Treatment of D2-receptor-expressing cells with specific drugs upregulates the receptor number at the cell surface independently of protein synthesis, leading to the concept of an intracellular receptor pool. However, how this pool is operating is still an enigma. Here, we report that a splice variant of the Gαi2 protein, protein sGαi2, plays a crucial role in the maintenance of this D2-receptor pool. Co-expression of sGi2 with D2 receptor reduced receptor localization to cell surface by one-third. This effect is associated with specific intracellular protein-protein interaction and the formation of a sGi2–D2-receptor complex. It has been suggested that the formation of this complex serves to prevent D2 receptors from reaching the cell membrane. Treatment of D2-receptor-expressing cells with agonists increased the number of cell surface D2 receptors and coincided with a reduction in these receptors from intracellular complexes, suggesting that agonist treatment released D2 receptors from the complex allowing them to localize to the cell membrane. Thus, in addition to elucidating how the intracellular pool of D2 receptor functions, our findings uncover a novel mechanism regulating the density of cell surface D2 receptors.

Drug-induced augmentation of D2 receptor on plasma membrane has been reported in a cell line that expresses endogenous D2 receptors (Ivins et al., 1991), as well as in HEK-293 cells (Boundy et al., 1995; Filtz et al., 1993), CHO cells (Itokawa et al., 1996; Zhang et al., 1994), C6 glioma cells and Ltk cells (Starr et al., 1995), and Sf9 cells (Ng et al., 1997) when expressing recombinant D2 receptors. This upregulation was not affected by the inhibition of protein synthesis (Filtz et al., 1993; Starr et al., 1995; Ng et al., 1997) and, therefore, it was proposed that this increase of receptors was owing to their recruitment from existing intracellular reservoir(s) (Ng et al., 1997). Increase in receptor numbers following drug treatments has also been observed in other transmitter-receptor systems (Creese and Sibley, 1981). The most recent work describing agonist exposure of cells that express D2 receptors found translocation of receptor from cytoplasm to plasma membrane (Ng et al., 1997), providing direct evidence of functional receptor pool in the cytoplasm. It is well known that the interaction of plasma-membrane-bound dopamine D2 receptor with Gαi2 protein is fundamental for signal transmission. However, we and others have shown previously that, in contrast to Gαi2 protein – which is localized at the cell surface, sGi2 is an intracellular protein and not found at plasma membrane (Khan and Gutierrez, 2004; Montmayeur and Borrelli, 1994). A Proline-rich motif at the C-terminus is thought to be crucial for the intracellular translocation of this protein (Picetti and Borrelli, 2000). The sGi2 transcript encodes a protein with a different C-terminus, in which a 24-amino acid (aa) strech is replaced by a 35-aa sequence. Like sGi2, other G-proteins have also been identified on intracellular membranes (Audigier et al., 1988; Weiss et al., 2001), where they participate in different functions, such as protein transport (Bomsel and Mostov, 1992; Erkolani et al., 1990; Helms, 1995; Pimplikar and Simons, 1993). On the basis of the interaction between non-spliced Gαi2 protein and dopamine D2 receptor (Gazi et al., 2003; Senogles, 1994), and the localization of sGi2 protein in brain dopaminergic cells (Khan and Gutierrez, 2004) where prominent expression of D2 receptors was also found (Khan et al., 1998a), we hypothesized that the sGi2 protein takes part in the translocation of dopamine D2 receptors to cell surface. To our surprise, sGi2 protein not only regulated the density of D2 receptor at cell surface but also participated in the formation of an intracellular reservoir of this receptor.

In intact BHK cells, co-expression of the long isoform of the D2 receptor (D2L) and the sGi2 protein led to a 31% decrease in the number (Bmax) of plasma-membrane-bound receptor (5.25±0.41 fmol per 106 cells of D2L alone versus 3.65±0.35 fmol per 106 cells of D2L with sGi2). This loss in D2L receptor was dependent on the amount of sGi2 protein co-expression (Fig. 1A). As the level of expression of sGi2 increased, more reduction in Bmax of D2L receptor was observed. However, no further reduction was observed at 30 μl of pseudovirions (not shown). The substitution of sGi2 with Gαi2 had no effect (Fig. 1A). This finding suggests that the receptor loss seen is specifically associated to sGi2 protein. To exclude the possibility that sGi2 inhibits protein expression, we checked the expression levels of both D2L and sGi2 proteins in these experiments by immunoblotting, by using affinity-purified antibodies [for data on specificity tests for D2L and D2S antibodies see Khan et al. (Khan et al., 1998a) and for sGi2 antibodies see supplementary material Fig. S1]. In fact, we found that the D2L receptor concentration was unchanged, whereas expression of sGi2 protein was increased as expected (Fig. 1B), indicating that reduction of D2L was not due to lower expression of this receptor. Next, we analyzed the binding affinity (Kd) of D2 receptor that might have been compromised during co-expression of both proteins, but no discernable change was observed (Kd=80±15 pM without sGi2 and 73±16 pM with sGi2). A similar effect of sGi2 protein on D2L as well as D2S receptors was also observed in other cell lines, including JEG-3 (Fig. 1C) and NG108-15 (Fig. 3B). These results not only confirm the observations made in BHK cells but also implicate that the regulation of the density of cell surface D2S and D2L receptors by sGi2 protein is a common characteristic among various cell types. The use of human carcinoma JEG-3 cells that lack D2 receptor and Gαi2 (Guiramand et al., 1995) also showed effects similar to other cell lines. It was observed that co-expression of either D2L or D2S receptors with sGi2 protein produced a loss of 32% or 35%, respectively (Fig. 1C). Furthermore, evidence from immunoblot analysis (Fig. 1B) suggests that, even though the total expression of D2L receptors (cell-surface-bound plus intracellular) was unchanged, localization of D2L receptor on the cell surface was reduced. Therefore, we further examined the localization of D2 receptors in these cells by double immunofluorescence labeling (Fig. 2) and found that, when D2S or D2L receptors were expressed alone, they mainly localized at plasma membrane (Fig. 2A), co-expression with sGi2, membrane localization of the receptor was noticeably reduced. This reduction in cell surface localization was probably due to accumulation of D2S-sGi2 protein complex seen in the intracellular membranes (Fig. 2B,C,D).

To further demonstrate that the loss of D2 receptor activity in intact cells was due to a reduction in functional cell surface receptor population, we performed in-vivo Ca2+-transient studies in these cells by Ca2+ imaging. It is known that antagonist-mediated blockade of D2 receptor augments intracellular Ca2+ flow via membrane-bound voltage-gated Ca2+ channels (Chronwall et al., 1995; Pauwels et al., 2001). Therefore, we used this paradigm expecting that a decrease in the cell-surface-associated D2 receptor after co-expression of sGi2 proportionally reduces the D2-antagonist-mediated Ca2+ rise. Because of downstream D2-signaling pathways, NG108-15 – a neuroblastoma/glioma cell line of neural origin (Pilon et al., 1994) – was also included in this study. Indeed, application of 15 μM raclopride, a D2 receptor antagonist, produced a transient increase in the intracellular Ca2+ concentration, which was significantly reduced (30-35%) in cells that co-expressed sGi2, similar to the observation made in ligand-binding experiments (D2L 2.1±0.1 versus D2L+sGi2 1.53±0.1 and D2S 2.2±0.2 versus D2S+sGi2 1.44±0.1; values are F/Fo ratios of Ca2+ changes after drug application from the baseline) (Fig. 3A-C). In control experiments, the use of neither Gαi2 protein in place of sGi2 nor D1 in place of D2S or D2L receptor produced any such effect (Fig. 3C). In contrast to antagonist, D2 receptor activation by agonist reduces Ca2+ flow (Lledo et al., 1990; Wolfe and Morris, 1999). Consistent with this, agonist treatment led to a decrease in Ca2+ levels in cells expressing D2 receptor; however, co-expression of sGi2 reduced this decrease (see supplementary material Fig. S2). These results suggest that a reduced efficiency of Ca2+ flow is associated with a lower number of D2 receptors at the plasma membrane. To rule out the participation of intracellular Ca2+ stores in our experiments, we used 2-aminoethoxydiphenyl borate (APB), an inhibitor of the IP3 receptor. Treatment of cells with APB did not influence the Ca2+ transients, whereas the thapsigargin-stimulated Ca2+ release from intracellular stores could still be observed. This observation suggested that intracellular Ca2+ stores were intact but did not participate in D2-modulated Ca2+ increase. In addition, the use of Ca2+-free medium or medium containing EGTA and CoCl2 (blockers of membrane Ca2+ channels) and APB yielded the same results, further supporting this notion (see supplementary material Fig. S3).

Direct evidence of interaction between D2 receptor and sGi2 protein came from the co-elution of a D2-sGi2 complex using affinity-columns (Fig. 4A). The columns were prepared with affinity-purified specific antibodies against D2L, D2S (Khan et al., 1998a; Khan et al., 2001) or sGi2 protein (see supplementary material Fig. S1 for evidence on antibody specificity). Both affinity-columns that were immobilized with antibodies against D2L and D2S co-eluted sGi2 protein (Fig. 4A). Using the sGi2 antibody affinity-column, we observed co-elution of D2S and D2L receptors but not of D1 receptor (Fig. 4A). To further demonstrate the functional interaction between sGi2 protein and active D2 receptors, solubilized proteins from cells were incubated with antiserum against sGi2 and presence of co-imunoprecipitated D2 receptor was determined. We observed 24.6±4.9% co-precipitation of D2 receptor with sGi2 antibodies (Fig. 4B). The fact that sGi2 antibody did not bind D1 receptor suggests again that the D2-sGi2 interaction is specific.

To find out whether the complex of D2-sGi2 also exists in brain tissues, we used extracts from substantia nigra, a region where most neurons express high number of dopamine D2S receptors (Khan et al., 1998a) and where we have also observed a strong immunolabeling of sGi2 protein (Khan and Gutierrez, 2004). Immunoaffinity co-elution experiments similar to those described above confirmed the existence of a D2S-sGi2 protein complex in this tissue (Fig. 5).

Furthermore, we performed deletion experiments with cDNA of sGi2 to dissect the site involved in their intracellular interaction. As indicated in Fig. 6A, deletion constructs of sGi2 were co-expressed with D2S receptor in BHK cells and D2 receptor density at cell surface was determined by whole-cell binding assays. Truncated protein constructs lacking 108 bp of the extreme 3′-terminal end (sGi2-N1, sGi2-N2 and sGi2-C2) lost the capability to interact (Fig. 6B); in contrast to sGi2 and its deletion construct (sGi2-C1) that both contain this extreme 3′-terminal, which retained this activity. These results suggest that the 36 amino acids C-terminal of sGi2 are essential for the binding with D2 receptor and necessary to invoke the effect of complex formation.

To further test the effect of D2 drugs on cell surface receptor density, we treated cells that expressed both D2S and sGi2 protein with D2-agonists for 30 minutes. Exposure with 10 μM of dopamine or 5 μM quinpirole led to an increase in the density of cell-surface-bound D2S receptors (Fig. 7A). These results are in agreement with earlier reports, in which an increase in the cell-surface-bound D2 receptor was observed after exposure to dopamine D2 drugs (Filtz et al., 1993; Starr et al., 1995; Ng et al., 1997). Treatment of the same cells with 5 μM raclopride, a D2 antagonist, had no effect (Fig. 7A). Next, we used the cells from the experiment described in Fig. 7A to isolate the sGi2-D2 complex with sGi2-immunoaffinity columns and to determine the concentration of D2S receptor bound to sGi2. The combination of immunoblots and optical-density measurements showed that dopamine and quinpirole treatment reduced the amount of D2 bound to sGi2, whereas levels of D2 in both raclopride-treated and untreated control cells was unchanged (Fig. 7B,C). The levels of sGi2 under these conditions were unchanged (Fig. 7B). Our findings suggest that D2-agonist-mediated activity has freed D2 receptors from sGi2-containing intracellular complexes, and that these D2 receptors then localized to the plasma membrane. Values obtained in the experiments using D2L or D2S were not significantly different from each other.

Here, we have presented evidence that the sGi2 protein participates in the formation of an intracellular D2 receptor pool by specific protein-protein interaction and in the regulation of the density of these receptors at the cell membrane. To our knowledge, this constitutes the first description of how the D2 receptor pool system functions within the cell. Our results of intracellular D2S- and D2L-receptor sequestration by sGi2 protein, and their translocation to cell surface after D2-agonist treatment explain how an increase in numbers of D2 receptor is possible while protein synthesis is blocked. An increase in the number of D2L receptors on the cell surface and a proportional decrease in the number of the same receptors in the cytoplasm were observed when D2L-expressing cells were exposed to agonist (Ng et al., 1997). In rats treated with haloperidol, dopamine challenge led to upregulation of the dopamine D2 receptor (Creese et al., 1976; Severson et al., 1984). However, the prevalence of steady-state D2 receptor RNA was unaffected (Goss et al., 1991). It was therefore, suggested that an intracellular D2 receptor pool is needed to upregulate D2 during dopamine challenge. It is likely that this pool supplies receptor to plasma membrane in conditions such as those described by `the law of denervation' and when normal protein synthesis is not capable to fulfill the requirement. Furthermore, the maintenance of the D2 receptor reservoir while protein synthesis is active (Starr et al., 1995), suggests that the machinery synthesizing dopamine D2 receptor is not a substitute for the D2 receptor pool. However, it remains to be explored whether, after synthesis, dopamine D2 receptors are first localized to the intracellular reservoir before being translocated to the cell surface, or whether they are transported directly to the plasma membrane when the reservoir is saturated. It is also reasonable to argue that this sGi2-driven reservoir might not only function as a stock room for D2 receptor but might also control the amount of D2-mediated signal to be transmitted inside the cell by regulating the cell surface density of this receptor.

The presence of sGi2-D2 receptor complex in brain tissues suggests their physiological importance in brain function. Apart from expression in neuronal intracellular compartments (Khan and Gutierrez, 2004), sGi2 proteins were also found in abundance in axons and spines, where the most nota observation was its frequent localization to or near to the neck of spines (70% of 41 labeled spines observed) (see supplementary material Fig. S4). In addition, these proteins are often localized not at, but in close proximity to, the synapse. Given the binding capability of sGi2 with D2 receptor, this strategic extra-synaptic presence of sGi2 proteins suggests that these proteins bind D2 receptors and prevent their localization at the synapse and, as a result, this binding can interrupt the full participation of D2 receptors in synaptic neurotransmission events. Therefore, this mechanism might fine-tune the D2-mediated synaptic transmission, depending on the requirement of local circuits. Although we have confirmed the non-binding to other dopamine, glycine and GABAA receptors, it remains to be determined whether sGi2 protein binds to other receptors and synaptic proteins to participate in similar processes.

The C-terminal of the Gαi2 protein is known to interact with the dopamine D2 receptor in order to transmit signals at the plasma membrane (Boundy et al., 1993; Damaj et al., 1996; Senogles et al., 2004); the third cytoplasmic loop of dopamine D2 receptor was found to be crucial for this Gi protein interaction (Malek et al., 1993). Though, in sGi2, this terminal end is replaced, our results using deletion constructs demonstrate that 36 amino acids of the C-terminal end are essential for the intracellular interaction with the D2 receptor. In contrast to sGi2, proteins that interact with the third intracellular loop of the dopamine D2 receptor, such as spinophilin (Smith et al., 1999), filamin A (Li et al., 2000; Lin et al., 2001) and heart fatty-acid-binding protein (Takeuchi and Fukunaga, 2003), have also been identified. In addition, the dimerization of D2 receptors through interaction has also been shown (Lee et al., 2003). Thus, these evidences point to the ability of dopamine D2 receptor to participate in protein-protein interaction with various cellular proteins and not only with sGi2 protein, as reported here.

In the central nervous system, cell surface dopamine D2 receptors are the major target of all effective antipsychotic drugs. Their interaction is considered to be the key event associated with improvements in patients (Kapur and Remington, 2001; Seeman and Kapur, 2000) and also in the generation of extrapyramidal side effects (Strange, 2001). Usually, antipsychotic drugs block dopamine D2 receptor signalling; therefore, intracellular sequestration of D2 receptor may offer an alternative in reducing the D2-mediated signaling without blocking its function. The advantage of this approach is that intracellular sequestration of D2 receptor might reduce inhibitory components of D2 signaling, one of the main causes for side effects in patients using antipsychotics.

In conclusion, our results demonstrate that the sGi2 protein and the dopamine D2 receptor form intracellular complexes that serve as of D2 receptor reservoir. Treatment with D2-specific drugs break down this protein complex and free D2 receptor can translocate to cell surface. We postulate that, in contrast to a long-term strategy where protein synthesis is essential, this mechanism is a short-term cellular strategy to cope with the demand for D2 receptor while the protein synthesis machinery is unable to respond.

cDNA

cDNA clone of Gαi2 (GenBank accession number, M17528) and Gαi3 (GenBank accession number M20713) was provided by Randall R. Reed (Johns Hopkins University, Baltimore, MD). Human D1, D2S and D2L cDNA clones were from Olivier Civelli (University of California, Irvine, CA). The full-length cDNA of sGi2 was obtained from human brain poly A+ RNA (Clontech) and was submitted to GenBank at the accession number AY677118.

Cloning

Genes were amplified by PCR using primers containing MluI and ApaI or MluI and XbaI restriction sites at their 5′ and 3′ ends, respectively. For the deletion constructs, PCR primers were designed to amplify the sGi2 DNA sequences containing restriction sites like those described above. The digestions were performed by incubating 3 μg of DNA with 40 units of MluI and ApaI or MluI and XbaI at 37° C for 4 hours. Digested DNA was separated on 1.4% agarose gels and recovered with Gel Extraction Kit (Qiagen). Digestion of pSinRep5 plasmid vector (Sindbis Expression System from Invitrogen) was also performed as above and was used for ligation. Gene DNA (0.5 μg) was added to 0.5 μg of digested plasmid DNA, ligated in the presence of 2 Weiss units of bacteriophage T4 DNA ligase and incubated for 1 hour at room temperature. The ligation mixture was directly transformed into a competent TOP10 One Shot cells (Invitrogen). Colonies were selected on LB agar plates containing 100 μg/ml ampicillin. After isolation of plasmid DNA with Wizard Plus Minipreps DNA Purification System (Promega) from several colonies, they were analyzed for the presence of gene by restriction digestion with MluI and ApaI and then by PCR using a combination of primers from both plasmid and gene. Usually, two to three colonies that showed correct size gene insert were then processed for large-quantity DNA isolation using Wizard Maxi-Plasmid Preparation System (Promega). The isolated DNA was quantified, aliquoted and stored at –20° C. These recombinant samples, including deletion constructs, were sequenced to confirm the DNA sequence.

In vitro transcription, transfection and preparation of pseudovirions

The above described recombinant genes were used as template to produce recombinant RNA with the InvitroScript Cap SP6 in vitro Transcription Kit (Invitrogen). Briefly, 5 μg of recombinant DNA was linearized with 50 units of NotI. The digest was extracted once with phenol-chloroform and 0.1 volumes of 5 M ammonium acetate and 2 volumes of ethanol were added and the mix was incubated at –20°C for 1 hour. After centrifugation, DNA pellet was suspended in RNase-free water to 0.5 μg/μl. The in vitro transcription reaction was set up at room temperature by mixing 1 μg of the linearized recombinant DNA with the SP6 transcription reagents as indicated by the Invitrogen protocol. The reaction was mixed gently and incubated for 2 hours at 37°C. A typical reaction yielded 10-20 μg of RNA. The RNA product was purified with phenol-chloroform extraction, quantified with spectrophotometer, aliquoted into 10 μl samples and stored at –80°C.

2×105 baby hamster kidney (BHK) cells were seeded into six-well culture plates in 2 ml of growth medium and incubated at 37° C in 5% CO2 for 12-18 hours until 80% confluency. Cells of each well were then washed with 2 ml OPTI-MEM I reduced-serum medium at room temperature. For liposome-mediated transfection, DMRIE-C reagent from Gibco was used. RNA-lipid complexes were prepared by adding 9 μl of liposome reagent (DMRIE-C), 9 μg of recombinant RNA and 9 μg of helper RNA to 1 ml of OPTI-MEM I in polystyrene tubes and were mixed briefly by vortexing. The lipid-RNA complexes were immediately added to the washed BHK cells and incubated for 4 hours at 37°C. Following the incubation, transfection medium was replaced with complete growth medium containing αMEM medium supplemented with 2 mM L-glutamine and 5% fetal bovine serum and the cells were incubated for an additional 36 hours. During this period, recombinant RNA are packaged into pseudovirion particles and then released into the medium. The medium from the cells was collected, aliquoted into 1 ml samples and stored at –80°C.

Infection of cells with recombinant pseudovirions

Cell lines used in this study were obtained from the American Type Culture Collection and they were cultured at 37°C in a 5% CO2 atmosphere. BHK cells were grown in αMEM medium supplemented with 2 mM L-glutamine and 5% fetal bovine serum. JEG-3 human carcinoma cells were cultured in Eagle's MEM with 2 mM L-glutamine and Earle's BSS containing 1.5 g/l sodium bicarbonate, 0.1 mM non-essential amino acids, 1.0 mM sodium pyruvate, and 10% fetal bovine serum. NG108-15 neuroblastoma/glioma cells were grown in DMEM with 4 mM L-glutamine without sodium pyruvate and modified to contain 4.5 g/l glucose, 1.5 g/l sodium bicarbonate, 0.005 mM pyridoxin-HCl, HAT supplement, and 10% fetal bovine serum. Cells for infection were grown to approximately 70-80% confluency in 60-mm tissue culture plates, and pseudovirions (0-30 μl) diluted to 450 μl was added to each well. After incubation at room temperature for 1 hour, 4 ml of medium were added and cells were incubated for 30-34 hours for expression of functional protein. The optimal amount of pseudovirions needed for maximal protein expression was determined by making serial dilutions of the stock.

Whole-cell binding assays

After infection with recombinant pseudovirions and protein expression, intact cells were harvested, counted and processed for the binding assays. As described previously (Khan et al., 1998b; Khan et al., 2001), the binding of [3H]sulpiride (NEN-PerkinElmer) to 105 cells was done by incubation with 0-1 nM of the radioligand for 1 hour at 24°C in a total volume of 0.5 ml. The reaction was terminated by rapid filtration through glass filters and counted for the retained radioactivity. This value was considered as total binding. Non-specific binding was determined with 1 μM of (+)-butaclamol-HCl or fluphenazine (RBI-Sigma). Specific binding was calculated by subtracting non-specific binding from total binding. Bmax and Kd values were calculated with Prism program (GraphPad Software). Data are presented as the mean ± s.d. from six to eight independent experiments.

To control the amount of receptor expression, cells were homogenized in parallel experiments and used for receptor binding, similar as described for whole-cells receptor binding, to determine the total number of receptors in each condition. Variations in the total number of receptor within the same experiment and also between experiments were below 3%. Furthermore, the amount of pseudovirions and number of cells used in experiments were kept constant throughout the study.

Antibodies

A peptide corresponding to sGi2 protein residues 343-354, LSGPDQHPHPSP (GenBank accession number AY677118), was synthesized and coupled to keyhole limpet hemocyanin (KLH). Peptide conjugation and rabbit immunizations were performed as described previously (Khan and Gutierrez, 2004; Khan et al., 1998a; Khan et al., 1998b). Development of immune response was monitored by ELISA using immobilized synthetic peptides. Antibodies against sGi2 were affinity-purified on the corresponding immobilized peptide as described in detail elsewhere (Khan et al., 1993). Briefly, peptide (5 mg) was coupled to 1 g of activated thiopropyl-Sepharose 6B (Pharmacia LKB). One milliliter of antiserum diluted fivefold in 10 mM phosphate-buffered saline (PBS) (10 mM Na2PO4, 0.14 M NaCl, 0.01 M KCl, pH 7.4) was circulated through the peptide column. After washing, the antibody was eluted with 50 mM glycine-HCl pH 2.3, and collected in 1-ml fractions. OD280 was determined for each fraction and fractions containing antibodies were pooled and dialyzed in PBS. Antibodies were stored as 50-μl aliquots at –20°C. Specificity of affinity-purified antibody was then determined (Khan and Gutierrez, 2004) (supplementary material Fig. S1). The isoform-specific D2S and D2L antibodies had been prepared earlier by us and their specificity have already been demonstrated (Khan et al., 1998a; Khan et al., 2001).

Immunoblots

Immunoblots were done as described previously (Khan et al., 1998a; Khan et al., 1993). Solubilization of proteins from harvested intact cells was done with solubilization buffer provided in Seize X Mammalian Immunoprecipitation kit (Pierce). These solubilized proteins were separated by 10% SDS-PAGE and transblotted to nitrocellulose membranes. Membranes containing proteins were incubated with 5 μg/ml antibodies to sGi2, D2S or D2L, followed by incubation with anti-rabbit IgG-HRP (1:2000; Amersham). Bands were visualized using an ECL kit (Amersham).

Calculation of percentile in immunoblots

The concentration of bands in blot experiment (Fig. 7B) was obtained by OD measurements (Fig. 7C). These results suggest that approximately 75% of D2 receptor was uncoupled from the D2-sGi2 complex after drug treatment. This calculation is based on the assumption that D2 receptor population bound to sGi2-D2 complex is 100% (Fig. 7B, C5) under normal conditions when D2 receptor and sGi2 are co-expressed. However, when comparing data of blots with binding experiments, this 100% of D2 receptor population of the D2-sGi2 complex represent approximately 30% in binding experiments (see Fig. 1). Therefore, normalizing the results of both blots and binding experiments to the same level (30%), the 75% value of blots comes down to 22%. A slightly lower value in the binding experiments might reflect the population of unbound receptor still in transit.

Fluorescence immunocytochemistry

After infection with recombinant pseudovirions and after protein expression, cells grown on Flask-style glass slides (Nunc) were fixed with 4% paraformaldehyde and 0.2% glutaraldehyde for 10 minutes and permeabilized with 0.3% Triton X-100. Immunofluorescence staining of cells was performed as described earlier (Khan et al., 1998a; Khan et al., 1998b; Khan et al., 2001; Khan and Gutierrez, 2004; Lopez-Aranda et al., 2006). Briefly, after incubation with sGi2 antibody (1:500), cells were incubated with FITC (green) coupled to anti-rabbit Fab2 fragment (1:100; Jackson), followed by incubation with D2S antibody (1:200) and Cy3-conjugated secondary antibody(red) (1:200; Jackson). Images were taken with Zeiss confocal microscope.

Ca2+ imaging

Cells infected with recombinant pseudovirions were grown on glass coverslips and incubated with 2 μM fluo-4-AM (fluo-4 acetoxymethyl ester from Molecular Probes) for 15-30 minutes. The fluorescence change in cells after application of 15 μM raclopride was measured with a Zeiss LSM 410 confocal laser scanning microscope system as described previously (Koulen et al., 1999). Images were acquired every 500 mseconds. Changes in fluorescence intensity were calculated by dividing the fluorescence intensity during drug application (F) by the average baseline fluorescence intensity (F0). Non-stimulus-related spontaneous changes in fluorescence were 1-3%. Data are presented as the mean ± s.d. of four independent experiments.

Immunoaffinity elution

Protein solubilization, antibody immobilized affinity column preparation and protein elution was performed as described in Seize X Mammalian Immunoprecipitation Kit (Pierce). In brief, 0.8 ml of gel-immobilized protein G was washed with PBS buffer (10 mM Na2PO4, 0.14 M NaCl, 0.01 M KCl, pH 7.4) by centrifugation and incubated with 2 mg of affinity-purified antibodies for 1 hour. The mixture was then transferred to spin cups and centrifuged. Flow-through was collected to determine the amount of antibody bound to resin. Approximately 80-90% of antibodies were bound; 1.3 mg of DSS crosslinker in DMSO was added to the resin and gently mixed by inversion for 1 hour. Resin was washed with Tris buffer (25 mM Tris, 0.15 M NaCl, pH 7.2) in spin cups and stored in 1 ml of PBS buffer containing 0.01% sodium azide.

After infection with recombinant pseudovirions, whole cells (2×106) were harvested and lysed with 2 ml of M-PER Mammalian Protein Extraction reagent (Pierce) for 10 minutes. After removing cell debris by centrifugation, clear supernatant was diluted 1:1 with PBS buffer and added to spin columns with resin bound to antibody. The samples were incubated for 2 hour at 4°C and eluted with 400 μl of elution buffer (pH 2.8). Immunoaffinity eluted proteins were then analyzed by immunoblots.

For the brain tissues, membrane was prepared as described earlier (Khan et al., 1998a; Khan et al., 1998b; Khan et al., 2001; Khan et al., 1993). The prepared membrane was then used for the protein solubilization, binding with affinity-column and elution as explained above.

Co-immunoprecipitation

Cells were infected with recombinant pseudovirions, harvested and their proteins were solubilized with 1% digitonin (Khan et al., 1998a; Khan et al., 1998b). After centrifugation, the supernatant was used for incubation with 20 μl of affinity-purified sGi2 antibody. The protein-antibody complexes were separated by incubation with 80 μl of proteinA-agarose (Sigma) followed by centrifugation. The non-immunoprecipitated supernatant was used for the binding assay using D2-specific ([3H]raclopride) and D1-specific ([3H]SCH 23390) radioligands as described elsewhere (Khan et al., 1993; Khan et al., 1998a; Khan et al., 1998b; Khan et al., 2001). For the binding assay, supernatant was incubated with 1 nM radioligand in total of 0.5 ml. Reaction was terminated by rapid filtration and retained radioactivity was counted as described above in detail in whole-cell binding assays. The amount of co-immunoprecipitated receptors was calculated by subtracting the binding values of supernatant from the total (100%) binding of radioligands. Incubation without addition of antiserum represented 100% binding.

Calculation of immunoprecipitation values

For calculation and deduction of immunoprecipitation values as in Fig. 4, proteins extract of cells was incubated with pre-immune serum or antiserum against sGi2 protein after termination of the experiment. The immunocomplex (sGi2-D2 complex) was precipitated using proteinA-agarose. The supernatant portion of this reaction was used for binding assays and cpm values were obtained after counting in scintillation counter. A value of 100% was assigned to the cpm values obtained from the extract treated with preimmune serum. The immunoprecipitation values were calculated by subtracting the values of supernatant originated from extract treated with antiserum to the 100% value, meaning amount of supernatant derived from extract treated with preimmune serum minus amount of supernatant derived from extract treated with antiserum equal the amount of immunoprecipitated receptor.

Drugs treatment

Following infection and protein expression, harvested intact cells were incubated with agonists (10 μM dopamine and 5 μM quinpirole, both from Sigma/RBI) and antagonist (5 μM raclopride from Sigma/RBI) for 30 minutes. After washing, cells were processed for whole-cell binding assays and immunoblots as described above.

We thank O. Civelli for cDNA of dopamine receptors; R. Reed for cDNA of G-proteins. This work was supported by BFI2003-03464, BFU2006-0306 and Ramón y Cajal program grants from MEC (to Z.U.K.), FIS PI060556 grant from MSC (to A.G.) and a Young Investigator Award from NARSAD (to P.K.).

Audigier, Y., Nigam, S. K. and Blobel, G. (
1988
). Identification of a G protein in rough endoplasmic reticulum of canine pancreas.
J. Biol. Chem.
263
,
16352
-16357.
Bomsel, M. and Mostov, K. (
1992
). Role of heterotrimeric G proteins in membrane traffic.
Mol. Biol. Cell
3
,
1317
-1328.
Boundy, V. A., Luedtke, R. R. and Molinoff, P. B. (
1993
). Development of polyclonal anti-D2 dopamine receptor antibodies to fusion proteins: inhibition of D2 receptor-G protein interaction.
J. Neurochem.
60
,
2181
-2191.
Boundy, V. A., Pacheco, M. A., Guan, W. and Molinoff, P. B. (
1995
). Agonists and antagonists differentially regulate the high affinity state of the D2L receptor in human embryonic kidney 293 cells.
Mol. Pharmacol.
48
,
956
-964.
Chronwall, B. M., Beatty, D. M., Sharma, P. and Morris, S. J. (
1995
). Dopamine D2 receptor regulate in vitro melanotroph L-type Ca2+ channel activity.
Endocrinology
136
,
614
-621.
Creese, I. and Sibley, D. R. (
1981
). Receptor adaptations to centrally acting drugs.
Annu. Rev. Pharmacol. Toxicol.
21
,
357
-391.
Creese, I., Burt, D. R. and Snyder, S. H. (
1976
). Dopamine receptor binding predicts clinical and pharmacological potencies of antischizophrenic drugs.
Science
192
,
481
-483.
Damaj, B. B., McColl, S. R., Mahana, W., Crouch, M. F. and Naccache, P. H. (
1996
). Physical association of Gi2alpha with interleukin-8 receptors.
J. Biol. Chem.
271
,
12783
-12789.
Ercolani, L., Stow, J. L., Boyle, J. F., Holtzman, E. J., Lin, H., Grove, J. R. and Ausiello, D. A. (
1990
). Membrane localization of the pertussis toxin-sensitive G-protein subunits alpha i-2 and alpha i-3 and expression of a metallothionein-alpha i-2 fusion gene in LLC-PK1 cells.
Proc. Natl. Acad. Sci. USA
87
,
4635
-4639.
Filtz, T. M., Artymyshyn, R. P., Guan, W. and Molinoff, P. B. (
1993
). Paradoxical regulation of dopamine receptors in transfected 293 cells.
Mol. Pharmacol.
44
,
371
-379.
Gazi, L., Nickolls, S. A. and Strange, P. G. (
2003
). Functional coupling of the human dopamine D2 receptor with G alpha i1, G alpha i2, G alpha i3 and G alpha o G proteins: evidence for agonist regulation of G protein selectivity.
Br. J. Pharmacol.
138
,
775
-786.
Goss, J. R., Kelly, A. B., Johnson, S. A. and Morgan, D. G. (
1991
). Haloperidol treatment increases D2 dopamine receptor protein independently of RNA levels in mice.
Life Sci.
48
,
1015
-1022.
Guiramand, J., Montmayeur, J. P., Ceraline, J., Bhatia, M. and Borrelli, E. (
1995
). Alternative splicing of the dopamine D2 receptor directs specificity of coupling to G-proteins.
J. Biol. Chem.
270
,
7354
-7358.
Helms, J. B. (
1995
). Role of heterotrimeric GTP binding proteins in vesicular protein transport: indications for both classical and alternative G protein cycles.
FEBS Lett.
369
,
84
-88.
Itokawa, M., Toru, M., Ito, K., Tsuga, H., Kameyama, K., Haga, T., Arinami, T. and Hamaguchi, H. (
1996
). Sequestration of the short and long isoforms of dopamine D2 receptors expressed in Chinese hamster ovary cells.
Mol. Pharmacol.
49
,
560
-566.
Ivins, K. J., Luedtke, R. R., Artymyshyn, R. P. and Molinoff, P. B. (
1991
). Regulation of dopamine D2 receptors in a novel cell line (SUP1).
Mol. Pharmacol.
39
,
531
-539.
Kapur, S. and Remington, G. (
2001
). Dopamine D2 receptors and their role in atypical antipsychotic action: still necessary and may even be sufficient.
Biol. Psychiatry
50
,
873
-883.
Khan, Z. U. and Gutierrez, A. (
2004
). Distribution of C-terminal splice variant of G alpha i2 in rat and monkey brain.
Neuroscience
127
,
833
-843.
Khan, Z. U., Fernando, L. P., Escriba, P., Busquets, X., Mallet, J., Miralles, C. P., Filla, M. and De Blas, A. L. (
1993
). Antibodies to the human γ2 subunit of the γ-aminobutyric acidA/benzodiazepine receptor.
J. Neurochem.
60
,
961
-971.
Khan, Z. U., Mrzljak, L., Gutierrez, A., de la Calle, A. and Goldman-Rakic, P. S. (
1998a
). Prominence of the dopamine D2 short isoform in dopaminergic pathways.
Proc. Natl. Acad. Sci. USA
95
,
7731
-7736.
Khan, Z. U., Gutierrez, A., Martin, R., Penafiel, A., Rivera, A. and De La Calle, A. (
1998b
). Differential regional and cellular distribution of dopamine D2-like receptors. An immunocytochemical study of subtype-specific antibodies in rat and human brain.
J. Comp. Neurol.
402
,
353
-371.
Khan, Z. U., Koulen, P., Rubinstein, M., Grandy, D. K. and Goldman-Rakic, P. S. (
2001
). An astroglia-linked dopamine D2-receptor action in prefrontal cortex.
Proc. Natl. Acad. Sci. USA
98
,
1964
-1969.
Koulen, P., Kuhn, R., Wassle, H. and Brandstatter, J. H. (
1999
). Modulation of the intracellular calcium concentration in photoreceptor terminals by a presynaptic metabotropic glutamate receptor.
Proc. Natl. Acad. Sci. USA
96
,
9909
-9914.
Lee, S. P., O'Dowd, B. F., Rajaram, R. D., Nguyen, T. and George, S. R. (
2003
). D2 dopamine receptor homodimerization is mediated by multiple sites of interaction, including an intermolecular interaction involving transmembrane domain 4.
Biochemistry
42
,
11023
-11031.
Li, M., Bermak, J. C., Wang, Z. W. and Zhou, Q. Y. (
2000
). Modulation of dopamine D(2) receptor signaling by actin-binding protein (ABP-280).
Mol. Pharmacol.
57
,
446
-452.
Lin, R., Karpa, K., Kabbani, N., Goldman-Rakic, P. S. and Levenson, R. (
2001
). Dopamine D2 and D3 receptors are linked to the actin cytoskeleton via interaction with filamin A.
Proc. Natl. Acad. Sci. USA
98
,
5258
-5263.
Lledo, P. M., Legendre, P., Israel, J. M. and Vincent, J. D. (
1990
). Dopamine inhibits two characterized voltage-dependent calcium currents in identified rat lactotroph cells.
Endocrinology
127
,
990
-1001.
Lopez-Aranda, M. F., Acevedo, M. J., Carballo, F. J., Gutierrez, A. and Khan, Z. U. (
2006
). Localization of GoLoco motif carrier regulator of G-protein signalling 12 and 14 proteins in monkey and rat brain.
Eur. J. Neurosci.
23
,
2971
-2982.
Malek, D., Munch, G. and Palm, D. (
1993
). Two sites in the third inner loop of the dopamine D2 receptor are involved in functional G protein-mediated coupling to adenylate cyclase.
FEBS Lett.
325
,
215
-219.
Montmayeur, J. P. and Borrelli, E. (
1994
). Targeting of G alpha i2 to the Golgi by alternative spliced carboxyl-terminal region.
Science
263
,
95
-98.
Ng, G. Y., Varghese, G., Chung, H. T., Trogadis, J., Seeman, P., O'Dowd, B. F. and George, S. R. (
1997
). Resistance of the dopamine D2L receptor to desensitization accompanies the up-regulation of receptors on to the surface of Sf9 cells.
Endocrinology
138
,
4199
-4206.
Pauwels, P. J., Tardiff, S., Wurch, T. and Colpaert, F. C. (
2001
). Real-time analysis of dopamine: antagonist interaction at recombinant human D2 long receptor upon modulation of its activation state.
Br. J. Pharmacol.
134
,
66
-97.
Picetti, R. and Borrelli, E. (
2000
). A region containing a proline-rich motif targets sG(i2) to the golgi apparatus.
Exp. Cell Res.
255
,
258
-269.
Pilon, C., Levesque, D., Dimitriadou, V., Griffon, N., Martres, M. P., Schwartz, J. C. and Sokoloff, P. (
1994
). Functional coupling of the human dopamine D2 receptor in a transfected NG 108-15 neuroblastoma-glioma hybrid cell line.
Eur. J. Pharmacol.
268
,
129
-139.
Pimplikar, S. W. and Simons, K. (
1993
). Regulation of apical transport in epithelial cells by a Gs class of heterotrimeric G protein.
Nature
362
,
456
-458.
Seeman, P. and Kapur, S. (
2000
). Schizophrenia: more dopamine, more D2 receptors.
Proc. Natl. Acad. Sci. USA
97
,
7673
-7675.
Senogles, S. E. (
1994
). D2 dopamine receptor isoforms signal through distinct Gi alpha proteins to inhibit adenylyl cyclase. A study with site-directed mutant Gi alpha proteins.
J. Biol. Chem.
269
,
23120
-23127.
Senogles, S. E., Heimert, T. L., Odife, E. R. and Quasney, M. W. (
2004
). A region of the third intracellular loop of the short form of the D2 dopamine receptor dictates Gi coupling specificity.
J. Biol. Chem.
279
,
1601
-1606.
Severson, J. A., Robinson, H. E. and Simpson, G. M. (
1984
). Neuroleptic-induced striatal dopamine receptor supersensitivity in mice: relationship to dose and drug.
Psychopharmacology Berl.
84
,
115
-119.
Smith, F. D., Oxford, G. S. and Milgram, S. L. (
1999
). Association of the D2 dopamine receptor third cytoplasmic loop with spinophilin, a protein phosphatase-1-interacting protein.
J. Biol. Chem.
274
,
19894
-19900.
Starr, S., Kozell, L. B. and Neve, K. A. (
1995
). Drug-induced up-regulation of dopamine D2 receptors on cultured cells.
J. Neurochem.
65
,
569
-577.
Strange, P. G. (
2001
). Antipsychotic drugs: importance of dopamine receptors for mechanisms of therapeutic actions and side effects.
Pharmacol. Rev.
53
,
119
-133.
Takeuchi, Y. and Fukunaga, K. (
2003
). Differential subcellular localization of two dopamine D2 receptor isoforms in transfected NG108-15 cells.
J. Neurochem.
85
,
1064
-1074.
Wedegaertner, P. B. (
2002
). Characterization of subcellular localization and stability of a splice variant of G alpha i2.
BMC Cell. Biol.
3
,
12
.
Weiss, T. S., Chamberlain, C. E., Takeda, T., Lin, P., Hahn, K. M. and Farquhar, M. G. (
2001
). Gαi3 binding to calnuc on Golgi membranes in living cells monitored by fluorescence resonance energy transfer of green fluorescent protein fusion proteins.
Proc. Natl. Acad. Sci. USA
98
,
14961
-14966.
Wolfe, S. E. and Morris, S. J. (
1999
). Dopamine D2 receptor isoforms expressed in AtT20 cells differentially couple to G proteins to acutely inhibit high voltage-activated calcium channels.
J. Neurochem.
73
,
2375
-2382.
Zhang, L. J., Lachowicz, J. E. and Sibley, D. R. (
1994
). The D2S and D2L dopamine receptor isoforms are differentially regulated in Chinese hamster ovary cells.
Mol. Pharmacol.
45
,
878
-889.

Supplementary information