Skip main navigation

G-Protein–Coupled Receptors in Heart Disease

Originally publishedhttps://doi.org/10.1161/CIRCRESAHA.118.311403Circulation Research. 2018;123:716–735

Abstract

GPCRs (G-protein [guanine nucleotide-binding protein]–coupled receptors) play a central physiological role in the regulation of cardiac function in both health and disease and thus represent one of the largest class of surface receptors targeted by drugs. Several antagonists of GPCRs, such as βARs (β-adrenergic receptors) and Ang II (angiotensin II) receptors, are now considered standard of therapy for a wide range of cardiovascular disease, such as hypertension, coronary artery disease, and heart failure. Although the mechanism of action for GPCRs was thought to be largely worked out in the 80s and 90s, recent discoveries have brought to the fore new and previously unappreciated mechanisms for GPCR activation and subsequent downstream signaling. In this review, we focus on GPCRs most relevant to the cardiovascular system and discuss traditional components of GPCR signaling and highlight evolving concepts in the field, such as ligand bias, β-arrestin–mediated signaling, and conformational heterogeneity.

Heart failure (HF)—a clinical condition resulting from the inability of the heart to sufficiently pump blood to meet the needs of the body—is a prevalent disease leading to high morbidity and mortality. It affects ≈6.5 million American adults and leads to 30 billion USD healthcare expenditure per year.1 Importantly, the 1- and 5-year mortality rate after HF hospitalization approaches 20% and 40%,1 and, therefore, understanding the fundamental mechanisms underlying the development of HF an imperative.

HF is most commonly caused by diseases, such as chronic hypertension, myocardial infarction because of coronary artery disease, cardiomyopathic processes, valvular heart disease, or viral myocarditis.2 These stimuli lead to impaired blood ejection or ventricular filling.3 From a cardiac performance perspective, HF is characterized as HF with reduced ejection fraction (also known as systolic HF) and caused by a failure of cardiac contraction or HF with preserved ejection fraction (also known as diastolic HF) with normal cardiac contraction but impaired filling.3 Both types of HF are associated with structural or functional abnormalities of the left ventricle that is primarily induced by the enlargement of cardiomyocytes as either an increase in cell length, width, or both, referred to as hypertrophic growth. Cardiomyocyte hypertrophy is initiated by a variety of intrinsic and extrinsic stimuli, such as mechanical stress, hormones, cytokines, and growth factors that are sensed by cardiomyocytes through a broad array of receptors on the cell membrane.4 Among the various surface receptors, a class of considerable importance is the family of GPCRs (G-protein [guanine nucleotide-binding protein]–coupled receptors), also known as 7TMRs (7-transmembrane receptors), which represents the largest and the most versatile family of cell surface receptors.5 GPCRs play important roles in a wide variety of physiological processes and, not surprisingly, are commonly targeted for medicinal therapeutics. For instance, chronic activation of βARs (β-adrenergic receptors) and AT1Rs (Ang II [angiotensin II] type 1 receptors) by their endogenous ligands, norepinephrine and Ang II, respectively, increases the workload of the heart, leading to detrimental effects, such as myocyte death and maladaptive cardiac remodeling.6 Therefore, drugs that block activation of these receptors, such as β-blockers, Ang II receptor blockers, and ACE (angiotensin-converting enzyme) inhibitors, are widely used in the treatment of HF. In this article, we will provide an overview of GPCR signaling, highlight the function of cardiac GPCRs in the most important cardiovascular cell types, and provide insights into our new appreciation for the complexity of GPCR signaling that has important implications when considering the development of future HF therapeutics.

Overview of GPCR Signaling

G-Protein–Mediated GPCR Signaling

In the classical paradigm, GPCRs transduce signaling through G proteins. G proteins are named for their ability to bind the nucleotides GTP and GDP. They act as molecular switches in transduction of intracellular signaling: when G proteins are bound to GTP, they are active (on) and when bound to GDP, they are inactive (off).7 Heterotrimeric G proteins are composed of α, β, and γ subunits. When an external signaling molecule binds a GPCR, it causes (or stabilizes) a receptor conformational change, triggering the recruitment of G proteins on the plasma membrane to exchange GDP for GTP on the Gα subunit leading to its activation. GDP-GTP exchange leads to the dissociation of the heterotrimeric G protein into 2 units: the GTP-bound Gα subunit and the dimeric Gβγ complex. Both can interact with a variety of signaling effectors, such as enzymes that produce second messengers and ion channels. The catalytic Gα subunit will hydrolyze the bound GTP to GDP, leading to its reassociation with the Gβγ subunits and termination of the G-protein activation cycle (Figure 1).

Figure 1.

Figure 1. Scheme of GPCR (G-protein–coupled receptor) signaling. On agonist ligand binding, GPCRs interact with heterotrimeric G proteins (guanine nucleotide-binding proteins). G proteins undergo a GDP-GTP exchange on the α subunit, leading to the dissociation of the α and βγ subunits and subsequent activation of downstream signaling effectors. G-protein–activated PKC (protein kinase C) and PKA (protein kinase A) in turn phosphorylates the receptor and turns off the G-protein signaling (heterologous desensitization, red line, and phosphate). GRK (GPCR kinase)-mediated GPCR phosphorylation leads to the recruitment of β-arrestins, resulting in desensitization by sterically interdicting G-protein interaction (homologous desensitization, purple line, and phosphate) and subsequent receptor internalization and ubiquitination. β-arrestin engagement with the receptor also initiates the activation of β-arrestin–mediated signaling. AC indicates adenylate cyclase; AKT, a serine/threonine kinase also known as protein kinase B; DAG, diacylglycerol; EGFR, epidermal growth factor receptor; IP3, inositol-1,4,5-trisphosphate; MAPK, mitogen-activated protein kinase; PI3K, phosphoinositide 3-kinase; PIP2, phosphatidylinositol 4,5-bisphosphate; and PLC, phospholipase C.

Currently, there are 21 Gα subunits, 6 Gβ subunits, and 12 Gγ subunits.7 Heterotrimeric G proteins are typically referred to by their Gα subunits and divided into 4 main classes: Gαstimulatory (Gαs), Gαinhibitory (Gαi), Gαq, and Gα12/13.7 This diversity in G-protein subfamilies allows for their distinct regulatory function in signaling transduction. The Gαs stimulates the effector enzyme, adenylyl cyclase, to produce the second messenger cAMP, leading to activation of PKA (protein kinase A) and phosphorylation of a wide variety of intracellular proteins that regulate cellular responses.8 In contrast, Gαi has an inhibitory effect on adenylyl cyclase, therefore, dampening intracellular cAMP.9q activates PLC (phospholipase C) leading to the cleavage of the membrane-bound phosphatidylinositol 4,5–bisphosphate to the second messengers inositol 1,4,5–triphosphate and diacylglycerol. Inositol 1,4,5–triphosphate promotes Ca2+ release from endoplasmic reticulum. Increased intracellular Ca2+ and diacylglycerol diffused from the plasma membrane will activate PKC (protein kinase C) stimulating cellular signaling.9,10 The G proteins Gα12/13 are known to activate the small GTPase Rho.11

In addition to cell surface receptors, GPCRs localized to other cellular compartments to activate downstream signaling. For instance, the β2AR (β2-adrenergic receptor) activates Gαs to promote cAMP production in early endosomes12; the β1AR (β1-adrenergic receptor) on cardiomyocyte nuclear membrane acts to activate adenylyl cyclase13; ET (endothelin) stimulates the nuclei-localized ET receptors to modulate nuclear Ca2+ concentration14; and the nuclear α1AR (α1-adrenergic receptor) stimulates inside-out signaling to activate ERK (extracellular signal-regulated kinase) localized to plasma membrane caveolae.15

GPCR Desensitization

Although GPCR signaling is essential for normal cellular function, sustained signaling can have detrimental effects on cell survival and maladaptive consequences, such as cardiac failure or tumorigenesis.16,17 Thus, tightly regulated termination of GPCR signaling is critical to maintain normal physiology. Several mechanisms are involved in the termination of receptor signaling, including the abovementioned G-protein inactivation, as well as receptor desensitization that uncouples the receptor from signaling even in the presence of a ligand. The process of receptor desensitization involves phosphorylation of the GPCR at certain serine and threonine residues on the third ICL (intracellular loop) and the carboxy-terminal tail (C terminus) and is termed heterologous and homologous desensitization, respectively.18 Heterologous desensitization is mediated by kinases activated by receptor-triggered second messenger signaling, such as PKA and PKC, which in turn phosphorylate amino acid residues primarily located within the third ICL and proximal C terminus, to uncouple their associated G proteins and thereby terminating signaling.18 PKA-mediated phosphorylation can also switch the receptor coupling to a distinct subtype of G proteins. For instance, PKA-mediated phosphorylation of the β2AR decreases the receptor affinity to Gαs, while promoting receptor coupling to Gαi,19 thus triggering Gαi-mediated pathways and turning off Gαs-activated cAMP production. In addition to inhibition of cAMP production, PKA- or PKC-mediated phosphorylation of cardiac βARs also promote degradation of cAMP by inducing the recruitment of the adaptor protein β-arrestin and PDE4 (phosphodiesterase 4).20 PDE4 is an important enzyme regulating βAR desensitization through hydrolyzing cAMP to maintain equilibrium of cAMP production and degradation under prolonged receptor stimulation21 and through enhancing PKA-mediated βAR phosphorylation and subsequent Gs-Gi switching.22 In contrast, homologous desensitization refers to another key mechanism by which GPCR desensitization occurs and involves receptor phosphorylation through a family of kinases known as GRKs (GPCR kinases).23 There are 7 members in the GRK family. Among them, GRK1 and GRK7 are specifically expressed in retina; GRK4 shows localized expression in spermatozoa and germinal cells, whereas the other 4 members (GRK 2, 3, 5, and 6) are ubiquitously expressed.23 GRKs mediate phosphorylation of serine and threonine residues primarily within the carboxyl tail of agonist-activated GPCRs and promote translocation of β-arrestin to the receptor. In addition to the GPCR-β-arrestin complex formation dependent on receptor tail phosphorylation, β-arrestin can also be activated by its transient engagement with GPCR transmembrane core.24

The binding of β-arrestin sterically interdicts further G-protein coupling to the receptor while promoting internalization of the activated receptor.25 β-arrestins interact with clathrin itself and the clathrin adaptor protein AP2 (adaptor protein 2), therefore, target the attached receptor to the clathrin-coated pits and lead to receptor internalization.26 Internalized receptors in early endosomes are rapidly recycled to the plasma membrane, or undergo prolonged internalization to late endosomes, from where they can be slowly recycled to the cell surface or proceed to lysosomes for degradation.9 Interestingly, recent studies suggest that the internalized GPCRs may continue to activate downstream pathways in endosomes leading to sustained signaling.12,27

β-arrestins may also regulate receptor internalization and degradation by regulating the ubiquitination of the receptors. β-arrestins can directly interact with an array of ubiquitin ligase and deubiquitinases, allowing a finely tuned regulation of the dynamics of ubiquitination/deubiquitination that determines the trafficking destination of the internalized receptors.28

β-Arrestin–Mediated GPCR Signaling and the Concept of Biased Agonism

β-arrestins are members of the arrestin family (arrestin 1–4). Arrestin1 and arrestin4 are expressed in retina, whereas arrestin2 (also named β-arrestin1) and arrestin3 (β-arrestin2) are widely expressed in other tissues.29 The 4 arrestin proteins share a high sequence and structural homology consisting of N- and C-terminal domains built of antiparallel β-strands and intervening loops.30,31 Despite their highly conserved structure, it seems that the 2 β-arrestin isoforms are sufficient to regulate the diverse array of agonist-activated GPCRs.29

β-arrestins are critical regulators of GPCR signaling that not only mediate desensitization and internalization of activated receptors but also function as signal transducers through their function as adaptor/scaffold proteins to activate a broad array of intracellular signaling pathways. For instance, β-arrestins can directly interact with the tyrosine kinase c-Src, leading to the formation of receptor-Src complexes and triggering the activation of ERK.32 They mediate the transactivation of the EGFR (epidermal growth factor receptor) by β1 AR and AT1Rs, which also leads to the ERK activation.33,34 β-arrestin–dependent ERK activation can regulate protein synthesis through MAPK (mitogen-activated protein kinase) interacting MNK1 (serine/threonine kinase 1),35 as well as mediate the antiapoptotic signaling by regulating the phosphorylation of BAD (BCL2-associated agonist of cell death).36 β-arrestins are also involved in cellular processes that initiate PI3K (phosphoinositide 3-kinase) and AKT signaling to lead to various cellular and physiological responses in the context of distinct GPCRs.3739 They inhibit the NF-κB (nuclear factor-κB)–targeted gene expression through binding to and stabilizing IκBα—the inhibitor of NF-κB.40

β-arrestin–mediated signaling seems to be kinetically and functionally divergent from that mediated by G proteins. Whereas G-protein signaling is rapid and transient, β-arrestin signaling is slower and more persistent.41 For example, G-protein–mediated activation of the ERK leads to its translocation into cell nucleus, where it phosphorylates and activates a variety of transcription factors. In contrast, the ERK activated through β-arrestin is retained in the cytosol, phosphorylating a distinct set of substrates and leading to different cellular responses.41,42

The ability of β-arrestins to independently transduce cellular signaling has led to the emerging concept known as biased agonism, which describes the ability of different ligands for one GPCR to activate distinct subsets of downstream signaling events.43 When β-arrestin was originally identified as a mediator of signaling, the prevailing concept at the time was that the binding of ligands to the orthosteric site of a GPCR will signal equally through G-protein– and β-arrestin–dependent pathways, that is, have balanced efficacy toward these two signaling cascades. Under this notion, ligands are classified into full agonists, partial agonists, inverse agonists, and neutral antagonists for both pathways.44 However, accumulating evidence has now led to a refinement of this conceptual framework whereby a given ligand can potentially activate a receptor to selectively engage either a G protein or β-arrestin as its transducer and stimulate only a subset of downstream signaling.44 Indeed, a biased ligand can preferentially activate β-arrestin signaling without activating, or even blocking, the G-protein pathway or vice versa.44 These ligands are, therefore, termed biased agonists (Figure 2).

Figure 2.

Figure 2. Modes of GPCR (G-protein–coupled receptor) signaling. A, Biased agonism of GPCRs and potential clinical implications. Biased agonists selectively activate G-protein– or β-arrestin–mediated signaling pathways. Allosteric modulators bind to distinct sites on the receptor and modulate the activity of orthosteric ligands in various manners. Previous studies suggest that sustained G-protein signaling activated by β1ARs (β1-adrenergic receptors) or AT1Rs (angiotensin II type 1 receptor) is associated with deleterious cardiac effects, whereas β-arrestin signaling may be beneficial for cardiac function. Therefore, β1AR and AT1R β-arrestin–biased agonists and allosteric modulators may block the detrimental G-protein activation while enhancing the cardioprotective effects. B, Scheme of the main features of GPCR structure involved in downstream signaling. Biophysical studies of several GPCRs show how on stimulation certain regions of the receptor are more prone to move allowing the binding of the effectors. In particular, the release of the ionic lock between TM (transmembrane domain) 3 and TM6 is critical for receptor activation; ICL (intracellular loop) 2, ICL3, TM5, and TM6 seem mainly involved in G-protein signaling initiation. β-arrestin binds to the receptor in 2 configurations: (1) interacting with the receptor tail to mediate receptor internalization and β-arrestin signaling or (2) interacting with the receptor transmembrane core to desensitize G-protein signaling.

In contrast to ligands that bind to the orthosteric site on the receptor (the primary site where endogenous ligands bind), allosteric ligands bind to a topographically distinct sites and modulate the effects of orthosteric ligands.45,46 Allosteric modulators can modulate the activity of orthosteric ligands in distinct manners: enhance (positive allosteric modulator), decrease (negative allosteric modulator), simply bind without any effect (silent allosteric modulator), or selectively regulate subsets of downstream signaling (biased allosteric modulator)47 (Figure 2A).

The concept of biased agonism at the level of the receptor and the discovery of allosteric modulators add important dimensions to our understanding of GPCR pharmacology and provide a framework for drug development that can yield improved therapeutic targeting of GPCRs based on ligand-directed selective signaling profiles. Increasingly, biased agonists are being identified for many GPCRs, and many of them are shown to have distinct physiological consequences from balanced agonists.48 For instance, the β-arrestin–biased ligands for AT1R, TRV120023, and TRV120027 have been shown to increase cardiomyocyte contractility, promote cell survival signaling, antagonize Ang II-induced blood pressure increase and cardiac hypertrophy, and improve cardiac performance.4952 However, 2 recent studies have called into question the notion of β-arrestin–biased signaling.53,54 Although β-arrestin siRNA knock down methods have been used in multiple studies to document β-arrestin–mediated signaling, recent studies using CRISPR/Cas9 (clustered regularly interspaced short palindromic repeats/clustered regularly interspaced short palindromic repeat–associated 9) gene editing to abolish β-arrestin1/2 expression in HEK293 cells showed that β-arrestins are dispensable for β2AR-mediated ERK activation.53 Additionally, several previously identified β-arrestin–biased agonists were unable to induce ERK signaling in zero function G cells where G proteins were deleted by gene editing or blocked by an inhibitor.54 Collectively, the authors of these 2 studies concluded that G proteins are the genuine drivers of GPCR signaling, whereas β-arrestins play a negligible role.53,54 To address these seemingly contradictory findings between the studies using CRISPR/Cas9 gene editing and 15 years of previous literature establishing β-arrestin–mediated GPCR signaling in vitro and in vivo, an international consortium of GPCR signaling laboratories comprehensively determined the role of β-arrestin in ERK activation downstream of several GPCRs.55 A range of approaches were used: multiple CRISPR/Cas9 gene-edited cell lines, siRNA knockdown, overexpression, and biased-agonist stimulation. In rigorous fashion, the data showed that siRNA-mediated β-arrestin knockdown attenuates ERK activation, whereas CRISPR/Cas9-mediated β-arrestin knockout has distinct effects depending on the cell lines and receptors tested.55 This study clearly demonstrated that the net effect of β-arrestin on ERK activation is determined by the balance of the inhibitory effect on G-protein signaling through receptor desensitization and the potentiating effect through β-arrestin–mediated pathways, which may vary between cell lines and physiological settings.55 Importantly, prolonged and complete deletion of β-arrestins by CRISPR/Cas9 leads to a cellular rewiring where cell growth and survival depends on the shift toward the activation of G-protein–mediated cellular processes.55

Structural Basis of GPCR Signaling

Despite the large number of family members and the incredible variety of signaling molecules they bind, GPCRs shared common structural motifs. They consist of 7 α-helical TMs (transmembrane domains), hence their 7TMR name, intervening intracellular and extracellular loop regions, an extracellular N terminus, and an intracellular C terminus.56

Remarkable progress has been made in the past decade in understanding the biophysical structure of GPCRs, revealing the 3-dimensional conformation of GPCRs and their signaling complexes. One of the most well characterized is the β2AR. In 2011, an active conformation of β2AR in complex with the agonist BI-167107 and the heterotrimeric Gs protein was solved.57 This structure showed that the intracellular side of the β2AR interacts with the N- and C-terminal α-helices of Gαs. Compared with the inactive state, the β2AR undergoes significant conformational changes on agonist activation: the outward shift of the C terminus of TM5, TM6, and the ICL 3, which forms an interface with the C terminus of the Gαs; the formation of a short helix in ICL 2, the second interface for Gαs binding; the inward shift of the C terminus of TM7 and the N terminus of helix 8, which may be associated with receptor phosphorylation and β-arrestin recruitment57 (Figure 2B). A recent structural study of the calcitonin receptor with cryo-electron microscopy revealed a similar interface of the cytoplasmic side of the receptor TM domains with the Gαs, suggesting a conserved receptor G-protein interface among GPCRs.58

Structural studies have also shed light on the interaction of GPCRs with β-arrestin. The crystal structure of constitutively active rhodopsin in complex with preactivated β-arrestin suggests that β-arrestin first interacts with the C terminus tail of rhodopsin and disrupts the intramolecular interaction of the N and C domain of β-arrestin.59 This promotes the full interaction of β-arrestin with the receptor intracellular pocket formed by the cytoplasmic side of rhodopsin TM domains. Studies of β2AR-β-arrestin complexes reveal that interaction of β-arrestin the receptor likely occurs in a 2-step process.25,60 This study used a chimeric receptor, the β2AR with C terminus replaced with the C tail of vasopressin type 2 receptor, which maintains the β2AR pharmacological properties but with a much higher affinity for β-arrestin.25 Analysis of EM images and 3-D reconstructions revealed that β-arrestin was bound to the receptor in 2 configurations: a weaker interaction involving the phosphorylated C terminus in ≈60% of the complexes and a tighter interaction with finger-loop region of β-arrestin engaged within the transmembrane core of the β2AR (Figure 2B).25 Interestingly, these 2 receptor-β-arrestin conformations appear to carry out distinct functions, with the tail interaction being able to mediate receptor internalization and β-arrestin signaling, whereas engagement with core conformation is likely needed to interdict G-protein activation, that is, desensitization (Figure 2B).61

The discovery of biased agonism implies that receptors can adopt multiple active conformations. Conceptually, GPCRs can be considered as oscillating among a variety of intermediate conformational states.62 Such conformational heterogeneity allows the receptor to differentially interact with distinct ligands and transducers and subsequently signal through a diverse set of pathways. Binding of ligands to the receptor shift the equilibrium of distinct receptor conformations, each ligand preferentially stabilizing unique conformational states that recruit selective signaling transducers.43 Accumulating evidence support the idea that different ligands stabilize distinct receptor conformations. Using fluorescence spectroscopy, it has been shown that ligands with different efficacy have distinct effects on the conformational change of β2AR, represented by the disruption of 2 molecular switches: ionic lock that links the cytoplasmic ends of TM3 and TM6 and whose disruption is a key feature of receptor activation; and a rotamer toggle switch in TM6 that lead to movement of TM6 and receptor activation.63 Using single-molecule fluorescence resonance energy transfer imaging movement of TM6 in the β2AR was monitored in the presence of different orthosteric ligands with varying efficacies (partial or full agonists) and showed that different agonists distinctively affect TM6 motion of Gs-coupled β2ARs highlighting that a β2AR in complex with Gs is structurally unique from that of the nucleotide-free β2ARs-Gs complex.64 The concept of ligand-specific receptor conformations has also been studied using chemical labeling and mass spectrometry and identified unique conformational changes of the β2AR when stabilized by a panel of ligands ranging from full agonists to a β-arrestin–biased agonist.65 These data are consistent with recent studies showing that G-protein– and β-arrestin–biased agonists stabilize distinct conformations for the vasopressin type 2 receptor66 and cholecystokinin-2 receptor.67

GPCRs in Cardiovascular Regulation and HF

GPCRs are widely expressed in the cardiovascular system and in a broad array of cell types, such as cardiomyocytes, fibroblasts, endothelial cells (ECs), and vascular smooth muscle cells (VSMCs). In this section, we will highlight several of the most well-studied and therapeutically targeted GPCRs and their regulation in distinct cardiac cell types.

β Adrenergic Receptors

Among the GPCRs, β1AR and β2AR are the predominant GPCR subtypes expressed in the heart of many mammalian species, including humans, and are the principal regulators of cardiovascular function.68 Under normal physiological conditions, the β1AR is the most abundant βAR subtype in cardiomyocytes, comprising ≈80% of total βARs, whereas the β2AR comprises ≈20%. The stoichiometry of the 2 βAR subtypes changes to ≈60:40 under conditions of HF, mainly caused by the selective downregulation of β1AR expression.69

βARs play an important role in the pathophysiology of human heart disease and are common therapeutic drug targets. βARs are traditionally activated by the catecholamine hormone epinephrine and neurotransmitter norepinephrine. They can also be modulated by a variety of pathways, such as vasopressin,70 insulin,71 TNF-α (tumor necrosis factor-α),72 and prostaglandin E.73 The β1AR primarily couples to Gαs, which activates the signaling effector adenylyl cyclase to promote the second messenger cAMP production and activates PKA, regulating a diverse array of intracellular responses and ultimately an increase in inotropic and chronotropic cardiac function.69 With biased ligand stimulation, the β1AR may selectively engage Gαi to activate β-arrestin–mediated pathways.74 The β2AR also primarily couples to Gαs but can also couple to Gαi through a G-protein switching mechanism induced by PKA-mediated receptor phosphorylation.19 Several studies suggest that excessive β1AR signaling promotes apoptosis of cardiomyocytes, activation of adverse signaling pathways, and exerts detrimental effects on the heart, whereas β2AR signaling has antiapoptotic and cardioprotective effects.7577

β-Blockers are one of the most widely used classes of drugs in numerous conditions, especially in cardiovascular diseases, such as hypertension, postacute myocardial infarction, and HF.78 For example, clinical trials have shown that treatment with the β-blockers, carvedilol, bisoprolol, or metoprolol significantly reduces morbidity and mortality in HF.7981 Treatment with β-blockers normalizes βAR signaling by preventing excessive receptor activation and reversing receptor downregulation and improves left ventricular contractile function. The main cardiovascular use of β-blockers is to block deleterious G-protein overactivation in the heart. The main side effects involve bronchial and blood vessel constriction, which are mainly caused by global inhibition of β2ARs in other tissues.82

The β-blocker carvedilol in particular has been identified as a β-arrestin–biased βAR ligand that preferentially activates β-arrestin–mediated pathways while having inverse agonism toward Gαs signaling.83,84 Carvedilol-stimulated βAR activates a diverse array of signaling events in β-arrestin–dependent manner, including microRNA processing, transactivation of EGFR, and induction of ERK.42,84,85 β-arrestin–mediated EGFR transactivation appears to have a cardioprotective effect as shown by increased apoptosis and cardiac dilation in transgenic mice overexpressing a mutant β1AR lacking GRK phosphorylation sites that cannot recruit β-arrestin and, therefore, unable to transactivate EGFRs.33 Thus, β-arrestin–dependent βAR signaling seems to be beneficial to the heart.

Among the many cell types that compose the mammalian heart, the most abundant are cardiomyocytes, cardiac fibroblasts, ECs, and VSMCs.86 In the following section, we will discuss unique roles of βARs in the various cell types relevant to cardiovascular pathophysiology (Figure 3A).

Figure 3.

Figure 3. Functional roles of GPCRs (G-protein–coupled receptors) in the cardiovascular system. A, Pathophysiological roles of βARs (β-adrenergic receptors) in cardiovascular cells. β1ARs (β1-adrenergic receptors) and β2ARs (β2-adrenergic receptors) are differentially expressed in the cardiovascular system. β1ARs are the predominant βAR subtype in cardiomyocytes, the activation of which increases cardiac contractility and promotes myocyte hypertrophy. β2ARs are most abundant in cardiac fibroblasts, endothelial cells, and vascular smooth muscle cells, where they play important roles in fibrosis and vasodilation. B, Schematic representation of AT1R (angiotensin II type 1 receptor) functions in cardiovascular cell types. AT1Rs regulate a complex array of responses in the cardiovascular system. Chronic AT1R activation promotes hypertrophy, fibrosis, and cardiac remodeling. C, Main effects of the other GPCRs. GPCRs participate at different levels in the regulation of cardiovascular pathophysiology. Each receptor class will have differential effects depending on the receptor subtype, cell type, and mode of stimulation, thus contributing in diverse ways to a specific phenotype. For specific functional roles of individual receptor subtypes, see the Table. αAR indicates α-adrenergic receptor; 5-HT, serotonin receptor; APJ, apelin receptor; AR, adenosine receptor; ECM, extracellular matrix; eNOS, endothelial NO synthase; ETR, endothelin receptor; HR, histamine receptor; MLCK, myosin light-chain kinase; MMP, metalloproteinase; MR, muscarinic receptor; RXFP, relaxin family peptide receptor; S1PR, sphingosine 1-phosphate receptor; VR, vasopressin receptor; and VSMC, vascular smooth muscle cell.

Cardiomyocyte

Cardiomyocytes, accounting for ≈30% of cell number and 70% of cell mass of the mammalian heart, compose the cardiac muscle and provide the contractile force for the heart.86 When stressed, they undergo hypertrophic and apoptotic responses that can lead to cardiomyopathy and ultimately to HF. While both the β1AR and the β2AR primarily couple to Gαs and are present on cardiomyocytes, they perform distinct effects on signaling pathways and cellular responses. The β1AR is the predominant subtype on cardiomyocytes. The stimulation of β1AR on cardiomyocytes increases cardiac contractility through PKA-mediated phosphorylation of several important regulatory proteins of intracellular Ca2+ level or myofilament-Ca2+ sensitivity, such as L-type Ca2+ channels, ryanodine receptor, phospholamban, and troponin I.69,87 The mice with cardiomyocyte-specific overexpression of β1ARs developed myocyte hypertrophy at a young age that progressively developed into HF.88 In contrast, β2AR stimulation exerts inhibitory effect on contractility and protects cardiomyocytes from apoptosis and hypertrophy.75,87 These distinct functional roles of the receptor subtypes may be attribute to 2 aspects: the β2AR-activated Gαi inhibits adenylyl cyclase while it activates PI3K-AKT pathway75 and the differential signaling compartmentation.89 Whereas the effect of β1AR activation is diffusive, the β2AR-induced cAMP accumulation and PKA-mediated protein phosphorylation are localized.89

Fibroblast

Cardiac fibroblasts are the most abundant nonmyocyte cell type in the heart, constituting ≈60% of cell number and 15% of cell mass of the tissue.90 They play key regulatory roles in cardiac remodeling, fibrosis, and hypertrophy through regulating cell proliferation, producing and remodeling extracellular matrix (ECM), as well as generating the autocrine and paracrine signaling molecules.91 The β2AR is the highest expressed βAR subtype in cardiac fibroblasts.91 Stimulation of β2ARs, but not β1ARs, promotes degradation of collagen and induces autophagy.92 β2ARs also induce ERK activation and cell proliferation through EGFR transactivation.91

Endothelial Cell

ECs regulate vascular functions, such as permeability, homeostasis, and angiogenesis.93 Endothelial dysfunction can increase cardiac oxidative stress, promote angiogenesis and fibrosis, and lead to impaired cardiac function in HF.93 β2ARs, but not β1ARs, are expressed in ECs.94 β2AR stimulation activates eNOS (endothelial NO synthase) and subsequently promotes NO-dependent vasodilation.95 In addition, stimulation or overexpression of the β2AR promotes EC proliferation through ERK activation.96

Vascular Smooth Muscle Cell

The VSMCs are involved in the regulation of vascular contractility and the production of ECM molecules.97 On stimulation, the β2AR expressed in VSMCs induces cAMP generation and decreases the activity of MLCK (myosin light-chain kinase)—an enzyme that phosphorylates smooth muscle myosin and enhances contractility.98 Therefore, the cAMP induced by the β2AR in VSMCs decreases contractility and promotes smooth muscle relaxation. It has been reported that β2AR-induced cAMP also inhibits VSMC migration.99

Angiotensin Receptors

Angiotensin receptors, particularly AT1R, play a pivotal role in heart pathophysiology. Cardiac AT1Rs are upregulated in response to hypertrophic stimuli, with ischemia, and promote adverse maladaptive cardiac remodeling in HF.69 Ang II—the endogenous ligand for AT1R—is a peptide hormone that regulates several important physiological processes, such as vascular smooth muscle contraction and aldosterone release. It is also a principle component of the renin-angiotensin system—a key regulatory system controlling blood pressure. AT1Rs couple to the Gαq family of G proteins to transduce signaling but also can mediate signaling through Gβγ subunits and β-arrestin.100,101 Interestingly, recent evidence suggests that for some β-arrestin–dependent signals, AT1Rs need to initially recruit Gαi to the AT1R-β-arrestin complex to fully transduce the signal.101

AT1R overexpression promotes cardiac fibrosis and cardiac hypertrophy, whereas knockout of AT1Rs shows enhanced cardiac function after myocardial infarction, suggesting that the AT1R has detrimental cardiac effects. Considering its central role in cardiovascular physiology, AT1Rs have become an attractive target for drug development for cardiovascular diseases.16 ARBs (angiotensin receptor blockers) and ACE inhibitors are widely used in the treatment of HF. Additionally, recent studies in experimental model systems suggest that selective activation of β-arrestin signaling, while blocking G-protein activation downstream of the AT1R, may provide additional benefit compared with current ARBs.4952 This has led to the development of AT1R-β-arrestin–biased ligands, such as the β-arrestin–biased ligands TRV120027 and TRV120023. In preclinical studies, TRV120027 has been shown to be able to increase cardiac performance in addition to the positive effect on preserving renal function observed also with conventional ARBs.49,50 Like ARBs, TRV120027 promotes vasodilation (via inhibition of the G-protein pathway) but also provides the additional benefit of enhancing cardiac contractility (via β-arrestin signaling). Clinical studies have been performed to evaluate the potential of this molecule102; even though the BLAST-AHF (Biased Ligand of the Angiotensin Receptor Study in Acute Heart Failure) failed in improving clinical status at 30-day follow-up in patients with acute HF,103 biased ligands still remain an attractive prospect for new drug design. Another member of the β-arrestin–biased ligand family, TRV120023, has been shown to inhibit Ang II-induced cardiac hypertrophy52 and promote cardiac contractility51,104 and cardiomyocyte survival after ischemic injury,51 whereas the β-arrestin–biased ligand, TRV120067, can improve cardiac function through the alteration of the myofilament-Ca2+ response in a mouse model of cardiomyopathy.105 All these studies suggest the potential for biased ligands as new therapeutic agents by selectively activating favorable signaling pathways while minimizing untoward activation of G-protein signaling know to be detrimental in cardiovascular disease.

AT1Rs are also known as a mechanosensors. Mechanical stress is a critical regulator of cardiac function and an important stimulus for the development of cardiac hypertrophy.106 Mechanical stress induces a variety of hypertrophic responses, such as regulating the expression of hypertrophic genes and increasing protein synthesis and the activity of multiple protein kinases. Treatment with AT1R blockers attenuates these mechanical stress-induced hypertrophic responses in cardiac myocytes.107 Mechanical stress can activate AT1Rs by promoting autocrine release of Ang II by cardiomyocytes, as well as through Ang II-independent pathways.108,109 Recent studies suggest that mechanical stress directly and specifically activates β-arrestin–biased signaling of AT1Rs, independent of ligands and G proteins, by allosterically stabilizing a unique β-arrestin–biased AT1R conformation.110,111 Biophysical analysis of the AT1R conformation suggests that, when the receptor is activated by mechanical stretch, TM7 of the AT1R undergoes an anticlockwise rotation that shifts it into the ligand-binding pocket.112 A recent study suggests that mechanical stress activates AT1R through a mechanism distinct from that activated by β-arrestin–biased ligands, where the receptors specifically engage Gαi to activate β-arrestin signaling.101 Importantly, mechanoactivated AT1R-mediated β-arrestin–biased signaling seems to be the mechanism by which the Frank-Starling law of the heart uses to generate force.104

While the effects of Ang II on AT1Rs have been well studied, it is now recognized that the AT2R (Ang II type 2 receptor) subtype plays an important role in cardiovascular physiology. AT2R levels are high in fetal tissue but rapidly decrease with aging, suggesting a more prominent role in developmental processes. Evidence suggests a potential role for AT2R in antagonizing AT1R signaling. In rat cardiomyocytes and fibroblasts, AT2R antagonizes AT1R-induced proliferation.113 Ang II-induced signaling that in cardiac fibroblasts activates both AT1R and AT2R to regulate cell growth and collagen secretion and is shown to be altered in HF.114 Nonetheless, the precise role and importance of AT2R in the cardiovascular pathophysiology still remains to be determined.

AT1Rs are expressed in different components of the cardiovascular system (Figure 3B). Depending on the cell type, the stimulated AT1R can elicit the activation of different pathways.115

Cardiomyocyte

In cardiomyocytes, AT1Rs have been shown to affect processes ranging from hypertrophy to apoptosis. Mice overexpressing AT1R specifically in cardiomyocytes display hearts significantly larger than age-matched controls, increased myocyte area, collagen deposition, and atrial natriuretic factor expression.116 Activation of Gαq signaling has been shown to contribute to the hypertrophic phenotypes.117 This phenotype is blocked by the AT1R antagonist losartan.118 Transgenic mice with inducible cardiomyocyte-specific expression of AT1R showed that receptor activation is responsible for blood pressure-independent hypertrophy and cardiac dysfunction.119 Interestingly, it was recently discovered that an endogenous peptide defined as Ang (1–7) (angiotensin [1–7]), previously known to be a natural Ang II competitive inhibitor,120 binds the AT1R acting as a β-arrestin–biased ligand, recruiting and activating β-arrestin to promote ERK phosphorylation.121 In a rat model of cardiac remodeling, Ang (1–7) infusion is sufficient to reduce the left ventricular wall thickness and the end-diastolic pressure.121 While the positive effect of the Ang (1–7) peptide has been demonstrated previously,120 it is now becoming clear that some of its action is derived from its properties as an AT1R-biased ligand.

Ang II-stimulated AT1Rs modulate cardiomyocyte apoptosis through multiple pathways.122,123q activation largely increases the intracellular Ca2+ level, which in turn increases intracellular endonuclease activity.122 In addition, AT1Rs can couple to Gα12/13 to promote the activation of Rho and Rac and, therefore, production of ROS (reactive oxygen species). These will lead to the activation of downstream pathways, such as JNK (c-Jun N-terminal kinase), p38, and MAPK signaling, as well as HSF1 (heat shock factor 1) acetylation and IGF-IIR (insulin-like growth factor II receptor) expression.123,124 Whereas the deleterious effects of AT1Rs on cardiac function seem to be G protein-dependent, the β-arrestin–mediated AT1R pathway confers cardioprotective effects, such as increased cardiomyocyte contractility.

Fibroblast

While cardiac remodeling that occurs as a result of chronic AT1R stimulation is prominently because of effects on cardiomyocytes, fibroblasts also contribute importantly to this process. AT1R activation leads to the transition of cardiac fibroblasts to active myofibroblasts resulting in the production of different components of the ECM, such as collagens, laminin, and fibronectin, and modifying the expression of MMPs (matrix metalloproteinases).125 This effect is, at least in part, regulated by an increase in TGF-β (transforming growth factor-β) production, which in turn promotes the translocation of Smad proteins to the nucleus driving the production of several profibrotic factors.126,127 Mice overexpressing AT1R mutant constructs, which are unable to signal via G proteins, display reduced cardiac fibrosis compared with mice that overexpress WT (wild type) AT1R, thereby indicating that the G-protein pathway contributes to fibrosis.128 However, there is evidence that β-arrestins also contribute to fibrosis in a range of disease settings.129 Although blocking AT1Rs has been shown to inhibit myocardial fibrosis,130 the precise contribution of β-arrestins to the development of fibrosis will require additional studies to evaluate the extent to which AT1R-biased agonists contribute to this process.

Endothelial Cell

In ECs, AT1R activation produces several effects, such as ROS production, increased apoptosis, and elevated thrombosis. Ang II stimulation, indeed, promotes eNOS-dependent NO production131 and exacerbates oxidative stress,132 leading to endothelial dysfunction—a relevant process in the context of hypertension. AT1R activation in ECs promotes microvascular permeability via an increase in intracellular calcium levels and endothelial dysfunction through activation of the calcium-dependent calpains.133 Blocking AT1Rs in ECs is, therefore, an attractive therapeutic option and is supported by data showing that ARBs can improve endothelial dysfunction in hypertensive patients.134

Vascular Smooth Muscle Cell

Changes observed in ECM production by fibroblasts are a key event in vascular remodeling and have implications for VSMC migration and adhesion as well.135 Ang II modulates vascular remodeling by directly stimulating VSMC to produce ECM components.115 In vitro, Ang II-stimulated AT1Rs promote PKC and Jak-STAT pathways, ROS production, and EGFR transactivation,34,101,136 all resulting in the stimulation of the MAPK cascade,137 as well as RhoA, Rho-kinase, c-fos, c-Src, and JNK.138,139 Altogether, these signals lead to a switch from a contractile to a proliferative phenotype and subsequent cell growth.115 The proproliferative effect has also been observed in vivo, where VSMC proliferation induced by vascular injury is strongly reduced in AT1R knockout mice compared with WT and is distinct from the effects mediated by the AT2R.140 Moreover, in both humans and rats, c-Src mediates Ang II contractile action, and this effect is exacerbated in hypertension.141 However, Ang II acts not only through the direct action on AT1Rs but also as an indirect effect on several cell signaling cascades by transactivating other receptor tyrosine kinases (PDGFR [platelet-derived growth factor receptor], EGFR, and IGFR [insulin-like growth factor receptor]). Among these, particularly interesting is the EGFR, which actively contributes to Ang II-induced ERK activation in cardiomyocytes and LV hypertrophy in vivo.142 In vivo, the inhibition of the EGFR transactivation attenuates hypertension and hypertrophy.142 In VSMCs, Ang II-mediated EGFR transactivation modulates hypertrophy and migration.143 Moreover, Ang II stimulation in VSMCs drives a robust proinflammatory response inducing ROS production, cytokine release, and upregulation of adhesion molecules and inflammatory genes.144 Taken together, these studies show the importance of Ang II and AT1Rs in vascular remodeling. Although AT1R-mediated EGFR transactivation seems to be important for cardiomyocyte growth, it may not be a critical in feature driving cardiac fibroblast proliferation.145

α Adrenergic Receptors

Similar to βARs, αARs (α-adrenergic receptors) also bind to and are activated by endogenous catecholamines. The αAR family is composed of α1AR (α1A, α1B, α1D) and α2AR (α2-adrenergic receptor; α2A, α2B, α2C) subfamilies (Table).146 α1ARs primarily couple to Gαq to activate PLC, leading to the generation of second messengers inositol 1,4,5–triphosphate and diacylglycerol that subsequently increase intracellular Ca2+ level.147 All 3 α1AR subtypes are expressed in the heart. Cardiomyocytes express both α1A and α1B, with α1B being predominant.147 α1DAR (α1D-adrenergic receptor) is the major α1AR subtype expressed in coronary smooth muscle cells.148 In response to hypertrophic stimuli or chronical stimulation, the expression of α1A is increased in cardiomyocytes, whereas the levels of α1B and α1D are decreased, with the overall α1AR level unaltered or increased.147 Unlike other Gq-coupled receptors, such as the AT1R of which the overexpression or chronical activation will exert adverse cardiac effect, α1ARs perform important cardioprotective functions, including physiological hypertrophy, increased contractility, and decreased apoptosis.149,150 In response to pressure overload by transverse aortic constriction, mice with double knockout of α1A and α1BAR (α1B-adrenergic receptor) had increased apoptosis, worse dilated cardiomyopathy, and reduced overall survival.151 These data suggest that some caution should be considered when using α1AR antagonists drugs for the treatment of hypertension or prostate disease. Indeed, although the precise mechanism is not entirely clear, the α1AR antagonists doxazosin and prazosin have been associated with increased incidence of HF and mortality in clinical trials.152,153 In contrast to cardiomyocytes, cardiac fibroblasts do not express α1ARs.154 The α1BAR is expressed on coronary artery endothelium cells and may regulate vasodilation and angiogenesis.155

Table. Representative GPCRs in Cardiovascular System

Effects of Receptor Activation on Different Cell Types
ReceptorsTransducersKnown Biased LigandsMain FunctionsExample of Clinical Agents for Cardiovascular DiseasesVSMCsECsFibroblastsCardiomyocytes
αARs
1A, α1B, α1D, α2A, α2B, α2C)
q
1A, α1B, α1D);
i
2A, α2B, α2C)
Desipramine (α2A)160Cardioprotective effect (α1A, α1B, α1D)149,150;
inhibit norepinephrine release (α2A, α2C)157,158
Phenoxybenzamine, doxazosin, terazosin, midodrine, carvedilol, bucindolol, labetalol, indoramin, phentolamine, prazosin, clonidine, methyldopaVasoconstriction (α1D, α2B)148Regulate EC proliferation, angiogenesis, and vasodilation (α1B)155Promote physiological hypertrophy147 and antiapoptosis; increase contractility (α1A, α1B)149,150
Muscarinic receptors
(M1, M2, M3, M4, M5)
q
(M1, M3, M5);
i
(M2, M4)
AF102B
(M1)161; pilocarpine
(M1, M3)162
Heart rate reduction163,164;
reduce intracellular cAMP and Ca2+; negative inotropic and chronotropic effect (M2)165169; stress response (M3)170
Atropine, isoproterenol, forskolin, acetylcholine, carbachol, atropine, cryptenamine; pirenzepine, AF-DX 116, methoctramine, flaxedil, methacholine, gallamine triethiodide,
pilocarpine
Reducing vasoconstriction171Vasodilatation;
intracellular Ca2+ overload (M3); release of NO (M1, M3)
172,173
Increase proliferation and collagen synthesis (M2)174Reduce heart rate; modulate contractility163,165;
induce NO production175
ET receptors
(ETAR, ETBR)
q/GαiBQ-123 (ETAR)176;
ICB2 (ETBR)177
Myocyte hypertrophy178; hypertension176,177,179Sitaxentan, ambrisentan, atrasentan, BQ-123, zibotentan (ETAR);
BQ-788, A192621 (ETBR);
bosentan, macitentan, tezosentan (ETAR, ETBR)
Increase intracellular Ca2+ levels; promote vasoconstriction (ETAR)180Promote NO and prostacyclin production181; vasorelaxant effect (ETBR)182Increase proliferation183 and
collagen synthesis183,184 (ETAR, ETBR)
Increase hypertrophy178
ARs (A1R, A2aR, A2bR, A3R)i
(A1R, A3R);
s
(A2aR, A2bR)
VCP746 (A1R)185;
capadenoson, inosine (A2aR)186,187
Modulate K+ and Ca2+ channels (A1R)163,188; cardioprotective effect (A2aR, A2bR)189192Caffeine, theophylline, theobromine rolofylline, BG-9928, SLV-320; binodenosonModulate cell proliferation and migration and monocyte adhesion (A2bR)193Promote NO production (A1R)194; mediate proliferation, apoptosis, and vasodilatation (A2aR, A2bR, A3R)188,195,196Reduce cell proliferation and collagen synthesis (A2aR)197Contrast hypertrophy, fibrosis, and oxidative stress (A1R)192
Serotonin receptors
(5-HT1A, 5-HT1B, 5-HT1D, 5-HT1E, 5-HT1F, 5-HT2A, 5-HT2B, 5-HT2C, 5-HT4, 5-HT5A, 5-HT5B, 5-HT6, 5-HT7)
i
(5-HT1, 5-HT5);
q
(5-HT2);
s
(5-HT4, 5-HT6, 5-HT7)
Lysergic acid diethylamide, ergotamine (5-HT2B)198Modulate heart rate (5-HT4)199;
vasoconstriction and vasodilation (5-HT1B, 5-HT2A, 5-HT2B, 5-HT4, 5-HT7)199
KetanserinPromote interleukin-6 synthesis (5-HT2A)200Promote NO release (5-HT2B)199Regulate cytokine production, increase oxidative stress (5-HT2B)201,202Differentiation, proliferation, and hypertrophy (5-HT2B)203,204
Histamine receptors
(H1, H2, H3, H4)
q
(H1);
s
(H2);
i
(H3, H4)
Famotidine, ranitidine, tiotidine (H2)205;
JNJ7777120, VUF10214 (H4)206
Vasoconstriction (H1);
increase heart rate and vasodilation (H2);
decrease norepinephrine release (H3)207,208
QuinidineIncrease intracellular Ca2+ levels (H1)209Promote NO release (H2)210Increase calcineurin and promote fibrosis (H2)211Induce positive chronotropic and inotropic responses; promote apoptosis (H2)211,212
APJi, GαqMM07, apelin-17(1–16), [Pyr1]apelin-13213Increase cardiac contractility214;
vasodilation215
Inhibit proliferation; promote apoptosis213Promote NO release213Induce positive inotropy214
Relaxin family peptide receptors (RXFP1, RXFP2, RXFP3, RXFP4)s, Gαi
(RXFP1, RXFP2);
i
(RXFP3, RXFP4)
B7-33, ML290 (RXFP1)216Induces positive inotropy and chronotropy;
vasodilation; angiogenesis;
antifibrotic217
SerelaxinVasodilation (RXFP1)217Promote NO release (RXFP1)217Inhibit proliferation, differentiation, and collagen synthesis (RXFP1)218Antiapoptotic (RXFP1)217
Vasopressin receptors
(V1a, V1b, V2)
q
(V1a, V1b);
s
(V2)
MCF14, SR121463 (V2)
66
Vasoconstriction;
promote fibrosis;
induce cardiomyopathy
(V1a)219
Conivaptan, mozavaptan, tolvaptan, lixivaptan, satavaptanIncrease intracellular Ca2+ levels; promote vasoconstriction (V1a)219Promote cell growth and proliferation (V1a)219Modulate βAR responsiveness and cellular metabolism; induce hypertrophy
(V1a)219
Sphingosine 1-phosphate receptors
(S1PR1, S1PR2, S1PR3, S1PR4, S1PR5)
i
(S1PR1);
s, Gαq, Gα12/13 (S1PR2);
i, Gαq, Gα12/13 (S1PR3);
i, Gα12/13 (S1PR4, S1PR5)
ApoM+ HDL-S1P
(S1PR1)220;
FTY720 (S1PR3)221
Cardioprotective effect
(S1PR2, S1PR3);
promote fibrosis
(S1PR1, S1PR3)
222
Inhibit migration
(S1PR2);
promote vasoconstriction
(S1PR3)
222
Enhance NO production; promote migration and adherens junction assembly (S1PR1);
promote angiogenesis (S1PR3)222
Increase interleukin-6 generation;
promote fibrosis
(S1PR1)223;
promote proliferation (S1PR3)222
Decrease myofilament contractility224;
required for heart development (S1PR1)225

αAR indicates α-adrenergic receptor; A1R, adenosine A1 receptor; A2aR, adenosine A2a receptor; A2bR, adenosine A2b receptor; A3R, adenosine A3 receptor; APJ, apelin receptor; AR, adenosine receptor; ET, endothelin; ETAR, ETA receptor; ETBR, ETB receptor; Gαi, Gαinhibitory;Gαs, Gαstimulatory;RXFP1, relaxin family peptide receptor 1; RXFP2, relaxin family peptide receptor 2; RXFP3, relaxin family peptide receptor 3; RXFP4, relaxin family peptide receptor 4; S1PR1, sphingosine-1-phosphate receptor 1; S1PR2, sphingosine-1-phosphate receptor 2; S1PR3, sphingosine-1-phosphate receptor 3; S1PR4, sphingosine-1-phosphate receptor 4; and S1PR5, sphingosine-1-phosphate receptor 5.

α2ARs couple to Gi to inhibit adenylyl cyclase, reducing the generation of cAMP and decreasing intracellular Ca2+ level. They occur both presynaptically and postsynaptically. Presynaptic α2ARs are key regulators of sympathetic and catecholamine releases.156 In the healthy heart, norepinephrine-activated α2ARs act as a negative feedback to inhibit norepinephrine release.156 Gene deletion of α2AAR (α2A-adrenergic receptor) or α2CAR (α2C-adrenergic receptor), or both of them, in mice leads to elevated norepinephrine levels and worsening cardiac failure.157,158 The α2BAR is present in VSMCs and regulates vasoconstriction.159

Muscarinic Receptors

Muscarinic receptors are ubiquitously expressed in human organs and are involved in several pathophysiological processes, ranging from smooth muscle contraction to glandular secretion.226 As a consequence, the possibility of targeting these receptors for therapeutic purposes has been extensively investigated, including for cardiovascular diseases. There are 5 major subtypes of muscarinic receptor (M1, M2, M3, M4, M5). Among the different subtypes, M1, M3, and M5 signal through Gq, whereas M2 and M4 couple to Gi,227 supporting their different roles in cardiac physiology (Table). In the heart, the predominant isoform is the M2 receptor, and its fundamental function is to mediate parasympathetic signaling, mainly by inhibiting the heart rate response.164 In response to acetylcholine, the activated M2 receptor binds Gi leading to a reduction of intracellular cAMP, reduced If current through the HCN (hyperpolarization-activated cyclic nucleotide-gated) channel,166,167 a decrease in intracellular Ca2+,168 with subsequent negative chronotropic and inotropic effects. Simultaneously, released Gβγ subunits activate the muscarinic-gated potassium channel further promoting the negative chronotropic effects.228

In patients with HF, M2 muscarinic receptor density is upregulated, whereas the other subtypes are unchanged,163,164,169,229 and activated M2 receptors may promote inotropic response through myosin light-chain phosphorylation.230 Functional M3 subtype also seems to play an important role in the heart.170

Muscarinic receptors were one of the first examples of biased signaling. Indeed, since the early 90s different analogs of acetylcholine have been screened for their ability of selectively activating a limited number of signaling pathways downstream the M1 receptor, as a possible therapy for Alzheimer disease.161 Recently, the discovery of mutations in both allosteric and orthosteric binding pockets, along with other sites, has aroused a new interest in the therapeutic potential of these receptors.231,232 In addition, well-known drugs are assuming a new role. An example is given by the pilocarpine—a muscarinic receptor agonist commonly used as cure for glaucoma and dry mouth, which now has been found as a β-arrestin–biased ligand for the M3 receptor.162 Moreover, β-arrestin itself seems to act as molecular switch at the M1 receptor that coordinates not only its desensitization but also diacylglycerol-dependent signaling ending.232 Hence, the use of biased agonist at these receptors seems particularly promising.

Endothelin Receptors

Of the 4 ETs (ET 1–4), ET-1 is the major isoform found in the cardiovascular system. Chronic stimulation with ET-1 is associated with adverse effects ranging from myocyte hypertrophy178 to hypertension179; and high levels of ET-1 in plasma have been found in both patients and experimental models of HF.233,234 ET-1 exerts its effects by binding to either the ETAR (ETA receptor) or the ETBR (ETB receptor; Table); however ETA seems to be the major player in the heart. In both atrial and ventricular tissue, ETARs couple to Gq promoting IP3 formation and activate MAPK signaling.235 In atria, ETARs could also lead to inhibition of adenylyl cyclase, likely through Gi coupling. Also, in human vessels have been found different molecules acting as biased ligands, which could be relevant in hypertension.176,177 Inhibition of ET signaling results in a prolonged survival in experimental HF,236 although ETR (endothelin receptor) blockade in clinical trials with human HF has not been shown to be beneficial.237,238

A recent meta-analysis of clinical trials that tested 4 ETR antagonists, bosentan, sitaxentan, macitentan, or ambrisentan, revealed various significant adverse effects of each drug.239 It is clear that although ETR antagonists confer protective effects on the cardiovascular systems, they also promote various unwanted side effects. The development of drugs that avoid these effects is, therefore, desirable. To this end, the crystal structure of the ETBR bound to bosentan was recently deciphered.240 This study revealed detailed interactions between the two, which are likely conserved in the closely related ETAR. The hope is that a sounder structural understanding of the relationship between this antagonist and the receptor will enable the rational design of therapeutically superior alternatives.

Adenosine Receptors

Adenosine is an adenine base purine and exerts its function in a lot of physiological processes, including those occurring in the cardiovascular system. Adenosine action is mediated via GPCRs known as ARs (adenosine receptors). There are 4 subtypes of ARs: A1R (adenosine A1 receptor), A2aR (adenosine A2a receptor), A2bR (adenosine A2b receptor), and A3R (adenosine A3 receptor; Table).241 Although all ARs bind adenosine, each receptor differs in several respects, including expression in various cell types and the transducers and effectors that they couple to. A1Rs and A3Rs transmit their signal mainly through Gi, inhibiting cAMP production, whereas A2Rs couple primarily to Gs. Gβγ subunits released by AR stimulation also play an important role in cell growth and remodeling, which is not surprising because Gβγ signaling has been shown for other GPCRs.228 A1R interacts with PLC, influencing inositol 1,4,5–triphosphate and Ca2+ release and indirectly modulating K+ and Ca2+ channels.163,188

As shown for other GPCRs, also the ARs display biased signaling, like the biased agonist VCP746, which has no effect on heart rate while reducing hypertrophy in rat,185 or capadenoson—a known A1R partial agonist able to contrast cardiac remodeling in an animal model of HF–has been found acting as a biased ligand for A2BR in both fibroblasts and cardiomyocytes.186 Another example is given by the inosine, which acts as a biased ligand on the A2aR in a unique way prolonging the adenosine stimulation.187

In the vasculature, A2R is considered the main subtype involved in adenosine signaling,188 whereas in the heart, adenosine-induced cardioprotective effects are mainly mediated by A1Rs, whose activation has been shown to protect from I/R injury; to counteract several processes associated with HF, like arrhythmogenesis, fibrosis, apoptosis, hypertrophy, and ventricular dysfunction; and to promote a positive effect on development of HF.189,191,192

A1Rs are considered potential therapeutic targets in HF, and several agonists and antagonists have been tested.242 Different studies have been done in human with A1R antagonists, such as rolofylline and KW-3902243,244; however, despite improved renal function and a positive effect on HF, various side effects, such as increase in stroke, have diminished the impact of this therapy.243 In contrast, A1R agonists have been used successfully to control heart rate for treatment of arrhythmias245 and are safe to use in patients with chronic HF.246 Nowadays, the possibility of using a partial agonist seems a attractive prospect,247 as well as the possibility of discovering a biased ligand, thereby activating the receptor avoiding negative effects because of broader spectrum action. However, considering that A1Rs exert their action mainly in the heart and A2R in the vasculature with a potentially different outcome, it will be important to determine whether a biased ligand shows selectivity for a specific pathway activation, as well as for each receptor subtype.

Other GPCRs

Over 200 GPCRs are expressed in the heart (Table).248 In addition to the receptors discussed above, several others have been identified to have important implications in the development or treatment of cardiovascular disease. For instance, the serotonin 5-HT2b receptor regulates cardiac development and function, and its activation promotes cardiac fibrosis and hypertrophy.204,249 Antagonists of the histamine H2 receptor improved the cardiac function of patients with chronic HF.250 Activation of APJ (apelin receptor) by apelin increases cardiac contractility,251 promotes vasodilation, and enhances cardiac output.252 The peptide hormone relaxin and its receptors exert antifibrosis and cardioprotective roles, and serelaxin (recombinant human relaxin 2) is under development as treatment for acute HF.216 Blockade of vasopressin V2 receptors prevents myocardial dysfunction and renal injury in experimental HF.253 Activation of cardiac sphingosine-1-phosphate receptors has been shown to confer cardioprotective effects.222 The main effects of these GPCRs on heart and blood vessels are outlined in Figure 3C.

Conclusions

GPCRs play vital roles in the physiological regulation of cardiac function and are major drug targets for the treatment of a wide variety of cardiovascular disease. Advances in our understanding of GPCR structure and function have led to a new wave of drug development that aims to enhance receptor specificity and promote distinct ligand-directed signaling efficacies. For instance, biophysical studies of GPCR structures now allow for in silico docking of chemical compounds and structure-based drug design. One of the major challenges of early GPCR structure studies was the instability of the receptor and the transient existence of the receptor-transducer signaling complexes. Recent advances in identifying various formats of GPCR-targeted antibodies to lock receptors in specific conformational states or in complex with signaling partners, such as Fab (antigen-binding fragment) and nanobodies, as well as advances in electron cryomicroscopy (CryoEM) technology now allow for atomic-resolution characterization of heterogeneous structural ensembles. These technological and methodological advances will continue to provide significant insights into the dynamic features of GPCR structure and function. Elucidating the mechanisms for selective activation of GPCR signaling components by biased agonists may yield new drugs that are able to more precisely enhance desired cardioprotective effects while blocking potential untoward detrimental actions or diminishing off target side effects. As new GPCR ligands and mechanisms of action are discovered, many potential therapeutic targets will become evident that one day may be translated into new therapeutic strategies and medicines.

Nonstandard Abbreviations and Acronyms

αAR

α-adrenergic receptor

βAR

β-adrenergic receptor

7TMR

7-transmembrane receptor

A1R

adenosine A1 receptor

A2aR

adenosine A2a receptor

A2bR

adenosine A2b receptor

A3R

adenosine A3 receptor

ACE

angiotensin-converting enzyme

Ang [1–7]

angiotensin [1–7]

Ang II

angiotensin II

APJ

apelin receptor

AR

adenosine receptor

ARB

angiotensin receptor blocker

AT1R

angiotensin II type 1 receptor

AT2R

Ang II type 2 receptor

BAD

BCL2-associated agonist of cell death

BLAST-AHF

Biased Ligand of the Angiotensin Receptor Study in Acute Heart Failure

C terminus

carboxy-terminal tail

CRISPR/Cas9

clustered regularly interspaced short palindromic repeats/clustered regularly interspaced short palindromic repeat–associated 9

EC

endothelial cell

ECM

extracellular matrix

EGFR

epidermal growth factor receptor

eNOS

endothelial NO synthase

ERK

extracellular signal-regulated kinase

ET

endothelin

ETAR

ETA receptor

ETBR

ETB receptor

ETR

endothelin receptor

Fab

antigen-binding fragment

Gαi

Gαinhibitory

Gαs

Gαstimulatory

G protein

guanine nucleotide-binding protein

GPCR

G-protein–coupled receptor

GRK

G-protein–coupled receptor kinase

HF

heart failure

ICL

intracellular loop

MAPK

mitogen-activated protein kinase

MLCK

myosin light-chain kinase

MMP

matrix metalloproteinase

MNK1

serine/threonine kinase 1

NF-κB

nuclear factor-κB

PDE4

phosphodiesterase 4

PI3K

phosphoinositide 3-kinase

PKA

protein kinase A

PKC

protein kinase C

PLC

phospholipase C

TGF-β

transforming growth factor-β

TM

transmembrane domain

TNF-α

tumor necrosis factor-α

VSMC

vascular smooth muscle cell

WT

wild type

Footnotes

Correspondence to Howard A. Rockman, MD, Department of Medicine, Duke University Medical Center, DUMC 3104, 226 CARL Bldg, Durham, NC 27710. Email

References

  • 1. Benjamin EJ, Virani SS, Callaway CWet al; American Heart Association Council on Epidemiology and Prevention Statistics Committee and Stroke Statistics Subcommittee. Heart disease and stroke statistics-2018 update: a report from the American Heart Association.Circulation. 2018; 137:e67–e492. doi: 10.1161/CIR.0000000000000558LinkGoogle Scholar
  • 2. Tanai E, Frantz S. Pathophysiology of heart failure.Compr Physiol. 2015; 6:187–214. doi: 10.1002/cphy.c140055CrossrefMedlineGoogle Scholar
  • 3. Yancy CW, Jessup M, Bozkurt Bet al. 2013 ACCF/AHA guideline for the management of heart failure: executive summary: a report of the American College of Cardiology Foundation/American Heart Association Task Force on practice guidelines.Circulation. 2013; 128:1810–1852. doi: 10.1161/CIR.0b013e31829e8807LinkGoogle Scholar
  • 4. Heineke J, Molkentin JD. Regulation of cardiac hypertrophy by intracellular signalling pathways.Nat Rev Mol Cell Biol. 2006; 7:589–600. doi: 10.1038/nrm1983CrossrefMedlineGoogle Scholar
  • 5. Lagerström MC, Schiöth HB. Structural diversity of G protein-coupled receptors and significance for drug discovery.Nat Rev Drug Discov. 2008; 7:339–357. doi: 10.1038/nrd2518CrossrefMedlineGoogle Scholar
  • 6. Zhu W, Woo AY, Zhang Y, Cao CM, Xiao RP. β-adrenergic receptor subtype signaling in the heart: from bench to the bedside.Curr Top Membr. 2011; 67:191–204. doi: 10.1016/B978-0-12-384921-2.00009-4CrossrefMedlineGoogle Scholar
  • 7. Oldham WM, Hamm HE. Heterotrimeric G protein activation by G-protein-coupled receptors.Nat Rev Mol Cell Biol. 2008; 9:60–71. doi: 10.1038/nrm2299CrossrefMedlineGoogle Scholar
  • 8. Gilman AG. G proteins and dual control of adenylate cyclase.Cell. 1984; 36:577–579.CrossrefMedlineGoogle Scholar
  • 9. Pierce KL, Premont RT, Lefkowitz RJ. Seven-transmembrane receptors.Nat Rev Mol Cell Biol. 2002; 3:639–650. doi: 10.1038/nrm908CrossrefMedlineGoogle Scholar
  • 10. Newton AC. Regulation of protein kinase C.Curr Opin Cell Biol. 1997; 9:161–167.CrossrefMedlineGoogle Scholar
  • 11. Kozasa T, Jiang X, Hart MJ, Sternweis PM, Singer WD, Gilman AG, Bollag G, Sternweis PC. p115 RhoGEF, a GTPase activating protein for Galpha12 and Galpha13.Science. 1998; 280:2109–2111.CrossrefMedlineGoogle Scholar
  • 12. Irannejad R, Tomshine JC, Tomshine JR, Chevalier M, Mahoney JP, Steyaert J, Rasmussen SG, Sunahara RK, El-Samad H, Huang B, von Zastrow M. Conformational biosensors reveal GPCR signalling from endosomes.Nature. 2013; 495:534–538. doi: 10.1038/nature12000CrossrefMedlineGoogle Scholar
  • 13. Boivin B, Lavoie C, Vaniotis G, Baragli A, Villeneuve LR, Ethier N, Trieu P, Allen BG, Hébert TE. Functional beta-adrenergic receptor signalling on nuclear membranes in adult rat and mouse ventricular cardiomyocytes.Cardiovasc Res. 2006; 71:69–78. doi: 10.1016/j.cardiores.2006.03.015CrossrefMedlineGoogle Scholar
  • 14. Boivin B, Chevalier D, Villeneuve LR, Rousseau E, Allen BG. Functional endothelin receptors are present on nuclei in cardiac ventricular myocytes.J Biol Chem. 2003; 278:29153–29163. doi: 10.1074/jbc.M301738200CrossrefMedlineGoogle Scholar
  • 15. Wright CD, Chen Q, Baye NL, Huang Y, Healy CL, Kasinathan S, O’Connell TD. Nuclear alpha1-adrenergic receptors signal activated ERK localization to caveolae in adult cardiac myocytes.Circ Res. 2008; 103:992–1000. doi: 10.1161/CIRCRESAHA.108.176024LinkGoogle Scholar
  • 16. Rockman HA, Koch WJ, Lefkowitz RJ. Seven-transmembrane-spanning receptors and heart function.Nature. 2002; 415:206–212. doi: 10.1038/415206aCrossrefMedlineGoogle Scholar
  • 17. Marinissen MJ, Gutkind JS. G-protein-coupled receptors and signaling networks: emerging paradigms.Trends Pharmacol Sci. 2001; 22:368–376.CrossrefMedlineGoogle Scholar
  • 18. Kelly E, Bailey CP, Henderson G. Agonist-selective mechanisms of GPCR desensitization.Br J Pharmacol. 2008; 153(suppl 1):S379–S388. doi: 10.1038/sj.bjp.0707604CrossrefMedlineGoogle Scholar
  • 19. Daaka Y, Luttrell LM, Lefkowitz RJ. Switching of the coupling of the beta2-adrenergic receptor to different G proteins by protein kinase A.Nature. 1997; 390:88–91. doi: 10.1038/36362CrossrefMedlineGoogle Scholar
  • 20. Shi Q, Li M, Mika D, Fu Q, Kim S, Phan J, Shen A, Vandecasteele G, Xiang YK. Heterologous desensitization of cardiac β-adrenergic signal via hormone-induced βAR/arrestin/PDE4 complexes.Cardiovasc Res. 2017; 113:656–670. doi: 10.1093/cvr/cvx036CrossrefMedlineGoogle Scholar
  • 21. Perry SJ, Baillie GS, Kohout TA, McPhee I, Magiera MM, Ang KL, Miller WE, McLean AJ, Conti M, Houslay MD, Lefkowitz RJ. Targeting of cyclic AMP degradation to beta 2-adrenergic receptors by beta-arrestins.Science. 2002; 298:834–836. doi: 10.1126/science.1074683CrossrefMedlineGoogle Scholar
  • 22. Baillie GS, Sood A, McPhee I, Gall I, Perry SJ, Lefkowitz RJ, Houslay MD. beta-Arrestin-mediated PDE4 cAMP phosphodiesterase recruitment regulates beta-adrenoceptor switching from Gs to Gi.Proc Natl Acad Sci USA. 2003; 100:940–945. doi: 10.1073/pnas.262787199CrossrefMedlineGoogle Scholar
  • 23. Rajagopal S, Shenoy SK. GPCR desensitization: acute and prolonged phases.Cell Signal. 2018; 41:9–16.CrossrefMedlineGoogle Scholar
  • 24. Eichel K, Jullié D, Barsi-Rhyne B, Latorraca NR, Masureel M, Sibarita JB, Dror RO, von Zastrow M. Catalytic activation of β-arrestin by GPCRs.Nature. 2018; 557:381–386. doi: 10.1038/s41586-018-0079-1CrossrefMedlineGoogle Scholar
  • 25. Shukla AK, Westfield GH, Xiao Ket al. Visualization of arrestin recruitment by a G-protein-coupled receptor.Nature. 2014; 512:218–222. doi: 10.1038/nature13430CrossrefMedlineGoogle Scholar
  • 26. Oakley RH, Laporte SA, Holt JA, Barak LS, Caron MG. Association of beta-arrestin with G protein-coupled receptors during clathrin-mediated endocytosis dictates the profile of receptor resensitization.J Biol Chem. 1999; 274:32248–32257.CrossrefMedlineGoogle Scholar
  • 27. Thomsen ARB, Plouffe B, Cahill TJet al. GPCR-G protein-β-arrestin super-complex mediates sustained G protein signaling.Cell. 2016; 166:907–919. doi: 10.1016/j.cell.2016.07.004CrossrefMedlineGoogle Scholar
  • 28. Shenoy SK, Lefkowitz RJ. β-arrestin-mediated receptor trafficking and signal transduction.Trends Pharmacol Sci. 2011; 32:521–533. doi: 10.1016/j.tips.2011.05.002CrossrefMedlineGoogle Scholar
  • 29. Shukla AK, Xiao K, Lefkowitz RJ. Emerging paradigms of β-arrestin-dependent seven transmembrane receptor signaling.Trends Biochem Sci. 2011; 36:457–469. doi: 10.1016/j.tibs.2011.06.003CrossrefMedlineGoogle Scholar
  • 30. Hirsch JA, Schubert C, Gurevich VV, Sigler PB. The 2.8 A crystal structure of visual arrestin: a model for arrestin’s regulation.Cell. 1999; 97:257–269.CrossrefMedlineGoogle Scholar
  • 31. Han M, Gurevich VV, Vishnivetskiy SA, Sigler PB, Schubert C. Crystal structure of beta-arrestin at 1.9 A: possible mechanism of receptor binding and membrane Translocation.Structure. 2001; 9:869–880.CrossrefMedlineGoogle Scholar
  • 32. Luttrell LM, Ferguson SS, Daaka Y, Miller WE, Maudsley S, Della Rocca GJ, Lin F, Kawakatsu H, Owada K, Luttrell DK, Caron MG, Lefkowitz RJ. Beta-arrestin-dependent formation of beta2 adrenergic receptor-Src protein kinase complexes.Science. 1999; 283:655–661.CrossrefMedlineGoogle Scholar
  • 33. Noma T, Lemaire A, Naga Prasad SV, Barki-Harrington L, Tilley DG, Chen J, Le Corvoisier P, Violin JD, Wei H, Lefkowitz RJ, Rockman HA. Beta-arrestin-mediated beta1-adrenergic receptor transactivation of the EGFR confers cardioprotection.J Clin Invest. 2007; 117:2445–2458. doi: 10.1172/JCI31901CrossrefMedlineGoogle Scholar
  • 34. Kim J, Ahn S, Rajagopal K, Lefkowitz RJ. Independent beta-arrestin2 and Gq/protein kinase Czeta pathways for ERK stimulated by angiotensin type 1A receptors in vascular smooth muscle cells converge on transactivation of the epidermal growth factor receptor.J Biol Chem. 2009; 284:11953–11962. doi: 10.1074/jbc.M808176200CrossrefMedlineGoogle Scholar
  • 35. DeWire SM, Kim J, Whalen EJ, Ahn S, Chen M, Lefkowitz RJ. Beta-arrestin-mediated signaling regulates protein synthesis.J Biol Chem. 2008; 283:10611–10620. doi: 10.1074/jbc.M710515200CrossrefMedlineGoogle Scholar
  • 36. Ahn S, Kim J, Hara MR, Ren XR, Lefkowitz RJ. {beta}-arrestin-2 mediates anti-apoptotic signaling through regulation of BAD phosphorylation.J Biol Chem. 2009; 284:8855–8865. doi: 10.1074/jbc.M808463200CrossrefMedlineGoogle Scholar
  • 37. Beaulieu JM, Sotnikova TD, Marion S, Lefkowitz RJ, Gainetdinov RR, Caron MG. An Akt/beta-arrestin 2/PP2A signaling complex mediates dopaminergic neurotransmission and behavior.Cell. 2005; 122:261–273. doi: 10.1016/j.cell.2005.05.012CrossrefMedlineGoogle Scholar
  • 38. Walters RW, Shukla AK, Kovacs JJ, Violin JD, DeWire SM, Lam CM, Chen JR, Muehlbauer MJ, Whalen EJ, Lefkowitz RJ. beta-Arrestin1 mediates nicotinic acid-induced flushing, but not its antilipolytic effect, in mice.J Clin Invest. 2009; 119:1312–1321. doi: 10.1172/JCI36806CrossrefMedlineGoogle Scholar
  • 39. Watson LJ, Alexander KM, Mohan ML, Bowman AL, Mangmool S, Xiao K, Naga Prasad SV, Rockman HA. Phosphorylation of Src by phosphoinositide 3-kinase regulates beta-adrenergic receptor-mediated EGFR transactivation.Cell Signal. 2016; 28:1580–1592. doi: 10.1016/j.cellsig.2016.05.006CrossrefMedlineGoogle Scholar
  • 40. Witherow DS, Garrison TR, Miller WE, Lefkowitz RJ. beta-Arrestin inhibits NF-kappaB activity by means of its interaction with the NF-kappaB inhibitor IkappaBalpha.Proc Natl Acad Sci USA. 2004; 101:8603–8607. doi: 10.1073/pnas.0402851101CrossrefMedlineGoogle Scholar
  • 41. Ahn S, Shenoy SK, Wei H, Lefkowitz RJ. Differential kinetic and spatial patterns of beta-arrestin and G protein-mediated ERK activation by the angiotensin II receptor.J Biol Chem. 2004; 279:35518–35525. doi: 10.1074/jbc.M405878200CrossrefMedlineGoogle Scholar
  • 42. Tilley DG, Kim IM, Patel PA, Violin JD, Rockman HA. beta-Arrestin mediates beta1-adrenergic receptor-epidermal growth factor receptor interaction and downstream signaling.J Biol Chem. 2009; 284:20375–20386. doi: 10.1074/jbc.M109.005793CrossrefMedlineGoogle Scholar
  • 43. Rajagopal S, Rajagopal K, Lefkowitz RJ. Teaching old receptors new tricks: biasing seven-transmembrane receptors.Nat Rev Drug Discov. 2010; 9:373–386. doi: 10.1038/nrd3024CrossrefMedlineGoogle Scholar
  • 44. Violin JD, Lefkowitz RJ. Beta-arrestin-biased ligands at seven-transmembrane receptors.Trends Pharmacol Sci. 2007; 28:416–422. doi: 10.1016/j.tips.2007.06.006CrossrefMedlineGoogle Scholar
  • 45. Ahn S, Pani B, Kahsai AWet al. Small-molecule positive allosteric modulators of the β2-adrenoceptor isolated from DNA-encoded libraries.Mol Pharmacol. 2018; 94:850–861. doi: 10.1124/mol.118.111948CrossrefMedlineGoogle Scholar
  • 46. Liu X, Ahn S, Kahsai AW, Meng KC, Latorraca NR, Pani B, Venkatakrishnan AJ, Masoudi A, Weis WI, Dror RO, Chen X, Lefkowitz RJ, Kobilka BK. Mechanism of intracellular allosteric β2AR antagonist revealed by X-ray crystal structure.Nature. 2017; 548:480–484. doi: 10.1038/nature23652CrossrefMedlineGoogle Scholar
  • 47. Wisler JW, Rockman HA, Lefkowitz RJ. Biased G protein-coupled receptor signaling: changing the paradigm of drug discovery.Circulation. 2018; 137:2315–2317. doi: 10.1161/CIRCULATIONAHA.117.028194LinkGoogle Scholar
  • 48. Violin JD, Soergel DG, Boerrigter G, Burnett JC, Lark MW. GPCR biased ligands as novel heart failure therapeutics.Trends Cardiovasc Med. 2013; 23:242–249. doi: 10.1016/j.tcm.2013.01.002CrossrefMedlineGoogle Scholar
  • 49. Violin JD, DeWire SM, Yamashita D, Rominger DH, Nguyen L, Schiller K, Whalen EJ, Gowen M, Lark MW. Selectively engaging β-arrestins at the angiotensin II type 1 receptor reduces blood pressure and increases cardiac performance.J Pharmacol Exp Ther. 2010; 335:572–579. doi: 10.1124/jpet.110.173005CrossrefMedlineGoogle Scholar
  • 50. Boerrigter G, Lark MW, Whalen EJ, Soergel DG, Violin JD, Burnett JC. Cardiorenal actions of TRV120027, a novel ß-arrestin-biased ligand at the angiotensin II type I receptor, in healthy and heart failure canines: a novel therapeutic strategy for acute heart failure.Circ Heart Fail. 2011; 4:770–778. doi: 10.1161/CIRCHEARTFAILURE.111.962571LinkGoogle Scholar
  • 51. Kim KS, Abraham D, Williams B, Violin JD, Mao L, Rockman HA. β-Arrestin-biased AT1R stimulation promotes cell survival during acute cardiac injury.Am J Physiol Heart Circ Physiol. 2012; 303:H1001–H1010. doi: 10.1152/ajpheart.00475.2012CrossrefMedlineGoogle Scholar
  • 52. Monasky MM, Taglieri DM, Henze M, Warren CM, Utter MS, Soergel DG, Violin JD, Solaro RJ. The β-arrestin-biased ligand TRV120023 inhibits angiotensin II-induced cardiac hypertrophy while preserving enhanced myofilament response to calcium.Am J Physiol Heart Circ Physiol. 2013; 305:H856–H866. doi: 10.1152/ajpheart.00327.2013CrossrefMedlineGoogle Scholar
  • 53. O’Hayre M, Eichel K, Avino S, Zhao X, Steffen DJ, Feng X, Kawakami K, Aoki J, Messer K, Sunahara R, Inoue A, von Zastrow M, Gutkind JS. Genetic evidence that beta-arrestins are dispensable for the initiation of beta2-adrenergic receptor signaling to ERK.Sci Signal. 2017; 10:eaal3395.CrossrefMedlineGoogle Scholar
  • 54. Grundmann M, Merten N, Malfacini Det al. Lack of beta-arrestin signaling in the absence of active G proteins.Nat Commun. 2018; 9:341. doi: 10.1038/s41467-017-02661-3CrossrefMedlineGoogle Scholar
  • 55. Luttrell LM, Wang J, Plouffe Bet al. Manifold roles of β-arrestins in GPCR signaling elucidated with siRNA and CRISPR/Cas9.Sci Signal. 2018; 11:eaat7650. doi: 10.1126/scisignal.aat7650.CrossrefMedlineGoogle Scholar
  • 56. Rosenbaum DM, Rasmussen SG, Kobilka BK. The structure and function of G-protein-coupled receptors.Nature. 2009; 459:356–363. doi: 10.1038/nature08144CrossrefMedlineGoogle Scholar
  • 57. Rasmussen SG, DeVree BT, Zou Yet al. Crystal structure of the β2 adrenergic receptor-Gs protein complex.Nature. 2011; 477:549–555. doi: 10.1038/nature10361CrossrefMedlineGoogle Scholar
  • 58. Liang YL, Khoshouei M, Radjainia Met al. Phase-plate cryo-EM structure of a class B GPCR-G-protein complex.Nature. 2017; 546:118–123. doi: 10.1038/nature22327CrossrefMedlineGoogle Scholar
  • 59. Kang Y, Zhou XE, Gao Xet al. Crystal structure of rhodopsin bound to arrestin by femtosecond X-ray laser.Nature. 2015; 523:561–567. doi: 10.1038/nature14656CrossrefMedlineGoogle Scholar
  • 60. Shukla AK, Manglik A, Kruse ACet al. Structure of active β-arrestin-1 bound to a G-protein-coupled receptor phosphopeptide.Nature. 2013; 497:137–141. doi: 10.1038/nature12120CrossrefMedlineGoogle Scholar
  • 61. Cahill TJ, Thomsen AR, Tarrasch JTet al. Distinct conformations of GPCR-β-arrestin complexes mediate desensitization, signaling, and endocytosis.Proc Natl Acad Sci USA. 2017; 114:2562–2567. doi: 10.1073/pnas.1701529114CrossrefMedlineGoogle Scholar
  • 62. Nygaard R, Zou Y, Dror ROet al. The dynamic process of β(2)-adrenergic receptor activation.Cell. 2013; 152:532–542. doi: 10.1016/j.cell.2013.01.008CrossrefMedlineGoogle Scholar
  • 63. Yao X, Parnot C, Deupi X, Ratnala VR, Swaminath G, Farrens D, Kobilka B. Coupling ligand structure to specific conformational switches in the beta2-adrenoceptor.Nat Chem Biol. 2006; 2:417–422. doi: 10.1038/nchembio801CrossrefMedlineGoogle Scholar
  • 64. Gregorio GG, Masureel M, Hilger D, Terry DS, Juette M, Zhao H, Zhou Z, Perez-Aguilar JM, Hauge M, Mathiasen S, Javitch JA, Weinstein H, Kobilka BK, Blanchard SC. Single-molecule analysis of ligand efficacy in β2AR-G-protein activation.Nature. 2017; 547:68–73. doi: 10.1038/nature22354CrossrefMedlineGoogle Scholar
  • 65. Kahsai AW, Xiao K, Rajagopal S, Ahn S, Shukla AK, Sun J, Oas TG, Lefkowitz RJ. Multiple ligand-specific conformations of the β2-adrenergic receptor.Nat Chem Biol. 2011; 7:692–700. doi: 10.1038/nchembio.634CrossrefMedlineGoogle Scholar
  • 66. Rahmeh R, Damian M, Cottet Met al. Structural insights into biased G protein-coupled receptor signaling revealed by fluorescence spectroscopy.Proc Natl Acad Sci USA. 2012; 109:6733–6738. doi: 10.1073/pnas.1201093109CrossrefMedlineGoogle Scholar
  • 67. Magnan R, Escrieut C, Gigoux V, De K, Clerc P, Niu F, Azema J, Masri B, Cordomi A, Baltas M, Tikhonova IG, Fourmy D. Distinct CCK-2 receptor conformations associated with β-arrestin-2 recruitment or phospholipase-C activation revealed by a biased antagonist.J Am Chem Soc. 2013; 135:2560–2573. doi: 10.1021/ja308784wCrossrefMedlineGoogle Scholar
  • 68. Woo AY, Xiao RP. β-Adrenergic receptor subtype signaling in heart: from bench to bedside.Acta Pharmacol Sin. 2012; 33:335–341. doi: 10.1038/aps.2011.201CrossrefMedlineGoogle Scholar
  • 69. Salazar NC, Chen J, Rockman HA. Cardiac GPCRs: GPCR signaling in healthy and failing hearts.Biochim Biophys Acta. 2007; 1768:1006–1018. doi: 10.1016/j.bbamem.2007.02.010CrossrefMedlineGoogle Scholar
  • 70. Tilley DG, Zhu W, Myers VDet al. β-adrenergic receptor-mediated cardiac contractility is inhibited via vasopressin type 1A-receptor-dependent signaling.Circulation. 2014; 130:1800–1811. doi: 10.1161/CIRCULATIONAHA.114.010434LinkGoogle Scholar
  • 71. Wang Q, Liu Y, Fu Q, Xu B, Zhang Y, Kim S, Tan R, Barbagallo F, West T, Anderson E, Wei W, Abel ED, Xiang YK. Inhibiting insulin-mediated beta2-adrenergic receptor activation prevents diabetes-associated cardiac dysfunction.Circulation. 2017; 135:73–88.LinkGoogle Scholar
  • 72. Vasudevan NT, Mohan ML, Gupta MK, Martelli EE, Hussain AK, Qin Y, Chandrasekharan UM, Young D, Feldman AM, Sen S, Dorn GW, Dicorleto PE, Naga Prasad SV. Gβγ-independent recruitment of G-protein coupled receptor kinase 2 drives tumor necrosis factor α-induced cardiac β-adrenergic receptor dysfunction.Circulation. 2013; 128:377–387. doi: 10.1161/CIRCULATIONAHA.113.003183LinkGoogle Scholar
  • 73. Liu S, Li Y, Kim S, Fu Q, Parikh D, Sridhar B, Shi Q, Zhang X, Guan Y, Chen X, Xiang YK. Phosphodiesterases coordinate cAMP propagation induced by two stimulatory G protein-coupled receptors in hearts.Proc Natl Acad Sci USA. 2012; 109:6578–6583. doi: 10.1073/pnas.1117862109CrossrefMedlineGoogle Scholar
  • 74. Wang J, Hanada K, Staus DP, Makara MA, Dahal GR, Chen Q, Ahles A, Engelhardt S, Rockman HA. Gαi is required for carvedilol-induced β1 adrenergic receptor β-arrestin biased signaling.Nat Commun. 2017; 8:1706. doi: 10.1038/s41467-017-01855-zCrossrefMedlineGoogle Scholar
  • 75. Communal C, Singh K, Sawyer DB, Colucci WS. Opposing effects of beta(1)-and beta(2)-adrenergic receptors on cardiac myocyte apoptosis: role of a pertussis toxin-sensitive G protein.Circulation. 1999; 100:2210–2212.LinkGoogle Scholar
  • 76. Chesley A, Lundberg MS, Asai T, Xiao RP, Ohtani S, Lakatta EG, Crow MT. The beta(2)-adrenergic receptor delivers an antiapoptotic signal to cardiac myocytes through G(i)-dependent coupling to phosphatidylinositol 3’-kinase.Circ Res. 2000; 87:1172–1179.LinkGoogle Scholar
  • 77. Yoo B, Lemaire A, Mangmool S, Wolf MJ, Curcio A, Mao L, Rockman HA. Beta1-adrenergic receptors stimulate cardiac contractility and CaMKII activation in vivo and enhance cardiac dysfunction following myocardial infarction.Am J Physiol Heart Circ Physiol. 2009; 297:H1377–H1386. doi: 10.1152/ajpheart.00504.2009CrossrefMedlineGoogle Scholar
  • 78. Baker JG. The selectivity of beta-adrenoceptor antagonists at the human beta1, beta2 and beta3 adrenoceptors.Br J Pharmacol. 2005; 144:317–322. doi: 10.1038/sj.bjp.0706048CrossrefMedlineGoogle Scholar
  • 79. Packer M, Bristow MR, Cohn JN, Colucci WS, Fowler MB, Gilbert EM, Shusterman NH. The effect of carvedilol on morbidity and mortality in patients with chronic heart failure. U.S. Carvedilol Heart Failure Study Group.N Engl J Med. 1996; 334:1349–1355. doi: 10.1056/NEJM199605233342101CrossrefMedlineGoogle Scholar
  • 80. Simon T, Mary-Krause M, Funck-Brentano C, Lechat P, Jaillon P. Bisoprolol dose-response relationship in patients with congestive heart failure: a subgroup analysis in the cardiac insufficiency bisoprolol study(CIBIS II).Eur Heart J. 2003; 24:552–559.CrossrefMedlineGoogle Scholar
  • 81. Kukin ML, Kalman J, Charney RH, Levy DK, Buchholz-Varley C, Ocampo ON, Eng C. Prospective, randomized comparison of effect of long-term treatment with metoprolol or carvedilol on symptoms, exercise, ejection fraction, and oxidative stress in heart failure.Circulation. 1999; 99:2645–2651.LinkGoogle Scholar
  • 82. Lewis RV, Lofthouse C. Adverse reactions with beta-adrenoceptor blocking drugs. An update.Drug Saf. 1993; 9:272–279.CrossrefMedlineGoogle Scholar
  • 83. Wisler JW, DeWire SM, Whalen EJ, Violin JD, Drake MT, Ahn S, Shenoy SK, Lefkowitz RJ. A unique mechanism of beta-blocker action: carvedilol stimulates beta-arrestin signaling.Proc Natl Acad Sci USA. 2007; 104:16657–16662. doi: 10.1073/pnas.0707936104CrossrefMedlineGoogle Scholar
  • 84. Kim IM, Tilley DG, Chen J, Salazar NC, Whalen EJ, Violin JD, Rockman HA. Beta-blockers alprenolol and carvedilol stimulate beta-arrestin-mediated EGFR transactivation.Proc Natl Acad Sci USA. 2008; 105:14555–14560. doi: 10.1073/pnas.0804745105CrossrefMedlineGoogle Scholar
  • 85. Kim IM, Wang Y, Park KM, Tang Y, Teoh JP, Vinson J, Traynham CJ, Pironti G, Mao L, Su H, Johnson JA, Koch WJ, Rockman HA. β-arrestin1-biased β1-adrenergic receptor signaling regulates microRNA processing.Circ Res. 2014; 114:833–844. doi: 10.1161/CIRCRESAHA.114.302766LinkGoogle Scholar
  • 86. Zhou P, Pu WT. Recounting cardiac cellular composition.Circ Res. 2016; 118:368–370. doi: 10.1161/CIRCRESAHA.116.308139LinkGoogle Scholar
  • 87. Devic E, Xiang Y, Gould D, Kobilka B. Beta-adrenergic receptor subtype-specific signaling in cardiac myocytes from beta(1) and beta(2) adrenoceptor knockout mice.Mol Pharmacol. 2001; 60:577–583.MedlineGoogle Scholar
  • 88. Engelhardt S, Hein L, Wiesmann F, Lohse MJ. Progressive hypertrophy and heart failure in beta1-adrenergic receptor transgenic mice.Proc Natl Acad Sci USA. 1999; 96:7059–7064.CrossrefMedlineGoogle Scholar
  • 89. Xiao RP, Cheng H, Zhou YY, Kuschel M, Lakatta EG. Recent advances in cardiac beta(2)-adrenergic signal transduction.Circ Res. 1999; 85:1092–1100.LinkGoogle Scholar
  • 90. MacKenna D, Summerour SR, Villarreal FJ. Role of mechanical factors in modulating cardiac fibroblast function and extracellular matrix synthesis.Cardiovasc Res. 2000; 46:257–263.CrossrefMedlineGoogle Scholar
  • 91. Kim J, Eckhart AD, Eguchi S, Koch WJ. Beta-adrenergic receptor-mediated DNA synthesis in cardiac fibroblasts is dependent on transactivation of the epidermal growth factor receptor and subsequent activation of extracellular signal-regulated kinases.J Biol Chem. 2002; 277:32116–32123. doi: 10.1074/jbc.M204895200CrossrefMedlineGoogle Scholar
  • 92. Aránguiz-Urroz P, Canales J, Copaja M, Troncoso R, Vicencio JM, Carrillo C, Lara H, Lavandero S, Díaz-Araya G. Beta(2)-adrenergic receptor regulates cardiac fibroblast autophagy and collagen degradation.Biochim Biophys Acta. 2011; 1812:23–31. doi: 10.1016/j.bbadis.2010.07.003CrossrefMedlineGoogle Scholar
  • 93. Shantsila E, Wrigley BJ, Blann AD, Gill PS, Lip GY. A contemporary view on endothelial function in heart failure.Eur J Heart Fail. 2012; 14:873–881. doi: 10.1093/eurjhf/hfs066CrossrefMedlineGoogle Scholar
  • 94. Broeders MA, Doevendans PA, Bekkers BC, Bronsaer R, van Gorsel E, Heemskerk JW, Egbrink MG, van Breda E, Reneman RS, van Der Zee R. Nebivolol: a third-generation beta-blocker that augments vascular nitric oxide release: endothelial beta(2)-adrenergic receptor-mediated nitric oxide production.Circulation. 2000; 102:677–684.LinkGoogle Scholar
  • 95. Iaccarino G, Cipolletta E, Fiorillo A, Annecchiarico M, Ciccarelli M, Cimini V, Koch WJ, Trimarco B. Beta(2)-adrenergic receptor gene delivery to the endothelium corrects impaired adrenergic vasorelaxation in hypertension.Circulation. 2002; 106:349–355.LinkGoogle Scholar
  • 96. Iaccarino G, Ciccarelli M, Sorriento Det al. Ischemic neoangiogenesis enhanced by beta2-adrenergic receptor overexpression: a novel role for the endothelial adrenergic system.Circ Res. 2005; 97:1182–1189. doi: 10.1161/01.RES.0000191541.06788.bbLinkGoogle Scholar
  • 97. Michel JB, Li Z, Lacolley P. Smooth muscle cells and vascular diseases.Cardiovasc Res. 2012; 95:135–137. doi: 10.1093/cvr/cvs172CrossrefMedlineGoogle Scholar
  • 98. Horman S, Morel N, Vertommen D, Hussain N, Neumann D, Beauloye C, El Najjar N, Forcet C, Viollet B, Walsh MP, Hue L, Rider MH. AMP-activated protein kinase phosphorylates and desensitizes smooth muscle myosin light chain kinase.J Biol Chem. 2008; 283:18505–18512. doi: 10.1074/jbc.M802053200CrossrefMedlineGoogle Scholar
  • 99. Johnson R, Webb JG, Newman WH, Wang Z. Regulation of human vascular smooth muscle cell migration by beta-adrenergic receptors.Am Surg. 2006; 72:51–54.CrossrefMedlineGoogle Scholar
  • 100. Ushio-Fukai M, Griendling KK, Akers M, Lyons PR, Alexander RW. Temporal dispersion of activation of phospholipase C-beta1 and –-gamma isoforms by angiotensin II in vascular smooth muscle cells. Role of alphaq/11, alpha12, and beta gamma G protein subunits.J Biol Chem. 1998; 273:19772–19777.CrossrefMedlineGoogle Scholar
  • 101. Wang J, Hanada K, Gareri C, Rockman HA. Mechanoactivation of the angiotensin II type 1 receptor induces β-arrestin-biased signaling through Gαi coupling.J Cell Biochem. 2018; 119:3586–3597. doi: 10.1002/jcb.26552CrossrefMedlineGoogle Scholar
  • 102. Felker GM, Butler J, Collins SP, Cotter G, Davison BA, Ezekowitz JA, Filippatos G, Levy PD, Metra M, Ponikowski P, Soergel DG, Teerlink JR, Violin JD, Voors AA, Pang PS. Heart failure therapeutics on the basis of a biased ligand of the angiotensin-2 type 1 receptor. Rationale and design of the BLAST-AHF study (Biased Ligand of the Angiotensin Receptor Study in Acute Heart Failure).JACC Heart Fail. 2015; 3:193–201. doi: 10.1016/j.jchf.2014.09.008CrossrefMedlineGoogle Scholar
  • 103. Pang PS, Butler J, Collins SPet al. Biased ligand of the angiotensin II type 1 receptor in patients with acute heart failure: a randomized, double-blind, placebo-controlled, phase IIB, dose ranging trial (BLAST-AHF).Eur Heart J. 2017; 38:2364–2373. doi: 10.1093/eurheartj/ehx196CrossrefMedlineGoogle Scholar
  • 104. Abraham DM, Davis RT, Warren CM, Mao L, Wolska BM, Solaro RJ, Rockman HA. β-Arrestin mediates the Frank-Starling mechanism of cardiac contractility.Proc Natl Acad Sci USA. 2016; 113:14426–14431. doi: 10.1073/pnas.1609308113CrossrefMedlineGoogle Scholar
  • 105. Ryba DM, Li J, Cowan CL, Russell B, Wolska BM, Solaro RJ. Long-term biased β-arrestin signaling improves cardiac structure and function in dilated cardiomyopathy.Circulation. 2017; 135:1056–1070. doi: 10.1161/CIRCULATIONAHA.116.024482LinkGoogle Scholar
  • 106. Dostal DE, Feng H, Nizamutdinov D, Golden HB, Afroze SH, Dostal JD, Jacob JC, Foster DM, Tong C, Glaser S, Gerilechaogetu F. Mechanosensing and regulation of cardiac function.J Clin Exp Cardiolog. 2014; 5:314. doi: 10.4172/2155-9880.1000314MedlineGoogle Scholar
  • 107. Yamazaki T, Komuro I, Kudoh S, Zou Y, Shiojima I, Mizuno T, Takano H, Hiroi Y, Ueki K, Tobe K. Angiotensin II partly mediates mechanical stress-induced cardiac hypertrophy.Circ Res. 1995; 77:258–265.LinkGoogle Scholar
  • 108. Sadoshima J, Xu Y, Slayter HS, Izumo S. Autocrine release of angiotensin II mediates stretch-induced hypertrophy of cardiac myocytes in vitro.Cell. 1993; 75:977–984.CrossrefMedlineGoogle Scholar
  • 109. Zou Y, Akazawa H, Qin Yet al. Mechanical stress activates angiotensin II type 1 receptor without the involvement of angiotensin II.Nat Cell Biol. 2004; 6:499–506. doi: 10.1038/ncb1137CrossrefMedlineGoogle Scholar
  • 110. Rakesh K, Yoo B, Kim IM, Salazar N, Kim KS, Rockman HA. beta-Arrestin-biased agonism of the angiotensin receptor induced by mechanical stress.Sci Signal. 2010; 3:ra46. doi: 10.1126/scisignal.2000769CrossrefMedlineGoogle Scholar
  • 111. Tang W, Strachan RT, Lefkowitz RJ, Rockman HA. Allosteric modulation of β-arrestin-biased angiotensin II type 1 receptor signaling by membrane stretch.J Biol Chem. 2014; 289:28271–28283. doi: 10.1074/jbc.M114.585067CrossrefMedlineGoogle Scholar
  • 112. Yasuda N, Miura S, Akazawa Het al. Conformational switch of angiotensin II type 1 receptor underlying mechanical stress-induced activation.EMBO Rep. 2008; 9:179–186. doi: 10.1038/sj.embor.7401157CrossrefMedlineGoogle Scholar
  • 113. van Kesteren CA, van Heugten HA, Lamers JM, Saxena PR, Schalekamp MA, Danser AH. Angiotensin II-mediated growth and antigrowth effects in cultured neonatal rat cardiac myocytes and fibroblasts.J Mol Cell Cardiol. 1997; 29:2147–2157. doi: 10.1006/jmcc.1997.0448CrossrefMedlineGoogle Scholar
  • 114. Tadevosyan A, Xiao J, Surinkaew S, Naud P, Merlen C, Harada M, Qi X, Chatenet D, Fournier A, Allen BG, Nattel S. Intracellular angiotensin-II interacts with nuclear angiotensin receptors in cardiac fibroblasts and regulates RNA synthesis, cell proliferation, and collagen secretion.J Am Heart Assoc. 2017; 6:e004965.LinkGoogle Scholar
  • 115. Mehta PK, Griendling KK. Angiotensin II cell signaling: physiological and pathological effects in the cardiovascular system.Am J Physiol Cell Physiol. 2007; 292:C82–C97. doi: 10.1152/ajpcell.00287.2006CrossrefMedlineGoogle Scholar
  • 116. Paradis P, Dali-Youcef N, Paradis FW, Thibault G, Nemer M. Overexpression of angiotensin II type I receptor in cardiomyocytes induces cardiac hypertrophy and remodeling.Proc Natl Acad Sci USA. 2000; 97:931–936.CrossrefMedlineGoogle Scholar
  • 117. Adams JW, Sakata Y, Davis MG, Sah VP, Wang Y, Liggett SB, Chien KR, Brown JH, Dorn GW. Enhanced Galphaq signaling: a common pathway mediates cardiac hypertrophy and apoptotic heart failure.Proc Natl Acad Sci USA. 1998; 95:10140–10145.CrossrefMedlineGoogle Scholar
  • 118. Rockman HA, Wachhorst SP, Mao L, Ross J. ANG II receptor blockade prevents ventricular hypertrophy and ANF gene expression with pressure overload in mice.Am J Physiol. 1994; 266:H2468–H2475. doi: 10.1152/ajpheart.1994.266.6.H2468MedlineGoogle Scholar
  • 119. Ainscough JF, Drinkhill MJ, Sedo A, Turner NA, Brooke DA, Balmforth AJ, Ball SG. Angiotensin II type-1 receptor activation in the adult heart causes blood pressure-independent hypertrophy and cardiac dysfunction.Cardiovasc Res. 2009; 81:592–600. doi: 10.1093/cvr/cvn230CrossrefMedlineGoogle Scholar
  • 120. Trask AJ, Ferrario CM. Angiotensin-(1-7): pharmacology and new perspectives in cardiovascular treatments.Cardiovasc Drug Rev. 2007; 25:162–174. doi: 10.1111/j.1527-3466.2007.00012.xCrossrefMedlineGoogle Scholar
  • 121. Teixeira LB, Parreiras-E-Silva LT, Bruder-Nascimento T, Duarte DA, Simões SC, Costa RM, Rodríguez DY, Ferreira PAB, Silva CAA, Abrao EP, Oliveira EB, Bouvier M, Tostes RC, Costa-Neto CM. Ang-(1-7) is an endogenous β-arrestin-biased agonist of the AT1 receptor with protective action in cardiac hypertrophy.Sci Rep. 2017; 7:11903. doi: 10.1038/s41598-017-12074-3CrossrefMedlineGoogle Scholar
  • 122. Cigola E, Kajstura J, Li B, Meggs LG, Anversa P. Angiotensin II activates programmed myocyte cell death in vitro.Exp Cell Res. 1997; 231:363–371. doi: 10.1006/excr.1997.3477CrossrefMedlineGoogle Scholar
  • 123. Nishida M, Tanabe S, Maruyama Y, Mangmool S, Urayama K, Nagamatsu Y, Takagahara S, Turner JH, Kozasa T, Kobayashi H, Sato Y, Kawanishi T, Inoue R, Nagao T, Kurose H. G alpha 12/13-and reactive oxygen species-dependent activation of c-Jun NH2-terminal kinase and p38 mitogen-activated protein kinase by angiotensin receptor stimulation in rat neonatal cardiomyocytes.J Biol Chem. 2005; 280:18434–18441. doi: 10.1074/jbc.M409710200CrossrefMedlineGoogle Scholar
  • 124. Huang CY, Kuo WW, Yeh YL, Ho TJ, Lin JY, Lin DY, Chu CH, Tsai FJ, Tsai CH, Huang CY. ANG II promotes IGF-IIR expression and cardiomyocyte apoptosis by inhibiting HSF1 via JNK activation and SIRT1 degradation.Cell Death Differ. 2014; 21:1262–1274. doi: 10.1038/cdd.2014.46CrossrefMedlineGoogle Scholar
  • 125. Berk BC, Fujiwara K, Lehoux S. ECM remodeling in hypertensive heart disease.J Clin Invest. 2007; 117:568–575. doi: 10.1172/JCI31044CrossrefMedlineGoogle Scholar
  • 126. Wang W, Huang XR, Canlas E, Oka K, Truong LD, Deng C, Bhowmick NA, Ju W, Bottinger EP, Lan HY. Essential role of Smad3 in angiotensin II-induced vascular fibrosis.Circ Res. 2006; 98:1032–1039. doi: 10.1161/01.RES.0000218782.52610.dcLinkGoogle Scholar
  • 127. Leask A. Potential therapeutic targets for cardiac fibrosis: TGFbeta, angiotensin, endothelin, CCN2, and PDGF, partners in fibroblast activation.Circ Res. 2010; 106:1675–1680. doi: 10.1161/CIRCRESAHA.110.217737LinkGoogle Scholar
  • 128. Zhai P, Yamamoto M, Galeotti Jet al. Cardiac-specific overexpression of AT1 receptor mutant lacking G alpha q/G alpha i coupling causes hypertrophy and bradycardia in transgenic mice.J Clin Invest. 2005; 115:3045–3056. doi: 10.1172/JCI25330CrossrefMedlineGoogle Scholar
  • 129. Gu YJ, Sun WY, Zhang S, Wu JJ, Wei W. The emerging roles of β-arrestins in fibrotic diseases.Acta Pharmacol Sin. 2015; 36:1277–1287. doi: 10.1038/aps.2015.74CrossrefMedlineGoogle Scholar
  • 130. Funabiki K, Onishi K, Dohi K, Koji T, Imanaka-Yoshida K, Ito M, Wada H, Isaka N, Nobori T, Nakano T. Combined angiotensin receptor blocker and ACE inhibitor on myocardial fibrosis and left ventricular stiffness in dogs with heart failure.Am J Physiol Heart Circ Physiol. 2004; 287:H2487–H2492. doi: 10.1152/ajpheart.00462.2004CrossrefMedlineGoogle Scholar
  • 131. Schena M, Mulatero P, Schiavone D, Mengozzi G, Tesio L, Chiandussi L, Veglio F. Vasoactive hormones induce nitric oxide synthase mRNA expression and nitric oxide production in human endothelial cells and monocytes.Am J Hypertens. 1999; 12:388–397.CrossrefMedlineGoogle Scholar
  • 132. Rajagopalan S, Kurz S, Münzel T, Tarpey M, Freeman BA, Griendling KK, Harrison DG. Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone.J Clin Invest. 1996; 97:1916–1923. doi: 10.1172/JCI118623CrossrefMedlineGoogle Scholar
  • 133. Scalia R, Gong Y, Berzins B, Freund B, Feather D, Landesberg G, Mishra G. A novel role for calpain in the endothelial dysfunction induced by activation of angiotensin II type 1 receptor signaling.Circ Res. 2011; 108:1102–1111. doi: 10.1161/CIRCRESAHA.110.229393LinkGoogle Scholar
  • 134. Kishi T, Hirooka Y, Konno S, Sunagawa K. Angiotensin II receptor blockers improve endothelial dysfunction associated with sympathetic hyperactivity in metabolic syndrome.J Hypertens. 2012; 30:1646–1655. doi: 10.1097/HJH.0b013e328355860eCrossrefMedlineGoogle Scholar
  • 135. Hartman CD, Isenberg BC, Chua SG, Wong JY. Vascular smooth muscle cell durotaxis depends on extracellular matrix composition.Proc Natl Acad Sci USA. 2016; 113:11190–11195. doi: 10.1073/pnas.1611324113CrossrefMedlineGoogle Scholar
  • 136. Griendling KK, Minieri CA, Ollerenshaw JD, Alexander RW. Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells.Circ Res. 1994; 74:1141–1148.LinkGoogle Scholar
  • 137. Xi XP, Graf K, Goetze S, Fleck E, Hsueh WA, Law RE. Central role of the MAPK pathway in ang II-mediated DNA synthesis and migration in rat vascular smooth muscle cells.Arterioscler Thromb Vasc Biol. 1999; 19:73–82.LinkGoogle Scholar
  • 138. Yamakawa T, Tanaka S, Numaguchi K, Yamakawa Y, Motley ED, Ichihara S, Inagami T. Involvement of Rho-kinase in angiotensin II-induced hypertrophy of rat vascular smooth muscle cells.Hypertension. 2000; 35:313–318.LinkGoogle Scholar
  • 139. Montezano AC, Callera GE, Yogi A, He Y, Tostes RC, He G, Schiffrin EL, Touyz RM. Aldosterone and angiotensin II synergistically stimulate migration in vascular smooth muscle cells through c-Src-regulated redox-sensitive RhoA pathways.Arterioscler Thromb Vasc Biol. 2008; 28:1511–1518. doi: 10.1161/ATVBAHA.108.168021LinkGoogle Scholar
  • 140. Suzuki J, Iwai M, Nakagami H, Wu L, Chen R, Sugaya T, Hamada M, Hiwada K, Horiuchi M. Role of angiotensin II-regulated apoptosis through distinct AT1 and AT2 receptors in neointimal formation.Circulation. 2002; 106:847–853.LinkGoogle Scholar
  • 141. Touyz RM, He G, Wu XH, Park JB, Mabrouk ME, Schiffrin EL. Src is an important mediator of extracellular signal-regulated kinase ½-dependent growth signaling by angiotensin II in smooth muscle cells from resistance arteries of hypertensive patients.Hypertension. 2001; 38:56–64.LinkGoogle Scholar
  • 142. Kagiyama S, Eguchi S, Frank GD, Inagami T, Zhang YC, Phillips MI. Angiotensin II-induced cardiac hypertrophy and hypertension are attenuated by epidermal growth factor receptor antisense.Circulation. 2002; 106:909–912.LinkGoogle Scholar
  • 143. Forrester SJ, Kawai T, O’Brien S, Thomas W, Harris RC, Eguchi S. Epidermal growth factor receptor transactivation: mechanisms, pathophysiology, and potential therapies in the cardiovascular system.Annu Rev Pharmacol Toxicol. 2016; 56:627–653. doi: 10.1146/annurev-pharmtox-070115-095427CrossrefMedlineGoogle Scholar
  • 144. Nguyen Dinh Cat A, Montezano AC, Burger D, Touyz RM. Angiotensin II, NADPH oxidase, and redox signaling in the vasculature.Antioxid Redox Signal. 2013; 19:1110–1120. doi: 10.1089/ars.2012.4641CrossrefMedlineGoogle Scholar
  • 145. Olson ER, Shamhart PE, Naugle JE, Meszaros JG. Angiotensin II-induced extracellular signal-regulated kinase ½ activation is mediated by protein kinase Cdelta and intracellular calcium in adult rat cardiac fibroblasts.Hypertension. 2008; 51:704–711. doi: 10.1161/HYPERTENSIONAHA.107.098459LinkGoogle Scholar
  • 146. Lymperopoulos A, Rengo G, Koch WJ. Adrenergic nervous system in heart failure: pathophysiology and therapy.Circ Res. 2013; 113:739–753. doi: 10.1161/CIRCRESAHA.113.300308LinkGoogle Scholar
  • 147. O’Connell TD, Jensen BC, Baker AJ, Simpson PC. Cardiac alpha1-adrenergic receptors: novel aspects of expression, signaling mechanisms, physiologic function, and clinical importance.Pharmacol Rev. 2014; 66:308–333. doi: 10.1124/pr.112.007203CrossrefMedlineGoogle Scholar
  • 148. Jensen BC, Swigart PM, Laden ME, DeMarco T, Hoopes C, Simpson PC. The alpha-1D Is the predominant alpha-1-adrenergic receptor subtype in human epicardial coronary arteries.J Am Coll Cardiol. 2009; 54:1137–1145. doi: 10.1016/j.jacc.2009.05.056CrossrefMedlineGoogle Scholar
  • 149. O’Connell TD, Ishizaka S, Nakamura A, Swigart PM, Rodrigo MC, Simpson GL, Cotecchia S, Rokosh DG, Grossman W, Foster E, Simpson PC. The alpha(1A/C)-and alpha(1B)-adrenergic receptors are required for physiological cardiac hypertrophy in the double-knockout mouse.J Clin Invest. 2003; 111:1783–1791. doi: 10.1172/JCI16100CrossrefMedlineGoogle Scholar
  • 150. McCloskey DT, Turnbull L, Swigart P, O’Connell TD, Simpson PC, Baker AJ. Abnormal myocardial contraction in alpha(1A)-and alpha(1B)-adrenoceptor double-knockout mice.J Mol Cell Cardiol. 2003; 35:1207–1216.CrossrefMedlineGoogle Scholar
  • 151. O’Connell TD, Swigart PM, Rodrigo MC, Ishizaka S, Joho S, Turnbull L, Tecott LH, Baker AJ, Foster E, Grossman W, Simpson PC. Alpha1-adrenergic receptors prevent a maladaptive cardiac response to pressure overload.J Clin Invest. 2006; 116:1005–1015. doi: 10.1172/JCI22811CrossrefMedlineGoogle Scholar
  • 152. Antihypertensive and Lipid-Lowering Treatment to Prevent Heart Attack Trial Collaborative Research Group. Diuretic versus alpha-blocker as first-step antihypertensive therapy: final results from the Antihypertensive and Lipid-Lowering Treatment to Prevent Heart Attack Trial (ALLHAT).Hypertension. 2003; 42:239–246. doi: 10.1161/01.HYP.0000086521.95630.5ALinkGoogle Scholar
  • 153. Cohn JN. The Vasodilator-Heart Failure Trials (V-HeFT). Mechanistic data from the VA cooperative studies. Introduction.Circulation. 1993; 87:VI1–VI4.MedlineGoogle Scholar
  • 154. Stewart AF, Rokosh DG, Bailey BA, Karns LR, Chang KC, Long CS, Kariya K, Simpson PC. Cloning of the rat alpha 1C-adrenergic receptor from cardiac myocytes. alpha 1C, alpha 1B, and alpha 1D mRNAs are present in cardiac myocytes but not in cardiac fibroblasts.Circ Res. 1994; 75:796–802.LinkGoogle Scholar
  • 155. Jensen BC, Swigart PM, Montgomery MD, Simpson PC. Functional alpha-1B adrenergic receptors on human epicardial coronary artery endothelial cells.Naunyn Schmiedebergs Arch Pharmacol. 2010; 382:475–482. doi: 10.1007/s00210-010-0558-xCrossrefMedlineGoogle Scholar
  • 156. Hein L, Altman JD, Kobilka BK. Two functionally distinct alpha2-adrenergic receptors regulate sympathetic neurotransmission.Nature. 1999; 402:181–184. doi: 10.1038/46040CrossrefMedlineGoogle Scholar
  • 157. Brede M, Wiesmann F, Jahns R, Hadamek K, Arnolt C, Neubauer S, Lohse MJ, Hein L. Feedback inhibition of catecholamine release by two different alpha2-adrenoceptor subtypes prevents progression of heart failure.Circulation. 2002; 106:2491–2496.LinkGoogle Scholar
  • 158. Brede M, Nagy G, Philipp M, Sorensen JB, Lohse MJ, Hein L. Differential control of adrenal and sympathetic catecholamine release by alpha 2-adrenoceptor subtypes.Mol Endocrinol. 2003; 17:1640–1646. doi: 10.1210/me.2003-0035CrossrefMedlineGoogle Scholar
  • 159. Link RE, Desai K, Hein L, Stevens ME, Chruscinski A, Bernstein D, Barsh GS, Kobilka BK. Cardiovascular regulation in mice lacking alpha2-adrenergic receptor subtypes b and c.Science. 1996; 273:803–805.CrossrefMedlineGoogle Scholar
  • 160. Cottingham C, Chen Y, Jiao K, Wang Q. The antidepressant desipramine is an arrestin-biased ligand at the α(2A)-adrenergic receptor driving receptor down-regulation in vitro and in vivo.J Biol Chem. 2011; 286:36063–36075. doi: 10.1074/jbc.M111.261578CrossrefMedlineGoogle Scholar
  • 161. Gurwitz D, Haring R, Heldman E, Fraser CM, Manor D, Fisher A. Discrete activation of transduction pathways associated with acetylcholine m1 receptor by several muscarinic ligands.Eur J Pharmacol. 1994; 267:21–31.CrossrefMedlineGoogle Scholar
  • 162. Pronin AN, Wang Q, Slepak VZ. Teaching an old drug new tricks: agonism, antagonism, and biased signaling of pilocarpine through M3 muscarinic acetylcholine receptor.Mol Pharmacol. 2017; 92:601–612. doi: 10.1124/mol.117.109678CrossrefMedlineGoogle Scholar
  • 163. Brodde OE, Michel MC. Adrenergic and muscarinic receptors in the human heart.Pharmacol Rev. 1999; 51:651–690.MedlineGoogle Scholar
  • 164. Giessler C, Dhein S, Pönicke K, Brodde OE. Muscarinic receptors in the failing human heart.Eur J Pharmacol. 1999; 375:197–202.CrossrefMedlineGoogle Scholar
  • 165. Harvey RD, Belevych AE. Muscarinic regulation of cardiac ion channels.Br J Pharmacol. 2003; 139:1074–1084. doi: 10.1038/sj.bjp.0705338CrossrefMedlineGoogle Scholar
  • 166. DiFrancesco D. The role of the funny current in pacemaker activity.Circ Res. 2010; 106:434–446. doi: 10.1161/CIRCRESAHA.109.208041LinkGoogle Scholar
  • 167. DiFrancesco D, Tortora P. Direct activation of cardiac pacemaker channels by intracellular cyclic AMP.Nature. 1991; 351:145–147. doi: 10.1038/351145a0CrossrefMedlineGoogle Scholar
  • 168. Giles W, Noble SJ. Changes in membrane currents in bullfrog atrium produced by acetylcholine.J Physiol. 1976; 261:103–123.CrossrefMedlineGoogle Scholar
  • 169. Le Guludec D, Cohen-Solal A, Delforge J, Delahaye N, Syrota A, Merlet P. Increased myocardial muscarinic receptor density in idiopathic dilated cardiomyopathy: an in vivo PET study.Circulation. 1997; 96:3416–3422.LinkGoogle Scholar
  • 170. Pönicke K, Heinroth-Hoffmann I, Brodde OE. Demonstration of functional M3-muscarinic receptors in ventricular cardiomyocytes of adult rats.Br J Pharmacol. 2003; 138:156–160. doi: 10.1038/sj.bjp.0704997CrossrefMedlineGoogle Scholar
  • 171. Lee KY, Choi HC. Acetylcholine-induced AMP-activated protein kinase activation attenuates vasoconstriction through an LKB1-dependent mechanism in rat aorta.Vascul Pharmacol. 2013; 59:96–102. doi: 10.1016/j.vph.2013.07.007CrossrefMedlineGoogle Scholar
  • 172. Zhao M, Jia HH, Liu LZ, Bi XY, Xu M, Yu XJ, He X, Zang WJ. Acetylcholine attenuated TNF-α-induced intracellular Ca2+ overload by inhibiting the formation of the NCX1-TRPC3-IP3R1 complex in human umbilical vein endothelial cells.J Mol Cell Cardiol. 2017; 107:1–12. doi: 10.1016/j.yjmcc.2017.04.001CrossrefMedlineGoogle Scholar
  • 173. Zuccolo E, Lim D, Kheder DA, Perna A, Catarsi P, Botta L, Rosti V, Riboni L, Sancini G, Tanzi F, D’Angelo E, Guerra G, Moccia F. Acetylcholine induces intracellular Ca2+ oscillations and nitric oxide release in mouse brain endothelial cells.Cell Calcium. 2017; 66:33–47. doi: 10.1016/j.ceca.2017.06.003CrossrefMedlineGoogle Scholar
  • 174. Haag S, Matthiesen S, Juergens UR, Racké K. Muscarinic receptors mediate stimulation of collagen synthesis in human lung fibroblasts.Eur Respir J. 2008; 32:555–562. doi: 10.1183/09031936.00129307CrossrefMedlineGoogle Scholar
  • 175. Balligand JL, Kelly RA, Marsden PA, Smith TW, Michel T. Control of cardiac muscle cell function by an endogenous nitric oxide signaling system.Proc Natl Acad Sci USA. 1993; 90:347–351.CrossrefMedlineGoogle Scholar
  • 176. Maguire JJ, Kuc RE, Pell VR, Green A, Brown M, Kumar S, Wehrman T, Quinn E, Davenport AP. Comparison of human ETA and ETB receptor signalling via G-protein and β-arrestin pathways.Life Sci. 2012; 91:544–549. doi: 10.1016/j.lfs.2012.03.021CrossrefMedlineGoogle Scholar
  • 177. Green DS, Rupasinghe C, Warburton R, Wilson JL, Sallum CO, Taylor L, Yatawara A, Mierke D, Polgar P, Hill N. A cell permeable peptide targeting the intracellular loop 2 of endothelin B receptor reduces pulmonary hypertension in a hypoxic rat model.pLoS One. 2013; 8:e81309. doi: 10.1371/journal.pone.0081309CrossrefMedlineGoogle Scholar
  • 178. Harada M, Itoh H, Nakagawa O, Ogawa Y, Miyamoto Y, Kuwahara K, Ogawa E, Igaki T, Yamashita J, Masuda I, Yoshimasa T, Tanaka I, Saito Y, Nakao K. Significance of ventricular myocytes and nonmyocytes interaction during cardiocyte hypertrophy: evidence for endothelin-1 as a paracrine hypertrophic factor from cardiac nonmyocytes.Circulation. 1997; 96:3737–3744.LinkGoogle Scholar
  • 179. Dhaun N, Goddard J, Kohan DE, Pollock DM, Schiffrin EL, Webb DJ. Role of endothelin-1 in clinical hypertension: 20 years on.Hypertension. 2008; 52:452–459. doi: 10.1161/HYPERTENSIONAHA.108.117366LinkGoogle Scholar
  • 180. Rich S, McLaughlin VV. Endothelin receptor blockers in cardiovascular disease.Circulation. 2003; 108:2184–2190. doi: 10.1161/01.CIR.0000094397.19932.78LinkGoogle Scholar
  • 181. Schneider MP, Boesen EI, Pollock DM. Contrasting actions of endothelin ET(A) and ET(B) receptors in cardiovascular disease.Annu Rev Pharmacol Toxicol. 2007; 47:731–759. doi: 10.1146/annurev.pharmtox.47.120505.105134CrossrefMedlineGoogle Scholar
  • 182. Brunner F, Brás-Silva C, Cerdeira AS, Leite-Moreira AF. Cardiovascular endothelins: essential regulators of cardiovascular homeostasis.Pharmacol Ther. 2006; 111:508–531. doi: 10.1016/j.pharmthera.2005.11.001CrossrefMedlineGoogle Scholar
  • 183. Kuruvilla L, Nair RR, Umashankar PR, Lal AV, Kartha CC. Endocardial endothelial cells stimulate proliferation and collagen synthesis of cardiac fibroblasts.Cell Biochem Biophys. 2007; 47:65–72.CrossrefMedlineGoogle Scholar
  • 184. Guarda E, Katwa LC, Myers PR, Tyagi SC, Weber KT. Effects of endothelins on collagen turnover in cardiac fibroblasts.Cardiovasc Res. 1993; 27:2130–2134.CrossrefMedlineGoogle Scholar
  • 185. Chuo CH, Devine SM, Scammells PJ, Krum H, Christopoulos A, May LT, White PJ, Wang BH. VCP746, a novel A1 adenosine receptor biased agonist, reduces hypertrophy in a rat neonatal cardiac myocyte model.Clin Exp Pharmacol Physiol. 2016; 43:976–982. doi: 10.1111/1440-1681.12616CrossrefMedlineGoogle Scholar
  • 186. Baltos JA, Vecchio EA, Harris MA, Qin CX, Ritchie RH, Christopoulos A, White PJ, May LT. Capadenoson, a clinically trialed partial adenosine A1 receptor agonist, can stimulate adenosine A2B receptor biased agonism.Biochem Pharmacol. 2017; 135:79–89. doi: 10.1016/j.bcp.2017.03.014CrossrefMedlineGoogle Scholar
  • 187. Welihinda AA, Kaur M, Greene K, Zhai Y, Amento EP. The adenosine metabolite inosine is a functional agonist of the adenosine A2A receptor with a unique signaling bias.Cell Signal. 2016; 28:552–560. doi: 10.1016/j.cellsig.2016.02.010CrossrefMedlineGoogle Scholar
  • 188. Mustafa SJ, Morrison RR, Teng B, Pelleg A. Adenosine receptors and the heart: role in regulation of coronary blood flow and cardiac electrophysiology.Handb Exp Pharmacol. 2009; 193:161–188.CrossrefGoogle Scholar
  • 189. Donato M, Gelpi RJ. Adenosine and cardioprotection during reperfusion–an overview.Mol Cell Biochem. 2003; 251:153–159.CrossrefMedlineGoogle Scholar
  • 190. Peart J, Headrick JP. Intrinsic A(1) adenosine receptor activation during ischemia or reperfusion improves recovery in mouse hearts.Am J Physiol Heart Circ Physiol. 2000; 279:H2166–H2175. doi: 10.1152/ajpheart.2000.279.5.H2166CrossrefMedlineGoogle Scholar
  • 191. Headrick JP, Hack B, Ashton KJ. Acute adenosinergic cardioprotection in ischemic-reperfused hearts.Am J Physiol Heart Circ Physiol. 2003; 285:H1797–H1818. doi: 10.1152/ajpheart.00407.2003CrossrefMedlineGoogle Scholar
  • 192. Puhl SL, Kazakov A, Müller A, Fries P, Wagner DR, Böhm M, Maack C, Devaux Y. Adenosine A1 receptor activation attenuates cardiac hypertrophy and fibrosis in response to α1 –-adrenoceptor stimulation in vivo.Br J Pharmacol. 2016; 173:88–102. doi: 10.1111/bph.13339CrossrefMedlineGoogle Scholar
  • 193. Grandoch M, Hoffmann J, Röck K, Wenzel F, Oberhuber A, Schelzig H, Fischer JW. Novel effects of adenosine receptors on pericellular hyaluronan matrix: implications for human smooth muscle cell phenotype and interactions with monocytes during atherosclerosis.Basic Res Cardiol. 2013; 108:340. doi: 10.1007/s00395-013-0340-6CrossrefMedlineGoogle Scholar
  • 194. Ray CJ, Marshall JM. The cellular mechanisms by which adenosine evokes release of nitric oxide from rat aortic endothelium.J Physiol. 2006; 570:85–96. doi: 10.1113/jphysiol.2005.099390CrossrefMedlineGoogle Scholar
  • 195. Tabrizchi R, Bedi S. Pharmacology of adenosine receptors in the vasculature.Pharmacol Ther. 2001; 91:133–147.CrossrefMedlineGoogle Scholar
  • 196. Ho MF, Low LM, Rose’Meyer RB. Pharmacology of the adenosine A3 receptor in the vasculature and essential hypertension.pLoS One. 2016; 11:e0150021. doi: 10.1371/journal.pone.0150021CrossrefMedlineGoogle Scholar
  • 197. Epperson SA, Brunton LL, Ramirez-Sanchez I, Villarreal F. Adenosine receptors and second messenger signaling pathways in rat cardiac fibroblasts.Am J Physiol Cell Physiol. 2009; 296:C1171–C1177. doi: 10.1152/ajpcell.00290.2008CrossrefMedlineGoogle Scholar
  • 198. Wacker D, Wang C, Katritch V, Han GW, Huang XP, Vardy E, McCorvy JD, Jiang Y, Chu M, Siu FY, Liu W, Xu HE, Cherezov V, Roth BL, Stevens RC. Structural features for functional selectivity at serotonin receptors.Science. 2013; 340:615–619. doi: 10.1126/science.1232808CrossrefMedlineGoogle Scholar
  • 199. Kaumann AJ, Levy FO. 5-hydroxytryptamine receptors in the human cardiovascular system.Pharmacol Ther. 2006; 111:674–706.CrossrefMedlineGoogle Scholar
  • 200. Ito T, Ikeda U, Shimpo M, Yamamoto K, Shimada K. Serotonin increases interleukin-6 synthesis in human vascular smooth muscle cells.Circulation. 2000; 102:2522–2527.LinkGoogle Scholar
  • 201. Jaffré F, Callebert J, Sarre A, Etienne N, Nebigil CG, Launay JM, Maroteaux L, Monassier L. Involvement of the serotonin 5-HT2B receptor in cardiac hypertrophy linked to sympathetic stimulation: control of interleukin-6, interleukin-1beta, and tumor necrosis factor-alpha cytokine production by ventricular fibroblasts.Circulation. 2004; 110:969–974. doi: 10.1161/01.CIR.0000139856.20505.57LinkGoogle Scholar
  • 202. Monassier L, Laplante MA, Jaffré F, Bousquet P, Maroteaux L, de Champlain J. Serotonin 5-HT(2B) receptor blockade prevents reactive oxygen species-induced cardiac hypertrophy in mice.Hypertension. 2008; 52:301–307. doi: 10.1161/HYPERTENSIONAHA.107.105551LinkGoogle Scholar
  • 203. Nebigil CG, Choi DS, Dierich A, Hickel P, Le Meur M, Messaddeq N, Launay JM, Maroteaux L. Serotonin 2B receptor is required for heart development.Proc Natl Acad Sci USA. 2000; 97:9508–9513.CrossrefMedlineGoogle Scholar
  • 204. Nebigil CG, Jaffré F, Messaddeq N, Hickel P, Monassier L, Launay JM, Maroteaux L. Overexpression of the serotonin 5-HT2B receptor in heart leads to abnormal mitochondrial function and cardiac hypertrophy.Circulation. 2003; 107:3223–3229. doi: 10.1161/01.CIR.0000074224.57016.01LinkGoogle Scholar
  • 205. Alonso N, Monczor F, Echeverría E, Davio C, Shayo C, Fernández N. Signal transduction mechanism of biased ligands at histamine H2 receptors.Biochem J. 2014; 459:117–126. doi: 10.1042/BJ20131226CrossrefMedlineGoogle Scholar
  • 206. Nijmeijer S, Vischer HF, Rosethorne EM, Charlton SJ, Leurs R. Analysis of multiple histamine H4 receptor compound classes uncovers Gαi protein-and β-arrestin2-biased ligands.Mol Pharmacol. 2012; 82:1174–1182. doi: 10.1124/mol.112.080911CrossrefMedlineGoogle Scholar
  • 207. Felix SB, Baumann G, Helmus S, Sattelberger U. The role of histamine in cardiac anaphylaxis; characterization of histaminergic H1-and H2-receptor effects.Basic Res Cardiol. 1988; 83:531–539.CrossrefMedlineGoogle Scholar
  • 208. Imamura M, Seyedi N, Lander HM, Levi R. Functional identification of histamine H3-receptors in the human heart.Circ Res. 1995; 77:206–210.LinkGoogle Scholar
  • 209. Shogakiuchi Y, Kanaide H, Nakamura M. Cytosolic calcium transients differ between porcine coronary arterial and aortic smooth muscle cells in primary culture.Circ Res. 1991; 68:818–826.LinkGoogle Scholar
  • 210. Kishi F, Nakaya Y, Ito S. Histamine H2-receptor-mediated nitric oxide release from porcine endothelial cells.J Cardiovasc Pharmacol. 1998; 32:177–182.CrossrefMedlineGoogle Scholar
  • 211. Zeng Z, Shen L, Li X, Luo T, Wei X, Zhang J, Cao S, Huang X, Fukushima Y, Bin J, Kitakaze M, Xu D, Liao Y. Disruption of histamine H2 receptor slows heart failure progression through reducing myocardial apoptosis and fibrosis.Clin Sci (Lond). 2014; 127:435–448. doi: 10.1042/CS20130716CrossrefMedlineGoogle Scholar
  • 212. Zhang J, Cai WK, Zhang Z, Wang P, Lin XQ, Feng J, Fu SC, He GH. Cardioprotective effect of histamine H2 antagonists in congestive heart failure: a systematic review and meta-analysis.Medicine (Baltimore). 2018; 97:e0409. doi: 10.1097/MD.0000000000010409MedlineGoogle Scholar
  • 213. Yang P, Maguire JJ, Davenport AP. Apelin, Elabela/Toddler, and biased agonists as novel therapeutic agents in the cardiovascular system.Trends Pharmacol Sci. 2015; 36:560–567. doi: 10.1016/j.tips.2015.06.002CrossrefMedlineGoogle Scholar
  • 214. Szokodi I, Tavi P, Földes G, Voutilainen-Myllylä S, Ilves M, Tokola H, Pikkarainen S, Piuhola J, Rysä J, Tóth M, Ruskoaho H. Apelin, the novel endogenous ligand of the orphan receptor APJ, regulates cardiac contractility.Circ Res. 2002; 91:434–440.LinkGoogle Scholar
  • 215. Chun HJ, Ali ZA, Kojima Y, Kundu RK, Sheikh AY, Agrawal R, Zheng L, Leeper NJ, Pearl NE, Patterson AJ, Anderson JP, Tsao PS, Lenardo MJ, Ashley EA, Quertermous T. Apelin signaling antagonizes Ang II effects in mouse models of atherosclerosis.J Clin Invest. 2008; 118:3343–3354. doi: 10.1172/JCI34871MedlineGoogle Scholar
  • 216. Devarakonda T, Salloum FN. Heart disease and relaxin: new actions for an old hormone.Trends Endocrinol Metab. 2018; 29:338–348. doi: 10.1016/j.tem.2018.02.008CrossrefMedlineGoogle Scholar
  • 217. Samuel CS, Hewitson TD. Relaxin in cardiovascular and renal disease.Kidney Int. 2006; 69:1498–1502. doi: 10.1038/sj.ki.5000264CrossrefMedlineGoogle Scholar
  • 218. Samuel CS, Unemori EN, Mookerjee I, Bathgate RA, Layfield SL, Mak J, Tregear GW, Du XJ. Relaxin modulates cardiac fibroblast proliferation, differentiation, and collagen production and reverses cardiac fibrosis in vivo.Endocrinology. 2004; 145:4125–4133. doi: 10.1210/en.2004-0209CrossrefMedlineGoogle Scholar
  • 219. Wasilewski MA, Myers VD, Recchia FA, Feldman AM, Tilley DG. Arginine vasopressin receptor signaling and functional outcomes in heart failure.Cell Signal. 2016; 28:224–233. doi: 10.1016/j.cellsig.2015.07.021CrossrefMedlineGoogle Scholar
  • 220. Galvani S, Sanson M, Blaho VA, Swendeman SL, Obinata H, Conger H, Dahlbäck B, Kono M, Proia RL, Smith JD, Hla T. HDL-bound sphingosine 1-phosphate acts as a biased agonist for the endothelial cell receptor S1P1 to limit vascular inflammation.Sci Signal. 2015; 8:ra79. doi: 10.1126/scisignal.aaa2581CrossrefMedlineGoogle Scholar
  • 221. Sensken SC, Stäubert C, Keul P, Levkau B, Schöneberg T, Gräler MH. Selective activation of G alpha i mediated signalling of S1P3 by FTY720-phosphate.Cell Signal. 2008; 20:1125–1133. doi: 10.1016/j.cellsig.2008.01.019CrossrefMedlineGoogle Scholar
  • 222. Means CK, Brown JH. Sphingosine-1-phosphate receptor signalling in the heart.Cardiovasc Res. 2009; 82:193–200. doi: 10.1093/cvr/cvp086CrossrefMedlineGoogle Scholar
  • 223. Ohkura SI, Usui S, Takashima SI, Takuwa N, Yoshioka K, Okamoto Y, Inagaki Y, Sugimoto N, Kitano T, Takamura M, Wada T, Kaneko S, Takuwa Y. Augmented sphingosine 1 phosphate receptor-1 signaling in cardiac fibroblasts induces cardiac hypertrophy and fibrosis through angiotensin II and interleukin-6.pLoS One. 2017; 12:e0182329. doi: 10.1371/journal.pone.0182329CrossrefMedlineGoogle Scholar
  • 224. Keul P, van Borren MM, Ghanem Aet al. Sphingosine-1-phosphate receptor 1 regulates cardiac function by modulating Ca2+ sensitivity and Na+/H+ exchange and mediates protection by ischemic preconditioning.J Am Heart Assoc. 2016; 5:e003393.LinkGoogle Scholar
  • 225. Clay H, Wilsbacher LD, Wilson SJ, Duong DN, McDonald M, Lam I, Park KE, Chun J, Coughlin SR. Sphingosine 1-phosphate receptor-1 in cardiomyocytes is required for normal cardiac development.Dev Biol. 2016; 418:157–165. doi: 10.1016/j.ydbio.2016.06.024CrossrefMedlineGoogle Scholar
  • 226. Matera C, Tata AM. Pharmacological approaches to targeting muscarinic acetylcholine receptors.Recent Pat CNS Drug Discov. 2014; 9:85–100.CrossrefMedlineGoogle Scholar
  • 227. Caulfield MP. Muscarinic receptors–characterization, coupling and function.Pharmacol Ther. 1993; 58:319–379.CrossrefMedlineGoogle Scholar
  • 228. Logothetis DE, Kurachi Y, Galper J, Neer EJ, Clapham DE. The beta gamma subunits of GTP-binding proteins activate the muscarinic K+ channel in heart.Nature. 1987; 325:321–326. doi: 10.1038/325321a0CrossrefMedlineGoogle Scholar
  • 229. Brodde OE, Michel MC, Zerkowski HR. Signal transduction mechanisms controlling cardiac contractility and their alterations in chronic heart failure.Cardiovasc Res. 1995; 30:570–584.CrossrefMedlineGoogle Scholar
  • 230. Hussain RI, Qvigstad E, Birkeland JA, Eikemo H, Glende A, Sjaastad I, Skomedal T, Osnes JB, Levy FO, Krobert KA. Activation of muscarinic receptors elicits inotropic responses in ventricular muscle from rats with heart failure through myosin light chain phosphorylation.Br J Pharmacol. 2009; 156:575–586. doi: 10.1111/j.1750-3639.2009.00016.xCrossrefMedlineGoogle Scholar
  • 231. Gregory KJ, Hall NE, Tobin AB, Sexton PM, Christopoulos A. Identification of orthosteric and allosteric site mutations in M2 muscarinic acetylcholine receptors that contribute to ligand-selective signaling bias.J Biol Chem. 2010; 285:7459–7474. doi: 10.1074/jbc.M109.094011CrossrefMedlineGoogle Scholar
  • 232. Nelson CD, Perry SJ, Regier DS, Prescott SM, Topham MK, Lefkowitz RJ. Targeting of diacylglycerol degradation to M1 muscarinic receptors by beta-arrestins.Science. 2007; 315:663–666. doi: 10.1126/science.1134562CrossrefMedlineGoogle Scholar
  • 233. Margulies KB, Hildebrand FL, Lerman A, Perrella MA, Burnett JC. Increased endothelin in experimental heart failure.Circulation. 1990; 82:2226–2230.LinkGoogle Scholar
  • 234. Sakai S, Miyauchi T, Sakurai T, Kasuya Y, Ihara M, Yamaguchi I, Goto K, Sugishita Y. Endogenous endothelin-1 participates in the maintenance of cardiac function in rats with congestive heart failure. Marked increase in endothelin-1 production in the failing heart.Circulation. 1996; 93:1214–1222.LinkGoogle Scholar
  • 235. Schorlemmer A, Matter ML, Shohet RV. Cardioprotective signaling by endothelin.Trends Cardiovasc Med. 2008; 18:233–239. doi: 10.1016/j.tcm.2008.11.005CrossrefMedlineGoogle Scholar
  • 236. Sakai S, Miyauchi T, Kobayashi M, Yamaguchi I, Goto K, Sugishita Y. Inhibition of myocardial endothelin pathway improves long-term survival in heart failure.Nature. 1996; 384:353–355. doi: 10.1038/384353a0CrossrefMedlineGoogle Scholar
  • 237. McMurray JJ, Teerlink JR, Cotter Get al; VERITAS Investigators. Effects of tezosentan on symptoms and clinical outcomes in patients with acute heart failure: the VERITAS randomized controlled trials.JAMA. 2007; 298:2009–2019. doi: 10.1001/jama.298.17.2009CrossrefMedlineGoogle Scholar
  • 238. Anand I, McMurray J, Cohn JN, Konstam MA, Notter T, Quitzau K, Ruschitzka F, Lüscher TF; EARTH investigators. Long-term effects of darusentan on left-ventricular remodelling and clinical outcomes in the EndothelinA Receptor Antagonist Trial in Heart Failure (EARTH): randomised, double-blind, placebo-controlled trial.Lancet. 2004; 364:347–354. doi: 10.1016/S0140-6736(04)16723-8CrossrefMedlineGoogle Scholar
  • 239. Wei A, Gu Z, Li J, Liu X, Wu X, Han Y, Pu J. Clinical adverse effects of endothelin receptor antagonists: insights from the meta-analysis of 4894 patients from 24 randomized double-blind placebo-controlled clinical trials.J Am Heart Assoc. 2016; 5:e003896.LinkGoogle Scholar
  • 240. Shihoya W, Nishizawa T, Yamashita K, Inoue A, Hirata K, Kadji FMN, Okuta A, Tani K, Aoki J, Fujiyoshi Y, Doi T, Nureki O. X-ray structures of endothelin ETB receptor bound to clinical antagonist bosentan and its analog.Nat Struct Mol Biol. 2017; 24:758–764. doi: 10.1038/nsmb.3450CrossrefMedlineGoogle Scholar
  • 241. Olah ME, Ren H, Stiles GL. Adenosine receptors: protein and gene structure.Arch Int Pharmacodyn Ther. 1995; 329:135–150.MedlineGoogle Scholar
  • 242. Müller CE, Jacobson KA. Recent developments in adenosine receptor ligands and their potential as novel drugs.Biochim Biophys Acta. 2011; 1808:1290–1308. doi: 10.1016/j.bbamem.2010.12.017CrossrefMedlineGoogle Scholar
  • 243. Teerlink JR, Iragui VJ, Mohr JPet al. The safety of an adenosine A(1)-receptor antagonist, rolofylline, in patients with acute heart failure and renal impairment: findings from PROTECT.Drug Saf. 2012; 35:233–244. doi: 10.2165/11594680-000000000-00000CrossrefMedlineGoogle Scholar
  • 244. Gottlieb SS, Givertz MM, Metra M, Gergich K, Bird S, Jones-Burton C, Massie B, Cotter G, Ponikowski P, Weatherley B, O’Connor C, Dittrich H. The effects of adenosine A(1) receptor antagonism in patients with acute decompensated heart failure and worsening renal function: the REACH UP study.J Card Fail. 2010; 16:714–719. doi: 10.1016/j.cardfail.2010.04.006CrossrefMedlineGoogle Scholar
  • 245. Wilbur SL, Marchlinski FE. Adenosine as an antiarrhythmic agent.Am J Cardiol. 1997; 79:30–37.CrossrefMedlineGoogle Scholar
  • 246. Voors AA, Düngen HD, Senni M, Nodari S, Agostoni P, Ponikowski P, Bax JJ, Butler J, Kim RJ, Dorhout B, Dinh W, Gheorghiade M. Safety and tolerability of neladenoson bialanate, a novel oral partial adenosine A1 receptor agonist, in patients with chronic heart failure.J Clin Pharmacol. 2017; 57:440–451. doi: 10.1002/jcph.828CrossrefMedlineGoogle Scholar
  • 247. Greene SJ, Sabbah HN, Butler J, Voors AA, Albrecht-Küpper BE, Düngen HD, Dinh W, Gheorghiade M. Partial adenosine A1 receptor agonism: a potential new therapeutic strategy for heart failure.Heart Fail Rev. 2016; 21:95–102. doi: 10.1007/s10741-015-9522-7CrossrefMedlineGoogle Scholar
  • 248. Hakak Y, Shrestha D, Goegel MC, Behan DP, Chalmers DT. Global analysis of G-protein-coupled receptor signaling in human tissues.FEBS Lett. 2003; 550:11–17.CrossrefMedlineGoogle Scholar
  • 249. Shyu KG. Serotonin 5-HT2B receptor in cardiac fibroblast contributes to cardiac hypertrophy: a new therapeutic target for heart failure?Circ Res. 2009; 104:1–3. doi: 10.1161/CIRCRESAHA.108.191122LinkGoogle Scholar
  • 250. Kim J, Ogai A, Nakatani S, Hashimura K, Kanzaki H, Komamura K, Asakura M, Asanuma H, Kitamura S, Tomoike H, Kitakaze M. Impact of blockade of histamine H2 receptors on chronic heart failure revealed by retrospective and prospective randomized studies.J Am Coll Cardiol. 2006; 48:1378–1384. doi: 10.1016/j.jacc.2006.05.069CrossrefMedlineGoogle Scholar
  • 251. Ashley EA, Powers J, Chen M, Kundu R, Finsterbach T, Caffarelli A, Deng A, Eichhorn J, Mahajan R, Agrawal R, Greve J, Robbins R, Patterson AJ, Bernstein D, Quertermous T. The endogenous peptide apelin potently improves cardiac contractility and reduces cardiac loading in vivo.Cardiovasc Res. 2005; 65:73–82. doi: 10.1016/j.cardiores.2004.08.018CrossrefMedlineGoogle Scholar
  • 252. Japp AG, Cruden NL, Barnes G, van Gemeren N, Mathews J, Adamson J, Johnston NR, Denvir MA, Megson IL, Flapan AD, Newby DE. Acute cardiovascular effects of apelin in humans: potential role in patients with chronic heart failure.Circulation. 2010; 121:1818–1827. doi: 10.1161/CIRCULATIONAHA.109.911339LinkGoogle Scholar
  • 253. Morooka H, Iwanaga Y, Tamaki Y, Takase T, Akahoshi Y, Nakano Y, Fujiki H, Miyazaki S. Chronic administration of oral vasopressin type 2 receptor antagonist tolvaptan exerts both myocardial and renal protective effects in rats with hypertensive heart failure.Circ Heart Fail. 2012; 5:484–492. doi: 10.1161/CIRCHEARTFAILURE.111.965392LinkGoogle Scholar

eLetters(0)

eLetters should relate to an article recently published in the journal and are not a forum for providing unpublished data. Comments are reviewed for appropriate use of tone and language. Comments are not peer-reviewed. Acceptable comments are posted to the journal website only. Comments are not published in an issue and are not indexed in PubMed. Comments should be no longer than 500 words and will only be posted online. References are limited to 10. Authors of the article cited in the comment will be invited to reply, as appropriate.

Comments and feedback on AHA/ASA Scientific Statements and Guidelines should be directed to the AHA/ASA Manuscript Oversight Committee via its Correspondence page.