AT2 expression in obese Zucker rats [191]. Also,supplementation with niacin-bound chromium at adequate dosesincreases insulin sensitivity while lowering the RAS activity throughACE inhibition [192]. In the future, research should focus oninvestigating how several foods or nutrients modify BP. For instance,magnesium, vitamin C, milk, lactic acid bacteria, tea, among others aresources with antihypertensive effects [193]. Many of these com-pounds could reduce BP through the modulation of the adipose orother local RAS and therefore, effort should be put on clarifyingsuch mechanisms.

In brief, carbohydrates, fatty acids, protein content and type, aswell as other food components affect whole body and tissue RAScomponents and activity. Therefore, dietary strategies that benefi-cially modulate the RAS are important tools to maintain vascular andtissue homeostasis for preventing obesity-associated hypertensionand other related abnormalities.

7. Molecular links between obesity, hypertension and diabetes:The role of adipose RAS

The obesity epidemic is related to hypertension and metabolicsyndrome in humans [194]. There is evidence that adipose RAS isimportant to associate RAS with hypertension and insulin resistance[195]. Evidence demonstrates that genes associated with the RAS areoverexpressed in visceral adipose tissue of overweight human

subjects, and ACE gene polymorphisms have been connected toobesity incidence [67,196,197]. A study in postmenopausal womendemonstrated that obese women had higher circulating AGT, renin,aldosterone and ACE levels than lean women [198]. Furthermore,studies in rats and humans have documented weight loss with ACEinhibitor administration, suggesting a role for Ang II in weight gain[199,200]. Visceral adipose tissue, which is a main player in metabolicsyndrome, has increased AGT expression compared with subcutane-ous adipose tissue [27]. Therefore, human obesity is a condition inwhich components of the RAS are exacerbated.

Moreover, the high blood AGT levels and hypertension that areobserved in obese patients may be because of increased fat mass [61].In fact, it has recently been demonstrated that adipocyte AGT plays acentral role in the development of hypertension, since mice deficientin adipocyte AGT fed a high fat diet did not develop high BP comparedto wild-type rats fed the same diet [201]. Consequently, systemic andtissue RAS can cause increased renal sodium reabsorption andchanges in pressure natriuresis [202]. Plasma aldosterone levels arehigher in obese subjects compared with lean subjects, and studiesinvolving humans and animals demonstrate that ACE inhibitors offeradvantages to control hypertension in obesity [203,204]. While therole of the RAS in hypertension and obesity has been established, theRAS is also implicated in impaired glucose homeostasis. Subjects onACE inhibitors or ARBs have a lower risk of developing type 2 diabetescompared with individuals on other anti-hypertensive medications[205]. Improvements in insulin sensitivity and glucose metabolismhave also been demonstrated in rats and humans that are treatedwithRAS antagonists [206,207].

Therefore, there is enough clinical evidence describing theassociation between systemic RAS, hypertension and insulin resis-tance. RAS is exacerbated during obesity, thus implicating the localadipose RAS as a molecular link among metabolic syndromecomponents. In genetic obesity or DIO, adipocytes have increasedRAS activity, which is not observed in the liver [208,209]. In rodents,adipose AGT might represent 60% of hepatic levels and couldcontribute to 30% of total circulating AGT levels [210]. Thedysregulated AGT and Ang II production is actually a featureof “dysfunctional adipose tissue”, which has also been called “sickfat” [72]. The autocrine and paracrine actions of adipose Ang IIworsen adipocyte dysfunction, leading to increased secretion ofadipokines as leptin and resistin and decreased production of theanti-inflammatory adipokine adiponectin. Ang II also induces pro-thrombotic factor plasminogen activator inhibitor-1 secretion byhuman adipocytes [211].

Ang II promotes leptin production in human adipocytes throughan AT1 and ERK1/2- dependent pathway [212] and in murineadipocytes via a prostaglandin-independent mechanism [213].Chronic hyperleptinemia promoted vasoconstriction and augmentedBP by increasing sympathetic nervous system activity [214,215]. Inobese mice, leptin infusion increases aldosterone and BP in theabsence of weight loss, indicating that despite leptin resistance,leptin-mediated vascular stimulation persists in obesity [216].Moreover, leptin has promotes ROS generation, which mightcontribute to hypertension [217]. In contrast with leptin, hypoadipo-nectinemia occurs concurrently with obesity and is an independentrisk factor for hypertension [218–220]. ARBs induce adiponectinproduction via PPAR-g activation [221], suggesting that the RAS isinvolved in reducing adiponectin concentration. Nevertheless, adipo-nectin can also improve BP by increasing nitric oxide generation andbioavailability [222,223].

FFA delivery from adipose tissue to the liver stimulates vascularsympathetic tone and increases BP [224]. Increased circulating FFAconcentrations are found in obesity, which have been associated withimpaired insulin signaling [225]. FFA induce AGT transcriptionalactivation [226]. Thus, FFA from accelerated lipolysis promotes BP

Page 9: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2011M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

elevation through increasing the RAS, although other mechanismshave been proposed. These include alpha (1)-adrenergic stimulation,endothelial dysfunction, and stimulating vascular smooth muscle cellgrowth and remodeling [227].

Glucocorticoids are not adipokines; however, 11 beta-hydroxys-teroid dehydrogenase type 1 (HSD1), which is the enzyme thatgenerates active cortisol from cortisone, is produced by omental fat[228] and other tissues. Recent studies revealed a prolonged half-lifeof cortisol in some patients with essential hypertension [229]. Micewith HSD1 overexpression are hypertensive, with apparent activationof the circulating RAS. These mice develop visceral obesity, insulinresistance, and dyslipidemia [230].

Taken together, overactivation of the adipose RAS is a keymediator of metabolic complications. Future studies on adipose-specific AGT, AT1 and AT2 animal models are needed to dissect theexact role of the adipose RAS on obesity and associated diseasepathogenesis. However, existing evidence indicates that adipocytedysfunction plays an important role. The increased adipose tissue RASfunction and its consequences on lipid and glucose metabolism,inflammation and impaired insulin signaling are evidence of therelationship between clinical hypertension and insulin resistance inobesity. In addition, adipocyte hypertrophy, pro-inflammatory adi-pokine over-secretion and decreased adiponectin production are allfeatures of adipocyte dysfunction, but are also consequences ofexacerbated adipose RAS, which adds to the already increasedadipose and systemic RAS.

Therefore, adipocyte dysfunction could be an early cause of theexacerbated RAS and/or could be a source that maintains and furtherincreases RAS components along with other factors to establishmetabolic syndrome.

8. Conclusions

The RAS is typically known for its biological actions on thecardiovascular system and kidney to promote vasoconstriction andincrease BP. The presence of a functional adipose tissue RAS is nowwell described. Over activation of the adipose RAS modulatesincreases in body weight and fat mass not only in animal studies,but also in clinical trials. When acting in an autocrine and paracrinefashion, this system promotes adipocyte triglyceride storage, adipo-cyte hypertrophy, and triggers pro-inflammatory adipokine secretion.Both AGT and Ang II are adipokines that are released in highconcentrations into the bloodstream during obesity, which signifi-cantly contribute to systemic levels. Thus, adipose RAS exertsendocrine effects that are important in blood pressure regulation,glucose homeostasis and energy balance. Interestingly, exacerbationof the RAS could be a cause, but also an effect of adipocytedysfunction. Therefore, such dysfunction represents an importantintersection linking metabolic disorders.

There is evidence of nutritional modulation of the adipose andsystemic RAS.While fructose and high fat-induced obesity exacerbatethe adipose and systemic RAS as well as the local RAS in other tissues,other food components such as soy protein and PUFAs attenuate RASpeptide expression in adipose tissue. Further research is essential toimprove our current knowledge on how nutrients modulate theadipose RAS. Using animal models with adipose-specific AGT and RAScomponent knockdown or overexpression in future investigationsshould provide evidence related to the interaction between nutri-tional regulation of the adipose RAS and its contribution to thecirculating RAS. Whole body consequences of the influence ofnutrition on adipose RAS will allow us to better understand the roleof this system in human obesity-related hypertension and insulinresistance. Finally, improved knowledge in these areas will lead tonovel therapeutic targets for metabolic syndrome treatment.

Acknowledgments

This work was supported by CONACYT Mexico (grant 154939 toA.R. Tovar). A Postdoctoral Fellowship from CONACYT was granted toM.E. Frigolet.

References

[1] Phillips MI, Schmidt-Ott KM. The Discovery of Renin 100 Years Ago. News inphysiological sciences : an international journal of physiology produced jointlyby the International Union of Physiological Sciences and the AmericanPhysiological Society. 1999;14:271–4.

[2] Naruse K, Inagami T, Celio MR, Workman RJ, Takii Y. Immunohistochemicalevidence that angiotensins I and II are formed by intracellular mechanism injuxtaglomerular cells. Hypertension 1982;4:70–4.

[3] Persson PB. Renin: origin, secretion and synthesis. J Physiol 2003;552:667–71.[4] Nguyen A, Cat D, Touyz RM. A new look at the renin–angiotensin system —

focusing on the vascular system. Peptides 2011;32:2141–50.[5] Beneteau-Burnat B, Baudin B. Angiotensin-converting enzyme: clinical applica-

tions and laboratory investigations on serum and other biological fluids. Crit RevClin Lab Sci 1991;28:337–56.

[6] Kuoppala A, Lindstedt KA, Saarinen J, Kovanen PT, Kokkonen JO. Inactivation ofbradykinin by angiotensin-converting enzyme and by carboxypeptidase N inhuman plasma. Am J Physiol Heart Circ Physiol 2000;278:H1069–74.

[7] Higuchi S, Ohtsu H, Suzuki H, Shirai H, Frank GD, Eguchi S. Angiotensin II signaltransduction through the AT1 receptor: novel insights into mechanisms andpathophysiology. Clin Sci (Lond) 2007;112:417–28.

[8] Lemarié CA, Schiffrin EL. The angiotensin II type 2 receptor in cardiovasculardisease. J Renin Angiotensin Aldosterone Syst 2010;11:19–31.

[9] Savoia C, Touyz RM, Endemann DH, Pu Q, Ko EA, De Ciuceis C, et al. Angiotensinreceptor blocker added to previous antihypertensive agents on arteries ofdiabetic hypertensive patients. Hypertension 2006;48:271–7.

[10] Horiuchi M, Lehtonen JY, Daviet L. Signaling mechanism of the AT2 angiotensin IIreceptor: crosstalk between AT1 and AT2 receptors in cell growth. TrendsEndocrinol Metab 1999;10:391–6.

[11] Putnam K, Shoemaker R, Yiannikouris F, Cassis LA. The renin–angiotensinsystem: a target of and contributor to dyslipidemias, altered glucose homeosta-sis, and hypertension of the metabolic syndrome. Am J Physiol Heart Circ Physiol2012;302:H1219–30.

[12] De Mello WC, Frohlich ED. On the local cardiac renin angiotensin system. Basicand clinical implications. Peptides 2011;32:1774–9.

[13] Kumar R, Thomas CM, Yong QC, Chen W, Baker KM. The intracrine renin–angiotensin system. Clin Sci (Lond) 2012;123:273–84.

[14] Balcells E, Meng QC, Johnson Jr WH, Oparil S, Dell'Italia LJ. Angiotensin IIformation from ACE and chymase in human and animal hearts: methods andspecies considerations. Am J Physiol 1997;273:H1769–74.

[15] Urata H, Boehm KD, Philip A, Kinoshita A, Gabrovsek J, Bumpus FM, et al. Cellularlocalization and regional distribution of an angiotensin II-forming chymase inthe heart. J Clin Invest 1993;91:1269–81.

[16] Johnston AP, Baker J, De Lisio M, Parise G. Skeletal muscle myoblasts possess astretch-responsive local angiotensin signalling system. J Renin AngiotensinAldosterone Syst 2011;12:75–84.

[17] Campbell DJ, Habener JF. Angiotensinogen gene is expressed and differentiallyregulated in multiple tissues of the rat. J Clin Invest 1986;78:31–9.

[18] Deschepper CF, Mellon SH, Cumin F, Baxter JD, Ganong WF. Analysis byimmunocytochemistry and in situ hybridization of renin and its mRNA inkidney, testis, adrenal, and pituitary of the rat. Proc Natl Acad Sci USA1986;83:7552–6.

[19] Ganten D, Minnich JL, Granger P, Hayduk K, Brecht HM, Barbeau A, et al.Angiotensin-forming enzyme in brain tissue. Science 1971;173:64–5.

[20] Hellmann W, Suzuki F, Ohkubo H, Nakanishi S, Ludwig G, Ganten D.Angiotensinogen gene expression in extrahepatic rat tissues: application of asolution hybridization assay. Naunyn Schmiedebergs Arch Pharmacol 1988;338:327–31.

[21] Bays HE, Gonzalez-Campoy JM, Bray GA, Kitabchi AE, Bergman DA, Schorr AB,et al. Pathogenic potential of adipose tissue and metabolic consequences ofadipocyte hypertrophy and increased visceral adiposity. Expert Rev CardiovascTher 2008;6:343–68.

[22] Janke J, Engeli S, Gorzelniak K, Luft FC, Sharma AM. Mature adipocytes inhibit invitro differentiation of human preadipocytes via angiotensin type 1 receptors.Diabetes 2002;51:1699–707.

[23] Engeli S, Negrel R, Sharma AM. Physiology and pathophysiology of the adiposetissue renin–angiotensin system. Hypertension 2000;35:1270–7.

[24] Cassis LA. Fat cell metabolism: insulin, fatty acids, and renin. Curr Hypertens Rep2000;2:132–8.

[25] Umemura S, Nyui N, Tamura K, Hibi K, Yamaguchi S, Nakamaru M, et al. Plasmaangiotensinogen concentrations in obese patients. Am J Hypertens 1997;10:629–33.

[26] Van Harmelen V, Ariapart P, Hoffstedt J, Lundkvist I, Bringman S, Arner P.Increased adipose angiotensinogen gene expression in human obesity. Obes Res2000;8:337–41.

Page 10: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2012 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

[27] Giacchetti G, Faloia E, Sardu C, Camilloni MA, Mariniello B, Gatti C, et al. Geneexpression of angiotensinogen in adipose tissue of obese patients. Int J Obes2000;24(Suppl 2):S142–3.

[28] Jonsson JR, Game PA, Head RJ, Frewin DB. The expression and localisation of theangiotensin-converting enzyme mRNA in human adipose tissue. Blood Press1994;3:72–5.

[29] Fowler JD, Johnson ND, Haroldson TA, Brintnall JA, Herrera JE, Katz SA, et al.Regulated renin release from 3T3-L1 adipocytes. Am J Physiol Endocrinol Metab2009;296:E1383–91.

[30] Saye JA, Ragsdale NV, Carey RM, Peach MJ. Localization of angiotensin peptide-forming enzymes of 3T3-F442A adipocytes. Am J Physiol 1993;264:C1570–6.

[31] Shenoy U, Cassis L. Characterization of renin activity in brown adipose tissue. AmJ Physiol 1997;272:C989–99.

[32] Weiland F, Verspohl EJ. Local formation of angiotensin peptides with paracrineactivity by adipocytes. J Pept Sci 2009;15:767–76.

[33] Achard V, Boullu-Ciocca S, Desbriere R, Nguyen G, Grino M. Renin receptorexpression in human adipose tissue. Am J Physiol Regul Integr Comp Physiol2007;292:R274–82.

[34] Takahashi K, Yamamoto H, Hirose T, Hiraishi K, Shoji I, Shibasaki A, et al.Expression of (pro)renin receptor in human kidneys with end-stage kidneydisease due to diabetic nephropathy. Peptides 2010;31:1405–8.

[35] Archard V, Tassistro V, Boullu-Giocca S, Grino M. Expression and nutritionalregulation of the (pro)renin receptor in rat visceral adipose tissue. J EndocrinolInvest 2011;34:840–6.

[36] Karlsson C, Lindell K, Ottosson M, Sjostrom L, Carlsson B, Carlsson LM. Humanadipose tissue expresses angiotensinogen and enzymes required for itsconversion to angiotensin II. J Clin Endocrinol Metab 1998;83:3925–9.

[37] Engeli S, Gorzelniak K, Kreutz R, Runkel N, Distler A, Sharma AM. Co-expressionof renin–angiotensin system genes in human adipose tissue. J Hypertens1999;17:555–60.

[38] Kalupahana NS, Moustaid-Moussa N. The renin–angiotensin system: a linkbetween obesity, inflammation and insulin resistance. Obes Rev 2012;13:136–49.

[39] Taleb S, Cancello R, Clement K, Lacasa D. Cathepsin s promotes humanpreadipocyte differentiation: possible involvement of fibronectin degradation.Endocrinology 2006;147:4950–9.

[40] Xiao Y, Junfeng H, Tianhong L, LuW, Shulin C, Yu Z, et al. Cathepsin K in adipocytedifferentiation and its potential role in the pathogenesis of obesity. J ClinEndocrinol Metab 2006;91:4520–7.

[41] Yang M, Zhang Y, Pan J, Sun J, Liu J, Libby P, et al. Cathepsin L activity controlsadipogenesis and glucose tolerance. Nat Cell Biol 2007;9:970–7.

[42] Crandall DL, Herzlinger HE, Saunders BD, Zolotor RC, Feliciano L, Cervoni P.Identification and characterization of angiotensin II receptors in rat epididymaladipocyte membranes. Metabolism 1993;42:511–5.

[43] Crandall DL, Armellino DC, Busler DE, McHendry-Rinde B, Kral JG. Angiotensin IIreceptors in human preadipocytes: role in cell cycle regulation. Endocrinology1999;140:154–8.

[44] Crandall DL, Herzlinger HE, Saunders BD, Armellino DC, Kral JG. Distribution ofangiotensin II receptors in rat and human adipocytes. J Lipid Res 1994;35:1378–85.

[45] Gorzelniak K, Engeli S, Janke J, Luft FC, Sharma AM. Hormonal regulation of thehuman adipose-tissue renin–angiotensin system: relationship to obesity andhypertension. J Hypertens 2002;20:965–73.

[46] Jones BH, Standridge MK, Moustaid N. Angiotensin II increases lipogenesis in3T3-L1 and human adipose cells. Endocrinology 1997;138:1512–9.

[47] Mallow H, Trindl A, Loffler G. Production of angiotensin II receptors type one(AT1) and type two (AT2) during the differentiation of 3T3-L1 preadipocytes.Horm Metab Res 2000;32:500–3.

[48] Juan CC, Chien Y,Wu LY, YangWM, Chang CL, Lai YH, et al. Angiotensin II enhancesinsulin sensitivity in vitro and in vivo. Endocrinology 2005;146:2246–54.

[49] Morris KL, Zemel MB. Effect of dietary carbohydrate source on the developmentof obesity in agouti transgenic mice. Obes Res 2005;13:21–35.

[50] Kim S, Iwao H. Molecular and cellular mechanisms of angiotensin II-mediatedcardiovascular and renal diseases. Pharmacol Rev 2000;52:11–34.

[51] Weiland F, Verspohl EJ. Variety of angiotensin receptors in 3T3-L1 preadiposecells and differentiated adipocytes. Horm Metab Res 2008;40:760–6.

[52] Benter IF, Yousif MH, Cojocel C, Al-Maghrebi M, Diz DI. Angiotensin-(1–7)prevents diabetes-induced cardiovascular dysfunction. Am J Physiol Heart CircPhysiol 2007;292:H666–72.

[53] Liu C, Lv XH, Li HX, Cao X, Zhang F, Wang L, et al. Angiotensin-(1–7) suppressesoxidative stress and improves glucose uptake via Mas receptor in adipocytes.Acta Diabetol 2012;49:291–9.

[54] Mario EG, Santos SH, Ferreira AV, Bader M, Santos RA, Botion LM. Angioten-sin-(1–7) Mas-receptor deficiency decreases peroxisome proliferator-activatedreceptor gamma expression in adipocytes. Peptides 2012;33:174–7.

[55] Caron AZ, Arguin G, Guillemette G. Angiotensin IV interacts with a juxtamem-brane site on AT(4)/IRAP suggesting an allosteric mechanism of enzymemodulation. Regul Pept 2003;113:9–15.

[56] Keller SR, Davis AC, Clairmont KB. Mice deficient in the insulin-regulatedmembrane aminopeptidase show substantial decreases in glucose transporterGLUT4 levels but maintain normal glucose homeostasis. J Biol Chem 2002;277:17677–86.

[57] Keller SR, Scott HM, Mastick CC, Aebersold R, Lienhard GE. Cloning andcharacterization of a novel insulin-regulated membrane aminopeptidase fromGlut4 vesicles. J Biol Chem 1995;270:23612–8.

[58] Marcus Y, Shefer G, Stern N. Adipose tissue renin–angiotensin -aldosteronesystem (RAAS) and progression of insulin resistance. Mol Cell Endocrinol2013;378:1–14.

[59] Cassis LA, Police SB, Yiannikouris F, Thatcher SE. Local adipose tissue renin–angiotensin system. Curr Hypertens Rep 2008;10:93–8.

[60] Faloia E, Gatti C, Camilloni MA, Mariniello B, Sardu C, Garrapa GG, et al.Comparison of circulating and local adipose tissue renin–angiotensin system innormotensive and hypertensive obese subjects. J Endocrinol Invest 2002;25:309–14.

[61] Massiera F, Bloch-Faure M, Ceiler D, Murakami K, Fukamizu A, Gasc JM, et al.Adipose angiotensinogen is involved in adipose tissue growth and bloodpressure regulation. FASEB J 2001;15:2727–9.

[62] Yasue S, Masuzaki H, Okada S, Ishii T, Kozuka C, Tanaka T, et al. Adipose tissue-specific regulation of angiotensinogen in obese humans and mice: impact ofnutritional status and adipocyte hypertrophy. Am J Hypertens 2010;23:425–31.

[63] Galvez-Prieto B, Bolbrinker J, Stucchi P, de Las Heras AI, Merino B, Arribas S, et al.Comparative expression analysis of the renin–angiotensin system componentsbetween white and brown perivascular adipose tissue. J Endocrinol 2008;197:55–64.

[64] Greenstein AS, Khavandi K, Withers SB, Sonoyama K, Clancy O, Jeziorska M, et al.Local inflammation and hypoxia abolish the protective anticontractile propertiesof perivascular fat in obese patients. Circulation 2009;119:1661–70.

[65] Lu C, Zhao AX, Gao YJ, Lee RM. Modulation of vein function by perivascularadipose tissue. Eur J Pharmacol 2011;657:111–6.

[66] Oliver JA, Sciacca RR. Local generation of angiotensin II as a mechanism ofregulation of peripheral vascular tone in the rat. J Clin Invest 1984;74:1247–51.

[67] Giacchetti G, Faloia E, Mariniello B, Sardu C, Gatti C, Camilloni MA, et al.Overexpression of the renin–angiotensin system in human visceral adiposetissue in normal and overweight subjects. Am J Hypertens 2002;15:381–8.

[68] Anderson PJ, Chan JC, Chan YL, Tomlinson B, Young RP, Lee ZS, et al. Visceral fatand cardiovascular risk factors in Chinese NIDDM patients. Diabetes Care1997;20:1854–8.

[69] Peiris AN, Sothmann MS, Hoffmann RG, Hennes MI, Wilson CR, Gustafson AB,et al. Adiposity, fat distribution, and cardiovascular risk. Ann Intern Med1989;110:867–72.

[70] Hajer GR, van Haeften TW, Visseren FL. Adipose tissue dysfunction in obesity,diabetes, and vascular diseases. Eur Heart J 2008;29:2959–71.

[71] Pausova Z. From big fat cells to high blood pressure: a pathway to obesity-associated hypertension. Curr Opin Nephrol Hypertens 2006;15:173–8.

[72] Iwai M, Horiuchi M. Role of renin–angiotensin system in adipose tissuedysfunction. Hypertens Res 2009;32:425–7.

[73] Boschmann M, Ringel J, Klaus S, Sharma AM. Metabolic and hemodynamicresponse of adipose tissue to angiotensin II. Obes Res 2001;9:486–91.

[74] Kouyama R, Suganami T, Nishida J, Tanaka M, Toyoda T, Kiso M, et al. Attenuationof diet-inducedweight gain and adiposity through increased energy expenditurein mice lacking angiotensin II type 1a receptor. Endocrinology 2005;146:3481–9.

[75] Iwai M, Chen R, Imura Y, Horiuchi M. TAK-536, a new AT1 receptor blocker,improves glucose intolerance and adipocyte differentiation. Am J Hypertens2007;20:579–86.

[76] Henriksen EJ, Jacob S, Kinnick TR, TeacheyMK, Krekler M. Selective angiotensin IIreceptor antagonism reduces insulin resistance in obese Zucker rats. Hyperten-sion 2001;38:884–90.

[77] MunozMC, Giani JF, Dominici FP, Turyn D, Toblli JE. Long-term treatment with anangiotensin II receptor blocker decreases adipocyte size and improves insulinsignaling in obese Zucker rats. J Hypertens 2009;27:2409–20.

[78] Henriksen EJ, Jacob S. Effects of captopril on glucose transport activity in skeletalmuscle of obese Zucker rats. Metabolism 1995;44:267–72.

[79] Jacob S, Henriksen EJ, Fogt DL, Dietze GJ. Effects of trandolapril and verapamil onglucose transport in insulin-resistant rat skeletal muscle. Metabolism 1996;45:535–41.

[80] Jauch KW, Hartl W, Guenther B, Wicklmayr M, Rett K, Dietze G. Captoprilenhances insulin responsiveness of forearm muscle tissue in non-insulin-dependent diabetes mellitus. Eur J Clin Invest 1987;17:448–54.

[81] Paolisso G, Gambardella A, Verza M, D'Amore A, Sgambato S, Varricchio M. ACEinhibition improves insulin-sensitivity in aged insulin-resistant hypertensivepatients. J Hum Hypertens 1992;6:175–9.

[82] Pollare T. Insulin sensitivity and blood lipids during antihypertensive treatmentwith special reference to ACE inhibition. J Diabet Complications 1990;4:75–8.

[83] Torlone E, Britta M, Rambotti AM, Perriello G, Santeusanio F, Brunetti P, et al.Improved insulin action and glycemic control after long-term angiotensin-converting enzyme inhibition in subjects with arterial hypertension and type IIdiabetes. Diabetes Care 1993;16:1347–55.

[84] Vuorinen-Markkola H, Yki-Jarvinen H. Antihypertensive therapy with enalaprilimproves glucose storage and insulin sensitivity in hypertensive patients withnon-insulin-dependent diabetes mellitus. Metabolism 1995;44:85–9.

[85] Lastra G, Habibi J, Whaley-Connell AT, Manrique C, Hayden MR, Rehmer J, et al.Direct renin inhibition improves systemic insulin resistance and skeletal muscleglucose transport in a transgenic rodent model of tissue renin overexpression.Endocrinology 2009;150:2561–8.

[86] Marchionne EM, Diamond-Stanic MK, Prasonnarong M, Henriksen EJ. Chronicrenin inhibition with aliskiren improves glucose tolerance, insulin sensitivity,and skeletal muscle glucose transport activity in obese Zucker rats. Am J PhysiolRegul Integr Comp Physiol 2012;302:R137–42.

[87] Folli F, Kahn CR, Hansen H, Bouchie JL, Feener EP. Angiotensin II inhibits insulinsignaling in aortic smooth muscle cells at multiple levels. A potential role for

Page 11: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2013M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

serine phosphorylation in insulin/angiotensin II crosstalk. J Clin Invest 1997;100:2158–69.

[88] Velloso LA, Folli F, Sun XJ, White MF, Saad MJ, Kahn CR. Cross-talk between theinsulin and angiotensin signaling systems. Proc Natl Acad Sci USA 1996;93:12490–5.

[89] Takayama S, White MF, Kahn CR. Phorbol ester-induced serine phosphorylationof the insulin receptor decreases its tyrosine kinase activity. J Biol Chem1988;263:3440–7.

[90] Tanti JF, Gremeaux T, van Obberghen E, Le Marchand-Brustel Y. Serine/threoninephosphorylation of insulin receptor substrate 1 modulates insulin receptorsignaling. J Biol Chem 1994;269:6051–7.

[91] Hirosumi J, Tuncman G, Chang L, Gorgun CZ, Uysal KT, Maeda K, et al. A centralrole for JNK in obesity and insulin resistance. Nature 2002;420:333–6.

[92] Zohn IE, Yu H, Li X, Cox AD, Earp HS. Angiotensin II stimulates calcium-dependent activation of c-Jun N-terminal kinase. Mol Cell Biol 1995;15:6160–8.

[93] Chang HH, Tsai PH, Liu CW, Ku HC, Kao CC, Kao YH. Cycloheximide stimulatessuppressor of cytokine signaling-3 gene expression in 3T3-L1 adipocytes via theextracellular signal-regulated kinase pathway. Toxicol Lett 2013;217:42–9.

[94] Emanuelli B, Peraldi P, Filloux C, Chavey C, Freidinger K, Hilton DJ, et al. SOCS-3inhibits insulin signaling and is up-regulated in response to tumor necrosisfactor-alpha in the adipose tissue of obese mice. J Biol Chem 2001;276:47944–9.

[95] Emanuelli B, Peraldi P, Filloux C, Sawka-Verhelle D, Hilton D, Van Obberghen E.SOCS-3 is an insulin-induced negative regulator of insulin signaling. J Biol Chem2000;275:15985–91.

[96] Ueki K, Kondo T, Kahn CR. Suppressor of cytokine signaling 1 (SOCS-1) andSOCS-3 cause insulin resistance through inhibition of tyrosine phosphorylationof insulin receptor substrate proteins by discrete mechanisms. Mol Cell Biol2004;24:5434–46.

[97] Hilton DJ, Richardson RT, Alexander WS, Viney EM, Willson TA, Sprigg NS, et al.Twenty proteins containing a C-terminal SOCS box form five structural classes.Proc Natl Acad Sci USA 1998;95:114–9.

[98] Velloso LA, Folli F, Perego L, Saad MJ. The multi-faceted cross-talk between theinsulin and angiotensin II signaling systems. Diabetes Metab Res Rev 2006;22:98–107.

[99] Limor R, Moataz U, Knol E, Stern N. Aldosterone increase oxidative stress indifferentiated but not in undifferentiated 3T3-L1 adipocytes via activation ofmineralocorticoid and glucocorticoid receptors. Florence: International Congressof Endocrinology; 2012.

[100] McCartyMF. ACE inhibitionmay decrease diabetes risk by boosting the impact ofbradykinin on adipocytes. Med Hypotheses 2003;60:779–83.

[101] Waeber B, Brunner HR. Cardiovascular hypertrophy: role of angiotensin II andbradykinin. J Cardiovasc Pharmacol 1996;27(Suppl 2):S36–40.

[102] Isami S, Kishikawa H, Araki E, Uehara M, Kaneko K, Shirotani T, et al. Bradykininenhances GLUT4 translocation through the increase of insulin receptor tyrosinekinase in primary adipocytes: evidence that bradykinin stimulates the insulinsignalling pathway. Diabetologia 1996;39:412–20.

[103] Beard KM, Lu H, Ho K, Fantus IG. Bradykinin augments insulin-stimulatedglucose transport in rat adipocytes via endothelial nitric oxide synthase-mediated inhibition of Jun NH2-terminal kinase. Diabetes 2006;55:2678–87.

[104] Imayama I, Ichiki T, Inanaga K, Ohtsubo H, Fukuyama K, Ono H, et al. Telmisartandownregulates angiotensin II type 1 receptor through activation of peroxisomeproliferator-activated receptor gamma. Cardiovasc Res 2006;72:184–90.

[105] Benson SC, Pershadsingh HA, Ho CI, Chittiboyina A, Desai P, Pravenec M, et al.Identification of telmisartan as a unique angiotensin II receptor antagonist withselective PPARgamma-modulating activity. Hypertension 2004;43:993–1002.

[106] Schupp M, Janke J, Clasen R, Unger T, Kintscher U. Angiotensin type 1 receptorblockers induce peroxisome proliferator-activated receptor-gamma activity.Circulation 2004;109:2054–7.

[107] Kintscher U, Law RE. PPARgamma-mediated insulin sensitization: theimportance of fat versus muscle. Am J Physiol Endocrinol Metab 2005;288:E287–91.

[108] Medina-Gomez G, Gray S, Vidal-Puig A. Adipogenesis and lipotoxicity: role ofperoxisome proliferator-activated receptor gamma (PPARgamma) and PPAR-gammacoactivator-1 (PGC1). Public Health Nutr 2007;10:1132–7.

[109] Fain JN, Madan AK, Hiler ML, Cheema P, Bahouth SW. Comparison of the releaseof adipokines by adipose tissue, adipose tissue matrix, and adipocytes fromvisceral and subcutaneous abdominal adipose tissues of obese humans.Endocrinology 2004;145:2273–82.

[110] Funakoshi Y, Ichiki T, Shimokawa H, Egashira K, Takeda K, Kaibuchi K, et al. Rho-kinase mediates angiotensin II-induced monocyte chemoattractant protein-1expression in rat vascular smooth muscle cells. Hypertension 2001;38:100–4.

[111] Asamizu S, Urakaze M, Kobashi C, Ishiki M, Norel Din AK, Fujisaka S, et al.Angiotensin II enhances the increase in monocyte chemoattractant protein-1production induced by tumor necrosis factor-{alpha} from 3T3-L1 preadipo-cytes. J Endocrinol 2009;202:199–205.

[112] Kurata A, Nishizawa H, Kihara S, Maeda N, Sonoda M, Okada T, et al. Blockade ofangiotensin II type-1 receptor reduces oxidative stress in adipose tissue andameliorates adipocytokine dysregulation. Kidney Int 2006;70:1717–24.

[113] HungWW, Hsieh TJ, Lin T, Chou PC, Hsiao PJ, Lin KD, et al. Blockade of the renin–angiotensin system ameliorates apelin production in 3T3-L1 adipocytes.Cardiovasc Drugs Ther 2011;25:3–12.

[114] Benter IF, Yousif MH, Anim JT, Cojocel C, Diz DI. Angiotensin-(1–7) preventsdevelopment of severe hypertension and end-organ damage in spontaneouslyhypertensive rats treated with L-NAME. Am J Physiol Heart Circ Physiol2006;290:H684–91.

[115] Santos SH, Braga JF, Mario EG, Porto LC, Rodrigues-Machado Mda G, Murari A,et al. Improved lipid and glucose metabolism in transgenic rats with increasedcirculating angiotensin-(1–7). Arterioscler Thromb Vasc Biol 2010;30:953–61.

[116] Cornelius P, MacDougald OA, Lane MD. Regulation of adipocyte development.Annu Rev Nutr 1994;14:99–129.

[117] Masson O, Prebois C, Derocq D, Meulle A, Dray C, Daviaud D, et al. Cathepsin-D, akey protease in breast cancer, is up-regulated in obese mouse and humanadipose tissue, and controls adipogenesis. PLoS One 2011;6:e16452.

[118] Vassaux G, Gaillard D, Ailhaud G, Negrel R. Prostacyclin is a specific effector ofadipose cell differentiation. Its dual role as a cAMP- and Ca(2+)-elevating agent.J Biol Chem 1992;267:11092–7.

[119] Darimont C, Vassaux G, Gaillard D, Ailhaud G, Negrel R. In situ microdialysis ofprostaglandins in adipose tissue: stimulation of prostacyclin release byangiotensin II. Int J Obes 1994;18:783–8.

[120] Hauner H. Adipocytes as a source and target of hormones: recent developmentsin adipose tissue physiology. Int Mon on EP & WC. 1996;5.

[121] Saint-Marc P, Kozak LP, Ailhaud G, Darimont C, Negrel R. Angiotensin II as atrophic factor of white adipose tissue: stimulation of adipose cell formation.Endocrinology 2001;142:487–92.

[122] Fuentes P, Acuna MJ, Cifuentes M, Rojas CV. The anti-adipogenic effect ofangiotensin II on human preadipose cells involves ERK1,2 activation and PPARGphosphorylation. J Endocrinol 2010;206:75–83.

[123] Aubert J, Ailhaud G, Negrel R. Evidence for a novel regulatory pathway activated by(carba)prostacyclin in preadipose and adipose cells. FEBS Lett 1996;397:117–21.

[124] Massiera F, Saint-Marc P, Seydoux J, Murata T, Kobayashi T, Narumiya S, et al.Arachidonic acid and prostacyclin signaling promote adipose tissue develop-ment: a human health concern? J Lipid Res 2003;44:271–9.

[125] Negrel R. Prostacyclin as a critical prostanoid in adipogenesis. ProstaglandinsLeukot Essent Fatty Acids 1999;60:383–6.

[126] Darimont C, Vassaux G, Ailhaud G, Negrel R. Differentiation of preadipose cells:paracrine role of prostacyclin upon stimulation of adipose cells by angiotensin-II.Endocrinology 1994;135:2030–6.

[127] Iwai M, Tomono Y, Inaba S, Kanno H, Senba I, Mogi M, et al. AT2 receptordeficiency attenuates adipocyte differentiation and decreases adipocyte numberin atherosclerotic mice. Am J Hypertens 2009;22:784–91.

[128] Brucher R, Cifuentes M, Acuna MJ, Albala C, Rojas CV. Larger anti-adipogeniceffect of angiotensin II on omental preadipose cells of obese humans. Obesity(Silver Spring) 2007;15:1643–6.

[129] Strazzullo P, Galletti F. Impact of the renin–angiotensin system on lipid andcarbohydrate metabolism. Curr Opin Nephrol Hypertens 2004;13:325–32.

[130] Ailhaud G, Teboul M, Massiera F. Angiotensinogen, adipocyte differentiation andfat mass enlargement. Curr Opin Clin Nutr Metab Care 2002;5:385–9.

[131] Tomono Y, Iwai M, Inaba S, Mogi M, Horiuchi M. Blockade of AT1 receptorimproves adipocyte differentiation in atherosclerotic and diabetic models. Am JHypertens 2008;21:206–12.

[132] Zorad S, Dou JT, Benicky J, Hutanu D, Tybitanclova K, Zhou J, et al. Long-termangiotensin II AT1 receptor inhibition produces adipose tissue hypotrophyaccompanied by increased expression of adiponectin and PPARgamma. Eur JPharmacol 2006;552:112–22.

[133] Cassis LA, Fettinger MJ, Roe AL, Shenoy UR, Howard G. Characterization andregulation of angiotensin II receptors in rat adipose tissue. Angiotensin receptorsin adipose tissue. Adv Exp Med Biol 1996;396:39–47.

[134] Kalupahana NS, Massiera F, Quignard-Boulange A, Ailhaud G, Voy BH,Wasserman DH, et al. Overproduction of angiotensinogen from adipose tissueinduces adipose inflammation, glucose intolerance, and insulin resistance.Obesity (Silver Spring) 2012;20:48–56.

[135] Yvan-Charvet L, Massiera F, Lamande N, Ailhaud G, Teboul M, Moustaid-MoussaN, et al. Deficiency of angiotensin type 2 receptor rescues obesity but nothypertension induced by overexpression of angiotensinogen in adipose tissue.Endocrinology 2009;150:1421–8.

[136] Carroll WX, Kalupahana NS, Booker SL, Siriwardhana N, Lemieux M, Saxton AM,et al. Angiotensinogen gene silencing reduces markers of lipid accumulation andinflammation in cultured adipocytes. Front Endocrinol 2013;4:10.

[137] Jurewicz M, McDermott DH, Sechler JM, Tinckam K, Takakura A, Carpenter CB,et al. Human T and natural killer cells possess a functional renin–angiotensinsystem: further mechanisms of angiotensin II-induced inflammation. J Am SocNephrol 2007;18:1093–102.

[138] Skurk T, van Harmelen V, Hauner H. Angiotensin II stimulates the release ofinterleukin-6 and interleukin-8 from cultured human adipocytes by activation ofNF-kappaB. Arterioscler Thromb Vasc Biol 2004;24:1199–203.

[139] Tsuchiya K, Yoshimoto T, Hirono Y, Tateno T, Sugiyama T, Hirata Y. Angiotensin IIinduces monocyte chemoattractant protein-1 expression via a nuclear factor-kappaB-dependent pathway in rat preadipocytes. Am J Physiol Endocrinol Metab2006;291:E771–8.

[140] Fujisaka S, Usui I, Kanatani Y, Ikutani M, Takasaki I, Tsuneyama K, et al.Telmisartan improves insulin resistance and modulates adipose tissue macro-phage polarization in high-fat-fed mice. Endocrinology 2011;152:1789–99.

[141] Cole BK, Keller SR, Wu R, Carter JD, Nadler JL, Nunemaker CS. Valsartan protectspancreatic islets and adipose tissue from the inflammatory and metabolicconsequences of a high-fat diet in mice. Hypertension 2010;55:715–21.

[142] Hakim J. [Reactive oxygen species and inflammation]. Comptes rendus desseances de la Societe de biologie et de ses filiales. 1993;187:286–95.

[143] Chabrashvili T, Kitiyakara C, Blau J, Karber A, Aslam S, Welch WJ, et al. Effects ofANG II type 1 and 2 receptors on oxidative stress, renal NADPH oxidase, and SODexpression. Am J Physiol Regul Integr Comp Physiol 2003;285:R117–24.

Page 12: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2014 M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

[144] Hussein O, Shneider J, Rosenblat M, Aviram M. Valsartan therapy has additiveanti-oxidative effect to that of fluvastatin therapy against low-densitylipoprotein oxidation: studies in hypercholesterolemic and hypertensivepatients. J Cardiovasc Pharmacol 2002;40:28–34.

[145] Mervaala EM, Cheng ZJ, Tikkanen I, Lapatto R, Nurminen K, Vapaatalo H, et al.Endothelial dysfunction and xanthine oxidoreductase activity in rats withhuman renin and angiotensinogen genes. Hypertension 2001;37:414–8.

[146] Elliott SS, Keim NL, Stern JS, Teff K, Havel PJ. Fructose, weight gain, and theinsulin resistance syndrome. Am J Clin Nutr 2002;76:911–22.

[147] Chou CL, Lai YH, Lin TY, Lee TJ, Fang TC. Aliskiren prevents and amelioratesmetabolic syndrome in fructose-fed rats. Arch Med Sci 2011;7:882–8.

[148] Rafiq K, Hitomi H, Nakano D, Ichihara A, Nishiyama A. Possible involvement ofthe (pro)renin receptor-dependent system in the development of insulinresistance. Front Biosci (Schol Ed) 2011;3:1478–85.

[149] Ishibashi S. The vascular renin–angiotensin system as a possible source ofvascular inflammation in fructose-fed rats. Hypertens Res 2007;30:375–6.

[150] Iyer SN, Katovich MJ, Raizada MK. Changes in angiotensin AT1 receptordensity during hypertension in fructose-fed rats. Adv Exp Med Biol 1996;396:49–58.

[151] Nyby MD, Abedi K, Smutko V, Eslami P, Tuck ML. Vascular angiotensin type 1receptor expression is associated with vascular dysfunction, oxidative stress andinflammation in fructose-fed rats. Hypertens Res 2007;30:451–7.

[152] Shinozaki K, Ayajiki K, Nishio Y, Sugaya T, Kashiwagi A, Okamura T. Evidence fora causal role of the renin–angiotensin system in vascular dysfunction associatedwith insulin resistance. Hypertension 2004;43:255–62.

[153] Pachori AS, Huentelman MJ, Francis SC, Gelband CH, Katovich MJ, Raizada MK.The future of hypertension therapy: sense, antisense, or nonsense? Hyperten-sion 2001;37:357–64.

[154] Giani JF, Mayer MA, Munoz MC, Silberman EA, Hocht C, Taira CA, et al. Chronicinfusion of angiotensin-(1–7) improves insulin resistance and hypertensioninduced by a high-fructose diet in rats. Am J Physiol Endocrinol Metab 2009;296:E262–71.

[155] Giacchetti G, Sechi LA, Griffin CA, Don BR, Mantero F, Schambelan M. The tissuerenin–angiotensin system in rats with fructose-induced hypertension: over-expression of type 1 angiotensin II receptor in adipose tissue. J Hypertens2000;18:695–702.

[156] Furuhashi M, Ura N, Takizawa H, Yoshida D, Moniwa N, Murakami H, et al.Blockade of the renin–angiotensin system decreases adipocyte size withimprovement in insulin sensitivity. J Hypertens 2004;22:1977–82.

[157] Freitas RR, Lopes KL, Carillo BA, Bergamaschi CT, Carmona AK, Casarini DE, et al.Sympathetic and renin–angiotensin systems contribute to increased bloodpressure in sucrose-fed rats. Am J Hypertens 2007;20:692–8.

[158] Noriega-Lopez L, Tovar AR, Gonzalez-Granillo M, Hernandez-Pando R, EscalanteB, Santillan-Doherty P, et al. Pancreatic insulin secretion in rats fed a soy proteinhigh fat diet depends on the interaction between the amino acid pattern andisoflavones. J Biol Chem 2007;282:20657–66.

[159] Palacios-Gonzalez B, Torres N, Tovar AR. Genistein stimulates AMPK phosphor-ylation and expression of genes involved in fatty acid oxidation in skeletalmuscle. FASEB J 2011;25:772.

[160] Tovar AR, Torre-Villalvazo I, Ochoa M, Elias AL, Ortiz V, Aguilar-Salinas CA, et al.Soy protein reduces hepatic lipotoxicity in hyperinsulinemic obese Zucker fa/farats. J Lipid Res 2005;46:1823–32.

[161] Frigolet ME, Torres N, Uribe-Figueroa L, Rangel C, Jimenez-Sanchez G, Tovar AR.White adipose tissue genome wide-expression profiling and adipocyte meta-bolic functions after soy protein consumption in rats. J Nutr Biochem 2011;22:118–29.

[162] Vasdev S, Stuckless J. Antihypertensive effects of dietary protein and itsmechanism. Int J Angiol 2010;19:e7-20.

[163] Frigolet ME, Torres N, Tovar AR. Soya protein attenuates abnormalities of therenin–angiotensin system in adipose tissue from obese rats. Br J Nutr 2012;107:36–44.

[164] LoWM, Li-Chan EC. Angiotensin I converting enzyme inhibitory peptides from invitro pepsin-pancreatin digestion of soy protein. J Agric Food Chem 2005;53:3369–76.

[165] Yang HY, Yang SC, Chen ST, Chen JR. Soy protein hydrolysate amelioratescardiovascular remodeling in rats with L-NAME-induced hypertension. J NutrBiochem 2008;19:833–9.

[166] Xu YY, Yang C, Li SN. Effects of genistein on angiotensin-converting enzyme inrats. Life Sci 2006;79:828–37.

[167] Ma LJ, Corsa BA, Zhou J, Yang H, Li H, Tang YW, et al. Angiotensin type 1 receptormodulates macrophage polarization and renal injury in obesity. Am J PhysiolRenal Physiol 2011;300:F1203–13.

[168] Massiera F, Seydoux J, Geloen A, Quignard-Boulange A, Turban S, Saint-Marc P,et al. Angiotensinogen-deficient mice exhibit impairment of diet-inducedweight gain with alteration in adipose tissue development and increasedlocomotor activity. Endocrinology 2001;142:5220–5.

[169] Boustany CM, Bharadwaj K, Daugherty A, Brown DR, Randall DC, Cassis LA.Activation of the systemic and adipose renin–angiotensin system in rats withdiet-induced obesity and hypertension. Am J Physiol Regul Integr Comp Physiol2004;287:R943–9.

[170] Gupte M, Boustany-Kari CM, Bharadwaj K, Police S, Thatcher S, Gong MC, et al.ACE2 is expressed in mouse adipocytes and regulated by a high-fat diet. Am JPhysiol Regul Integr Comp Physiol 2008;295:R781–8.

[171] Sun X, Pan H, Tan H, Yu Y. High free fatty acids level related with cardiacdysfunction in obese rats. Diabetes Res Clin Pract 2012;95:251–9.

[172] Tian YQ, Li SS, Su XD, Zhang GZ, Zhao JJ, Li GW, et al. Effects of pioglitazone onhigh-fat-diet-induced ventricular remodeling and dysfunction in rats. J Cardi-ovasc Pharmacol Ther 2012;17:223–8.

[173] Boustany CM, Brown DR, Randall DC, Cassis LA. AT1-receptor antagonismreverses the blood pressure elevation associated with diet-induced obesity. Am JPhysiol Regul Integr Comp Physiol 2005;289:R181–6.

[174] de las Heras N, Martin-Fernandez B, Miana M, Ballesteros S, Oubina MP, Lopez-Farre AJ, et al. The protective effect of irbesartan in rats fed a high fat diet isassociated with modification of leptin-adiponectin imbalance. J Hypertens Suppl2009;27:S37–41.

[175] Stucchi P, Cano V, Ruiz-Gayo M, Fernandez-Alfonso MS. Aliskiren reduces body-weight gain, adiposity and plasma leptin during diet-induced obesity. Br JPharmacol 2009;158:771–8.

[176] Weisinger RS, Stanley TK, Begg DP, Weisinger HS, Spark KJ, Jois M. Angiotensinconverting enzyme inhibition lowers body weight and improves glucosetolerance in C57BL/6J mice maintained on a high fat diet. Physiol Behav2009;98:192–7.

[177] Chung S, Park CW, Shin SJ, Lim JH, Chung HW, Youn DY, et al. Tempol orcandesartan prevents high-fat diet-induced hypertension and renal damage inspontaneously hypertensive rats. Nephrol Dial Transplant 2010;25:389–99.

[178] Li X, Ren H, Xu ZR, Liu YJ, Yang XP, Liu JQ. Prevalence and risk factors of prolongedQTc interval among Chinese patients with type 2 diabetes. Exp Diabetes Res2012;2012:234084.

[179] Yuan L, Li X, Li J, Li HL, Cheng SS. Effects of renin–angiotensin system blockade onthe islet morphology and function in rats with long-term high-fat diet. ActaDiabetol 2013;50:479–88.

[180] Vinh A, Widdop RE, Drummond GR, Gaspari TA. Chronic angiotensin IVtreatment reverses endothelial dysfunction in ApoE-deficient mice. CardiovascRes 2008;77:178–87.

[181] Northcott CA, Fink GD, Garver H, Haywood JR, Laimon-Thomson EL, McClain JL,et al. The development of hypertension and hyperaldosteronism in a rodentmodel of life-long obesity. Endocrinology 2012;153:1764–73.

[182] Takeda M, Yamamoto K, Takemura Y, Takeshita H, Hongyo K, Kawai T, et al. Lossof ACE2 exaggerates high-calorie diet-induced insulin resistance by reduction ofGLUT4 in mice. Diabetes 2013;62:223–33.

[183] Kurukulasuriya LR, Stas S, Lastra G, Manrique C, Sowers JR. Hypertension inobesity. Endocrinol Metab Clin North Am 2008;37:647–62 ix.

[184] Engler MM, Engler MB, Erickson SK, Paul SM. Dietary gamma-linolenic acidlowers blood pressure and alters aortic reactivity and cholesterol metabolism inhypertension. J Hypertens 1992;10:1197–204.

[185] Engler MM, Schambelan M, Engler MB, Ball DL, Goodfriend TL. Effects of dietarygamma-linolenic acid on blood pressure and adrenal angiotensin receptors inhypertensive rats. Proc Soc Exp Biol Med 1998;218:234–7.

[186] Jayasooriya AP, Begg DP, Chen N, Mathai ML, Sinclair AJ, Wilkinson-Berka J, et al.Omega-3 polyunsaturated fatty acid supplementation reduces hypertension inTGR(mRen-2)27 rats. Prostaglandins Leukot Essent Fatty Acids 2008;78:67–72.

[187] Wyrwoll CS, Mark PJ, Waddell BJ. Developmental programming of renalglucocorticoid sensitivity and the renin–angiotensin system. Hypertension2007;50:579–84.

[188] DeClercq V, Taylor CG, Zahradka P. Isomer-specific effects of conjugated linoleicacid on blood pressure, adipocyte size and function. Br J Nutr 2012;107:1413–21.

[189] Ferraro PM, Ferraccioli GF, Gambaro G, Fulignati P, Costanzi S. Combinedtreatment with renin–angiotensin system blockers and polyunsaturated fattyacids in proteinuric IgA nephropathy: a randomized controlled trial. NephrolDial Transplant 2009;24:156–60.

[190] Pinterova L, Zelezna B, Fickova M, Macho L, Krizanova O, Jezova D, et al. ElevatedAT1 receptor protein but lower angiotensin II-binding in adipose tissue of ratswith monosodium glutamate-induced obesity. Horm Metab Res 2001;33:708–12.

[191] Samuel P, Ali Q, Sabuhi R, Wu Y, Hussain T. High Na intake increases renalangiotensin II levels and reduces expression of the ACE2-AT(2)R-MasR axis inobese Zucker rats. Am J Physiol Renal Physiol 2012;303:F412–9.

[192] Perricone NV, Bagchi D, Echard B, Preuss HG. Long-term metabolic effects ofdifferent doses of niacin-bound chromium on Sprague–Dawley rats. Mol CellBiochem 2010;338:91–103.

[193] Huang WY, Davidge ST, Wu J. Bioactive natural constituents from food sources-potential use in hypertension prevention and treatment. Crit Rev Food Sci Nutr2013;53:615–30.

[194] Steckelings UM, Rompe F, Kaschina E, Unger T. The evolving story of the RAAS inhypertension, diabetes and CV disease: moving from macrovascular tomicrovascular targets. Fundam Clin Pharmacol 2009;23:693–703.

[195] Kalupahana NS, Moustaid-Moussa N. The adipose tissue renin–angiotensinsystem and metabolic disorders: a review of molecular mechanisms. Crit RevBiochem Mol Biol 2012;47:379–90.

[196] Kramer H, Wu X, Kan D, Luke A, Zhu X, Adeyemo A, et al. Angiotensin-convertingenzyme gene polymorphisms and obesity: an examination of three blackpopulations. Obes Res 2005;13:823–8.

[197] Strazzullo P, Iacone R, Iacoviello L, Russo O, Barba G, Russo P, et al. Geneticvariation in the renin–angiotensin system and abdominal adiposity in men: theOlivetti Prospective Heart Study. Ann Intern Med 2003;138:17–23.

[198] Engeli S, Bohnke J, Gorzelniak K, Janke J, Schling P, Bader M, et al. Weight loss andthe renin–angiotensin–aldosterone system. Hypertension 2005;45:356–62.

[199] Campbell DJ, Kladis A, Valentijn AJ. Effects of losartan on angiotensin andbradykinin peptides and angiotensin-converting enzyme. J Cardiovasc Pharma-col 1995;26:233–40.

Page 13: The renin–angiotensin system in adipose tissue and its metabolic consequences during obesity

2015M.E. Frigolet et al. / Journal of Nutritional Biochemistry 24 (2013) 2003–2015

[200] Setoguchi Y, Ohnuki T, Rashid M, Nakamura T, Hattori K, Nagatomo T, et al.Effects of chronic administration of sarpogrelate on systolic blood pressure ofspontaneously hypertensive rats: comparison with quinapril. Pharmacology2002;64:71–5.

[201] Yiannikouris F, Gupte M, Putnam K, Thatcher S, Charnigo R, Rateri DL, et al.Adipocyte deficiency of angiotensinogen prevents obesity-induced hypertensionin male mice. Hypertension 2012;60:1524–30.

[202] Hall JE, BrandsMW, Henegar JR. Mechanisms of hypertension and kidney diseasein obesity. Ann N Y Acad Sci 1999;892:91–107.

[203] de Paula RB, da Silva AA, Hall JE. Aldosterone antagonism attenuates obesity-induced hypertension and glomerular hyperfiltration. Hypertension 2004;43:41–7.

[204] Reisin E, Weir MR, Falkner B, Hutchinson HG, Anzalone DA, Tuck ML. Lisinoprilversus hydrochlorothiazide in obese hypertensive patients: a multicenterplacebo-controlled trial. Treatment in Obese Patients With Hypertension(TROPHY) Study Group. Hypertension 1997;30:140–5.

[205] Vermes E, Ducharme A, Bourassa MG, Lessard M, White M, Tardif JC. Enalaprilreduces the incidence of diabetes in patients with chronic heart failure: insightfrom the Studies Of Left Ventricular Dysfunction (SOLVD). Circulation 2003;107:1291–6.

[206] Nagel JM, Tietz AB, Goke B, Parhofer KG. The effect of telmisartan on glucose andlipidmetabolism in nondiabetic, insulin-resistant subjects. Metabolism 2006;55:1149–54.

[207] Steen MS, Foianini KR, Youngblood EB, Kinnick TR, Jacob S, Henriksen EJ.Interactions of exercise training and ACE inhibition on insulin action in obeseZucker rats. J Appl Physiol 1999;86:2044–51.

[208] Frederich Jr RC, Kahn BB, Peach MJ, Flier JS. Tissue-specific nutritional regulationof angiotensinogen in adipose tissue. Hypertension 1992;19:339–44.

[209] Hainault I, Nebout G, Turban S, Ardouin B, Ferre P, Quignard-Boulange A. Adiposetissue-specific increase in angiotensinogen expression and secretion in the obese(fa/fa) Zucker rat. Am J Physiol Endocrinol Metab 2002;282:E59–66.

[210] Lu H, Boustany-Kari CM, Daugherty A, Cassis LA. Angiotensin II increases adiposeangiotensinogen expression. Am J Physiol Endocrinol Metab 2007;292:E1280–7.

[211] Skurk T, Lee YM, Hauner H. Angiotensin II and its metabolites stimulate PAI-1protein release from human adipocytes in primary culture. Hypertension2001;37:1336–40.

[212] Skurk T, van Harmelen V, Blum WF, Hauner H. Angiotensin II promotes leptinproduction in cultured human fat cells by an ERK1/2-dependent pathway. ObesRes 2005;13:969–73.

[213] Kim S, Whelan J, Claycombe K, Reath DB, Moustaid-Moussa N. Angiotensin IIincreases leptin secretion by 3T3-L1 and human adipocytes via a prostaglandin-independent mechanism. J Nutr 2002;132:1135–40.

[214] Dunbar JC, Hu Y, Lu H. Intracerebroventricular leptin increases lumbar andrenal sympathetic nerve activity and blood pressure in normal rats. Diabetes1997;46:2040–3.

[215] Shek EW, Brands MW, Hall JE. Chronic leptin infusion increases arterial pressure.Hypertension 1998;31:409–14.

[216] de Chantemele EJ Belin, Mintz JD, Rainey WE, Stepp DW. Impact of leptin-mediated sympatho-activation on cardiovascular function in obese mice.Hypertension 2011;58:271–9.

[217] Quehenberger P, Exner M, Sunder-Plassmann R, Ruzicka K, Bieglmayer C, EndlerG, et al. Leptin induces endothelin-1 in endothelial cells in vitro. Circ Res2002;90:711–8.

[218] Adamczak M,Wiecek A, Funahashi T, Chudek J, Kokot F, Matsuzawa Y. Decreasedplasma adiponectin concentration in patients with essential hypertension. Am JHypertens 2003;16:72–5.

[219] Francischetti EA, Celoria BM, Duarte SF, da Silva EG, Santos IJ, Cabello PH, et al.Hypoadiponectinemia is associated with blood pressure increase in obeseinsulin-resistant individuals. Metabolism 2007;56:1464–9.

[220] Iwashima Y, Katsuya T, Ishikawa K, Ouchi N, Ohishi M, Sugimoto K, et al.Hypoadiponectinemia is an independent risk factor for hypertension. Hyper-tension 2004;43:1318–23.

[221] Clasen R, Schupp M, Foryst-Ludwig A, Sprang C, Clemenz M, Krikov M, et al.PPARgamma-activating angiotensin type-1 receptor blockers induce adiponec-tin. Hypertension 2005;46:137–43.

[222] Chen H, Montagnani M, Funahashi T, Shimomura I, Quon MJ. Adiponectinstimulates production of nitric oxide in vascular endothelial cells. J Biol Chem2003;278:45021–6.

[223] Motoshima H, Wu X, Mahadev K, Goldstein BJ. Adiponectin suppressesproliferation and superoxide generation and enhances eNOS activity inendothelial cells treated with oxidized LDL. Biochem Biophys Res Commun2004;315:264–71.

[224] Grekin RJ, Vollmer AP, Sider RS. Pressor effects of portal venous oleate infusion. Aproposed mechanism for obesity hypertension. Hypertension 1995;26:193–8.

[225] Boden G, Shulman GI. Free fatty acids in obesity and type 2 diabetes: definingtheir role in the development of insulin resistance and beta-cell dysfunction. EurJ Clin Invest 2002;32(Suppl 3):14–23.

[226] Safonova I, Aubert J, Negrel R, Ailhaud G. Regulation by fatty acids of angiotensino-gen gene expression in preadipose cells. Biochem J 1997;322(Pt 1):235–9.

[227] Sarafidis PA, Bakris GL. Non-esterified fatty acids and blood pressure elevation: amechanism for hypertension in subjects with obesity/insulin resistance? J HumHypertens 2007;21:12–9.

[228] Bujalska IJ, Kumar S, Stewart PM. Does central obesity reflect “Cushing's diseaseof the omentum”? Lancet 1997;349:1210–3.

[229] Ferrari P, Lovati E, Frey FJ. The role of the 11beta-hydroxysteroid dehydrogenasetype 2 in human hypertension. J Hypertens 2000;18:241–8.

[230] Masuzaki H, Yamamoto H, Kenyon CJ, Elmquist JK, Morton NM, Paterson JM, et al.Transgenic amplification of glucocorticoid action in adipose tissue causes highblood pressure in mice. J Clin Invest 2003;112:83–90.