Next Article in Journal
Literature Review: The sFlt1/PlGF Ratio and Pregestational Maternal Comorbidities: New Risk Factors to Predict Pre-Eclampsia
Next Article in Special Issue
Functional Role of Arrestin-1 Residues Interacting with Unphosphorylated Rhodopsin Elements
Previous Article in Journal
Local Microtubule and F-Actin Distributions Fully Constrain the Spatial Geometry of Drosophila Sensory Dendritic Arbors
Previous Article in Special Issue
α1-Adrenergic Receptors: Insights into Potential Therapeutic Opportunities for COVID-19, Heart Failure, and Alzheimer’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unveiling the Differences in Signaling and Regulatory Mechanisms between Dopamine D2 and D3 Receptors and Their Impact on Behavioral Sensitization

Department of Pharmacology, College of Pharmacy, Chonnam National University, Gwang-Ju 61186, Republic of Korea
Int. J. Mol. Sci. 2023, 24(7), 6742; https://doi.org/10.3390/ijms24076742
Submission received: 10 March 2023 / Revised: 25 March 2023 / Accepted: 28 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue G Protein-Coupled Receptors: Signaling and Regulation)

Abstract

:
Dopamine receptors are classified into five subtypes, with D2R and D3R playing a crucial role in regulating mood, motivation, reward, and movement. Whereas D2R are distributed widely across the brain, including regions responsible for motor functions, D3R are primarily found in specific areas related to cognitive and emotional functions, such as the nucleus accumbens, limbic system, and prefrontal cortex. Despite their high sequence homology and similar signaling pathways, D2R and D3R have distinct regulatory properties involving desensitization, endocytosis, posttranslational modification, and interactions with other cellular components. In vivo, D3R is closely associated with behavioral sensitization, which leads to increased dopaminergic responses. Behavioral sensitization is believed to result from D3R desensitization, which removes the inhibitory effect of D3R on related behaviors. Whereas D2R maintains continuous signal transduction through agonist-induced receptor phosphorylation, arrestin recruitment, and endocytosis, which recycle and resensitize desensitized receptors, D3R rarely undergoes agonist-induced endocytosis and instead is desensitized after repeated agonist exposure. In addition, D3R undergoes more extensive posttranslational modifications, such as glycosylation and palmitoylation, which are needed for its desensitization. Overall, a series of biochemical settings more closely related to D3R could be linked to D3R-mediated behavioral sensitization.

1. Introduction

Dopamine (DA), an ethylamine with an attached catechol group (a phenyl group with two hydroxyl groups attached to meta and para positions) is released from nerve endings upon axonal stimulation. Released DA acts on postsynaptic and presynaptic receptors at the synapse and is mostly taken up back into nerve endings by the dopamine transporter protein, which belongs to solute carrier transporter family.
DA modulates the neuronal activities in the brain regions that are innervated by three major dopaminergic pathways: nigrostriatal, mesocorticolimbic, and tuberoinfundibular. Drugs acting on D2R and D3R have traditionally been utilized for the treatment of various disorders including Parkinson’s disease [1], schizophrenia [2], and hyperprolactinemia [3].
The dopamine receptors are pharmacologically classified into D1-like and D2-like families [4,5,6]. D1-like receptors comprise D1 and D5 receptors (D1R, D5R) [7,8,9], whereas D2-like receptors consist of D2, D3, and D4 receptors (D2R, D3R, D4R) [10,11,12]. D1R and D5R mediate the stimulation of adenylyl cyclase by coupling to Gs, whereas D2R, D3R, and D4R exert inhibitory effects on this enzyme by coupling to Gi/o. D2-like receptors have larger third cytoplasmic loops and smaller carboxyl tails compared with D1-like receptors.
The architectures of D2R and D3R are similar, with 46% overall amino acid homology and 78% identity in the transmembrane domains [13]. Similarly, D2R and D3R share many signaling properties when they are expressed in mammalian cells. For example, both receptors inhibit adenylyl cyclase [14,15,16] and stimulate extracellular acidification (Na+/H+ exchange) [14,17,18], mitogenesis [14,19], MAPK activation [20,21], dopamine release [22], and the opening of inward rectifier potassium channels [23,24].
Whereas D2R and D3R exhibit a high degree of similarity in sequence homology and signaling properties, they differ in some other respects. For example, the transcript levels of D3R are much lower than those of D2R, differing by several orders of magnitude in their expression levels [25]. However, D3R shows about 150 times higher affinity for DA [26]. In addition, the distribution of D3R in the brain is more restricted than that of D2R, that is, D3R shows preferential expression in brain regions responsible for emotional and cognitive functions, such as the nucleus accumbens and the islands of Calleja and has been suggested as a therapeutic target for the affective and mood-associated diseases [27,28,29]. Moreover, significant variations exist between D2R and D3R with respect to their endocytosis, desensitization, and posttranslational modifications [30,31,32].
The main objective of this review is to comprehend the functional implications of the distinct regulatory processes of D2R and D3R by highlighting their differences. Functional in vivo roles and ligands of D2R and D3R are described in other excellent reviews [33,34,35,36,37]. This review focuses on the regulatory functions of D2R and D3R in biochemical pathways.

2. Desensitization and Intracellular Trafficking

2.1. Desensitization

Desensitization of G protein-coupled receptors (GPCRs) represents a gradual attenuation of receptor responsiveness by continuous or repeated exposure to agonists [38,39,40]. GPCR desensitization is generally classified into two categories in which either GRKs/arrestins or second messenger-regulated kinases are involved. The former desensitizes only agonist-activated receptors, whereas the latter can desensitize receptors both in an activation-dependent and activation-independent manner [40]. Typical second messenger-regulated kinases are protein kinase A (PKA) and protein kinase C (PKC) [41,42,43].
The desensitization of GPCRs has been most extensively characterized by the conventional or steric-hindrance-based uncoupling model. When these receptors are activated by agonists, they undergo conformational changes that involve the outward movement of transmembrane helix VI. This results in the formation of a cavity that allows for the binding of G proteins [44,45,46]. The (E/D)x(I/L)xxxGL motif, which is commonly present in the C-termini of Gα and arrestins, is believed to participate in their interaction with the active GPCRs at the same binding crevice [47,48]. The binding of arrestins to GPCRs is likely to occur in competition with G proteins. By binding to the same binding crevice, arrestins can obstruct further coupling of G proteins via steric hindrance, which ultimately leads to a decrease in receptor signaling [49].
Even though D2R and D3R have similar structural features and signaling pathways, they display distinct properties of desensitization and intracellular trafficking. In response to agonistic stimulation, D2R is phosphorylated in a manner dependent on GRK2/3 and in part on GRK5 and mediates arrestin translocation to the plasma membrane accompanied by receptor endocytosis [30]. In contrast, D3R is rarely phosphorylated and undergoes only a minute amount of endocytosis.
According to the steric-hindrance-based uncoupling model, D2R but not D3R would show agonist-induced desensitization. However, subsequent studies have reported the desensitization of D3R but not D2R [50,51,52], suggesting that the steric-hindrance-based uncoupling model may not be applicable to the desensitization of D2R and D3R or desensitized D2R might be rapidly resensitized [51,53].
Recent studies conducted on D2-like receptors (D2R, D3R, D4R) and the β2 adrenoceptor (β2AR) indicate that receptors that undergo desensitization have a predisposition to form a stable complex with Gβγ and enable a basal interaction between Gβγ and arrestins [54]. This arrangement leads to the recruitment of Mdm2 to the cytoplasm by these receptors in their basal state, ultimately resulting in the constitutive ubiquitination of arrestins in the cytoplasm [55].
In a recent study, it was suggested that certain GPCRs, including D3R and β2AR, undergo desensitization via a cascade involving Src, PDK1, and Akt. Akt then interacts with USP33, an arrestin deubiquitinase, to facilitate the deubiquitination of arrestins [32]. Following deubiquitination, arrestins form a complex with Gβγ and translocate to the nucleus via an importin complex. Here, they sequester Gβγ from the receptor and Gα, thus dampening receptor signaling. Therefore, it is plausible that appropriate desensitization measures are in place to properly control the cellular responses depending on the characteristics of receptors, agonists, and other cellular factors.
In the case of D2R, which does not display noticeable desensitization under the same experimental conditions where D3R undergoes desensitization [50,52], GRK-mediated receptor phosphorylation is needed for the recycling of endocytosed receptors [53]. This recycling accompanies the dissociation of arrestins from the receptor, which is believed to be the cellular process required for the resensitization of D2R [51].
In contrast to agonist-induced desensitization, which is mainly associated with D3R, PKC-mediated desensitization affects both D2R and D3R [43,56].
The amino acid residues S145/146 and C147 are responsible for the agonist-induced desensitization of D3R [50,52]; S325 for the PKC-mediated desensitization of D2SR (short isoform of the D2R generated by alternative splicing) [56]; and S229 and S257 for the PKC-mediated desensitization of D3R [43] (Table 1, Figure 1 and Figure 2).

2.2. Endocytosis

The process of endocytosis involves the uptake of extracellular materials by cells through the inward budding of vesicles that originate from the plasma membrane [60]. Endocytosis of the receptor occurs without failure when agonist stimulation induces receptor phosphorylation and arrestin translocation. Translocated arrestins mediate receptor endocytosis [61,62] and additional components of endocytic machinery, such as dynamin and β2-adaptin, are also involved in this process [63,64]. The interaction of arrestins with receptors in endocytic vesicles is dependent on the affinity between the two, with varying levels of association observed [65]. Class A receptors, such as the β2AR and D1R, exhibit lower affinity towards arrestins compared with class B receptors. As a result, these receptors tend to dissociate from arrestins when they are incorporated into endocytic vesicles. In contrast, class B receptors, such as the angiotensin II type 1A receptor and vasopressin V2 receptor, maintain their interaction with arrestins during endocytosis. This ability of arrestins to remain associated with class B receptors is facilitated by specific clusters of serine and threonine residues located in the carboxyl-terminal tails of the receptors [66].
Similar to β2AR, D2R undergoes agonist-induced endocytosis. GRK2/3 and GRK5 in part mediate the phosphorylation of D2R. In contrast, GRK-mediated phosphorylation rarely occurs in D3R, which undergoes only a minute amount of endocytosis. It has been reported that S/T residues located in the second (T134, T144, S147, S148) and the third (T225) intracellular loops of D2R are involved in agonist-induced D2R endocytosis (Cho et al., 2010) (Figure 1). Another study has shown that some other S/T residues located within the third intracellular loop of D2R (S256, S257, T258, S259, T264, S282, S288, S292) are also phosphorylated in a GRK-mediated manner and that they are required for the agonist-induced (GRK-mediated) recycling of endocytosed D2R [53] (Table 1, Figure 1). It is interesting that different sets of S/T residues mediate distinct intracellular trafficking processes. The resensitization of D2R occurs when the receptor dissociates from arrestin via agonist-induced endocytosis [51].
In addition to GRK/arrestin-mediated regulation, D2R and D3R are also regulated by PKC. Robust phosphorylation of D2R is induced by PMA (phorbol 12-myristate 13-acetate) but the resulting endocytosis is weaker than the agonist-induced endocytosis of D2R [43,51]. The phosphorylation, endocytosis, and desensitization of D3R induced by PMA are more robust than those of D2R [43,51,56]. It has been reported that treatment with PMA results in the ubiquitination of D3R and its subsequent degradation through the lysosomal pathway. The degradation of D3R by PKC is dependent on clathrin-mediated endocytosis. However, this process is distinct from the desensitization of D3R and is not affected by its inhibition [59]. The D2R residues T225, S228, S229, T322, T324, and S325 are responsible for PMA-induced receptor phosphorylation and endocytosis [56,67] (Table 1, Figure 1). PKC-mediated phosphorylation and endocytosis of D3R are mediated by the S229 and S257 residues [43] (Table 1, Figure 2).
Even though D3R does not undergo typical endocytosis, which involves the movement of receptors from the plasma membrane to the cytosol, it undergoes another type of intracellular trafficking called pharmacological sequestration [52,68]. Upon desensitization of D3R induced by an adequate amount of agonist, it translocates to the hydrophobic region of the membrane where the hydrophilic agonist has reduced binding efficiency. The time course of D3R desensitization and pharmacological sequestration are similar, and both depend on the presence of Gβγ and arrestin, indicating that the phenomenon of pharmacological sequestration may account for D3R desensitization.
The movement of GFP-tagged receptors from the plasma membrane to the cytosol upon agonist treatment is a convenient way to confirm the endocytosis of receptor proteins. Alternatively, biochemical assays such as fluorescence-activated single cell sorting (FACS) or enzyme-linked immunosorbent assay (ELISA) can be employed using constructs with HA or FLAG epitopes attached to the N-terminus of the receptor. Hydrophilic radioligands can also be used to assess receptor endocytosis as they cannot bind receptors in the cytoplasm. However, caution should be exercised when dealing with receptors that have high affinity for the agonist, such as DA and D3R. In such cases, it is difficult to completely wash away the agonist from cells expressing D3R unless the cells are washed under harsh conditions, such as with a low pH buffer. If cells are treated with a radioligand under conditions where a pre-applied agonist is not completely washed, it may give a false impression that receptor endocytosis is occurring (approximately 30%) because the radioligand cannot properly bind to the receptor that is already occupied by the tightly bound agonist.

2.3. Roles of Small G Proteins in the Signaling and Endocytosis of D2R and D3R

There are two families of GTP binding proteins: heterotrimeric large G proteins composed of three subunits (α, β, and γ) and small G proteins with a single subunit. Small G proteins mediate more versatile functions than trimeric G proteins due to their larger number in the cell [69,70]. Small GTPases can be categorized into five main families, namely Ras, Rho, Rab, ADP-ribosylation factor (Arf), and Ran, based on both sequence and functional criteria [70].
Small GTPases, including Rab and ARF proteins, as well as the large GTPase dynamin, play regulatory roles in vesicular transport [71]. Several Rab proteins, including Rab5 and Rab23, are involved in the regulation of membrane trafficking and recycling of cell surface proteins from the Golgi to the plasma membrane [72,73].
Whereas RalA and ARF6 small G proteins are known to regulate D2R and D3R [67,74,75], the involvement of Rab5 and Rab23 in the endocytosis of D2R remains controversial [67,76].

2.3.1. RalA

RalA belongs to the Ras family of small G proteins. The GTP-bound form (G23V, active) of RalA is known to inhibit the endocytosis of D2R in a manner that is independent of the previously reported downstream effectors of RalA, such as Ral-binding protein 1 and phospholipase D [74]. The endocytosis of GPCRs, including D2R, can be inhibited by active RalA through its high affinity for GRK2, which sequesters GRK2 away from the GPCRs. RalA is converted from its active to inactive state by the translocation of Ral GDP dissociation stimulator-like (RGL), a guanine nucleotide exchange factor, from the plasma membrane to the cytosol in a complex with Gβγ. Thus, the agonist-induced Gβγ-mediated conversion of RalA from its GTP-bound to GDP-bound form is thought to trigger and facilitate the endocytosis of GPCRs induced by agonist stimulation [74].
Filamin A (FLNA), an actin-binding protein, functions as a signaling and intracellular trafficking scaffold for various GPCRs, including D2R and D3R. FLNA plays an essential role in maintaining the proper expression of D2R on the plasma membrane [77] and is also necessary for signaling through both D2R and D3R [78,79,80]. FLNA is known to constitutively inhibit RalA activity, thus ensuring proper receptor trafficking and signaling. In the absence of FLNA, RalA is activated and sequesters GRK2 from the receptor, resulting in desensitization of D2R due to its inhibition of recycling [51,53]. On the other hand, active RalA inhibits the signaling of D3R, which does not undergo agonist-induced endocytosis, in an arrestin-dependent manner. Thus, it is suggested that active RalA modulates receptor signaling differently depending on whether GRK2 or arrestin is involved in the functional regulation of the respective receptor.

2.3.2. ARF6

The ARF (ADP-ribosylation factor) family of small GTPases comprises six members, among which ARF6 is the most extensively studied with regards to its role in intracellular trafficking of membrane proteins. ARF6 is known to accumulate in clathrin-coated pits (CCPs) in a GTP-dependent manner and facilitate the rapid recycling of plasma membrane receptors [81,82]. Furthermore, research suggests that the activation of ARF6 via arrestin-mediated mechanisms is involved in the endocytosis of β2AR [83].
According to a report, the endocytosis of D2R is not influenced by ARF6, as neither constitutively active nor a dominant-negative mutant of ARF6 has any effect [67]. However, the recycling of endocytosed D2R is inhibited by a constitutively GTP-bound mutant of ARF6 (Q67L), indicating that the conversion of ARF6 from the GTP-bound to GDP-bound form is crucial for the proper recycling of endocytosed vesicles.

2.4. Roles of Regulators of G Protein Signaling (RGS) in the Signaling of D2R and D3R

In its inactive state, the Gα subunit is bound to GDP, whereas activation of the GPCR by an agonist triggers the exchange of GDP for GTP on the Gα subunit, resulting in its dissociation from the Gβγ heterodimer [84]. Both the Gα and Gβγ subunits are capable of regulating downstream effector proteins. The Gα subunit possesses intrinsic GTPase activity, which converts the GTP-bound Gα to its inactive form, Gα-GDP, allowing it to re-associate with the Gβγ heterodimers and effectively terminate signaling via both Gα-GTP and Gβγ.
The rapid physiological timing of GPCR signal transduction observed in vivo contrasts with the slow rates of GTP hydrolysis exhibited by purified Gα subunits in vitro. This discrepancy has been explained by the existence of the GTPase-accelerating protein (GAP) family [85,86]. Unlike the five main families of GAPs that regulate the Ras superfamily, regulators of G protein signaling (RGSs) specifically target heterotrimeric G proteins [87]. RGS proteins bind to and regulate the Gαi and Gαq subfamilies of proteins by increasing their GTPase activity. However, the Gαs subfamily already exhibits a sufficiently high intrinsic GTPase activity and is not sensitive to RGS protein-mediated GAP activity.
RGS2, RGS4, and RGS9-2 are RGS subtypes that have been associated with the dopaminergic nervous system. In a previous study utilizing in situ hybridization, it was suggested that D1R and D2R co-localize with RGS2 and RGS4, respectively [88]. It has also been shown that D1R and D2R regulate the expression of RGS2 and RGS4 in opposite directions. For instance, quinpirole, a D2-like receptor agonist, induces the downregulation of RGS2 and upregulation of RGS4 [89]. Interestingly, male rats require RGS4, but not RGS2, for the rewarding effects of cocaine [90]. Further studies have demonstrated that the N-terminal domain of RGS4 interacts with D2R and D3R and inhibits receptor signaling via the RGS domain [57].
RGS9-2 exhibits high levels of expression in the striatum and nucleus accumbens, regions that also express D2-like receptors. Studies suggest that RGS9-2 plays a functional role in regulating D2-like receptors. Specifically, introducing RGS9-2 through viral expression in brain regions such as the nucleus accumbens or introducing RGS9-2 proteins into striatal cholinergic interneurons reduces the behavioral or electrophysiological response to D2-like receptor stimulation [91,92,93]. In contrast, when RGS9-2 is knocked out, there is an augmentation in the behavioral response to D2-like receptor activation [91,93].
RGS9-2 is a member of the R7 RGS subfamily and possesses three domains: RGS, disheveled-EGL10-pleckstrin (DEP) homology, and G-gamma-like (GGL). The RGS domain binds to the Gα subunit and facilitates GAP activity. The DEP domain anchors the protein to the membrane and allows for interaction with the C-terminal tail of GPCRs. Finally, the GGL domain ensures protein stability by dimerizing with Gβ5 [94,95,96,97].
In HEK-293 cells or C6 glioma cells, the expression of RGS9-2 specifically hinders D3R signaling while leaving D2R/D4R signaling unaffected. This can be attributed to the varying affinities of the receptors for arrestin3, which facilitates the creation of a complex consisting of RGS9’s DEP domain, Gβ5, R7-binding protein (R7BP), and D3R [98].

3. Biased Signaling

GPCRs carry out their functions through various signaling pathways, which can be categorized into G protein-dependent and G protein-independent pathways (Wisler et al., 2018). The G protein-independent pathways mainly involve arrestins, which were previously believed to be involved in agonist-triggered receptor desensitization and endocytosis [61,99,100]. Unlike conventional balanced agonists that stimulate both G protein and arrestin pathways, newly synthesized ligands have been reported to selectively act on one of the two pathways [101,102,103]. It is speculated that biased agonists stabilize GPCRs in a specific conformation [104,105].
There has been great interest in the development of biased ligands because they are proposed to be superior to traditional balanced agonists in terms of efficacy and adverse effects in the treatment of certain disorders [106,107]. Usually, biased ligands have been designed based on data obtained through structure–activity analyses of currently available ligands [35,108]. Biased signaling can occur not only in response to tailored ligands but also through genetically modified GPCRs, which can provide new insights for designing novel biased ligands. Several biased D2Rs have been reported, and some of them are discussed below (Figure 1, Table 2).
Structure–activity analyses of D2R agonists revealed that a hydrophobic pocket at the interface of the second extracellular loop and fifth transmembrane segment of D2R is involved in biased signaling [108]. One of the point mutations examined in D2R, specifically the mutation at F189, which is equivalent to F5.38 according to Ballesteros–Weinstein GPCR numbering system [112], showed a preference for G protein coupling. When tested for G protein signaling, the DA potency for F189A-D2R decreased about 500 times compared with that for WT-D2R; however, the interaction with arrestin2 was almost completely abolished, resulting in impaired arrestin recruitment while maintaining G protein signaling to some extent.
The Asp–Arg–Tyr (DRY) motif is located at the beginning of the second intracellular loop. The Arg residue within this motif plays vital roles in G protein coupling, receptor phosphorylation, and arrestin recruitment by forming intramolecular interactions [113,114]. To create a biased D2R, a point mutation was introduced at R132 in the DRY motif (R3.50), so that both G protein coupling and arrestin translocation were suppressed. An additional mutation was then introduced at L3.41 (L123W), which determines the thermostability [115], to increase the potency for DA [109]. The resulting D2R mutant (R132L/L123W) produced displayed biased signaling through G proteins. This is because the L123W mutation, in addition to R132L, significantly restored the signaling ability of R132L-D2R through G proteins while having a lesser impact on its ability to recruit arrestins.
Research also revealed that modifying four amino acid residues within the N-terminal region of the third intracellular loop (212IYIV215) resulted in a signaling-biased receptor. The receptor maintained its ability to bind ligands and couple with and activate G proteins but demonstrated impaired ability to mediate arrestin3 translocation to the plasma membrane [110].
Using the evolutionary trace (ET) method [116], another study aimed to develop a biased D2R. Peterson et al. successfully identified D2R mutants that exhibited selective signaling through either G protein or arrestins [111,117]. The team discovered two D2R mutants with dual amino acid substitutions, L125N/Y133L and A135R/M140D, which showed a preference for G protein-mediated and arrestin-dependent pathways, respectively. These modified D2R mutants were then used for in vivo studies to investigate the roles of the arrestin pathway in amphetamine-induced locomotion potentiation [117].
Identifying the sites that determine biased signaling in D3R might not be straightforward, as agonist treatment rarely induces arrestin recruitment [30]. To measure arrestin-dependent signaling, the carboxyl terminus of the vasopressin type 2 receptor is typically attached to the receptor for stable arrestin translocation [118]. In the case of D2R, this approach leads to a complementary interaction with the DRY motif, whereas in the case of D3R it causes a shift in the arrestin translocation pattern from type A to type B [113,114]. In type A GPCRs, arrestin is released from the receptor as endocytosis progresses, whereas in type B GPCRs arrestin remains bound to the receptor and accompanies it in the endocytic vesicle [65].
A previous study has demonstrated that a G protein-biased ligand interacts with Asp110 on transmembrane 3 and His349 on transmembrane 6, but it is unclear whether these amino acids are the decisive factors for the biased signaling of D3R [119]. Although the involvement of the arrestin-dependent pathway has been suggested in D3R signaling, identifying the amino acid residues responsible for biased signaling and creating new biased agonists will require both conceptual and methodological advancements [120].
A previous study showed that the knockdown of arrestins does not affect D2R-mediated ERK activation [121], which is contradictory to recent publications indicating the existence of biased mutant D2Rs and ligands [35,108,109,110,111,122,123].
There are several potential reasons for the discrepancy observed in the effects of arrestin depletion on ERK activation. One possible explanation is that the G protein and arrestin pathways are complementary for ERK activation, meaning that the remaining pathway can compensate for the absence of the other. Another possibility is the dual role of arrestins in mediating both their own pathway, such as ERK activation, and desensitizing the G protein pathway. Therefore, removing endogenous arrestin may block arrestin-dependent ERK activation but enhance G protein-dependent ERK activation. Another consideration is the technical limitations of siRNA- or shRNA-mediated knockdown, which may not completely remove the target protein. If a small amount of arrestin is sufficient to support the arrestin signaling pathway, ERK activation may not be affected even with significant arrestin depletion.
The identification of the factors that determine ligand binding and understanding the 3D structure of D2R and D3R can greatly aid in developing selective or biased ligands for these receptors. Specifically, the structural analyses of D2R or D3R in complex with G protein or arrestin can have significant implications for developing biased ligands. Recent research has revealed the crystal structures of D2R bound to risperidone and D3R bound to eticlopride [124,125], as well as the 3D structures of D2R or D3R in complex with Gi through cryo-electron microscopy analysis [126,127]. Additional studies linking computational techniques with experimental results have identified factors involved in ligand binding, further advancing structure-based drug design [128,129]. Together, these findings enhance our understanding of the structural and functional properties of D2R and D3R and underscore the potential of structure-based drug design to develop novel therapeutics targeting these receptors.

4. Dimerization of Dopamine D2 and D3 Receptors

GPCRs have the ability to form dimers or oligomers with similar or different types of GPCRs. Coactivation of D1R and D2R, which couple to Gs and Gi/o, respectively, has been demonstrated to activate Gq protein, leading to intracellular calcium release through the activation of phospholipase C and IP3 [130,131,132].
The D1R and D2R interact through adjacent glutamic acids in the carboxyl tail of D1R (404E/405E) and two adjacent arginine residues in the third intracellular loop of D2R (245R/246R) [133]. Previously, the third intracellular loop of D2R, specifically the region 217RRRRKR222, was identified as a potential site for heteromer interaction, forming heteromers with D1R [134], 5-HT2AR [135], or adenosine A2A receptors [136].
Studies have demonstrated both direct and functional interactions between D1R and D3R. Depending on the cell type or signaling, the two receptor subtypes may affect neurons in either a synergistic or opposing manner [137,138,139]. Furthermore, D3R stimulation has been shown to enhance D1R agonist affinity, which potentiates D1R-mediated behavioral effects [140,141].
Dimerization between these two receptors has also been observed to alter endocytic properties. Heterodimerization with D3R, for instance, abolishes agonist-induced endocytosis of D1R but allows the endocytosis of the D1R/D3R complex in response to simultaneous agonistic stimulation of both receptors [141]. It is noteworthy that G protein-independent signaling can occur with stimulation of D1R/D3R heteromers in the nucleus accumbens [142].
There have been several studies indicating the functional interactions between D3R and nAChR in vivo. For instance, D3R and α4β2 nAChR form heteromers in dopaminergic neurons that are crucial for the neurotrophic effects of nicotine [143]. Furthermore, the development of nicotine sensitization is accompanied by an increase in the expression of D3R [144,145], and it has been suggested that D3R ligands can be used to treat tobacco dependence [146]. A recent in vitro study showed that nicotinic stimulation of α4β2 nAChR leads to Src activation in an arrestin2- and 14-3-3η-dependent manner. The activated Src phosphorylates the tyrosine residue(s) on Syk molecules that then interact with phospholipase Cγ1 to trigger the translocation of PKCβII to the cell membrane by elevating cellular diacylglycerol levels [147].

5. Posttranslational Modifications

Posttranslational modifications (PTMs) are chemical modifications of amino acid side chains that occur after protein synthesis is complete. These modifications usually take place in the endoplasmic reticulum (ER) and the Golgi apparatus [148]. In the case of GPCRs, PTMs can occur in all regions of the receptor except for the transmembrane domains. These modifications are important for regulating receptor folding, maturation, trafficking, and signaling, which increases functional diversity and fine-tunes signaling pathways [149,150]. The most extensively characterized PTMs include glycosylation, phosphorylation, ubiquitination, and palmitoylation [150,151]. The sites and functional roles of PTMs of D2R and D3R are summarized in Table 1 and Figure 3.

5.1. Glycosylation

Glycosylation takes place prevalently at the N-termini or extracellular loops of GPCRs. N-glycosylation, the major form of glycosylation in GPCRs, links a sugar molecule to the nitrogen of the Asn (N) residue in the consensus motif N-X-S/T (X ≠ P) [152,153]. Detection of N-glycosylation in target proteins has traditionally involved enzymatic cleavage with peptide N-glycosidase F (PNGase F) or endoglycosidase H (Endo H) along with site-directed mutagenesis in the consensus sequence.
The process of glycosylation commences with dolichol, which is a polymerized isoprene molecule. In the ER membrane, dolichol, with its hydroxyl group facing the cytoplasmic side, is converted to dolichol phosphate [154]. Subsequently, N-acetylglucosamine (NAG)-UDP and β-D-mannopyranose are added in succession to dolichol phosphate. Dolichol then flips so that the entire glycosyl moiety faces the ER lumen and additional mannose (M) moieties along with glucose (G) molecules are added to form a large complex of dolichol diphosphate–NAG-NAG-M/Gs, which is now ready for the glycosylation of the target protein. Glycosylation takes place when the NAG-NAG-M/G moiety, detached from dolichol diphosphate, interacts with the nitrogen atom on the side chain of the asparagine residue in the target protein.
Glycosylation is involved in the modulation of various aspects of receptor function, such as maturation, trafficking, ligand binding, and cell signaling, as reported in several studies [155,156,157]. The D2R possesses three N-linked glycosylation sites at its N-terminus (N5, N17, N23), whereas the D3R has four N-linked glycosylation sites. Two of these sites are located at the N-terminus (N12, N19), whereas the other two are present in the second and third extracellular loops (N97 and N173, respectively).
Surface expression of both D2R and D3R is commonly affected by glycosylation at the N-terminus, with D2R undergoing caveolae endocytosis and D3R undergoing clathrin-mediated endocytosis. In the case of D3R, glycosylation at the N-terminus is necessary for basal signaling and desensitization, whereas glycosylation at N97 and N173 in the second and third extracellular domains, respectively, is involved in receptor endocytosis [57,58,158].

5.2. Phosphorylation

Following agonist binding, GPCRs undergo conformational changes and interact with a specific heterotrimeric G protein, which enhances its GTPase activity, resulting in the separation of the G protein into α and βγ subunits that subsequently modulate downstream signaling pathways [84]. Concurrently, regulatory processes are initiated following receptor activation, with receptor phosphorylation playing a crucial role [159,160].
The intracellular regions of GPCRs contain numerous phosphorylation sites that are predominantly targeted by two classes of serine/threonine kinases: GPCR kinases (GRKs) and second-messenger-dependent protein kinases, such as protein kinase A (PKA) and protein kinase C (PKC) [40,42,43,161]. Techniques such as mass spectrometry, phosphor-specific immunoblotting, and site-directed mutagenesis of the consensus sites in the target proteins have been utilized to identify GPCR phosphorylation sites [162,163].
Protein kinases, including PKA and PKC, recognize specific substrate sequences independent of receptor activation [49,164]. Furthermore, the activation of these kinases is reliant on second messengers that can spread throughout the cell [42,43,165]. As a result, it is likely that both occupied and unoccupied neighboring receptors are phosphorylated by these kinases.
GRKs have a preference for phosphorylating GPCRs when the receptors are in an active state, which means that they are occupied by agonists [166]. This selectivity is due to the dynamic interactions between GRKs and GPCRs, as both proteins undergo conformational changes upon agonist stimulation of GPCRs [167,168]. The interaction and activation of GRKs are mainly mediated by the overall topologic structure of the activated receptor rather than the amino acid sequence surrounding the phosphorylation site [167,169,170,171]. Activated receptors also enhance the enzymatic activity of GRKs, making them both a substrate and activator of GRKs [167,170,172].
The intracellular regions of D2R and D3R contain a number of serine and threonine residues. The short isoform of D2R (D2SR) has three threonine residues located in the first intracellular loop, two serine and two threonine residues in the second loop, and a total of thirteen serine and ten threonine residues in the third loop. On the other hand, D3R includes three threonine residues in the first intracellular loop, two serine and two threonine residues in the second loop, ten serine and seven threonine residues in the third loop, and an additional serine residue in the carboxyl terminal region.
Phosphorylation of D2R is induced by agonists at various serine and threonine residues, including S256, S257, T258, S259, T264, S282, S288, and S292 [53]. PKC-mediated phosphorylation of D2SR, which is inferred from D2LR results, mainly occurs at residues 228, 229, and 325, with additional contributions of threonine residues at 322 and 324 [56].
D3R is phosphorylated at S229 and S259 upon PMA stimulation, and these phosphorylations play a significant role in regulating desensitization and intracellular trafficking of the receptor [43].

5.3. Ubiquitination

Ubiquitin is a polypeptide consisting of 76 amino acids that is found in eukaryotes. The process of ubiquitylation involves the formation of an isopeptide bond that connects an internal lysine residue of the target protein with the C-terminal glycine (glycine 76) of ubiquitin (Hershko, 2005). Ubiquitination requires three sets of enzymes: Ub-activating (E1), Ub-conjugating (E2), and Ub-ligating enzymes (E3) [173]. E1 and E2 sequentially form thioester bonds with ubiquitin, which involves condensation and conjugation between the thiol groups (usually of cysteine) of E1 and E2 and the carboxyl group of Ub (glycine 76). E2 then associates with an E3, through which ubiquitin is transferred from E2 to E3. Finally, E3 catalyzes the conjugation between the amino group of a lysine residue on the target substrate and the carboxyl group of Ub via an isopeptide bond. E3 plays a crucial role in ubiquitination by conferring substrate specificity [174].
There are three different ways in which substrate proteins can be attached to Ub, namely monoubiquitination, multi-monoubiquitination, and polyubiquitination. To form polyubiquitin chains, eight residues in ubiquitin can be employed, including K6, K11, K27, K29, K33, K48, K63, and M1 (linear). Depending on the type of linkage, the modified proteins are directed towards diverse cellular outcomes [175,176].
Regulating biosynthesis, endocytosis, lysosomal sorting, degradation, and cellular signaling of GPCRs heavily rely on ubiquitination [150,177]. Typically, intracellular lysine residues can undergo either monoubiquitination or polyubiquitination in response to agonist-dependent or agonist-independent stimuli [178,179].
Although PMA treatment has been found to induce ubiquitination and lysosomal degradation of D3R [59], there is no evidence of agonist-induced ubiquitination of D2R and D3R.

5.4. Palmitoylation

The initial step of palmitoylation involves creating a thioester linkage between the carboxyl group of palmitic acid (hexadecanoic acid, C16) and the thiol group of coenzyme A (CoA). S-palmitoylation, which is more frequent than N-palmitoylation, involves the thiol group of a serine residue in the target protein attacking the carbonyl group of the thioester linkage formed in the first step [180]. This results in the formation of a fresh thioester linkage between the target protein and palmitic acid. Meanwhile, CoA is released and can be reused. The precise consensus sequences surrounding the cysteinyl residues that undergo palmitoylation have yet to be determined [181].
When methionine, the first amino acid of a polypeptide, is removed, N-palmitoylation occurs when the first amino acid is either cysteine, glycine, or lysine [182]. Similar to S-palmitoylation, the free amino group of these amino acids in the target protein reacts with the carbonyl group of the thioester linkage established in the first step. As a result, a new amide linkage is formed between the target protein and palmitic acid.
S-palmitoylation can occur spontaneously or can be catalyzed by multi-span transmembrane integral proteins known as protein acyl transferases (PATs). PATs possess zinc-finger and aspartate–histidine–histidine–cysteine (zDHHC) domains [183]. The intracellular DHHC motif plays a crucial role in the S-acylating activity of PATs, allowing protein palmitoylation to occur at the cytoplasmic face of membranes in the secretory pathway, including the ER and Golgi apparatus as well as the plasma membrane.
Palmitoylation is a reversible process, and depalmitoylation is mediated by protein thioesterases or depalmitoylases. Currently, three classes of depalmitoylases have been identified; acyl-protein thioesterases (APTs), α/β hydrolase domain-containing 17 proteins (ABHD17s), and palmitoyl-protein thioesterases (PPTs) [184]. APTs located in the cytosol are responsible for depalmitoylating targets of Gα and Ras [185,186]. In contrast, PPTs found in the lysosome play a role in regulating target protein degradation [187].
During the biosynthesis of GPCRs, palmitoylation often occurs at their C-termini and sometimes at intracellular loops. This modification is usually a basal process but can also be induced by agonist stimulation in some cases [188]. Palmitoylation is essential for various functions of GPCRs, including trafficking, cell surface localization, dimerization, and signaling [188,189].
D2R and D3R share a conserved cysteine residue at their carboxyl termini and have a highly similar sequence for the last 10 amino acid residues except for the second-to-last residue (histidine in D2R and serine in D3R). Despite this similarity, D3R is constitutively palmitoylated at C400, whereas D2SR is not palmitoylated in HEK-293 cells [31]. Palmitoylation is a critical modification for proper localization of D3R to the cell surface, as well as for PKC-mediated endocytosis, agonist affinity, and desensitization of the receptor [31].
Studies have shown that the palmitoylation of D2R may vary depending on the expression system. D2SR was not palmitoylated in HEK-293 cells [31], whereas other studies have reported that both D2SR and D2LR are palmitoylated when expressed in the baculovirus/Sf9 cell system [190,191]. Moreover, conflicting results have been reported on the palmitoylation site of D2R, with one study suggesting C443 as the palmitoylation site [192] and another study reporting no palmitoylation at this site [31]. These discrepancies may have resulted from differences in experimental strategies and conditions, such as differences in the concentration and duration of 2-bromopalmitate treatment and the use of deletion or point mutation strategies to disrupt palmitoylation.
Taking into account the variations in experimental conditions, it is probable that only D3R (not D2R) undergoes palmitoylation at the C-terminus. Furthermore, although palmitoylation may occur in D2R, it appears that D3R is more heavily palmitoylated. Additionally, palmitoylation seems to have a more crucial impact on the regulation of D3R’s functions than those of D2R.
Although the consensus sequence for palmitoylation is not yet fully established, a comparison between D2R and D3R yielded some insight. Specifically, the presence of a Ser residue preceding the Cys residue appears to be crucial for palmitoylation of D3R. For instance, when the C-terminal “HC” sequence of D2R, which is typically not palmitoylated, is modified to “SC” as in D3R, D2R becomes palmitoylated, and the reverse is also true [31].

6. Conclusions

Establishing a correlation between the in vivo roles of certain GPCRs and their biochemical characteristics in vitro is challenging. Both D2R and D3R are involved in regulating locomotion, reward, and addiction, but D3Rs appear to have a more specific role in regulating emotional and cognitive processes that involve behavioral sensitization [37]. Evidence suggests that behavioral sensitization is closely related to the plasticity of limbic neurochemical systems [193,194], which are also implicated in a range of psychiatric and substance use disorders.
The roles of D3R in the development of behavioral sensitization have been proposed based on various observations [195]. For instance, D3R antagonists can block behavioral sensitization, and both transcriptional and translational activities involved in D3R expression are reduced following behavioral sensitization [144,196].
Previous reports have demonstrated that D3R has inhibitory effects on the synthesis and release of DA in the striatum [197,198,199]. D3R also appears to mediate the inhibition of neuronal firing [200,201], and selective D3R agonists can lower extracellular levels of DA, leading to the inhibition of locomotion [202,203].
Although the signaling pathways of D2R and D3R largely overlap, the intensity of D2R signaling is stronger than that of D3R, except in the inhibition of dopamine release [22,204,205]. Therefore, the development of the desensitization of D3R after repeated neuronal stimulation may result in the loss of a D3R-mediated ‘brake’ on dopamine release, leading to increases in locomotion and behavioral sensitization [195,206].
Posttranslational modifications play critical roles in maintaining the desensitizing properties of D2R and D3R. Among the regulatory processes, such as agonist-induced receptor phosphorylation, arrestin translocation, and receptor endocytosis, only those related to intracellular trafficking occur more strongly with D2R than with D3R. Considering that receptor endocytosis is necessary for the resensitization of desensitized receptors, D3R is likely more susceptible to desensitization than D2R. In contrast, glycosylation and palmitoylation of D3R are necessary for maintaining its capacity to undergo desensitization.
Overall, D3R is predominantly expressed in limbic brain regions that are associated with various psychiatric and substance use disorders, where the plasticity of limbic neurochemical systems plays a significant role. In vitro studies have demonstrated that D3R undergoes dynamic desensitization that necessitates specific PTMs. As D3R appears to function as a negative regulator for various behavioral aspects that are linked to the dopaminergic nervous system, it is justifiable to propose a theoretical model of behavioral sensitization that views it as the desensitization of inhibitory systems, where D3R has a vital function.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (KRF-2020R1I1A3062151).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, T.; Seeman, P.; Rajput, A.; Farley, I.J.; Hornykiewicz, O. Receptor basis for dopaminergic supersensitivity in Parkinson’s disease. Nature 1978, 273, 59–61. [Google Scholar] [CrossRef]
  2. Seeman, P. Dopamine receptors and the dopamine hypothesis of schizophrenia. Synapse 1987, 1, 133–152. [Google Scholar] [CrossRef] [PubMed]
  3. Cunnah, D.; Besser, M. Management of prolactinomas. Clin. Endocrinol. 1991, 34, 231–235. [Google Scholar] [CrossRef] [PubMed]
  4. Brown, J.H.; Makman, M.H. Stimulation by dopamine of adenylate cyclase in retinal homogenates and of adeno-sine-3′:5′-cyclic monophosphate formation in intact retina. Proc. Natl. Acad. Sci. USA 1972, 69, 539–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. De Camilli, P.; Macconi, D.; Spada, A. Dopamine inhibits adenylate cyclase in human prolactin-secreting pituitary adenomas. Nature 1979, 278, 252–254. [Google Scholar] [CrossRef]
  6. Kebabian, J.W.; Calne, D.B. Multiple receptors for dopamine. Nature 1979, 277, 93–96. [Google Scholar] [CrossRef]
  7. Dearry, A.; Gingrich, J.A.; Falardeau, P.; Bates, M.D.; Caron, M.G. Molecular cloning and expression of the gene for a human D1 dopamine receptor. Nature 1990, 347, 72–76. [Google Scholar] [CrossRef]
  8. Sunahara, R.K.; Niznik, H.B.; Weiner, D.M.; Stormann, T.M.; Brann, M.R.; Kennedy, J.L.; Gelernter, J.E.; Rozmahel, R.; Yang, Y.; Israel, Y.; et al. Human dopamine D1 receptor encoded by an intronless gene on chromosome 5. Nature 1990, 347, 80–83. [Google Scholar] [CrossRef]
  9. Sunahara, R.K.; Guan, H.-C.; O’Dowd, B.F.; Seeman, P.; Laurier, L.G.; Ng, G.; George, S.R.; Torchia, J.; Van Tol, H.H.M.; Niznik, H.B. Cloning of the gene for a human dopamine D5 receptor with higher affinity for dopamine than D1. Nature 1991, 350, 614–619. [Google Scholar] [CrossRef]
  10. Bunzow, J.R.; Van Tol, H.H.M.; Grandy, D.K.; Albert, P.; Salon, J.; Christie, M.; Machida, C.A.; Neve, K.A.; Civelli, O. Cloning and expression of a rat D2 dopamine receptor cDNA. Nature 1988, 336, 783–787. [Google Scholar] [CrossRef]
  11. Sokoloff, P.; Giros, B.; Martres, M.P.; Bouthenet, M.L.; Schwartz, J.C. Molecular cloning and characterization of a novel dopamine receptor (D3) as a target for neuroleptics. Nature 1990, 347, 146–151. [Google Scholar] [CrossRef]
  12. Van Tol, H.H.M.; Bunzow, J.R.; Guan, H.-C.; Sunahara, R.K.; Seeman, P.; Niznik, H.B.; Civelli, O. Cloning of the gene for a human dopamine D4 receptor with high affinity for the antipsychotic clozapine. Nature 1991, 350, 610–614. [Google Scholar] [CrossRef] [PubMed]
  13. Giros, B.; Martres, M.P.; Sokoloff, P.; Schwartz, J.C. Gene cloning of human dopaminergic D3 receptor and identification of its chromosome. C R Acad. Sci. III 1990, 311, 501–508. [Google Scholar] [PubMed]
  14. Chio, C.L.; Lajiness, E.M.; Huff, R.M. Activation of heterologously expressed D3 dopamine receptors: Comparison with D2 dopamine receptors. Mol. Pharmacol. 1994, 45, 51–60. [Google Scholar] [PubMed]
  15. Freedman, S.B.; Patel, S.; Marwood, R.; Emms, F.; Seabrook, G.R.; Knowles, M.R.; McAllister, G. Expression and pharmaco-logical characterization of the human D3 dopamine receptor. J. Pharmacol. Exp. Ther. 1994, 268, 417–426. [Google Scholar]
  16. Robinson, S.W.; Caron, M.G. Selective Inhibition of Adenylyl Cyclase Type V by the Dopamine D3Receptor. Mol. Pharmacol. 1997, 52, 508–514. [Google Scholar] [CrossRef]
  17. Ganz, M.B.; Pachter, A.J.; Barber, D.L. Multiple receptors coupled to adenylate cyclase regulate Na-H exchange independent of cAMP. J. Biol. Chem. 1990, 265, 8989–8992. [Google Scholar] [CrossRef]
  18. Neve, K.A.; Kozlowski, M.R.; Rosser, M.P. Dopamine D2 receptor stimulation of Na+/H+ exchange assessed by quantifica-tion of extracellular acidification. J. Biol. Chem. 1992, 267, 25748–25753. [Google Scholar] [CrossRef]
  19. Pilon, C.; Levesque, D.; Dimitriadou, V.; Griffon, N.; Martres, M.P.; Schwartz, J.C.; Sokoloff, P. Functional coupling of the hu-man dopamine D3 receptor in a transfected NG 108-15 neuroblastoma-glioma hybrid cell line. Eur. J. Pharmacol. 1994, 268, 129–139. [Google Scholar] [CrossRef]
  20. Welsh, G.I.; Hall, D.A.; Warnes, A.; Strange, P.G.; Proud, C.G. Activation of microtubule-associated protein kinase (Erk) and p70 S6 kinase by D2 dopamine receptors. J. Neurochem. 2002, 70, 2139–2146. [Google Scholar] [CrossRef] [Green Version]
  21. Cussac, D.; Newman-Tancredi, A.; Pasteau, V.; Millan, M.J. Human Dopamine D3Receptors Mediate Mitogen-Activated Protein Kinase Activation via a Phosphatidylinositol 3-Kinase and an Atypical Protein Kinase C-Dependent Mechanism. Mol. Pharmacol. 1999, 56, 1025–1030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Tang, L.; Todd, R.D.; O’Malley, K.L. Dopamine D2 and D3 receptors inhibit dopamine release. Experiment 1994, 270, 475–479. [Google Scholar]
  23. Werner, P.I.A.; Hussy, N.; Buell, G.; Jones, K.A.; North, R.A. D2, D3, and D4 dopamine receptors couple to G protein-regulated potassium channels in Xenopus oocytes. Mol. Pharmacol. 1996, 49, 656–661. [Google Scholar] [PubMed]
  24. Kuzhikandathil, E.V.; Yu, W.; Oxford, G.S. Human dopamine D3 and D2L receptors couple to inward rectifier potassium channels in mammalian cell lines. Mol. Cell. Neurosci. 1998, 12, 390–402. [Google Scholar] [CrossRef]
  25. Bouthenet, M.-L.; Souil, E.; Martres, M.-P.; Sokoloff, P.; Giros, B.; Schwartz, J.-C. Localization of dopamine D3 receptor mRNA in the rat brain using in situ hybridization histochemistry: Comparison with dopamine D2 receptor mRNA. Brain Res. 1991, 564, 203–219. [Google Scholar] [CrossRef]
  26. Robinson, S.W.; Jarvie, K.R.; Caron, M.G. High affinity agonist binding to the dopamine D3 receptor: Chimeric receptors delineate a role for intracellular domains. Mol. Pharmacol. 1994, 46, 352–356. [Google Scholar]
  27. Joyce, J.N. Dopamine D3 receptor as a therapeutic target for antipsychotic and antiparkinsonian drugs. Pharmacol. Ther. 2001, 90, 231–259. [Google Scholar] [CrossRef]
  28. Joyce, J.N.; Millan, M.J. Dopamine D3 receptor antagonists as therapeutic agents. Drug Discov. Today 2005, 10, 917–925. [Google Scholar] [CrossRef]
  29. Newman, A.H.; Grundt, P.; Nader, M.A. Dopamine D3 Receptor Partial Agonists and Antagonists as Potential Drug Abuse Therapeutic Agents. J. Med. Chem. 2005, 48, 3663–3679. [Google Scholar] [CrossRef]
  30. Kim, K.M.; Valenzano, K.J.; Robinson, S.R.; Yao, W.D.; Barak, L.S.; Caron, M.G. Differential regulation of the dopamine D2 and D3 receptors by G protein-coupled receptor kinases and beta-arrestins. J. Biol. Chem. 2001, 276, 37409–37414. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, X.; Le, H.T.; Zhang, X.; Zheng, M.; Choi, B.-G.; Kim, K.-M. Palmitoylation on the carboxyl terminus tail is required for the selective regulation of dopamine D2 versus D3 receptors. Biochim. Biophys. Acta (BBA) Biomembr. 2016, 1858, 2152–2162. [Google Scholar] [CrossRef] [PubMed]
  32. Min, X.; Sun, N.; Wang, S.; Zhang, X.; Kim, K.M. Sequestration of Gbetagamma by deubiquitinated arrestins into the nucleus as a novel desensitization mechanism of G protein-coupled receptors. Cell Commun. Signal. 2023, 21, 11. [Google Scholar] [CrossRef] [PubMed]
  33. Boeckler, F.; Gmeiner, P. The structural evolution of dopamine D3 receptor ligands: Structure–activity relationships and selected neuropharmacological aspects. Pharmacol. Ther. 2006, 112, 281–333. [Google Scholar] [CrossRef] [PubMed]
  34. Leggio, G.M.; Bucolo, C.; Platania, C.B.M.; Salomone, S.; Drago, F. Current drug treatments targeting dopamine D3 receptor. Pharmacol. Ther. 2016, 165, 164–177. [Google Scholar] [CrossRef]
  35. McCorvy, J.D.; Butler, K.V.; Kelly, B.; Rechsteiner, K.; Karpiak, J.; Betz, R.M.; Kormos, B.L.; Shoichet, B.K.; Dror, R.O.; Jin, J.; et al. Structure-inspired design of beta-arrestin-biased ligands for aminergic GPCRs. Nat. Chem. Biol. 2018, 14, 126–134. [Google Scholar] [CrossRef]
  36. Kiss, B.; Laszlovszky, I.; Kramos, B.; Visegrady, A.; Bobok, A.; Levay, G.; Lendvai, B.; Roman, V. Neuronal Dopamine D3 Re-ceptors: Translational Implications for Preclinical Research and CNS Disorders. Biomolecules 2021, 11, 104. [Google Scholar] [CrossRef]
  37. Sokoloff, P.; Le Foll, B. A Historical Perspective on the Dopamine D3 Receptor. Curr. Top Behav. Neurosci. 2022, 1–28. [Google Scholar] [CrossRef]
  38. Gainetdinov, R.R.; Premont, R.T.; Bohn, L.M.; Lefkowitz, R.J.; Caron, M.G. Desensitization of g Protein–Coupled Receptors and Neuronal Functions. Annu. Rev. Neurosci. 2004, 27, 107–144. [Google Scholar] [CrossRef] [Green Version]
  39. Williams, J.T.; Ingram, S.L.; Henderson, G.; Chavkin, C.; von Zastrow, M.; Schulz, S.; Koch, T.; Evans, C.J.; Christie, M.J. Regula-tion of mu-opioid receptors: Desensitization, phosphorylation, internalization, and tolerance. Pharmacol. Rev. 2013, 65, 223–254. [Google Scholar] [CrossRef] [Green Version]
  40. Sun, N.; Kim, K.-M. Mechanistic diversity involved in the desensitization of G protein-coupled receptors. Arch. Pharmacal Res. 2021, 44, 342–353. [Google Scholar] [CrossRef]
  41. Nambi, P.; Peters, J.R.; Sibley, D.R.; Lefkowitz, R.J. Desensitization of the turkey erythrocyte beta-adrenergic receptor in a cell-free system. Evidence that multiple protein kinases can phosphorylate and desensitize the receptor. J. Biol. Chem. 1985, 260, 2165–2171. [Google Scholar] [CrossRef] [PubMed]
  42. Benovic, J.L.; Pike, L.J.; Cerione, R.A.; Staniszewski, C.; Yoshimasa, T.; Codina, J.; Caron, M.G.; Lefkowitz, R.J. Phosphoryla-tion of the mammalian beta-adrenergic receptor by cyclic AMP-dependent protein kinase. Regulation of the rate of receptor phos-phorylation and dephosphorylation by agonist occupancy and effects on coupling of the receptor to the stimulatory guanine nucleo-tide regulatory protein. J. Biol. Chem. 1985, 260, 7094–7101. [Google Scholar] [PubMed]
  43. Cho, E.-Y.; Cho, D.-I.; Park, J.H.; Kurose, H.; Caron, M.G.; Kim, K.-M. Roles of Protein Kinase C and Actin-Binding Protein 280 in the Regulation of Intracellular Trafficking of Dopamine D3 Receptor. Mol. Endocrinol. 2007, 21, 2242–2254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Farrens, D.L.; Altenbach, C.; Yang, K.; Hubbell, W.L.; Khorana, H.G. Requirement of Rigid-Body Motion of Transmembrane Helices for Light Activation of Rhodopsin. Science 1996, 274, 768–770. [Google Scholar] [CrossRef] [PubMed]
  45. Scheerer, P.; Park, J.H.; Hildebrand, P.W.; Kim, Y.J.; Krauß, N.; Choe, H.-W.; Hofmann, K.P.; Ernst, O.P. Crystal structure of opsin in its G-protein-interacting conformation. Nature 2008, 455, 497–502. [Google Scholar] [CrossRef]
  46. Rasmussen, S.G.F.; DeVree, B.T.; Zou, Y.; Kruse, A.C.; Chung, K.Y.; Kobilka, T.S.; Thian, F.S.; Chae, P.S.; Pardon, E.; Calinski, D.; et al. Crystal structure of the β2 adrenergic receptor–Gs protein complex. Nature 2011, 477, 549–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Szczepek, M.; Beyrière, F.; Hofmann, K.P.; Elgeti, M.; Kazmin, R.; Rose, A.; Bartl, F.J.; von Stetten, D.; Heck, M.; Sommer, M.E.; et al. Crystal structure of a common GPCR-binding interface for G protein and arrestin. Nat. Commun. 2014, 5, 4801. [Google Scholar] [CrossRef] [Green Version]
  48. Kumari, P.; Srivastava, A.; Banerjee, R.; Ghosh, E.; Gupta, P.; Ranjan, R.; Chen, X.; Gupta, B.; Gupta, C.; Jaiman, D.; et al. Functional competence of a partially engaged GPCR-beta-arrestin complex. Nat. Commun. 2016, 7, 13416. [Google Scholar] [CrossRef] [Green Version]
  49. Hausdorff, W.P.; Caron, M.G.; Lefkowitz, R.J. Turning off the signal: Desensitization of beta-adrenergic receptor function. FASEB J. 1990, 4, 2881–2889. [Google Scholar] [CrossRef]
  50. Westrich, L.; Kuzhikandathil, E.V. The tolerance property of human D3 dopamine receptor is determined by specific ami-no acid residues in the second cytoplasmic loop. Biochim. Biophys. Acta 2007, 1773, 1747–1758. [Google Scholar] [CrossRef] [Green Version]
  51. Cho, D.; Zheng, M.; Min, C.; Ma, L.; Kurose, H.; Park, J.H.; Kim, K.M. Agonist-induced endocytosis and receptor phosphoryla-tion mediate resensitization of dopamine D(2) receptors. Mol. Endocrinol. 2010, 24, 574–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Min, C.; Zheng, M.; Zhang, X.; Caron, M.G.; Kim, K.M. Novel roles for beta-arrestins in the regulation of pharmacological sequestration to predict agonist-induced desensitization of dopamine D3 receptors. Br. J. Pharmacol. 2013, 170, 1112–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Namkung, Y.; Dipace, C.; Javitch, J.; Sibley, D.R. G Protein-coupled Receptor Kinase-mediated Phosphorylation Regulates Post-endocytic Trafficking of the D2 Dopamine Receptor. J. Biol. Chem. 2009, 284, 15038–15051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Min, C.; Zhang, X.; Zheng, M.; Sun, N.; Acharya, S.; Zhang, X.; Kim, K.-M. Molecular Signature That Determines the Acute Tolerance of G Protein-Coupled Receptors. Biomol. Ther. 2017, 25, 239–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zheng, M.; Zhang, X.; Min, X.; Sun, N.; Kim, K.-M. Cytoplasmic recruitment of Mdm2 as a common characteristic of G pro-tein-coupled receptors that undergo desensitization. Biochem. Biophys. Res. Commun. 2020, 530, 181–188. [Google Scholar] [CrossRef]
  56. Namkung, Y.; Sibley, D.R. Protein Kinase C Mediates Phosphorylation, Desensitization, and Trafficking of the D2 Dopamine Receptor. J. Biol. Chem. 2004, 279, 49533–49541. [Google Scholar] [CrossRef] [Green Version]
  57. Cho, D.I.; Min, C.; Jung, K.S.; Cheong, S.Y.; Zheng, M.; Cheong, S.J.; Oak, M.H.; Cheong, J.H.; Lee, B.K.; Kim, K.-M. The N-terminal region of the dopamine D2 receptor, a rhodopsin-like GPCR, regulates correct integration into the plasma membrane and endocytic routes. Br. J. Pharmacol. 2012, 166, 659–675. [Google Scholar] [CrossRef] [Green Version]
  58. Min, C.; Zheng, M.; Zhang, X.; Guo, S.; Kwon, K.J.; Shin, C.Y.; Kim, H.S.; Cheon, S.H.; Kim, K.M. N-linked Glycosylation on the N-terminus of the dopamine D2 and D3 receptors determines receptor association with specific microdomains in the plasma membrane. Biochim. Biophys. Acta 2015, 1853, 41–51. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, X.; Sun, N.; Zheng, M.; Kim, K.-M. Clathrin-mediated endocytosis is responsible for the lysosomal degradation of dopamine D3 receptor. Biochem. Biophys. Res. Commun. 2016, 476, 245–251. [Google Scholar] [CrossRef]
  60. Zhang, X.; Kim, K.-M. Multifactorial Regulation of G Protein-Coupled Receptor Endocytosis. Biomol. Ther. 2017, 25, 26–43. [Google Scholar] [CrossRef] [Green Version]
  61. Ferguson, S.S.; Downey, W.E.; Colapietro, A.M., 3rd; Barak, L.S.; Menard, L.; Caron, M.G. Role of beta-arrestin in mediating agonist-promoted G protein-coupled receptor internalization. Science 1996, 271, 363–366. [Google Scholar] [CrossRef] [PubMed]
  62. Ferguson, S.S.; Caron, M.G. G protein-coupled receptor adaptation mechanisms. Semin. Cell Dev. Biol. 1998, 9, 119–127. [Google Scholar] [CrossRef]
  63. Hinshaw, J.E.; Schmid, S.L. Dynamin self-assembles into rings suggesting a mechanism for coated vesicle budding. Nature 1995, 374, 190–192. [Google Scholar] [CrossRef] [PubMed]
  64. Laporte, S.A.; Oakley, R.H.; Holt, J.A.; Barak, L.S.; Caron, M.G. The interaction of beta-arrestin with the AP-2 adaptor is re-quired for the clustering of beta 2-adrenergic receptor into clathrin-coated pits. J. Biol. Chem. 2000, 275, 23120–23126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Oakley, R.H.; Laporte, S.A.; Holt, J.A.; Caron, M.G.; Barak, L.S. Differential affinities of visual arrestin, beta arrestin1, and beta arrestin2 for G protein-coupled receptors delineate two major classes of receptors. J. Biol. Chem. 2000, 275, 17201–17210. [Google Scholar] [CrossRef] [Green Version]
  66. Oakley, R.H.; Laporte, S.A.; Holt, J.A.; Barak, L.S.; Caron, M.G. Molecular determinants underlying the formation of stable intracellular G protein-coupled receptor-beta-arrestin complexes after receptor endocytosis. J. Biol. Chem. 2001, 276, 19452–19460. [Google Scholar] [CrossRef] [Green Version]
  67. Cho, D.I.; Zheng, M.; Min, C.; Kwon, K.J.; Shin, C.Y.; Choi, H.K.; Kim, K.M. ARF6 and GASP-1 are post-endocytic sorting pro-teins selectively involved in the intracellular trafficking of dopamine D(2) receptors mediated by GRK and PKC in transfected cells. Br. J. Pharmacol. 2013, 168, 1355–1374. [Google Scholar] [CrossRef] [Green Version]
  68. Mostafapour, S.; Kobilka, B.K.; von Zastrow, M. Pharmacological sequestration of a chimeric beta 3/beta 2 adrenergic re-ceptor occurs without a corresponding amount of receptor internalization. Recept. Signal Transduct. 1996, 6, 151–163. [Google Scholar]
  69. Burgoyne, R.D. Small GTP-binding proteins. Trends Biochem. Sci. 1989, 14, 394–396. [Google Scholar] [CrossRef]
  70. Takai, Y.; Sasaki, T.; Matozaki, T. Small GTP-binding proteins. Physiol. Rev. 2001, 81, 153–208. [Google Scholar] [CrossRef]
  71. Segev, N. GTPases in intracellular trafficking: An overview. Semin. Cell Dev. Biol. 2011, 22, 1–2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Stenmark, H.; Olkkonen, V.M. The Rab GTPase family. Genome Biol. 2001, 2, REVIEWS3007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Evans, T.; Ferguson, C.; Wainwright, B.; Parton, R.; Wicking, C. Rab23, a Negative Regulator of Hedgehog Signaling, Localizes to the Plasma Membrane and the Endocytic Pathway. Traffic 2003, 4, 869–884. [Google Scholar] [CrossRef] [PubMed]
  74. Zheng, M.; Zhang, X.; Guo, S.; Zhang, X.; Min, C.; Cheon, S.H.; Oak, M.-H.; Kim, Y.R.; Kim, K.-M. Agonist-induced changes in RalA activities allows the prediction of the endocytosis of G protein-coupled receptors. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2016, 1863, 77–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Zheng, M.; Zhang, X.; Sun, N.; Min, C.; Zhang, X.; Kim, K.M. RalA employs GRK2 and beta-arrestins for the filamin A-mediated regulation of trafficking and signaling of dopamine D and D receptor. Biochim. Biophys. Acta 2016, 1863, 2072–2083. [Google Scholar] [CrossRef]
  76. Iwata, K.; Ito, K.; Fukuzaki, A.; Inaki, K.; Haga, T. Dynamin and rab5 regulate GRK2-dependent internalization of dopamine D2 receptors. JBIC J. Biol. Inorg. Chem. 1999, 263, 596–602. [Google Scholar] [CrossRef] [Green Version]
  77. Lin, R.; Karpa, K.; Kabbani, N.; Goldman-Rakic, P.; Levenson, R. Dopamine D2 and D3 receptors are linked to the actin cy-toskeleton via interaction with filamin A. Proc. Natl. Acad. Sci. USA 2001, 98, 5258–5263. [Google Scholar] [CrossRef] [Green Version]
  78. Li, M.; Bermak, J.C.; Wang, Z.W.; Zhou, Q.Y. Modulation of Dopamine D2Receptor Signaling by Actin-Binding Protein (ABP-280). Mol. Pharmacol. 2000, 57, 446–452. [Google Scholar] [CrossRef] [Green Version]
  79. Li, M.; Li, C.; Weingarten, P.; Bunzow, J.R.; Grandy, D.K.; Zhou, Q.Y. Association of dopamine D(3) receptors with ac-tin-binding protein 280 (ABP-280). Biochem. Pharmacol. 2002, 63, 859–863. [Google Scholar] [CrossRef]
  80. Kim, K.M.; Gainetdinov, R.R.; Laporte, S.A.; Caron, M.G.; Barak, L.S. 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. 2005, 280, 12774–12780. [Google Scholar] [CrossRef] [Green Version]
  81. Radhakrishna, H.; Donaldson, J.G. ADP-Ribosylation Factor 6 Regulates a Novel Plasma Membrane Recycling Pathway. J. Cell Biol. 1997, 139, 49–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Montagnac, G.; de Forges, H.; Smythe, E.; Gueudry, C.; Romao, M.; Salamero, J.; Chavrier, P. Decoupling of Activation and Effector Binding Underlies ARF6 Priming of Fast Endocytic Recycling. Curr. Biol. 2011, 21, 574–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Claing, A.; Chen, W.; Miller, W.E.; Vitale, N.; Moss, J.; Premont, R.T.; Lefkowitz, R.J. beta-Arrestin-mediated ADP-ribosylation factor 6 activation and beta 2-adrenergic receptor endocytosis. J. Biol. Chem. 2001, 276, 42509–42513. [Google Scholar] [CrossRef] [Green Version]
  84. Gilman, A.G. G proteins: Transducers of receptor-generated signals. Annu. Rev. Biochem. 1987, 56, 615–649. [Google Scholar] [CrossRef] [PubMed]
  85. Angleson, J.K.; Wensel, T.G. A GTPase-accelerating factor for transducin, distinct from its effector cGMP phosphodiester-ase, in rod outer segment membranes. Neuron 1993, 11, 939–949. [Google Scholar] [CrossRef] [PubMed]
  86. Arshavsky, V.Y.; Pugh, E.N. Lifetime Regulation of G Protein–Effector Complex: Emerging Importance of RGS Proteins. Neuron 1998, 20, 11–14. [Google Scholar] [CrossRef] [Green Version]
  87. Ross, E.M.; Wilkie, T.M. GTPase-Activating Proteins for Heterotrimeric G Proteins: Regulators of G Protein Signaling (RGS) and RGS-Like Proteins. Annu. Rev. Biochem. 2000, 69, 795–827. [Google Scholar] [CrossRef] [Green Version]
  88. Taymans, J.-M.; Kia, H.K.; Claes, R.; Cruz, C.; Leysen, J.; Langlois, X. Dopamine receptor-mediated regulation of RGS2 and RGS4 mRNA differentially depends on ascending dopamine projections and time. Eur. J. Neurosci. 2004, 19, 2249–2260. [Google Scholar] [CrossRef]
  89. Taymans, J.-M.; Leysen, J.E.; Langlois, X. Striatal gene expression of RGS2 and RGS4 is specifically mediated by dopamine D1 and D2 receptors: Clues for RGS2 and RGS4 functions. J. Neurochem. 2003, 84, 1118–1127. [Google Scholar] [CrossRef]
  90. Rorabaugh, B.R.; Rose, M.J.; Stoops, T.S.; Stevens, A.A.; Seeley, S.L.; D’Souza, M.S. Regulators of G-protein signaling 2 and 4 differentially regulate cocaine-induced rewarding effects. Physiol. Behav. 2018, 195, 9–19. [Google Scholar] [CrossRef]
  91. Rahman, Z.; Schwarz, J.; Gold, S.J.; Zachariou, V.; Wein, M.N.; Choi, K.-H.; Kovoor, A.; Chen, C.-K.; DiLeone, R.; Schwarz, S.C.; et al. RGS9 Modulates Dopamine Signaling in the Basal Ganglia. Neuron 2003, 38, 941–952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Cabrera-Vera, T.M.; Hernandez, S.; Earls, L.R.; Medkova, M.; Sundgren-Andersson, A.K.; Surmeier, D.J.; Hamm, H.E. RGS9-2 modulates D 2 dopamine receptor-mediated Ca 2+ channel inhibition in rat striatal cholinergic interneurons. Proc. Natl. Acad. Sci. USA 2004, 101, 16339–16344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kovoor, A.; Seyffarth, P.; Ebert, J.; Barghshoon, S.; Chen, C.K.; Schwarz, S.; Axelrod, J.D.; Cheyette, B.N.; Simon, M.I.; Lester, H.A.; et al. D2 dopamine receptors colocalize regulator of G-protein signaling 9-2 (RGS9-2) via the RGS9 DEP domain, and RGS9 knock-out mice develop dyskinesias associated with dopamine pathways. J. Neurosci. 2005, 25, 2157–2165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Watson, A.; Katz, A.; Simon, M. A fifth member of the mammalian G-protein beta-subunit family. Expression in brain and activation of the beta 2 isotype of phospholipase C. J. Biol. Chem. 1994, 269, 22150–22156. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, S.; Hamm, H.E. DEP Domains: More Than Just Membrane Anchors. Dev. Cell 2006, 11, 436–438. [Google Scholar] [CrossRef] [Green Version]
  96. Cheever, M.L.; Snyder, J.T.; Gershburg, S.; Siderovski, D.P.; Harden, T.K.; Sondek, J. Crystal structure of the multifunctional Gbeta5-RGS9 complex. Nat. Struct. Mol. Biol. 2008, 15, 155–162. [Google Scholar] [CrossRef] [Green Version]
  97. Masuho, I.; Wakasugi-Masuho, H.; Posokhova, E.N.; Patton, J.R.; Martemyanov, K.A. Type 5 G protein beta subunit (Gbeta5) controls the interaction of regulator of G protein signaling 9 (RGS9) with membrane anchors. J. Biol. Chem. 2011, 286, 21806–21813. [Google Scholar] [CrossRef] [Green Version]
  98. Zheng, M.; Cheong, S.Y.; Min, C.; Jin, M.; Cho, D.I.; Kim, K.M. Beta-arrestin2 plays permissive roles in the inhibitory activities of RGS9-2 on G protein-coupled receptors by maintaining RGS9-2 in the open conformation. Mol. Cell Biol. 2011, 31, 4887–4901. [Google Scholar] [CrossRef] [Green Version]
  99. Benovic, J.L.; Kuhn, H.; Weyand, I.; Codina, J.; Caron, M.G.; Lefkowitz, R.J. Functional desensitization of the isolated be-ta-adrenergic receptor by the beta-adrenergic receptor kinase: Potential role of an analog of the retinal protein arrestin (48-kDa pro-tein). Proc. Natl. Acad. Sci. USA 1987, 84, 8879–8882. [Google Scholar] [CrossRef] [Green Version]
  100. Lohse, M.J.; Benovic, J.L.; Codina, J.; Caron, M.G.; Lefkowitz, R.J. Beta-Arrestin: A protein that regulates beta-adrenergic re-ceptor function. Science 1990, 248, 1547–1550. [Google Scholar] [CrossRef]
  101. Wei, H.; Ahn, S.; Shenoy, S.K.; Karnik, S.S.; Hunyady, L.; Luttrell, L.M.; Lefkowitz, R.J. Independent beta-arrestin 2 and G pro-tein-mediated pathways for angiotensin II activation of extracellular signal-regulated kinases 1 and 2. Proc. Natl. Acad. Sci. USA 2003, 100, 10782–10787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Urban, J.D.; Clarke, W.P.; von Zastrow, M.; Nichols, D.E.; Kobilka, B.; Weinstein, H.; Javitch, J.A.; Roth, B.L.; Christopoulos, A.; Sexton, P.M.; et al. Functional Selectivity and Classical Concepts of Quantitative Pharmacology. Experiment 2006, 320, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Violin, J.D.; DeWire, S.M.; Yamashita, D.; Rominger, D.H.; Nguyen, L.; Schiller, K.; Whalen, E.J.; Gowen, M.; Lark, M.W. Selec-tively engaging beta-arrestins at the angiotensin II type 1 receptor reduces blood pressure and increases cardiac performance. J. Pharmacol. Exp. Ther. 2010, 335, 572–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Reiter, E.; Ahn, S.; Shukla, A.K.; Lefkowitz, R.J. Molecular mechanism of beta-arrestin-biased agonism at sev-en-transmembrane receptors. Annu. Rev. Pharmacol. Toxicol. 2012, 52, 179–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Gurevich, V.V.; Gurevich, E.V. Biased GPCR signaling: Possible mechanisms and inherent limitations. Pharmacol. Ther. 2020, 211, 107540. [Google Scholar] [CrossRef] [PubMed]
  106. Soergel, D.G.; Subach, R.A.; Burnham, N.; Lark, M.W.; James, I.E.; Sadler, B.M.; Skobieranda, F.; Violin, J.D.; Webster, L.R. Bi-ased agonism of the mu-opioid receptor by TRV130 increases analgesia and reduces on-target adverse effects versus morphine: A randomized, double-blind, placebo-controlled, crossover study in healthy volunteers. Pain 2014, 155, 1829–1835. [Google Scholar] [CrossRef]
  107. Violin, J.D.; Crombie, A.L.; Soergel, D.G.; Lark, M.W. Biased ligands at G-protein-coupled receptors: Promise and progress. Trends Pharmacol. Sci. 2014, 35, 308–316. [Google Scholar] [CrossRef]
  108. Sanchez-Soto, M.; Verma, R.K.; Willette, K.A.B.; Gonye, E.C.; Moore, A.M.; Moritz, A.E.; Boateng, C.A.; Yano, H.; Free, R.B.; Shi, L.; et al. A structural basis for how ligand binding site changes can allosterically regulate GPCR signaling and engender func-tional selectivity. Sci. Signal. 2020, 13, eaaw5885. [Google Scholar] [CrossRef]
  109. Donthamsetti, P.; Gallo, E.F.; Buck, D.C.; Stahl, E.L.; Zhu, Y.; Lane, J.R.; Bohn, L.M.; Neve, K.A.; Kellendonk, C.; Javitch, J.A. Ar-restin recruitment to dopamine D2 receptor mediates locomotion but not incentive motivation. Mol. Psychiatry 2020, 25, 2086–2100. [Google Scholar] [CrossRef]
  110. Lan, H.; Liu, Y.; Bell, M.I.; Gurevich, V.V.; Neve, K.A. A dopamine D2 receptor mutant capable of G protein-mediated signaling but deficient in arrestin binding. Mol. Pharmacol. 2009, 75, 113–123. [Google Scholar] [CrossRef]
  111. Peterson, S.M.; Pack, T.F.; Wilkins, A.D.; Urs, N.M.; Urban, D.J.; Bass, C.E.; Lichtarge, O.; Caron, M.G. Elucidation of G-protein and beta-arrestin functional selectivity at the dopamine D2 receptor. Proc. Natl. Acad. Sci. USA 2015, 112, 7097–7102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Ballesteros, J.A.; Weinstein, H. Integrated methods for the construction of three-dimensional models and computational probing of structure-function relations in G protein-coupled receptors. Methods Neurosci. 1995, 25, 366–428. [Google Scholar]
  113. Kim, K.M.; Caron, M.G. Complementary roles of the DRY motif and C-terminus tail of GPCRS for G protein coupling and beta-arrestin interaction. Biochem. Biophys. Res. Commun. 2008, 366, 42–47. [Google Scholar] [CrossRef] [PubMed]
  114. Kim, J.-H.; Cho, E.-Y.; Min, C.; Park, J.H.; Kim, K.-M. Characterization of functional roles of DRY motif in the 2nd intracellular loop of dopamine D2 and D3 receptors. Arch. Pharmacal Res. 2008, 31, 474–481. [Google Scholar] [CrossRef] [PubMed]
  115. Roth, C.B.; Hanson, M.A.; Stevens, R.C. Stabilization of the human beta2-adrenergic receptor TM4-TM3-TM5 helix interface by mutagenesis of Glu122(3.41), a critical residue in GPCR structure. J. Mol. Biol. 2008, 376, 1305–1319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Lichtarge, O.; Bourne, H.R.; Cohen, F.E. An Evolutionary Trace Method Defines Binding Surfaces Common to Protein Families. J. Mol. Biol. 1996, 257, 342–358. [Google Scholar] [CrossRef] [Green Version]
  117. Peterson, S.M.; Pack, T.F.; Caron, M.G. Receptor, Ligand and Transducer Contributions to Dopamine D2 Receptor Functional Selectivity. PLoS ONE 2015, 10, e0141637. [Google Scholar] [CrossRef]
  118. Kroeze, W.K.; Sassano, M.F.; Huang, X.-P.; Lansu, K.; McCorvy, J.D.; Giguère, P.M.; Sciaky, N.; Roth, B.L. PRESTO-Tango as an open-source resource for interrogation of the druggable human GPCRome. Nat. Struct. Mol. Biol. 2015, 22, 362–369. [Google Scholar] [CrossRef] [Green Version]
  119. Xu, W.; Wang, X.; Tocker, A.M.; Huang, P.; Reith, M.E.; Liu-Chen, L.Y.; Smith, A.B.; Kortagere, S., 3rd. Functional Characterization of a Novel Series of Biased Signaling Dopamine D3 Receptor Agonists. ACS Chem. Neurosci. 2017, 8, 486–500. [Google Scholar] [CrossRef] [Green Version]
  120. Xu, W.; Bearoff, F.; Kortagere, S. G-protein biased signaling agonists of Dopamine D3 receptor promote distinct activation patterns of ERK1/2. Pharmacol. Res. Off. J. Ital. Pharmacol. Soc. 2022, 179, 106223. [Google Scholar] [CrossRef]
  121. Quan, W.; Kim, J.H.; Albert, P.R.; Choi, H.; Kim, K.M. Roles of G protein and beta-arrestin in dopamine D2 receptor-mediated ERK activation. Biochem. Biophys. Res. Commun. 2008, 377, 705–709. [Google Scholar] [CrossRef] [PubMed]
  122. Chen, X.; McCorvy, J.D.; Fischer, M.G.; Butler, K.V.; Shen, Y.; Roth, B.L.; Jin, J. Discovery of G Protein-Biased D2 Dopamine Receptor Partial Agonists. J. Med. Chem. 2016, 59, 10601–10618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Bonifazi, A.; Yano, H.; Guerrero, A.M.; Kumar, V.; Hoffman, A.F.; Lupica, C.R.; Shi, L.; Newman, A.H. Novel and Potent Do-pamine D(2) Receptor Go-Protein Biased Agonists. ACS Pharmacol. Transl. Sci. 2019, 2, 52–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Chien, E.Y.; Liu, W.; Zhao, Q.; Katritch, V.; Han, G.; Hanson, W.M.; Shi, A.; Newman, L.A.; Javitch, H.J.; Cherezov, A.V.; et al. Structure of the human dopamine D3 receptor in complex with a D2/D3 selective antagonist. Science 2010, 330, 1091–1095. [Google Scholar] [CrossRef] [Green Version]
  125. Wang, S.; Che, T.; Levit, A.; Shoichet, B.K.; Wacker, D.; Roth, B.L. Structure of the D2 dopamine receptor bound to the atypical antipsychotic drug risperidone. Nature 2018, 555, 269–273. [Google Scholar] [CrossRef]
  126. Yin, J.; Chen, K.-Y.M.; Clark, M.J.; Hijazi, M.; Kumari, P.; Bai, X.-C.; Sunahara, R.K.; Barth, P.; Rosenbaum, D.M. Structure of a D2 dopamine receptor–G-protein complex in a lipid membrane. Nature 2020, 584, 125–129. [Google Scholar] [CrossRef]
  127. Xu, P.; Huang, S.; Mao, C.; Krumm, B.E.; Zhou, X.E.; Tan, Y.; Huang, X.P.; Liu, Y.; Shen, D.D.; Jiang, Y.; et al. Structures of the human dopamine D3 receptor-G(i) complexes. Mol. Cell. 2021, 81, 1147–1159.e4. [Google Scholar] [CrossRef]
  128. Platania, C.B.M.; Salomone, S.; Leggio, G.M.; Drago, F.; Bucolo, C. Homology Modeling of Dopamine D2 and D3 Receptors: Molecular Dynamics Refinement and Docking Evaluation. PLoS ONE 2012, 7, e44316. [Google Scholar] [CrossRef]
  129. Lane, J.R.; Abramyan, A.M.; Adhikari, P.; Keen, A.C.; Lee, K.H.; Sanchez, J.; Verma, R.K.; Lim, H.D.; Yano, H.; Javitch, J.A.; et al. Distinct inactive conformations of the dopamine D2 and D3 receptors correspond to different extents of inverse agonism. Elife 2020, 9, e52189. [Google Scholar] [CrossRef]
  130. Lee, S.P.; So, C.H.; Rashid, A.J.; Varghese, G.; Cheng, R.; Lança, A.J.; O’Dowd, B.F.; George, S.R. Dopamine D1 and D2 Receptor Co-activation Generates a Novel Phospholipase C-mediated Calcium Signal. J. Biol. Chem. 2004, 279, 35671–35678. [Google Scholar] [CrossRef] [Green Version]
  131. Rashid, A.J.; So, C.H.; Kong, M.M.; Furtak, T.; El-Ghundi, M.; Cheng, R.; O’Dowd, B.F.; George, S.R. D1-D2 dopamine receptor heterooligomers with unique pharmacology are coupled to rapid activation of Gq/11 in the striatum. Proc. Natl. Acad. Sci. USA 2007, 104, 654–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Hasbi, A.; Fan, T.; Alijaniaram, M.; Nguyen, T.; Perreault, M.L.; O’Dowd, B.F.; George, S.R. Calcium signaling cascade links dopamine D1–D2 receptor heteromer to striatal BDNF production and neuronal growth. Proc. Natl. Acad. Sci. USA 2009, 106, 21377–21382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. O’Dowd, B.F.; Ji, X.; Nguyen, T.; George, S.R. Two amino acids in each of D1 and D2 dopamine receptor cytoplasmic regions are involved in D1–D2 heteromer formation. Biochem. Biophys. Res. Commun. 2012, 417, 23–28. [Google Scholar] [CrossRef] [Green Version]
  134. Łukasiewicz, S.; Faron-Górecka, A.; Dobrucki, J.; Polit, A.; Dziedzicka-Wasylewska, M. Studies on the role of the receptor protein motifs possibly involved in electrostatic interactions on the dopamine D1 and D2 receptor oligomerization. FEBS J. 2008, 276, 760–775. [Google Scholar] [CrossRef] [PubMed]
  135. Łukasiewicz, S.; Polit, A.; Kędracka-Krok, S.; Wędzony, K.; Maćkowiak, M.; Dziedzicka-Wasylewska, M. Hetero-dimerization of serotonin 5-HT2A and dopamine D2 receptors. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2010, 1803, 1347–1358. [Google Scholar] [CrossRef] [Green Version]
  136. Ciruela, F.; Burgueno, J.; Casado, V.; Canals, M.; Marcellino, D.; Goldberg, S.R.; Bader, M.; Fuxe, K.; Agnati, L.F.; Lluis, C.; et al. Combining mass spectrometry and pull-down techniques for the study of receptor heteromerization. Direct epitope-epitope electrostatic interactions between adenosine A2A and dopamine D2 receptors. Anal. Chem. 2004, 76, 5354–5363. [Google Scholar] [CrossRef] [PubMed]
  137. Ridray, S.; Griffon, N.; Mignon, V.; Souil, E.; Carboni, S.; Diaz, J.; Schwartz, J.-C.; Sokoloff, P. Coexpression of dopamine D1and D3receptors in islands of Calleja and shell of nucleus accumbens of the rat: Opposite and synergistic functional interactions. Eur. J. Neurosci. 1998, 10, 1676–1686. [Google Scholar] [CrossRef]
  138. Schwartz, J.C.; Diaz, J.; Bordet, R.; Griffon, N.; Perachon, S.; Pilon, C.; Ridray, S.; Sokoloff, P. Functional implications of multi-ple dopamine receptor subtypes: The D1/D3 receptor coexistence. Brain Res. Brain Res. Rev. 1998, 26, 236–242. [Google Scholar] [CrossRef]
  139. Cho, D.-I.; Quan, W.; Oak, M.-H.; Choi, H.J.; Lee, K.Y.; Kim, K.-M. Functional interaction between dopamine receptor subtypes for the regulation of c-fos expression. Biochem. Biophys. Res. Commun. 2007, 357, 1113–1118. [Google Scholar] [CrossRef]
  140. Marcellino, D.; Ferre, S.; Casado, V.; Cortes, A.; Le Foll, B.; Mazzola, C.; Drago, F.; Saur, O.; Stark, H.; Soriano, A.; et al. Identification of dopamine D1-D3 receptor heteromers. Indications for a role of synergistic D1-D3 receptor interactions in the striatum. J. Biol. Chem. 2008, 283, 26016–26025. [Google Scholar] [CrossRef] [Green Version]
  141. Fiorentini, C.; Busi, C.; Gorruso, E.; Gotti, C.; Spano, P.; Missale, C. Reciprocal Regulation of Dopamine D1 and D3 Receptor Function and Trafficking by Heterodimerization. Mol. Pharmacol. 2008, 74, 59–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Guitart, X.; Moreno, E.; Rea, W.; Sánchez-Soto, M.; Cai, N.-S.; Quiroz, C.; Kumar, V.; Bourque, L.; Cortés, A.; Canela, E.I.; et al. Biased G Protein-Independent Signaling of Dopamine D1-D3 Receptor Heteromers in the Nucleus Accumbens. Mol. Neurobiol. 2019, 56, 6756–6769. [Google Scholar] [CrossRef] [PubMed]
  143. Bontempi, L.; Savoia, P.; Bono, F.; Fiorentini, C.; Missale, C. Dopamine D3 and acetylcholine nicotinic receptor heteromeriza-tion in midbrain dopamine neurons: Relevance for neuroplasticity. Eur. Neuropsychopharmacol. 2017, 27, 313–324. [Google Scholar] [CrossRef] [PubMed]
  144. Le Foll, B.; Diaz, J.; Sokoloff, P. Increased dopamine D3 receptor expression accompanying behavioral sensitization to nicotine in rats. Synapse 2003, 47, 176–183. [Google Scholar] [CrossRef]
  145. Smith, L.N.; Bachus, S.E.; McDonald, C.G.; Smith, R.F. Role of the D3 dopamine receptor in nicotine sensitization. Behav. Brain Res. 2015, 289, 92–104. [Google Scholar] [CrossRef]
  146. Le Foll, B.; Goldberg, S.R.; Sokoloff, P. Dopamine D3receptor ligands for the treatment of tobacco dependence. Expert Opin. Investig. Drugs 2006, 16, 45–57. [Google Scholar] [CrossRef]
  147. Acharya, S.; Kundu, D.; Choi, H.J.; Kim, K.M. Metabotropic signaling cascade involved in alpha4beta2 nicotinic acetylcho-line receptor-mediated PKCbetaII activation. Biochim. Biophys. Acta Mol. Cell Res. 2020, 1867, 118721. [Google Scholar] [CrossRef]
  148. Ramazi, S.; Zahiri, J. Posttranslational modifications in proteins: Resources, tools and prediction methods. Database 2021, 2021, baab012. [Google Scholar] [CrossRef]
  149. Krishna, R.G.; Wold, F. Post-translational modification of proteins. Adv. Enzymol. Relat. Areas Mol. Biol. 1993, 67, 265–298. [Google Scholar]
  150. Patwardhan, A.; Cheng, N.; Trejo, J. Post-Translational Modifications of G Protein–Coupled Receptors Control Cellular Signaling Dynamics in Space and Time. Pharmacol. Rev. 2020, 73, 120–151. [Google Scholar] [CrossRef]
  151. Duarte, M.L.; Devi, L.A. Post-translational Modifications of Opioid Receptors. Trends Neurosci. 2020, 43, 417–432. [Google Scholar] [CrossRef] [PubMed]
  152. Haselbeck, A.; Tanner, W. Dolichyl phosphate-mediated mannosyl transfer through liposomal membranes. Proc. Natl. Acad. Sci. USA 1982, 79, 1520–1524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Bause, E. Structural requirements of N-glycosylation of proteins. Studies with proline peptides as conformational probes. Biochem. J. 1983, 209, 331–336. [Google Scholar] [CrossRef] [Green Version]
  154. Cantagrel, V.; Lefeber, D.J. From glycosylation disorders to dolichol biosynthesis defects: A new class of metabolic diseases. J. Inherit. Metab. Dis. 2011, 34, 859–867. [Google Scholar] [CrossRef] [Green Version]
  155. Chen, Q.; Miller, L.J.; Dong, M. Role of N-linked glycosylation in biosynthesis, trafficking, and function of the human glu-cagon-like peptide 1 receptor. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E62–E68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lackman, J.J.; Goth, C.K.; Halim, A.; Vakhrushev, S.Y.; Clausen, H.; Petaja-Repo, U.E. Site-specific O-glycosylation of N-terminal serine residues by polypeptide GalNAc-transferase 2 modulates human delta-opioid receptor turnover at the plasma membrane. Cell. Signal. 2018, 42, 184–193. [Google Scholar] [CrossRef]
  157. Wang, T.; Nakagawa, S.; Miyake, T.; Setsu, G.; Kunisue, S.; Goto, K.; Hirasawa, A.; Okamura, H.; Yamaguchi, Y.; Doi, M. Identi-fication and functional characterisation of N-linked glycosylation of the orphan G protein-coupled receptor Gpr176. Sci. Rep. 2020, 10, 4429. [Google Scholar] [CrossRef] [Green Version]
  158. Free, R.B.; Hazelwood, L.A.; Cabrera, D.M.; Spalding, H.N.; Namkung, Y.; Rankin, M.L.; Sibley, D.R. D1 and D2 Dopamine Receptor Expression Is Regulated by Direct Interaction with the Chaperone Protein Calnexin. J. Biol. Chem. 2007, 282, 21285–21300. [Google Scholar] [CrossRef] [Green Version]
  159. Wilden, U.; Kuhn, H. Light-dependent phosphorylation of rhodopsin: Number of phosphorylation sites. Biochemistry 1982, 21, 3014–3022. [Google Scholar] [CrossRef]
  160. Stadel, J.M.; Nambi, P.; Shorr, R.G.; Sawyer, D.F.; Caron, M.G.; Lefkowitz, R.J. Catecholamine-induced desensitization of tur-key erythrocyte adenylate cyclase is associated with phosphorylation of the beta-adrenergic receptor. Proc. Natl. Acad. Sci. USA 1983, 80, 3173–3177. [Google Scholar] [CrossRef] [Green Version]
  161. Benovic, J.L.; Strasser, R.H.; Caron, M.G.; Lefkowitz, R.J. Beta-adrenergic receptor kinase: Identification of a novel protein kinase that phosphorylates the agonist-occupied form of the receptor. Proc. Natl. Acad. Sci. USA 1986, 83, 2797–2801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Lawrence, R.T.; Searle, B.C.; Llovet, A.; Villén, J. Plug-and-play analysis of the human phosphoproteome by targeted high-resolution mass spectrometry. Nat. Methods 2016, 13, 431–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Liu, J.J.; Sharma, K.; Zangrandi, L.; Chen, C.; Humphrey, S.J.; Chiu, Y.-T.; Spetea, M.; Liu-Chen, L.-Y.; Schwarzer, C.; Mann, M. In vivo brain GPCR signaling elucidated by phosphoproteomics. Science 2018, 360, eaao4927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Pinna, L.A.; Ruzzene, M. How do protein kinases recognize their substrates? Biochim. Biophys. Acta (BBA) Mol. Cell Res. 1996, 1314, 191–225. [Google Scholar] [CrossRef] [Green Version]
  165. Sibley, D.R.; Peters, J.R.; Nambi, P.; Caron, M.G.; Lefkowitz, R.J. Desensitization of turkey erythrocyte adenylate cyclase. Be-ta-adrenergic receptor phosphorylation is correlated with attenuation of adenylate cyclase activity. J. Biol. Chem. 1984, 259, 9742–9749. [Google Scholar]
  166. Shichi, H.; Somers, R. Light-dependent phosphorylation of rhodopsin. Purification and properties of rhodopsin kinase. J. Biol. Chem. 1978, 253, 7040–7046. [Google Scholar] [CrossRef]
  167. Palczewski, K.; Buczyłko, J.; Kaplan, M.; Polans, A.; Crabb, J. Mechanism of rhodopsin kinase activation. J. Biol. Chem. 1991, 266, 12949–12955. [Google Scholar] [CrossRef]
  168. Samama, P.; Cotecchia, S.; Costa, T.; Lefkowitz, R. A mutation-induced activated state of the beta 2-adrenergic receptor. Extending the ternary complex model. J. Biol. Chem. 1993, 268, 4625–4636. [Google Scholar] [CrossRef]
  169. Benovic, J.; Onorato, J.; Lohse, M.; Dohlman, H.; Staniszewski, C.; Caron, M.; Lefkowitz, R. Synthetic peptides of the hamster beta 2-adrenoceptor as substrates and inhibitors of the beta-adrenoceptor kinase. Br. J. Clin. Pharmacol. 1990, 30, 3S–12S. [Google Scholar] [CrossRef]
  170. Chen, C.Y.; Dion, S.B.; Kim, C.M.; Benovic, J.L. Beta-adrenergic receptor kinase. Agonist-dependent receptor binding pro-motes kinase activation. J. Biol. Chem. 1993, 268, 7825–7831. [Google Scholar] [CrossRef]
  171. Komolov, K.E.; Benovic, J.L. G protein-coupled receptor kinases: Past, present and future. Cell. Signal. 2018, 41, 17–24. [Google Scholar] [CrossRef] [PubMed]
  172. Brown, N.G.; Fowles, C.; Sharma, R.; Akhtar, M. Mechanistic studies on rhodopsin kinase. Eur. J. Biochem. 1992, 208, 659–667. [Google Scholar] [CrossRef] [PubMed]
  173. Pickart, C.M. Mechanisms underlying ubiquitination. Annu. Rev. Biochem. 2001, 70, 503–533. [Google Scholar] [CrossRef] [PubMed]
  174. Ardley, H.C.; Robinson, P.A. E3 ubiquitin ligases. Essays Biochem. 2005, 41, 15–30. [Google Scholar] [CrossRef] [PubMed]
  175. Suresh, B.; Lee, J.; Kim, H.; Ramakrishna, S. Regulation of pluripotency and differentiation by deubiquitinating enzymes. Cell Death Differ. 2016, 23, 1257–1264. [Google Scholar] [CrossRef] [Green Version]
  176. Dittmar, G.; Winklhofer, K.F. Linear Ubiquitin Chains: Cellular Functions and Strategies for Detection and Quantification. Front. Chem. 2020, 7, 915. [Google Scholar] [CrossRef]
  177. Kennedy, J.E.; Marchese, A. Regulation of GPCR Trafficking by Ubiquitin. Prog. Mol. Biol. Transl. Sci. 2015, 132, 15–38. [Google Scholar] [CrossRef] [Green Version]
  178. Imai, Y.; Soda, M.; Inoue, H.; Hattori, N.; Mizuno, Y.; Takahashi, R. An Unfolded Putative Transmembrane Polypeptide, which Can Lead to Endoplasmic Reticulum Stress, Is a Substrate of Parkin. Cell 2001, 105, 891–902. [Google Scholar] [CrossRef] [Green Version]
  179. Marchese, A.; Benovic, J.L. Agonist-promoted Ubiquitination of the G Protein-coupled Receptor CXCR4 Mediates Lysosomal Sorting. J. Biol. Chem. 2001, 276, 45509–45512. [Google Scholar] [CrossRef] [Green Version]
  180. Aicart-Ramos, C.; Valero, R.A.; Rodriguez-Crespo, I. Protein palmitoylation and subcellular trafficking. Biochim. Biophys. Acta (BBA) Biomembr. 2011, 1808, 2981–2994. [Google Scholar] [CrossRef] [Green Version]
  181. Resh, M.D. Membrane Targeting of Lipid Modified Signal Transduction Proteins. Sub Cell. Biochem. 2004, 37, 217–232. [Google Scholar] [CrossRef]
  182. Fukata, Y.; Bredt, D.S.; Fukata, M. Protein Palmitoylation by DHHC Protein Family. In The Dynamic Synapse: Molecular Methods in Ionotropic Receptor Biology; Kittler, J.T., Moss, S.J., Eds.; Taylor & Francis: Boca Raton, FL, USA, 2006. [Google Scholar]
  183. Dunphy, J.T.; Schroeder, H.; Leventis, R.; Greentree, W.K.; Knudsen, J.K.; Silvius, J.R.; Linder, M.E. Differential effects of acyl-CoA binding protein on enzymatic and non-enzymatic thioacylation of protein and peptide substrates. Biochim. Biophys. Acta 2000, 1485, 185–198. [Google Scholar] [CrossRef] [PubMed]
  184. Koster, K.P.; Yoshii, A. Depalmitoylation by Palmitoyl-Protein Thioesterase 1 in Neuronal Health and Degeneration. Front. Synaptic Neurosci. 2019, 11, 25. [Google Scholar] [CrossRef] [Green Version]
  185. Wang, A.; Yang, H.-C.; Friedman, P.; Johnson, C.; Dennis, E. A specific human lysophospholipase: cDNA cloning, tissue distribution and kinetic characterization. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 1999, 1437, 157–169. [Google Scholar] [CrossRef]
  186. Rocks, O.; Peyker, A.; Kahms, M.; Verveer, P.J.; Koerner, C.; Lumbierres, M.; Kuhlmann, J.; Waldmann, H.; Wittinghofer, A.; Bastiaens, P.I.H. An Acylation Cycle Regulates Localization and Activity of Palmitoylated Ras Isoforms. Science 2005, 307, 1746–1752. [Google Scholar] [CrossRef] [PubMed]
  187. Verkruyse, L.A.; Hofmann, S.L. Lysosomal Targeting of Palmitoyl-protein Thioesterase. J. Biol. Chem. 1996, 271, 15831–15836. [Google Scholar] [CrossRef] [Green Version]
  188. Qanbar, R.; Bouvier, M. Role of palmitoylation/depalmitoylation reactions in G-protein-coupled receptor function. Pharmacol. Ther. 2002, 97, 1–33. [Google Scholar] [CrossRef]
  189. Chini, B.; Parenti, M. G-protein-coupled receptors, cholesterol and palmitoylation: Facts about fats. J. Mol. Endocrinol. 2009, 42, 371–379. [Google Scholar] [CrossRef] [Green Version]
  190. Ng, G.Y.K.; O’Dowd, B.F.; Caron, M.; Dennis, M.; Brann, M.R.; George, S.R. Phosphorylation and Palmitoylation of the Human D2L Dopamine Receptor in Sf9 Cells. J. Neurochem. 2002, 63, 1589–1595. [Google Scholar] [CrossRef]
  191. Grunewald, S.; Haase, W.; Reilander, H.; Michel, H. Glycosylation, palmitoylation, and localization of the human D2S receptor in baculovirus-infected insect cells. Biochemistry 1996, 35, 15149–15161. [Google Scholar] [CrossRef]
  192. Ebersole, B.; Petko, J.; Woll, M.; Murakami, S.; Sokolina, K.; Wong, V.; Stagljar, I.; Lüscher, B.; Levenson, R. Effect of C-Terminal S-Palmitoylation on D2 Dopamine Receptor Trafficking and Stability. PLoS ONE 2015, 10, e0140661. [Google Scholar] [CrossRef]
  193. Wolf, M.; Dahlin, S.; Hu, X.-T.; Xue, C.-J.; White, K. Effects of lesions of prefrontal cortex, amygdala, or fornix on behavioral sensitization to amphetamine: Comparison with N-methyl-d-aspartate antagonists. Neuroscience 1995, 69, 417–439. [Google Scholar] [CrossRef] [PubMed]
  194. Cador, M.; Bjijou, Y.; Cailhol, S.; Stinus, L. d-Amphetamine-induced behavioral sensitization: Implication of a glutamatergic medial prefrontal cortex–ventral tegmental area innervation. Neuroscience 1999, 94, 705–721. [Google Scholar] [CrossRef] [PubMed]
  195. Richtand, N.M. Behavioral Sensitization, Alternative Splicing, and D3 Dopamine Receptor-Mediated Inhibitory Function. Neuropsychopharmacology 2006, 31, 2368–2375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Bordet, R.; Ridray, S.; Carboni, S.; Diaz, J.; Sokoloff, P.; Schwartz, J.-C. Induction of dopamine D 3 receptor expression as a mechanism of behavioral sensitization to levodopa. Proc. Natl. Acad. Sci. USA 1997, 94, 3363–3367. [Google Scholar] [CrossRef] [Green Version]
  197. Damsma, G.; Bottema, T.; Westerink, B.H.; Tepper, P.G.; Dijkstra, D.; Pugsley, T.A.; MacKenzie, R.G.; Heffner, T.G.; Wikstrom, H. Pharmacological aspects of R-(+)-7-OH-DPAT, a putative dopamine D3 receptor ligand. Eur. J. Pharmacol. 1993, 249, R9–R10. [Google Scholar] [CrossRef]
  198. Meller, E.; Bohmaker, K.; Goldstein, M.; Basham, D.A. Evidence that striatal synthesis-inhibiting autoreceptors are dopa-mine D3 receptors. Eur. J. Pharmacol. 1993, 249, R5–R6. [Google Scholar] [CrossRef]
  199. Koeltzow, T.E.; Xu, M.; Cooper, D.C.; Hu, X.-T.; Tonegawa, S.; Wolf, M.E.; White, F.J. Alterations in Dopamine Release but Not Dopamine Autoreceptor Function in Dopamine D3Receptor Mutant Mice. J. Neurosci. 1998, 18, 2231–2238. [Google Scholar] [CrossRef] [Green Version]
  200. Kreiss, D.S.; Bergstrom, D.A.; Gonzalez, A.M.; Huang, K.-X.; Sibley, D.R.; Walters, J.R. Dopamine receptor agonist potencies for inhibition of cell firing correlate with dopamine D3 receptor binding affinities. Eur. J. Pharmacol. 1995, 277, 209–214. [Google Scholar] [CrossRef]
  201. Piercey, M.F.; Hoffmann, W.E.; Smith, M.W.; Hyslop, D.K. Inhibition of dopamine neuron firing by pramipexole, a dopamine D3 receptor-preferring agonist: Comparison to other dopamine receptor agonists. Eur. J. Pharmacol. 1996, 312, 35–44. [Google Scholar] [CrossRef]
  202. Daly, S.A.; Waddington, J.L. Behavioural effects of the putative D-3 dopamine receptor agonist 7-OH-DPAT in relation to other “D-2-like” agonists. Neuropharmacology 1993, 32, 509–510. [Google Scholar] [CrossRef] [PubMed]
  203. Boer, P.; Enrico, P.; Wright, J.; Wise, L.D.; Timmerman, W.; Moor, E.; Dijkstra, D.; Wikstrom, H.V.; Westerink, B.H. Charac-terization of the effect of dopamine D3 receptor stimulation on locomotion and striatal dopamine levels. Brain Res. 1997, 758, 83–91. [Google Scholar] [CrossRef] [PubMed]
  204. Waters, N.; Haadsma-Svensson, S.; Svensson, K.; Sonesson, C.; Carlsson, A. Differential effects of dopamine D2 and D3 receptor antagonists in regard to dopamine release, in vivo receptor displacement and behaviour. J. Neural Transm. 1994, 98, 39–55. [Google Scholar] [CrossRef] [PubMed]
  205. Sotnikova, T.D.; Gainetdinov, R.R.; Grekhova, T.V.; Rayevsky, K.S. Effects of intrastriatal infusion of D2 and D3 dopamine receptor preferring antagonists on dopamine release in rat dorsal striatum (in vivo microdialysis study). Pharmacol. Res. 2001, 43, 283–290. [Google Scholar] [CrossRef]
  206. Li, T.; Hou, Y.; Yan, C.X.; Chen, T.; Zhao, Y.; Li, S.B. Dopamine D3 receptor knock-out mice display deficits in locomotor sen-sitization after chronic morphine administration. Neurosci. Lett. 2010, 485, 256–260. [Google Scholar] [CrossRef]
Figure 1. Major posttranslational modification sites of dopamine D2 receptor and their functional roles. S/T residues highlighted in red indicate GRK-mediated phosphorylation sites; circles filled with yellow and S/T residues highlighted in purple indicate the amino acids responsible for GRK-mediated D2R endocytosis; S/T residues highlighted in blue indicate PKC-mediated phosphorylation sites.
Figure 1. Major posttranslational modification sites of dopamine D2 receptor and their functional roles. S/T residues highlighted in red indicate GRK-mediated phosphorylation sites; circles filled with yellow and S/T residues highlighted in purple indicate the amino acids responsible for GRK-mediated D2R endocytosis; S/T residues highlighted in blue indicate PKC-mediated phosphorylation sites.
Ijms 24 06742 g001
Figure 2. Major posttranslational modification sites of dopamine D3 receptor and their functional roles. Circles filled with yellow indicate the amino acids responsible for agonist-induced desensitization; the two S residues highlighted in red indicate PKC-mediated regulatory processes.
Figure 2. Major posttranslational modification sites of dopamine D3 receptor and their functional roles. Circles filled with yellow indicate the amino acids responsible for agonist-induced desensitization; the two S residues highlighted in red indicate PKC-mediated regulatory processes.
Ijms 24 06742 g002
Figure 3. Sequence alignment of dopamine D2 and D3 receptors. D2LR represents alternatively spliced long form of dopamine D2 receptor. The asparagine residues (N) indicated by orange letters located within the N-terminal tail or within the extracellular loops represent potential N-linked glycosylation sites. Asterisks indicate amino acid residues involved in biased signaling. For details, please refer to the text.
Figure 3. Sequence alignment of dopamine D2 and D3 receptors. D2LR represents alternatively spliced long form of dopamine D2 receptor. The asparagine residues (N) indicated by orange letters located within the N-terminal tail or within the extracellular loops represent potential N-linked glycosylation sites. Asterisks indicate amino acid residues involved in biased signaling. For details, please refer to the text.
Ijms 24 06742 g003
Table 1. Roles of posttranslational modification of dopamine D2 and D3 receptors.
Table 1. Roles of posttranslational modification of dopamine D2 and D3 receptors.
PTMReceptorSites (Region)Functional Roles
GlycosylationD2RN5, N17, N23 (N-terminus)Caveolar endocytosis [57,58]
D3RN12, N19 (N-terminus)Clathrin-mediated endocytosis [58]
N97 (ECL1), N173 (ECL2)Receptor endocytosis [58]
PhosphorylationD2RS256, S257, T258, S259, T264, S282, S288, S292Agonist-induced phosphorylation, receptor recycling [51,53]
S228, S229, T322, T324, S325PKC-mediated phosphorylation, receptor endocytosis; desensitization (S325) [56]
D3RS229, S257PKC-mediated phosphorylation, endocytosis, desensitization [43]
UbiquitinationD2RNone
D3R?PMA-induced lysosomal downregulation [59]
PalmitoylationD2RNone
D3RC400Cell-surface expression, PKC-mediated endocytosis, agonist affinity, desensitization [31]
Table 2. Regions involved in the biased signaling of dopamine D2 receptor.
Table 2. Regions involved in the biased signaling of dopamine D2 receptor.
SitesSpecificity (G Protein vs. Arrestin)
F198A (F5.38)G protein > Arrestin [108]
R132L/L123W (R3.50/L3.41)G protein > Arrestin [109]
212IYIV215 → AAAAG protein > Arrestin [110]
L125N/Y133LG protein > Arrestin [111]
A135R/M140DG protein < Arrestin [111]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kim, K.-M. Unveiling the Differences in Signaling and Regulatory Mechanisms between Dopamine D2 and D3 Receptors and Their Impact on Behavioral Sensitization. Int. J. Mol. Sci. 2023, 24, 6742. https://doi.org/10.3390/ijms24076742

AMA Style

Kim K-M. Unveiling the Differences in Signaling and Regulatory Mechanisms between Dopamine D2 and D3 Receptors and Their Impact on Behavioral Sensitization. International Journal of Molecular Sciences. 2023; 24(7):6742. https://doi.org/10.3390/ijms24076742

Chicago/Turabian Style

Kim, Kyeong-Man. 2023. "Unveiling the Differences in Signaling and Regulatory Mechanisms between Dopamine D2 and D3 Receptors and Their Impact on Behavioral Sensitization" International Journal of Molecular Sciences 24, no. 7: 6742. https://doi.org/10.3390/ijms24076742

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop