422

Biophysical Journal

Volume 107

July 2014

422–434

Article Differences in Allosteric Communication Pipelines in the Inactive and Active States of a GPCR Supriyo Bhattacharya1 and Nagarajan Vaidehi1,* 1

Division of Immunology, Beckman Research Institute of the City of Hope, Duarte, California

ABSTRACT G-protein-coupled receptors (GPCRs) are membrane proteins that allosterically transduce the signal of ligand binding in the extracellular (EC) domain to couple to proteins in the intracellular (IC) domain. However, the complete pathway of allosteric communication from the EC to the IC domain, including the role of individual amino acids in the pathway is not known. Using the correlation in torsion angle movements calculated from microseconds-long molecular-dynamics simulations, we elucidated the allosteric pathways in three different conformational states of b2-adrenergic receptor (b2AR): 1), the inverseagonist-bound inactive state; 2), the agonist-bound intermediate state; and (3), the agonist- and G-protein-bound fully active state. The inactive state is less dynamic compared with the intermediate and active states, showing dense clusters of allosteric pathways (allosteric pipelines) connecting the EC with the IC domain. The allosteric pipelines from the EC domain to the IC domain are weakened in the intermediate state, thus decoupling the EC domain from the IC domain and making the receptor more dynamic compared with the other states. Also, the orthosteric ligand-binding site becomes the initiator region for allosteric communication in the intermediate state. This finding agrees with the paradigm that the nature of the agonist governs the specific signaling state of the receptor. These results provide an understanding of the mechanism of allosteric communication in class A GPCRs. In addition, our analysis shows that mutations that affect the ligand efficacy, but not the binding affinity, are located in the allosteric pipelines. This clarifies the role of such mutations, which has hitherto been unexplained.

INTRODUCTION G-protein-coupled receptors (GPCRs) are seven helical transmembrane (TM) proteins that mediate intracellular (IC) signaling and form the largest class of drug targets. GPCRs act as allosteric machines in communicating the event of ligand binding in the extracellular (EC) domain to the IC domain to initiate the coupling of effector proteins. Binding of an agonist or an inverse agonist leads to stabilization of specific conformational states of the GPCR, and binding of both an agonist and a G-protein stabilizes the receptor in the fully active state, as evidenced by recently published crystal structures (1–6). The key question of how the ligand binding in the EC domain is communicated to residues in the IC domain has not been completely answered and is difficult to resolve from experiments. Allosteric modulators that modulate (increase or decrease) the activity of the orthosteric agonist are emerging as a cutting-edge strategy in drug design for GPCRs (7). Many allosteric modulators affect activation without any direct contact with the orthosteric agonist or the G-protein site. One way these modulators could affect activation is by interacting with the residues that mediate allosteric communication between the orthosteric and the G-protein site. Mapping the allosteric communication pathways in GPCRs would thus provide insights into allostery and aid the discovery of allosteric modulators.

Submitted February 10, 2014, and accepted for publication June 9, 2014. *Correspondence: [email protected]

In class A GPCRs, the distance between the orthosteric ligand binding site and the signaling protein coupling site ˚ (3). Therefore, allosteric communication plays a is >30 A vital role in propagating the activation signal from the orthosteric site to the IC region of the receptor. Biophysical experiments have provided insights into the activation mechanism of mainly class A GPCRs, such as rhodopsin and b2 adrenergic receptor (b2AR) (8,9). It has been proposed that minor perturbations at the agonist-binding site are transmitted to the IC interface via intermediate residue contacts (10,11). This hypothesis is corroborated by the fact that mutations of residues distant from both the ligand-binding site and the G-protein interface affect receptor activation. The role of such mutations is not well understood. Here, we used mutual information (MI) in internal coordinates, calculated from microseconds-long, all-atom molecular-dynamics (MD) simulations of human b2AR in explicit membrane and water, to elucidate the allosteric communication pathways. Computational approaches for mapping allosteric networks in proteins range from bioinformatics (12,13) to elastic network models (14,15), analysis of contact maps (16), force distribution (17), and correlated residue motions (18–22). One drawback of bioinformatics-based methods, such as evolutionary trace analysis, is that they cannot distinguish functional residues from those that govern structural stability. Previous studies used correlated residue motions to map allosteric networks in soluble proteins by

Editor: David Sept. Ó 2014 by the Biophysical Society 0006-3495/14/07/0422/13 $2.00

http://dx.doi.org/10.1016/j.bpj.2014.06.015

Allosteric Pipelines in GPCR Activation

employing MI in Cartesian coordinates (19,21) calculated from MD simulations. MI-based network analyses with accelerated MD (AMD) trajectories were used to map allosteric communication between the ligand-binding site and G-protein site of the M2 muscarinic acetylcholine receptor (M2-AcR), a class A GPCR (22). The use of Cartesian coordinates in the above works could introduce noise into the correlation calculation due to high-frequency motions (18,20), leading to spurious correlations among allosteric residues. Here, we used correlations in the torsional space, which eliminates high-frequency noise (20). In this work, we derived the allosteric communication pathways for inactive and active conformations of human b2AR using correlations in torsion angle movements (calculated using MI) obtained from long-timescale MD trajectories. When multiple allosteric communication pathways show significant overlap, they form what we term allosteric pipelines. The residues in the pipeline through which multiple pathways pass are termed allosteric hubs. Our main goals in this work were to (1), investigate the role of allosteric pipelines in conferring stability to inactive and active states of the receptor; (2), identify allosteric hub residues that play an important role in allosteric communication; and (3), identify allosteric sites that can aid in the design of allosteric modulators (7). MATERIALS AND METHODS Description of the MD protocol The MD trajectories for the three states of b2AR, i.e., starting from the crystal structures of the inactive state, the active state with the nanobody (G-protein mimic) removed, and the nanobody-bound active state, were obtained from DE Shaw Research (23,24). Inactive-state simulations were performed for 16 ms on the inactive b2AR crystal structure (Protein Data Bank (PDB) ID: 2RH1) bound to the inverse agonist carazolol. Active-state simulations (5.9 ms) were performed on the active b2AR crystal structure (PDB ID: 3P0G) bound to nanobody and the agonist BI-167017. Some of the simulations that were started from the active-state structure (3P0G), but with the nanobody removed, showed a collapse of the structure to the inactive state. These simulations showed a pathway to the deactivation of the receptor. We extracted the ensemble for the agonist-bound intermediate state from these simulations as detailed in Text S1 of the Supporting Material. All simulations were performed in the presence of palmitoyl oleoyl phosphatidyl choline (POPC) lipids, water, and ions using the CHARMM27 force field (25) on the ANTON supercomputer.

Calculation of allosteric pathways We calculated the MI using the torsional degrees of freedom as described in Text S2. We calculated allosteric pathways between each pair of residues ˚ apart in that showed above-average MI (MI > MIavg) and were >10 A the receptor structure. We first constructed an undirected graph using interresidue contacts, where the residues formed nodes and the interresidue contacts formed the edges of the network. An interresidue contact was iden˚ of one another. tified if the Ca atoms of the residue pair were within 10 A The edge weights were calculated as MImax – MIab, where MImax is the maximum MI among all residue pairs in the receptor and MIab is the MI between the terminal residues of the edge. Moreover, to avoid selecting path-

423 ways with weak MI, the weights of all edges with MI < MIavg were set to zero. For a given residue pair, the allosteric pathway is defined to be the connecting route between the two residues, which minimizes the number of intermediate nodes and maximizes the sum of edge MIs of the connecting route. The allosteric pathways were calculated using the shortest-path algorithm by Dijkstra (26), as implemented in the Bioinformatics ToolBox in MATLAB (The MathWorks, Natick, MA).

Clustering of allosteric pathways into pipelines The allosteric pathways were sorted by the MI between their termini residues, and the top 500 pathways were selected for further analysis. These pathways were clustered based on their mutual proximity in the receptor structure. To define the proximity of two pathways, we defined a parameter called overlap, which is the fraction of nodes of both pathways that are ˚ ) of one another. Thus, nearby pathways within a cutoff distance (10 A have strong overlap and distant pathways have zero overlap. Using overlap as a similarity metric, we clustered the pathways using the hierarchical clustering routine in MATLAB. To determine the optimal number of clusters, we calculated 1), intraoverlap, defined as the average overlap between pathways within each cluster; and 2), interoverlap, defined as the average overlap between clusters. The optimal number of clusters was chosen as that which maximized intraoverlap while minimizing interoverlap. More details regarding the clustering procedure are given in Text S3 and Fig. S8.

RESULTS We calculated the allosteric pipelines from the MI calculated from extensive MD simulations of three distinct conformational states of b2AR: 1) the inactive state of the receptor bound to the inverse agonist carazolol; 2) agonist-bound b2AR, a highly dynamic state that is referred to as the intermediate state (10,11) because it is on the activation pathway to the fully active state; and 3) the fully active state of b2AR with both the full agonist BI-167017 and a G-protein mimetic nanobody bound. For details regarding these MD trajectories, see the Materials and Methods section. We first characterized the number and nature of the substates present in the ensemble of the three MD trajectories. Analysis of crystal structures of the active (PDB ID: 3SN6) and the inactive (PDB ID: 2RH1) states of b2AR showed a movement of the IC end of TM6 away from TM3 in going from the inactive state to the active state. To characterize the ensemble of substates sampled in the MD simulations of the inactive, intermediate, and fully active states of b2AR, we compared the distance between the IC ends of TM3 and TM6 (shown in Fig. 1 A) among the different substates, as well as the root mean-square deviation (RMSD) of the NPxxY motif on TM7 from the inactive state (Fig. S1 f) (24). Fig. 1 B shows the distribution of the distance between TM3 and TM6 for the inactive, intermediate, and active states from the three ensembles. Whereas the active- and inactive-state receptors show sharper peaks close to their respective crystal structures, the intermediate agonist-bound state has a broader distribution of TM3–TM6 distance, with a peak in between the inactive and active states. Also in the intermediate state, the NPxxY motif assumes an inactivelike conformation (Fig. S1 f). The broad distribution of Biophysical Journal 107(2) 422–434

424

Bhattacharya and Vaidehi

FIGURE 1 (A) The TM3–TM6 distance is measured between the geometric centers of the residues highlighted in orange on TM3 and TM6. Due to the high flexibility of helical ends, the last four residues from the IC end of each helix were excluded from the geometric center calculation. (B) Population distributions of TM3–TM6 distance in simulations of the inactive (blue), intermediate (green), and active (red) states. This distance in the crystal structures of the inactive (PDB ID: 2RH1) and G-protein-bound active (PDB ID: 3SN6) states are shown by dotted lines. (C) Average interdihedral MI as a function of the pairwise interdihedral distance. A dihedral is defined as a set of four sequentially bonded atoms that form a torsion angle. The interdihedral distance is the distance between the geometric centers of two dihedrals. (D) Average MI and ratio of allosteric to direct contact MI for different b2AR states. To see this figure in color, go online.

TM3–TM6 distance in the intermediate state agrees well with the finding that the order parameter from NMR studies of b2AR showed lower order for the intermediate state than for the active or inactive state (10). We also decomposed the distance distribution for each state into its component Gaussian distributions (Fig. S1). The inactive- and activestate distance distributions show populations around two distinct substates, representing the peaks in the individual Gaussians, whereas the intermediate state shows three substates (Fig. S1). The high number of substates further demonstrates the dynamic nature of the agonist-bound intermediate state. Using this ensemble of states, we calculated the MI in torsional space as described in the Materials and Methods section and Text S2.

˚. sion MI decreases, but then rises slowly beyond 10 A Therefore, the trend shown in Fig. 1 C is evidence of the presence of allosteric communication in these dynamic trajectories of b2AR. The allosteric long-distance communication is weakest in the intermediate state and increases in both the inactive and active states. We quantified the average MI in different receptor states using the ratio of the allosteric MI to the direct MI in Fig. 1 D. It is evident from this figure that the intermediate state shows both lower direct MI and allosteric MI, and therefore the residue movements are more uncorrelated, leading to a more dynamic state of the receptor. This is in agreement with the recent NMR results for b2AR that show that the agonist-bound intermediate state is more dynamic than the inactive or active state of b2AR (10,11).

The inactive state shows more correlated movement of residues than the intermediate and active states

Allosteric communication pathways, or allosteric pipelines

We identified the pairs of residues that showed strong allosteric communication (sorted by their strength of MI). Fig. 1 C shows the time-averaged MI in torsion angles for all torsion pairs as a function of the intertorsion distance (i.e., the distance between the geometric centers of the two torsional angles; also see Text S4). The higher the MI value in torsion angles, the greater is the correlated movement of the atoms that make up the two torsion angles. If the two torsions in a ˚ ), the correlated movement pair are neighbors (within 5 A will be high, as would be expected due to the direct contact ˚ , the intertorof the atoms in the two torsions. Beyond 5 A

For a residue pair with a high level of correlated movement (e.g., residues A and B), we calculated the shortest allosteric communication pathway to get from residue A to B while also maximizing the MI in this pathway, using graph theory algorithms as described in Materials and Methods. Hereafter, residues A and B are referred to as termini. By mapping the pathways between all distant (Ca-Ca distance > ˚ ) residue pairs that have MI above a certain cutoff 10 A (top 500 sorted by the calculated MI), we constructed a network of allosteric pathways in each conformational state of b2AR. Allosteric pathways that connected the same

Biophysical Journal 107(2) 422–434

Allosteric Pipelines in GPCR Activation

termini with a high overlap (see Materials and Methods) were clustered into channels, which we term allosteric pipelines. Fig. 2, A–C, show the different allosteric pipelines in the inactive, intermediate, and active conformations of b2AR. The thickness of each pipeline is proportional to the number of overlapping allosteric communication pathways. To validate the statistical significance of the differences in allosteric communication pipelines among the three b2AR conformations, we analyzed MD trajectories of control simulations. We compared the allosteric pipelines calculated from two MD simulations starting from the same receptor conformation and with just one residue changed. The two simulations started from the crystal structure of the inactive state (PDB ID: 2RH1) with a single residue D1303.49 protonated (neutral) and deprotonated (charged) (Fig. S2). These two simulations reflect the effect of mutating a charged residue to a neutral one. In these two simulations, the top-scoring allosteric pipeline is the same as that shown in Fig. S2. However, there are other, thinner pipelines connecting D1303.49 that show differences between the two simulations. This is to be expected due to the change of the charged residue to a neutral residue. A comparison of the number of allosteric communication pathways in each pipeline in the three conformational states of b2AR is shown in Table S1. For a given allosteric pipeline, the termini residues of the pathways that are close to one another can be clustered together into termini patches of residues. Thus, the EC residues form a terminus patch communicating with the patch of residues in the G-protein-coupling site, and there could be one thick allosteric

425

pipeline connecting these two patches of residues. This concept is illustrated in Fig. S3. Comparison of the allosteric pipelines for the three conformational states of the b2AR Among the inactive, intermediate, and active states of the receptor, the allosteric pipelines are thickest for the inactive state, followed by the active and intermediate states (shown in Fig. 2, A and C). In the inactive state, a thick allosteric pipeline passes through TM6 (green pipeline in Fig. 2 A), stabilizing it against independent movement and thus keeping the receptor in the inactive conformation. This allosteric pipeline is absent in the intermediate and active states (Fig. 2, B and C). The IC end of TM6 becomes less correlated with the rest of the receptor upon agonist binding, facilitating activation. The narrow but distinct pipeline that passes through TM6 in the active state (gray pipeline in Fig. 2 C) could stabilize the receptor in the active conformation. The active and intermediate conformations show thick pipelines passing through TM7 (cyan and blue pipelines in Fig. 2, B and C, respectively), which indicates that TM7 is increasingly correlated in the active state, unlike TM6, which is more dynamic. In the crystal structures of active b2AR and rhodopsin (3,27), the IC part of TM7 is oriented closer to the receptor core compared with the inactive state, facilitating the outward tilt of TM6. The blue pipeline could thus stabilize TM7 in the active-like conformation. The intermediate and active states of the receptor also show fairly thick allosteric pipelines (shown in cyan and

FIGURE 2 (A–C) Comparison of allosteric pipelines in three conformational states of b2AR. (A) Inverse agonist-bound inactive state. (B) Agonist-bound intermediate state. (C) Agonistand nanobody-bound active state. Each allosteric pipeline is colored differently and the thickness of the pipeline is proportional to the number of allosteric pathways included in it. The number of pipelines and quantitative details are given in Table S1. (D) Schematic view of b2AR showing the difference in allosteric communication between the inverse agonist-bound inactive and agonist-bound intermediate states. The colored arrows indicate significant differences in allosteric pathways among the different functional domains (e.g., orthosteric binding site and G-protein interface). The red arrow indicates an increase in the number of allosteric pathways going from the inactive to the agonist-bound state, and the blue arrow denotes a decrease in allosteric communication between the two states. To see this figure in color, go online.

Biophysical Journal 107(2) 422–434

426

brown in Fig. 2, B and C) from the orthosteric ligand-binding site to helix 8. This is a putative b-arrestin or G-protein receptor kinase-mediated allosteric pipeline that requires further verification from experiments. The putative b-arrestin-mediating allosteric pipeline (cyan pipeline in Fig. 2 B) connecting the agonist-binding pocket with the C terminus is the thickest in the intermediate state. Also, the allosteric pipelines connecting the agonist-binding pocket with the G-protein interface are weakened in the intermediate state compared with the inactive state. These two observations are in agreement with recent NMR studies that showed that conformational coupling between the agonistbinding site and the G-protein interface is weakened in the intermediate state (11). Moreover, the conformational heterogeneity of the intermediate state could facilitate signaling through multiple signaling pathways. In our observations, the cyan pipeline in the intermediate state could facilitate such noncanonical signaling pathways. To compare the allosteric pipelines among the inactive, intermediate, and active states, we plotted the population of each pipeline in decreasing order of magnitude (Fig. S4; Table S1). Whereas the pipeline population in the inactive state showed a rapid decay with the number of pipelines, both the intermediate and active states showed slower decays. Thus, the inactive state shows fewer allosteric pipelines, but each pipeline has a higher population than those in the active state. The intermediate and active states show increased numbers of allosteric pipelines, but each pipeline has a lower population than the inactive-state pipelines. Additionally, the pipelines in the intermediate state show less correlated movement compared with both the inactive and active states, as shown in Fig. 1 D. Thus, the intermediate state with just an agonist bound is a more dynamic state, which shows an overall decrease in correlated movement within the receptor. We also observed the presence of an allosteric pipeline connecting the residues within the IC regions (as shown in blue in Fig. 2 A), which was more dominant in the inactive state of the receptor compared with the intermediate agonist-bound state. This allosteric pipeline could be useful for maintaining the IC regions in a tight conformation to prevent the G-protein binding to the inactive state of the receptor. We also observed such IC allosteric pipelines in the fully active state, possibly to retain the bound G-protein (28). To identify which specific regions of the receptor show correlated movement in the three states, we partitioned the receptor into five regions: 1), the EC loop region; 2), the EC part of the TM region; 3), the orthosteric ligand-binding site; 4), the IC part of the TM region; and 5), the IC loop region. These different regions are shown in Fig. 2 D. We calculated the difference in the number of allosteric communication pathways in the inactive-state ensemble to the intermediate-state ensemble, as shown in Fig. 2 D and Table S2. The red arrows show an increase in the number of allosteric pathways in the intermediate agonist-bound state compared Biophysical Journal 107(2) 422–434

Bhattacharya and Vaidehi

with the inactive state, and the blue arrows show a decrease. In the inactive state of the receptor, the allosteric pipelines are initiated by residues in the EC loop region and terminated by residues in the IC loop regions, thus transmitting the signal through pipelines that keep the receptor intact in the inactive state. However, when an agonist is bound to the orthosteric site, the residues in the orthosteric ligandbinding site become the initiators of allosteric communication to other parts of the receptor, as shown by the red arrows emerging from the ligand-binding site to the EC and IC regions. Simultaneously, there is a weakening of allosteric communication between the EC loops and EC part of the TM region with the G-protein-coupling site. Thus, in the intermediate state, the top half of the receptor becomes decoupled from the bottom half of the receptor, and the receptor becomes more dynamic than in either the active or inactive state. The intermediate state was observed to be dynamic with a low order parameter in recent NMR measurements on agonist-bound b2AR (10). Also in the intermediate state, the G-protein-binding region becomes decoupled from the rest of the receptor and thus acquires additional flexibility, possibly to facilitate the insertion of G-protein. Mutations that affect the ligand efficacy, but not the binding affinity, are identified to be in the allosteric pipelines in b2AR Several single-point mutations that are distant from the orthosteric ligand-binding site but show a substantial increase or decrease in agonist-mediated receptor activation (or constitutive activation) were observed previously in experiments, but the mechanism by which these residues affect the ligand efficacy has hitherto remained unexplained. These mutations and the references from which they were taken (along with the mutation positions that have been shown to increase the thermostability of a specific inactive or active state of the receptor on avian b1AR and human adenosine A2A receptors (29,30)) are shown in Table S3. The criteria for selecting the mutation data are explained in Text S5. We observed that many of these mutations are located along the major allosteric pipelines in one of the three functional states of b2AR. There are experimental mutations leading to inactivation or constitutive activity of the receptor that are not located in the allosteric pipelines or near any of the allosteric pipelines. These mutations could not be explained by the allosteric model, possibly because mutated receptors can give rise to newer allosteric pipelines that are not predicted using MD simulations on the wild-type receptor. The residues from Table S3 are highlighted in Fig. 3, A and B. Fig. 3, A and B, shows the activating and deactivating mutations and the allosteric pipelines mediated by these residues in the inactive and active states, respectively. We postulated that residues that show a decrease in agonist efficacy upon mutation should stabilize the active state of the receptor, and residues whose mutations increase the agonist efficacy

Allosteric Pipelines in GPCR Activation

427

FIGURE 3 (A) Positions of mutations in the literature that increase the agonist-mediated activation, constitutive activity, or thermostability of the active state in class A GPCRs. The associated allosteric pipelines in the inactive state are shown. (B) Mutations that suppress receptor activation or thermostabilize the inactive state and the associated allosteric pipelines in the active state. (C) Number of pathways mediated through each mutation position in b2AR in the three conformations. Green bars represent activating mutations and red bars represent inactivating mutations. (D and E) Location of residues that increase receptor activation upon mutation and the number of pathways mediated through these residues in the (D) inactive and (E) active states. (F and G) Locations of inactivating mutations and the number of pathways mediated in the (E) inactive and (F) active states. To see this figure in color, go online.

should stabilize the inactive state of the receptor. For a given state of the receptor, we identified the residues that mediate multiple pathways in each allosteric pipeline, i.e., the allosteric hubs. Fig. 3 C shows the number of allosteric pathways that pass through the residues listed in Table S3. To test the robustness of these results (if the experimental mutations listed in Table S3 can be placed near any arbitrary pipeline), we reshuffled the residue pairwise MI and

computed allosteric pipelines using the reshuffled MI matrix (Fig. S5 A). Fig. S5 B shows the allosteric hub score (the number of allosteric pathways mediated) of each residue in the receptor for both the true allosteric pipelines and the pipelines computed using the reshuffled MI (termed randomized pipelines in the figure). All of the experimentally mutated residues show significantly higher allosteric hub scores for the true pipelines compared with the randomized Biophysical Journal 107(2) 422–434

428

pipelines. Thus, the location of the experimental mutations near the allosteric pipelines is a significant observation. Seven out of nine mutations that increase the agonistinduced activation, constitutive activity, or thermostability of the active state in class A GPCRs (L2.46, V2.52, F3.27, I3.40, L3.43, S4.53, and F6.44) are located near major allosteric pipelines in the inactive state. Here, we used the Ballesteros-Weinstein residue numbering system for class A GPCRs (31). The first number is the TM helix in which the amino acid is present and the second number is the position of this residue with respect to the most conserved residue in that helix, which is numbered 50. Both L2.46 and V2.52 are located close to D2.50, which is a highly conserved residue in class A GPCRs and is known to be a sodium-binding site (allosteric modulator) in the highresolution adenosine A2A receptor crystal structure (32). Residues I3.40, L3.43, F6.44, and M6.41 form a tight hydrophobic cluster in the middle of the TM region in inactive b2AR (4). These residues mediate major allosteric pipelines in the inactive state of b2AR. In a1BAR, mutating S4.53 to Ala increased inositol phosphate accumulation by 125% without affecting agonist binding. The number of allosteric communication pathways mediated by residues L2.46, V2.52, F3.27, I3.40, L3.43, S4.53, and F6.44 is higher in the inactive state of b2AR, showing that these residues, which are farther away from the ligand-binding site, stabilize the inactive state by mediating allosteric communication. Eight out of 13 mutations that reduce the activity of the receptor show an increase in the number of allosteric pathways in the intermediate or active state compared with the inactive state. Among these residues, T2.37, Y5.58, G7.42, N7.49, and Y7.53 show high allosteric involvement in the active state, whereas T4.56, C6.47, and L7.38 show high allosteric involvement in the intermediate state. The former set of residues could be important for stabilizing the active state once the G-protein binds, and the latter set could be involved in triggering the conformational changes upon agonist binding that lead to activation. Experimentally, mutating both sets of residues leads to a reduction in receptor activation. Five of the 13 inactivating mutations (M2.53, E3.41, L6.37, I7.41, and L7.51) show increased allosteric pathway involvement in the inactive state, which suggests that the mutations should stabilize the active state. Experimentally, however, these mutations all lead to loss of activity of the receptor. E3.41 is located in the middle of TM3 facing TM4. Mutating this residue to Trp increases the thermostability of the inactive state of b2AR (33). Mutating L6.37 to Ala increases agonist-induced activation in a1bAR (34). Both E3.41 and L6.37 mediate allosteric pathways in all three conformational states in b2AR, with higher involvement in the inactive state. Thus, mutating E3.41 or L6.37 can affect the stability of both the inactive and active states. L7.51 is part of the highly conserved NPxxY motif on TM7, which changes conformation by moving inward upon activation. The NPxxY motif is more Biophysical Journal 107(2) 422–434

Bhattacharya and Vaidehi

involved in allosteric communication in the intermediate and active states compared with the inactive state (Fig. 2, A–C). Thus, mutating any residue in the NPxxY region could adversely affect activation. In contrast, the roles of M2.53 and I7.41 are less clear. Mutating M2.53 to Leu increases the thermostability of the inactive state in b1AR. In recent NMR studies, M2.53 was shown to switch between multiple rotameric states in the inactive state of b2AR, and these transitions were suggested to occur on a millisecond timescale (11). Therefore, microseconds of MD simulations may not be sufficient to reproduce allosteric correlations involving M2.53. The effect of mutating I7.41 was weak. In b1AR, mutating I7.41 to Ala slightly increased antagonist binding, and no activation data were reported. Fig. 3 D shows several mutations that increase the activity of the receptor, namely, L2.46, I3.40, L3.43, S4.53, and F6.44, and the associated allosteric pathways in the inactive state of b2AR, and Fig. 3 E shows similar pathways in the active state. These residues are mainly located on TM3 and TM6 and they form a tight hydrophobic cluster that stabilizes the inactive state. The number of pathways mediated by these residues is higher in the inactive state compared with the active state. This explains why mutating these residues leads to increased activation of the mutant receptors. In contrast, many of the inactivating mutations are located on TM7, as shown in Fig. 3, F and G. These residues mediate an increasing number of allosteric pathways in the intermediate and active states compared with the inactive state. Thus, our analysis shows that mutations that affect the ligand efficacy, but not the binding affinity, are located in the allosteric communication pathways. Thus, our findings provide insights into the role of such mutations, which has hitherto been unexplained. Role of residues in the orthosteric site in the allosteric communication pathways We first identified the binding-site residues that are present in the orthosteric binding site of the agonist BI-167017 and the inverse agonist carazolol. The next step was to identify which of these residues are also major participants in allosteric communication by calculating the number of pathways mediated by these residues. Fig. S6, A–D, show the binding-site residues and the allosteric communication pathways they mediate in the intermediate state of the receptor. We chose the agonist-bound intermediate state because this is the state that is dynamic and is in the pathway to full activation. Fig. S6, E–H, show the binding-site residues that mediate more allosteric communication pathways in the active state compared with the inactive state. T195 in EC loop 2 (ECL2) is placed directly above the orthosteric ligand-binding site. This residue is located in multiple allosteric pathways of the intermediate state that connect the EC loop region to the G-protein-coupling region. T195 is also located in the allosteric pathway of the active state that connects the EC region to helix 8, which is part of the carboxy

Allosteric Pipelines in GPCR Activation

terminus that couples to b-arrestin (35). T1103.29, which is located in the agonist-binding site one turn above D1133.32, is involved in extensive allosteric networks involving the EC loops and the G-protein-coupling region (Fig. S6 B). The other two residues that show allosteric pathway involvement in the intermediate state are Y3167.43 and S2075.46. Y3167.43 is also involved in allosteric communication in the active state. This residue directly interacts with the agonist by forming a hydrogen bond with the protonated amine of the ligand. Our observation explains why the mutation of these residues affects the receptor activity (36,37). S2045.43 shows allosteric activity in the active state. S2035.42 and S2075.46 are major agonist contacts in b2AR2 (2,3). S2045.43 does not contact the agonist directly and instead forms a hydrogen bond with N2936.55. When mutated to nonpolar amino acids, both S2035.42 and S2045.43 impair the binding and activity of catecholamine agonists (38,39). F2896.51 and F2906.52, located in the orthosteric site, form aromatic packing with the biphenyl/catechol moieties of b2AR agonists. Mutating F2896.51 to Ala and F2906.52 to Met reduced agonist binding by 1000- and 10-fold, respectively, but did not affect antagonist binding (40). These two residues are involved in allosteric communication with the G-protein-binding site in the active state. Identification of allosteric sites for finding allosteric modulators Designing allosteric modulators has become a current therapeutic strategy for specifically targeting receptors with sub-

429

type specificity (7,41). Computational methods can be used to identify potential binding sites that are different from orthosteric sites in GPCRs (42). However, there is no method to identify whether targeting these potential void spaces in the protein would be effective for allosteric modulation. We have shown in this work that mutation results could be explained by identifying the allosteric hubs that mediate multiple allosteric communication pathways, and the location of these residues must be involved in allosteric communication. Any stable and significant void spaces that form during the dynamics of the receptor and are close to these allosteric hubs could be potential allosteric sites that can be used to design allosteric modulators. Fig. 4 A shows all of the void spaces (blue wireframe) that are detected in the active-state crystal structure of b2AR (PDB ID: 3P0G) and could accommodate small-molecule binders. All of these sites are water accessible as well (Fig. S7). One of these is the orthosteric site (shown in pink). All of the other void spaces correspond to allosteric modulator binding pockets observed in related class A GPCRs. These allosteric pockets are interconnected by allosteric pipelines, shown in Fig. 4 A. Right above the orthosteric site is the allosteric binding site in M2-AcR (PDB ID: 4MQT) (43). In b2AR, there are two major allosteric hubs that are located in this region: F194ECL2 and T195ECL2 (Fig. 4 B). In addition, we also identified five different void spaces that are distant from the orthosteric agonist-binding site and located near the G-protein interface (Fig. 4 C). One of these identified sites (pocket 5) was previously verified experimentally as the Zn2þ-binding site, with Zn2þ being an allosteric

FIGURE 4 (A) Allosteric pockets identified in the active conformation of b2AR, along with the allosteric pipelines that mediate communication among them. The known allosteric binding sites from other class A GPCRs are overlaid. These include a Zn2þ-binding site in b2AR, Naþ-binding site in A2A adenosine receptor, and allosteric modulator LY2119620 in M2-AcR. (B) Expanded view of the allosteric region located in the EC loops of b2AR, showing the key allosteric hub residues. (C) Allosteric pockets near the G-protein interface in b2AR. The residues that define these pockets are highlighted. To see this figure in color, go online.

Biophysical Journal 107(2) 422–434

430

modulator of b2AR activity (44). Pocket 1 is below the orthosteric site and the other pockets are farther down. Pocket 2 is next to the NPxxY motif on TM7, pocket 3 is between TM3 and TM7, and pocket 4 is near TM5 and TM6. Pocket 2 corresponds to the Naþ-binding site in the A2A adenosine receptor (32), which was recently suggested to bind other modulators, such as amiloride (45). Fig. 4 C also shows the allosteric pipelines that are in close proximity to these pockets. The potential small-molecule-binding pockets 2, 3, 4, and 5, which are modulated by the allosteric pipelines, could be allosteric binding pockets. The allosteric hub residues that modulate the interactions of the allosteric pipelines with these pockets are F2826.44, N3227.49, and Y2195.58. Small-molecule modulators designed to bind to these pockets could potentially enhance the activation of b2AR because they could directly affect the allosteric pipelines. Similar allosteric sites identified by this method could be used to design or screen for allosteric modulators of b2AR or other class A GPCRs. DISCUSSION Among the three conformational states of b2AR (inactive, active, and intermediate), the intermediate state shows thinner allosteric pipelines than the inactive state and exchanges less MI among the TM residues compared with the other two states. Thus, the intermediate state is more dynamic compared with the inactive or active state. The receptor dynamics in the agonist-bound intermediate state reveals molecular motion similar to the activation process, i.e., movement of TM5 and TM6 away from TM3 (10,46). The decrease in correlated movement between the EC region and the G-protein-coupling site in the agonist-bound intermediate state makes the IC interface flexible, thus facilitating insertion of the G-protein. The dynamic nature of the

Bhattacharya and Vaidehi

intermediate state agrees with the observation that the orientational order parameter observed in 19F NMR experiments for b2AR is lower in the agonist-bound intermediate state of b2AR compared with the other states (10). When the receptor is in the inactive state, the allosteric pipelines are initiated by residues in the EC regions and they communicate to residues in the IC region of the receptor as shown in Fig. 5 A. This keeps the receptor in the inactive state, with minimal movement of TM6 away from TM3. The involvement of the residues in the orthosteric ligandbinding site in allosteric communication is minimal in the inactive state, as depicted by the schematic picture in Fig. 5 A. Binding of the agonist leads to residues in the agonist-binding site, initiating allosteric communication to the EC regions as well as IC regions of the receptor. In this intermediate state, the agonist-binding site becomes the command center, communicating with different parts of the receptor. In the intermediate state, the receptor is decoupled into the top EC half and bottom IC half and becomes more dynamic, as shown in the schematic in Fig. 5 B. To study the early events in the dynamics of the agonistbound receptor leading to activation, we analyzed the propagation of allosteric signal from the orthosteric site to the G-protein-coupling site in the agonist-bound intermediate receptor conformation. Fig. 6 A shows the allosteric network near the orthosteric ligand-binding site in the intermediate receptor conformation. T1103.29, S2035.42, S2075.46, and Y3167.43 in the orthosteric site receive the allosteric signal from the agonist-binding site and propagate it to several residues located on TM3, TM5, TM6, and TM7. These residues (I1213.40, E1223.41, L2125.51, F2826.44, C2856.47, and F3217.48), which we call the connector region (marked as connector 2 in Fig. 6 D), are highlighted in Fig. 6 A. Among these six residues, I1213.40 and F2826.44 have also been previously shown to have concerted rotamer motion in a

FIGURE 5 Model of b2AR showing the dynamics and variation of allosteric pipelines in the TM domains in the (A) inactive state and (B) agonist-bound intermediate state. The diffusion of agonist into the orthosteric site is schematically depicted as the triggering event for the uncorrelated motions in the EC and IC domains of the TM regions that lead to activation. The arrows indicate the mode of dynamics of the TM regions in the two states. In the inactive state, the strong MI correlation between the EC and IC domains suggests that the entire TM regions move as a whole. While in the agonist-bound state, the EC and IC domains move independently and their motions are governed by the orthosteric site. To see this figure in color, go online.

Biophysical Journal 107(2) 422–434

Allosteric Pipelines in GPCR Activation

431

FIGURE 6 (A) Allosteric network of residues in the intermediate state between the ligand-binding site and connector residues in TM3, TM5, TM6, and TM7. (B) Allosteric network in the active state between TM5 and TM6. (C) Initiator domain in the EC loops with the allosteric communication network involving T195 shown by the orange lines. The crystal structure of M2-AcR is overlaid for comparison. (D) Residues involved in the mechanism of agonist-mediated allosteric communication starting at the EC loops and ending at the IC interface of b2AR. To see this figure in color, go online.

microsecond-timescale dynamics simulation on nanobodybound active b2AR (24). Also, mutating C2856.47 to Ser suppressed activation without affecting agonist binding (47). The allosteric model thus explains the effects of these residues. The residues involved in the allosteric pipelines near TM5 and TM6 are distinctly different in the active state compared with the intermediate state. In the intermediate state, the IC domain of TM6 shows very little allosteric-correlated movement. However, TM6 in the active conformation shows a distinct allosteric pipeline that communicates with TM5 (Fig. 6 B). We also observed that in the active state, M2796.41 on TM6 allosterically communicates with Y2195.58 and L2125.51 on TM5. While we were working on this manuscript, allosteric networks for M2-AcR (22) calculated using accelerated MD simulations were published. In the MD simulations on M2-AcR apoprotein, two major clusters of communicating residues were observed: one involving TM3 and TM5, and the other including TM6 and TM7. These two allosteric clusters were connected through communication networks such as Y2065.58–L3936.41, which correspond to Y2195.58 and M2796.41 in b2AR, and P1985.50–V1113.40, which correspond to P2115.50 and I1213.40 in b2AR. Besides Y2195.58 and M2796.41, I1213.40 forms a connector residue in our calculations that receives the signal from the orthosteric site. Thus, our results correlate with other simulation results (22). Notably, in the M2-AcR crystal structure (48), Y2065.58 (M2-AcR sequence numbering, which is Y2195.58 in b2AR) interacts with the residues in the VTIL motif on TM6, which is one turn below L3936.41 (M2796.41 in b2AR). It was shown by site-directed mutagenesis that the interaction of Y2065.58 with VTIL is critical for G-protein selectivity (preference for Gi/o over Gq) (49). Therefore, the allosteric network involving Y2195.58 and M2796.41 is important not only for the stability of the G-protein-bound receptor conformation but also for the specificity of the G-protein-coupling conformation.

In the biogenic amine family of GPCRs, the EC loops possibly form the ligand entry and exit domain. The EC loops are widely speculated to contain a putative allosteric-modulator-binding site; however, so far, only the muscarinic AcRs have been found to bind allosteric modulators at this site (50). In b2AR, the EC surface has been shown to react differently to agonists and antagonists by modulating a salt bridge between K3057.32 and D192ECL2 (51). It was proposed that F193ECL2 next to D192ECL2 interacts directly with antagonists and thus propagates allosteric signal from the EC loops to the orthosteric site. In our calculations, T195 in ECL2 is a major initiator of allosteric communication in the intermediate and active receptor conformations. T195ECL2 is close to F193ECL2 and shows correlated torsion angle motion with F193ECL2. In turn, F193ECL2 communicates with T1103.29 and D1133.32 in the agonist-binding orthosteric site (Fig. 6 C). To correlate the allosteric network of residues we identified in ECL2 with the allosteric modulator site speculated in M2-AcR, we overlaid the M2-AcR crystal structure (PDB ID: 3UON) over b2AR (PDB ID: 3P0G) shown in Fig. 6 C. We found that T195ECL2 and F193ECL2 are in the same locations as F181ECL2 and Y177ECL2 in M2-AcR. In microseconds of MD simulations with M2-AcR, the agonist tiotropium transiently binds to an allosteric site consisting of F181 and Y177 before diffusing into the orthosteric site (52). Moreover, Y177 is a known key contact for modulating agonist binding in M2-AcR (50) and breaking of the salt bridge between K3057.32 and D192ECL2 in b2AR (51). Also, in recent MD simulations by Dror et al. (53), the residues Y177ECL2, W4227.35, Y802.61, and Y832.64 were shown to be the major contacts for allosteric modulators in M2-AcR. In b2AR, residue Y3087.35 is located near major allosteric hubs that communicate with the G-protein interface. However, residues G902.61 and H932.64 show more allosteric involvement in the inactive state compared with the active state. Together, these results indicate that allosteric communication originates at the EC loops at residues such as Biophysical Journal 107(2) 422–434

432

T195ECL2, F193ECL2, and Y3087.35, propagating to the orthosteric site via V872.58, F892.60, T1644.56, and S1654.57, which form the connector 1 region in Fig. 6 D. From here, the communication signal reaches the distribution hub in the orthosteric site consisting of T1103.29, S2035.42, S2075.46, and Y3167.43. The signal is then distributed to the connector 2 region consisting of I1213.40, E1223.41, L2125.51, F2826.44, C2856.47, and F3217.48 (Fig. 6 D), and communicated to L752.46 I1544.46, M2155.54, Y2195.58, I2786.40, M2796.41, and Y3267.53 before terminating at the G-protein-coupling region and helix 8 or the carboxy terminus. Thus, a complete chain of communication exists between the EC loops and the G-protein-coupling region at the IC interface, modulated in the middle by residues in the agonist-binding orthosteric site. This is schematically depicted in Fig. 6 D. CONCLUSIONS In this work, we mapped the allosteric communication pipelines in b2AR that transmit the activation signal from the orthosteric site to the IC interface of the receptor. We compared the allosteric pipelines among three receptor states, termed the inverse agonist-bound inactive, agonist-bound intermediate, and agonist- and nanobody (G-protein mimic)-bound active states. Whereas the inactive state shows a few thick allosteric pipelines, the intermediate and active states show many thin pipelines. There is an overall decrease in allosteric-correlated movement between residues in the intermediate state, indicating the dynamic nature of the intermediate state. The allosteric pipelines extend from the EC to the IC regions, keeping the receptor in the inactive state. These pipelines are weakened in the agonist-bound intermediate state, with more pipelines emerging from the orthosteric ligandbinding site to the rest of the receptor in this state. Thus, the intermediate state is more dynamic than the inactive or active state of the receptors. This increased dynamics would be necessary to trigger the conformational changes in the receptor that leads to activation. We identified the key residues and residue networks involved in the continuous communication pipelines from the EC loops to the G-protein-coupling site that leads to receptor activation (Fig. 6 D). The allosteric signal originates at residues T195 and F193 in ECL2, which corresponds to the transient agonist (and allosteric modulator)-binding site in M2-AcR. A major portion of the allosteric signal then propagates to the IC ends of TM3 and TM6, which contain the engineered Zn2þ-binding site, a positive allosteric modulator in b2AR. These results provide an understanding of the mechanism of allosteric communication in class A GPCRs. Several residues in the ligand-binding orthosteric site allosterically communicate with the G-protein interface in the intermediate state. We observed that these residues play a key role in allosteric communication. In addition, our analysis showed that mutations that affect Biophysical Journal 107(2) 422–434

Bhattacharya and Vaidehi

the ligand efficacy, but not the binding affinity, are hub residues or located near hub residues in the allosteric communication pipelines. Thus, our findings provide insight into the role of such mutations, which has hitherto been unexplained, and the resulting information can now be used to stabilize the receptor in specific signaling states. Since many class A GPCRs have structural homology, these results could also provide insight into other class A GPCRs. SUPPORTING MATERIAL Supporting Materials and Methods, eight figures, three tables, and four equations are available at http://www.biophysj.org/biophysj/supplemental/ S0006-3495(14)00624-9. The authors thank Dr. Christofer Tautermann and Michiel Niesen for helpful suggestions regarding the analysis of MI. The Python programs for calculating MI were written by Jeffrey Wagner and Michiel Niesen. We thank Dr. Ron Dror and members of DE Shaw Research who shared the MD simulation trajectories and provided some of their scripts for converting the trajectories. This work was supported by the National Institutes of Health (grant NIHRO1-GM097261).

SUPPORTING CITATIONS References (54–63) appear in the Supporting Material.

REFERENCES 1. Dore´, A. S., N. Robertson, ., F. H. Marshall. 2011. Structure of the adenosine A(2A) receptor in complex with ZM241385 and the xanthines XAC and caffeine. Structure. 19:1283–1293. 2. Rasmussen, S. G. F., H. J. Choi, ., B. K. Kobilka. 2011. Structure of a nanobody-stabilized active state of the b(2) adrenoceptor. Nature. 469:175–180. 3. Rasmussen, S. G. F., B. T. DeVree, ., B. K. Kobilka. 2011. Crystal structure of the b2 adrenergic receptor-Gs protein complex. Nature. 477:549–555. 4. Rosenbaum, D. M., V. Cherezov, ., B. K. Kobilka. 2007. GPCR engineering yields high-resolution structural insights into b2-adrenergic receptor function. Science. 318:1266–1273. 5. Warne, T., M. J. Serrano-Vega, ., G. F. X. Schertler. 2008. Structure of a b1-adrenergic G-protein-coupled receptor. Nature. 454:486–491. 6. White, J. F., N. Noinaj, ., R. Grisshammer. 2012. Structure of the agonist-bound neurotensin receptor. Nature. 490:508–513. 7. Conn, P. J., A. Christopoulos, and C. W. Lindsley. 2009. Allosteric modulators of GPCRs: a novel approach for the treatment of CNS disorders. Nat. Rev. Drug Discov. 8:41–54. 8. Audet, M., and M. Bouvier. 2012. Restructuring G-protein-coupled receptor activation. Cell. 151:14–23. 9. Kobilka, B. K. 2007. G protein coupled receptor structure and activation. Biochim. Biophys. Acta. 1768:794–807. 10. Kim, T. H., K. Y. Chung, ., R. S. Prosser. 2013. The role of ligands on the equilibria between functional states of a G protein-coupled receptor. J. Am. Chem. Soc. 135:9465–9474. 11. Nygaard, R., Y. Zou, ., B. K. Kobilka. 2013. The dynamic process of b(2)-adrenergic receptor activation. Cell. 152:532–542. 12. Su¨el, G. M., S. W. Lockless, ., R. Ranganathan. 2003. Evolutionarily conserved networks of residues mediate allosteric communication in proteins. Nat. Struct. Biol. 10:59–69.

Allosteric Pipelines in GPCR Activation 13. Madabushi, S., A. K. Gross, ., O. Lichtarge. 2004. Evolutionary trace of G protein-coupled receptors reveals clusters of residues that determine global and class-specific functions. J. Biol. Chem. 279:8126– 8132. 14. Chennubhotla, C., and I. Bahar. 2007. Signal propagation in proteins and relation to equilibrium fluctuations. PLOS Comput. Biol. 3:1716– 1726. 15. Zheng, W., B. R. Brooks, and D. Thirumalai. 2006. Low-frequency normal modes that describe allosteric transitions in biological nanomachines are robust to sequence variations. Proc. Natl. Acad. Sci. USA. 103:7664–7669. 16. Daily, M. D., T. J. Upadhyaya, and J. J. Gray. 2008. Contact rearrangements form coupled networks from local motions in allosteric proteins. Proteins. 71:455–466. 17. Stacklies, W., F. Xia, and F. Gra¨ter. 2009. Dynamic allostery in the methionine repressor revealed by force distribution analysis. PLOS Comput. Biol. 5:e1000574. 18. Pandini, A., A. Fornili, ., J. Kleinjung. 2012. Detection of allosteric signal transmission by information-theoretic analysis of protein dynamics. FASEB J. 26:868–881. 19. Rivalta, I., M. M. Sultan, ., V. S. Batista. 2012. Allosteric pathways in imidazole glycerol phosphate synthase. Proc. Natl. Acad. Sci. USA. 109:E1428–E1436. 20. McClendon, C. L., G. Friedland, ., M. P. Jacobson. 2009. Quantifying correlations between allosteric sites in thermodynamic ensembles. J. Chem. Theory Comput. 5:2486–2502. 21. Sethi, A., J. Eargle, ., Z. Luthey-Schulten. 2009. Dynamical networks in tRNA:protein complexes. Proc. Natl. Acad. Sci. USA. 106:6620– 6625. 22. Miao, Y., S. E. Nichols, ., J. A. McCammon. 2013. Activation and dynamic network of the M2 muscarinic receptor. Proc. Natl. Acad. Sci. USA. 110:10982–10987. 23. Dror, R. O., D. H. Arlow, ., D. E. Shaw. 2009. Identification of two distinct inactive conformations of the b2-adrenergic receptor reconciles structural and biochemical observations. Proc. Natl. Acad. Sci. USA. 106:4689–4694.

433 the helix 3/helix 6 interface in receptor activation. Mol. Pharmacol. 61:1025–1032. 35. Krasel, C., U. Zabel, ., M. J. Lohse. 2008. Dual role of the b2-adrenergic receptor C terminus for the binding of b-arrestin and receptor internalization. J. Biol. Chem. 283:31840–31848. 36. Strader, C. D., I. S. Sigal, and R. A. Dixon. 1989. Structural basis of b-adrenergic receptor function. FASEB J. 3:1825–1832. 37. Cavalli, A., F. Fanelli, ., S. Cotecchia. 1996. Amino acids of the a1b-adrenergic receptor involved in agonist binding: differences in docking catecholamines to receptor subtypes. FEBS Lett. 399:9–13. 38. Sato, T., H. Kobayashi, ., H. Kurose. 1999. Ser203 as well as Ser204 and Ser207 in fifth transmembrane domain of the human b2-adrenoceptor contributes to agonist binding and receptor activation. Br. J. Pharmacol. 128:272–274. 39. Liapakis, G., J. A. Ballesteros, ., J. A. Javitch. 2000. The forgotten serine. A critical role for Ser-2035.42 in ligand binding to and activation of the b2-adrenergic receptor. J. Biol. Chem. 275:37779–37788. 40. Dixon, R. A., I. S. Sigal, and C. D. Strader. 1988. Structure-function analysis of the b-adrenergic receptor. Cold Spring Harb. Symp. Quant. Biol. 53:487–497. 41. Wang, C. I. A., and R. J. Lewis. 2013. Emerging opportunities for allosteric modulation of G-protein coupled receptors. Biochem. Pharmacol. 85:153–162. 42. Ivetac, A., and J. A. McCammon. 2010. Mapping the druggable allosteric space of G-protein coupled receptors: a fragment-based molecular dynamics approach. Chem. Biol. Drug Des. 76:201–217. 43. Kruse, A. C., A. M. Ring, ., B. K. Kobilka. 2013. Activation and allosteric modulation of a muscarinic acetylcholine receptor. Nature. 504:101–106. 44. Swaminath, G., T. W. Lee, and B. Kobilka. 2003. Identification of an allosteric binding site for Zn2þ on the b2 adrenergic receptor. J. Biol. Chem. 278:352–356. 45. Gutie´rrez-de-Tera´n, H., A. Massink, ., R. C. Stevens. 2013. The role of a sodium ion binding site in the allosteric modulation of the A(2A) adenosine G protein-coupled receptor. Structure. 21:2175–2185.

24. Dror, R. O., D. H. Arlow, ., D. E. Shaw. 2011. Activation mechanism of the b2-adrenergic receptor. Proc. Natl. Acad. Sci. USA. 108:18684– 18689.

46. Niesen, M. J., S. Bhattacharya, and N. Vaidehi. 2011. The role of conformational ensembles in ligand recognition in G-protein coupled receptors. J. Am. Chem. Soc. 133:13197–13204.

25. MacKerell, A. D., D. Bashford, ., M. Karplus. 1998. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B. 102:3586–3616.

47. Fraser, C. M. 1989. Site-directed mutagenesis of b-adrenergic receptors. Identification of conserved cysteine residues that independently affect ligand binding and receptor activation. J. Biol. Chem. 264:9266–9270.

26. Dijkstra, E. W. 1959. A note on two problems in connexion with graphs. Numer. Math. 1:269–271. 27. Choe, H. W., Y. J. Kim, ., O. P. Ernst. 2011. Crystal structure of metarhodopsin II. Nature. 471:651–655.

48. Haga, K., A. C. Kruse, ., T. Kobayashi. 2012. Structure of the human M2 muscarinic acetylcholine receptor bound to an antagonist. Nature. 482:547–551.

28. Shenoy, S. K., and R. J. Lefkowitz. 2011. b-Arrestin-mediated receptor trafficking and signal transduction. Trends Pharmacol. Sci. 32: 521–533.

49. Liu, J., B. R. Conklin, ., J. Wess. 1995. Identification of a receptor/ G-protein contact site critical for signaling specificity and G-protein activation. Proc. Natl. Acad. Sci. USA. 92:11642–11646.

29. Serrano-Vega, M. J., F. Magnani, ., C. G. Tate. 2008. Conformational thermostabilization of the b1-adrenergic receptor in a detergent-resistant form. Proc. Natl. Acad. Sci. USA. 105:877–882.

50. Bock, A., N. Merten, ., K. Mohr. 2012. The allosteric vestibule of a seven transmembrane helical receptor controls G-protein coupling. Nat. Commun. 3:1044.

30. Lebon, G., K. Bennett, ., C. G. Tate. 2011. Thermostabilisation of an agonist-bound conformation of the human adenosine A(2A) receptor. J. Mol. Biol. 409:298–310.

51. Bokoch, M. P., Y. Zou, ., B. K. Kobilka. 2010. Ligand-specific regulation of the extracellular surface of a G-protein-coupled receptor. Nature. 463:108–112.

31. Ballesteros, J., and H. Weinstein. 1995. [19] Integrated methods for the construction of three-dimensional models and computational probing of structure-function relations in G protein-coupled receptors. Methods Neurosci. 25:366–428.

52. Kruse, A. C., J. Hu, ., B. K. Kobilka. 2012. Structure and dynamics of the M3 muscarinic acetylcholine receptor. Nature. 482:552–556.

32. Liu, W., E. Chun, ., R. C. Stevens. 2012. Structural basis for allosteric regulation of GPCRs by sodium ions. Science. 337:232–236.

53. Dror, R. O., H. F. Green, ., D. E. Shaw. 2013. Structural basis for modulation of a G-protein-coupled receptor by allosteric drugs. Nature. 503:295–299.

33. Hanson, M. A., V. Cherezov, ., R. C. Stevens. 2008. A specific cholesterol binding site is established by the 2.8 A structure of the human b2-adrenergic receptor. Structure. 16:897–905.

54. Barak, L. S., L. Me´nard, ., M. G. Caron. 1995. The conserved seven-transmembrane sequence NP(X)2,3Y of the G-protein-coupled receptor superfamily regulates multiple properties of the b2-adrenergic receptor. Biochemistry. 34:15407–15414.

34. Greasley, P. J., F. Fanelli, ., S. Cotecchia. 2002. Mutagenesis and modelling of the alpha(1b)-adrenergic receptor highlight the role of

55. Beukers, M. W., J. van Oppenraaij, ., A. P. IJzerman. 2004. Random mutagenesis of the human adenosine A2B receptor followed by growth Biophysical Journal 107(2) 422–434

434

Bhattacharya and Vaidehi selection in yeast. Identification of constitutively active and gain of function mutations. Mol. Pharmacol. 65:702–710.

56. Kikkawa, H., M. Isogaya, ., H. Kurose. 1998. The role of the seventh transmembrane region in high affinity binding of a b2-selective agonist TA-2005. Mol. Pharmacol. 53:128–134. 57. Killian, B. J., J. Y. Kravitz, ., M. K. Gilson. 2009. Configurational entropy in protein-peptide binding: computational study of Tsg101 ubiquitin E2 variant domain with an HIV-derived PTAP nonapeptide. J. Mol. Biol. 389:315–335. 58. Savarese, T. M., C. D. Wang, and C. M. Fraser. 1992. Site-directed mutagenesis of the rat m1 muscarinic acetylcholine receptor. Role of conserved cysteines in receptor function. J. Biol. Chem. 267:11439– 11448. 59. Shannon, C. E. 1948. The mathematical theory of communication. Bell Syst. Tech. J. 27:379–423.

Biophysical Journal 107(2) 422–434

60. Steuer, R., J. Kurths, ., J. Selbig. 2002. The mutual information: detecting and evaluating dependencies between variables. Bioinformatics. 18 (Suppl 2):S231–S240. 61. Sugimoto, Y., R. Fujisawa, ., H. Kurose. 2002. b (1)-Selective agonist (-)-1-(3,4-dimethoxyphenetylamino)-3-(3,4-dihydroxy)-2-propanol [(-)-RO363] differentially interacts with key amino acids responsible for b (1)-selective binding in resting and active states. J. Pharmacol. Exp. Ther. 301:51–58. 62. Tao, Y. X., A. N. Abell, ., D. L. Segaloff. 2000. Constitutive activation of G protein-coupled receptors as a result of selective substitution of a conserved leucine residue in transmembrane helix III. Mol. Endocrinol. 14:1272–1282. 63. Warne, T., R. Moukhametzianov, ., C. G. Tate. 2011. The structural basis for agonist and partial agonist action on a b(1)-adrenergic receptor. Nature. 469:241–244.

Differences in allosteric communication pipelines in the inactive and active states of a GPCR.

G-protein-coupled receptors (GPCRs) are membrane proteins that allosterically transduce the signal of ligand binding in the extracellular (EC) domain ...
2MB Sizes 0 Downloads 3 Views