Abstract
Orexins are hypothalamic neuropeptides primarily involved in regulating the sleep/wakefulness cycle and circadian rhythm. They bind to the orexin receptor type 1 (OX1) and type 2 (OX2), well-known drug targets in the treatment of sleep disorders, that have recently been shown to play a significant role in different cancers. Lemborexant is one of a few orexin receptor antagonists that have been approved for the treatment of insomnia. Despite being classified as an antagonist, lemborexant may display agonist-like behavior in the non-canonical signaling pathway of the orexin receptors, as confirmed recently in cancer cell models. Here, we generated a model of OX2 in complex with the full-length Gq protein and used it in the molecular dynamics (MD) study. We compared the impact of lemborexant and the OX2-selective, potent agonist compound 1 on OX2 activation and subsequent guanosine diphosphate (GDP) to guanosine triphosphate (GTP) exchange in the Gαq subunit. These 2 µs MD simulations showed that both ligands evoke similar, activation-like conformational changes in OX2 and explained the observed lemborexant-mediated apoptosis of cancer cells. In addition, MD simulations of the active-state OX2-Gq complexes allowed us to uncover a sequence of micro- and macroscale events during the activation of Gq and to detect important micro- and macroswitches in the Gα subunit.
Keywords: Orexin receptors, OX2, Gq protein activation, GDP/GTP exchange, Signal transduction, Lemborexant, Molecular dynamics
Subject terms: Cellular signalling networks, Drug development, Computational chemistry
Introduction
Orexins1 were discovered in 1998 by two independent research groups, de Lecea et al.2 and Sakurai et al.3 These hypothalamic neuropeptides exist in two forms: orexin-A (OxA, 33 amino acids) and orexin-B (OxB, 28 amino acids)4, and are involved in regulating the central nervous system5. Both are formed from the same precursor via proteolytic processing3 but share only 46% of the same sequence4. Best known for regulating the sleep/wakefulness cycle and circadian rhythm, orexin deficiencies can cause sleep disorders6. They also influence blood pressure, locomotion, hormone secretion7, drug addiction, reward seeking, energy homeostasis5, and peripheral organ function8.
Orexins bind to two G protein-coupled receptors (GPCRs), OX1 and OX29, which recognize both OxA and OxB with poor selectivity10. However, other known ligands demonstrate varied subtype selectivity, e.g., seltorexant (OX2)11, ACT-335827 (OX1)12, with some binding to both receptors13. Agonist binding activates a canonical signal cascade involving the Gq protein, triggering intracellular Ca2+ release. While orexin receptors additionally couple with Gi/o and Gs14, our recent data indicated that OX1 is coupled exclusively to Gq in recombinant human embryonic kidney-OX1 (HEK-OX1) cells15. A second non-canonical signaling pathway of orexin receptors leads to apoptosis. Upon activation by orexins, two immunoreceptor tyrosine-based inhibitory motifs (ITIMs) of OX1 or OX2 are phosphorylated, leading to the recruitment of Src homology region 2 domain-containing phosphatase-2 SHP2. This triggers a signaling cascade involving the activation of the p38 mitogen-activated protein (MAP) kinase, the translocation of the apoptosis regulator Bax in mitochondria, cytochrome C release, and the activation of caspases-3 and -7, ultimately causing cell death5.
Orexin receptors are promising drug targets for sleep disorders like narcolepsy or insomnia16, with small-molecule antagonists such as suvorexant, lemborexant, or daridorexant already on the market. While orexins as agonists promote wakefulness, antagonists treat insomnia17,18. Additionally, the orexin system plays a role in different cancers, e.g., OX1 in glioblastoma, colon, pancreas, and prostate cancers, and OX2 in rat pancreas cancer, endometrial carcinomas, or cortical adenomas19–21. Both orexins can induce apoptosis, thereby displaying antitumoral properties20,22; similar results have been reported for almorexant23. Recent studies have shown that suvorexant, lemborexant, and daridorexant induce apoptosis in human colon cancer cells expressing OX1, noticeably inhibiting OxA-induced intracellular calcium release15. They also prompted apoptosis in recombinant HEK cells expressing OX215. Additionally, lemborexant induced a partial dissociation of Gq, leading to its quasi-agonistic properties on apoptosis induction in cancer cells15.
As the structure of the active-state OX1 has not yet been solved, we used OX2 in complex with Gq and either a potent, OX2-selective, small molecule agonist (compound 1)24 or an antagonist (lemborexant)25 to study the activation of the orexin system by graphics processing unit (GPU)-accelerated all-atom MD simulations. Notably, the activation of OX2 by orexins induced pro-apoptotic effects in cancer cells similar to the one induced by OX115,21. MD is commonly used to observe the evolution of molecular systems in time26, and can be used to study both micro- and macroscale cellular processes27, also in the context of GPCRs28–32. In many such cases MD is complemented with other computational methods, especially those involving the use of artificial intelligence33 that have found applications in cancer diagnosis, treatment, and drug discovery, thereby accelerating the development of new treatment strategies34.
The orexin system has already been included in several in-silico studies. In 2016, Yin et al. performed molecular dynamics simulations to elucidate the basis for receptor subtype selectivity in regard to small-molecule agonists35. They determined which residues affected the binding affinity of the antagonist SB-674042 and confirmed it with radioligand binding experiments35. Computational approaches such as quantitative structure–activity relationship (QSAR) models; molecular docking; and absorption, distribution, metabolism, and excretion (ADME) predictions have also been used in the design of novel OX1 antagonists28, while molecular mechanics have been used to compute the binding free energies of dual orexin receptor antagonists such as lemborexant and suvorexant using previously obtained crystal structures36. Additionally, homology modeling and MD simulations were used to predict the binding mode of OxA to OX2 before the corresponding Protein Data Bank (PDB) structure was released37.
In our MD study, accompanied by the experimental study of Gratio et al.15, we observed similarities between the classic OX2 agonist compound 1 and lemborexant-including systems, suggesting the agonist-like behaviour of lemborexant in the recombinant HEK293T cell line expressing OX215 despite its well-known role as a dual OX1/OX2 antagonist36. In principle, an antagonist tightly binds to the orthosteric binding site of the GPCR receptor, preventing an endogenous agonist binding and the subsequent receptor activation. Since the lemborexant orthosteric binding site in OX1/OX2 is known and well-described36, and no other binding sites, e.g., allosteric, were detected15, its surprising pro-apoptotic effect induced by the receptor activation could not be explained other than as evoking activation as an agonist. Additionally, the preceding experimental study15 showed that the orexin receptor was activated by lemborexant in the absence of orexin, which excludes the possibility of an allosteric mode of action of lemborexant in tested cell lines. This is further supported by the fact that the pre-treatment of HEK cells expressing OX1 with antagonists such as almorexant, lemborexant, suvorexant, and daridorexant completely abolished the induction of calcium release by OxA15. The ligand-dependent preferential activation of selected pathways in different tissues (biased agonism or functional selectivity) was previously observed for many GPCR receptors, e.g., for opioid receptors and β-adrenergic receptors38.
Here, we explained the agonist-like behaviour of lemborexant on the molecular level using MD. The simulation systems included OX2 bound to either lemborexant or the OX2-selective, classic agonist compound 1, and the full-length Gαq, which substituted the mini-Gsqi construct present in the template (PDB: 7L1V)15,24,39 and used in our previous study. Performed MD simulations additionally provided insights into the molecular basis of the OX2 activation, allowing us to propose a sequence of conformational changes that the receptor and Gαq undergo during activation, e.g., the opening of the Gαq helical domain and subsequent release of GDP. Due to a high sequence identity (67% according to Clustal2.1) between two orexin receptors, our results regarding receptor activation could be extrapolated to OX1, for which even more detailed experimental evidence has recently been provided15.
Methods
Preparation of simulation systems
Models of the OX2–Gq systems were prepared using Modeller v. 10.440, with the 7L1V24 PDB entry used as a template for the receptor and G protein complex, and 7SQ241 used as the template for the Gα subunit. The positions of conserved residues in the orexin receptor with respect to other GPCRs were confirmed (see Supplementary Fig. S1). Two separate sets of models were generated—one containing compound 1 in the orthosteric binding site, and another containing lemborexant. Compound 1 was present in the active-state 7L1V structure of OX2. Lemborexant was positioned in the orthosteric binding site by superposing the inactive-state 7XRR PDB structure36 onto the apo 7L1V structure and transferring the lemborexant molecule from the inactive-state structure to the active-state structure in order to retain its binding mode observed in 7XRR. Both simulation systems contained GDP bound to the Gα subunit. The crystal water molecules in the 7SQ2 structure were retained during the model building by Modeller but removed before building the membrane with CHARMM-GUI. The target sequence consisted of the UNIPROT42 entries for OX2 (O43614), Gαq (P50148), Gβ1 (P62873), and Gγ2 (P59768). The UNIPROT entry for OX2 was compared in terms of sequence identity (67%) with OX1 (O43613) using ClustalOmega (https://www.ebi.ac.uk/jdispatcher/msa/clustalo). In order to make sure the N- and C-termini of the OX2 models were comparable to the 7L1V structure, all of the residues before Y1.31 and after F7.68 were removed. Additionally, residues G5.72 to V6.20 of intracellular loop 3 (ICL3) were removed (33 residues) to simplify the model. Although ICL3 is involved in signaling through the G protein43, this region is highly flexible and predicted to be intrinsically disordered44,45, and is absent in most GPCR structures in PDB43. It is assumed to affect the receptor specificity for the G protein type43, and then cAMP accumulation after GDP release46, neither of which were considered in this study.
For each set, 500 preliminary models were generated. Ten iterations of loop refinements were then performed for each model, resulting in a total of 5000 models per set. Their quality was evaluated by comparing DOPE scores47, and the 30 lowest-energy models were further assessed in PyMOL version 2.548. Three lemborexant-OX2-Gq and three compound 1-OX2-Gq models were selected for the generation of MD simulation systems, resulting in a total number of six replicas that were simulated in parallel. The protein preparation wizard in Maestro (Schrödinger Suite release 2022-3)49 was used to preprocess the models, i.e., add hydrogens and optimize bonds.
Molecular dynamics simulations
Inputs for the molecular dynamics simulations were prepared using CHARMM-GUI’s50 Membrane Builder51 (specifically, the Bilayer Builder). Each model was exported from Maestro in the PDB file format and uploaded into CHARMM-GUI. Information about the locations of disulphide bonds in the proteins was obtained from the templates and all crystal water molecules were removed. The ligands were exported from Maestro in sdf format and uploaded into CHARMM-GUI, which then automatically generated CHARMM topology and parameter files using the CHARMM General Force Field (CGenFF). Using the replacement method, the complexes were placed in a lipid bilayer consisting of a 3:1 ratio of POPC (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine) to cholesterol. Such a membrane composition has been used in other MD studies performed for GPCRs52. The periodic rectangular water box (TIP3P) was fitted to the complex, and each system was neutralized by adding Na+ and Cl- ions at a concentration of 0.15 M. The number of atoms in each system was around 240,000–250,000, and the initial size of the simulation box was ca. 122 Å × 122 Å × 179 Å.
The MD simulations were performed using the GPU-accelerated version of NAMD (NAMD 3.0 Alpha)53 with the Charmm36m force field. The equilibration step included 10,000 steps of the steepest descent minimization, then 25,000 steps of the conjugated gradients minimization. The equilibration simulation was performed in NVT (constant particle number, volume, and temperature) using Langevin dynamics (303.15 K). The time integration step in the equilibration and production runs was set to 2 fs. The production run in NPT (constant particle number, pressure, and temperature) was performed using the Langevin piston Nose–Hoover method (1 bar, 303.15 K) and lasted for 2 µs. Every tenth frame of the computed trajectories was combined using CatDCD54, then wrapped using PBCTools55 and analyzed in VMD v.1.9.356. Further analyses were performed using PyMOL and Maestro.
Results
Similarities in binding modes between the classic OX2 agonist compound 1 and lemborexant
MD simulations of both the classic agonist compound 124 and lemborexant-including complexes of the active-state OX2 with the full-length Gαq were analyzed in terms of variance in the position of ligands and complex components (Fig. 1, Supplementary Fig. S2). As a dual OX1/OX2 antagonist57,58, lemborexant was expected to display significant RMSD fluctuations in the binding site of the active-state OX2 structure, especially with additional intracellular stabilization by a G protein. However, its position remained stable, with insignificant RMSD fluctuations comparable to compound 1. This suggests there could be conditions under which lemborexant may induce an active-like conformation of OX2 and behave similarly to the classic agonist. The maintained active state of OX2 during simulations was validated by analyzing DRW motif residues (corresponding to the DRY ‘ionic lock’) (Supplementary Figures S3, S4). In class A GPCRs, R3.50 interacts with T2.39 (inactive state, e.g., 7XRR33) or Y5.58 (active state, e.g., 7L1U, 7L1V24)59. The conformations of the DRW motif residues in the compound 1-including complexes remained ‘on’ in replicas 1 and 3 but shifted into an intermediate-like position in replica 2 (Supplementary Fig. S3). In the lemborexant-including complexes, active-state interactions were also formed, with fewer inactive-state interactions than in compound 1-including replicas (Supplementary Fig. S4). Additional microswitches were analyzed: PVF, CYLP, and NPIIY (corresponding to the PIF, CWxP, and NPxxY motifs, respectively; Supplementary Figure S5). In the majority of systems, V3.40 in the PVF motif changed from an active-like conformation to an intermediate conformation resembling neither the active nor inactive state. The Y6.48 toggle switch of the CYLP motif changed conformations in the final frame of all of the lemborexant replicas but only one compound 1 replica. Although the inactive-state 7XRR and active-state 7L1V structures include the same Y6.48 conformation, the changes of this microswitch are reported to be temporary60. The Y7.53 toggle switch of the NPIIY motif remained in an active state in all but one system, where it adopted an intermediate conformation. Additionally, transmembrane helix 6—recognized as a key macroswitch involved in GPCR activation—did not undergo global rearrangements in any of the simulation replicas. This was assessed by measuring the distance between the Cɑ atoms of R3.50 and A6.34 in the models, 7L1V (11.8 Å), and 7XRR (8.3 Å)61. In all models, these distances remained greater than 11.4 Å, thus comparable to 7L1V. These results further confirmed that OX2 remained active throughout the simulations.
Fig. 1.
Similarities in binding modes between the classic OX2 agonist compound 1 and lemborexant. A comparison of the binding modes of compound 1 (left, replica 1) and lemborexant (right, replica 1). The first frame of the simulation is shown in gray with a yellow ligand, and the final frame in green with a red ligand. The final frame is meant to represent the ligand conformation at the end of the simulations and includes the key receptor residues described in Supplementary Figures S4 and S5.
A detailed analysis of compound 1 and lemborexant binding modes (Fig. 1) aligns with previous studies24,36,57,58. In the 7L1V structure, the phenyl rings of compound 1 were surrounded by a lipophilic patch consisting of P3.29, M4.64, T2.61, and V2.6424. Lemborexant was similarly surrounded by P3.29, T2.61, and V2.64, though Asada et al.36 proposed interactions with T2.61 to be rather irrelevant for lemborexant binding to the inactive-state receptor, suggesting that it demonstrated a slightly different binding mode for the active-state. H7.39 interacted with compound 1 (Supplementary Figs. S6, S7), as indicated by Hong et al.24 and also stabilized lemborexant in 7XRR33 but through π–π stacking instead of a hydrogen bond. Both ligands closely interacted with F5.42 through π–π stacking and formed hydrogen bonds with Q3.32. However, extracellular loop 2 (ECL2) closure, observed previously for OX2–mini-Gsqi complexes15, caused both ligands to move further inside the receptor. Consequently, the Q3.32 side chain also moved towards the inside of the receptor (see Fig. 1) in contrast to what can be observed in 7L1V24. Hong et al. showed that Q3.32 remained stably oriented outward during a 1 μs MD simulation with OxB24. In our 2 μs MD simulations with the small-molecule agonist compound 1, we observed the closing of ECL2 after 1 μs, shifting Q3.32 towards the inside of the receptor after 2 μs. Perhaps the Q3.32 conformation was influenced by ligand bulkiness (small-molecule vs. peptide) and correlated with the closing of ECL2, rather than strictly by ligand type (agonist vs. antagonist) as suggested by Hong et al.24 Since ECL2 remained open for peptide agonists, Q3.32 could be oriented outwards24.
Activation-like changes of microswitches in Gαq
Ham et al. described residue F341G.H5.08 (labeled there as F336) in helix H5 of Gα to be a microswitch affecting the GDP binding energy via switch I and the P loop (Supplementary Fig. S8), as shown in mutagenesis studies62. The F341G.H5.08A mutation or a change of F341G.H5.08 to the ‘on’ position caused fast GDP/GTP exchange rates62. In our homology models, this microswitch was initially in the ‘off’ position, as expected for an inactive G protein (Fig. 2A)62. During the simulations, its change to the ‘on’ position required a slight counterclockwise rotation. This was observed mostly for the lemborexant-including replicas (1 and 2) representing the furthest stage of Gα activation. In previous 2 µs simulations involving mini-Gsqi15, F341G.H5.08 was also in the ‘on’ position at the end of the simulation. This suggested its transition may not be directly related to the opening of Gα, as these systems lacked the alpha helical domain (AHD) of Gα. Instead, this change may have been induced by receptor activation alone.
Fig. 2.
Changes of microswitches D277G.HG.02 and F341G.H5.08 and macroswitches I–III in OX2 in comparison to the rhodopsin-mediated activation of Gαi (magenta). (A) The D277G.HG.02 and F341G.H5.08 microswitches described by Ham et al.62 in the ‘off’ position in the first frame of the compound 1 replica 2 simulation (top, left) and 1GP2 PDB structure (bottom, left). The D277G.HG.02 and F341G.H5.08 microswitches described by Ham et al.62 in the ‘on’ position in the last frame of the compound 1 replica 1 simulation (top, center), the last frame of the lemborexant replica 1 simulation (top, right), the 6CMO PDB structure (bottom, center), and the 6OS9 PDB structure (bottom, right). (B) A comparison between the positions of macroswitches I–III, the P loop, and HC and HD between the active state 6CMO PDB structure70 (magenta), the first frame of compound 1 replica 1 (gray, yellow switches), and the final 2 μs frame of the compound 1 replica 1 (green, red switches) (Left). A comparison between the positions of macroswitches I–III, the P loop, and HC and HD in the active state 6CMO PDB structure70 (magenta), the first frame of the lemborexant replica 1 (gray, yellow switches), and in the final 2 μs frame of the lemborexant replica 1 (green, red switches) (Right).
Ham et al.62 also described the D277G.HG.02 microswitch (labeled there as D272) as important for GDP/GTP exchange. In our case, most replicas demonstrated a breaking of the GDP–D277G.HG.02 interactions regardless of ligand type (Fig. 2, Supplementary Figs. S9, S10). D277G.HG.02 stabilized GDP binding by interacting with the guanine ring62. The breaking of this interaction was required for GDP dissociation, which changed the D277G.HG.02 microswitch into the ‘on’ position—seen as the χ1 dihedral angle (N–Cα–Cβ–Cγ) change from ca. 60° to − 60° (Supplementary Figs. S7, S8). We observed the movement of the phosphoryl group, not the guanine ring, which imposed only slight changes on D277G.HG.02. Notably, GDP dissociated from Gα via its phosphate side63. D277G.HG.02 was indeed in the ‘on’ position at the end of the lemborexant-including simulations. In the compound 1-including replicas, D277G.HG.02 fluctuated, changing to the ‘on’ position only at around 1.70 μs for replica 3, while in the other two replicas, it switched to the ‘off’ position at around 1.00 μs.
Activation-like changes of macroswitches in Gαq inducing the GDP/GTP exchange
The microsecond timescale used for MD simulations allow to observe first stages of the GPCR signal transduction involving both changes in the ligand–receptor complex and the interacting G protein or arrestin31. This required the full length Gαq to be present. To our knowledge, none of the active-state structures of GPCRs in PDB include the full-length Gαq, though the full-length Gαi is present in e.g., the rhodopsin PDB structure 6CMO (Fig. 2). Only the P loop and switch II are present in their entirety in mini-Gsqi, as observed when compared to the Gα topology described by Calebiro et al.64 (Supplementary Fig. S8). For this reason, we constructed models with the full-length Gαq to obtain a fully functional OX2-G protein complex.
Though switch II remained rather unchanged in the previous mini-Gsqi simulations15, in most of the full-length Gαq simulations changes in that region were more visible, e.g., a subsequent movement of helix H2. This altered Gα–Gβγ interactions, as Gα macroswitches I and II, and the Gα N-terminal helix form a hydrophobic cleft fitting Gβγ64–66. These differences between results for the full-length Gαq and mini-Gsqi simulations suggest that switch II movement might be connected specifically to the presence of the AHD in the former.
In our MD simulations of OX2 with a full-length Gαq, a noticeable difference was observed for compound 1 vs. lemborexant. At the end of the simulations, the most open Gα conformations were observed for the lemborexant-including simulations (Supplementary Figs. S11, S12). These opening movements of the Gα subunit, required for the GDP/GTP exchange and confirmed experimentally for other GPCR receptors67, were hardly observed for compound 1 (Fig. 3). This may suggest that Gα opens faster in the lemborexant-including systems.
Fig. 3.
Changes in Gα macroswitches inducing the opening of the GDP/GTP exchange tunnel upon compound 1 binding to OX2. The opening of a tunnel was observed near the phosphate side of GDP. From left to right: interactions between nearby residues, a surface view of the Gα subunit and GDP, hydrogen bonds determined by VMD to form between R183G.hfs2.20 (switch I) and E49G.s1h1.04 (P loop), and R183G.hfs2.20 and E241G.s4h3.10 (switch III) at a maximum 4 Å distance at an angle cutoff of 20º.
The GDP position changed during the simulations regardless of the OX2 ligand (Supplementary Fig. S13). Notably, the terminal phosphoryl group of GDP lost its interaction with R183G.hfs2.02 (GproteinDb68 numbering scheme) in switch I on behalf of the second phosphoryl group. The formation of hydrogen bonds between the oxygen atoms in the second phosphoryl group and R183G.hfs2.02 was analyzed (Supplementary Fig. S14). In the compound 1-including complexes, hydrogen bonds between R183G.hfs2.02 and GDP were more retained than in the case of lemborexant. However, the average RMSD changes for the switch I (ca. 4 Å) were comparable for both systems (Supplementary Fig. S15).
The beginning of the opening of the ‘GDP/GTP exchange tunnel’ (Supplementary Files S3, S4) near the phosphate side of GDP was observed in both simulation systems, and the same way of GDP dissociation was described by Louet et al.63 This was most visible for compound 1 replica 1 and lemborexant replica 1 (Figs. 3, 4). This opening of the exchange tunnel, along with the opening of Gα, suggested that these replicas represent the furthest stage of G protein activation regarding GDP release. In the remaining replicas, this tunnel was blocked by residues from the P loop and switch I. Namely, the formation of a salt bridge between R183G.hfs2.02 (switch I) and E49G.s1h1.04 (P loop) blocked the tunnel and prevented GDP from leaving the binding site (lemborexant replica 3, compound 1 replicas 2 and 3; Figs. 3, 4). The disruption of this salt bridge was necessary for Gα to open69. In our simulations, during Gα activation, R183G.hfs2.02 in switch I moved to the outer side to interact with E241G.s4h3.10 in switch III instead (lemborexant replica 2). Presumably, switch III was responsible for stabilizing R183G.hfs2.02 at this stage, preventing it from interfering with GDP leaving the binding site. Interestingly, at the beginning of every simulation, E241G.s4h3.10 in switch III interacted with two residues of switch II: R210G.H2.01 and Q209G.s3h2.03 (to a lesser extent). Thus, the release of E241G.s4h3.10 from interactions with switch II might also be required69.
Fig. 4.
Changes in Gα macroswitches inducing the opening of the GDP/GTP exchange tunnel upon lemborexant binding to OX2. The opening of a tunnel was observed near the phosphate side of GDP. From left to right: interactions in nearby residues, a surface view of the Gα subunit and GDP, hydrogen bonds determined by VMD to form between R183G.hfs2.20 (switch I) and E49G.s1h1.04 (P loop), and R183G.hfs2.20 and E241G.s4h3.10 (switch III) at a maximum 4 Å distance at an angle cutoff of 20º.
To verify if the described above direction of the observed changes in Gα is activation-like, we compared our results with the open conformation of Gα bound to rhodopsin (6CMO)70 (Fig. 2) in terms of Gα local and global conformational changes. A further comparison of 6CMO with 1GP266 (the inactive-state G protein with bound GDP) showed a similar opening of D277G.HG.02 and F341G.H5.08 as observed in our simulations (Fig. 2A). Furthermore, there was a relocation of H63G.H1.12 near M59G.H1.08 (mentioned by Ham et al.62) from helix H1 to an adjacent loop resulting from the change in F341G.H5.08. The extension of the adjacent loop was followed by the opening of Gα (observed in full in 6CMO). This movement of H63G.H1.12 was not seen in our simulations as it would require much longer simulation timescale to move AHD for such long distance observed in 6CMO (Fig. 2B). However, the beginning of this H63G.H1.12 translocation was observed in most simulations (Fig. 2A).
The release of GDP is certainly associated with the opening of Gα (6CMO vs. 1GP2). It requires the translation of the AHD to the right and bottom, in accordance with Fig. 2B, which can be seen in the movement of helices HA and HD of 6CMO. Before Gα opening, residues in switches I–III (especially switch I) had to start moving in the same direction as in 6CMO. Both simulation systems showed macroswitch changes resembling those in 6CMO (Fig. 2B), confirming the beginning of the G protein activation.
The average RMSD values of switch III were higher for the lemborexant-including simulations than for the compound 1-including simulations (Supplementary Fig. S16). These larger conformational fluctuations correlated with the downward movement of HD and HA, causing the whole AHD to move downward, as seen in 6CMO. This movement appeared crucial for the opening of Gα (Fig. 2B, Supplementary Fig. S11).
Global conformational changes in Gq
Noticeable fluctuations of the AHD in both simulation systems suggest the first stage of the opening of Gα (Fig. 2). Helix HD of Gα displayed fluctuations to the left or right (Fig. 2B) with a simultaneous HA movement. In some replicas, HD moved right preceding Gα opening, like in 6CMO. HD movement to the left seemed to represent fluctuations of the relative position of the RD and AHD domains that did not lead to the GDP dissociation. After 1 μs, two lemborexant-including replicas showed HD moving right, but this was observed only in one compound 1-including replica. By 2 μs, only one lemborexant-including replica but two compound 1-including replicas displayed this movement. This observation indicated that fluctuations of the AHD, including HD or HA movements, likely precede the opening of Gα as this seems to be a highly dynamic process involving the simultaneous exchange of other molecules in its binding site70. Nevertheless, the simulation replicas that demonstrated rightward HD and HA fluctuations like in 6CMO also demonstrated higher fluctuations of GDP RMSD (Supplementary Figure S13), while those with leftward movement had low fluctuations of GDP RMSD. This confirmed that the global rearrangements of HD and HA observed in our simulations were indeed associated with the GDP dissociation.
As mentioned above, the lemborexant systems seemed to demonstrate a faster opening of the Gα subunit, associated with the GDP dissociation, than the compound 1 systems (Supplementary Fig. 11). The way this could occur, i.e., via a hinge involving helix HF, seemed to be consistent with previously described results71. In the lemborexant-including simulations, a surface view of Gαq showed AHD gradually separating from RD, creating the GDP/GTP exchange tunnel. This was less noticeable in the compound 1 replica 1 (Fig. 3, Supplementary Fig. S12). AHD remained closed in other replicas. Interestingly, the opening of Gα in the lemborexant-including simulations was accompanied by a slight rotation of the N-terminal helix of Gα, like described by Ahn et al.72 on the example of two cryo-EM structures of active-state NTSR1 (Supplementary Fig. S17).
Conclusions and discussion
The results of the all-atom MD simulations, including both mini-Gsqi15 and full-length Gαq, suggest that lemborexant could behave as a non-typical agonist under certain conditions, rather than solely as an antagonist57,58. Minimal fluctuations in its position in the orthosteric binding site of OX2 indicated that it could stabilize the active-state receptor conformation, similarly to the known OX2R agonist compound 1. The maintained active state of OX2 during 2 μs of simulations was confirmed through a conformational analysis of selected microswitches. Recently, Tyson et al.73 suggested a new molecular mechanism for inverse agonists, through which they can bind to and act through GPCR-G protein complexes. This raises the question of whether lemborexant could instead be an inverse agonist. The authors observed that the microswitches present in the structures of inverse agonists bound to the κ-opioid receptor were in a mix of active (the PIF and DRW motifs), inactive (the orthosteric and Na + binding pockets, and the CWxP motif), and intermediate states (the NPxxY motif)73. The microswitch conformations in the final frames of our simulations (Supplementary Fig. S5) were not in agreement with these observations but rather indicated the maintained active state of the receptor. On this basis, we concluded that lemborexant is unlikely to be an inverse agonist. However, there may be differences between the κ-opioid and orexin receptors that contribute to variations in inverse agonist behavior, and our simulations may not fully capture all possible receptor conformations, especially under different physiological conditions.
We analyzed and summarized the observed conformational changes (Table 1). Larger conformational changes of switch III correlated with the opening of Gα. This caused higher GDP position fluctuations in the lemborexant-including simulation systems in comparison to the compound 1 systems. Notably, during the first microsecond of the simulations, fluctuations of the GDP binding region were observed mostly for the lemborexant-OX2-G protein multicomplexes and only slight fluctuations were noticed for the compound 1 complexes. This suggests differences between the assumed agonist-like behavior of lemborexant and the canonical mode of action of compound 1 in Gq signaling. It cannot be undoubtedly stated whether these differences refer only to the activation-like conformational changes of the studied multicomplex or also to the rate of the GDP/GTP exchange. However, we could suggest that GDP dissociation is faster for the lemborexant complexes whereas compound 1 seems to better stabilize the Gαq-GDP complex. The impact of this on the rate of the GDP/GTP exchange and subsequent Gq-mediated signaling is unclear.
Table 1.
A comparison of the conformational changes observed in microsecond MD simulations including lemborexant and compound 1.
| Conformational change | Compound 1 replica 1/2/3* | Lemborexant replica 1/2/3 |
|---|---|---|
| Ligand position in the receptor binding site | No change | No change/no change/slight change |
| ECL2 closing | Yes/yes/yes | Yes/yes/yes |
| Changes of macroswitch III | Yes/no/no | Yes/yes/no |
| Changes of AHD position (opening of Gα) | Yes/no/yes | Yes/yes/yes |
| Formation of the GDP/GTP exchange tunnel | Yes/no/no | Yes/yes/no |
| Active-like direction of rotation of N-terminal helix of Gα like in NTSR1 | Yes/no/no | No/no/no |
| Changes of macroswitch I to ‘on’ with R183G.hfs2.02 movement away from GDP | Yes/yes/yes | Yes/yes/yes |
| Changes of F341G.H5.08 microswitch to ‘on’ | Yes/no/no | Yes/yes/no |
| Changes of D277G.HG.02 microswitch to ‘on’ | Yes/yes/yes | Yes/yes/yes |
*If differences between replicas were observed.
In addition, there is the question of whether the simulations performed with an active-state apo receptor would yield similar Gα activation observations30. Though no such study has been performed for the orexin system, de Lima et al.74 performed 1 µs MD simulations for an active-state apo µ-opioid receptor in complex with Gi-GDP. They observed that when the receptor was bound to Gi in a closed conformation, it switched to an inactive state and remained that way until the end of the simulation74. The authors therefore concluded that GDP kept Gi in a closed conformation, and its activation and subsequent opening could be triggered by either the presence of an agonist or a larger simulation timescale74. Previously, we also observed that the presense of an agonist is required to rebuild the receptor conformation into the fully active state, not intermediate, confirmed later with the cryo-EM structure of C5aR130. Considering the above studies, we first observed changes in G protein microswitch conformations within 1 μs of our simulations, leading to the conclusion that it was indeed the presence of an agonist that led to Gq activation. This was further confirmed by the experimental results obtained by Gratio et al.15.
The results presented here show only the first stages of the GDP dissociation and Gq activation accessible to MD timescales. In addition to the known role of switches I and II shown recently in metadynamics simulations of the receptor-free spontaneous G protein activation69, we described the mechanism of the beginning of the opening of Gαq and GDP dissociation. The opening of Gαq involves two major movements of AHD rightward and downward (consistent with Fig. 2), in the direction known previously from the structure of the rhodopsin–Gi multicomplex (6CMO), wherein Gα is in its fully open conformation70. As we observed, one movement of AHD (rightward) could require changes of the F341G.H5.08 microswitch, while the second movement (downward) could require changes in switch III. Slight changes in switch I (Fig. 2), as shown in our simulations and in 6CMO, induced the translation of AHD to both sides. In addition, we analyzed the GDP binding site to determine which residues and interactions may contribute to its dissociation.
The GDP dissociation required the opening of Gα and it seemed to be associated mostly with the movement of AHD to the right (Fig. 4). The GDP release was mainly triggered by conformational changes near its phosphoryl end, i.e., a loss of interactions of GDP with R183G.hfs2.02 located in switch I. Thus, no clear and definite changes in the D277G.HG.02 microswitch were observed. Based on the above, we conclude that the sequence of conformational changes during Gαq activation mediated by an OX2 agonist could be as follows (Fig. 5):
changes in the F341G.H5.08 microswitch,
a loss of interactions between the phosphoryl end of GDP and R183G.hfs2.02 (switch I) caused by changes in switch I,
the formation of temporary interactions between R183G.hfs2.02 and E49G.s1h1.04 (P loop),
a loss of interactions between R183G.hfs2.02 and E49G.s1h1.04,
a loss of interactions between E241G.s4h3.10 (switch III) and two residues of switch II: R210G.H2.01 and Q209G.s3h2.03 (to a lesser extent), releasing E241G.s4h3.10,
the formation of interactions between R183G.hfs2.02 and E241G.s4h3.10, stabilizing the large distance between R183G.hfs2.02 and GDP,
the opening of the GDP/GTP exchange tunnel near the phosphoryl end of GDP,
a loss of interactions between R183G.hfs2.02 and E241G.s4h3.10 and E49G.s1h1.04
changes in the D277G.HG.02 microswitch,
a loss of interactions between the guanine end of GDP and D277G.HG.02 caused by changes in the D277G.HG.02 rotamer,
changes in helices HD and HA accompanied by the opening of Gα,
GDP release accompanied by a further opening of Gα.
Fig. 5.
The suggested sequence of conformational changes that take place during the activation of Gαq. The initial conformation is presented in gray, the suggested conformational change that takes place is presented in green, and the yellow dashed lines represent polar contacts.
In the canonical Gq signaling pathway, the GDP/GTP exchange initiates Gαq protein activation, followed by the breaking of most Gα–Gβγ interactions. Gα activates phospholipase C (PLC)75, which in turn converts phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol 1,4,5-trisphosphate (IP3) and diacyl glycerol (DAG). IP3, acting in the cytosol, induces Ca2+ release from the endoplasmic reticulum, while DAG, attached to the cell membrane, activates protein kinase C (PKC). Recent experimental results15 have proven that lemborexant induces cancer cell apoptosis through a non-canonical, Ca2+-independent pathway. This suggests that lemborexant does not activate the OX1/OX2-Gq complex in a classic way to release Gα for the subsequent PLC activation. Instead, Gβγ could be released or partially released to activate the Src kinase, which in turn phosphorylates the ITIM site in OX1/OX2, leading to SHP2 phosphatase recruitment, p38 MAP kinase activation, and subsequently caspase 3/7 activation. Other examples involving Gβγ have already been observed, e.g., for the Gi-coupled M2 muscarinic acetylcholine, α2-adrenergic, and somatostatin 2 receptors, and for Gs-coupled β-adrenergic receptors, which induced the ERK pathway through Gβγ-mediated stimulation of tyrosine kinases such as Src76,77. In contrast, the Gq/11-coupled muscarinic acetylcholine receptor M1 and α1-adrenergic receptor activated the ERK-mediated apoptosis signaling by Gq/1176,77. In the case of Gq-coupled P2Y1 receptor, the stress-activated protein kinase (SAPK) activation (without involving p38) also evoked caspase activity, leading to apoptosis76.
Here, we observed a clear difference in the GDP/GTP exchange rate leading to the opening of Gαq. The faster activation of Gq by the lemborexant-OX2R complexes could be an impulse to evoke the Gβγ-mediated activation of Src. Presumably, these differences in the rate of the GDP/GTP exchange between lemborexant and compound 1 complexes affect the activation of Gq, leading to the activation of signaling pathways mediated by different subunits (Gβγ vs. Gα, respectively). However, the timescale currently accessible to MD simulations is not enough to observe the evolution of the second messenger system activated by orexin receptor effectors.
Supplementary Information
Acknowledgements
We acknowledge the High-Performance Computing Infrastructure of the University of Warsaw Biological and Chemical Research Center. We gratefully acknowledge High-Performance Computing Infrastructure PLGrid in Poland (HPC Center: ACK Cyfronet AGH) for providing computer facilities and support (PLG/2024/017063) and the LUMI supercomputer (PLL/2023/04/016464), owned by the EuroHPC Joint Undertaking, hosted by CSC (Finland) and the LUMI consortium through PLGrid, Poland. We acknowledge funds for a short-term scientific mission within the ERNEST COST Action (CA18133) (E-COST-GRANT-CA18133-37c87d4f).
Abbreviations
- GPCR
G protein-coupled receptor
- OX1
Orexin receptor type 1
- OX2
Orexin receptor type 2
- OxA
Orexin-A
- OxB
Orexin-B
- GDP
Guanosine diphosphate
- GTP
Guanosine triphosphate
- MD
Molecular dynamics
- AHD
Alpha helical domain
- RD
Ras-like domain
- MD
Molecular dynamics
Author contributions
This computational study was designed by D.L. Protein modeling, molecular dynamics simulations, and the analysis of simulation trajectories were performed by P.D. and D.L. The first draft of the manuscript was written by P.D. and D.L. Figures were designed and prepared by P.D. and D.L. Movies were prepared by D.L. The manuscript was edited and reviewed by all of the authors. All authors have given approval to the final version of the manuscript.
Data availability
PDB structures used in the current study are available at: https://www.rcsb.org (ID: 7L1V, 7SQ2, 6OSA, 6OS9, 6CMO). Protein sequences used in the current study are available at: https://www.uniprot.org (ID: P50148, O43613, O43614). Protein models and simulation trajectories generated during the current study are available at: 10.5281/zenodo.14035283 and 10.5281/zenodo.14035739.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Change history
9/29/2025
The original online version of this Article was revised: The accompanying Supplementary Information files have been updated.
Supplementary Information
The online version contains supplementary material available at 10.1038/s41598-025-03857-0.
References
- 1.Kukkonen, J. P., Jacobson, L. H., Hoyer, D., Rinne, M. K. & Borgland, S. L. International union of basic and clinical pharmacology CXIV: Orexin receptor function, nomenclature and pharmacology. Pharmacol. Rev.76(5), 625–688. 10.1124/pharmrev.123.000953 (2024). [DOI] [PubMed] [Google Scholar]
- 2.de Lecea, L. et al. The hypocretins: Hypothalamus-specific peptides with neuroexcitatory activity. Proc. Natl. Acad. Sci.95(1), 322–327. 10.1073/pnas.95.1.322 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Sakurai, T. et al. Orexins and orexin receptors: a family of hypothalamic neuropeptides and G Protein-coupled receptors that regulate feeding behavior. Cell92(4), 573–585. 10.1016/S0092-8674(00)80949-6 (1998). [DOI] [PubMed] [Google Scholar]
- 4.Janto, K., Prichard, J. R. & Pusalavidyasagar, S. An update on dual orexin receptor antagonists and their potential role in insomnia therapeutics. J. Clin. Sleep Med.14(08), 1399–1408. 10.5664/jcsm.7282 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Couvineau, A. et al. The anti-tumoral properties of orexin/hypocretin hypothalamic neuropeptides: an unexpected therapeutic role. Front. Endocrinol.9, 573. 10.3389/fendo.2018.00573 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Chemelli, R. M. et al. Narcolepsy in orexin knockout mice: molecular genetics of sleep regulation. Cell98(4), 437–451. 10.1016/S0092-8674(00)81973-X (1999). [DOI] [PubMed] [Google Scholar]
- 7.Sutcliffe, J. G. & de Lecea, L. The hypocretins: Excitatory neuromodulatory peptides for multiple homeostatic systems, including sleep and feeding. J. Neurosci. Res.62(2), 161–168 (2000). [DOI] [PubMed] [Google Scholar]
- 8.Heinonen, M. V., Purhonen, A. K., Mäkelä, K. A. & Herzig, K. H. Functions of orexins in peripheral tissues. Acta Physiol.192(4), 471–485. 10.1111/j.1748-1716.2008.01836.x (2008). [DOI] [PubMed] [Google Scholar]
- 9.Sakurai, T. The role of orexin in motivated behaviours. Nat. Rev. Neurosci.15(11), 719–731. 10.1038/nrn3837 (2014). [DOI] [PubMed] [Google Scholar]
- 10.Laburthe, M. & Voisin, T. The orexin receptor OX1R in colon cancer: A promising therapeutic target and a new paradigm in G protein-coupled receptor signalling through ITIMs. Br. J. Pharmacol.165(6), 1678–1687. 10.1111/j.1476-5381.2011.01510.x (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bonaventure, P. et al. Characterization of JNJ-42847922, a selective orexin-2 receptor antagonist, as a clinical candidate for the treatment of insomnia. J. Pharmacol. Exp. Ther.354(3), 471–482. 10.1124/jpet.115.225466 (2015). [DOI] [PubMed] [Google Scholar]
- 12.Steiner, M. A. et al. Discovery and characterization of ACT-335827, an orally available, brain penetrant orexin receptor type 1 selective antagonist. ChemMedChem8(6), 898–903. 10.1002/cmdc.201300003 (2013). [DOI] [PubMed] [Google Scholar]
- 13.Pałasz, A. et al. Dual orexin receptor antagonists—promising agents in the treatment of sleep disorders. Int. J. Neuropsychopharmacol.17(1), 157–168. 10.1017/S1461145713000552 (2014). [DOI] [PubMed] [Google Scholar]
- 14.Kukkonen, J. P. & Leonard, C. S. Orexin/hypocretin receptor signalling cascades. Br. J. Pharmacol.171(2), 314–331. 10.1111/bph.12324 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Gratio, V., Dragan, P., Garcia, L., Saveanu, L., Nicole, P., Voisin, T., Latek, D., & Couvineau, A. (2024). Pharmacodynamics of the orexin type 1 (OX1) receptor in colon cancer cell models: A two‐sided nature of antagonistic ligands resulting from partial dissociation of Gq. Br. J. Pharmacol. bph.17422. 10.1111/bph.17422 [DOI] [PubMed]
- 16.Kodadek, T. & Cai, D. Chemistry and biology of orexin signaling. Mol. BioSyst.6(8), 1366–1375. 10.1039/C003468A (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Roch, C., Bergamini, G., Steiner, M. A. & Clozel, M. Nonclinical pharmacology of daridorexant: A new dual orexin receptor antagonist for the treatment of insomnia. Psychopharmacology238(10), 2693–2708. 10.1007/s00213-021-05954-0 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Roth, T., Roehrs, T. & Pies, R. Insomnia: Pathophysiology and implications for treatment. Sleep Med. Rev.11(1), 71–79. 10.1016/j.smrv.2006.06.002 (2007). [DOI] [PubMed] [Google Scholar]
- 19.Couvineau, A., Nicole, P., Gratio, V. & Voisin, T. Orexins/hypocretins and cancer: A neuropeptide as emerging target. Molecules26(16), Article 16. 10.3390/molecules26164849 (2021). [Google Scholar]
- 20.Yang, S. et al. Multiomics integration reveals the effect of Orexin A on glioblastoma. Front. Pharmacol.14, 1096159. 10.3389/fphar.2023.1096159 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Voisin, T., Firar, A. E., Avondo, V. & Laburthe, M. Orexin-induced apoptosis: The key role of the seven-transmembrane domain orexin type 2 receptor. Endocrinology147(10), 4977–4984. 10.1210/en.2006-0201 (2006). [DOI] [PubMed] [Google Scholar]
- 22.Voisin, T. et al. Aberrant expression of OX1 receptors for orexins in colon cancers and liver metastases: An openable gate to apoptosis. Can. Res.71(9), 3341–3351. 10.1158/0008-5472.CAN-10-3473 (2011). [DOI] [PubMed] [Google Scholar]
- 23.Dayot, S., et al. In vitro, in vivo and ex vivo demonstration of the antitumoral role of hypocretin-1/orexin-A and almorexant in pancreatic ductal adenocarcinoma. Oncotarget. 9(6), 6952–6967. 10.18632/oncotarget.24084 (2018). [DOI] [PMC free article] [PubMed]
- 24.Hong, C. et al. Structures of active-state orexin receptor 2 rationalize peptide and small-molecule agonist recognition and receptor activation. Nat. Commun.12(1), 815. 10.1038/s41467-021-21087-6 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Scott, L. J. Lemborexant: First approval. Drugs80(4), 425–432. 10.1007/s40265-020-01276-1 (2020). [DOI] [PubMed] [Google Scholar]
- 26.Hollingsworth, S. A. & Dror, R. O. Molecular dynamics simulation for all. Neuron99(6), 1129–1143. 10.1016/j.neuron.2018.08.011 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Kmiecik, S. et al. Coarse-grained protein models and their applications. Chem. Rev.116(14), 7898–7936. 10.1021/acs.chemrev.6b00163 (2016). [DOI] [PubMed] [Google Scholar]
- 28.Feng, Z., Ma, S., Hu, G. & Xie, X.-Q. Allosteric binding site and activation mechanism of class C G-protein coupled receptors: metabotropic glutamate receptor family. AAPS J.17(3), 737–753. 10.1208/s12248-015-9742-8 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Wang, J. & Miao, Y. Mechanistic insights into specific G protein interactions with adenosine receptors. J. Phys. Chem. B123(30), 6462–6473. 10.1021/acs.jpcb.9b04867 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Wisniewski, S., Dragan, P., Makal, A. & Latek, D. Helix 8 in chemotactic receptors of the complement system. PLoS Comput. Biol.18(7), e1009994. 10.1371/journal.pcbi.1009994 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Rodríguez-Espigares, I. et al. GPCRmd uncovers the dynamics of the 3D-GPCRome. Nat. Methods17(8), 777–787. 10.1038/s41592-020-0884-y (2020). [DOI] [PubMed] [Google Scholar]
- 32.David, Aranda-García Tomasz Maciej et al. Large scale investigation of GPCR molecular dynamics data uncovers allosteric sites and lateral gateways Nature Communications 16(1). https://doi.org/10.1038/s41467-025-57034-y (2025). [DOI] [PMC free article] [PubMed]
- 33.Dorota, Latek Khushil, Prajapati Paulina, Dragan Matthew, Merski Przemysław, Osial (2025) GPCRVS - AI-driven Decision Support System for GPCR Virtual Screening International Journal of Molecular Sciences 26(5) 2160-10.3390/ijms26052160 [DOI] [PMC free article] [PubMed]
- 34.Sanmukh, S. G., Krzykawska-Serda, M., Dragan, P., Baron, S., Lobaccaro, J.-M. A., & Latek, D. Is Cancer Our Equal or Our Better? Artificial Intelligence in Cancer Drug Discovery (Springer International Publishing, 2024). 10.1007/16833_2024_326.
- 35.Yin, J. et al. Structure and ligand-binding mechanism of the human OX1 and OX2 orexin receptors. Nat. Struct. Mol. Biol.23(4), Article 4. 10.1038/nsmb.3183 (2016). [DOI] [PubMed] [Google Scholar]
- 36.Asada, H. et al. Molecular basis for anti-insomnia drug design from structure of lemborexant-bound orexin 2 receptor. Structure30(12), 1582-1589.e4. 10.1016/j.str.2022.11.001 (2022). [DOI] [PubMed] [Google Scholar]
- 37.Karhu, L., Magarkar, A., Bunker, A. & Xhaard, H. Determinants of orexin receptor binding and activation—a molecular dynamics study. J. Phys. Chem. B123(12), 2609–2622. 10.1021/acs.jpcb.8b10220 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Dragan, P. & Latek, D. The two-sided impact of beta-adrenergic receptor ligands on inflammation. Curr. Opin. Physiol.41, 100779. 10.1016/j.cophys.2024.100779 (2024). [Google Scholar]
- 39.Nehmé, R. et al. Mini-G proteins: Novel tools for studying GPCRs in their active conformation. PLoS ONE12(4), e0175642. 10.1371/journal.pone.0175642 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Eswar, N. et al. Comparative protein structure modeling using modeller. Curr. Protoc. Bioinform.10.1002/0471250953.bi0506s15 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Waldo, G. L. et al. Kinetic scaffolding mediated by a phospholipase C–β and Gq signaling complex. Science330(6006), 974–980. 10.1126/science.1193438 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.The UniProt Consortium. UniProt: The universal protein knowledgebase in 2023. Nucleic Acids Res.51(D1), D523–D531. 10.1093/nar/gkac1052 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Sadler, F. et al. Autoregulation of GPCR signalling through the third intracellular loop. Nature615(7953), 734–741. 10.1038/s41586-023-05789-z (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Jaakola, V.-P., Prilusky, J., Sussman, J. L. & Goldman, A. G protein-coupled receptors show unusual patterns of intrinsic unfolding. Protein Eng. Des. Sel.18(2), 103–110. 10.1093/protein/gzi004 (2005). [DOI] [PubMed] [Google Scholar]
- 45.Deupi, X. Relevance of rhodopsin studies for GPCR activation. Biochim. Biophys. Acta BBA Bioenerget.1837(5), 674–682. 10.1016/j.bbabio.2013.09.002 (2014). [DOI] [PubMed] [Google Scholar]
- 46.Qiu, X. et al. Coupling and activation of the β1 adrenergic receptor—the role of the third intracellular loop. J. Am. Chem. Soc.146(41), 28527–28537. 10.1021/jacs.4c11250 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Shen, M. & Sali, A. Statistical potential for assessment and prediction of protein structures. Protein Sci. Publ. Protein Soc.15(11), 2507–2524. 10.1110/ps.062416606 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.The PyMOL Molecular Graphics System, Schrödinger (Version 2.4.0). (n.d.). [Computer software]. Schrödinger, LLC.
- 49.Maestro. (2022). [Computer software]. Schrödinger, LLC.
- 50.Lee, J. et al. CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. J. Chem. Theory Comput.12(1), 405–413. 10.1021/acs.jctc.5b00935 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Wu, E. L. et al. CHARMM-GUI membrane builder toward realistic biological membrane simulations. J. Comput. Chem.35(27), 1997–2004. 10.1002/jcc.23702 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Sengupta, D. & Chattopadhyay, A. Identification of cholesterol binding sites in the Serotonin1A receptor. J. Phys. Chem. B116(43), 12991–12996. 10.1021/jp309888u (2012). [DOI] [PubMed] [Google Scholar]
- 53.Phillips, J. C. et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J. Chem. Phys.153(4), 044130. 10.1063/5.0014475 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Gullingsrud, J. (version 4.0). CatDCD (Version 4.0) [Computer software]. Theoretical and Computational Biophysics Group.
- 55.Giorgino, T., Henin, J., Lenz, O., Mura, C., & Saam, J. (version 2.7). PBCTools (Version 2.7) [Computer software]. Theoretical and Computational Biophysics Group.
- 56.Humphrey, W., Dalke, A. & Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph.14(1), 33–38. 10.1016/0263-7855(96)00018-5 (1996). [DOI] [PubMed] [Google Scholar]
- 57.Beuckmann, C. T. et al. In vitro and in silico characterization of lemborexant (E2006), a novel dual orexin receptor antagonist. J. Pharmacol. Exp. Ther.362(2), 287–295. 10.1124/jpet.117.241422 (2017). [DOI] [PubMed] [Google Scholar]
- 58.Yoshida, Y. et al. Design, synthesis, and structure–activity relationships of a series of novel N-aryl-2-phenylcyclopropanecarboxamide that are potent and orally active orexin receptor antagonists. Bioorg. Med. Chem.22(21), 6071–6088. 10.1016/j.bmc.2014.08.034 (2014). [DOI] [PubMed] [Google Scholar]
- 59.Hauser, A. S. et al. GPCR activation mechanisms across classes and macro/microscales. Nat. Struct. Mol. Biol.28(11), Article 11. 10.1038/s41594-021-00674-7 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Trzaskowski, B. et al. Action of molecular switches in GPCRs—theoretical and experimental studies. Curr. Med. Chem.19(8), 1090–1109. 10.2174/092986712799320556 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kjølbye, L. R. et al. Lipid modulation of a class B GPCR: elucidating the modulatory role of PI(4,5)P2 lipids. J. Chem. Inf. Model.62(24), 6788–6802. 10.1021/acs.jcim.2c00635 (2022). [DOI] [PubMed] [Google Scholar]
- 62.Ham, D. et al. Conformational switch that induces GDP release from Gi. J. Struct. Biol.213(1), 107694. 10.1016/j.jsb.2020.107694 (2021). [DOI] [PubMed] [Google Scholar]
- 63.Louet, M., Martinez, J. & Floquet, N. GDP release preferentially occurs on the phosphate side in heterotrimeric G-proteins. PLoS Comput. Biol.8(7), e1002595. 10.1371/journal.pcbi.1002595 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Calebiro, D., Koszegi, Z., Lanoiselée, Y., Miljus, T. & O’Brien, S. G protein-coupled receptor-G protein interactions: A single-molecule perspective. Physiol. Rev.101(3), 857–906. 10.1152/physrev.00021.2020 (2021). [DOI] [PubMed] [Google Scholar]
- 65.Lambright, D. G. et al. The 2.0 Å crystal structure of a heterotrimeric G protein. Nature379(6563), 311–319. 10.1038/379311a0 (1996). [DOI] [PubMed] [Google Scholar]
- 66.Wall, M. A. et al. The structure of the G protein heterotrimer Giα1β1γ2. Cell83(6), 1047–1058. 10.1016/0092-8674(95)90220-1 (1995). [DOI] [PubMed] [Google Scholar]
- 67.Dohlman, H. G. & Jones, J. C. Signal activation and inactivation by the Gα helical domain: A long-neglected partner in G protein signaling. Sci. Signal.5(226), re2. 10.1126/scisignal.2003013 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Pándy-Szekeres, G. et al. The G protein database, GproteinDb. Nucleic Acids Res.50(D1), D518–D525. 10.1093/nar/gkab852 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Sun, X., Singh, S., Blumer, K. J. & Bowman, G. R. Simulation of spontaneous G protein activation reveals a new intermediate driving GDP unbinding. Elife7, e38465. 10.7554/eLife.38465 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Kang, Y. et al. Cryo-EM structure of human rhodopsin bound to an inhibitory G protein. Nature558(7711), Article 7711. 10.1038/s41586-018-0215-y (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Dror, R. O. et al. Structural basis for nucleotide exchange in heterotrimeric G proteins. Science348(6241), 1361–1365. 10.1126/science.aaa5264 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Ahn, D., Ham, D. & Chung, K. Y. The conformational transition during G protein-coupled receptor (GPCR) and G protein interaction. Curr. Opin. Struct. Biol.69, 117–123. 10.1016/j.sbi.2021.03.013 (2021). [DOI] [PubMed] [Google Scholar]
- 73.Tyson, A. S., et al. Molecular mechanisms of inverse agonism via κ-opioid receptor–G protein complexes. Nat. Chem. Biol. 1–12. 10.1038/s41589-024-01812-0 (2025). [DOI] [PMC free article] [PubMed]
- 74.Lima, M. R. P. D., Bezerra, R. F. S., Serafim, D. D. B. & Sena Junior, D. M. Dynamics of the Apo µ-opioid receptor in complex with Gi protein. Int. J. Mol. Sci.24(17), 13430. 10.3390/ijms241713430 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Suh, P.-G. et al. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep.41(6), 415–434. 10.5483/BMBRep.2008.41.6.415 (2008). [DOI] [PubMed] [Google Scholar]
- 76.Sellers, L. A. et al. Adenosine nucleotides acting at the human P2Y1 receptor stimulate mitogen-activated protein kinases and induce apoptosis. J. Biol. Chem.276(19), 16379–16390. 10.1074/jbc.M006617200 (2001). [DOI] [PubMed] [Google Scholar]
- 77.Yanamadala, V., Negoro, H. & Denker, B. Heterotrimeric G proteins and apoptosis: intersecting signaling pathways leading to context dependent phenotypes. Curr. Mol. Med.9(5), 527–545. 10.2174/156652409788488784 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
PDB structures used in the current study are available at: https://www.rcsb.org (ID: 7L1V, 7SQ2, 6OSA, 6OS9, 6CMO). Protein sequences used in the current study are available at: https://www.uniprot.org (ID: P50148, O43613, O43614). Protein models and simulation trajectories generated during the current study are available at: 10.5281/zenodo.14035283 and 10.5281/zenodo.14035739.





