Skip to main content

Beyond the G protein α subunit: investigating the functional impact of other components of the Gαi3 heterotrimers

Abstract

Background

Specific interactions between G protein-coupled receptors (GPCRs) and G proteins play a key role in mediating signaling events. While there is little doubt regarding receptor preference for Gα subunits, the preferences for specific Gβ and Gγ subunits and the effects of different Gβγ dimer compositions on GPCR signaling are poorly understood. In this study, we aimed to investigate the subcellular localization and functional response of Gαi3-based heterotrimers with different combinations of Gβ and Gγ subunits.

Methods

Live-cell imaging microscopy and colocalization analysis were used to investigate the subcellular localization of Gαi3 in combination with Gβ1 or Gβ2 heterotrimers, along with representative Gγ subunits. Furthermore, fluorescence lifetime imaging microscopy (FLIM-FRET) was used to investigate the nanoscale distribution of Gαi3-based heterotrimers in the plasma membrane, specifically with the dopamine D2 receptor (D2R). In addition, the functional response of the system was assessed by monitoring intracellular cAMP levels and conducting bioinformatics analysis to further characterize the heterotrimer complexes.

Results

Our results show that Gαi3 heterotrimers mainly localize to the plasma membrane, although the degree of colocalization is influenced by the accompanying Gβ and Gγ subunits. Heterotrimers containing Gβ2 showed slightly lower membrane localization compared to those containing Gβ1, but certain combinations, such as Gαi3β2γ8 and Gαi3β2γ10, deviated from this trend. Examination of the spatial arrangement of Gαi3 in relation to D2R and of changes in intracellular cAMP level showed that the strongest functional response is observed for those trimers for which the distance between the receptor and the Gα subunit is smallest, i.e. complexes containing Gβ1 and Gγ8 or Gγ10 subunit. Deprivation of Gαi3 lipid modifications resulted in a significant decrease in the amount of protein present in the cell membrane, but did not always affect intracellular cAMP levels.

Conclusion

Our studies show that the composition of G protein heterotrimers has a significant impact on the strength and specificity of GPCR-mediated signaling. Different heterotrimers may exhibit different conformations, which further affects the interactions of heterotrimers and GPCRs, as well as their interactions with membrane lipids. This study contributes to the understanding of the complex signaling mechanisms underlying GPCR-G-protein interactions and highlights the importance of the diversity of Gβ and Gγ subunits in G-protein signaling pathways.

Video Abstract

Background

G protein-coupled receptors (GPCRs), which are encoded by more than 800 genes in the human genome, constitute a huge group of eukaryotic transmembrane proteins. Their primary function is to transmit extracellular signals into the cell, allowing cells to communicate with each other and sense the extracellular environment. Unlike other types of receptors, GPCRs rely on interactions with G proteins, which further transmit signals to membrane-bound effectors. G proteins play a key role in providing the high flexibility, sensitivity, and specificity observed in GPCR signaling. Upon activation by various ligands, GPCRs can regulate a wide range of signaling pathways by engaging small G-proteins and heterotrimeric guanine nucleotide-binding proteins, referred to as G-proteins hereafter. The heterotrimeric G-protein complex consists of one of 16 Gα subunits, one of 5 Gβ subunits, and one of 12 Gγ subunits [1]. Considering that all possible combinations of these three subunits can likely form within cells, the number of potential heterotrimeric combinations is extensive. The diverse signaling properties of GPCRs account for the multitude of stimuli and cellular responses they are involved in. This diversity is achieved by the diversity in the varied composition of G protein heterotrimers, thereby contributing to the complexity of the human organism. However, it also gives rise to intriguing questions, such as how the appropriate receptor interacts with the specific G-protein trimer in the precise cellular context.

G protein heterotrimers function as regulatory GTP hydrolases. Based on the sequence similarities of the Gα subunit, they can be categorized into four groups: Gαs, Gαi/o, Gαq/11, and Gα12/13. Each family member within these groups acts on distinct signaling pathways, thereby modulating various physiological processes in response to extracellular stimuli. The Gαs family subunits stimulate adenylate cyclases (ACs), while different Gαi/o subunits exert an inhibitory effect on specific ACs. The Gαq/11 subfamily regulates phospholipase C activity, and Gα12/13 can interact with Rho nucleotide exchange factors [2]. In the GDP-bound conformation, the βγ dimer prevents the release of the nucleotide, thereby stabilizing the inactive form of the Gαβγ heterotrimer. When an activated GPCR receptor binds to it, a rapid release of GDP and its replacement by GTP is induced. This results in the dissociation of both the GPCR–Gαβγ complex and the rearrangement of the G-protein complex into the free Gα subunit and the Gβγ complex, which can independently activate different signal transduction pathways, leading to specific physiological effects. Once GTP is hydrolyzed to GDP by the Gα subunit, the now inactive Gα subunit can reassociate with the Gβγ dimer.

The diversity of signals sent into the cell following GPCR activation arises from the GPCRs’ capacity to activate multiple G proteins [3]. Most GPCRs interact with specific Gα subunits, which in turn determine distinct patterns of engagement with effector molecules [2]. However, GPCRs can activate any G protein, albeit with varying efficiencies [4]. The picture gets even more complicated when we consider that Gγ subunits regulate the spatial–temporal organization of G proteins. Since all Gβγ complexes can dissociate from the cell membrane, the membrane dissociation of Gβγ complexes serves as an additional mechanism to control their availability within the submembrane space. Consequently, cells expressing different levels of Gγ subunits exhibit distinct characteristics in their response to GPCR stimulation.

The existence of so many combinations of Gβγ complexes and their function in cells remains unexplained, as for many years this complex has been shadowed by research on the Gα subunit. But this is the Gβγ complex, which—upon heterotrimer activation—changes its localization and is responsible for effects in areas distant from the membrane. As it turns out, different combinations of Gβ and Gγ subunits can combine to form complexes with different affinities, the ability to move to distinct cellular compartments, and generate different outcomes [4]. So, it seems likely that the diversity in the composition of Gβγ complexes is responsible for the plasticity of GPCRs to generate intracellular responses. Transfer of the Gβγ complex from the plasma membrane determines signaling to intracellular organelles and can proceed in two ways. Rapidly dissociating farnesylated Gγ promptly diffuses into the endoplasmic reticulum (ER), Golgi apparatus, mitochondrion, and early endosomes. Geranylated Gγ subunits can only reach the ER and mitochondrion by slow diffusion, or with active transport to the early endosomes and Golgi [4]. The rate of translocation of human Gγ-subunits of G proteins from the plasma membrane to inner membranes has, incidentally, been used as the basis for one of two classifications of these proteins, according to which it distinguishes fast-moving subunits (about 10 s), subunits moving at an intermediate rate (about 60 s) and subunits moving slowly (more than 2 min) [5]. The second classification method, which we follow in the present study, assigns these proteins to five classes based on sequence homology, and so Class I includes Gγ1, Gγ9, and Gγ11, Class II: Gγ2, Gγ8, Gγ3, and Gγ4, Class III: Gγ7 and Gγ12, Class IV: Gγ5 and Gγ10 and Class V: Gγ13 [6].

The precise determinants of the interaction between receptors and G proteins are not yet fully elucidated. However, it is well-established that the second and third intracellular loops of the receptor, as well as the third, fifth, and sixth transmembrane helices, exhibit significant structural diversity among GPCRs. These regions are responsible for the selective binding of the Gα subunits [7]. On the G protein side, the C-terminal and N-terminal helices, along with two loops of the Gα subunit, are involved in the interaction with receptors. The interaction between the receptor immobilized in the membrane and the Gαi subunit is facilitated by lipid modifications that all G protein α-subunits undergo. Depending on the protein class, myristic and/or palmitic acid residues are attached. In the case of Gαi subunits, N-myristoylation, and S-palmitoylation take place. N-myristoylation is an irreversible process that anchors the subunit to the cell membrane. S-palmitoylation, on the other hand, is reversible and enables the regulation of protein accessibility to the receptor at different stages of protein transfer [8]. Additionally, all Gα subunits undergo co-translational irreversible N-terminal acylation. Furthermore, all Gγ subunits are modified at their C-terminus with a farnesyl or geranylgeranyl moiety [4]. Although Gα and Gγ subunits do not directly interact, their modified ends remain in close proximity and are responsible for the association of the complex with the membrane.

The Gβγ complex also binds to GPCRs, and this interaction stabilizes the GPCR-Gαβγ interface, presumably by placing Gα in a conformation suitable for receptor binding [9]. Experimental evidence has demonstrated the direct interaction of GPCRs with the C-terminus of the Gβ subunit and the farnesylated C-terminus of Gγ [10, 11]. Following activation, Gβγ dimers translocate to intracellular organelles to propagate the signaling from the cell membrane to these locations. The destination, kinetics, and efficiency of this translocation process are influenced by the specific Gγ subunit involved [4, 12, 13].

Recent evidence suggests that in addition to structural compatibility, the interaction between these proteins is influenced by various factors. These factors include the specific ligand used to activate the receptor, receptor oligomerization, and the lipid composition of the cell membrane [5, 14, 15]. This intricate signaling mechanism allows for diverse messages to be conveyed from the same receptor, leading to distinct signaling pathways and cellular responses. Different cell types exhibit variations in lipid composition and the expression of enzymes involved in lipid modification. Furthermore, within a single cell, different regions of the cell membrane can show heterogeneity in protein and lipid composition, including cholesterol and phospholipids, which can impact GPCR signaling [16, 17].

In the research described in the present studies, we focused on examining the impact of individual components of the heterotrimeric complex on the signaling of dopamine D2 receptors (D2R), which belong to the class A GPCRs. The stimulation of D2R leads to a reduction in cAMP levels through its interaction with Gαi/o class proteins. D2R is expressed in various brain regions, and it has been observed that Gαi subunits are also present in these regions without indicating regional specificity. According to the Human Protein Atlas, Gβ and Gγ subunits are also found in all brain structures without showing regional specificity, although their expression levels vary. Among the Gβ subunits, Gβ1 and Gβ2 have been reported to have the highest expression levels, while Gγ3, Gγ8, and Gγ7 are among the Gγ subunits with the highest expression levels (www.proteinatlas.org; [18]).

Using confocal microscopy and functional studies in living cells, as well as bioinformatics tools, we showed the composition of the heterocomplex, particularly the Gβ and Gγ subunits as well as lipid modifications. Our findings revealed that these factors are crucial in determining the localization of the heterocomplex within the cell membrane, the spatial orientation of the Gαi3 subunit relative to the dopamine D2R, and the cellular response upon receptor stimulation.

Materials and methods

Plasmid vectors and protein constructs

Plasmid pcDNA3.1+ vectors, namely GNB01, GNG01, GNG02, and GNG11, were obtained from the UMR cDNA Resource Center. Plasmids GNB02, GNG08, GNG09, and GNG13 were generously provided by Narasimhan Gautam (Addgene plasmids #42,182, #36,106, #64,203, #36,110). Plasmids GNG07, GNG10, and GNG12 were kindly provided by Catherine Berlot (Addgene plasmids #54,473, #55,192, #55,194) [5, 19]. Additionally, previously prepared and described plasmids were used in the studies described in this paper: pmCherry-N1DRD2, pcDNA3.1+ GNAI3Citrine, and pcDNA3.1+ GNASCitrine [20,21,22]. Plasmids containing GNAI3 and GNAS were modified to eliminate lipid moiety attachment through point mutations (G2A or C3A) using the QuikChange approach [23]. Sigma-Aldrich prepared starters (Poznań, Poland). The modified plasmids were confirmed by sequencing performed at Genomed (Warsaw, Poland).

Cell culture and transfection

HEK293 cells (ATCC, Manassas, VA, USA) were cultured following previously described methods. For microscopy experiments, the cells were cultured on sterile glass coverslips in 30 mm plates. For cAMP and RT-qPCR assays, the cells were cultured in six-well plates, for cAMP assays additionally coated with 0.5% gelatin [21]. Transient transfection of the cells was performed using the TransITX2® Dynamic Delivery System (Mirus Bio, Madison, WI, USA) according to the manufacturer’s instructions. The transfection involved the use of pcDNA3.1+ GNAI3Citrine or GNASCitrine plasmids, as well as plasmids encoding Gβ and Gγ proteins, and pmCherry-N1DRD2. The DNA amounts for Gα, Gβ, and Gγ were equimolar in all experiments. The DNA ratio of Gαβγ with the D2R (used for cAMP level measurements) was 1–1.25. The total amount of DNA used for cell imaging experiments was 0.1–0.3 μg per dish, while for cAMP level determination and real-time PCR analysis, it was 1.7 μg per well.

Cell membrane staining was performed using the Cell Plasma Membrane Staining Kit Deep Red Fluorescence Cytopainter (ab219942; Abcam, Cambridge, GB) immediately before live-cell imaging according to the manufacturer’s instructions. After staining, the cell culture was washed with sterile PBS (BioShop Canada Inc.) and maintained in Dulbecco’s Modified Eagle Medium (DMEM F-12) without phenol red (ThermoFisher Scientific, Inc., Waltham, MA, USA) supplemented with 2% FBS (Sigma-Aldrich, Poznań, Poland).

For ER visualization the CellLight™ ER-RFP BacMam 2.0 (C10591; Invitrogen ™, Thermo Fisher Scientific, Inc., Waltham, MA, USA) was used. Cell infection with the ER-RFP BacMam 2.0 was performed two days before microscopic observation, following the manufacturer’s protocol.

Live-cell imaging microscopy

Cell imaging was conducted using a Leica SP5 II SMD confocal microscope (Leica Microsystems, Mannheim, Germany) equipped with a Leica HCX Plan Apo 63 × lens (1.4 numerical aperture). The fluorescence signal was acquired in sequential line scanning mode using an argon ion laser (488 nm) for Citrine excitation and a laser diode (594 nm) for cell membrane and ER markers. The line average was three, and the scanning speed was 400 Hz. The fluorescence emission range was 495–570 nm for Citrine or 610–720 nm for Deep Red Fluorescence Cytopainter and CellLight™ ER-RFP. All imaging was performed on living cells in an air–steam cube incubator at 37 °C in DMEM F-12 without phenol red supplemented with 2% FBS.

Colocalization analysis

The Coloc2 plugin in ImageJ software was used for colocalization analysis. The entire cell displaying signals from both fluorophores was selected as the ROI in all images. Only images that met the Nyquist criterion were included in the analysis, with a voxel size of xy below 93 nm for Citrine and 110 nm for RFP/Deep Red fluorophore. Background subtraction was performed using the rolling ball method with a radius of 50 pixels. The point-spread function (PSF) was determined using the equation PSF = d/ “pixel spacing,” where d = λ/2NA. The “pixel spacing” represents the distance between the centers of two adjacent pixels, and NA refers to the objective numerical aperture. The Pearson Correlation coefficient was calculated using the default threshold for background correction. A total of 3–4 independent experiments were conducted, and images were collected for analysis.

FLIM-FRET measurements

Fluorescent lifetime imaging microscopy was conducted on cells transiently transfected with Gαi3 Citrine (donor) or Gαi3 Citrine and D2R mCherry (donor–acceptor) constructs. The FLIM acquisition was performed using a confocal laser scanning microscope (Leica SP5 II SMD, Germany) equipped with the PicoHarp 300 Time-Correlated Single Photon Counting (TCSPC) module (PicoQuant, Berlin, Germany). A pulsed laser diode at 470 nm (Leica, 40 MHz, Germany) was used to excite the fluorescence of the energy donor (Citrine). The emission was collected in the range of 500–550 nm using a bandpass filter, and an avalanche photodiode was employed for detection. The acquisition time for each measurement was 3–4 min, and the images were recorded in a 512 × 512 format.

Before measuring the Citrine fluorescence intensity decay, the fluorescence intensity levels of the fluorophores in cells expressing either the donor alone or the donor–acceptor pair were confirmed (n = 5). The analysis focused on the fluorescence signal originating exclusively from the plasma membrane, and a double exponential decay function was used for analysis in the SymPhoTime software (PicoQuant, Berlin, Germany). The reduction in fluorescence lifetime, observed as a result of FRET, was primarily evident in the short lifetime component (τ1), while the other component (τ2) remained relatively stable. This indicates that energy transfer occurred only from one donor species characterized by the lifetime τ1. Therefore, only the τ1 component was considered for calculating the FRET efficiency.

The FRET efficiency was determined using the equation: E = 1 − (τDA/τD), where τDA represents the lifetime of the donor in the presence of the acceptor, and τD represents the lifetime of the donor without the FRET acceptor.

cAMP levels measurements

The intracellular cAMP levels were determined using the cAMP ELISA chemiluminescent kit (STA-500, Cell Biolabs Inc, San Diego, CA, USA), following previously described methods [20]. In brief, cells transfected with vectors encoding Gαi3, Gβγ dimer combinations, and D2R were stimulated for 10 min in a medium containing 1 μM sumanirole maleate and phosphodiesterase inhibitors (Sigma-Aldrich, Poznań, Poland). Nontransfected cells that were stimulated served as control conditions for HEK293 cells, aiming to minimize the impact of endogenous G proteins and GPCRs. The concentration of cAMP was normalized to control values in each experiment. Samples were measured in duplicates (n = 4, or n = 1–2 for Gβ2 trimers).

RT-qPCR

The experiments were conducted following our previously published protocol [21]. HEK293 cells were prepared in a manner analogous to microscopy preparations. Total RNA was extracted using RNA Extractol reagent (EURx). Reverse transcription was performed with RevertAid reverse transcriptase (Thermo Scientific) according to the manufacturer's instructions, utilizing oligo(dT)16 primers (Genomed). The resulting cDNA was then utilized in a qPCR reaction using an Eco Real-Time PCR apparatus (Illumina). The reaction mixture consisted of 10 μl, including 5 μl of Luminaris Hi Green (Thermo Scientific), 1 μl of the resulting cDNA, and 0.3 μM specific primers (Table S1, Genomed). The reaction conditions were as follows: 95 °C for 10 min, followed by 40 cycles of 95 °C for 15 s, 60 °C for 30 s, and 72 °C for 30 s. Melting curves were analyzed for each pair of primers (95 °C for 15 s, 55 °C for 15 s, and temperature increase to 95 °C by 1 °C every 3 s). Transfections, RNA extractions, and RT-qPCR experiments were performed in three independent replicates. Each qPCR reaction was performed in duplicate. The expression levels of the β and γ subunits of G proteins were determined by analyzing the threshold cycle for each set of primers in material derived from transfected HEK293 cells, comparing it to nontransfected HEK293 cells. Data were analyzed using the ΔΔCq method with a Pfaffl modification [24] to account for PCR reaction efficiency. The results were presented as fold change in gene expression normalized against the GAPDH gene and compared to endogenous expression in the nontransfected control.

Bioinformatic analysis

The docking of the Gβ1γ1 or Gβ1γ2 dimer to Gαi3 or Gαs was performed using the HADDOCK 2.4 webserver [25, 26]. PDB files of Gαi3 (7E9H), Gαs (7F55), Gβ1γ1 (1TBG), and Gβ1γ2 (5UZ7) were prepared according to the software developers. Residual water, ions, ligands, double occupancies of amino acid residues (keeping only the first conformation), and other amino acid chains such as receptors or other partners present in the structure were removed using PyMol. For each trimer, two scenarios were analyzed: interaction between the N-helix of the Gα subunit and amino acids N88 and N89 of the Gβ subunit, or interaction between the N-helix and residues 205–215 of the Gα subunit and amino acids N88, N89, L117, D228, D246, and W332 of the Gβ subunit [27]. The results of all investigated arrangements were visually verified by comparison with published structures, and the model with the highest score was presented. The visual representation was prepared using Discovery Studio software, version 4.0 (BIOVIA, D. S., San Diego, CA, USA, 2015).

The human Gβ and Gγ protein sequences were aligned using Clustal Omega (CLUSTAL O(1.2.4) EMBL-EBI) [28, 29]. The amino acid sequences were obtained from the UniProtKB database: Gβ: P62873, P62879, P16520, Q9HAV0, O14775; Gγ: P63211, P59768, P63215, P50150, P63218, O60262, Q9UK08, O14610, P50151, P61952, Q9UBI6, Q9P2W3. The results are presented as a Percent Identity Matrix created by Clustal 12.1.

Statistical analysis

In all experiments, outliers were identified using Grubbs’s test and subsequently excluded from the analysis. The distribution of the data was assessed using the Shapiro–Wilk test, analysis of skewness and kurtosis, and the equality of variances was evaluated using Levene’s test. For data that followed a normal distribution, an unpaired t-test was conducted, and the results are presented as mean ± SEM. In cases where the data did not exhibit a normal distribution or when there were unequal sample sizes between groups, the Mann–Whitney U test was performed, and the values are reported as median ± MAD. The statistical analysis was performed using the Statistica program (data analysis software system), version 13 (TIBCO Software Inc., Palo Alto, CA, USA, 2017; http://statistica.io).

Results

The composition of the Gβγ dimer subunits influences the proportion of Gαi3 heterotrimers reaching the plasma membrane

We conducted experiments to characterize the preferences of Gαi3 protein for different Gβγ dimers formed by Gβ1 or Gβ2 and representative Gγ subunits from each of the five different families [6]. These include Gγ1, Gγ9, and Gγ11 from class I, Gγ2 and Gγ8 from class II, Gγ7 and Gγ12 from class III, Gγ5 and Gγ10 from class IV, and Gγ13 from class V. We tested all possible combinations of Gαi3 with the mentioned Gβ and Gγ subunits, except for Gαi3β2γ1 and Gαi3β1γ2 heterotrimers. It has been postulated that the Gβγ dimers in the first combination do not form a functional complex [30,31,32]. On the other hand, the Gαi3β1γ2 heterotrimer, which includes the extensively studied canonical Gβγ dimer, has already been characterized in terms of subcellular localization and functional role [20].

Previous studies have reported that all the combinations of Gβ and Gγ subunits tested in our study can form stable dimers in HEK293 cells [4]. We optimized the transfection conditions to ensure similar expression levels of all components of the heterotrimer, including Gαi3, Gβ, and Gγ subunits. Expression analysis using quantitative PCR for fluorescently unlabeled Gβ and Gγ subunits showed that the expression levels of Gβ1 and Gβ2 subunits in all tested combinations containing the receptor, Gαi3, Gβ1, or Gβ2, and the various Gγ subunits, were similar, with the relative quantity of mRNA oscillating around 1 (Fig. S1 in the supplementary information). The expression levels of Gγ subunits varied more due to the presence of different amounts of endogenous proteins in HEK293 cells. Our results are consistent with the data presented by Atwood et al. [33]. The relative amount of mRNA obtained in our experiment was higher when the endogenous mRNA of Gγ subunits was present in lower quantities in HEK293 cells. For almost all Gγ subunits, the relative mRNA amount exceeded that of Gβ subunits, since this cell line contained more endogenous mRNA of Gβ1 and Gβ2 subunits than Gγ. However, an exception was observed for Gγ10, whose mRNA was found to be present in the highest amount among endogenous mRNA in HEK293 cells. In this case, the result we obtained was less than zero, indicating that the expression levels we are working with are comparable to the endogenous levels.

Following the expression level examination of the Gαi3-based heterotrimers, we proceeded to investigate their subcellular localization. In some cases, we also assessed the inhibitory potential of intracellular cAMP concentration upon D2R activation. To begin our studies, we examined the ability of all the investigated Gαi3βγ heterotrimers to target the plasma membrane. We evaluated the colocalization of Citrine-tagged Gαi3 with the plasma membrane marker, Deep Red. The use of Deep Red fluorophore, which exhibits a significant redshift, reduced the likelihood of signal overlap between the two fluorophores. We confirmed this by imaging HEK293 cells labeled only with Deep Red, where no signal was detected in the wavelength range specific to Citrine fluorescence. Representative images for the investigated combinations are presented in Figs. 1A-B and S2. We also analyzed the colocalization between the Gαi3 overexpressed alone and the plasma membrane marker as a reference for the Gαi3βγ combinations. The colocalization results between Deep Red and the Citrine-labeled Gαi3βγ are shown in Fig. 2A-B, presenting only the Pearson’s correlation coefficient values due to their insensitivity to variations in pixel intensity and offset settings between images [34].

Fig. 1
figure 1

Cells transiently transfected with Gαi3-Citrine variants show different intracellular localization. Cells were stained with Cytopainter membrane Deep Red fluorophore (A, B) or CellLight™ ER-RFP (C). The scatter plot at the lower panel facilitated PCC estimation with linear regression, where the y-axis shows pixels in the Deep Red channel and the x-axis is the Citrine channel. The scale bar corresponds to 5 μm

Fig. 2
figure 2

Colocalization analysis of the wild type Gαi3, Gαs and their muteins with various Gβγ dimers with the cell membrane or ER represented as Pearson correlation coefficient (PCC). A-E HEK293 cells transfected with Gαi3/Gαs-Citrine encoding vector alone or together with different Gβγ vectors without fluorescent protein were imaged after cell membrane staining (Deep Red dye). Images were collected on living cells subsequently for Citrine and Deep Red dye fluorescence. Cells transfected only with Gαi3/Gαs-Citrine vector represent the membrane localization of the Gαi3/Gαs without overexpression of the Gβγ dimer. F Similarly, colocalization of the Gαi3 C3A-Citrine with ER was performed. ER was stained using the CellLight™ system, and images were acquired two days after infection. Data were collected from at least two independent experiments (n = 2) and are presented as mean ± 95% confidence interval (CI). All obtained mean PCC were compared with the WT Gαi3(A-D)/Gαs (E) PCC with unpaired t-test (black color, * – p < 0.05, ** – p < 0.01, *** – p < 0.005, **** – p < 0.001).Additionally differences in the PCC values between Gαi3 muteins G2A vs C3A (red color ****—p < 0.001) or Gαi3 C3A β1γ2 vs Gαi3β1γ8 (black color) all with unpaired t-test (F)

Live-cell imaging of the investigated combinations of Gαi3 with Gβ and Gγ confirmed that all heterotrimers reached the plasma membrane. However, we observed that the fraction of proteins reaching the plasma membrane depended on the subunit composition (Fig. S2). Interestingly, the degree of colocalization between Deep Red and Gαi3 in the heterotrimeric form was not significantly different from that observed for Gαi3 overexpressed alone. Specifically, for the complexes containing Gβ2, the membrane-bound fraction was slightly lower than for Gβ1-containing heterotrimers and Gαi3 only. This was particularly evident for Gαi3β2γ2 and Gαi3β2γ7 (p < 0.005). The estimated PCC values for Gβ1 heterotrimers were comparable to those of the Gαi3 when overexpressed alone, indicating similar levels of membrane localization. However, it is worth noting that this pattern was not observed for Gγ8 and Gγ10, which belong to class II and IV of Gγ, respectively (Figs. 2 and S2). In fact, their heterotrimers with Gβ1 showed slightly poorer membrane localization (p < 0.05) compared to other Gαi3β1γ combinations. Conversely, the Gβ2 complexes with these Gγ subunits exhibited colocalization with Deep Red on the plasma membrane at the same level as the Gαi3 overexpressed alone. A portion of the Gαi3 protein is probably complexed with endogenous Gβγ proteins. However, looking at the fold change in expression levels of Gβ and Gγ subunits (Fig. S1), as well as fold change in Gαi3 mRNA [21], we can surmise that this constitutes only a small fraction. Nevertheless, we do not underestimate their potential contribution, and we compare the results obtained after transfecting cells with all components of the heterocomplex to the results obtained after transfection with only the plasmid containing the Gαi3 coding sequence. Overall, these results suggest that all Gαi3β2γ complexes, except those composed of Gγ8 and Gγ10, bind to the plasma membrane with slightly lower efficiency compared to Gαi3β1γ’s.

Effect of various Gβγ complexes on cellular localization of Gαi3βγ heterotrimers

To explore the effects of the Gβγ dimer on the membrane localization of Gαi3, we generated modified versions of Citrine-labeled Gαi3 by introducing specific mutations to disrupt its lipidation. Two point mutations were introduced to replace critical amino acid residues involved in lipid attachment with alanine (G2A or C3A), resulting in the elimination of lipid moieties. Previous research has indicated that myristoylation is a crucial step for the palmitoylation of the Gαi/o subunit [35, 36]. Specifically, the G2A mutation has been shown to block the attachment of a myristic moiety by NMT transferases to the N-terminal glycine of Gα, when myristoylation is eliminated, both myristoylation and palmitoylation are expected to be removed [36,37,38]. However, it's worth noting that myristoylation may not always be an absolute requirement for palmitoylation, as demonstrated in certain cases. The presence of the Gβγ dimer or the presence of polybasic motifs at the N-terminus of Gα can effectively support the association of Gα subunits with the cell membrane, even when myristoylation is absent, and in some cases, it can even restore palmitoylation [35, 39, 40].

The Gαs subunit was employed as a reference protein, as extensive studies have examined the roles of lipidation and the polybasic region in its cellular membrane localization [41, 42]. The corresponding lipidation-disrupting mutations were introduced in the Gαs subunit labelled with Citrine. As depicted in Fig. 2C and D and Figs. S3–S5, all four mutant proteins (Gαi3 C3A, Gαi3 G2A, Gαs C3A, and Gαs G2A) exhibited significantly weakened plasma membrane localization, as indicated by PCC values ranging from 0.08 to 0.2. This effect was particularly prominent in the case of Gαi3, where the wild-type protein exhibited a mean PCC value of 0.55, whereas both mutants displayed significantly reduced plasma membrane localization.

Interestingly, a notable difference (p < 7.3E − 6) was observed between the mutant lacking myristoylation (Gαi3 G2A) and those lacking palmitoylation (Gαi3 C3A). The mutant lacking myristoylation exhibited a lower proportion of proteins reaching the plasma membrane compared to the mutants lacking palmitoylation alone. This finding supports the widely accepted hypothesis that myristoylation alone is insufficient for attaching Gαi/o proteins to the membrane. Although this study cannot confirm the absence of palmitoylation in the Gαi3 G2A mutant, the poorer membrane localization observed suggests the potential absence of both lipid anchors. However, previous studies by Degtyarev et al. have demonstrated the presence of Gαi1 G2A in the membrane fraction and the incorporation of [3H]palmitate when Gβ1γ2 is present [39]. Therefore, the small portion of Gαi3 G2A mutant observed in the plasma membrane may be attributed to the formation of heterotrimers with endogenous Gβγ dimers in HEK293 cells and subsequent palmitoylation.

In the case of Gαs (Figs. 2E and S5), even the wild-type protein exhibited weaker plasma membrane localization compared to Gαi3, with a PCC of 0.3. Nonetheless, impaired membrane docking was observed for both Gαs mutants (Gαs C3A and Gαs G2A), with p < 0.05 for Gαs C3A and p < 0.005 for Gαs G2A.

In summary, removing lipidations from Gαi3 cause more significant changes in this protein population residing at the plasma membrane, as compared to Gαs. Gαi3 exhibits stronger membrane localization than Gαs, as evidenced by the difference in PCC values between the wild-type proteins. Cotransfection of Gαi3 muteins with selected Gβ1γ or Gβ2γ dimers resulted in variations in membrane docking. Nevertheless, as shown in Figs. 2C, D, S3, and S4 none of the studied heterotrimers showed a significant improvement in the membrane-bound fraction compared to the Gαi3 G2A and Gαi3 C3A subunits when expressed alone, with comparable PCC values. The heterotrimers Gαi3 C3A with Gβ2γ2 and Gβ2γ8, as well as myristoylation-deficient Gαi3 G2A with Gβ2γ2, displayed slightly higher membrane-bound fractions, but the differences were not substantial. In contrast, in the case of Gαs heterotrimers, the PCC values substantially increased from 0.3 for Gαs alone to 0.5 for Gαsβ1γ2. Similarly, the muteins of Gαs with impaired lipidation, especially the N-terminal palmitoylation-deficient Gαs G2A, showed restored membrane localization when cotransfected with Gβ1γ2, as indicated by a PCC of 0.43.

The myristoylation-deficient Gαi3 G2A mutant, when overexpressed alone or in combination with various Gβγ dimers, showed a predominant random localization in the cytosol (Fig. S3). In contrast, the palmitoylation-deficient Gαi3 C3A mutant exhibited specific subcellular localization, particularly in the ER, when cotransfected with certain Gβ1-containing dimers (Fig. S4). Figures 1C and 2F depict the perinuclear localization of Citrine-labelled Gαi3 C3A with Gβ1γ2, Gβ1γ8, or Gβ1γ11, which colocalized with the ER marker CellLight™ ER-RFP. Live cell imaging confirmed this ER localization and allowed for the calculation of PCC values, which indicated a high colocalization between Gαi3 C3A and the ER with the aforementioned Gβ1γ dimers (PCC 0.42–0.5). Notably, this ER localization was not observed when Gβ2-containing heterotrimers were coexpressed with Gαi3 C3A (Fig. S3A).

Effect of different Gβγ dimers on intracellular cAMP levels induced by dopamine D2 receptor activation

To assess the functional insights, we evaluated the ability of selected Gαi3βγ heterotrimers to transmit GPCR signals at the plasma membrane by evaluating their inhibitory influence on cyclic AMP production in response to stimulation of the dopamine D2R by sumanirole. We specifically focused on four Gαi3β1-containing heterotrimers with Gγ2, Gγ8, Gγ9, or Gγ10, as well as two heterotrimers containing Gαi3β2 with Gγ8 or Gγ10. We noticed noteworthy differences not only between complexes that differed in the Gβ subunit but also those that differed solely in the Gγ subunit. Interestingly, cells expressing Gαi3β2γ8 or Gαi3β2γ10 heterotrimers did not show significantly greater activity compared to control cells expressing only endogenous proteins, indicating that the interaction of these heterotrimers with the D2R can be neglected. As shown in Fig. 3A, Gβ1γ8, and Gβ1γ10 heterotrimers exhibited even more efficient reduction of intracellular cAMP levels compared to the canonical Gβ1γ2 dimer (p < 0.001). In contrast, Gαi3β1γ9 displayed the least effective ACs inhibition among the examined Gβ1-containing heterotrimers. The inhibition of ACs achieved by Gαi3β1γ8, Gαi3β1γ10, Gαi3β1γ2, and Gαi3β1γ9 was 11.9, 14.4, 20, and 70.4% of the control, respectively. Notably, complexes containing Gγ8 and Gγ10 with the β1 subunit showed significantly higher inhibitory responses compared to the corresponding complexes containing the β2 subunit, despite having slightly lower membrane localization. These findings suggest that the dopamine D2R exhibits varying preferences for heterotrimers containing the same Gα subunit but different Gβγ dimers.

Fig. 3
figure 3

Intracellular cAMP level in the HEK293 cells lysates overexpressing Gαi3 with different Gβγ and for Gαi3 muteins after stimulation of the D2 receptor with full agonist (sumanirole). HEK293 cells transiently transfected with plasmids encoding different variants of the Gαi3-Citrine, selected Gβ and Gγ subunits and D2R-mCherry receptor were stimulated with sumanirole. The intracellular cAMP levels were determined in the cell lysates. Results are presented as a percentage of cAMP levels in control (stimulated nontransfected cells), as a mean ± 95% confidence interval (CI). Experiments were performed in duplicates n = 4, except for Gαi3β2γ8 (n = 2) and Gαi3β2γ10 (n = 1). An unpaired t-test was performed to evaluate differences between Gαi3β1γ2 vs other Gβγ dimers as well as between wild type of Gαi3 and muteins without lipidations in the presence of the Gβ1γ8 dimer (** – p < 0.01, *** – p < 0.005, **** – p < 0.001)

Furthermore, for the set of Gβγ subunits most effectively inhibiting cAMP accumulation, that is Gβ1γ8, we investigated the effect of mutations that prevent lipidation of Gαi3. We found that the Gαi3 G2A mutant (Fig. 3), which lacks myristoylation, exhibited a weaker response in inhibiting ACs compared to the wild-type Gαi3 (p < 0.01) and the Gαi3 C3A mutant. Interestingly, the palmitoylation-deficient Gαi3 C3A mutant showed equal effectiveness in reducing intracellular cAMP levels compared to the wild-type protein, and no impairment was observed. This finding is surprising considering that the membrane localization of heterotrimers containing the lipid-deficient Gαi3 mutants was significantly lower than that of the wild-type protein. Additionally, a substantial portion of the palmitoylation-deficient Gαi3 C3A mutant was found to be retained in the ER. Despite these limitations, the palmitoylation-deficient Gαi3 C3A mutant still exhibited an effective reduction of intracellular cAMP levels when complexed with Gβ1γ8 dimers.

Different Gαi3βγ heterotrimers show differences in nanoscale distribution in the plasma membrane

Although colocalization analysis enables the investigation of interactions of the studied macromolecules in their cellular context, it does not provide the spatial resolution required to evaluate the precise nanoscale arrangement of molecules within the plasma membrane. Therefore, in order to further investigate and verify the differences observed in the colocalization analysis, we employed the FLIM-FRET method. This technique allowed us to study the membrane organization of selected Gαi3βγ heterotrimers relative to the D2R and determine the distances between them within the plasma membrane.

In this experimental setup, we monitored the fluorescence lifetime of the donor fluorophore (Citrine-labeled Gαi3) in the absence and presence of the acceptor (mCherry labeled D2R), as described in the Materials and methods section (FLIM-FRET measurements). HEK293 cells were cotransfected with the corresponding Gβ and Gγ subunits to reproduce a fully functional system. We previously confirmed the appropriate cellular localization and activity of the mCherry-labeled dopamine D2R [20, 21]. We focused on four heterotrimers: Gαi3β2γ8, which showed a high level of colocalization with the plasma membrane marker; Gαi3β1γ8 and Gαi3β1γ9, which showed the lowest level of colocalization; and the canonical trimer containing the Gβ1γ2 dimer. The decrease in the fluorescence lifetime of the donor fluorophore indicates the close proximity of proteins when both the energy donor and acceptor are present in the system.

The FLIM images collected from cells cotransfected with Gαi3βγ and D2 showed a reduction in the apparent fluorescence lifetime of the donor (depicted as change in color toward the blue hues across all pixels), compared to those expressing only Citrine-labelled Gαi3βγ or Gαi3 (Fig. 4B). Figure 4A presents the fluorescence lifetime distributions of the fluorescence donor for all tested arrangements, including Gαi3 alone or in the presence of various Gβγ complexes, with or without the dopamine D2R. The minimal reduction in donor lifetime in FRET pairs was observed for Gαi3β2γ8 or Gαi3β1γ9 with D2R. Importantly, these changes were not significantly different from those observed between the Gαi3 was overexpressed with D2R only, indicating a greater distance between these heterotrimers and D2R compared to the other two heterotrimers and D2R. The calculated efficiencies of energy transfer (Fig. 4C) indicate that the spatial distribution of closely related Gαi3 heterotrimers differs. If the Gβ1γ2 or Gβ1γ8 dimer is present in the heterotrimer, the FRET signal is more pronounced, suggesting that Gαi3 is located in closer proximity to D2R within the membrane compared to its complex with Gβ2γ8 (where only the Gβ subunit is changed) or Gβ1γ9 dimer. Notably, these FRET results align with the measurements of intracellular cAMP levels, where the inhibition of ACs was most effective for Gαi3β1γ2 or Gαi3β1γ8. Lower FRET efficiency was observed for Gαi3 heterotrimers that showed less effective ACs inhibition.

Fig. 4
figure 4

FLIM-FRET results. A Fluorescence lifetimes for the donor (Gαi3-Citrine) and the donor in the presence of the acceptor (D2R-mCherry). Data were collected from at least five independent experiments (n = 5) and are presented as mean ± 95% confidence interval (CI). The mean fluorescence lifetime of donor Gαi3-Citrine was compared with the donor fluorescence lifetime in the presence of the acceptor D2R-mCherry with an unpaired t-test (**** – p < 0.001). Because of the varying number of repeats, the median donor lifetime in the Gαi3βγ setup was compared to donor–acceptor systems with different Gβγ dimers using the Mann–Whitney U test (**** – p < 0.001). B Representative fluorescence lifetime images for one of the FRET donor (Gαi3-Citrine β1γ8; B1) and FRET donor–acceptor pair (Gαi3-Citrine β1γ8 – D2R-mCherry; B2). C Energy transfer efficiency (% E) between investigated Gαi3-Citrine – D2R–mCherry pairs; error estimated with exact differential

One of the key findings of this study is that the nanoscale arrangement of G proteins within the plasma membrane influences the effectiveness of signal transduction. The distribution of these proteins within the membrane is more significant than the overall amount of heterotrimer bound to the membrane. Despite the Gαi3β1γ8 complex showing the lowest degree of colocalization with the plasma membrane marker among the Gβ1-containing heterotrimers, it was found to be the most effective in reducing intracellular cAMP concentration. FRET measurements revealed that, for this complex, the distance between the receptor and the Gαi3 subunit was the shortest, indicating a close spatial arrangement that facilitates efficient signal transduction.

The Gβγ dimer affects the conformation of the entire heterotrimer complex

The observed differences in the interaction of heterotrimers with the plasma membrane or D2R can be attributed to variations in the individual subunits of the G protein complex. While the Gβ1 and Gβ2 subunits share a high similarity of 90.3% in their amino acid sequences, the significance of the Gβγ dimer becomes even more apparent when comparing the similarities between different Gγ subunits. These Gγ subunits not only differ in their attached lipid anchor but also in their amino acid sequences, particularly in the N-terminal helix region. Sequence analysis of Gγ subunits reveals a greater degree of diversity among the representatives compared to the relatively higher similarity observed in the case of Gβ subunits (Fig. S6B).

The three classes of Gγ subunits: class I (Gγ1, Gγ9, and Gγ11), class II (Gγ2, Gγ3, Gγ4, and Gγ8), and class III (Gγ7 and Gγ12) exhibit similarities of around 70% within their respective groups (except for Gγ3 in class II). However, for class IV, containing Gγ5 and Gγ10, the similarity drops to around 53%. Furthermore, the sequence of Gγ13 significantly differs from that of other Gγ subunits, and the similarity of class I (Gγ1, Gγ9, and Gγ11) to other classes is also relatively low.

To identify potential binding sites and determine the binding affinities of the docked poses of Gαi3 with Gβ1, we utilized HADDOCK, a molecular docking prediction server. Additionally, we performed a comparative analysis by docking the Gαs subunit. We examined four different complexes: Gαi3β1γ2, Gαi3β1γ1, Gαsβ1γ2, and Gαsβ1γ1. Our analysis focused solely on the amino acid residues within the Gβ subunit, excluding any residues within Gγ that could potentially contribute to heterotrimer formation. The selection of Gβ interface residues was based on mutational analysis data of Gβ1 [27]. Specifically, residues N88 and K89 were chosen due to their proximity to the Gα N-helix, while residues L117, D228, D246, and W332 were selected for their proximity to the helical fragment within the Gα helical domain (Fig. 5). We conducted two docking procedures, targeting either the N-helix of Gα that participates in the interaction with Gβ residues N88 and K89, or the entire sequence of Gα along with Gβ residues N88, K89, L117, D228, D246, and W332. Interestingly, the docking analysis of Gαs and Gαi3 complexes with Gβ1γ2 and Gβ1γ1 revealed slight differences in the orientation of Gβ1 between each complex. The best scoring pose for each complex is shown in Fig. S6A.

Fig. 5
figure 5

Representation of Gαsβ1γ2 docking results for interactions with all proposed Gβ1 binding sites. Gαs is shown as a molecular surface colored based on the hydrophobicity of the amino acid residues, the polar residues are colored blue and the hydrophobic residues are colored brown. The Gβ1γ2 dimer is colored green with selected active Gβ1 residues marked in red

Table 1 presents HADDOCK scores indicating a stronger interaction between Gαs and Gβ1γ2 compared to Gαs and Gβ1γ1 (179.5 ± 9.8 for Gβ1γ2 vs 122.8 ± 9.3 for Gβ1γ1). Additionally, other important docking evaluation parameters such as cluster size, RMSD, and Z score also favor Gβ1γ2 (Table S2). While the docking analysis of both full-length Gα and the N-terminal helix alone suggests the significance of both Gα interfaces in forming a complex with Gβ1, our findings suggest that the N-helix of Gαsβ1γ2 plays a more prominent role in the interaction. Interestingly, this is not observed in the case of the Gαsβ1γ1 complex, where both surfaces are similarly involved in the interaction. However, for the Gαsβ1γ2 heterotrimer, the interaction between the Gα subunit and the Gβγ dimer is the strongest (HADDOCK score 179.5 ± 9.8) and preferred over the Gβ1γ1 dimer (HADDOCK score 122.8 ± 9.3). Furthermore, our results indicate that the Gαsβ1γ2 heterotrimer is slightly more stable and preferable than Gαi3β1γ2 (HADDOCK score 141.5 ± 10.2). However, the analysis of the docking in the case of Gαi3β1γ2 or Gαi3β1γ1 (HADDOCK score 138.2 ± 8.2) shows negligible preferences between these two dimers for Gαi3. This may be due to weaker preferences of Gαi3 towards a specific Gβγ dimer or the possibility that the Gβγ dimers used in the docking analysis may not be the most suitable choice. The analysis presented in this study highlights the potential of computational tools like AlphaFold to predict new Gβγ dimer structures for use in docking studies. These results suggest that despite considering only the amino acid residues of Gβ, the results for the Gβ1γ1 and Gβ1γ2 dimers differ. Overall, further investigation is required to fully comprehend the preferences of Gαi3 towards different Gβγ dimers and the potential implications of these preferences on G protein signaling.

Table 1 Docking results of trimer formation for Gαs and Gαi3 with Gβ1γ2 or Gβ1γ1 dimers

Discussion

The precise mechanisms underlying the signaling diversity of GPCRs are not yet fully understood. One important contributing factor is the ability of GPCRs to interact with different types of G proteins, which can activate distinct downstream signaling pathways. Additionally, the existence of multiple subtypes of G protein subunits further enhances signaling diversity. However, whether GPCRs exhibit a preference for specific Gβ and Gγ subunits has not been extensively investigated.

The specificity of GPCRs for G proteins is primarily determined by the Gα subunit, but multiple isoforms of both Gβ and Gγ subunits add to the diversity of signaling. The question of whether GPCRs demonstrate a preference for particular Gβ and Gγ subunits remains unanswered. Emerging evidence suggests that GPCRs exhibit unique specificity for G proteins, not only favoring specific Gα subunits but also specific Gβγ dimers [4, 43, 44]. As previously reported, heterotrimers composed of Gαi3β1γ9 exhibited less efficient modulation of basal cAMP levels in HEK293 cells upon D2R activation compared to Gαi3β1γ2 [20]. The diversity of heterotrimer compositions is vast due to the abundance of proteins in each subunit group, allowing for numerous Gαβγ combinations in theory. However, not all heterotrimers are physiologically relevant, likely due to variations in expression levels or tissue localization, and the reduced stability and proven dissociation of some Gα and Gβγ combinations over extended periods of time [45, 46]. Despite recent discoveries, data on many G protein heterotrimers is still lacking. Recent studies have demonstrated that Gβ and Gγ exhibit certain preferences in forming Gβγ complexes, and the composition of the Gβγ dimer influences the kinetics and efficacy of GPCR responses [4]. As expected, each Gα subunit also demonstrates a preference for different Gβγ dimers.

In this study, we present, to our knowledge, the first comprehensive analysis of the cellular localization of Gαi3-based heterotrimers formed by either Gβ1 or Gβ2 in combination with representative Gγ subunits, as well as their interactions with the dopamine D2R. Initially, we conducted a direct comparison of the subcellular localization of Gαi3-based heterotrimers using the two most predominant Gβ types in cells, namely Gβ1 and Gβ2, along with several Gγ subunits from different classes. This approach was chosen regardless of downstream signaling, as it can be challenging to reliably compare all heterotrimers based solely on downstream signals.

To assess the subcellular localization of the studied heterotrimers, we utilize live cell fluorescence imaging and colocalization strategies. In order to generate heterotrimers with a specific subunit composition, we utilized the strategy of overexpressing the complex components in HEK293 cells. It is important to note that we ensured the level of overproduction was not excessively high but sufficient to preferentially produce heterotrimers with the desired composition. G proteins are typically transported to the plasma membrane as full heterotrimeric complexes, which assemble on the cytoplasmic surface of the Golgi apparatus or ER, and it is widely accepted that heterotrimer assembly occurs before the acylation of the Gα subunit [47, 48]. Previous studies have indicated that all combinations of the analyzed Gβ and Gγ subunits interact with other members of the Gαi/o family, particularly the Gαi1 subunit [45]. Furthermore, Gβ subunits demonstrate distinct preferences for Gγ subunits. Additionally, certain combinations of Gβγ dimers have been observed to be relatively less stable and tend to dissociate into their individual Gβ and Gγ components when isolated from the cellular environment [4, 45, 49, 50].

Our data reveals that all examined Gαi3 heterotrimers primarily localize at the plasma membrane, with minimal fluorescence signal detected from within the cell. However, we observed a slightly lower fraction of Gβ2-containing complexes bound to the plasma membrane compared to Gβ1, despite their highly similar structure (with a sequence identity of 90%) and generally interchangeable molecular interactions [51]. Previous studies have demonstrated that Gβ1 can interact equally well with most Gγ subunits in the presence of exogenous GαoA. On the other hand, Gβ2 exhibits greater selectivity and significant differences in binding to specific Gγ subunits, including Gγ1, Gγ7, Gγ8, and Gγ11–13 [4]. Another study utilizing the yeast two-hybrid assay showed that the examined Gγ subunits (Gγ1–5 and Gγ7) could interact with both Gβ1 and Gβ2, albeit the interaction with Gβ2 was relatively weaker [52].

The reduced plasma membrane localization of heterotrimers composed of Gβ2 observed in our experiments could potentially be attributed to differences in their interaction capabilities, with a few exceptions. For instance, the Gαi3β2γ8 and Gαi3β2γ10 heterotrimers exhibited greater efficiency in binding to the plasma membrane, even when compared to their complexes with Gβ1. Conversely, combinations such as Gαi3β2γ2, Gαi3β2γ7, and Gαi3β2γ13 showed the lowest membrane localization. Interestingly, these Gγ subunits belong to different classes and are all geranylgeranylated at the C-terminus. Recent demonstrations have shown that the GαoA subunit exhibits the strongest binding affinity for Gβ2γ7 and Gβ2γ8 while exhibiting weaker binding to Gβ2γ1, Gβ2γ11, and Gβ2γ13 [4]. Previous studies have reported that Gβ5γ dimers can distinguish between different defined Gα subunits [4, 53,54,55]. For example, they interact more preferentially with the Gαi1 subunit and exhibit weaker binding to Gαs [55]. Conversely, the Gβ1γ2 and Gβ1γ9 dimers preferentially bind to Gαi3 over Gαs [20].

The interactions between Gα and Gβγ primarily occur through two interface regions [56, 57]. The first interface is formed between the top of the Gβ propeller and Switch I and Switch II of Gα, while the second interface is located between blade 1 of the Gβ propeller and the N-terminus helix of Gα. Crystal structures of G-protein heterotrimers and Gβγ alone indicate that Gβγ subunits do not undergo significant conformational changes upon heterotrimer formation, but the available molecular data is somewhat limited, as crystal structures of G proteins have only included combinations of Gβ1γ1 or Gβ1γ2 dimers. The contact interface between Gγ and Gβ subunits is located in regions composed of residues that are generally highly conserved in both proteins [58].

Our analysis of docking heterotrimers composed of Gαs or Gαi3 with Gβ1γ2 or Gβ1γ1 revealed that the orientation of Gβ1 in heterotrimers can differ. Additionally, HADDOCK docking parameters suggest differences in the binding affinity of Gα subunits to Gβγ dimers. While Gα subunits exhibit a high degree of sequence and structural conservation, there are minor differences in the Gβγ contact interface regions between them. Gαi3 and Gαs differ primarily in their N-terminus helix sequence, with some differences also present within the conserved Switch I and Switch II regions. Consequently, these amino acid residue differences within the interface regions of both Gα and Gβ subunits may potentially impact their interaction. Unlike Gβγ, Gα undergoes structural changes upon heterotrimer formation. The binding of Gβγ induces a rearrangement of Switch II and induces conformational changes within Switch I and Switch II [59]. Furthermore, in the heterotrimeric state, the N-terminal helix of Gα remains intact, as it is stabilized by interactions with Gβ [56, 60].

The G protein complex is targeted to the plasma membrane through multiple binding signals that act synergistically to achieve high affinity and specificity for interaction with membrane lipids. These signals involve a combination of covalent lipid modifications on both the Gγ and Gα subunits, as well as the presence of positively charged residues in the C-terminus (preCaaX region) of Gγ and the N-terminus of Gα [5, 21, 61]. The prenylation of Gγ subunits, along with the polybasic motif, plays a crucial role in directing to the membrane localization and dissociation of the Gβγ dimer as well as facilitating the interaction with Gα and downstream effectors [60, 62, 63]. Similarly, the myristoylation of Gα enhances its affinity to Gβγ [64, 65]. Previous studies have demonstrated that Gα is necessary for the plasma membrane localization of Gβγ. When different combinations of Gβ and Gγ are overexpressed without Gα, the Gβγ dimers exhibit weak localization at the plasma membrane and tend to accumulate in intracellular structures, predominantly in the ER [66,67,68].

The observation that even Gαi3 protein overexpressed alone localizes to the plasma membrane with similar efficiency as its heterotrimers (Figs. 2 and S2) supports the notion that Gαi3 plays a crucial role in directing the complete heterotrimeric complex to the membrane. This is further supported by the finding that heterotrimers containing the Gαi3 G2A mutein, which is resistant to lipid modification, are predominantly retained in the cytosol unless coexpressed with Gβ2γ2, which partially restores membrane localization. This suggests that the myristoylation of Gαi3 may not be as critical for the association of this specific heterotrimer as it is for others. It is still possible that palmitoylation of the Gαi3 G2A mutant occurs after complex formation, as previously observed in the case of Gαi1 G2A Gβ1γ2 [39]. However, the results obtained with the palmitoylation-deficient Gαi3 C3A mutant, which displayed a similar level of membrane association, lack conclusive evidence due to the significant dispersion of the PCC value. Nevertheless, it can be reasonably concluded that the localization of this heterotrimer to the plasma membrane relies on protein–protein interactions. Among the tested combinations of Gβγ dimers, only two Gβ2γ dimers, including Gγ2 and Gγ8 from the same class, were effective at rescuing the membrane localization of the Gαi3 C3A mutein. Importantly, none of the tested dimers containing Gβ1 were able to restore the membrane localization of the heterotrimer with the palmitoylation-deficient Gαi3. In summary, these findings suggest that palmitoylation is essential for the stable binding of Gαi3 heterotrimers to the plasma membrane, as previously proposed [38, 69, 70]. However, the significance of protein–protein interactions should not be overlooked. Overall, Gαi3 appears to be the driving force behind the effective targeting of Gαi3 heterotrimers to the plasma membrane, resulting from the combined effects of fatty acid modification and protein–protein interactions.

The inhibition of Gαi3 palmitoylation leads to the accumulation of a significant fraction of Gαi3 C3A in the ER when certain Gβ1γ combinations are present. This observation is intriguing because overexpressed Gαi3 C3A mutein alone typically exhibits random localization in the cytosol. These findings align with previous research showing that mutations abolishing palmitoylation cause Gα subunits to shift to the cytosolic fraction, as observed in the cases of Gαo, Gαz, and Gαi1 [38, 71, 72]. Another study demonstrated that when palmitoylation of Gαi2 is prevented, either through mutation of the palmitoylated cysteine residue to serine or treatment with 2-bromopalmitate (2-BP), the Gβ1γ2 dimer accumulates in the ER while the heterotrimer remains in the Golgi [67].

The trafficking pathway for G proteins involves the preassembly of heterotrimers before palmitoylation of the Gα subunit, followed by delivery of the complex to the plasma membrane [73]. A conserved family of enzymes known as DHHC-Cysteine Rich Domain (DHHC-CRD) enzymes has been identified as at least one of the enzymes responsible for the S-palmitoylation of Gα proteins. Most DHHCs localize at the Golgi, with some distributed among the ER, plasma membrane, and endosomes [74]. Previous studies have shown that DHHC3 and 7 enzymes acylate Gαq, Gαs, Gαi2 and Gαo subunits mainly in the Golgi [37, 75]. However, other studies have shown that DHHC3 and 7 knockout, as well as DHHC7 overexpression, had only a minor effect on Gαi1-3 acylation [76]. Instead, it was suggested that the acylation of these subunits may be mediated by different acyltransferases, likely localized outside the Golgi apparatus. Our experiments revealed that Gβ1γ2, Gβ1γ8, and Gβ1γ11 were capable of retaining palmitoylation-deficient Gαi3 C3A in the ER, indicating the involvement of ER-localized DHHC acyltransferases in the palmitoylation of the Gαi3 subunit. These findings suggest that multiple enzymes redundantly participate in the acylation of the Gαi family. Although S-palmitoylation is generally considered nonspecific and proximity-based, recent studies have shown that DHHCs can differentiate between substrates even with minor differences in amino acid composition [37]. The cellular distribution of palmitoylation-deficient Gαi3 C3A depends on the specific Gβγ dimer with which it is coexpressed. This led to the consideration that the particular Gβγ dimer incorporated into the heterotrimer may influence the positioning of the N-terminus of Gα relative to the membrane, thereby affecting its susceptibility to specific acyltransferases.

In contrast to Gαi3, Gαs exhibit weaker plasma membrane localization when expressed alone. The presence of Gβγ is crucial for targeting Gαs to the plasma membrane. Although only one Gβγ dimer composition, Gβ1γ2, was examined, this observation aligns with previous findings indicating that the membrane affinity of the Gαs subunit is determined by the specific Gβγ dimer [68]. Studies by Evanko and colleagues demonstrated that Gβ2–4γ2–3 dimers were effective in promoting more pronounced plasma membrane localization of Gαs. Based on the available evidence, it appears that Gαs do not play a major role in anchoring heterotrimers to the membrane, unlike Gαi3. Instead, there may be reciprocal cooperation between Gαs and Gβγ that facilitates the targeting of complex components in the plasma membrane.

Mutation of the N-terminal glycine or adjacent cysteine in Gαs resulted in a significant reduction in membrane localization compared to wild-type Gαs, which was partially restored by coexpression with Gβ1γ2. However, this effect was more pronounced for the N-terminal palmitoylation-deficient Gαs G2A mutein than for the second palmitoylation-deficient Gαs C3A mutant. It has been proposed that myristoylation and a polybasic motif act as complementary signals for initial membrane targeting in Gα subunits [40]. The polybasic motif, composed of positively charged amino acids on the protein surface, is more prominent in Gαs than in myristoylated members of the Gαi family (reviewed in [15]). Such positively charged regions on a protein’s surface act as membrane-binding signals through electrostatic interactions with the negatively charged headgroups of membrane lipids. Removal of positive charges from this motif leads to a shift in the localization of Gαs from the plasma membrane to the cytosol [41]. Therefore, the effectiveness of Gβ1γ2 in rescuing the membrane binding loss of the N-terminal palmitoylation-deficient Gαs G2A mutein suggests that the presence of the polybasic motif in Gαs compensates for the lack of palmitoylation. It is worth noting that in heterotrimer complexes, the importance of the second palmitoyl anchor appears to outweigh that of the N-terminal one, as evidenced by the reduced relative amount of heterotrimers containing the Gαs C3A mutein that bind to the membrane. Thus, even in the absence of N-terminal palmitoylation, the polybasic motif alone in Gαs is sufficient for the formation of a complex with Gβ1γ2 and subsequent palmitoylation at the N-terminal C3 residue. This provides direct evidence for the importance of the interaction with the Gβγ dimer and subsequent palmitoylation of Gαs at the C3 residue for the membrane targeting of Gαs heterotrimers.

In our efforts to identify Gαi3 heterotrimer combinations that effectively bind to the dopamine D2R and inhibit ACs upon activation by sumanirole, a selective full-efficacy agonist for the D2R [77] we discovered that certain Gβγ combinations were more successful. The most effective combination consisted of Gβ1 with either Gγ8 or Gγ10. Slightly less effective combinations included Gγ2, while the least preferred dimers were Gβ1γ9, Gβ2γ8, and Gβ2γ10. This suggests that the specific association of Gβ1, rather than Gβ2, with Gγ8 or Gγ10 enables a favorable interaction with the receptor. Therefore, the Gβ subunit plays a crucial role in presenting the Gα subunit in a suitable conformation for effective receptor interaction. Recent evidence also supports the idea that direct interaction between Gβ and the receptor acts as a scaffold, facilitating Gα subunit recruitment and localization of the G protein at the active receptor [78, 79]. Interestingly, we found no correlation between the levels of Gβγ dimers on the plasma membrane and the efficiency of ACs inhibition by the D2R. Instead, our FRET experiments proved that the spatial distribution of G proteins and D2R molecules on the plasma membrane played a more significant role. Specifically, there was a strong correlation between the involvement of specific Gαi3 heterotrimers and D2R in the cell membrane and the efficiency of ACs inhibition.

Our studies also confirmed the substantial role of the Gγ subunit in the interaction between the G protein complex and the activated receptor, consistent with previous research [6, 80,81,82]. Although there is limited structural data demonstrating the direct interaction of Gβγ with GPCRs [83] the crucial role of Gγ subunits in signaling is well-established. Different Gβγ complexes primarily vary in the speed and efficiency of G protein activation upon receptor stimulation. Following receptor activation, the Gβγ complex relocates from the plasma membrane to intracellular membranes [4, 82, 84]. The C-terminal helices of the examined Gγ subunits display wide sequence similarity scores ranging from 30 to 67% [5]. These subunits differ in terms of their lipidation patterns and adjacent stretches of basic and hydrophobic amino acid residues, which contribute to their ability to bind to the membrane and potentially interact with GPCRs. Among the examined subunits, only Gγ9 belongs to the rapidly translocating class I. Our findings for Gβ1γ9 support the hypothesis that fast-translocating Gβγ complexes are less effective in activating effectors at the plasma membrane compared to slower-moving Gβγ complexes [85].

Our findings indicate that even in the absence of post-translational palmitoylation of Gαi3, the Gβ1γ8-complexed heterotrimer retains its ability to effectively reduce intracellular cAMP levels, similar to the wild-type protein. However, it is noteworthy that this particular heterotrimer remains localized in the ER membrane. Interestingly, the D2R has been identified in the ER membrane [86]) and retains its ability to initiate G-protein signaling, even within the ER [87]. This is possible because key components of signaling pathways, including ACs, are also present in the ER and Golgi apparatus [88]. There is growing evidence suggesting that G proteins are not only present but also functional in intracellular compartments such as the ER, nucleus, endosomes, and Golgi complex [89, 90]. It is important to note that activation of GPCRs localized within these internal membranes can lead to distinct effects compared to those occurring at the cell surface [91]. In this study, we focused on assessing intracellular cAMP levels resulting from D2R activation and observed comparable outcomes to the stimulation of the receptor fraction located in the cell membrane. However, it is crucial to consider that overall signaling may differ between these two receptor populations, as G proteins can bind to diverse effectors to activate distinct signaling pathways. Further investigations are necessary to gain deeper insight into the specific Gβγ-effector signaling in both the subcellular localizations of the D2R.

Conclusion

Our study confirmed that the composition of heterotrimers, including all subunits (Gα, Gβ, and Gγ), has a significant impact on the strength and specificity of GPCR-mediated signaling. It is crucial to recognize that an interaction between a GPCR and Gα cannot be generalized to all complexes, not only within a specific class of Gα but also among different heterotrimers of the same subunit. Each heterotrimeric complex has the potential to exhibit distinct variations in overall conformation, primarily determined by the specific combination of Gα, Gβ, and Gγ subunits. This variation has implications for the interaction between heterotrimeric complexes and GPCRs, as well as their interactions with membrane lipids.

Recent studies have postulated that selective GPCRs likely engage specific G proteins through shared contacts and further stabilize the complex by forming selective contacts located at the periphery of the GPCR:G protein interface [92]. Our findings strongly support this notion and add to the growing body of evidence emphasizing the fundamental importance of specific compositions of Gαβγ heterotrimers in GPCR signaling. Both the Gα and Gβγ subunits contribute to the modulation of downstream signaling events, highlighting their cooperative role in mediating GPCR signaling pathways.

Availability of data and materials

All data are contained within the article and supporting information. The datasets used and analyzed data during this study are available from the corresponding author on reasonable request.

References

  1. Downes GB, Gautam N. The G Protein Subunit Gene Families. Genomics. 1999;62:544–52.

    PubMed  CAS  Google Scholar 

  2. Wettschureck N, Offermanns S. Mammalian G Proteins and Their Cell Type Specific Functions. Physiol Rev. 2005;85:1159–204.

    PubMed  CAS  Google Scholar 

  3. Hermans E. Biochemical and pharmacological control of the multiplicity of coupling at G-protein-coupled receptors. Pharmacol Ther. 2003;99:1.

    Google Scholar 

  4. Masuho I, Skamangas NK, Muntean BS, Martemyanov KA. Diversity of the Gβγ complexes defines spatial and temporal bias of GPCR signaling. Cell Syst. 2021;12:4.

    Google Scholar 

  5. O’Neill PR, Karunarathne WKA, Kalyanaraman V, Silvius JR, Gautam N. G-protein signaling leverages subunit-dependent membrane affinity to differentially control translocation to intracellular membranes. Proc Natl Acad Sci. 2012;109:51.

    Google Scholar 

  6. Chen H, Leung TC, Giger KE, Stauffer AM, Humbert JE, Sinha S, et al. Expression of the G protein γT1 subunit during zebrafish development. Gene Expr Patterns. 2007;7:5.

    Google Scholar 

  7. Venkatakrishnan AJ, Deupi X, Lebon G, Heydenreich FM, Flock T, Miljus T, et al. Diverse activation pathways in class A GPCRs converge near the G-protein-coupling region. Nature. 2016;536:7617.

    Google Scholar 

  8. Vögler O, Barceló JM, Ribas C, Escribá PV. Membrane interactions of G proteins and other related proteins. Biochim Biophys Acta. 2008;1778:7–8.

    Google Scholar 

  9. Oldham WM, Hamm HE. Structural basis of function in heterotrimeric G proteins. Q Rev Biophys. 2006;39:2.

    Google Scholar 

  10. Kisselev O, Pronin A, Ermolaeva M, Gautam N. Receptor-G protein coupling is established by a potential conformational switch in the βγ complex. Proc Natl Acad Sci U S A. 1995;92:9102–6.

    PubMed  PubMed Central  CAS  Google Scholar 

  11. Taylor JM, Jacob-Mosier GG, Lawton RG, VanDort M, Neubig RR. Receptor and Membrane Interaction Sites on Gβ. J Biol Chem. 1996;271:7.

    Google Scholar 

  12. Ajith Karunarathne WK, PR O’Neill, Martinez-Espinosa PL, Kalyanaraman V, Gautam N. All G protein βγ complexes are capable of translocation on receptor activation. Biochem Biophys Res Commun. 2012;421:3.

    Google Scholar 

  13. Akgoz M, Kalyanaraman V, Gautam N. Receptor-mediated Reversible Translocation of the G Protein Gβγ Complex from the Plasma Membrane to the Golgi Complex. J Biol Chem. 2004;279:49.

    Google Scholar 

  14. Masuho I, Ostrovskaya O, Kramer GM, Jones CD, Xie K, Martemyanov KA. Distinct profiles of functional discrimination among G proteins determine the actions of G protein-coupled receptors. Science Signaling. 2015;8:405.

    Google Scholar 

  15. Polit A, Mystek P, Błasiak E. Every detail matters . That is , how the interaction between G α proteins and membrane affects their function. 2021;

  16. van Meer G, Voelker RD, Feigenson GW. Membrane lipids: where they are and how they behave. Nat Rev Mol Cell Biol. 2008;9:2.

    Google Scholar 

  17. Balla T. Phosphoinositides: Tiny Lipids With Giant Impact on Cell Regulation. Physiol Rev. 2013;93:1019–137.

    PubMed  PubMed Central  CAS  Google Scholar 

  18. Sjöstedt E, Zhong W, Fagerberg L, Karlsson M, Mitsios N, Adori C, et al. An atlas of the protein-coding genes in the human, pig, and mouse brain. Science. 2020;367:1090.

    Google Scholar 

  19. Hynes TR, Tang L, Mervine SM, Sabo JL, Yost EA, Devreotes PN, et al. Visualization of G protein βγ dimers using bimolecular fluorescence complementation demonstrates roles for both β and γ in subcellular targeting. J Biol Chem. 2004;279:29.

    Google Scholar 

  20. Mystek P, Rysiewicz B, Gregrowicz J, Dziedzicka-Wasylewska M, Polit A. Gγ and Gα Identity Dictate a G-Protein Heterotrimer Plasma Membrane Targeting. Cells. 2019;8:1246.

    PubMed  PubMed Central  CAS  Google Scholar 

  21. Polit A, Rysiewicz B, Mystek P, Błasiak E, Wasylewska MD. The Gαi protein subclass selectivity to the dopamine ­D2 receptor is also decided by their location at the cell membrane. Cell Commun Signal. 2020;18:189.

    PubMed  PubMed Central  CAS  Google Scholar 

  22. Mystek P, Tworzydło M, Dziedzicka-Wasylewska M, Polit A. New insights into the model of dopamine D1 receptor and G-proteins interactions. Biochim Biophys Acta. 2015;1853:3.

    Google Scholar 

  23. Jajesniak P, Wong TS. QuickStep-Cloning: a sequence-independent, ligation-free method for rapid construction of recombinant plasmids. J Biol Eng. 2015;9:15.

    PubMed  PubMed Central  Google Scholar 

  24. Pfaffl MW. A new mathematical model for relative quantification in real-time RT– PCR. Nucleic Acids Res. 2001;29:9.

    Google Scholar 

  25. Honorato RV, Koukos PI, Jiménez-garcía B, Bonvin AMJJ. Structural Biology in the Clouds: The WeNMR-EOSC Ecosystem. Front Mol Biosci. 2021;8:729513.

    PubMed  PubMed Central  Google Scholar 

  26. Van Zundert GCP, Rodrigues JPGLM, Trellet M, Schmitz C, Kastritis PL, Karaca E, et al. The HADDOCK2.2 Web Server: User-Friendly Integrative Modeling of Biomolecular Complexes. J Mol Biol. 2015;428:4.

    Google Scholar 

  27. Li Y, Sternweis PM, Charnecki S, Smith TF, Gilman AG, Neer EJ, et al. Sites for Gα Binding on the G Protein β Subunit Overlap with Sites for Regulation of Phospholipase Cβ and Adenylyl Cyclase. J Biol Chem. 1998;273:26.

    Google Scholar 

  28. Goujon M, Mcwilliam H, Li W, Valentin F, Squizzato S, Paern J, et al. A new bioinformatics analysis tools framework at EMBL– EBI. Nucleic Acids Res. 2010;38:W695-9.

    PubMed  PubMed Central  CAS  Google Scholar 

  29. Sievers F, Wilm A, Dineen D, Gibson TJ, Karplus K, Li W, et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Syst Biol. 2011;7:539.

    PubMed  PubMed Central  Google Scholar 

  30. Schmidt CJ, Thomas TC, Levinell MA, Neers EJ. Specificity of G Protein β and γ. J Biol Chem. 1992;267:20.

    Google Scholar 

  31. Pronin AN, Gautam N. Interaction between G-protein 13 and. Proc Natl Acad Sci USA. 1992;89:July.

  32. Hawesj BE, Van BT, Koch WJ, Lefkowitz RJ, Luttrell LM. Distinct Pathways of Gi- and Gq-mediated Mitogen-activated Protein Kinase Activation. J Biol Chem. 1995;270:29.

    Google Scholar 

  33. Atwood BK, Lopez J, Wager-Miller J, Mackie K, Straiker A. Expression of G protein-coupled receptors and related proteins in HEK293, AtT20, BV2, and N18 cell lines as revealed by microarray analysis. BMC Genomics. 2011;12:1.

    Google Scholar 

  34. Adler J, Parmryd I. Quantifying Colocalization by Correlation : The Pearson Correlation Coefficient is Superior to the Mander ’ s Overlap Coefficient Quantifying Colocalization by Correlation : The Pearson Correlation Coefficient is Superior to the Mander ’ s Overlap Coeffic. Cytometry A. 2010;77:8.

    Google Scholar 

  35. Wang Y, Windh RT, Chen CA, Manning DR. N-myristoylation and βγ play roles beyond anchorage in the palmitoylation of the G protein α(o) subunit. J Biol Chem. 1999;274:52.

    Google Scholar 

  36. Mumby SM, Heukeroth RO, Gordon JI, Gilman AG. G-protein α-subunit expression, myristoylation, and membrane association in COS cells. Proc Natl Acad Sci USA. 1990;87:2.

    Google Scholar 

  37. Solis GP, Kazemzadeh A, Abrami L, Valnohova J, Alvarez C, van der Goot FG, et al. Local and substrate-specific S-palmitoylation determines subcellular localization of Gαo. Nat Commun. 2022;13:1.

    Google Scholar 

  38. Galbiati F, Guzzi F, Magee AI, Milligan G, Parenti M. N-terminal fatty acylation of the α-subunit of the G-protein G(i)1: Only the myristoylated protein is a substrate for palmitoylation. Biochem J. 1994;303:3.

    Google Scholar 

  39. Degtyarev MY, Spiegel AM, Jones TLZ. Palmitoylation of a G protein α(i) subunit requires membrane localization not myristoylation. J Biol Chem. 1994;269:49.

    Google Scholar 

  40. Kosloff M, Elia N, Selinger Z. Structural homology discloses a bifunctional structural motif at the N-termini of Gα proteins. Biochemistry. 2002;41:49.

    Google Scholar 

  41. Crouthamel M, Thiyagarajan MM, Evanko DS, Wedegaertner PB. N-terminal polybasic motifs are required for plasma membrane localization of Gαs and Gαq. Cell Signal. 2008;20:10.

    Google Scholar 

  42. Wedegaertner PB, Chu DH, Wilson PT, Levis MJ, Bourne HR. Palmitoylation is required for signaling functions and membrane attachment of G(q)α and G(s)α. J Biol Chem. 1993;268:33.

    Google Scholar 

  43. Yim YY, Betke KM, McDonald WH, Gilsbach R, Chen Y, Hyde K, et al. The in vivo specificity of synaptic Gβ and Gγ subunits to the α2a adrenergic receptor at CNS synapses. Sci Rep. 2019;9:1.

    CAS  Google Scholar 

  44. Tennakoon M, Senarath K, Kankanamge D, Ratnayake K, Wijayaratna D, Olupothage K, et al. Subtype-dependent regulation of Gβγ signalling. Cellular Signalling. 2021;82:December 2020.

  45. Hillenbrand M, Schori C, Schöppe J, Plückthun A. Comprehensive analysis of heterotrimeric G-protein complex diversity and their interactions with GPCRs in solution. Proc Natl Acad Sci. 2015;112:11.

    Google Scholar 

  46. Dingus J, Wells CA, Campbell L, Cleator JH, Robinson K, Hildebrandt JD. G protein βγ dimer formation: Gβ and Gγ differentially determine efficiency of in vitro dimer formation. Biochemistry. 2005;44:35.

    Google Scholar 

  47. Dupré DJ, Robitaille M, Rebois VR, Hébert TE. The Role of Gβγ Subunits in the Organization, Assembly, and Function of GPCR Signaling Complexes. Annu Rev Pharmacol Toxicol. 2009;49:31–56.

    PubMed  PubMed Central  Google Scholar 

  48. Mizuno-yamasaki E, Rivera-molina F, Novick P. GTPase Networks in Membrane Traffic. Annu Rev Biochem. 2012;81:637–59.

    PubMed  PubMed Central  CAS  Google Scholar 

  49. Morishita R, Kato K, Asano T. A brain-specific y subunit of G protein freed from the corresponding β subunit under non-denaturing conditions. FEBS Lett. 1994;337:23–6.

    PubMed  CAS  Google Scholar 

  50. Blake BL, Wing MR, Zhou JY, Lei Q, Hillmann JR, Behe CI, et al. Gβ Association and Effector Interaction Selectivities of the Divergent Gγ Subunit Gγ13. J Biol Chem. 2001;276:52.

    Google Scholar 

  51. Simon MI, Strathmann MP, Gautam N. Diversity of G proteins in signal transduction. Science. 1991;252:1971.

    Google Scholar 

  52. Yan K, Kalyanaraman V, Gautam N. Differential ability to form the G protein βγ complex among members of the β and γ subunit families. J Biol Chem. 1996;271:12.

    Google Scholar 

  53. Fletcher JE, Lindorfer MA, DeFilippo JM, Yasuda H, Guilmard M, Garrison JC. The G protein β5 subunit interacts selectively with the G(q) α subunit. J Biol Chem. 1998;273:1.

    Google Scholar 

  54. Yoshikawa DM, Hatwar M, Smrcka AV. G protein β5 subunit interactions with α subunits and effectors. Biochemistry. 2000;39:37.

    Google Scholar 

  55. Hillenbrand M, Schori C, Schöppe J, Plückthun A. Comprehensive analysis of heterotrimeric G-protein complex diversity and their interactions with GPCRs in solution. Proc Natl Acad Sci USA. 2015;112:11.

    Google Scholar 

  56. Lambright DG, Sondek J, Bohm A, Skiba NP, Hamm HE, Sigler PB. The 2.0 Å crystal structure of a heterotrimeric G protein. Nature. 1996;379:6563.

    Google Scholar 

  57. Ford CE, Skiba NP, Bae H, Daaka Y, Reuveny E, Shekter LR, et al. Molecular basis for interactions of G protein betagamma subunits with effectors Molecular Basis for Interactions of G Protein βγ Subunits with Effectors. Science. 1998;280:5367.

    Google Scholar 

  58. Hillenbrand M, Schori C, Schöppe J, Plückthun A. Comprehensive analysis of heterotrimeric G-protein complex diversity and their interactions with GPCRs in solution. 2015;

  59. Berghuis AM, Lee E, Raw AS, Gilman AG, Sprang SR. Structure of the GDP – Pi complex of Gly203 Ala Gαi1: a mimic of the ternary product complex of Gα-catalyzed GTP hydrolysis. Structure. 1996;4:1277–90.

    PubMed  CAS  Google Scholar 

  60. Wall MA, Coleman DE, Lee E, Iaiguez-lluhi JA, Posner BA, Gilman AG, et al. The Structure of the G Protein Heterotrimer Giα1β1γ2. Cell. 1995;83:1047–58.

    PubMed  CAS  Google Scholar 

  61. Mumby SM, Casey PJ, Gilman AG, Gutowski S, Sternweis PC. G protein y subunits contain a 20-carbon isoprenoid. Proc Natl Acad Sci USA. 1990;87:5873–7.

    PubMed  PubMed Central  CAS  Google Scholar 

  62. Gautam N, Downes GB, Yan K, Kisselev O. The G-Protein βγ Complex. Cell Signal. 1998;10:7.

    Google Scholar 

  63. Myung C, Yasuda H, Liu WW, Harden TK, Garrison JC, Hill C, et al. Role of Isoprenoid Lipids on the Heterotrimeric G Protein γ Subunit in Determining Effector Activation. J Biol Chem. 1999;274:23.

    Google Scholar 

  64. Linder ME, Pang I-H, Duronio RJ, Gordon JI, Sternweis PC, Gilman AG. Lipid modifications of G protein subunits. J Biol Chem. 1991;266:7.

    Google Scholar 

  65. Jones TLZ, Simonds WF, Merendino JJ Jr, Brann MR, Spiegel AM. Myristoylation of an inhibitory GTP-binding protein α subunit is essential for its membrane attachment. Proc Natl Acad Sci U S A. 1990;87:2.

    Google Scholar 

  66. Takida S, Wedegaertner PB. Heterotrimer Formation, Together with Isoprenylation, Is Required for Plasma Membrane Targeting of Gβγ. J Biol Chem. 2003;278:19.

    Google Scholar 

  67. Michaelson D, Ahearn I, Bergo M, Young S, Philips M, Francisco S. Membrane Trafficking of Heterotrimeric G Proteins via the Endoplasmic Reticulum and Golgi. Mol Biol Cell. 2002;13:3294–302.

    PubMed  PubMed Central  CAS  Google Scholar 

  68. Evanko DS, Thiyagarajan MM, Siderovski DP, Wedegaertner PB. Gβγ Isoforms Selectively Rescue Plasma Membrane Localization and Palmitoylation of Mutant Gαs and Gαq. J Biol Chem. 2001;276:26.

    Google Scholar 

  69. Wedegaertner PB. Lipid Modifications and Membrane Targeting of Gα. Biol Signals Recept. 1998;7:125–35.

    PubMed  CAS  Google Scholar 

  70. Peitzsch RM, McLaughlin S. Binding of Acylated Peptides and Fatty Acids to Phospholipid Vesicles: Pertinence to Myristoylated Proteins. Biochemistry. 1993;32:39.

    Google Scholar 

  71. Wilson PT, Bourne HR. Fatty acylation of α(z). Effects of palmitoylation and myristoylation on α(z) signaling. J Biol Chem. 1995;270:16.

    Google Scholar 

  72. Mumby SM, Kleuss C, Gilman AG. Receptor regulation of G-protein palmitoylation. Proc Natl Acad Sci USA. 1994;91:7.

    Google Scholar 

  73. Marrari Y, Crouthamel M, Irannejad R, Wedegaertner PB. Assembly and Trafficking of Heterotrimeric G Proteins. Biochemistry. 2007;46:26.

    Google Scholar 

  74. Ohno Y, Kihara A, Sano T, Igarashi Y. Intracellular localization and tissue-specific distribution of human and yeast DHHC cysteine-rich domain-containing proteins. Biochim Biophys Acta. 2006;1761:4.

    Google Scholar 

  75. Tsutsumi R, Fukata Y, Noritake J, Iwanaga T, Perez F, Fukata M. Identification of G Protein α Subunit-Palmitoylating Enzyme. Mol Cell Biol. 2009;29:2.

    Google Scholar 

  76. Nůsková H, Serebryakova MV, Ferrer-Caelles A, Sachsenheimer T, Lüchtenborg C, Miller AK, et al. Stearic acid blunts growth-factor signaling via oleoylation of GNAI proteins. Nat Commun. 2021;12:1.

    Google Scholar 

  77. McCall RB, Lookingland KJ, Bédard PJ, Huff RM. Sumanirole, a highly dopamine D2-selective receptor agonist: In vitro and in vivo pharmacological characterization and efficacy in animal models of Parkinson’s disease. J Pharmacol Exp Ther. 2005;314:3.

    Google Scholar 

  78. Tsai C, Marino J, Adaixo R, Pamula F, Muehle J, Maeda S, et al. Cryo-EM structure of the rhodopsin-Gαi-βγ complex reveals binding of the rhodopsin C-terminal tail to the gβ subunit. eLife. 2019;8:e46041.

    PubMed  PubMed Central  Google Scholar 

  79. García-Nafría J, Lee Y, Bai X, Carpenter B, Tate CG. Cryo-EM structure of the adenosine A2A receptor coupled to an engineered heterotrimeric G protein. eLife. 2018;7:e35946.

    PubMed  PubMed Central  Google Scholar 

  80. Kisselev OG, Ermolaeva M, Gautam N. A farnesylated domain in the G protein gamma subunit is a specific determinant of receptor coupling. J Biol Chem. 1994;269:34.

    Google Scholar 

  81. Hou Y, Azpiazu I, Smrcka A, Gautam N. Selective role of G protein γ subunits in receptor interaction. J Biol Chem. 2000;275:50.

    Google Scholar 

  82. Senarath K, Payton JL, Kankanamge D, Siripurapu P, Tennakoon M, Karunarathne A. Gγ identity dictates efficacy of Gβγ signaling and macrophage migration. J Biol Chem. 2018;293:8.

    Google Scholar 

  83. Mcintire WE. A model for how G βγ couples G α to GPCR. J Gen Physiol. 2022;154:5.

    Google Scholar 

  84. Saini DK, Kalyanaraman V, Chisari M, Gautam N. A family of G protein βγ subunits translocate reversibly from the plasma membrane to endomembranes on receptor activation. J Biol Chem. 2007;282:33.

    Google Scholar 

  85. Kankanamge D, Tennakoon M, Karunarathne A, Gautam N. G protein gamma subunit, a hidden master regulator of GPCR signaling. J Biol Chem. 2022;298:12.

    Google Scholar 

  86. Sesack SR, Aoki C, Pickel VM. Ultrastructural localization of D2 receptor-like immunoreactivity in midbrain dopamine neurons and their striatal targets. J Neurosci. 1994;14:1.

    Google Scholar 

  87. Prou D, Gu W, Crom S Le, Vincent J, Salamero J, Vernier P. Intracellular retention of the two isoforms of the D 2 dopamine receptor promotes endoplasmic reticulum disruption. Journal of cell sciencecience. 2001;114.

  88. Yamamoto S, Kawamura K, James TN. Intracellular distribution of adenylate cyclase in human cardiocytes determined by electron microscopic cytochemistry. Microsc Res Tech. 1998;40:6.

    Google Scholar 

  89. Hewavitharana T, Wedegaertner PB. Non-canonical signaling and localizations of heterotrimeric G proteins. Cell Signal. 2012;24:1.

    Google Scholar 

  90. Saini DK, Chisari M, Gautam N. Shuttling and translocation of heterotrimeric G proteins and Ras. Trends Pharmacol Sci. 2009;30:6.

    Google Scholar 

  91. Plouffe B, Thomsen ARB, Irannejad R. Emerging Role of Compartmentalized G Protein-Coupled Receptor Signaling in the Cardiovascular Field. ACS Pharmacology & Translational Science. 2020;3.

  92. Sandhu M, Cho A, Ma N, Mukhaleva E, Namkung Y, Lee S, et al. Dynamic spatiotemporal determinants modulate GPCR: G protein coupling selectivity and promiscuity. Nat Commun. 2022;13:1.

    Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by a grant awarded by the Polish National Center for Science (NCN) no. 2016/23/B/NZ1/00530. The open-access publication of this article was funded by the program “Excellence Initiative–Research University” at the Faculty of Biochemistry, Biophysics and Biotechnology of the Jagiellonian University in Kraków, Poland.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: A.P. and B.R. Funding acquisition: A.P. Investigation and data analysis: Genetics constructs: B.R., E.B., P.M. Live-cell imaging microscopy: B.R. RT-qPCR: E.B. FLIM–FRET experiments: A.P. and B.R. cAMP production experiments: B.R. Bioinformatic analysis: B.R. Writing—original draft: A.P., E.B., B.R. Writing—review and editing: A.P., E.B., M.D.-W., P.M. All authors read and approved the final version of the manuscript.

Corresponding author

Correspondence to Agnieszka Polit.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rysiewicz, B., Błasiak, E., Mystek, P. et al. Beyond the G protein α subunit: investigating the functional impact of other components of the Gαi3 heterotrimers. Cell Commun Signal 21, 279 (2023). https://doi.org/10.1186/s12964-023-01307-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12964-023-01307-w

Keywords