Abstract
G-protein coupled receptor (GPCR) kinases (GRKs) and β-arrestins play key roles in GPCR and non-GPCR cellular responses. In fact, GRKs and arrestins are involved in a plethora of pathways vital for physiological maintenance of inter- and intracellular communication. Here we review decades of research literature spanning from the discovery, identification of key structural elements, and findings supporting the diverse roles of these proteins in GPCR-mediated pathways. We then describe how GRK2 and β-arrestins partake in non-GPCR signaling and briefly summarize their involvement in various pathologies. We conclude by presenting gaps in knowledge and our prospective on the promising pharmacological potential in targeting these proteins and/or downstream signaling. Future research is warranted and paramount for untangling these novel and promising roles for GRK2 and arrestins in metabolism and disease progression.
Keywords: GPCR regulation, arrestins, GRK, diseases, metabolism, cancer, pain, heart failure
1. GPCR signaling and receptor barcoding
G protein-coupled receptors (GPCRs) are seven transmembrane domain proteins that upon agonist binding activate several downstream pathways involved in various cellular processes. GPCRs compose one of the largest protein superfamilies which importantly account for the targets of over 35% of FDA-approved pharmacological agents [1]. Importantly, a vast number of GPCRs are orphaned receptors, that is, an agonist and/or function remains unknown. Thus, there is a dire ongoing need to better understand GPCR-mediated mechanisms in the context of both disease and pharmacological intervention and to unravel novel GPCRs and GPCR-regulating proteins. The extensive involvement of this receptor family and its regulators in various pathologies (section 4) highlights the impact of GPCR-pathways in regulating cellular function and in many cases driving pathological signaling.
GPCRs can be classified into seven main subfamilies. Class A (Rhodopsin-like receptor family) is the largest GPCR subfamily which includes adrenergic receptors. Class B (Secretin receptor family) comprises peptide hormones like glucagon-like peptide-1 receptor (GLP1R) [2]. Class C (Metabotropic glutamate receptor family) consists of glutamate and GABAB receptors that form special allosteric dimers [3]. Class D includes fungal mating pheromone receptors. Class E notably includes cAMP receptors; Class F (Frizzled (FZD) / Taste receptor family) have been described for their role in regulating transcription, cytoskeleton, and calcium [4–6]; and lastly the adhesion receptor family [7,8].
GPCR signaling patterns are principally determined by the binding of ligands in the extracellular domain and the interaction with cytosolic signaling partners on the intracellular interface (Fig. 1). Each GPCR has a unique binding profile for one or more ligands, each of which with specific receptor interaction dynamics. This review will focus on the important regulatory events in the intracellular side of GPCR activation which controls its localization and capacity to induce second messenger signaling cascades. The unique signaling pattern of a GPCR is informed by the specific heterotrimeric G-protein complex to which it is coupled [9]. The intracellular guanine nucleotide heterotrimeric G protein complex that couples the GPCR includes Gα, Gβ, and Gγ subunits [10] (Fig. 1). The Gα subunit can be further characterized into four sub-families, Gαs, Gαi/o, Gαq/11, and Gα12/13, based on structure similarity and activity [11]. The human proteome contains 23 Gα, 5 Gβ, and 12 Gγ [12], which theoretically could form more than 1300 heterotrimeric complexes giving GPCRs a vast ability to fine-tune signaling cascades based on specific heterotrimeric G-protein coupling. Upon stimulation, the active-inactive cycle of GPCR begins with the dissociation of guanosine diphosphate (GDP, inactive form) from the Gα subunit which allows for guanosine triphosphate (GTP, active form) to bind [13–15]. The Gα subunit then dissociates from the Gβγ dimer and activates second messengers including the adenylyl cyclase (AC)/cyclic AMP (cAMP) catalyzation cascade, inositol triphosphate (IP3), and/or ion channels. G-protein signaling is terminated by GTP hydrolysis catalyzed by regulator of G protein signaling (RGS) proteins. This allows G-proteins to return to their unstimulated GPCR-bound state [16] (Fig. 1). Moreover, coupling of specific GPCRs to G-protein complexes can be cell type-specific. For instance, D1-like dopamine receptors canonically couple to Gαs while D2-like dopamine receptors canonically couple to Gαi; however, in certain brain regions and in macrophages it has been documented that both classes can primarily couple to Gαq [17,18]. This allows for an incredible number of GPCR-G protein coding and integration permitting a very finely-tune signaling pattern that may differ in specific cell types as well as allowing for different cellular responses to the same stimuli.
Fig. 1.

Schematic representation of GPCR-G-protein signaling. Created with BioRender.
Notably, these cell type and cell state deviations in signaling have also been proposed to be associated with post-translational modifications to the GPCR’s intracellular surface. An important and increasingly well-characterized domain of post-translational modification to GPCRs is referred to as “GPCR barcoding”. GPCR barcoding refers to the multitude of regulatory phosphorylation and dephosphorylation events mediated by kinases and phosphatases on the intracellular side of an activated GPCR. These phosphorylation patterns inform the energetic favorability for binding of intracellular proteins. For instance, a rapid turn-off mechanism can be accomplished by GPCR-kinase (GRK) receptor phosphorylation which promotes binding of β-arrestins [19,20]. The recruitment of β-arrestins allows for clathrin-mediated endocytic internalization of the receptor for recycling and/or degradation (Fig. 2). This review will focus on the role of GRKs and β-arrestins in cellular function and diseased conditions. Noteworthy, other kinases are well known to participate in receptor barcoding such as Protein Kinase A (PKA) [21] and Protein Kinase C (PKC) [22]. For instance, increased PKA-mediated phosphorylation of the β2-adrenergic receptor (β2AR) can direct its signaling from Gαs to Gαi [23]. These regulatory phosphorylation events can be classified as second messenger dependent (PKA and PKC) or as ligand binding dependent (GRK family). Finally, it is important to note that, while these two classes of kinases both regulate GPCR signaling through barcoding, they can also directly interact with each other, further integrating second messenger-dependent and ligand-dependent GPCR barcoding. For instance, PKA is known to phosphorylate GRK2 and thereby enhance its desensitization of the β2-adrenergic receptors and potentially other receptors [24]. However, their contributions to the GPCR barcode remain distinguishable as PKA phosphorylation sites do not appear to influence β-arrestin recruitment, whereas GRKs are the primary driving force towards this GPCR fate [25].
Fig. 2.

GRK2 and β-arrestin signaling in receptor internalization. Created with BioRender.
1.1. GRK Family
Phosphorylation of activated GPCRs by GRKs was first described almost five decades ago where rhodopsin, a classical GPCR, was found to be phosphorylated by rhodopsin kinase, now known as GRK1 [26]. The β2AR was subsequently found to be phosphorylated and desensitized by β-adrenergic receptor kinase 1 (also known as GRK2), and found to be altered in various pathologies [27]. Several additional GRK isoforms have since been identified and they are now divided into three main groups: (i) the rhodopsin kinase subfamily which includes GRK1 and GRK7; (ii) the β-adrenergic receptor kinase subfamily, which includes GRK2, and GRK3; and (iii) the GRK4 subfamily includes GRK4, GRK5, and GRK6 [28]. GRK1 is expressed in rods (low/dim light) and cones (bright light with color distinction) and can inactivate vertebrate photoreceptors and regulate the phototransduction cascade. GRK7 is expressed primarily in cones and regulates the same processes [29]. GRK3, also known as β-adrenergic receptor kinase 2, is the least abundant cytosolic GRK, but it is widely expressed in the brain [30]. GRK4 plays an important role in the phosphorylation and regulation of dopamine D1 receptors, and when overexpressed induces hypertension [31]; although, GRK4 is mainly expressed in the testis, cerebellum, and kidneys [32]. GRK4 and GRK6 are post-translationally modified to localize to the membrane [33]. GRK2 and GRK5 are unevenly expressed in different tissues throughout the body and are well characterized for their roles in heart function [34]. Notably, GRK5 is associated with the membrane through PIP2 and phospholipid binding [35]. Due to its broad expression pattern, potential for pharmacological targeting, and involvement in a spectrum of diseases, this review will focus on GRK2.
1.2. Molecular determinants of GRK2 signaling
GRK family members have similar structures including a conserved central catalytic domain, an N-terminal domain that recognizes and localizes the kinases to lipid membranes, and the variable C-terminal domain which regulates their translocation [36,37]. However, there are three distinct regions in GRK2 that play key roles in determining its function: an amino terminal RGS homology (RH) domain, a pleckstrin homology (PH) domain, and a unique C-terminal domain. The RH domain of GRK2 [38] has been shown to specifically interact with members of the Gαq and Gα11 family and is able to inhibit Gαq-mediated phospholipase C activity [39]. The PH domain participates in its translocation to the plasma membrane as it is implicated in binding phospholipids and Gβγ subunits. Table 1 shows a subset of known GPCRs with their respective primary functions that are negatively regulated by GRK2-β-arrestin interactions. Thus, upregulation of GRK2 or β-arrestin activity by natural or pharmacological agents would detrimentally impact receptor activity while the reverse would enhance its receptor primary responses.
Table 1.
A subset of GPCRs known to be negatively regulated by GRK2 phosphorylation.
| GPCR | Primary Functions | Reference (s) |
|---|---|---|
| β2-Adrenergic Receptor | Increases cardiac and skeletal muscle contractility in response to sympathetic stimuli. | [40] |
| D1-like Dopamine Receptors | Regulate central nervous system reward circuitry and contribute to immune responses [41]. | [42] |
| D2-like Dopamine Receptors | [43] | |
| μ-Opioid Receptor | Key target of natural and pharmacological pain relief. | [44] |
| Muscarinic Receptors | Broadly expressed receptor family which response to acetylcholine especially from the parasympathetic nervous system. | [45] |
| CCR9 | Regulates immune cell migration particularly for T cells to the GI tract. | [46,47] |
| CXCR1 | Induce leukocyte recruitment and activation at sites of inflammation. | [46] |
| CXCR2 | ||
| CXCR4 | Chemokine receptors involved in immune development, hematopoiesis, and vascularization. HIV co-receptor. | [48] |
| CCR5 | Chemokine receptor involved in the migration and activation of leukocytes. HIV co-receptor | [49] |
| Glucagon like peptide 1 receptor | Incretin hormone involved in regulating insulin secretion, satiety, and cardiac rhythm. | [50,51] |
GRK2 induces the desensitization of GPCRs through multiple phosphorylation events at unique sites specific to each GPCR. Distinguishing which phosphorylation sites are specific to GRK2 relative to other GRKs or second messenger-dependent protein kinases requires using multiple site-specific mutations, functional analysis, and well validated phospho-specific antibodies. Such studies have been performed for some key GRK2-regulated receptors including the β2AR, where GRK2 phosphorylates threonine 360 and serine residues 364, 396, 401, 407, and 411 within the C-terminal tail [52]. Conversely, in dopamine receptors, since specific sites are not well established, the GPCR-GRK2 interaction is studied by mutating all intracellular serine and threonine residues to alanine [53,54]. Some studies have attempted to predict GRK2 phosphorylation sites by computational methods using available crystal structures of GRK2 and its interacting partners, considering residue accessibility and structural flexibility [55]. While much has been detailed regarding certain subtypes of GPCRs and its GRK2 interactions, many of GRK2’s preference for specific GPCRs remain to be further detailed.
1.3. The complexity of GRK2 signaling beyond GPCR phosphorylation
In addition to regulating processes via phosphorylation, GRK2 can regulate cellular responses in a non-phosphorylation-dependent manner, such as through structural interactions with proteins involved in signal transduction and transport. This includes facilitating the recruitment of phosphoinositide 3-kinase (PI3K) to the membrane and aiding in receptor desensitization [56]. The binding of serine-threonine kinase Akt through the C-terminus of GRK2 can result in inhibiting phosphorylation of Akt [57]. The role of upregulated GRK2 on decreasing extracellular signal-regulated kinase (ERK) activity is independent of receptor phosphorylation, in a proposed mechanism involving direct or coordinate downstream interaction with the chemokine-driven mitogen-activated protein kinase (MEK) [58]. The interaction between Raf and MEK could be dissociated by the Raf kinase inhibitor protein (RKIP), which was also shown to activate GRK2, and further inhibits the kinase signaling through the Ras-Raf-MEK-ERK pathway [59]. Clathrin binds to GRK2 through a clathrin box located in the C-terminal domain of GRK2 and induces the internalization of certain GPCRs through β-arrestin recruitment [58,60]. Protein-coupled receptor kinase-interacting protein (GIT1), a GTPase-activating protein (GAP), was shown to interact with GRK2 [61]. The inhibitory binding of calmodulin to GRK2 could be almost completely abolished by PKC phosphorylation at serine 29, and resulted in decreased kinase activity of GRK2 [62]. GRK2 also associates with caveolin, α-actinin, proto-oncogene tyrosine-protein kinase Src (c-SRC), IkBα, and heat shock protein 90 (Hsp90), a folding regulator of GRK2 [36,63–65]. Notably, GRK2 phosphorylates non–GPCR proteins, such as the insulin receptor substrate 1 (IRS1), an important regulator of insulin sensitivity and glucose uptake [66]. Inhibition of GRK2 enhanced endothelial nitric oxide synthase (eNOS) signaling, which reduced myocyte loss and improved ischemic injury in mice [67]. Thus, GRK2 is an important regulator of cell death, cell-cell interaction, cytoskeletal networks through tubulin and actin [68], immunity and other intracellular functions via mechanisms that also involve its non-GPCR non-kinase function.
Just as receptor phosphorylation by GRKs can trigger internalization and desensitization, residue specific dephosphorylation can oppose this process and promote re-sensitization. Receptors internalized by endocytosis must undergo dephosphorylation to be meaningfully recycled to the plasma membrane (Fig. 2) as a GPCR barcoded for internalization would be short-lived at the plasma membrane [69]. Dephosphorylation may also occur at the plasma membrane prior to endocytic engulfment, thus preventing internalization [70]. As with the GRK family of kinases, phosphatases that regulate this process do not solely interact with GPCRs [71] and they possess functional redundancy (e.g. multiple phosphatases are able to regulate one receptor). Delineating what causes a receptor to be recycled or degraded is of great importance. For instance, a therapeutic strategy that favored central nervous system opioid receptor recycling over degradation has been sought after for some time [72]. In addition to phosphatase activity, ubiquitination and de-ubiquitination can contribute to this process as an ubiquitin-tagged protein will be targeted for degradation [73]. As such, the details of how GPCR post-internalization sorting occurs merits its own review [74].
2. Arrestins in Cellular Signaling
β-Arrestins are a family of regulatory proteins that recognize and bind to activated GPCRs with high affinity to promote receptor desensitization, turning off continued signaling [75,76]. The visual arrestin-1 (also known as S-Arrestin) was first described in 1986 as a 48-kDa protein that phosphorylates and suppresses rhodopsin’s cGMP phosphodiesterase activating capacity [76]. A second arrestin was then found to bind competitively with phosphorylated rhodopsin and phosphorylated β2AR. For this non-visual interaction, it garnered the name β-arrestin; it was later termed β-arrestin1 following the discovery of the second non-visual (third in total) arrestin. The third arrestin homolog (β-arrestin 2) was isolated from bovine brain cDNAs [77,78]. β-arrestin 1 and 2 have more than 70% conserved amino acid identity and considerable functional overlap, yet with some functional distinctions [79,80]. Despite being named for their initial association with the β2AR, it is now known that β-arrestins interact with more than 100 proteins and many other GPCRs [81]. A fourth arrestin, a visual arrestin with restricted expression in photoreceptors cells, was later reported where cDNA was isolated from cone photoreceptor of ascidian Ciona intestinalis [82]. Focus of this review will be on non-visual β-arrestin 1 and β-arrestin 2 as it relates to GRK signaling and diseases.
2.1. Molecular determinants of β-arrestin signaling
Crystal structures for β-arrestin 1 [83] and for β-arrestin 1 and 2 in complex with a number of GPCRs [84] or GPCR fragments [85] have revealed key structural determinants of these proteins. β-arrestins have two separate noteworthy binding domains, where one domain recognizes phosphorylation sites on the C-terminal tail of GPCRs and a second domain recognizes whether the receptor is in an active conformation. These two domains have been proposed to induce an active state of β-arrestin [86], where both separate binding events are thought to contribute to β-arrestin activity and GPCR desensitization. Upon binding an activated and phosphorylated GPCR, β-arrestin undergoes a conformational change that rearranges both its C- and N-termini exposing a third critical domain, the clathrin-binding domain. The exposure of this domain in its active form initiates the formation of the clathrin-coated endocytic vesicle [87], which is key to receptor desensitization processes.
2.2. β-arrestin in biased and unbiased signaling
In the standard model of GPCR signaling, agonist binding induces signaling through the heterotrimeric G-protein complex equivalently, and thus one agonist binding event will in equal measure induce second messenger signaling and receptor desensitization. However, many GPCRs have multiple natural and synthetic agonists, some of which can favor the induction of signaling or desensitization of the receptor relative to the primary cognate agonist, through a process known as biased agonism. The model for biased agonism is based on the principle that GPCRs exhibit ligand-dependent conformational changes resulting in ligand-dependent signal transduction [88]. For instance, there are a number of biased agonists for the β2AR either towards greater G-protein signaling or greater desensitization each with distinct clinical utility which was recently reviewed by Ippolito and colleagues [89]. For instance, inhaled β2AR agonists biased towards G-protein signaling have greater utility in asthma than desensitization-biased agonists. GRKs and arrestins binding to an activated GPCR is a key determinant of whether an agonist will exhibit this biased effect. As mentioned above, bias is driven by ligand-dependent conformational changes, and indeed some ligands induce conformational changes that energetically favor GRK2 binding. This is well demonstrated by FRET and BRET studies showing real time differential GRK2 [90] and/or β-arrestin colocalization to GPCRs in response to biased and non-biased ligands [91]. From a functional standpoint, it has been demonstrated that by prohibiting phosphorylation by GRKs through site-specific mutagenesis, a non-biased agonist can exhibit the effects of a G-protein biased agonist [92]. Exploring signaling one way or another and determining key events leading to β-arrestin biased signaling may be pivotal for developing targeted pharmacological agents to treat specific pathologies.
3. GRK2 and β-arrestin signaling
β-arrestins can sterically limit the attachment of G-protein heterotrimers with the receptor and therefore prevent its reassembly with the receptor. GPCR phosphorylation by GRK2 favors the recruitment of β-arrestin and in turn, induce receptors desensitization, promoting rapid receptor internalization [93]. Other subtypes of GRKs involved in desensitization and β-arrestin recruitment have different efficiencies. For instance, GRK2 and GRK3 are more effective in promoting receptor endocytosis through the β-arrestin pathway than GRK5 and GRK6 [25]. New studies have proposed key events and cross-talk between GRK2 and arrestins of which we summarize in the following subsections.
3.1. Non-GPCR signaling of GRK2/β-arrestin
In addition to the inhibitory effect of GRK2/arrestins in GPCR signal transduction, emerging evidence suggests that GRK2 and β-arrestins regulate intracellular signaling independently of GPCRs by affecting non-GPCR receptors or by direct interaction with various cellular target molecules involved in signal transduction. β-arrestins can be used as scaffolds to bring GPCR-GRK2 complexes in close proximity with different signal molecules, including JNK-3, IGF-1R, and IGF-1, PDE4, cytoskeleton Ral-GDS modulator, Nuclear factor-κB (NF-κB) signal pathway, Mdm2, etc [94–96]. GRK2 and β-arrestins can be localized in other cellular compartments (i.e. mitochondria) where binding partners and cellular functional impact remain to be fully elucidated. Thus, GRK2 interaction with β-arrestins is essential for “turning-off” GPCRs but may also be key to overall cellular signal cascade that depend or not of GPCR responses.
3.2. GRK2/β-arrestin in metabolic regulation
Mitochondrial localization of GRK2 was first evidenced in rat cerebrovascular tissue and has been proposed to contribute to the progression of Alzheimer’s disease (AD) [97]. The kinase role of GRK2 in the mitochondria was then reported to regulate biological processes and ATP generation in primary cultured mouse aorta cells [98]. Recently, emerging evidence supported the metabolic regulatory role for the non-canonically mitochondrial localized GRK2 where it participates in cellular metabolism post cardiac ischemia reperfusion injury [99]. Supporting data suggests that indeed GRK2 translocates to the mitochondria via ERK phosphorylation of GRK2 and subsequent interaction of GRK2 with Hsp90 [100] where it inhibits glucose oxidation post-IR via decreased pyruvate dehydrogenase activity [99]. Although functionally mitochondrial GRK2 has been shown to negatively impact metabolic function and promote ROS-formation [101] clearly defined mitoGRK2-protein interactions within the mitochondria remains to be fully detailed. These interactions could be important to regulating mitochondrial substrate utilization particularly in diseased conditions such as myocardial infarction.
In addition to critically regulating metabolic processes, mitochondria can induce cell death responses to stress by releasing cytochrome c into the cytosol and promoting caspase activation [102]. Interestingly, GRK2 has been shown to induce cell death by increasing cleaved caspase 3 post-IR, whereas preventing its translocation leads to decreased CC3 and cell death post-IR [100,103]. Recent studies have also shown that apoptosis-induced caspases cleave β-arrestin 2 which then translocates to mitochondria to increase the release of cytochrome C into the cytosol, enhancing apoptotic signaling [104]. In addition, β-arrestin 1 mediates the activation of PI3K, subsequently activating Akt-induced apoptosis [105]. GRK2 and β-arrestins have been proposed to regulate mitochondrial function in cardiac cells through mechanisms involving mitochondrial superoxide production via NADPH oxidase 4 (Nox4) [106,107]. Noteworthy, β-arrestins have been shown to affect several proteins involved in mitochondrial respiration and metabolism like GAPDH in the glycolysis pathway and ATP synthases [81]. Exact mechanisms of action are not fully known albeit necessary to further understand the role of these proteins in metabolic regulation.
3.3. Beta-arrestin as a transcription factor
β-arrestins regulate gene transcription in the nucleus through cytosolic and nuclear interactions. β-arrestins bind to IkB and sequester the complex IkB-NF-κB in the cytosol, thus inhibiting NF-κB transcriptional activity, critical for inflammasome priming [108,109]. Although the ability of β-arrestin 1 to alter cellular transcription was well accepted, whether β-arrestin exerted this function via direct nuclear interactions was largely debated until an active nuclear localization signal was identified in β-arrestin 1 [110]. Since then, several studies have reported β-arrestin 1 as an epigenetic regulator in various cell types [111–113]. Importantly, β-arrestin 1 localization in the nucleus was dependent on GPCR activation [110] suggesting that its nuclear activity may be dependent on GRK initiation of GPCR-β-Arrestin interactions and scaffolding complexes. Further studies are needed to clarify these mechanisms.
3.4. Mutations and differential expression of GRKs and β-arrestin in animal disease models
Using targeted knockout and heterozygous mice, GRK2 has been shown to play an important role in atherosclerosis and other heart diseases [114–116], lymphocyte chemotaxis [117], and autoimmune diseases [118]. There are various genetic and chemically induced mouse models utilized to study GRK2/β-arrestin signaling. Studies have shown that cardiac overexpression of GRK2, as evidenced in human heart failure, is detrimental to myocardial function, particularly post-cardiac injury or pathological hypertrophy [119]. Global small molecule GRK2 inhibition is available through several compounds (including paroxetine, compound 101, balonal, etc), each posing a unique degree of selectivity and potency [120]. Alternatively, tissue-specific inhibition has been accomplished using expression of the small peptide termed βARKct [121].
GRK2 knockout strategies were performed by targeting exon 8, where homozygous GRK2 knockout mice were embryonically lethal [122]. Particularly, a lethality was linked to abnormal cardiac development, possibly due to markedly hypoplastic ventricular myocardium with abnormally lacking ventricular myocardium organization or differentiation during early embryogenesis, leading to heart failure. This mouse model supported the notion that GRK2 plays a critical role in development, particularly of the heart [122]. Interestingly, the global heterozygous GRK2 knockout mice were viable, indicating a minimal protein expression level that reached threshold for successful cellular differentiation and development. Cardiac-specific GRK2 knockout mice were generated using the Cre-lox technology, where mice were viable with normal heart structure and function in early development. Nonetheless, β-adrenergic stimulation revealed increased sensitivity of inotropic and lusitropic tachyphylaxis in the adult heart, suggesting the importance of cardiac GRK2 in protective effects in cardiomyopathy induced by catecholamine toxicity [115]. Further studies, using pancreatic-specific GRK2 knockout mice revealed a decrease in glucose-mediated insulin secretion linked to decreased calcium influx in pancreatic islets [123]. Whether this is developmentally linked remains to be further investigated but may be an important contributor to diabetes disease progression.
Murine knockout of β-arrestin 1 were generated by genetic disruption via homologous recombination at mouse chromosome 7. The study showed a similar survival rate of the β-arrestin 1 KO mice with wild-type mice, accompanied with similar fertile and offspring generation. Unchanged body function was reported in various tissues including brain, kidney, lung, and intestine. No abnormalities were linked to cardiovascular functions such as heart rate, blood pressure, and blood chemistry. Nevertheless, loss of β-arrestin 1 in the heart induced higher cardiac performance as measured by ejection fraction in response to isoproterenol, a βAR agonist, suggesting its role on β-adrenergic receptors [124]. Homologous recombination was used to inactivate the gene encoding for β-arrestin 2 [122]. Similarly to β-arrestin 1 knockout mice, loss of β-arrestin 2 in mice did not impact viability nor organ morphology. Notably, the analgesic effect of morphine was significantly enhanced in the knockout animals due to μ-opioid receptor signaling [125].
4. GRK2 and β arrestin in diseases
4.1. In cardiovascular and metabolic pathology
GRK2 and GRK5 are the prominent GRKs in the heart and elevated in heart disease [126]. GRK2 is paramount for both normal physiological cardiac function and disease progression. Upregulation of cardiac GRK2 has been reported in human HF [127,128] and diabetes [129]. In its canonical role, the accumulation of cardiac GRK2 internalizes βARs, leading to βAR insensitivity. This is particularly important to cardiac physiology as βARs are key to chronotropic and ionotropic regulation of the cardiovascular system in response to norepinephrine [130]. Increase in circulating catecholamines and upregulation in cardiac GRK2 are often observed in HF patients, both of which promote βAR down-regulation, cardiac hypertrophy, and myocyte apoptosis [128]. Notably, in response to chronic isoproterenol stimulation, β-arrestin-mediated EGFR transactivation, independent of G-protein activation, confered cardioprotection when compared to control animals [131]. Thus, drugs that selectively maintain cytoprotective mechanisms could be beneficial to HF patients.
Mechanistically, studies involving GRK2 in cardiovascular and metabolic pathologies have mainly stemmed from genetic mouse models. Although GRK2 deletion in mice is embryonically lethal, heterozygous GRK2 knockout mice have been used to investigate the role of GRK2 in HF progression in response to β-adrenergic stimulation [132]. Indeed, isoproterenol stimulated inotropic activity was more sensitive in cardiac-specific GRK2 knockout mice, where it restored inotropic and lusitropic responsiveness in acute isoproterenol-induced tachyphylaxis but led to accumulated catecholamine toxicity [115]. Studies have also found that cardiac-specific GRK2 ablation protected the heart from myocardial infarction (MI), reversed pathological left ventricular remodeling [116], and enhanced Ca2+ handling while diminishing tissue remodeling [133]. Circulating catecholamines are regulated by adrenal GRK2 which has been implicated in the pathophysiological features of HF, in particular regulating cardiac rate and contractility [134]. Studies have found that adrenal gland-specific inhibition of GRK2 decreased the level of plasma catecholamines and upregulated cardiac βAR signaling, improving cardiac function in HF [135]. Overall preventing GRK2 upregulation in the heart appears to be a viable strategy to treat HF. Notably, recent studies have shown that whole pancreas GRK2 knockdown leads to decreased glucose-mediated insulin secretion, increased weight gain when in a high fat regimen, and decreased cardiac function [123]. Kidney inhibition of GRK2 has also been reported to negatively impact kidney function and blood pressure [136]. GRK2 accumulation in the heart turns-off insulin signaling and inhibits glucose uptake [137]. Noteworthy are the studies showing that high-fat diet (HFD) regimen in cardiac-specific GRK2 inhibitor peptide mice elicited an obesogenic phenotype [138]. Key to untangling these confounding phenotypes are future studies detailing the link between GRK2 and metabolism. Overall, it is important that pre-clinical studies aimed at GRK2 inhibition for HF consider the impact in other organs and how dietary regimens may overall lead to physiological alterations.
Both β-arrestin 1 and 2 are abundantly expressed in cardiomyocytes and contribute to HF progression by enhancing PKA signaling of cardiac βAR [23]. β-arrestins were found to induce transactivation of Epidermal Growth Factor Receptor (EGFR) signaling via β1AR [131]. One of the regulators of blood homeostasis, angiotensin II (Ang II) receptor, plays an important role in enhancing the renin-angiotensin system (RAS) and is involved in the progression of cardiac hypertension and HF [139]. Studies have found that cardiac-specific overexpression of Ang II receptor type 1 (AT1) induces cardiomyocyte apoptosis by activation of Src and translocation of nuclear phospho-ERKs leading to cytoplasmic accumulation of phospho-ERKs [140], and the cytoplasmic accumulation of phospho-ERKs in response to Ang II in cultured cardiac myocytes could be enhanced by the interaction with β-arrestin and mediate a hypertrophic response [141,142]. Cardiomyocytes from heterozygous GRK2 KO mice were also found to have enhanced fractional shortening in response to “biased agonist” of the AT1 receptor (SII) by inducing β-arrestin 2 recruitment [143]. In addition, it has been reported that overexpression of β-arrestins increases mitochondrial superoxide production, while the knockdown decreases ROS production in failing cardiac fibroblasts and human HF [106,144,145]. Metabolically, the heart relies on exogenous glucose and fatty acids, which are tightly controlled by insulin signaling. Insulin resistance plays a key role in the pathogenesis of HF [146]. Studies have found that inhibiting β-arrestin 1 attenuates glucagon-like peptide 1 (GLP1) signaling in cultured pancreatic β cells, resulting in lower cAMP levels, decreased activity of ERK1/2, and CREB, followed by impaired insulin secretion [147]. β-arrestin 1 can alter insulin signaling by inhibiting insulin-induced proteasomal degradation of IRS-1 and the inhibition of beta-arrestin-1 leads to enhanced IRS-1 degradation and accentuated cellular insulin resistance [148]. Similarly, β-arrestin 2-KO mice developed insulin-resistance, followed by decreased insulin-induced phosphorylation of Akt, GSK-3β, and FOXO1 [149]. Conversely, adipocyte-specific β-arrestin 2 knockout mice showed reduced adiposity and striking metabolic improvements when consuming excess calories [150]. In fact, in the adipose tissue β-arrestin 2 is proposed to be a potent negative regulator of β3ARs. As for GRK2 signaling, better understanding specific mechanisms involving β-arrestins in various tissues will guide future treatment strategies and lead to clarifications on the role of arrestins in metabolic regulation.
4.2. In Neurodegenerative pathology
Dopaminergic neurotransmission in the central nervous system (CNS) regulates behavioral responses such as locomotor activity and reward responses. The loss of dopaminergic cells participates in the progression of neurodegenerative pathologies like Parkinson’s disease and AD. Several studies have shown evidence that β-arrestins signaling downstream of internalized receptors are important for the regulation of dopamine-dependent behaviors [151,152]. Dopamine (D) receptors are GPCRs that share similarities with βARs [153]. Both D2 and D3 receptors couple to Gαi and are the most abundant receptor in the brain, D2 receptors, undergo rapid GRK phosphorylation and arrestin binding, which leads to receptor internalization [43]. Moreover, using different antipsychotics, the D2 receptor was shown to participate in the formation of β-arrestin-mediated signaling complexes by negatively regulating Akt activity [154,155].
GRK2 is expressed in multiple brain regions, including major dopaminergic regions [30]. D3 receptors are localized in the limbic cortex, striatum, hippocampus, and other brain regions that are associated with schizophrenia [156] and found to affect schizophrenia-like behavior of rats [157]. It is reported that the phosphorylation of the D3 receptor by GRK2 disrupts the interaction between the D3 receptor and filamin A (FLNA), potentially through a direct interaction with β-arrestin 2 and regulation of D3 receptors in lipid rafts [158]. Interestingly, internalization of D3 receptors stimulated by a novel compound SK609, was GRK2-dependent without the involvement of β-arrestin 1 or 2 with a proposed mechanism involving clathrin [159]. Conversely, another study reported agonist-induced receptor internalization dependent on β-arrestin 2 but independent of GRK2-mediated receptor phosphorylation [160]. Thus, how GRK2 and β-arrestins participate in D3 receptor internalization and signaling remains debatable.
GRK2 has also been implicated in AD. In a rat model of AD, GRK2 was upregulated in the cerebrovascular system at an early stage [97], and GRK2 immunoreactivity in AD human samples was significantly increased. The increase in GRK2 was thought to be linked to increased oxidative stress in response to chronic hypoperfusion, resulting in dysregulated cerebrovascular metabolism homeostasis [161]. Thus, GRK2-mediated mitochondrial localization found in AD patients has been proposed to be a marker of brain damage caused by early hypoperfusion. Chronic hypoperfusion causes oxidative stress and detrimentally impacts the brain and cerebrovascular homeostasis and metabolism [97,161]. Whether GRK2 upregulation in the AD brain is causative or consequential remains to be determined.
4.3. In pain and inflammation
Multiple immune processes involve GPCR-GRK signaling pathways. Leukocytes are known to express a vast number of GPCRs which are known to tightly regulate immune responses and intercellular interactions [162,163]. Higher expression of GRK2 and GRK5 in neutrophils was found in both LPS-induced sepsis in vitro and in septic patients which induces the phosphorylation of chemotactic receptors including CXCR1 and leads to suppression of neutrophil migration.[164]. Chemokine receptors CCR2B and CXCR4 play an important role in pain and inflammation by promoting chemotaxis via the binding of monocyte chemoattractant proteins and the recruitment of immune cells like macrophages [165,166]. GRK2 interacts with chemokine receptor CCR2B and chemokine CXCR4 receptor to negatively regulate ERK activation [58]. Splenocytes from mice with deficient β-arrestin 2 expression showed impaired CCR4-mediated chemotaxis in both trans-well and trans-endothelial migration tests suggesting a positive role for β-arrestin 2 in mediating the chemotactic responses of T and B lymphocytes [167]. In addition, the ERK pathway can be induced by a variety of inflammatory mediators including LPS, TNFα, interleukin-1 (IL-1), IL-8 and prostaglandin (E2) as extracellular signals, where the overexpression of GRK2 inhibited β-arrestin 2-mediated ERK activation [168]. GRK2 can negatively regulate LPS-induced ERK pathway in macrophages and decrease cytokine production. In fact, myeloid-specific knockdown of GRK2 exaggerates inflammatory cytokine and chemokine production in response to LPS stimulation via the activation of MEK-ERK pathway, and results in organ injury in mice [169], as well as the Akt signaling via inhibited CCL2 [170]. Moreover, LPS-induced cytokine release is increased in GRK2 heterozygous mice through the phosphorylation of p38 at site Thr-123 [171], which suggested a direct interaction between GRK2 and p38. GRK2 expression has also been implicated in microglia and macrophage function, particularly in ischemic brain damage. In fact, GRK2 has been postulated to regulate p38-induced release of TNFα in response to LPS during the hypoxic-ischemic brain damage [172]. Interestingly, a recent study suggested that GRK2 inactivation of p38 MAPK pathway involve microRNA-15a/16 epigenetic regulation of neuropathic pain [173]. Additionally, p38 MAPK signaling can also directly phosphorylate GRK2 and inhibit its translocation to the membrane, thereby preventing the internalization of CCR2, an important receptor signaling pathway in migration and activation of monocytes [174]. Taken together, these studies indicate that GRK2 participates in regulating the MAPK pathway in the immune system.
GRK2 is an important regulator of NF-κB, which is inextricably linked to the inflammatory system via the interaction with IκBα [175] and p105 (another member of IκB family) [169]. β-arrestin 2 has been shown to negatively regulate migration of dendritic cells (DCs), and play a role in inflammatory diseases [176], including asthma [177]. Inflammation can cause hyperalgesia and allodynia. This is due to the increased excitability of peripheral nociceptive sensory nerves caused by the stimulation of inflammatory mediators like peptides, chemokines, cytokines, and neurotransmitters [178]. Inflammatory associated chronic pain has been linked to GRK2 signaling. In neuropathic pain, the level of GRK2 was reduced in the spinal cord in mice [179], underlining the importance of GRK2 in regulating inflammatory hyperalgesia. In fact, knockdown of neuronal GRK2 prolonged prostaglandin E2 (PGE2)- or cAMP- induced hyperalgesia, in an ERK-dependent pathway [180], and either the reduction of GRK2 in microglia/macrophages or in peripheral sensory neurons can both lead to enhanced severity and duration of pain in PGE2 induced pathological conditions [179]. Overall, further detailing GRK2 and β-arrestin mechanism of action are warranted to better understand this signaling pathway and to evaluate the pharmacological potential of targeting these proteins in the treatment of acute inflammatory responses, autoimmune diseases, and pain.
4.4. In cancer
Tumor angiogenesis is a hallmark of cancer and plays an essential role in tumor initiation, progression, and metastasis. Growth and migration of tumors include arrestin signaling via different signaling pathways like CXCR4·CXCR7 complex [181] and β2AR signaling [182]. Analysis of clinical samples and in vivo experiments in mice support a key role for GRK2 in regulating angiogenesis in a variety of tumors [183,184]. Chemokine receptor activation is known to promote angiogenesis [185]. Unlike the classic chemokine receptor CXCR4, recruitment of β-arrestin 2 is different with atypical chemokine receptors (ACKRs) such as CXCR7 [186]. Studies have confirmed that CXCR4 activation induced Gαi signaling and recruitment of β-arrestin 2 through GRK2, where it facilitated stimulation of ERK1/2 signaling. Interestingly, activation of CXCR7 activates GRK2 through Gβ1 and subsequent β-arrestin 2, leading to receptor phosphorylation [187,188]. CXCR7, also known as ACKR3, is key in the progression of various tumors including lung, glioma, and breast cancer [198]. Internalization of CXCR7 is thought to modulate the chemokine potential via the internalization of its agonist CXCL12 [199]. CXCR7 recruits both β-arrestin 1 and 2 via interactions with GRK2 [189].
Many biased-GPCR signaling events involved in tumor growth are linked to GRK/β-arrestin pathways regulating cell growth and apoptosis. Somatostatin receptors (SSTRs), commonly found in neuroendocrine tumors [189], were shown to recruit β-arrestin 2 and in turn lead to robust SSTR2 internalization and accumulation in endosomes which may further reduce the differentiation rate in cancer cells [190,191]. Interestingly, differential β-arrestin recruitment was observed in cells expressing SSTR2 or SSTR2/SSTR5. Co-expression of SSTR2 and SSTR5, disrupted binding of β-arrestin to the receptor, delaying SSTR2 internalization, prolonging GPCR signaling [190] and activating other downstream pathways including MAPK [192].
Evidence also suggests that expression of GRK2 and β-arrestins are corelated with pituitary adenomas [193]. GPR54, a GPCR that can be activated by the neuropeptide kisspeptin, regulates placentation and tumor metastases [194]. Following activation, GPR54 stimulate phosphorylation of ERK1/2 and p38 MAPK pathway via β-arrestin 2 [195]. Overexpression of GRK2 enhanced desensitization of GPR54 by directly interacting with the receptor in HEK 293 cells [196]. Thus GPR54, a key regulator of the hypothalamic-pituitary-gonadal axis, is a promising target for developing therapies to treat endocrine-related disorders [194]. Moreover, inhibition of GRK2 enhanced degradation of cancer-relevant insulin-like growth factor-1 receptor (IGF1R) and prevented its interaction with recruited β-arrestin 2 [197], restraining malignant cell growth. As malignant tumor growth rely on increased bioenergetic demand, the impact of GRK2 on IGF1R signaling warrants further inverstigation as it holds pharmacological promise to various cancer types. Cellular metabolism is key to all cancers as the bioenergetic potential will greatly influence the ability of cancerous cells to migrate, proliferate, and metastasize. GRK2 and β-arrestin have been recently implicated in various metabolic regulatory signaling pathways, yet how GRK2 and β-arrestin participate in altering oncogenic metabolic landscape and impact proliferation and metastasis remain largely understudied. Overall, data supports the notion that GRKs, especially GRK2/β-arrestin signaling, also holds a promise for anticancer therapy.
5. Conclusions
GPCR signaling impacts various cell types and cellular functions. It is not surprising that about one-third of FDA-approved drugs target GPCR signaling in one way or another. As there are so many orphaned GPCRs, one can only imagine the pharmacological promise of what still remains unknown. GPCR availability at the plasma membrane is strongly dependent on GRKs’ ability to dynamically phosphorylate these receptors and subsequently recruit arrestins. In addition to the canonical role of these proteins at the plasma membrane, GRK2 and arrestins have been shown to perform a gamut of other non-receptor functions. Focusing on the GRK2/β-arrestin pathway and their link to various diseases (Fig. 3), we reviewed current literature embracing different scientific fields and human diseases. There is a strong consensus that GRK2 and β-arrestins can regulate intracellular signaling proteins in a GPCR-dependent and independent manner. These activities affect various organelles within a cell and exert strong functional and sometimes diverging effects in different organ systems. These observations are particularly important for pre-clinical studies aiming to target GRK2 and/or β-arrestins to treat specific pathologies. It is imperative to delineate further the newly found role of GRK2 and arrestins in metabolism especially as GRK2 emerges as a promising target for HF. Are there unbeknownst outcomes for these potential interventions? Can we design new pharmacological targets that can beneficially target a specific arm of GRK2/β-arrestin signaling? Could we explore this new knowledge to develop personalized medicine strategies? As we untangle this biological “knitted sweater” of signaling network, one thing is certain, the future is bright, exciting, and promising for GRK2 and arrestin pharmacological strategies in the treatment of various human pathologies.
Fig. 3.

GRK2 and β-arrestin major reported signaling alterations and related pathologies.
Funding
This work was supported by the State of Pennsylvania CURE foundation award, W.W. Smith Foundation Award (H2105), Mary Dewitt Pettit Fellowship, and The National Institutes of Health (R56HL149887 and R01HL163666) and a University of PennsylvaniaDiabetes Research Center Pilot Project Award (P30-DK19525) to (P.Y.S.)
Footnotes
Disclosures
All authors report no financial or personal relationships that may be perceived as influencing their work.
References
- [1].Sriram K, Insel PA, G protein-coupled receptors as targets for approved drugs: how many targets and how many drugs? Mol. Pharmacol 93 (4) (2018) 251–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [2].Nadkarni P, Chepurny OG, Holz GG, Regulation of glucose homeostasis by GLP-1, Prog. Mol. Biol. Transl. Sci 121 (2014) 23–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Kniazeff J, Prezeau L, Rondard P, Pin JP, Goudet C, Dimers and beyond: The functional puzzles of class C GPCRs, Pharmacol. Ther 130 (1) (2011) 9–25. [DOI] [PubMed] [Google Scholar]
- [4].Gubb D, Garcia-Bellido A, A genetic analysis of the determination of cuticular polarity during development in Drosophila melanogaster, J. Embryol. Exp. Morpholog 68 (1982) 37–57. [PubMed] [Google Scholar]
- [5].Wang HY, Malbon CC, Wnt signaling, Ca2+, and cyclic GMP: visualizing Frizzled functions, Science. 300 (5625) (2003) 1529–1530. [DOI] [PubMed] [Google Scholar]
- [6].Wang Y, Chang H, Rattner A, Nathans J, Frizzled receptors in development and disease, Curr. Top. Dev. Biol 117 (2016) 113–139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Fredriksson R, Lagerstrom MC, Lundin LG, Schioth HB, The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints, Mol. Pharmacol 63 (6) (2003) 1256–1272. [DOI] [PubMed] [Google Scholar]
- [8].Langenhan T, Adhesion G protein-coupled receptors-Candidate metabotropic mechanosensors and novel drug targets, Basic Clin. Pharmacol. Toxicol 126 (Suppl. 6) (2020) 5–16. [DOI] [PubMed] [Google Scholar]
- [9].Nickoloff-Bybel EA, Mackie P, Runner K, Matt SM, Khoshbouei H, Gaskill PJ, Dopamine increases HIV entry into macrophages by increasing calcium release via an alternative signaling pathway, Brain Behav. Immun 82 (2019) 239–252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Lambright DG, Noel JP, Hamm HE, Sigler PB, Structural determinants for activation of the alpha-subunit of a heterotrimeric G protein, Nature. 369 (6482) (1994) 621–628. [DOI] [PubMed] [Google Scholar]
- [11].Gudermann T, Kalkbrenner F, Schultz G, Diversity and selectivity of receptor-G protein interaction, Annu. Rev. Pharmacol. Toxicol 36 (1996) 429–459. [DOI] [PubMed] [Google Scholar]
- [12].McCudden CR, Hains MD, Kimple RJ, Siderovski DP, Willard FS, G-protein signaling: back to the future, Cell. Mol. Life Sci 62 (5) (2005) 551–577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Noel JP, Hamm HE, Sigler PB, The 2.2 A crystal structure of transducin-alpha complexed with GTP gamma S, Nature. 366 (6456) (1993) 654–663. [DOI] [PubMed] [Google Scholar]
- [14].Berghuis AM, Lee E, Raw AS, Gilman AG, Sprang SR, Structure of the GDP-Pi complex of Gly203–>Ala gialpha1: a mimic of the ternary product complex of galpha-catalyzed GTP hydrolysis, Structure. 4 (11) (1996) 1277–1290. [DOI] [PubMed] [Google Scholar]
- [15].Sugden PH, Clerk A, Regulation of the ERK subgroup of MAP kinase cascades through G protein-coupled receptors, Cell. Signal 9 (5) (1997) 337–351. [DOI] [PubMed] [Google Scholar]
- [16].Knight KM, Ghosh S, Campbell SL, Lefevre TJ, Olsen RHJ, Smrcka AV, et al. , A universal allosteric mechanism for G protein activation, Mol. Cell 81 (7) (2021) 1384–96 e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Felder CC, Jose PA, Axelrod J, The dopamine-1 agonist, SKF 82526, stimulates phospholipase-C activity independent of adenylate cyclase, J. Pharmacol. Exp. Ther 248 (1) (1989) 171–175. [PubMed] [Google Scholar]
- [18].Jin LQ, Wang HY, Friedman E, Stimulated D(1) dopamine receptors couple to multiple Galpha proteins in different brain regions, J. Neurochem 78 (5) (2001) 981–990. [DOI] [PubMed] [Google Scholar]
- [19].Carman CV, Benovic JL, G-protein-coupled receptors: turn-ons and turn-offs, Curr. Opin. Neurobiol 8 (3) (1998) 335–344. [DOI] [PubMed] [Google Scholar]
- [20].Gurevich VV, Benovic JL, Visual arrestin interaction with rhodopsin. Sequential multisite binding ensures strict selectivity toward light-activated phosphorylated rhodopsin, J. Biol. Chem 268 (16) (1993) 11628–11638. [PubMed] [Google Scholar]
- [21].Gardner LA, Delos Santos NM, Matta SG, Whitt MA, Bahouth SW, Role of the cyclic AMP-dependent protein kinase in homologous resensitization of the beta1-adrenergic receptor, J. Biol. Chem 279 (20) (2004) 21135–21143. [DOI] [PubMed] [Google Scholar]
- [22].Burns RN, Singh M, Senatorov IS, Moniri NH, Mechanisms of homologous and heterologous phosphorylation of FFA receptor 4 (GPR120): GRK6 and PKC mediate phosphorylation of Thr(3)(4)(7), Ser(3)(5)(0), and Ser(3)(5)(7) in the C-terminal tail, Biochem. Pharmacol 87 (4) (2014) 650–659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [23].Daaka Y, Luttrell LM, Lefkowitz RJ, Switching of the coupling of the beta2-adrenergic receptor to different G proteins by protein kinase A, Nature. 390 (6655) (1997) 88–91. [DOI] [PubMed] [Google Scholar]
- [24].Cong M, Perry SJ, Lin FT, Fraser ID, Hu LA, Chen W, et al. , Regulation of membrane targeting of the G protein-coupled receptor kinase 2 by protein kinase A and its anchoring protein AKAP79, J. Biol. Chem 276 (18) (2001) 15192–15199. [DOI] [PubMed] [Google Scholar]
- [25].Violin JD, Ren XR, Lefkowitz RJ, G-protein-coupled receptor kinase specificity for beta-arrestin recruitment to the beta2-adrenergic receptor revealed by fluorescence resonance energy transfer, J. Biol. Chem 281 (29) (2006) 20577–20588. [DOI] [PubMed] [Google Scholar]
- [26].Kuhn H, Light-regulated binding of rhodopsin kinase and other proteins to cattle photoreceptor membranes, Biochemistry. 17 (21) (1978) 4389–4395. [DOI] [PubMed] [Google Scholar]
- [27].Sibley DR, Strasser RH, Benovic JL, Daniel K, Lefkowitz RJ, Phosphorylation/dephosphorylation of the beta-adrenergic receptor regulates its functional coupling to adenylate cyclase and subcellular distribution, Proc. Natl. Acad. Sci. U. S. A 83 (24) (1986) 9408–9412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Penela P, Ribas C, Mayor F Jr., Mechanisms of regulation of the expression and function of G protein-coupled receptor kinases, Cell. Signal 15 (11) (2003) 973–981. [DOI] [PubMed] [Google Scholar]
- [29].Osawa S, Weiss ER, A tale of two kinases in rods and cones, Adv. Exp. Med. Biol 723 (2012) 821–827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Arriza JL, Dawson TM, Simerly RB, Martin LJ, Caron MG, Snyder SH, et al. , The G-protein-coupled receptor kinases beta ARK1 and beta ARK2 are widely distributed at synapses in rat brain, J. Neurosci 12 (10) (1992) 4045–4055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [31].Chen K, Fu C, Chen C, Liu L, Ren H, Han Y, et al. , Role of GRK4 in the regulation of arterial AT1 receptor in hypertension, Hypertension. 63 (2) (2014) 289–296. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Watanabe H, Xu J, Bengra C, Jose PA, Felder RA, Desensitization of human renal D1 dopamine receptors by G protein-coupled receptor kinase 4, Kidney Int. 62 (3) (2002) 790–798. [DOI] [PubMed] [Google Scholar]
- [33].Vatter P, Stoesser C, Samel I, Gierschik P, Moepps B, The variable C-terminal extension of G-protein-coupled receptor kinase 6 constitutes an accessorial autoregulatory domain, FEBS J. 272 (23) (2005) 6039–6051. [DOI] [PubMed] [Google Scholar]
- [34].Koch WJ, Rockman HA, Samama P, Hamilton RA, Bond RA, Milano CA, et al. , Cardiac function in mice overexpressing the beta-adrenergic receptor kinase or a beta ARK inhibitor, Science. 268 (5215) (1995) 1350–1353. [DOI] [PubMed] [Google Scholar]
- [35].Thiyagarajan MM, Stracquatanio RP, Pronin AN, Evanko DS, Benovic JL, Wedegaertner PB, A predicted amphipathic helix mediates plasma membrane localization of GRK5, J. Biol. Chem 279 (17) (2004) 17989–17995. [DOI] [PubMed] [Google Scholar]
- [36].Ribas C, Penela P, Murga C, Salcedo A, Garcia-Hoz C, Jurado-Pueyo M, et al. , The G protein-coupled receptor kinase (GRK) interactome: role of GRKs in GPCR regulation and signaling, Biochim. Biophys. Acta 1768 (4) (2007) 913–922. [DOI] [PubMed] [Google Scholar]
- [37].Pitcher JA, Inglese J, Higgins JB, Arriza JL, Casey PJ, Kim C, et al. , Role of beta gamma subunits of G proteins in targeting the beta-adrenergic receptor kinase to membrane-bound receptors, Science. 257 (5074) (1992) 1264–1267. [DOI] [PubMed] [Google Scholar]
- [38].Lodowski DT, Barnhill JF, Pitcher JA, Capel WD, Lefkowitz RJ, Tesmer JJ, Purification, crystallization and preliminary X-ray diffraction studies of a complex between G protein-coupled receptor kinase 2 and Gbeta1gamma2, Acta Crystallogr D Biol Crystallogr. 59 (Pt 5) (2003) 936–939. [DOI] [PubMed] [Google Scholar]
- [39].Carman CV, Parent JL, Day PW, Pronin AN, Sternweis PM, Wedegaertner PB, et al. , Selective regulation of Galpha(q/11) by an RGS domain in the G protein-coupled receptor kinase, GRK2, J. Biol. Chem 274 (48) (1999) 34483–34492. [DOI] [PubMed] [Google Scholar]
- [40].Freeman K, Lerman I, Kranias EG, Bohlmeyer T, Bristow MR, Lefkowitz RJ, et al. , Alterations in cardiac adrenergic signaling and calcium cycling differentially affect the progression of cardiomyopathy, J. Clin. Invest 107 (8) (2001) 967–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Gurevich EV, Gainetdinov RR, Gurevich VV, G protein-coupled receptor kinases as regulators of dopamine receptor functions, Pharmacol. Res 111 (2016) 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Tiberi M, Nash SR, Bertrand L, Lefkowitz RJ, Caron MG, Differential regulation of dopamine D1A receptor responsiveness by various G protein-coupled receptor kinases, J. Biol. Chem 271 (7) (1996) 3771–3778. [DOI] [PubMed] [Google Scholar]
- [43].Kim KM, Valenzano KJ, Robinson SR, Yao WD, Barak LS, Caron MG, Differential regulation of the dopamine D2 and D3 receptors by G protein-coupled receptor kinases and beta-arrestins, J. Biol. Chem 276 (40) (2001) 37409–37414. [DOI] [PubMed] [Google Scholar]
- [44].Mouledous L, Froment C, Dauvillier S, Burlet-Schiltz O, Zajac JM, Mollereau C, GRK2 protein-mediated transphosphorylation contributes to loss of function of mu-opioid receptors induced by neuropeptide FF (NPFF2) receptors, J. Biol. Chem 287 (16) (2012) 12736–12749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Kameyama K, Haga K, Haga T, Kontani K, Katada T, Fukada Y, Activation by G protein beta gamma subunits of beta-adrenergic and muscarinic receptor kinase, J. Biol. Chem 268 (11) (1993) 7753–7758. [PubMed] [Google Scholar]
- [46].Raghuwanshi SK, Su Y, Singh V, Haynes K, Richmond A, Richardson RM, The chemokine receptors CXCR1 and CXCR2 couple to distinct G protein-coupled receptor kinases to mediate and regulate leukocyte functions, J. Immunol 189 (6) (2012) 2824–2832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [47].Tong X, Zhang L, Zhang L, Hu M, Leng J, Yu B, et al. , The mechanism of chemokine receptor 9 internalization triggered by interleukin 2 and interleukin 4, Cell. Mol. Immunol 6 (3) (2009) 181–189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [48].Busillo JM, Armando S, Sengupta R, Meucci O, Bouvier M, Benovic JL, Site-specific phosphorylation of CXCR4 is dynamically regulated by multiple kinases and results in differential modulation of CXCR4 signaling, J. Biol. Chem 285 (10) (2010) 7805–7817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Aramori I, Ferguson SS, Bieniasz PD, Zhang J, Cullen B, Cullen MG, Molecular mechanism of desensitization of the chemokine receptor CCR-5: receptor signaling and internalization are dissociable from its role as an HIV-1 co-receptor, EMBO J. 16 (15) (1997) 4606–4616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Ang R, Mastitskaya S, Hosford PS, Basalay M, Specterman M, Aziz Q, et al. , Modulation of Cardiac Ventricular Excitability by GLP-1 (Glucagon-Like Peptide-1), Circ. Arrhythm. Electrophysiol 11 (10) (2018), e006740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [51].Arcones AC, Vila-Bedmar R, Mirasierra M, Cruces-Sande M, Vallejo M, Jones B, et al. , GRK2 regulates GLP-1R-mediated early phase insulin secretion in vivo, BMC Biol. 19 (1) (2021) 40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [52].Nobles KN, Xiao K, Ahn S, Shukla AK, Lam CM, Rajagopal S, et al. , Distinct phosphorylation sites on the beta(2)-adrenergic receptor establish a barcode that encodes differential functions of beta-arrestin, Sci. Signal 4 (185) (2011) ra51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Zhang X, Zheng M, Kim KM, GRK2-mediated receptor phosphorylation and Mdm2-mediated beta-arrestin2 ubiquitination drive clathrin-mediated endocytosis of G protein-coupled receptors, Biochem. Biophys. Res. Commun 533 (3) (2020) 383–390. [DOI] [PubMed] [Google Scholar]
- [54].Cho D, Zheng M, Min C, Ma L, Kurose H, Park JH, et al. , Agonist-induced endocytosis and receptor phosphorylation mediate resensitization of dopamine D (2) receptors, Mol. Endocrinol 24 (3) (2010) 574–586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [55].Arthur JW, Sanchez-Perez A, Cook DI, Scoring of predicted GRK2 phosphorylation sites in Nedd4-2, Bioinformatics. 22 (18) (2006) 2192–2195. [DOI] [PubMed] [Google Scholar]
- [56].Naga Prasad SV, Laporte SA, Chamberlain D, Caron MG, Barak L, Rockman HA, Phosphoinositide 3-kinase regulates beta2-adrenergic receptor endocytosis by AP-2 recruitment to the receptor/beta-arrestin complex, J. Cell Biol 158 (3) (2002) 563–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [57].Liu S, Premont RT, Kontos CD, Zhu S, Rockey DC, A crucial role for GRK2 in regulation of endothelial cell nitric oxide synthase function in portal hypertension, Nat. Med 11 (9) (2005) 952–958. [DOI] [PubMed] [Google Scholar]
- [58].Jimenez-Sainz MC, Murga C, Kavelaars A, Jurado-Pueyo M, Krakstad BF, Heijnen CJ, et al. , G protein-coupled receptor kinase 2 negatively regulates chemokine signaling at a level downstream from G protein subunits, Mol. Biol. Cell 17 (1) (2006) 25–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Hagan S, Garcia R, Dhillon A, Kolch W, Raf kinase inhibitor protein regulation of raf and MAPK signaling, Methods Enzymol. 407 (2006) 248–259. [DOI] [PubMed] [Google Scholar]
- [60].Shiina T, Arai K, Tanabe S, Yoshida N, Haga T, Nagao T, et al. , Clathrin box in G protein-coupled receptor kinase 2, J. Biol. Chem 276 (35) (2001) 33019–33026. [DOI] [PubMed] [Google Scholar]
- [61].Premont RT, Claing A, Vitale N, Freeman JL, Pitcher JA, Patton WA, et al. , beta2-Adrenergic receptor regulation by GIT1, a G protein-coupled receptor kinase-associated ADP ribosylation factor GTPase-activating protein, Proc. Natl. Acad. Sci. U. S. A 95 (24) (1998) 14082–14087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [62].Krasel C, Dammeier S, Winstel R, Brockmann J, Mischak H, Lohse MJ, Phosphorylation of GRK2 by protein kinase C abolishes its inhibition by calmodulin, J. Biol. Chem 276 (3) (2001) 1911–1915. [DOI] [PubMed] [Google Scholar]
- [63].Sorriento D, Santulli G, Franco A, Cipolletta E, Napolitano L, Gambardella J, et al. , Integrating GRK2 and NFkappaB in the Pathophysiology of Cardiac Hypertrophy, J. Cardiovasc. Transl. Res 8 (8) (2015) 493–502. [DOI] [PubMed] [Google Scholar]
- [64].Luo J, Benovic JL, G protein-coupled receptor kinase interaction with Hsp90 mediates kinase maturation, J. Biol. Chem 278 (51) (2003) 50908–50914. [DOI] [PubMed] [Google Scholar]
- [65].Penela P, Elorza A, Sarnago S, Mayor F Jr., Beta-arrestin- and c-Src-dependent degradation of G-protein-coupled receptor kinase 2, EMBO J. 20 (18) (2001) 5129–5138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Huang ZM, Gao E, Fonseca FV, Hayashi H, Shang X, Hoffman NE, et al. , Convergence of G protein-coupled receptor and S-nitrosylation signaling determines the outcome to cardiac ischemic injury, Sci. Signal 6 (299) (2013) ra95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Cipolletta E, Campanile A, Santulli G, Sanzari E, Leosco D, Campiglia P, et al. , The G protein coupled receptor kinase 2 plays an essential role in beta-adrenergic receptor-induced insulin resistance, Cardiovasc. Res 84 (3) (2009) 407–415. [DOI] [PubMed] [Google Scholar]
- [68].Pitcher JA, Hall RA, Daaka Y, Zhang J, Ferguson SS, Hester S, et al. , The G protein-coupled receptor kinase 2 is a microtubule-associated protein kinase that phosphorylates tubulin, J. Biol. Chem 273 (20) (1998) 12316–12324. [DOI] [PubMed] [Google Scholar]
- [69].Murphy JE, Roosterman D, Cottrell GS, Padilla BE, Feld M, Brand E, et al. , Protein phosphatase 2A mediates resensitization of the neurokinin 1 receptor, Am. J. Phys. Cell Phys 301 (4) (2011) C780–C791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Doll C, Poll F, Peuker K, Loktev A, Gluck L, Schulz S, Deciphering micro-opioid receptor phosphorylation and dephosphorylation in HEK293 cells, Br. J. Pharmacol 167 (6) (2012) 1259–1270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Abdul Khaliq S, Umair Z, Baek MO, Chon SJ, Yoon MS, C-peptide promotes cell migration by controlling matrix metallopeptidase-9 activity through direct regulation of beta-catenin in human endometrial stromal cells, Front Cell Dev Biol. 10 (2022), 800181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Kaye AD, Edinoff AN, Babin KC, Hebert CM, Hardin JL, Cornett EM, et al. , Pharmacological advances in opioid therapy: a review of the role of oliceridine in pain management, Pain Ther. 10 (2) (2021) 1003–1012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [73].Kaur S, Chen Y, Shenoy SK, Agonist-activated glucagon receptors are deubiquitinated at early endosomes by two distinct deubiquitinases to facilitate Rab4a-dependent recycling, J. Biol. Chem 295 (49) (2020) 16630–16642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Gupta MK, Mohan ML, Naga Prasad SV, G protein-coupled receptor resensitization paradigms, Int. Rev. Cell Mol. Biol 339 (2018) 63–91. [DOI] [PubMed] [Google Scholar]
- [75].Gurevich EV, Benovic JL, Gurevich VV, Arrestin2 expression selectively increases during neural differentiation, J. Neurochem 91 (6) (2004) 1404–1416. [DOI] [PubMed] [Google Scholar]
- [76].Wilden U, Hall SW, Kuhn H, Phosphodiesterase activation by photoexcited rhodopsin is quenched when rhodopsin is phosphorylated and binds the intrinsic 48-kDa protein of rod outer segments, Proc. Natl. Acad. Sci. U. S. A 83 (5) (1986) 1174–1178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [77].Lohse MJ, Andexinger S, Pitcher J, Trukawinski S, Codina J, Faure JP, et al. , Receptor-specific desensitization with purified proteins. Kinase dependence and receptor specificity of beta-arrestin and arrestin in the beta 2-adrenergic receptor and rhodopsin systems, J. Biol. Chem 267 (12) (1992) 8558–8564. [PubMed] [Google Scholar]
- [78].Rapoport B, Kaufman KD, Chazenbalk GD, Cloning of a member of the arrestin family from a human thyroid cDNA library, Mol. Cell. Endocrinol 84 (3) (1992) R39–R43. [DOI] [PubMed] [Google Scholar]
- [79].Srivastava A, Gupta B, Gupta C, Shukla AK, Emerging functional divergence of beta-arrestin isoforms in GPCR function, Trends Endocrinol. Metab 26 (11) (2015) 628–642. [DOI] [PubMed] [Google Scholar]
- [80].Kohout TA, Lin FS, Perry SJ, Conner DA, Lefkowitz RJ, beta-Arrestin 1 and 2 differentially regulate heptahelical receptor signaling and trafficking, Proc. Natl. Acad. Sci. U. S. A 98 (4) (2001) 1601–1606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].Xiao K, McClatchy DB, Shukla AK, Zhao Y, Chen M, Shenoy SK, et al. , Functional specialization of beta-arrestin interactions revealed by proteomic analysis, Proc. Natl. Acad. Sci. U. S. A 104 (29) (2007) 12011–12016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [82].Nakagawa M, Orii H, Yoshida N, Jojima E, Horie T, Yoshida R, et al. , Ascidian arrestin (Ci-arr), the origin of the visual and nonvisual arrestins of vertebrate, Eur. J. Biochem 269 (21) (2002) 5112–5118. [DOI] [PubMed] [Google Scholar]
- [83].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. 9 (9) (2001) 869–880. [DOI] [PubMed] [Google Scholar]
- [84].Yin W, Li Z, Jin M, Yin YL, de Waal PW, Pal K, et al. , A complex structure of arrestin-2 bound to a G protein-coupled receptor, Cell Res. 29 (12) (2019) 971–983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Min K, Yoon HJ, Park JY, Baidya M, Dwivedi-Agnihotri H, Maharana J, et al. , Crystal Structure of beta-Arrestin 2 in Complex with CXCR7 Phosphopeptide, Structure. 28 (9) (2020) 1014–23 e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [86].Nobles KN, Guan Z, Xiao K, Oas TG, Lefkowitz RJ, The active conformation of beta-arrestin1: direct evidence for the phosphate sensor in the N-domain and conformational differences in the active states of beta-arrestins1 and −2, J. Biol. Chem 282 (29) (2007) 21370–21381. [DOI] [PubMed] [Google Scholar]
- [87].Xiao K, Shenoy SK, Nobles K, Lefkowitz RJ, Activation-dependent conformational changes in {beta}-arrestin 2, J. Biol. Chem 279 (53) (2004) 55744–55753. [DOI] [PubMed] [Google Scholar]
- [88].Gether U, Lin S, Kobilka BK, Fluorescent labeling of purified beta 2 adrenergic receptor. Evidence for ligand-specific conformational changes, J. Biol. Chem 270 (47) (1995) 28268–28275. [DOI] [PubMed] [Google Scholar]
- [89].Ippolito M, Benovic JL, Biased agonism at beta-adrenergic receptors, Cell. Signal 80 (2021), 109905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [90].Miess E, Gondin AB, Yousuf A, Steinborn R, Mosslein N, Yang Y, et al. , Multisite phosphorylation is required for sustained interaction with GRKs and arrestins during rapid mu-opioid receptor desensitization, Sci. Signal 11 (539) (2018). [DOI] [PubMed] [Google Scholar]
- [91].Mancini AD, Bertrand G, Vivot K, Carpentier E, Tremblay C, Ghislain J, et al. , beta-arrestin recruitment and biased agonism at free fatty acid Receptor 1, J. Biol. Chem 290 (34) (2015) 21131–21140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [92].Choi M, Staus DP, Wingler LM, Ahn S, Pani B, Capel WD, et al. , G protein-coupled receptor kinases (GRKs) orchestrate biased agonism at the beta2-adrenergic receptor, Sci. Signal 11 (544) (2018). [DOI] [PubMed] [Google Scholar]
- [93].Shenoy SK, Lefkowitz RJ, Receptor regulation: beta-arrestin moves up a notch, Nat. Cell Biol 7 (12) (2005) 1159–1161. [DOI] [PubMed] [Google Scholar]
- [94].Gesty-Palmer D, Chen M, Reiter E, Ahn S, Nelson CD, Wang S, et al. , Distinct beta-arrestin- and G protein-dependent pathways for parathyroid hormone receptor-stimulated ERK1/2 activation, J. Biol. Chem 281 (16) (2006) 10856–10864. [DOI] [PubMed] [Google Scholar]
- [95].Wisler JW, DeWire SM, Whalen EJ, Violin JD, Drake MT, Ahn S, et al. , A unique mechanism of beta-blocker action: carvedilol stimulates beta-arrestin signaling, Proc. Natl. Acad. Sci. U. S. A 104 (42) (2007) 16657–16662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Hupfeld CJ, Olefsky JM, Regulation of receptor tyrosine kinase signaling by GRKs and beta-arrestins, Annu. Rev. Physiol 69 (2007) 561–577. [DOI] [PubMed] [Google Scholar]
- [97].Obrenovich ME, Smith MA, Siedlak SL, Chen SG, de la Torre JC, Perry G, et al. , Overexpression of GRK2 in Alzheimer disease and in a chronic hypoperfusion rat model is an early marker of brain mitochondrial lesions, Neurotox. Res 10 (1) (2006) 43–56. [DOI] [PubMed] [Google Scholar]
- [98].Fusco A, Santulli G, Sorriento D, Cipolletta E, Garbi C, Dorn GW 2nd, et al. , Mitochondrial localization unveils a novel role for GRK2 in organelle biogenesis, Cell. Signal 24 (2) (2012) 468–475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [99].Sato PY, Chuprun JK, Grisanti LA, Woodall MC, Brown BR, Roy R, et al. , Restricting mitochondrial GRK2 post-ischemia confers cardioprotection by reducing myocyte death and maintaining glucose oxidation, Sci. Signal 11 (560) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Chen M, Sato PY, Chuprun JK, Peroutka RJ, Otis NJ, Ibetti J, et al. , Prodeath signaling of G protein-coupled receptor kinase 2 in cardiac myocytes after ischemic stress occurs via extracellular signal-regulated kinase-dependent heat shock protein 90-mediated mitochondrial targeting, Circ. Res 112 (8) (2013) 1121–1134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [101].Sato PY, Chuprun JK, Ibetti J, Cannavo A, Drosatos K, Elrod JW, et al. , GRK2 compromises cardiomyocyte mitochondrial function by diminishing fatty acid-mediated oxygen consumption and increasing superoxide levels, J. Mol. Cell. Cardiol 89 (Pt B) (2015) 360–364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [102].Kasahara A, Cipolat S, Chen Y, Dorn GW 2nd, Scorrano L, Mitochondrial fusion directs cardiomyocyte differentiation via calcineurin and Notch signaling, Science. 342 (6159) (2013) 734–737. [DOI] [PubMed] [Google Scholar]
- [103].Brinks H, Boucher M, Gao E, Chuprun JK, Pesant S, Raake PW, et al. , Level of G protein-coupled receptor kinase-2 determines myocardial ischemia/reperfusion injury via pro- and anti-apoptotic mechanisms, Circ. Res 107 (9) (2010) 1140–1149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [104].Kook S, Zhan X, Cleghorn WM, Benovic JL, Gurevich VV, Gurevich EV, Caspase-cleaved arrestin-2 and BID cooperatively facilitate cytochrome C release and cell death, Cell Death Differ. 21 (1) (2014) 172–184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [105].Povsic TJ, Kohout TA, Lefkowitz RJ, Beta-arrestin1 mediates insulin-like growth factor 1 (IGF-1) activation of phosphatidylinositol 3-kinase (PI3K) and anti-apoptosis, J. Biol. Chem 278 (51) (2003) 51334–51339. [DOI] [PubMed] [Google Scholar]
- [106].Philip JL, Razzaque MA, Han M, Li J, Theccanat T, Xu X, et al. , Regulation of mitochondrial oxidative stress by beta-arrestins in cultured human cardiac fibroblasts, Dis. Model. Mech 8 (12) (2015) 1579–1589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [107].Zhang J, Xiao H, Shen J, Wang N, Zhang Y, Different roles of beta-arrestin and the PKA pathway in mitochondrial ROS production induced by acute beta-adrenergic receptor stimulation in neonatal mouse cardiomyocytes, Biochem. Biophys. Res. Commun 489 (4) (2017) 393–398. [DOI] [PubMed] [Google Scholar]
- [108].Luan B, Zhang Z, Wu Y, Kang J, Pei G, Beta-arrestin2 functions as a phosphorylation-regulated suppressor of UV-induced NF-kappaB activation, EMBO J. 24 (24) (2005) 4237–4246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].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. U. S. A 101 (23) (2004) 8603–8607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [110].Hoeppner CZ, Cheng N, Ye RD, Identification of a nuclear localization sequence in beta-arrestin-1 and its functional implications, J. Biol. Chem 287 (12) (2012) 8932–8943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [111].Kang J, Shi Y, Xiang B, Qu B, Su W, Zhu M, et al. , A nuclear function of beta-arrestin1 in GPCR signaling: regulation of histone acetylation and gene transcription, Cell. 123 (5) (2005) 833–847. [DOI] [PubMed] [Google Scholar]
- [112].Tao Y, Ma L, Liao Z, Le Q, Yu J, Liu X, et al. , Astroglial beta-Arrestin1-mediated nuclear signaling regulates the expansion of neural precursor cells in adult hippocampus, Sci. Rep 5 (2015) 15506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Barella LF, Rossi M, Pydi SP, Meister J, Jain S, Cui Y, et al. , beta-Arrestin-1 is required for adaptive beta-cell mass expansion during obesity, Nat. Commun 12 (1) (2021) 3385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [114].Lymperopoulos A, Rengo G, Gao E, Ebert SN, Dorn GW 2nd, Koch WJ, Reduction of sympathetic activity via adrenal-targeted GRK2 gene deletion attenuates heart failure progression and improves cardiac function after myocardial infarction, J. Biol. Chem 285 (21) (2010) 16378–16386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [115].Matkovich SJ, Diwan A, Klanke JL, Hammer DJ, Marreez Y, Odley AM, et al. , Cardiac-specific ablation of G-protein receptor kinase 2 redefines its roles in heart development and beta-adrenergic signaling, Circ. Res 99 (9) (2006) 996–1003. [DOI] [PubMed] [Google Scholar]
- [116].Raake PW, Vinge LE, Gao E, Boucher M, Rengo G, Chen X, et al. , G protein-coupled receptor kinase 2 ablation in cardiac myocytes before or after myocardial infarction prevents heart failure, Circ. Res 103 (4) (2008) 413–422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Vroon A, Heijnen CJ, Lombardi MS, Cobelens PM, Mayor F Jr., Caron MG, et al. , Reduced GRK2 level in T cells potentiates chemotaxis and signaling in response to CCL4, J. Leukoc. Biol 75 (5) (2004) 901–909. [DOI] [PubMed] [Google Scholar]
- [118].Parvataneni S, Gonipeta B, Packiriswamy N, Lee T, Durairaj H, Parameswaran N, Role of myeloid-specific G-protein coupled receptor kinase-2 in sepsis, Int. J. Clin. Exp. Med 4 (4) (2011) 320–330. [PMC free article] [PubMed] [Google Scholar]
- [119].Sato PY, Chuprun JK, Schwartz M, Koch WJ, The evolving impact of g protein-coupled receptor kinases in cardiac health and disease, Physiol. Rev 95 (2) (2015) 377–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Murga C, Arcones AC, Cruces-Sande M, Briones AM, Salaices M, Mayor F Jr., G Protein-Coupled Receptor Kinase 2 (GRK2) as a Potential Therapeutic Target in Cardiovascular and Metabolic Diseases, Front. Pharmacol 10 (2019) 112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Cannavo A, Koch WJ, GRK2 as negative modulator of NO bioavailability: Implications for cardiovascular disease, Cell. Signal 41 (2018) 33–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [122].Jaber M, Koch WJ, Rockman H, Smith B, Bond RA, Sulik KK, et al. , Essential role of beta-adrenergic receptor kinase 1 in cardiac development and function, Proc. Natl. Acad. Sci. U. S. A 93 (23) (1996) 12974–12979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [123].Snyder J, Lackey AI, Brown GS, Diaz M, Yuzhen T, Sato PY, GRK2 contributes to glucose mediated calcium responses and insulin secretion in pancreatic islet cells, Sci. Rep 11 (1) (2021) 11129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Conner DA, Mathier MA, Mortensen RM, Christe M, Vatner SF, Seidman CE, et al. , beta-Arrestin1 knockout mice appear normal but demonstrate altered cardiac responses to beta-adrenergic stimulation, Circ. Res 81 (6) (1997) 1021–1026. [DOI] [PubMed] [Google Scholar]
- [125].Bohn LM, Lefkowitz RJ, Gainetdinov RR, Peppel K, Caron MG, Lin FT, Enhanced morphine analgesia in mice lacking beta-arrestin 2, Science. 286 (5449) (1999) 2495–2498. [DOI] [PubMed] [Google Scholar]
- [126].Yi XP, Gerdes AM, Li F, Myocyte redistribution of GRK2 and GRK5 in hypertensive, heart-failure-prone rats, Hypertension. 39 (6) (2002) 1058–1063. [DOI] [PubMed] [Google Scholar]
- [127].Iaccarino G, Tomhave ED, Lefkowitz RJ, Koch WJ, Reciprocal in vivo regulation of myocardial G protein-coupled receptor kinase expression by beta-adrenergic receptor stimulation and blockade, Circulation. 98 (17) (1998) 1783–1789. [DOI] [PubMed] [Google Scholar]
- [128].Ungerer M, Bohm M, Elce JS, Erdmann E, Lohse MJ, Altered expression of beta-adrenergic receptor kinase and beta 1-adrenergic receptors in the failing human heart, Circulation. 87 (2) (1993) 454–463. [DOI] [PubMed] [Google Scholar]
- [129].Lai S, Fu X, Yang S, Zhang S, Lin Q, Zhang M, et al. , G protein-coupled receptor kinase-2: A potential biomarker for early diabetic cardiomyopathy, J Diabetes. 12 (3) (2020) 247–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [130].Rockman HA, Chien KR, Choi DJ, Iaccarino G, Hunter JJ, Ross J Jr., et al. , Expression of a beta-adrenergic receptor kinase 1 inhibitor prevents the development of myocardial failure in gene-targeted mice, Proc. Natl. Acad. Sci. U. S. A 95 (12) (1998) 7000–7005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [131].Noma T, Lemaire A, Naga Prasad SV, Barki-Harrington L, Tilley DG, Chen J, et al. , Beta-arrestin-mediated beta1-adrenergic receptor transactivation of the EGFR confers cardioprotection, J. Clin. Invest 117 (9) (2007) 2445–2458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Rockman HA, Choi DJ, Akhter SA, Jaber M, Giros B, Lefkowitz RJ, et al. , Control of myocardial contractile function by the level of beta-adrenergic receptor kinase 1 in gene-targeted mice, J. Biol. Chem 273 (29) (1998) 18180–18184. [DOI] [PubMed] [Google Scholar]
- [133].Volkers M, Weidenhammer C, Herzog N, Qiu G, Spaich K, Wegner FV, et al. , The inotropic peptide betaARKct improves betaAR responsiveness in normal and failing cardiomyocytes through G(betagamma)-mediated L-type calcium current disinhibition, Circ. Res 108 (1) (2011) 27–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Rengo G, Lymperopoulos A, Zincarelli C, Donniacuo M, Soltys S, Rabinowitz JE, et al. , Myocardial adeno-associated virus serotype 6-betaARKct gene therapy improves cardiac function and normalizes the neurohormonal axis in chronic heart failure, Circulation. 119 (1) (2009) 89–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [135].Lymperopoulos A, Rengo G, Funakoshi H, Eckhart AD, Koch WJ, Adrenal GRK2 upregulation mediates sympathetic overdrive in heart failure, Nat. Med 13 (3) (2007) 315–323. [DOI] [PubMed] [Google Scholar]
- [136].Tutunea-Fatan E, Abd-Elrahman KS, Thibodeau JF, Holterman CE, Holleran BJ, Leduc R, et al. , GRK2 knockdown in mice exacerbates kidney injury and alters renal mechanisms of blood pressure regulation, Sci. Rep 8 (1) (2018) 11415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Ciccarelli M, Chuprun JK, Rengo G, Gao E, Wei Z, Peroutka RJ, Gold JI, Gumpert A, Chen M, Otis NJ, Dorn GW 2nd, Trimarco B, Iaccarino G, Koch WJ. G protein-coupled receptor kinase 2 activity impairs cardiac glucose uptake and promotes insulin resistance after myocardial ischemia. Circulation. 2011. May 10;123(18): 1953–62. doi: 10.1161/CIRCULATIONAHA.110.988642. Epub 2011 Apr 25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [138].Woodall BP, Gresham KS, Woodall MA, Valenti MC, Cannavo A, Pfleger J, et al. , Alteration of myocardial GRK2 produces a global metabolic phenotype. JCI, Insight. 5 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [139].Baker KM, Booz GW, Dostal DE, Cardiac actions of angiotensin II: Role of an intracardiac renin-angiotensin system, Annu. Rev. Physiol 54 (1992) 227–241. [DOI] [PubMed] [Google Scholar]
- [140].Seta K, Nanamori M, Modrall JG, Neubig RR, Sadoshima J, AT1 receptor mutant lacking heterotrimeric G protein coupling activates the Src-Ras-ERK pathway without nuclear translocation of ERKs, J. Biol. Chem 277 (11) (2002) 9268–9277. [DOI] [PubMed] [Google Scholar]
- [141].Zhai P, Yamamoto M, Galeotti J, Liu J, Masurekar M, Thaisz J, et 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 115 (11) (2005) 3045–3056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [142].Tohgo A, Pierce KL, Choy EW, Lefkowitz RJ, Luttrell LM, beta-Arrestin scaffolding of the ERK cascade enhances cytosolic ERK activity but inhibits ERK-mediated transcription following angiotensin AT1a receptor stimulation, J. Biol. Chem 277 (11) (2002) 9429–9436. [DOI] [PubMed] [Google Scholar]
- [143].Rajagopal K, Whalen EJ, Violin JD, Stiber JA, Rosenberg PB, Premont RT, et al. , Beta-arrestin2-mediated inotropic effects of the angiotensin II type 1A receptor in isolated cardiac myocytes, Proc. Natl. Acad. Sci. U. S. A 103 (44) (2006) 16284–16289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [144].Packer M, Fowler MB, Roecker EB, Coats AJ, Katus HA, Krum H, et al. , Effect of carvedilol on the morbidity of patients with severe chronic heart failure: results of the carvedilol prospective randomized cumulative survival (COPERNICUS) study, Circulation. 106 (17) (2002) 2194–2199. [DOI] [PubMed] [Google Scholar]
- [145].Zhang Z, Hao J, Zhao Z, Ben P, Fang F, Shi L, et al. , beta-Arrestins facilitate ubiquitin-dependent degradation of apoptosis signal-regulating kinase 1 (ASK1) and attenuate H2O2-induced apoptosis, Cell. Signal 21 (7) (2009) 1195–1206. [DOI] [PubMed] [Google Scholar]
- [146].Rutter MK, Parise H, Benjamin EJ, Levy D, Larson MG, Meigs JB, et al. , Impact of glucose intolerance and insulin resistance on cardiac structure and function: sex-related differences in the Framingham Heart Study, Circulation. 107 (3) (2003) 448–454. [DOI] [PubMed] [Google Scholar]
- [147].Sonoda N, Imamura T, Yoshizaki T, Babendure JL, Lu JC, Olefsky JM, Beta-Arrestin-1 mediates glucagon-like peptide-1 signaling to insulin secretion in cultured pancreatic beta cells, Proc. Natl. Acad. Sci. U. S. A 105 (18) (2008) 6614–6619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [148].Usui I, Imamura T, Huang J, Satoh H, Shenoy SK, Lefkowitz RJ, et al. , beta-arrestin-1 competitively inhibits insulin-induced ubiquitination and degradation of insulin receptor substrate 1, Mol. Cell. Biol 24 (20) (2004) 8929–8937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [149].Luan B, Zhao J, Wu H, Duan B, Shu G, Wang X, et al. , Deficiency of a beta-arrestin-2 signal complex contributes to insulin resistance, Nature. 457 (7233) (2009) 1146–1149. [DOI] [PubMed] [Google Scholar]
- [150].Pydi SP, Jain S, Tung W, Cui Y, Zhu L, Sakamoto W, et al. , Adipocyte beta-arrestin-2 is essential for maintaining whole body glucose and energy homeostasis, Nat. Commun 10 (1) (2019) 2936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [151].Urs NM, Daigle TL, Caron MG, A dopamine D1 receptor-dependent beta-arrestin signaling complex potentially regulates morphine-induced psychomotor activation but not reward in mice, Neuropsychopharmacology. 36 (3) (2011) 551–558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].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. 122 (2) (2005) 261–273. [DOI] [PubMed] [Google Scholar]
- [153].Dixon RA, Kobilka BK, Strader DJ, Benovic JL, Dohlman HG, Frielle T, et al. , Cloning of the gene and cDNA for mammalian beta-adrenergic receptor and homology with rhodopsin, Nature. 321 (6065) (1986) 75–79. [DOI] [PubMed] [Google Scholar]
- [154].Beaulieu JM, Marion S, Rodriguiz RM, Medvedev IO, Sotnikova TD, Ghisi V, et al. , A beta-arrestin 2 signaling complex mediates lithium action on behavior, Cell. 132 (1) (2008) 125–136. [DOI] [PubMed] [Google Scholar]
- [155].Masri B, Salahpour A, Didriksen M, Ghisi V, Beaulieu JM, Gainetdinov RR, et al. , Antagonism of dopamine D2 receptor/beta-arrestin 2 interaction is a common property of clinically effective antipsychotics, Proc. Natl. Acad. Sci. U. S. A 105 (36) (2008) 13656–13661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [156].Guma E, Rocchetti J, Devenyi GA, Tanti A, Mathieu AP, Lerch JP, et al. , Role of D3 dopamine receptors in modulating neuroanatomical changes in response to antipsychotic administration, Sci. Rep 9 (1) (2019) 7850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Watson DJ, Loiseau F, Ingallinesi M, Millan MJ, Marsden CA, Fone KC, Selective blockade of dopamine D3 receptors enhances while D2 receptor antagonism impairs social novelty discrimination and novel object recognition in rats: a key role for the prefrontal cortex, Neuropsychopharmacology. 37 (3) (2012) 770–786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [158].Kim KM, Gainetdinov RR, Laporte SA, Caron MG, Barak LS, G protein-coupled receptor kinase regulates dopamine D3 receptor signaling by modulating the stability of a receptor-filamin-beta-arrestin complex. A case of autoreceptor regulation, J. Biol. Chem 280 (13) (2005) 12774–12780. [DOI] [PubMed] [Google Scholar]
- [159].Xu W, Reith MEA, Liu-Chen LY, Kortagere S, Biased signaling agonist of dopamine D3 receptor induces receptor internalization independent of beta-arrestin recruitment, Pharmacol. Res 143 (2019) 48–57. [DOI] [PubMed] [Google Scholar]
- [160].Namkung Y, Dipace C, Javitch JA, Sibley DR, G protein-coupled receptor kinase-mediated phosphorylation regulates post-endocytic trafficking of the D2 dopamine receptor, J. Biol. Chem 284 (22) (2009) 15038–15051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [161].Obrenovich ME, Palacios HH, Gasimov E, Leszek J, Aliev G, The GRK2 Overexpression Is a Primary Hallmark of Mitochondrial Lesions during Early Alzheimer Disease, Cardiovasc. Psychiatry Neurol 2009 (2009), 327360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [162].Lammermann T, Germain RN, The multiple faces of leukocyte interstitial migration, Semin. Immunopathol 36 (2) (2014) 227–251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [163].Kehrl JH, The impact of RGS and other G-protein regulatory proteins on Galphai-mediated signaling in immunity, Biochem. Pharmacol 114 (2016) 40–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [164].Arraes SM, Freitas MS, da Silva SV, de Paula Neto HA, Alves-Filho JC, Auxiliadora Martins M, et al. , Impaired neutrophil chemotaxis in sepsis associates with GRK expression and inhibition of actin assembly and tyrosine phosphorylation, Blood. 108 (9) (2006) 2906–2913. [DOI] [PubMed] [Google Scholar]
- [165].Weisberg SP, Hunter D, Huber R, Lemieux J, Slaymaker S, Vaddi K, et al. , CCR2 modulates inflammatory and metabolic effects of high-fat feeding, J. Clin. Invest 116 (1) (2006) 115–124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [166].Menichella DM, Abdelhak B, Ren D, Shum A, Frietag C, Miller RJ, CXCR4 chemokine receptor signaling mediates pain in diabetic neuropathy, Mol. Pain 10 (2014) 42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [167].Fong AM, Premont RT, Richardson RM, Yu YR, Lefkowitz RJ, Patel DD, Defective lymphocyte chemotaxis in beta-arrestin2- and GRK6-deficient mice, Proc. Natl. Acad. Sci. U. S. A 99 (11) (2002) 7478–7483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [168].Kim J, Ahn S, Ren XR, Whalen EJ, Reiter E, Wei H, et al. , Functional antagonism of different G protein-coupled receptor kinases for beta-arrestin-mediated angiotensin II receptor signaling, Proc. Natl. Acad. Sci. U. S. A 102 (5) (2005) 1442–1447. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [169].Patial S, Saini Y, Parvataneni S, Appledorn DM, Dorn GW 2nd, Lapres JJ, et al. , Myeloid-specific GPCR kinase-2 negatively regulates NF-kappaB1p105-ERK pathway and limits endotoxemic shock in mice, J. Cell. Physiol 226 (3) (2011) 627–637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [170].Kleibeuker W, Jurado-Pueyo M, Murga C, Eijkelkamp N, Mayor F Jr., Heijnen CJ, et al. , Physiological changes in GRK2 regulate CCL2-induced signaling to ERK1/2 and Akt but not to MEK1/2 and calcium, J. Neurochem 104 (4) (2008) 979–992. [DOI] [PubMed] [Google Scholar]
- [171].Peregrin S, Jurado-Pueyo M, Campos PM, Sanz-Moreno V, Ruiz-Gomez A, Crespo P, et al. , Phosphorylation of p38 by GRK2 at the docking groove unveils a novel mechanism for inactivating p38MAPK, Curr. Biol 16 (20) (2006) 2042–2047. [DOI] [PubMed] [Google Scholar]
- [172].Nijboer CH, Heijnen CJ, Willemen HL, Groenendaal F, Dorn GW 2nd, van Bel F, et al. , Cell-specific roles of GRK2 in onset and severity of hypoxic-ischemic brain damage in neonatal mice, Brain Behav. Immun 24 (3) (2010) 420–426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [173].Li T, Wan Y, Sun L, Tao S, Chen P, Liu C, et al. , Inhibition of MicroRNA-15a/16 Expression Alleviates Neuropathic Pain Development through Upregulation of G Protein-Coupled Receptor Kinase 2, Biomol Ther (Seoul). 27 (4) (2019) 414–422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [174].Liu Z, Jiang Y, Li Y, Wang J, Fan L, Scott MJ, et al. , TLR4 Signaling augments monocyte chemotaxis by regulating G protein-coupled receptor kinase 2 translocation, J. Immunol 191 (2) (2013) 857–864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [175].Parameswaran R, Lunning M, Mantha S, Devlin S, Hamilton A, Schwartz G, et al. , Romiplostim for management of chemotherapy-induced thrombocytopenia, Support Care Cancer 22 (5) (2014) 1217–1222. [DOI] [PubMed] [Google Scholar]
- [176].Cai Y, Yang C, Yu X, Qian J, Dai M, Wang Y, et al. , Deficiency of beta-Arrestin 2 in Dendritic Cells Contributes to Autoimmune Diseases, J. Immunol 202 (2) (2019) 407–420. [DOI] [PubMed] [Google Scholar]
- [177].Walker JK, Fong AM, Lawson BL, Savov JD, Patel DD, Schwartz DA, et al. , Beta-arrestin-2 regulates the development of allergic asthma, J. Clin. Invest 112 (4) (2003) 566–574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [178].Linley JE, Rose K, Ooi L, Gamper N, Understanding inflammatory pain: ion channels contributing to acute and chronic nociception, Pflugers Arch. 459 (5) (2010) 657–669. [DOI] [PubMed] [Google Scholar]
- [179].Eijkelkamp N, Heijnen CJ, Willemen HL, Deumens R, Joosten EA, Kleibeuker W, et al. , GRK2: a novel cell-specific regulator of severity and duration of inflammatory pain, J. Neurosci 30 (6) (2010) 2138–2149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [180].Eijkelkamp N, Wang H, Garza-Carbajal A, Willemen HL, Zwartkruis FJ, Wood JN, et al. , Low nociceptor GRK2 prolongs prostaglandin E2 hyperalgesia via biased cAMP signaling to Epac/Rap1, protein kinase Cepsilon, and MEK/ERK, J. Neurosci 30 (38) (2010) 12806–12815. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [181].Decaillot FM, Kazmi MA, Lin Y, Ray-Saha S, Sakmar TP, Sachdev P, CXCR7/CXCR4 heterodimer constitutively recruits beta-arrestin to enhance cell migration, J. Biol. Chem 286 (37) (2011) 32188–32197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [182].Zhang P, He X, Tan J, Zhou X, Zou L, beta-arrestin2 mediates beta-2 adrenergic receptor signaling inducing prostate cancer cell progression, Oncol. Rep 26 (6) (2011) 1471–1477. [DOI] [PubMed] [Google Scholar]
- [183].Kuai J, Han C, Wei W, Potential Regulatory Roles of GRK2 in Endothelial Cell Activity and Pathological Angiogenesis, Front. Immunol 12 (2021), 698424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [184].Rivas V, Carmona R, Munoz-Chapuli R, Mendiola M, Nogues L, Reglero C, et al. , Developmental and tumoral vascularization is regulated by G protein-coupled receptor kinase 2, J. Clin. Invest 123 (11) (2013) 4714–4730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [185].Carmeliet P, Jain RK, Molecular mechanisms and clinical applications of angiogenesis, Nature. 473 (7347) (2011) 298–307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [186].Ulvmar MH, Hub E, Rot A, Atypical chemokine receptors, Exp. Cell Res 317 (5) (2011) 556–568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [187].Nguyen HT, Reyes-Alcaraz A, Yong HJ, Nguyen LP, Park HK, Inoue A, et al. , CXCR7: a beta-arrestin-biased receptor that potentiates cell migration and recruits beta-arrestin2 exclusively through Gbetagamma subunits and GRK2, Cell Biosci. 10 (1) (2020) 134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [188].Zarca A, Perez C, van den Bor J, Bebelman JP, Heuninck J, de Jonker RJF, et al. , Differential Involvement of ACKR3 C-Tail in beta-Arrestin Recruitment, Trafficking and Internalization, Cells. 10 (3) (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- [189].Reubi JC, Kvols LK, Waser B, Nagorney DM, Heitz PU, Charboneau JW, et al. , Detection of somatostatin receptors in surgical and percutaneous needle biopsy samples of carcinoids and islet cell carcinomas, Cancer Res. 50 (18) (1990) 5969–5977. [PubMed] [Google Scholar]
- [190].Grant M, Alturaihi H, Jaquet P, Collier B, Kumar U, Cell growth inhibition and functioning of human somatostatin receptor type 2 are modulated by receptor heterodimerization, Mol. Endocrinol 22 (10) (2008) 2278–2292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [191].Zou Y, Tan H, Zhao Y, Zhou Y, Cao L, Expression and selective activation of somatostatin receptor subtypes induces cell cycle arrest in cancer cells, Oncol. Lett 17 (2) (2019) 1723–1731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [192].Florio T, Morini M, Villa V, Arena S, Corsaro A, Thellung S, et al. , Somatostatin inhibits tumor angiogenesis and growth via somatostatin receptor-3-mediated regulation of endothelial nitric oxide synthase and mitogen-activated protein kinase activities, Endocrinology. 144 (4) (2003) 1574–1584. [DOI] [PubMed] [Google Scholar]
- [193].Gatto F, Feelders R, van der Pas R, Kros JM, Dogan F, van Koetsveld PM, et al. , beta-Arrestin 1 and 2 and G protein-coupled receptor kinase 2 expression in pituitary adenomas: role in the regulation of response to somatostatin analogue treatment in patients with acromegaly, Endocrinology. 154 (12) (2013) 4715–4725. [DOI] [PubMed] [Google Scholar]
- [194].de Roux N, Genin E, Carel JC, Matsuda F, Chaussain JL, Milgrom E, Hypogonadotropic hypogonadism due to loss of function of the KiSS1-derived peptide receptor GPR54, Proc. Natl. Acad. Sci. U. S. A 100 (19) (2003) 10972–10976. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [195].Kotani M, Detheux M, Vandenbogaerde A, Communi D, Vanderwinden JM, Le Poul E, et al. , The metastasis suppressor gene KiSS-1 encodes kisspeptins, the natural ligands of the orphan G protein-coupled receptor GPR54, J. Biol. Chem 276 (37) (2001) 34631–34636. [DOI] [PubMed] [Google Scholar]
- [196].Pampillo M, Camuso N, Taylor JE, Szereszewski JM, Ahow MR, Zajac M, et al. , Regulation of GPR54 signaling by GRK2 and {beta}-arrestin, Mol. Endocrinol 23 (12) (2009) 2060–2074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [197].Crudden C, Shibano T, Song D, Dragomir MP, Cismas S, Serly J, et al. , Inhibition of G Protein-Coupled Receptor Kinase 2 Promotes Unbiased Downregulation of IGF1 Receptor and Restrains Malignant Cell Growth, Cancer Res. 81 (2) (2021) 501–514. [DOI] [PubMed] [Google Scholar]
