Downloads

Song, Y., WOO, A. Y.-H., Zhang, Y., & Xiao, R. Cardiac β-Adrenoceptor Signaling: The New Insight on An Old Target in the Therapy of Cardiovascular Disease. International Journal of Drug Discovery and Pharmacology. 2022, 1(1), 3. doi: https://doi.org/10.53941/ijddp.v1i1.177

Review

Cardiac β-Adrenoceptor Signaling: The New Insight on An Old Target in the Therapy of Cardiovascular Disease

Ying Song 1, Anthony Yiu-Ho Woo 2,*, Yan Zhang 3,4,*, and Ruiping Xiao 5,6,7,8

1 Center of Basic Medical Research, Institute of Medical Innovation and Research, Peking University Third Hospital, Beijing, 100191, China.

2 School of Life Sciences and Biopharmaceutics, Shenyang Pharmaceutical University, Shenyang, 110016, China.

3 Institute of Cardiovascular Sciences and Key Laboratory of Molecular Cardiovascular Sciences, School of Basic Medical Sciences, Ministry of Education, Peking University Health Science Center, Beijing, 100191, China.

4 Beijing Key Laboratory of Cardiovascular Receptors Research, Peking University, Beijing, 100191, China.

5 State Key Laboratory of Membrane Biology, Institute of Molecular Medicine, College of Future Technology, Peking University, Beijing, 100871, China.

6 Peking-Tsinghua Center for Life Sciences, Peking University, Beijing, China.

7 Beijing City Key Laboratory of Cardiometabolic Molecular Medicine, Peking University, Beijing, 100871, China.

8 PKU-Nanjing Joint Institute of Translational Medicine, Nanjing, 210000, China.

* Correspondence: yiuhowoo@syphu.edu.cn (Anthony Yiu-Ho Woo), Tel.: +86-24-23986375; zhangyan9876@pku.edu.cn (Yan Zhang), Tel.: +86-10-82805945.

 

 

Received: 19 October 2022

Accepted: 28 November 2022

Published: 21 December 2022

 

Abstract: A variety of G protein-coupled receptors (GPCRs) are involved in the regulation of cardiovascular function. The β-adrenoceptors (β-ARs), with three subtypes, are the dominant receptor species in the heart, in which the β1-AR and the β2-AR are considered functional. Stimulation of the β-ARs produces myocardial inotropy via activation of the Gs-cAMP-PKA signaling cascade. Prolonged stimulation of the β1-AR is cardiac harmful because the stimulated β1-AR couples only to Gs proteins and it mediates a cardiotoxic signal. On the other hand, the β2-AR couples dually to both Gs and Gi proteins and the β2-AR-Gi pathway is antiapoptotic. The activated Gi signal also counteracts the β-AR-Gs-promoted positive inotropic effect. Other key players in cardiac β-AR signaling include Ca2+/calmodulin-dependent protein kinases (CaMKs), GPCR kinases (GRKs), β-arrestins and phosphodiesterases. During heart failure, excessive sympathetic stimulation results in the activation of the cardiotoxic β1-AR-CaMKIIδ pathway and the upregulation of GRK2 and Gi in the heart. GRK2 promotes the desensitization of β-ARs and enhances a β2-AR-mediated Gi signaling. These signal transduction processes accompanying the downregulation of the β1-AR are involved in cardiac dysfunction, maladaptive cardiac remodeling, and the progression of chronic heart failure. β-Blockers are widely used in the treatment of cardiovascular disease. They have established their position as one of the “four pillars of heart failure” thirty years ago. In the present review, we provide an overview of the recent progress in the basic research of GPCRs focusing on cardiac β-AR signal transduction.

Keywords:

heart receptors, adrenergic, beta signal transduction heart failure

1.GPCRs and Cardiovascular Disease

1.1. Brief Introduction of GPCRs

G protein-coupled receptors (GPCRs), or seven-fold transmembrane receptors, are the largest class of membrane proteins in the human genome. GPCRs share a common structure consisting of peptides with an extracellular N-terminus, an intracellular C-terminus (CT), and seven hydrophobic transmembrane helices linked by three extracellular and three intracellular loops (ICLs). The extracellular part of the receptor is usually glycosylated, and the highly conserved cysteine residues in the extracellular loops can form disulfide bonds to stabilize the spatial structure of the receptor. The CT and the ICLs are binding sites for G proteins, G protein-coupled receptor kinases (GRKs), arrestins and other downstream signaling molecules. GRKs and arrestins are key molecules participating in the desensitization of GPCRs. In addition, adenylyl cyclase (AC), phospholipase C, phosphodiesterases (PDEs), cyclic adenosine monophosphate (cAMP), cyclic guanosine monophosphate (cGMP), inositol triphosphate, Ca2+, etc. are also involved in GPCR signal transduction. Classical theory suggests that upon binding to a ligand, a GPCR undergoes a conformational change that promotes interaction of its intracellular part with a G protein. The exchange of GDP for GTP in the G protein results in the dissociation of the heterotrimeric G protein into the Gα subunit and the Gβγ subunits. Both of which interact with various effectors to regulate second messenger production or ion channel opening. Over the past two decades, numerous studies have shown that receptor-bound arrestins trigger a second round of signaling that is quite different from that mediated by G proteins [1]. Some GPCR ligands are ‘balanced’ while some others are ‘biased’ in terms of their preferences on activating one type of G protein or arrestin signal over another [1]. Adding to this complexity is the capability of GPCRs to form heterodimers, and the new receptor species thus formed can have signaling properties different from those of their parent receptors. Thus, GPCR heterodimers constitute very interesting targets for future drugs [2,3].

The human GPCR superfamily have more than 800 members. Some GPCRs have sensory function mediating smell, taste, vision or pheromone signals. GPCRs without sensory function participate in growth, development, reproduction and other physiological processes. They are also involved in the development of diseases including cardiovascular diseases, diabetes, cancer, immune disorders, infectious diseases, neurological and mental diseases. GPCRs can be divided into five categories according to their structures, as shown in Table 1. Such taxonomy is known as the GRAFS nomenclature system [4].

Table 1. GPCR Classes.

The orphan receptors in Table 1 refer to GPCRs lacking endogenous ligands. Although molecular biology and bioinformatic techniques have enabled the identification of orphan receptors, it has been a great challenge to find their endogenous ligands. Over 300 GPCRs have been de-orphaned in the last few decades. These receptors are potential cardiovascular disease targets. For example, the Gi protein-coupled receptor APJ, once considered an orphan receptor, is found to be a receptor with its endogenous ligand apelin, which exerts peripheral and central cardiovascular effects by promoting vasodilation, angiogenesis and contractility of the heart [5]. For receptors that are not de-orphaned, endogenous ligands can be bypassed to develop alternative ligand drugs. Modulators of GPR119, GPR35, GPR55, MAS and GPR84 have already entered clinical trials [6].

GPCRs are among the most well-established drug targets. With the development of structural biology and other technologies, drugs targeting GPCRs are rapidly developed and utilized. New GPCR targets continue to emerge, bringing new opportunities for drug development.

1.2.GPCR and Cardiovascular Disease

Cardiovascular diseases are the leading cause of mortality and morbidity worldwide, accounting for a quarter of global deaths every year [7]. Myocardial infarction, cardiomyopathy, valvular heart disease, viral myocarditis, chronic hypertension, coronary artery disease, heart injuries, etc. can eventually lead to heart failure. Heart failure is a complex clinical syndrome characterized by a reduced pumping and/or refilling capacity of the heart. Physiologically, heart failure can be defined as insufficiency of the cardiac output to meet metabolic demands. Globally, about 26 million adults suffer from heart failure, and therefore heart failure is defined as a global pandemic. The five-year mortality rate of heart failure is as high as 42.3% [8]. Patients with heart failure usually require long and repeated hospitalization. A rising patient population in heart failure brings serious economic, social and personal burdens.

GPCRs are widely expressed in the cardiovascular system and the corresponding cell types, such as cardiomyocytes, fibroblasts, endothelial cells and vascular smooth muscle cells. Over 200 GPCRs participate in the regulation of cardiovascular function. GPCRs present in the cardiovascular system are involved in the sensing and responses to various stimuli (such as mechanical stress, hormones, cytokines and growth factors). They also participate in the development of cardiovascular diseases, comprising 24% of the drug targets in the cardiovascular system [9]. The most well-studied ones include adrenoceptors (ARs), angiotensin receptors (ATRs), muscarinic receptors, endothelin receptors, and adenosine receptors.

ATRs, especially the AT1R, play a key role in the pathophysiology of the cardiovascular system. AT1Rs are expressed in the heart as well as in blood vessels. Angiotensin II acts on vascular AT1Rs to produce vasoconstriction causing blood pressure to rise. Overexpression of the AT1R promotes myocardial fibrosis and hypertrophy while its knockdown enhances cardiac function after myocardial infarction, suggesting that AT1R mediates a harmful effect on the heart. Myocardial AT1R is upregulated during ischemia and is associated with adverse cardiac remodeling in heart failure [10]. Therefore, AT1R antagonists (or angiotensin II receptor blockers, ARBs) are widely used in the treatment of cardiovascular diseases such as hypertension, coronary artery disease and heart failure. The AT1R couples to Gαq proteins and it may also mediate signals via Gβγ subunits and β-arrestins [11]. Recent studies have suggested that activation of β-arrestin signaling in addition to the blockade of G protein signaling downstream of the AT1R may provide a beneficial cardiotonic effect in heart failure. This may lead to the development of a superior class of ARBs [12,13]. Thus, similar basic studies on GPCR signal transduction may open up new opportunities for the development of future cardiovascular drugs.

2.β-AR Signaling in the Cardiovascular System

2.1.β-ARs in the Cardiovascular System

Vital body functions such as heart beat and blood pressure maintenance are under the control of the autonomic nervous system composed of the sympathetic nervous system (SNS) and the parasympathetic nervous system. Target organs of the cardiovascular system such as the heart and peripheral arteries respond to sympathetic stimulation by means of ARs expressed on the plasma membranes of target cells. Norepinephrine secreted from sympathetic nerve terminals and circulatory catecholamines (mainly epinephrine with a small amount of norepinephrine) are the endogenous agonists of these receptors. In vascular smooth muscles, catecholamines stimulate α‑ARs to cause vasoconstriction and β2-ARs to cause vasodilation. In the heart, stimulation of β‑ARs increases heart rate and myocardial contractility. Adaptive activation of the SNS increases blood pressure and cardiac output to meet the metabolic requirements of the body.

β-ARs are the most abundant GPCR members expressed in mammalian hearts. All the three subtypes of the β-ARs, β1, β2 and β3, exist in the human heart. The β1-AR has the highest abundance followed by the β2-AR. Under normal conditions, the expression level of the β3-AR is much lower [14]. The expression patterns of β-ARs in specific cells is worth exploring and refining. The most abundant cell types in the mammalian heart are cardiomyocytes, cardiac fibroblasts, endothelial cells and vascular smooth muscle cells. Among them, cardiomyocytes account for about 30% in quantity and 70% in mass and volume, constituting the myocardium and providing contractile force for the heart [15]. Recent data show that β1-ARs are expressed in every cardiomyocyte in mice [16]. Cardiac fibroblasts, endothelial cells and vascular smooth muscle cells generally express β2-ARs. With the increase of age, the level of β1-ARs in human ventricular muscles decreases, especially in women [17]. But sex and age have no impact on β-AR-mediated inotropy in isolated human atrial muscles [18].

2.2.Cardiac β-AR Signaling

β-ARs are Gs protein-coupled receptors. Signal transduction begins with the binding of an adrenergic agonist to a β-AR causing it to couple to intracellular Gs proteins. The activated Gs protein enhances cAMP synthesis by AC, thereby activating cAMP-dependent protein kinase (PKA) which subsequently phosphorylates various PKA substrates to produce a cellular response. In cardiomyocytes, these PKA substrates include proteins responsible for intracellular Ca2+ regulation (e.g. phospholamban (PLB), L-type calcium channel (LTCC) and ryanodine receptor (RyR)), other cation channels involved in the generation of an action potential (e.g. the delayed rectifier potassium channel IKs) and regulatory proteins in the contractile machinery (e.g. cardiac troponin I and cardiac myosin-binding protein C).

Excitation-contraction coupling is the mechanism to relay electrical activity in a muscle cell to the process of cellular contraction. The transport of Ca2+ across different membranes play a central role in this process. LTCC on the plasma membrane (sarcolemma) is responsible for the voltage-dependent influx of extracellular Ca2+, which is spatially coupled with a cluster of RyRs on subsarcolemmal cisternae of sarcoplasmic reticulum (SR) responsible for the release of stored Ca2+ into the cytosol. An action potential in a cardiomyocyte triggers a small influx of Ca2+ (ICa,L) through LTCC. The increase in local Ca2+ levels induces a large efflux of stored Ca2+ via RyRs in a process known as calcium-induced calcium release [19], generating Ca2+ sparks. Spatial synchronization of these Ca2+ sparks results in a large surge of cytosolic Ca2+ concentration ([Ca2+]i). Binding of Ca2+ to cardiac troponin C activates myofilaments and initiates cell contraction. In the repolarization phase of the action potential, actions of the sarcolemmal sodium-calcium exchanger to transport intracellular Ca2+ out of the cell and that of the sarcoplasmic/endoplasmic reticulum calcium ATPase (SERCA) to refill the Ca2+ pool restore the resting [Ca2+]i level, allowing the cardiomyocyte to relax.

Acute β-adrenergic stimulation increases myocardial contractility through enhancement of Ca2+ transport, myofilament Ca2+ sensitivity and membrane repolarization. These effects are the result of increased PKA-dependent phosphorylation of proteins. Studies on isolated cardiomyocytes have shown that β-adrenergic stimulation produces a positive inotropic effect and increases IKs and both of these effects depends on PKA activation [20,21] (Figure 1).

Figure 1. Schematic representation of β‑AR signaling in cardiomyocytes. The β1-AR and the β2-AR mediate the biological effects of catecholamines (CA) on cardiomyocytes. These effects include one on contractility enhancement (i.e. a positive inotropic effect) and another one on cell survival and death regulation. Though both of these receptors by means of coupling to Gs mediate a positive inotropic effect, the promiscuously Gi-coupled β2-AR also mediates a negative effect on contractility and a pro-survival effect. The positive inotropic effect of acute β1 or β2 stimulation depends on the activation of PKA and the resultant phosphorylation of proteins (LTCC, PLB, RyR, cMBP-C, etc.) that promote the mobilization and efficiency of cellular Ca2+. The β2-AR-Gs-cAMP signal is compartmentalized near t-tubules and caveolae due to subcellular localization of the β2-ARs and PDE4 (which hydrolyzes cAMP) at these membrane compartments. On the other hand, the β1-AR-Gs-cAMP signal is a global one thanks to the even distribution of β1-ARs on the sarcolemma. ‘Chronic’ (long-term) stimulation of the β1-AR turns on a β1-AR-CaMKII signal, which mediates both a positive inotropic effect and a pro-apoptotic effect. GRKs and β-arrestins are important regulators of cardiac β-AR-signaling. They take part in homologous desensitization that terminates CA-triggered β-AR-Gs signaling. GRK2 is the most abundant GRK in the heart and has been proven to be a strong negative regulator of β-AR-Gs signaling. The role of β-arrestin is more diverse because β-arrestin also mediates signaling of its own and it has been implicated in the activation of the β1-AR-CaMKII signal as well as in heterologous desensitization of the β2-AR and cardioprotection (not shown). Heart diseases, particularly heart failure, is manifested by excessive CA stimulation and changes of the expression pattern of proteins in this signaling network, causing disturbances on the normal balance of cellular function and well-being. Thus, targeting on the cardiac β-AR signaling provides an important means to treat heart diseases. See the text for details. AC, adenylyl cyclase; Akt, protein kinase B; β1-AR, β1-adrenoceptor; β2-AR, β2-adrenoceptor; βγ, Gβγ subunits; CA, catecholamine; CaMKII, Ca2+/calmodulin-dependent protein kinase II; cAMP, cyclic adenosine monophosphate; cMBP-C, cardiac myosin-binding protein C; EPAC, exchange protein activated by cAMP; GRK2, G protein-coupled receptor kinase 2; Gαs, α subunit of the stimulatory G protein; Gs, stimulatory G protein; Gi, inhibitory G protein; LTCC, L-type calcium channel; PDE, phosphodiesterase; PI3K, phosphatidylinositol 3 kinase; PKA, cAMP-dependent protein kinase; PLB, phospholamban; RyR, ryanodine receptor; SERCA, sarcoplasmic/endoplasmic reticulum calcium ATPase; SR, sarcoplasmic reticulum.

The β1-AR couples to Gs protein while the β2-AR and the β3-AR couple to both Gs and Gi proteins [22]. The β2-AR can ‘switch’ its G protein-coupling preference from Gs to Gi and this is dependent on the phosphorylation of the β2-AR by PKA [23]. Gi inhibits the activity of the Gs-activated AC in synthesizing cAMP. In cardiomyocytes, β2-AR-Gi signaling restricts the inotropic response mediated by the Gs-AC-cAMP-PKA signaling cascade [24] (Figure 1). Studies on rabbit, eel, frog, rat, dog, human and other species have found that β3-ARs generally mediate a negative inotropic effect or are nonfunctional in the heart [14,25]. The Gi and the nitric oxide (NO) pathways have been implicated in the negative effect of the β3-AR but details of the signal transduction mechanism (e.g. the link between Gi protein-coupling and the production of NO and the number of cell types involved) are still unclear [25].

β-Adrenergic stimulation is a powerful means to increase cardiac output in the fight-or-flight state but if the stimulation is continued, it will cause harm to the heart by promoting detrimental β‑AR signaling. Dysregulation of the cardiac β-AR signaling is central to the pathogenesis of heart failure. In compensated heart failure, cardiac output and blood pressure is maintained, despite with a lower cardiac reserve, by activation of neuroendocrine axes, for instance the SNS and the renin-angiotensin-aldosterone system. A sustained increase of catecholamine levels is a leading feature of this mechanism [18]. In the failing hearts, chronic sympathetic stimulation ultimately leads to cardiomyocyte death and maladaptive cardiac remodeling, precipitating decompensation [26].

β1-AR signaling and β2-AR signaling are diverged in their effects on cardiomyocyte survival and death. Studies have implicated the central role of the β1-AR in heart failure. Of particular importance is that chronic stimulation of the β1-AR activates a cardiotoxic (pro-apoptotic) Ca2+/calmodulin-dependent protein kinase II (CaMKII)‑dependent signal transduction pathway [27]. In contrast, β2-AR signaling is largely cardioprotective. In particular, β2-AR signaling involves a Gi-Gβγ-phosphatidylinositol 3 kinase (PI3K)-protein kinase B (Akt) antiapoptotic pathway [28] (Figure 1). This is, however, a greatly simplified picture. As will be discussed in the following section, regulatory modification of β-AR signaling can have profound impacts on the biological effects produced.

Persistent stimulation of the cardiac β-AR induces cardiac hypertrophy and fibrosis and cardiomyocyte apoptosis [29], and these pathological changes are the basis of cardiac remodeling in heart failure. Evidence suggests that paracrine interactions between cardiomyocytes, cardiac fibroblasts and other cell types are key to this process. Cardiac fibroblasts play an important role in cardiac remodeling, fibrosis and hypertrophy due to their involvement in cell proliferation, extracellular matrix production and autocrine/paracrine signaling. A study on a transgenic mouse model has shown that sustained activation of the β2-AR-Gs-PKA pathway in myofibroblasts promotes cardiac hypertrophy possibly via a paracrine effect [30]. In addition, activation of the β2-AR in cardiac fibroblasts upregulates the pleotropic pro-inflammatory cytokine interleukin-6, also a hypertrophic factor in the heart, via different signaling mechanisms [31,32]. Conversely, paracrine factors released from cardiomyocytes in response to sustained β-adrenergic stimulation may promote the growth of cardiac fibroblasts and consequently induce myocardial fibrosis [33]. A recent study has identified the adhesive cell surface protein galectin-3 as a key regulatory component in promoting inflammatory cell mobilization, fibroblast activation and myocardial fibrosis [34]. Galectin-3 is upregulated in cardiac-restricted β2-AR transgenic mice [34], and its serum levels are also increased in human heart failure [35] and post-myocardial infarction [36]. Another study has found that β-adrenergic stimulation enhances cardiac expression of galectin-3 through the Hippo signaling pathway [37]. Interestingly, β2-adrenergic stimulation in cardiac fibroblasts alone promotes autophagy and this is correlated with an enhanced degradation of collagen. Thus, β2-AR signaling in cardiac fibroblasts may actually reduce cardiac fibrosis [38].

2.3.Regulation of Cardiac β-AR Signaling and Its Implications

In heart failure, chronic elevation of catecholamines leads to downregulation of the cardiac β1-AR while the expression level of the cardiac β2-AR remains unchanged [39]. Reduced transcription of the β1-AR is the primary cause of the downregulation [40]. Hence, the ratio of β1-AR to β2-AR drops from about 4:1 in the normal human heart to about 3:2 in the failing heart. This is accompanied by subsensitivity of the β1-AR and β2-AR signal transduction pathways, a leading cause of cardiac insufficiency in heart failure [39].

Downregulation of the cardiac β1-AR is viewed as a protective mechanism to counteract excessive sympathetic stimulation but it also aggravates contractile dysfunction in heart failure. Despite of the downregulation, residual β1-ARs still mediate a significant cardiotoxic effect of catecholamines. The result is a progressive loss of myocytes and myocardial contractile function in the failing heart [41,42]. Given the important role of this process in the pathogenesis of heart failure, the use of β1-AR antagonists becomes logical. Indeed, clinical data have suggested that β1-AR antagonists not only blocked the cardiotoxic β1-AR pathway and improved survival in heart failure [43], but also restored the number of downregulated cardiac β1-ARs and improved cardiac function in the long run [44,45].

Heart failure is also associated with upregulation of other proteins involved in β‑AR signaling, for instance Gi proteins and GRK2 [44,46]. Gi proteins and GRK2 are negative regulators of β‑AR-Gs-cAMP signaling in cardiomyocytes (Figure 1). Their upregulation in heart failure contributes to contractile dysfunction as detailed below.

Manipulation of β-AR levels in genetic animal models has provided important insights into the role of β-ARs in heart failure. In cardiac-specific overexpression mouse models, β2-ARs at moderate levels (30-fold over normal) reduced the incidence of heart failure while β2-ARs at high levels (300-fold or higher) led to heart failure [47,48]. Similarly, overexpression of the β1-AR in mouse hearts for 15-fold led to spontaneous heart failure and lower levels of the β1-AR (5-fold) led to cardiac hypertrophy, consistent with the detrimental effect of persistent β1 stimulation [49].

Upon β-adrenergic stimulation, β-AR signaling is attenuated through cooperative actions of a series of regulatory molecules. This regulatory process, known as desensitization, reduces the responsiveness of a GPCR to continuous or repeated stimuli. In one of the most well-characterized processes called homologous desensitization, activation of β-ARs (β1 and β2) promotes their own phosphorylation by GRKs at specific sites on their CTs. β-Arrestins are then recruited to the phosphorylated receptor sites to form complexes with the β-ARs. β‑Arrestins sterically inhibit the coupling of β‑ARs with G proteins, thus weakening G protein signaling. They also interact with proteins in the endocytic machinery such as clathrin to initiate receptor internalization. Internalized receptors are degraded or recycled, and thus losing their abilities temporally or permanently to interact with agonists. Agonist-induced receptor internalization is more rapid and obvious in β2-ARs, but less sensitive in β1-ARs [50]. Overall, homologous desensitization reduces G protein-mediated signal transduction, and is believed to produce a beneficial effect on the failing heart by restricting the β1-AR-mediated cardiotoxic pathway [22], but it may also, to a lesser extent, contribute to the functional decline of the failing heart by causing β2-AR subsensitivity [39].

Apart from GRKs, phosphorylation of GPCRs by second messenger-dependent protein kinases (e.g. PKA and protein kinase C (PKC)) provides another means for GPCR signal regulation, that is, heterologous desensitization. The β2-AR is among the most well-studied GPCRs in the process of heterologous desensitization. Stimulation of Gs protein- or Gq protein-coupled receptors promotes phosphorylation of the β2-AR by PKA and PKC, respectively, leading to functional uncoupling of the β2-AR from the Gs protein-mediated pathway without receptor internalization [51]. Heterologous desensitization of the β2-AR is a common mechanism for the suppression of β‑AR-stimulated myocyte contractile function by different hormones [51]. Importantly, it is believed to play a major role in causing β2-AR subsensitivity in heart failure due to upregulation of Gi proteins and, hence, augmentation of the inhibitory effect on Gs-cAMP signaling [44]. Studies have shown that other proteins (e.g. PDE4D and β‑arrestin2, see below) also participate in heterologous desensitization of the β2-AR [51].

Cardiomyocytes display a highly sophisticated membrane structure. Small pits (caveolae) and invaginations (transverse or T-tubules) exist on the surface of the plasma membrane of a cardiomyocyte. T-tubules extend from the cell surface deep into the cell and connected with longitudinal tubules to form the transverse-axial tubules system, which serves to conduct action potential and extracellular Ca2+ to the vicinity of each of the smallest unit of contraction – the sarcomere. Integrity of this membrane structure is critical for spatial synchronization of calcium-induced calcium release and, hence, coordinated contraction of sarcomeres throughout the cardiomyocyte upon induction by an action potential. Heart failure is manifested by an aberrant change of this membrane structure and relocation of proteins in it. This can cause important changes on the compartmentalized signals of cAMP and Ca2+ as discussed below. In late-stage heart failure, disordered Ca2+ waves occur more readily in cardiomyocytes, promoting arrhythmogenesis [52].

In a healthy cardiomyocyte, β1-ARs are evenly distributed on the sarcolemma while β2-ARs are located on T-tubules and caveolae [53]. In addition, β2-ARs are generally located within signal complexes containing LTCC, PDEs, Gi, and other proteins involved in signal transduction [51,54,55]. In particular, PDE4 limits diffusion of cAMP, so that the β2-AR-mediated cAMP signal is generally confined to the vicinities of T-tubules and caveolae as opposed to the β1-AR-mediated cAMP signal which is global [52,54] (Figure 1). Our recent study has found that β2-adrenergic stimulation transforms a β1-AR-mediated cAMP signal from the global mode to the local mode at nanometer level [56]. We call this PDE4-dependent β2-AR signal localization spillover effect ‘offside’ compartmentalization. It prevents the β1-stimulated PKA activity from acting on RyRs adjacent to activated β2-ARs, which may to some extent limit the β1-AR-mediated cardiotoxicity and sharpen the transient β1-AR response of sympathetic stimulation. Moreover, the β2-AR is directly associated with an LTCC in a signal complex that also contains a Gs protein, AC, PKA, and the counterbalancing phosphatase PP2A (type 2A protein phosphatase) [57]. Thus, the compartmentalized (offside or otherwise) β2-AR signaling is uncoupled from a global cAMP signal and phosphorylation of most PKA substrates in a cardiomyocyte but can still effectively enhance ICa,L and a contractile response [58].

Heart failure is characterized by progressive loss of T-tubules in cardiomyocytes. β2-ARs originally located on T-tubules are relocated to cell crests of cardiomyocytes. This leads to the loss of cAMP compartmentation with the production of a cAMP signal similar to the one elicited by β1 stimulation [52]. LTCCs also exhibit similar relocation from T-tubules to cell crests in heart failure and this change has been implicated as a pathological basis for arrhythmogenesis at single cell level [59]. Moreover, β2-ARs and LTCCs have been modelled to locate in cell crests and T-tubules of a cardiomyocyte in silico, which reveals enhanced phosphorylation of LTCC by PKA and, hence, increased ICa,L when they are present in cell crests [60]. Thus, spatial redistribution of cardiac β2-ARs in heart failure may enhance the pro-arrhythmogenic effect of β-AR-cAMP signaling.

GRKs and β-arrestins can regulate GPCR signaling by mechanisms other than homologous desensitization. The GRK family regulates various GPCRs. Among them, GRK2, GRK3 and GRK5 are highly expressed in the human heart. GRK2 and GRK5 are expressed in almost all cardiac cells and upregulated in heart failure, whereas GRK3 is detected only in cardiomyocytes [61]. GRK2 and GRK5 phosphorylate different sites on the CT of the β2-AR and this has biological implications [62]. GRK3 phosphorylates α1-ARs in the heart and plays no role on β-AR signaling [63].

Numerous studies have pinpointed GRK2 as a main culprit for cardiodepression, cardiac hypertrophy and maladaptive cardiac remodeling in the pathogenesis of heart failure [64–70], as upregulation of GRK2 occurs early in human heart diseases and is partly responsible for the diminished responses to adrenergic stimulation [71]. Particularly, in a murine model of pressure overload-induced heart failure, cardiac-specific overexpression of β2-ARs with deletion of all PKA phosphorylation sites (β2-AR PKA-) showed earlier disease-onset, more severe structural and functional cardiac damage and higher cardiac expression of GRK2 and Gi proteins as compared with mice overexpressing wild-type human β2-ARs or β2-ARs with deletion of all GRK phosphorylation sites [64]. Importantly, the blunted β-AR-mediated positive inotropic effect in the β2-AR PKA- mice could be rescued by the Gi inhibitor pertussis toxin. These results indicate that GRK2-induced phosphorylation of β2-ARs results in Gi-biased β2-AR signaling, linking upregulation of GRK2 to cardiodepression, maladaptive cardiac remodeling and heart failure. In heart failure, the GRK2-promoted Gi signaling not only inhibits the β2-AR-mediated positive inotropic effect but also cross-inhibits the β1-AR-mediated positive inotropic effect. Increased cardiac expression of Gi plays a central role in the cross-inhibition. Recent structural biology data have revealed the structural basis of receptor-coupling to different G proteins by both the β1- and β2-AR [72,73]. Under normal conditions, the coupling preference to Gi is much lower than that to Gs for both β‑AR subtypes. Nevertheless, current evidence is in favor that the β2-AR-Gi pathway plays an important role in the pathophysiology of human heart failure and Takotsubo syndrome [44,74,75]. Activation of this pathway induces cardiodepression, but at the same time produces a cardioprotective effect [74,75], corroborating the perception that β2-AR-Gi signaling is equivalent to an endogenous β1-AR antagonist in the heart [31]. Similar to the use of β-blockers in heart failure where low doses of them improve cardiac function in the long run whereas high doses of β‑blockers induce acute cardiac decompensation, therefore, the GRK2-promoted β2-AR-Gi signaling needs to be kept in check to avoid functional impairment while allowing delivery of a right amount of protective signals to the failing heart.

Increased levels of GRK5 in the failing hearts has been suggested to be a maladaptive mechanism due to the key role of GRK5 in cardiac hypertrophy [76,77]. Interestingly, carriers of the GRK5-Leu41 allele, which confers more β-AR desensitization and protection against catecholamine-induced cardiomyopathy in mice, have lower mortality than the wild-type GRK5-Gln4 allele carriers in African Americans with heart failure or cardiac ischemia [78]. Additionally, GRK5 has been implicated in β1-AR-mediated β-arrestin2 signaling and the resultant cardiac fibrosis-inducing effect of the β-blocker metoprolol in wild-type mice and in neonatal rat cardiomyocytes [79]. In this study, the β-blocker acted as an antagonist for the G protein-dependent pathway and as an agonist for the β-arrestin-dependent pathway. Phosphorylation of β1-ARs by GRK5 is the trigger for the recruitment of β‑arrestins that set off signal transduction (presumably via activation of ERK) in a G protein-independent manner. This represents a classical example of β‑arrestin-biased signaling, and metoprolol is defined as a β‑arrestin-biased ligand in this paradigm. However, it should be note that β‑blockers atenolol, metoprolol and propranolol also prevent cardiac fibrosis induced by β‑adrenergic stimulation in cardiomyocytes [33]. The β‑blocker carvedilol has also been suggested to be a β‑arrestin-biased ligand at the β1-AR and the GRK5/GRK6-dependent β‑arrestin-biased signaling it induced is protective to cardiomyocytes via upregulation of multiple cardioprotective microRNAs [80]. GRK5 exhibits both harmful and protective effects on the heart, which needs to be further studied.

Interestingly, GRK2 and β-arrestin1 have been found to be involved in the offside compartmentalization effect of β2-AR stimulation on β1-AR signaling [56]. Activation of the β2-AR leads to the translocation of GRK2 to the plasma membrane. GRK2 can also act on a β1-AR in close proximity to the activated β2-AR. Phosphorylation of the β1-AR at the CT by GRK2 then triggers recruitment of β-arrestin1 to the β1-AR, and the β-arrestin1 subsequently attracted PDEs to come close to the β1-AR. The latter limits β1-AR signaling by hydrolyzing cAMP. In contrast, heterologous desensitization of the β2-AR by activation of other GPCRs involves β-arrestin2 but not GRKs. Phosphorylation of the β2-AR by second messenger kinases likely triggers recruitment of β-arrestin2 that brings along PDE4D to limit β2-AR-cAMP signaling [51].

CaMKII is an important molecule for the transmission of Ca2+ signals. Four isoforms of CaMKII exist, in which the δ isoform is the major one in mammalian hearts. CaMKII positively regulates Ca2+ signaling in cardiomyocytes through phosphorylation of LTCCs, RyRs and PLB [81,82]. Knockout of CaMKIIδ decreases β-AR-stimulated ICa,L response in ventricular myocytes and reduces cardiac reserve and heart rate in response to β-adrenergic stimulation in mice, suggesting that CaMKIIδ is involved in the β-AR-mediated cardiac stimulatory effect [83].

We have found that β1-adrenergic stimulation produces an early but rapidly desensitized cAMP-PKA signal, followed by a slow-onset but sustained CaMKII signal and both of these signals could facilitate biological processes such as SR Ca2+ refilling in cardiomyocytes [84]. Recently, a negative feedback mechanism on CaMKII activity regulation was identified in cardiomyocytes, in which CaMKII phosphorylates and activates PDE4D, causing a reduction in cAMP signaling which positively regulates CaMKII activity [85]. This may provide a mechanistic explanation for the rapid desensitization of the β1-AR-cAMP signal in our observation. Our study suggests that the β1-AR-mediated myocardial contractile response is sustained by switching from cAMP-PKA signaling to CaMKII signaling upon prolonged stimulation [84] (Figure 1).

Activation of CaMKII has been implicated in the detrimental effects of β-AR signaling on the heart. Particularly, sustained β1-adrenergic stimulation has been shown to induce cardiomyocyte apoptosis through a CaMKII-dependent but PKA-independent pathway [86]. The involvement of CaMKII in β‑AR-mediated cardiomyocyte apoptosis has also been demonstrated in vivo [87]. Moreover, CaMKIIδ also participates in β-AR-mediated SR Ca2+ leak in cardiomyocytes by enhancing RyR2 phosphorylation and this mechanism may be involved in the induction of cardiomyopathy in mice subjected to chronic catecholamine stimulation [82]. The β1-AR-activated CaMKII signal is also implicated in catecholamine-induced cardiac hypertrophy in neonatal rat cardiomyocytes [88]. In the failing hearts, activated CaMKII phosphorylates RyRs and subsequently promotes SR Ca2+ leakage, delayed afterdepolarization and tachycardia [89]. Inhibition of CaMKII but not PKA decreases RyR2 activity and arrhythmogenic Ca2+ release in response to β‑adrenergic stimulation [90,91].

Different signal transduction mechanisms have been proposed to relay sustained β‑adrenergic stimulation to the activation of CaMKII. These mechanisms may be categorized according to the involvement of exchange protein activated by cAMP (Epac)‑dependent and/or PKA. Epac is a signaling molecule responsive to intracellular cAMP elevation independent of PKA [92,93]. In one of the most extensively studied Epac-dependent mechanism, stimulation of the β1-AR induces the formation of a complex containing β-arrestin, Epac1 and CaMKII, leading to CaMKII activation [94]. It is proposed that the β-arrestin that binds to the CT of the β1-AR assumes a specific conformation which keeps CaMKII and Epac in a stable complex with the receptor, whereas the conformation of the β2-AR-bound β-arrestin is incapable of recruiting CaMKII and Epac. This explains why the β1-AR, but not the β2-AR, is the relevant receptor for activating CaMKII signaling [94]. Further studies in cardiomyocytes suggests that β-AR-CaMKII signaling involves NO, and a β‑AR-cAMP-Epac-PI3K-Akt-NOS1-CaMKII signaling cascade has been proposed to link β‑adrenergic stimulation to increased SR Ca2+ leaks [95,96]. In the PKA-dependent mechanism, increased [Ca2+]i leads to CaMKII activation and the source of Ca2+ is cAMP-PKA-PLB-SR-mediated Ca2+ pool refilling [97]. This pathway has been shown to transmit a pro-cell death signal of β-adrenergic stimulation in cardiomyocytes. Interestingly, the same study has also discovered a cardioprotective β1-AR-cAMP-Epac2-Rap1-Rac-ERK pathway. The regulatory effects of PKA on CaMKII expression and activity have also been demonstrated in vivo in transgenic mice with cardiac-specific overexpressed PKI (a peptide inhibitor of PKA). In particular, acute β-stimulation-induced CaMKII signaling that leads to heart rate and cardiomyocyte contractility increment was sensitive to PKI [20]. A recent study has proposed yet another mechanism of β1-AR-stimulated CaMKII activation in cardiomyocytes that involves PKA, GRK5, SAP97, β-arrestin2 and Epac2 [27].

Recent experimental progress also holds promise for CaMKII inhibition in the treatment of heart failure and related myocardial abnormalities [98–100].

Acute β‑adrenergic stimulation with dobutamine produced positive inotropic effects in mouse hearts while sustained β‑adrenergic stimulation with isoproterenol resulted in cardiac hypertrophy in mice. Both effects could be attenuated in transgenic mice lacking B56α, a PP2A regulatory subunit highly expressed in the heart, indicating that PP2A is a potential regulator of cardiac function after β-AR agitation [101]. Another study has found that overexpression of the Raf kinase inhibitor protein produced a well-tolerated, persistent increase in cardiac contractility mediated by the β1-AR via simultaneous activation of the β2-AR subtype [102]. Signal transducers and activators of transcription 3 has been suggested to be a key transcriptional regulator of β‑AR-mediated cardiac stress adaptation, pathological remodeling and heart failure [103]. Recently, we have found that 5-HT2B receptors formed heterodimers with β2-ARs, promoting cardioprotective Gi signaling with β2-adrenergic stimulation [2]. We have also found that β1-ARs and receptors of advanced glycation end products physically interacted with each other. Stimulation with either of their agonists generated a pro-cell death CaMKII signal with significant signal crosstalk, leading to cardiomyocyte injury and cardiac remodeling [104]. The G protein-coupled estrogen receptor 1, a Gi protein-coupled receptor, has been suggested to moderate myocardial Ca2+ dynamics and cardiomyocyte contraction by limiting β1-AR-cAMP-PKA signaling [105]. This study may provide a mechanistic basis for the effects of estrogen on cardiac function regulation, and possibly cardioprotection.

Non-coding micro single-stranded RNA (microRNA, miRNA) is a class of small-molecular transcription regulator. Studies have shown that miR-133 regulates multiple components of β1-AR signaling and has a cardioprotective effect by reducing apoptosis and fibrosis in pressure overload-induced heart failure [106]. miR-145 improves cardiac dilatation and fibrosis and alleviates heart failure-related cardiac remodeling by upregulating β2-ARs and downregulating CaMKII [107]. Downregulation of non-coding RNA circ-HIPK3 attenuates fibrosis and maintains myocardial function after myocardial infarction in mice through miR-17-3p [108].

3.Interventions of β-AR Signaling and Cardiovascular Disease

3.1.β-Blockers

β-Blockers are widely used in tachyarrhythmias, coronary artery disease, hypertension and heart failure. Their use is considered to be a milestone in the treatment of heart failure [109]. Over the past six decades, three generations of β-blockers have been developed: the first generation nonselective β-blockers, the second generation cardioselective β-blockers (selectively antagonizing the β1-AR), and the third generation vasodilating β-blockers (exhibiting vasodilating effects as a result of α1-AR blockade or endothelial nitric oxide synthase (eNOS) activation via β2-AR or β3-AR agonism). Clinical studies have shown that the cardioselective β-blockers bisoprolol and metoprolol and the vasodilating β-blockers carvedilol and nebivolol significantly reduce morbidity and mortality from heart failure [110]. The reason causing the different efficacies of β-blockers in heart failure treatment is not totally clear, but it is quite certain that the side effect profiles of nonselective β-blockers are the key issue disfavoring their long-term use in heart failure and essential hypertension.

Propranolol, which has been widely used in clinical practice to treat arrhythmias, is a representative of the first generation β-blockers. They have similar affinities for the β1-AR and the β2-AR, so they are called nonselective β-blockers. Blocking of the β1-AR in the heart is beneficial in both heart failure and myocardial ischemia [111], but blocking of the β2- or β3-AR is not beneficial because it antagonizes β2-AR-mediated bronchodilation and the mechanisms of cardiac protection and vasodilation mediated by both the β2- and β3-AR [28,112]. In blood vessels, β2- and β3-ARs are present in vascular smooth muscle cells and endothelial cells and promote vasodilation through the eNOS-NO-cGMP-protein kinase G pathway. The first generation β-blockers may cause fatal side effects such as bronchospasm due to the antagonism of the β2-AR. Therefore, the use of these blockers is strictly prohibited in patients with asthma. A recent experimental study has suggested that propranolol may be useful in the treatment of congenital heart disease by promoting cardiomyocyte cytokinesis in neonates and could, therefore, increase the number of cardiomyocytes and confer benefit after myocardial infarction in adulthood [113].

In order to find drugs that avoid the bronchial side effects of the first generation β-blockers, the second generation β-blockers have been developed. The most representative drugs are atenolol and metoprolol [114,115]. Their affinities for the β1 subtype are higher than those for the β2 and β3 subtypes. At low doses, they inhibit β1-AR-mediated effects in the heart, but not β2-AR-mediated vasodilation or bronchodilation. At higher doses, the selectivity to the β1-AR is lost. Therefore, the use of the second generation β‑blockers should still be considered with caution in patients with respiratory diseases.

In addition to its effects on cardiomyocytes, new research has shown that metoprolol also acts on the hematopoietic system. Metoprolol reduced neutrophil infiltration by blocking the β1-ARs on neutrophils. In the absence of neutrophils, metoprolol lost its protective effect on acute myocardial infarction, indicating that neutrophils are one of the targets of metoprolol in the treatment of acute myocardial infarction [116]. Other studies have shown that the functional improvement of metoprolol after myocardial infarction is dependent on the β3-AR [117]. Therefore, patients with β3-AR dysfunction may be insensitive to metoprolol. In addition, the treatment benefit of metoprolol is not correlated with a change in CaMKII activity in experimental and human heart failure [118]. Therefore, direct inhibition of CaMKII on top of β1-AR blockade might bring additional benefit to heart failure therapy.

The third generation β‑blockers are drugs with vasodilating properties. This vasodilating activity is beneficial because it reduces peripheral vascular resistance and lowers blood pressure while maintaining or improving cardiac output and left ventricular function. Carvedilol and labetalol produce their vasodilating effects by antagonizing the α1-AR in vascular smooth muscle cells. Nebivolol produces NO-mediated vasodilation by means of activating the β2- and/or β3-ARs in endothelial cells [119]. Vasodilating β-blockers also have neutral (labetalol and nebivolol) or beneficial (carvedilol) effects on glucose and lipid metabolism, whereas non-vasodilating β-blockers tend to have negative effects on glucose and lipid metabolism [120].

Carvedilol has the most evidence for reducing morbidity and mortality in patients with ischemic heart failure [110]. Its effectiveness might be due to its unique biological activity. Carvedilol possesses an antioxidant activity, and could prevent cardiomyocyte apoptosis induced by doxorubicin and the reduction of β1-AR expression induced by hydrogen peroxide or doxorubicin [121]. One possible mechanism is that carvedilol contains a large aromatic amine substituent that is not present in other β-blockers. This large group forms a unique contact with the β1-AR ligand binding pocket, causing a conformational rearrangement on the extracellular surface, thus burying the redox-sensitive disulfide bond in the receptor structure. Another possible explanation for its beneficial effect on heart failure is that carvedilol promotes the accumulation of N-terminal truncated β1-ARs, which constitutively (without ligand induction) activate non-classical Gi protein-mediated Akt signaling in carvedilol-treated cardiomyocytes [121]. Unlike the classical β1-AR-coupled Gs protein-mediated signal transduction pathway, this study and some others suggest that carvedilol promotes the recruitment of Gi proteins to the β1-AR (but not β2-AR) and initiates unique signal transduction [80,122]. Carvedilol also reversed the deterioration of cardiac function in diabetes which seems to be related to β-arrestins [123].

Specific blocking of the β1-AR in Golgi bodies of cardiomyocytes has recently been found to prevent cardiac hypertrophy [124], which has a potential to be developed into new-generation β-blockers.

β-Blockers differ a lot in their pharmacological properties, for instance receptor subtype specificity, intrinsic sympathomimetic activity, membrane-stabilizing activity, lipophilicity, side effects and effectiveness in treating heart failure. Thus, they need to be studied individually rather than by category [125].

3.2.Other Potential β-AR-related Therapies

Recent evidence has shown that β2-AR agonists may be potentially useful in cardiovascular diseases [126,127]. Moreover, β2-AR agonists with high specificity and activities have been designed and synthesized continuously [128,129]. Biased ligands and positive or negative allosteric modulators for β-ARs have also been designed and reported [130–136]. In addition, some researchers have activated β2-AR-mediated β‑arrestin-biased signaling by expressing allosteric regulatory peptides without the participation of agonists, and have confirmed the protective effect of this signal on cardiomyocytes [137]. It should be noted that long-term use of β2-AR agonists has been shown to increase the risk of heart failure in some meta-analysis although the overall cardiovascular safety of β2-AR agonists has been confirmed in many studies [138]. Whether β2-AR ligands should be used alone or in combination with other drugs such as β1-AR blockers in treating cardiovascular disease requires more clinical validation.

In addition, gold nanoparticles, a widely used drug transport carrier, have a property to accumulate in the heart through blood circulation, and have been found to play an anti-myocardial hypertrophy role by downregulating the β1-AR and its downstream ERK pathway [139]. Thus, gold nanoparticles may be a potential multifunctional material for drug transport and anti-myocardial hypertrophy therapy. One possibility is to use gold nanoparticles to deliver small molecule inhibitors, microRNAs or gene therapy to combat the deleterious effect of GRKs in heart diseases [61]. Other new development includes light-controlled drugs with a high level of β1-/β2-AR selectivity [140]. Given the presence of multiple phosphorylation sites on GPCRs that are associated with distinct signal transduction pathways, mechanisms that generate different receptor phosphorylation patterns may be exploited for the development of pathway-selective drugs.

4.Future Prospects

Cardiovascular disease is a leading public health problem worldwide. In particular, heart failure is one of the most common complications of cardiovascular disease and an important cause of hospitalization and immature death. There is a large and growing demand of better and more personalized cardiovascular drugs. β‑AR signal transduction plays an important role in cardiac physiology and the pathophysiology of heart failure, and it is also a therapeutic target with a long history. Technological advances, especially functional genomics and structure-based drug discovery, have enabled the continual discovery of existing and novel GPCR targets (i.e. orphan receptors) and GPCR-targeting drugs. Especially, emerging studies have suggested that biased GPCR signaling may be exploited as new therapeutic targets. Recently, the structural biology approach has begun to illuminate the structural basis of GPCR activation by different ligands or allosteric modulators, and hold promise for elucidating the structure-function relationships of GPCR signal complexes. It is anticipated that new drugs based on these structural understandings will be developed in the near future. Therefore, new disease models and transgenic systems need to be created along the way to promote basic research on cardiovascular science and drug development.

Author Contributions: All authors discussed the materials and the content of the manuscript. Y.S, A.Y.H.W and Y.Z wrote the manuscript text. Y.S and A.Y.H.W prepared the figure. All authors reviewed and edited the manuscript.

Funding: This work was financially supported by Beijing Municipal Natural Science Foundation (7224348), National Key R&D Program of China (2018YFA0800501) and the National Natural Science Foundation of China (92168114).

Conflicts of Interest: The authors declare no conflict of interest.

References

  1. Kenakin T. Biased receptor signaling in drug discovery. Pharmacol.Rev., 2019, 71(2): 267-315. DOI: https://doi.org/10.1124/pr.118.016790
  2. Song Y.; Xu C.J.; Liu J.F.; et al. Heterodimerization with 5-HT2BR is indispensable for β2AR-Mediated cardioprotection. Circ.Res.,2021, 128(2): 262-277. DOI: https://doi.org/10.1161/CIRCRESAHA.120.317011
  3. Wnorowski A.;Jozwiak K.Homo- and hetero-oligomerization of β2-adrenergic receptor in receptor trafficking, signaling pathways and receptor pharmacology. Cell Signal., 2014, 26(10): 2259-2265. DOI: https://doi.org/10.1016/j.cellsig.2014.06.016
  4. Alexander S.; Christopoulos A.; Davenport P.P.; et al. THE CONCISE GUIDE TO PHARMACOLOGY 2019/20:G protein-coupled receptors, Br. J.Pharmacol., 2019, 176(Suppl 1): S21-S141. DOI: https://doi.org/10.1111/bph.14748
  5. Grisanti A.A.; Schumacher M.M.; Tilley G.G.; et al. Designer approaches for G Protein-Coupled receptor modulation for cardiovascular disease. JACC Basic Transl. Sci., 2018, 3(4): 550-562. DOI: https://doi.org/10.1016/j.jacbts.2017.12.002
  6. Laschet C.; Dupuis N.; Hanson J. The G protein-coupled receptors deorphanization landscape. Biochem.Pharmacol.,2018, 153: 62-74. DOI: https://doi.org/10.1016/j.bcp.2018.02.016
  7. Mouat A.A.; Coleman L.J.L.J.; Smith J.J. GPCRs in context: sexual dimorphism in the cardiovascular system. Br. J. Pharmacol., 2018, 175(21): 4047-4059. DOI: https://doi.org/10.1111/bph.14160
  8. Virani S.S.; Alonso A.; Benjamin J.J.; et al. Heart disease and stroke statistics-2020 update: a report from the american heart association. Circulation, 2020, 141(9): e139-e596.
  9. Li P.; Fu Y.X.; Ru J.L.; et al. Insights from systems pharmacology into cardiovascular drug discovery and therapy. BMC Syst. Biol.,2014,8: 141. DOI: https://doi.org/10.1186/s12918-014-0141-z
  10. Huang B.S.; Chen A.; Ahmad M.; et al.Mineralocorticoid and AT1 receptors in the paraventricular nucleus contribute to sympathetic hyperactivity and cardiac dysfunction in rats post myocardial infarct. J. Physiol., 2014, 592(15): 3273-3286. DOI: https://doi.org/10.1113/jphysiol.2014.276584
  11. Wang J.L.; Hanada K.J.; Gareri C.; et al. Mechanoactivation of the angiotensin II type 1 receptor induces β-arrestin-biased signaling through Gαi coupling. J. Cell.Biochem., 2018, 119(4): 3586-3597. DOI: https://doi.org/10.1002/jcb.26552
  12. Woo A.Y.H.; Komuro I.; Xiao R.P. Biased agonism at the angiotensin receptor: blocker and Calcium sensitizer at the same time. Circulation, 2017, 135(11): 1071-1074. DOI: https://doi.org/10.1161/CIRCULATIONAHA.117.027276
  13. Ryba M.M.; Li J.L.; Cowan L.L.;et al. Long-Term biased β-Arrestin signaling improves cardiac structure and function in dilated cardiomyopathy. Circulation, 2017, 135(11): 1056-1070. DOI: https://doi.org/10.1161/CIRCULATIONAHA.116.024482
  14. Michel M.C.; Harding S.E.; Bond R.A. Are there functional β3-adrenoceptors in the human heart?. Br. J.Pharmacol.,2011, 162(4): 817-822. DOI: https://doi.org/10.1111/j.1476-5381.2010.01005.x
  15. Zhou P.Z.; Pu T.T. Recounting cardiac cellular composition. Circ.Res., 2016, 118(3): 368-370. DOI: https://doi.org/10.1161/CIRCRESAHA.116.308139
  16. Myagmar B.E.; Flynn J.M.; Cowley P.M.; et al. Adrenergic receptors in individual ventricular myocytes: the beta-1 and alpha-1B are in all cells, the alpha-1a is in a subpopulation, and the beta-2 and beta-3 are mostly absent. Circ.Res., 2017, 120(7): 1103-1115. DOI: https://doi.org/10.1161/CIRCRESAHA.117.310942
  17. Lindenfeld J.; Cleveland J.C.Jr.; Kao D.P.;et al. Sex-related differences in age-associated downregulation of human ventricular myocardial β1-adrenergic receptors. J. Heart Lung Transplant., 2016, 35(3): 352-361. DOI: https://doi.org/10.1016/j.healun.2015.10.021
  18. Pecha S.; Geelhoed B.; Kempe R.; et al. No impact of sex and age on beta-adrenoceptor-mediated inotropy in human right atrial trabeculae. Acta physiol., 2021, 231(3): e13564. DOI: https://doi.org/10.1111/apha.13564
  19. Bers M.M. Sarcoplasmic reticulum Ca release in intact ventricular myocytes. Front. Biosci., 2002, 7: d1697-d1711. DOI: https://doi.org/10.2741/A873
  20. Zhang Y.; Wang E.E.; Zhang X.Y.; et al. Cardiomyocyte PKA ablation enhances basal contractility while eliminates cardiac β-Adrenergic response without adverse effects on the heart. Circ. Res., 2019, 124(12): 1760-1777. DOI: https://doi.org/10.1161/CIRCRESAHA.118.313417
  21. Mi X.Y.; Ding W.G.; Toyoda F.; et al. Selective activation of adrenoceptors potentiates IKs current in pulmonary vein cardiomyocytes through the protein kinase A and C signaling pathways. J. Mol. Cell.Cardiol., 2021, 161: 86-97. DOI: https://doi.org/10.1016/j.yjmcc.2021.08.004
  22. Xing G.; Woo Y.H.Y.H.; Pan L.; et al. Recent advances in β2-Agonists for treatment of chronic respiratory diseases and heart failure. J.Med. Chem., 2020, 63(24): 15218-15242. DOI: https://doi.org/10.1021/acs.jmedchem.0c01195
  23. Daaka Y.; Luttrell M.M.; Lefkowitz J.J. Switching of the coupling of the beta2-adrenergic receptor to different G proteins by protein kinase A. Nature, 1997, 390(6655): 88-91. DOI: https://doi.org/10.1038/36362
  24. Xiao R.P.; Ji X.; LakattaE.G. Functional coupling of the beta 2-adrenoceptor to a pertussis toxin-sensitive G protein in cardiac myocytes. Mol.Pharmacol., 1995, 47(2): 322-329.
  25. Arioglu-Inan E.; Kayki-Mutlu G.; Michel M.C. Cardiac β3-adrenoceptors-A role in human pathophysiology?. Br. J.Pharmacol., 2019, 176(14): 2482-2495. DOI: https://doi.org/10.1111/bph.14635
  26. Ali D.C.; Naveed M.; Gordon A.; et al. β-Adrenergic receptor, an essential target in cardiovascular diseases. Heart Fail Rev., 2020, 25(2): 343-354. DOI: https://doi.org/10.1007/s10741-019-09825-x
  27. Xu B.; Li M.H.; Wang Y.; et al. GRK5 controls SAP97-Dependent cardiotoxic β1 adrenergic Receptor-CaMKII signaling in heart failure. Circ.Res., 2020, 127(6): 796-810. DOI: https://doi.org/10.1161/CIRCRESAHA.119.316319
  28. Huang J.; Li C.Z.; Song Y.; et al. ADRB2 polymorphism Arg16Gly modifies the natural outcome of heart failure and dictates therapeutic response to beta-blockers in patients with heart failure. Cell Discov., 2018, 4: 57. DOI: https://doi.org/10.1038/s41421-018-0058-6
  29. Shin E.; KoK.S.; Rhee B.D.; et al. Different effects of prolonged β-adrenergic stimulation on heart and cerebral artery. Integr. Med. Res., 2014, 3(4): 204-210. DOI: https://doi.org/10.1016/j.imr.2014.10.002
  30. Imaeda A.; Tanaka S.; Tonegawa K.; et al. Myofibroblast β2 adrenergic signaling amplifies cardiac hypertrophy in mice. Biochem.Biophys. Res. Commun., 2019, 510(1): 149-155. DOI: https://doi.org/10.1016/j.bbrc.2019.01.070
  31. Tanaka S.; Imaeda A.; Matsumoto K.; et al. β2-adrenergic stimulation induces interleukin-6 by increasing arid5a, a stabilizer of mRNA, through cAMP/PKA/CREB pathway in cardiac fibroblasts. Pharmacol.Rese.Perspect., 2020, 8(2): e00590. DOI: https://doi.org/10.1002/prp2.590
  32. Bageghni S.A.; Hemmings K.E.; Zava N.; et al. Cardiac fibroblast-specific p38α MAP kinase promotes cardiac hypertrophy via a putative paracrine interleukin-6 signaling mechanism. FASEB J., 2018, 32(9): 4941-4954. DOI: https://doi.org/10.1096/fj.201701455RR
  33. Nuamnaichati N.; Sato V.H.; Moongkarndi P.; et al. Sustained β-AR stimulation induces synthesis and secretion of growth factors in cardiac myocytes that affect on cardiac fibroblast activation. Life Sci., 2018, 193: 257-269. DOI: https://doi.org/10.1016/j.lfs.2017.10.034
  34. She G.; HouM.C.; Zhang Y.; et al. Gal-3 (Galectin-3) and KCa3.1 Mediate Heterogeneous Cell Coupling and Myocardial Fibrogenesis Driven by betaAR (beta-Adrenoceptor) Activation. Hypertension, 2020, 75(2): 393-404. DOI: https://doi.org/10.1161/HYPERTENSIONAHA.119.13696
  35. Khadeja Bi A.; Santhosh V.; Sigamani K. Levels of galectin-3 in chronic heart failure: a Case-Control study. Cureus, 2022, 14(8): e28310. DOI: https://doi.org/10.7759/cureus.28310
  36. Al-Salam S.; Hashmi S.; JagadeeshG.S.; et al. Galectin-3: a cardiomyocyte antiapoptotic mediator at 24-Hour post myocardial infarction. Cell. Physiol. Biochem., 2020, 54(2): 287-302. DOI: https://doi.org/10.33594/000000220
  37. Zhao W.B.; Lu Q.; Nguyen N.N.; et al. Stimulation of β-adrenoceptors up-regulates cardiac expression of galectin-3 and BIM through the Hippo signalling pathway. Br. J. Pharmacol., 2019, 176(14): 2465-2481. DOI: https://doi.org/10.1111/bph.14674
  38. Aránguiz-Urroz P.; Canales J.; Copaja M.; et al. Beta(2)-adrenergic receptor regulates cardiac fibroblast autophagy and collagen degradation. Biochim.Biophys.Acta, 2011, 1812(1): 23-31. DOI: https://doi.org/10.1016/j.bbadis.2010.07.003
  39. Bristow M.R.; Hershberger R.E.; Port J.D.; et al. Beta 1- and beta 2-adrenergic receptor-mediated adenylate cyclase stimulation in nonfailing and failing human ventricular myocardium. Mol. Pharmacol., 1989, 35(3): 295-303.
  40. Bristow M.R.; Minobe W.A.; Raynolds M.V.; et al. Reduced beta 1 receptor messenger RNA abundance in the failing human heart. J. Clin. Invest., 1993, 92(6): 2737-2745. DOI: https://doi.org/10.1172/JCI116891
  41. Lymperopoulos A.; Rengo G.; Koch W.J. Adrenergic nervous system in heart failure: pathophysiology and therapy. Circ. Res., 2013, 113(6): 739-753. DOI: https://doi.org/10.1161/CIRCRESAHA.113.300308
  42. de LuciaC.; Eguchi A.; Koch W.J. New insights in cardiac beta-Adrenergic signaling during heart failure and aging. Front. Pharmacol., 2018, 9: 904. DOI: https://doi.org/10.3389/fphar.2018.00904
  43. Effect of metoprolol CR/XL in chronic heart failure:Metoprolol CR/XL Randomised Intervention Trial in Congestive Heart Failure (MERIT-HF). Lancet, 1999, 353(9169): 2001-2007. DOI: https://doi.org/10.1016/S0140-6736(99)04440-2
  44. Sigmund M.; Jakob H.; Becker H.; et al. Effects of metoprolol on myocardial beta-adrenoceptors and Gi alpha-proteins in patients with congestive heart failure. Eur. J. Clin. Pharmacol., 1996, 51(2): 127-132. DOI: https://doi.org/10.1007/s002280050172
  45. Kukin M.L.; Kalman J.; Charney R.H.; et al. 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(20): 2645-2651. DOI: https://doi.org/10.1161/01.CIR.99.20.2645
  46. Ungerer M.; Böhm M.; Elce J.S.; et al. Altered expression of beta-adrenergic receptor kinase and beta 1-adrenergic receptors in the failing human heart. Circulation, 1993, 87(2): 454-463. DOI: https://doi.org/10.1161/01.CIR.87.2.454
  47. Nguyen N.N.; Kiriazis H.; Ruggiero D.; et al. Spontaneous ventricular tachyarrhythmias in β2-adrenoceptor transgenic mice in relation to cardiac interstitial fibrosis. Am. J. Physiol., 2015, 309(5): H946-H957. DOI: https://doi.org/10.1152/ajpheart.00405.2015
  48. Dorn G.W. 2nd.;Tepe NM.; Lorenz J.N.;et al. Low- and high-level transgenic expression of beta2-adrenergic receptors differentially affect cardiac hypertrophy and function in Galphaq-overexpressing mice. Proc. Natl. Acad. Sci. U. S. A., 1999, 96(11): 6400-6405. DOI: https://doi.org/10.1073/pnas.96.11.6400
  49. Engelhardt S.; Hein L.; Wiesmann F.; et al. Progressive hypertrophy and heart failure in beta1-adrenergic receptor transgenic mice. Proc. Natl. Acad. Sci. U. S. A., 1999, 96(12): 7059-7064. DOI: https://doi.org/10.1073/pnas.96.12.7059
  50. Liang W.; Austin S.; Hoang Q.; et al. Resistance of the human beta 1-adrenergic receptor to agonist-mediated down-regulation. J. Biol. Chem., 2003, 278(41): 39773-39781. DOI: https://doi.org/10.1074/jbc.M304482200
  51. Shi Q.; Li M.H.; Mika D.; et al. Heterologous desensitization of cardiac β-adrenergic signal via hormone-induced βAR/arrestin/PDE4 complexes. Cardiovasc. Res., 2017, 113(6): 656-670. DOI: https://doi.org/10.1093/cvr/cvx036
  52. Judina A.; Gorelik J.; Wright P.T. Studying signal compartmentation in adult cardiomyocytes. Biochem. Soc. Trans., 2020, 48(1): 61-70. DOI: https://doi.org/10.1042/BST20190247
  53. Steinberg S.F. Beta(2)-Adrenergic receptor signaling complexes in cardiomyocyte caveolae/lipid rafts. J. Mol. Cell. Cardiol., 2004, 37(2): 407-415. DOI: https://doi.org/10.1016/j.yjmcc.2004.04.018
  54. Xiang Y.K. Compartmentalization of beta-adrenergic signals in cardiomyocytes. Circ. Res., 2011, 109(2): 231-244. DOI: https://doi.org/10.1161/CIRCRESAHA.110.231340
  55. Xiao R.P.; Cheng H.; Zhou Y.Y.; et al. Recent advances in cardiac beta(2)-adrenergic signal transduction. Circ. Res., 1999, 85(11): 1092-1100. DOI: https://doi.org/10.1161/01.RES.85.11.1092
  56. Yang H.Q.; Wang L.P.; Gong Y.Y.; et al. β2-Adrenergic stimulation compartmentalizes β1 signaling into nanoscale local domains by targeting the C-Terminus of β1-Adrenoceptors. Circ. Res., 2019, 124(9): 1350-1359. DOI: https://doi.org/10.1161/CIRCRESAHA.118.314322
  57. Davare M.A.; Avdonin V.; Hall D.D.; et al. A beta2 adrenergic receptor signaling complex assembled with the Ca2+ channel Cav1.2. Science, 2001, 293(5527): 98-101. DOI: https://doi.org/10.1126/science.293.5527.98
  58. Zhou Y.Y.; Cheng H.; BogdanovK.Y.;et al. Localized cAMP-dependent signaling mediates beta 2-adrenergic modulation of cardiac excitation-contraction coupling. Am. J. Physiol., 1997, 273(3 Pt 2): H1611-H1618. DOI: https://doi.org/10.1152/ajpheart.1997.273.3.H1611
  59. Sanchez-Alonso J.L.; Bhargava A.; O’Hara T,; et al.Microdomain-specific modulation of L-Type calcium channels leads to triggered ventricular arrhythmia in heart failure. Circ. Res., 2016, 119(8): 944-955. DOI: https://doi.org/10.1161/CIRCRESAHA.116.308698
  60. Loucks A.D.; O’Hara T.; Trayanova N.A. Degradation of t-tubular microdomains and altered cAMP compartmentation lead to emergence of arrhythmogenic triggers in heart failure myocytes: an in silico study. Front. Physiol., 2018, 9: 1737. DOI: https://doi.org/10.3389/fphys.2018.01737
  61. Hullmann J.; Traynham C.J.; Coleman R.C.; et al. The expanding GRK interactome: Implications in cardiovascular disease and potential for therapeutic development. Pharmacol. Res., 2016, 110: 52-64. DOI: https://doi.org/10.1016/j.phrs.2016.05.008
  62. Nobles K.N.; Xiao K.H.; Ahn S.;et al. Distinct phosphorylation sites on the β(2)-adrenergic receptor establish a barcode that encodes differential functions of β-arrestin. Sci. Signal., 2011, 4(185): ra51. DOI: https://doi.org/10.1126/scisignal.2001707
  63. Eckhart A.D.; Duncan S.J.; Penn D.B.; et al. Hybrid transgenic mice reveal in vivo specificity of G protein-coupled receptor kinases in the heart. Circ. Res., 2000, 86(1): 43-50. DOI: https://doi.org/10.1161/01.RES.86.1.43
  64. Zhu W.Z.; Petrashevskaya N.; Ren S.X.; et al. Gi-Biased β2AR Signaling Links GRK2 Upregulation to Heart Failure. Circ. Res., 2012, 110(2): 265-274. DOI: https://doi.org/10.1161/CIRCRESAHA.111.253260
  65. Schlegel P.; Reinkober J.; Meinhardt E.; et al. G protein-coupled receptor kinase 2 promotes cardiac hypertrophy. PloS one, 2017, 12(7): e0182110. DOI: https://doi.org/10.1371/journal.pone.0182110
  66. Travers J.G.;Kamal F.A.;Valiente-Alandi I.; et al. Activated Fibroblast Targeting of Gbetagamma-GRK2 After Myocardial Ischemia Attenuates Heart Failure Progression. J. Am. Coll. Cardiol., 2017, 70(8): 958-971. DOI: https://doi.org/10.1016/j.jacc.2017.06.049
  67. Woodall M.C.; Woodall B.P.; Gao E.H.; et al. Cardiac fibroblast GRK2 deletion enhances contractility and remodeling following ischemia/reperfusion injury. Circ.Res., 2016, 119(10): 1116-1127. DOI: https://doi.org/10.1161/CIRCRESAHA.116.309538
  68. AbdAlla J.; Graemer M.; Fu X.B.; et al. Inhibition of G-protein-coupled Receptor Kinase 2 Prevents the Dysfunctional Cardiac Substrate Metabolism in Fatty Acid Synthase Transgenic Mice. J. Biol. Chem., 2016, 291(6): 2583-2600. DOI: https://doi.org/10.1074/jbc.M115.702688
  69. Cannavo A.; Marzano F.; Elia A.; et al. Aldosterone jeopardizes myocardial insulin and beta-Adrenergic receptor signaling via G Protein-Coupled receptor kinase 2. Front. Pharmacol., 2019, 10: 888. DOI: https://doi.org/10.3389/fphar.2019.00888
  70. Powell J.M.; Ebin E.; Borzak S.; et al. Hypothesis: paroxetine, a G Protein-Coupled receptor kinase 2 (GRK2) inhibitor reduces morbidity and mortality in patients with heart failure. J. Cardiovasc. Pharmacol.Ther., 2017, 22(1): 51-53. DOI: https://doi.org/10.1177/1074248416644350
  71. Santulli G.; Campanile A.; Spinelli L.; et al.G protein-coupled receptor kinase 2 in patients with acute myocardial infarction. Am. J. Cardiol., 2011, 107(8): 1125-1130. DOI: https://doi.org/10.1016/j.amjcard.2010.12.006
  72. Ma X.Y.; Hu Y.F.; Batebi H.; et al. Analysis of β2AR-Gs and β2AR-Gi complex formation by NMR spectroscopy. Proc. Natl. Acad. Sci. U. S. A., 2020, 117(37): 23096-23105. DOI: https://doi.org/10.1073/pnas.2009786117
  73. Alegre K.O.; Paknejad N.; Su M.F.; et al. Structural basis and mechanism of activation of two different families of G proteins by the same GPCR. Nat. Struct. Mol. Biol., 2021, 28(11): 936-944. DOI: https://doi.org/10.1038/s41594-021-00679-2
  74. Nakano T.; Onoue K.J.; Nakada Y.; et al. Alteration of β-Adrenoceptor signaling in left ventricle of acute phase takotsubo syndrome: a human study. Sci. Rep., 2018, 8(1): 12731. DOI: https://doi.org/10.1038/s41598-018-31034-z
  75. Woo A.Y.; Song Y.; Xiao R.P.; et al. Biased β2-adrenoceptor signalling in heart failure: pathophysiology and drug discovery. Br. J. Pharmacol., 2015, 172(23): 5444-5456. DOI: https://doi.org/10.1111/bph.12965
  76. Martini J.S.; Raake P.; VingeL.E.; et al. Uncovering G protein-coupled receptor kinase-5 as a histone deacetylase kinase in the nucleus of cardiomyocytes. Proc. Natl. Acad. Sci. U. S. A., 2008, 105(34): 12457-12462. DOI: https://doi.org/10.1073/pnas.0803153105
  77. Gold J.I.; Gao E.H.; Shang X.Y.;et al. Determining the absolute requirement of G protein-coupled receptor kinase 5 for pathological cardiac hypertrophy: short communication. Circ. Res., 2012, 111(8): 1048-1053. DOI: https://doi.org/10.1161/CIRCRESAHA.112.273367
  78. Liggett S.B.; Cresci S.; Kelly R.J.; et al. A GRK5 polymorphism that inhibits beta-adrenergic receptor signaling is protective in heart failure. Nat. Med., 2008, 14(5): 510-517. DOI: https://doi.org/10.1038/nm1750
  79. Nakaya M.C.; Chikura S.; Watari K.J.; et al. Induction of cardiac fibrosis by β-blocker in G protein-independent and G protein-coupled receptor kinase 5/β-arrestin2-dependent Signaling pathways. J.Bio. Chem., 2012, 287(42): 35669-35677. DOI: https://doi.org/10.1074/jbc.M112.357871
  80. Teoh J.P.; BayoumiA.S.; Aonuma T.; et al. β-arrestin-biased agonism of β-adrenergic receptor regulates dicer-mediated microRNA maturation to promote cardioprotective signaling. J. Mol. Cell. Cardiol.,2018, 118: 225-236. DOI: https://doi.org/10.1016/j.yjmcc.2018.04.001
  81. Colomer J.M.; Mao L.; Rockman H.A.; et al. Pressure overload selectively up-regulates Ca2+/calmodulin-dependent protein kinase II in vivo. Mol. Endocrinol., 2003, 17(2): 183-192. DOI: https://doi.org/10.1210/me.2002-0350
  82. Grimm M.; Ling H.Y.; Willeford A.; et al. CaMKIIδ mediates β-adrenergic effects on RyR2 phosphorylation and SR Ca(2+) leak and the pathophysiological response to chronic β-adrenergic stimulation. J. Mol. Cell. Cardiol., 2015, 85: 282-291. DOI: https://doi.org/10.1016/j.yjmcc.2015.06.007
  83. Xu L.; Lai D.W.; Cheng J.; et al. Alterations of L-type Calcium current and cardiac function in CaMKII{delta} knockout mice. Circ. Res., 2010, 107(3): 398-407. DOI: https://doi.org/10.1161/CIRCRESAHA.110.222562
  84. Wang W.; Zhu W.Z.; Wang S.Q.; et al. Sustained beta1-adrenergic stimulation modulates cardiac contractility by Ca2+/calmodulin kinase signaling pathway. Circ. Res., 2004, 95(8): 798-806. DOI: https://doi.org/10.1161/01.RES.0000145361.50017.aa
  85. Mika D.; Richter W.; Conti M. A CaMKII/PDE4D negative feedback regulates cAMP signaling. Proc. Natl. Acad. Sci. U. S. A., 2015, 112(7): 2023-2028. DOI: https://doi.org/10.1073/pnas.1419992112
  86. Zhu W.Z.; Wang S.Q.; Chakir K.; et al. Linkage of beta1-adrenergic stimulation to apoptotic heart cell death through protein kinase A-independent activation of Ca2+/calmodulin kinase II. J. Clin. Inves., 2003, 111(5): 617-625. DOI: https://doi.org/10.1172/JCI200316326
  87. Yang Y.B.; Zhu W.Z.; Joiner M.L.; et al. Calmodulin kinase II inhibition protects against myocardial cell apoptosis in vivo. Am. J. Physiol., 2006, 291(6): H3065-H3075. DOI: https://doi.org/10.1152/ajpheart.00353.2006
  88. Sucharov C.C.; Mariner P.D.; Nunley K.R.; et al. A beta1-adrenergic receptor CaM kinase II-dependent pathway mediates cardiac myocyte fetal gene induction. Am. J. Physiol., 2006, 291(3): H1299-H1308. DOI: https://doi.org/10.1152/ajpheart.00017.2006
  89. Dybkova N.; Sedej S.; Napolitano C.; et al. Overexpression of CaMKIIδc in RyR2R4496C+/- knock-in mice leads to altered intracellular Ca2+ handling and increased mortality. J. Am. Coll. Cardiol., 2011, 57(4): 469-479. DOI: https://doi.org/10.1016/j.jacc.2010.08.639
  90. Sadredini M.; Haugsten Hansen M.; Frisk M.; et al. CaMKII inhibition has dual effects on spontaneous Ca2+ release and Ca2+ alternans in ventricular cardiomyocytes from mice with a gain-of-function RyR2 mutation. Am. J. Physiol, 2021, 321(2): H446-H460. DOI: https://doi.org/10.1152/ajpheart.00011.2021
  91. Curran J.; Hinton M.J.; Ríos E.; et al. Beta-adrenergic enhancement of sarcoplasmic reticulum Calcium leak in cardiac myocytes is mediated by Calcium/calmodulin-dependent protein kinase. Circ. Res., 2007, 100(3): 391-398. DOI: https://doi.org/10.1161/01.RES.0000258172.74570.e6
  92. Bos J.L. Epac proteins: multi-purpose cAMP targets. Trends Biochem. Sci., 2006, 31(12): 680-686. DOI: https://doi.org/10.1016/j.tibs.2006.10.002
  93. Ponsioen B.; Gloerich M.; Ritsma L.; et al. Direct spatial control of Epac1 by cyclic AMP. Mol. Cell. Biol., 2009, 29(10): 2521-2531. DOI: https://doi.org/10.1128/MCB.01630-08
  94. Mangmool S.; Shukla A.K.; Rockman H.A. beta-Arrestin-dependent activation of Ca(2+)/calmodulin kinase II after beta(1)-adrenergic receptor stimulation. J. Cell Biol., 2010, 189(3): 573-587. DOI: https://doi.org/10.1083/jcb.200911047
  95. Curran J.; Tang L.F..; Roof S.R.;et al. Nitric oxide-dependent activation of CaMKII increases diastolic sarcoplasmic reticulum Calcium release in cardiac myocytes in response to adrenergic stimulation. PloS one, 2014, 9(2): e87495. DOI: https://doi.org/10.1371/journal.pone.0087495
  96. Pereira L.; Bare D.J.; Galice S.; et al. β-Adrenergic induced SR Ca2+ leak is mediated by an Epac-NOS pathway. J. Mol. Cell. Cardiol., 2017, 108: 8-16. DOI: https://doi.org/10.1016/j.yjmcc.2017.04.005
  97. Zhang X.Y.; Szeto C.; Gao E.; et al. Cardiotoxic and cardioprotective features of chronic β-adrenergic signaling. Circ. Res., 2013, 112(3): 498-509. DOI: https://doi.org/10.1161/CIRCRESAHA.112.273896
  98. Beauverger P.; Ozoux M.L.; Bégis G.; et al. Reversion of cardiac dysfunction by a novel orally available Calcium/calmodulin-dependent protein kinase II inhibitor, RA306, in a genetic model of dilated cardiomyopathy. Cardiovasc. Res., 2020, 116(2): 329-338. DOI: https://doi.org/10.1093/cvr/cvz097
  99. Mustroph J.; Wagemann O.; Lücht C.M.; et al. Empagliflozin reduces Ca/calmodulin-dependent kinase II activity in isolated ventricular cardiomyocytes. ESC Heart Failure, 2018, 5(4): 642-648. DOI: https://doi.org/10.1002/ehf2.12336
  100. Chen J.J.; Xu S.N.; Zhou W.; et al. Exendin-4 reduces ventricular arrhythmia activity and Calcium Sparks-Mediated sarcoplasmic reticulum Ca leak in rats with heart failure. Int. Heart J., 2020, 61(1): 145-152. DOI: https://doi.org/10.1536/ihj.19-327
  101. Puhl S.L.; Weeks K.L.; Güran A.; et al. Role of type 2A phosphatase regulatory subunit B56α in regulating cardiac responses to β-adrenergic stimulation in vivo. Cardiovasc. Res., 2019, 115(3): 519-529. DOI: https://doi.org/10.1093/cvr/cvy230
  102. Schmid E.; Neef S.; Berlin C.; et al. Cardiac RKIP induces a beneficial β-adrenoceptor-dependent positive inotropy. Nat. Med., 2015, 21(11): 1298-1306. DOI: https://doi.org/10.1038/nm.3972
  103. Zhang W.J.; Qu X.X.; Chen B.Y.; et al. Critical roles of STAT3 in β-Adrenergic functions in the heart. Circulation, 2016, 133(1): 48-61. DOI: https://doi.org/10.1161/CIRCULATIONAHA.115.017472
  104. Zhu W.Z.; Tsang S.; Browe D.M.; et al. Interaction of β1-adrenoceptor with RAGE mediates cardiomyopathy via CaMKII signaling. JCI Insight, 2016, 1(1): e84969. DOI: https://doi.org/10.1172/jci.insight.84969
  105. Whitcomb V.; Wauson E.; Christian D.; et al. Regulation of beta adrenoceptor-mediated myocardial contraction and Calcium dynamics by the G protein-coupled estrogen receptor 1. Biochem. Pharm., 2020, 171: 113727. DOI: https://doi.org/10.1016/j.bcp.2019.113727
  106. Castaldi A.; Zaglia T.; Di Mauro V.; et al. MicroRNA-133 modulates the β1-adrenergic receptor transduction cascade. Circ. Res., 2014, 115(2): 273-283. DOI: https://doi.org/10.1161/CIRCRESAHA.115.303252
  107. Liu Z.B.; Tao B.; Fan S.Z.; et al. Over-expression of microRNA-145 drives alterations in β-adrenergic signaling and attenuates cardiac remodeling in heart failure post myocardial infarction. Aging, 2020, 12(12): 11603-11622. DOI: https://doi.org/10.18632/aging.103320
  108. Deng Y.F.; Wang J.; Xie G.J.; et al. Circ-HIPK3 strengthens the effects of adrenaline in heart failure by MiR-17-3p - ADCY6 axis. Int. J. Biol. Sci., 2019, 15(11): 2484-2496. DOI: https://doi.org/10.7150/ijbs.36149
  109. Johnson J.A.; Liggett S.B. Cardiovascular pharmacogenomics of adrenergic receptor signaling: clinical implications and future directions. Clin. Pharmacol. Thera., 2011, 89(3): 366-378. DOI: https://doi.org/10.1038/clpt.2010.315
  110. DiNicolantonio J.J.; Fares H.; Niazi A.K.; et al. β-Blockers in hypertension, diabetes, heart failure and acute myocardial infarction: a review of the literature. Open Heart, 2015, 2(1): e000230. DOI: https://doi.org/10.1136/openhrt-2014-000230
  111. Poirier L.; Tobe S.W. Contemporary use of β-blockers: clinical relevance of subclassification. Canadi. J. Cardio., 2014, 30(5 Suppl): S9-S15. DOI: https://doi.org/10.1016/j.cjca.2013.12.001
  112. Moens A.L.; Leyton-Mange J.S.; Niu X.L.; et al. Adverse ventricular remodeling and exacerbated NOS uncoupling from pressure-overload in mice lacking the beta3-adrenoreceptor. J. Mol. Cell. Cardiol., 2009, 47(5): 576-585. DOI: https://doi.org/10.1016/j.yjmcc.2009.06.005
  113. Liu H.H.; Zhang C.H.; Ammanamanchi N.; et al. Control of cytokinesis by β-adrenergic receptors indicates an approach for regulating cardiomyocyte endowment. Sci. Transl. Med., 2019, 11(513): eaaw6419. DOI: https://doi.org/10.1126/scitranslmed.aaw6419
  114. Frishman W.H. Fifty years of beta-adrenergic blockade: a golden era in clinical medicine and molecular pharmacology. Am. J. Med., 2008, 121(11): 933-934. DOI: https://doi.org/10.1016/j.amjmed.2008.06.025
  115. do Vale G.T.; Ceron C.S.; Gonzaga N.A.; et al. Three generations of β-blockers: history, class differences and clinical applicability. Curr. Hypertens. Rev., 2019, 15(1): 22-31. DOI: https://doi.org/10.2174/1573402114666180918102735
  116. García-Prieto J.; Villena-Gutiérrez R.; Gómez M.; et al. Neutrophil stunning by metoprolol reduces infarct size. Nat. Commun., 2017, 8: 14780. DOI: https://doi.org/10.1038/ncomms14780
  117. Cannavo A.; Rengo G.; Liccardo D.; et al. β1-Blockade prevents Post-Ischemic myocardial decompensation via β3AR-Dependent protective sphingosine-1 phosphate signaling. J. Am. Coll. Cardiol., 2017, 70(2): 182-192. DOI: https://doi.org/10.1016/j.jacc.2017.05.020
  118. Dewenter M.; Neef S.; Vettel C.; et al. Calcium/calmodulin-dependent protein kinase II activity persists during chronic β-adrenoceptor blockade in experimental and human heart failure. Circ.: Heart Failure, 2017, 10(5): e003840. DOI: https://doi.org/10.1161/CIRCHEARTFAILURE.117.003840
  119. Broeders M.A.; Doevendans P.A.; Bekkers B.C.; et al. Nebivolol: a third-generation beta-blocker that augments vascular nitric oxide release: endothelial beta(2)-adrenergic receptor-mediated nitric oxide production. Circulation, 2000, 102(6): 677-684. DOI: https://doi.org/10.1161/01.CIR.102.6.677
  120. Fonseca V.A. Effects of beta-blockers on glucose and lipid metabolism. Curr. Med. Res. Opin., 2010, 26(3): 615-629. DOI: https://doi.org/10.1185/03007990903533681
  121. Park M.; Steinberg S.F. Carvedilol prevents redox inactivation of cardiomyocyte β1-adrenergic receptors. JACC. Basic Transl. Sci., 2018, 3(4): 521-532. DOI: https://doi.org/10.1016/j.jacbts.2018.06.002
  122. Wang J.L.; Hanada K.J.; Staus D.P.; et al. Gαi is required for carvedilol-induced β1 adrenergic receptor β-arrestin biased signaling. Nat. Commun., 2017, 8(1): 1706. DOI: https://doi.org/10.1038/s41467-017-01855-z
  123. Güven B.; Kara Z.; Onay-Beşikci A. Metabolic effects of carvedilol through β-arrestin proteins: investigations in a streptozotocin-induced diabetes rat model and in C2C12 myoblasts. Br. J. Pharmacol., 2020, 177(24): 5580-5594. DOI: https://doi.org/10.1111/bph.15269
  124. Nash C.A.; Wei W.H.; Irannejad R.; et al. Golgi localized β1-adrenergic receptors stimulate Golgi PI4P hydrolysis by PLCε to regulate cardiac hypertrophy. Elife, 2019, 8: e48167. DOI: https://doi.org/10.7554/eLife.48167
  125. Koracevic G.; Micic S.; Stojanovic M.; et al. Compelling indications should be listed for individual beta-blockers(due to diversity),not for the whole class. Curr. Vasc. Pharmacol., 2021, 19(4): 343-346. DOI: https://doi.org/10.2174/1570161118666200518113833
  126. Germano N.; Summerfield D.; Johnson B. A mini review of inhaled beta 2 agonists in acute decompensated heart failure requiring respiratory support. Pulm. Crit. Care. Med., 2019, 4(3): 10.15761/pccm.1000161. DOI: https://doi.org/10.15761/PCCM.1000161
  127. Reddy Y.N.V.; Obokata M.; Koepp K.E.; et al. The β-Adrenergic agonist albuterol improves pulmonary vascular reserve in heart failure with preserved ejection fraction. Circ. Res., 2019, 124(2): 306-314. DOI: https://doi.org/10.1161/CIRCRESAHA.118.313832
  128. Ge X.A.; Woo A.Y.H.; Xing G.; et al. Synthesis and biological evaluation of β2-adrenoceptor agonists bearing the 2-amino-2-phenylethanol scaffold. Eur. J. Med. Chem., 2018, 152: 424-435. DOI: https://doi.org/10.1016/j.ejmech.2018.04.041
  129. Ge X.Y.; Mo Y.M.; Xing G.; et al. Synthesis, biological evaluation and molecular modeling of 2-amino-2-phenylethanol derivatives as novel β2-adrenoceptor agonists. Bioorg. Chem., 2018, 79: 155-162. DOI: https://doi.org/10.1016/j.bioorg.2018.04.017
  130. Ahn S.; Kahsai A.W.; Pani B.; et al. Allosteric “beta-blocker” isolated from a DNA-encoded small molecule library. Proc. Natl. Acad. Sci. U. S. A., 2017, 114(7): 1708-1713. DOI: https://doi.org/10.1073/pnas.1620645114
  131. Liu X.Y.; Ahn S.; Kahsai A.W.; et al. Mechanism of intracellular allosteric β2AR antagonist revealed by X-ray crystal structure. Nature, 2017, 548(7668): 480-484. DOI: https://doi.org/10.1038/nature23652
  132. Meng K.C.; Shim P.; Wang Q.; et al. Design, synthesis, and functional assessment of Cmpd-15 derivatives as negative allosteric modulators for the β2-adrenergic receptor. Bioorg. Med. Chem., 2018, 26(9): 2320-2330. DOI: https://doi.org/10.1016/j.bmc.2018.03.023
  133. Ahn S.; Pani B.; Kahsai A.W.; et al. Small-Molecule positive allosteric modulators of the β2-Adrenoceptor isolated from DNA-Encoded libraries. Mol. Pharmacol., 2018, 94(2): 850-861. DOI: https://doi.org/10.1124/mol.118.111948
  134. Liu X.Y.; Masoudi A.; Kahsai A.W.; et al. Mechanism of β2AR regulation by an intracellular positive allosteric modulator. Science, 2019, 364(6447): 1283-1287. DOI: https://doi.org/10.1126/science.aaw8981
  135. Liu X.Y.; Kaindl J.; Korczynska M.; et al. An allosteric modulator binds to a conformational hub in the β2 adrenergic receptor. Nat. Chem. Biol., 2020, 16(7): 749-755. DOI: https://doi.org/10.1038/s41589-020-0549-2
  136. Woo A.Y.H.; Ge X.Y.; PanL.; et al. Discovery of β-arrestin-biased β2-adrenoceptor agonists from 2-amino-2-phenylethanol derivatives. Acta Pharmacol. Sin., 2019, 40(8): 1095-1105. DOI: https://doi.org/10.1038/s41401-018-0200-x
  137. Carr R. 3rd.; Schilling J.; Song J.L.;et al. β-arrestin-biased signaling through the β2-adrenergic receptor promotes cardiomyocyte contraction. Proc. Natl. Acad. Sci. U. S. A., 2016, 113(28): E4107-E4116. DOI: https://doi.org/10.1073/pnas.1606267113
  138. Li C.X.; Cheng W.K.; Guo J.; et al. Relationship of inhaled long-acting bronchodilators with cardiovascular outcomes among patients with stable COPD: a meta-analysis and systematic review of 43 randomized trials. Int. J. Chronic Obstruct. Pulm. Dis., 2019, 14: 799-808. DOI: https://doi.org/10.2147/COPD.S198288
  139. Qiao Y.H.; Zhu B.L.; Tian A.J.; et al. PEG-coated Gold nanoparticles attenuate β-adrenergic receptor-mediated cardiac hypertrophy. Int. J. Nanomed., 2017, 12: 4709-4719. DOI: https://doi.org/10.2147/IJN.S130951
  140. Duran-Corbera A.; Faria M.; Ma Y.Y.; et al. A photoswitchable ligand targeting the β1-Adrenoceptor enables Light-Control of the cardiac rhythm**. Angew. Chem., Int. Ed., 2022, 61(30): e202203449. DOI: https://doi.org/10.1002/anie.202203449