adrenergic receptor

0 downloads 0 Views 1MB Size Report
May 17, 2011 - Both CL3 conformations are observed (in different monomers) in the carazolol-bound and in the iodocyanopindolol- bound structures ...
Two distinct conformations of helix 6 observed in antagonist-bound structures of a β1-adrenergic receptor Rouslan Moukhametzianov, Tony Warne, Patricia C. Edwards, Maria J. Serrano-Vega, Andrew G. W. Leslie, Christopher G. Tate1, and Gebhard F. X. Schertler,1,2 Medical Research Council Laboratory of Molecular Biology, Hills Road, Cambridge CB2 0QH, United Kingdom Edited by Martin J. Lohse, Universitat Wurzburg Germany, Wurzburg, Germany, and accepted by the Editorial Board March 25, 2011 (received for review January 5, 2011)

The β1 -adrenergic receptor (β1 AR) is a G-protein-coupled receptor whose inactive state structure was determined using a thermostabilized mutant (β1 AR–M23). However, it was not thought to be in a fully inactivated state because there was no salt bridge between Arg139 and Glu285 linking the cytoplasmic ends of transmembrane helices 3 and 6 (the R3.50 − D∕E6.30 “ionic lock”). Here we compare eight new structures of β1 AR–M23, determined from crystallographically independent molecules in four different crystals with three different antagonists bound. These structures are all in the inactive R state and show clear electron density for cytoplasmic loop 3 linking transmembrane helices 5 and 6 that had not been seen previously. Despite significantly different crystal packing interactions, there are only two distinct conformations of the cytoplasmic end of helix 6, bent and straight. In the bent conformation, the Arg139Glu285 salt bridge is present, as in the crystal structure of darkstate rhodopsin. The straight conformation, observed in previously solved structures of β-receptors, results in the ends of helices 3 and 6 being too far apart for the ionic lock to form. In the bent conformation, the R3.50 − E6.30 distance is significantly longer than in rhodopsin, suggesting that the interaction is also weaker, which could explain the high basal activity in β1 AR compared to rhodopsin. Many mutations that increase the constitutive activity of G-protein-coupled receptors are found in the bent region at the cytoplasmic end of helix 6, supporting the idea that this region plays an important role in receptor activation. membrane protein ∣ seven-transmembrane receptor ∣ conformational thermostabilization

T

he G-protein-coupled receptor (GPCR) superfamily is responsible for the transduction of a wide variety of extracellular signals, and they are found in almost all eukaryotes, with over 800 different GPCRs in humans (1, 2). Their common seventransmembrane-domain architecture has been preserved and evolved to sense agonists with remarkable structural diversity and to couple to a variety of specific signaling pathways. The structures of rhodopsin (3–7), the β1 - and β2 -adrenergic receptors (β1 AR and β2 AR) (8–11). the adenosine A2A receptor (A2A R) (12), the dopamine D3 receptor (13), and the CXCR4 chemokine receptor (14) show that there is a similar overall architecture for all family A GPCRs (reviewed in refs. 15–17), and there is probably a corresponding underlying conservation of the activation process (18, 19). With this wealth of structural information, the challenge is now to correlate the pharmacological properties of the hormone-binding receptors with their structure. One functional aspect that still requires a structural explanation is the phenomenon of basal activity—i.e., most GPCRs show some ability to activate G proteins even in the absence of agonists (20, 21) —and this will be the focus of this report. Our current understanding of the structural basis for receptor activation is based largely on the structures determined for the dark-state of rhodopsin (R state) (3–5) and its light-activated conformation, opsin (R* state) (6, 7). Rhodopsin and its close 8228–8232 ∣ PNAS ∣ May 17, 2011 ∣ vol. 108 ∣ no. 20

homologues are unique amongst GPCRs because they contain a covalently bound inverse agonist, 11-cis retinal, which is converted by a photon of light to all-trans retinal, which is a full agonist for the receptor. Thus rhodopsin functions as a binary switch, the photon of light converting an inactive form of receptor to one that is fully competent to activate G proteins (reviewed in ref. 22). It is essential that rhodopsin shows no G-protein-coupling activity in the absence of light and, therefore, it has evolved to have absolutely no basal activity within the lifetime of rhodopsin in the cell. Consequently, it is accepted that the structure of dark-state rhodopsin represents the fully inactive state of a GPCR. Upon activation by light, numerous conformational changes around the retinal binding pocket result in the cytoplasmic end of transmembrane helix 6 (H6) straightening and moving 6 Å away from the center of the molecule, accompanied by a smaller movement and rotation of transmembrane helix 5 (H5). The movements of H5 and H6 create a cleft in the cytoplasmic face of the opsin that allows the binding of the C-terminal helix of Gα, thus activating the G protein. A key feature of rhodopsin activation is the outward movement of H6 away from H3 (6, 7, 23). In the rhodopsin structure, there is a salt-bridge R3.50 − E6.30 (Ballesteros-Weinstein numbering system is in superscripts; ref. 24), which links the cytoplasmic ends of H6 and H3. Both R3.50 and D∕E6.30 are highly conserved throughout the GPCR superfamily and, prior to the structure determination of rhodopsin, had already been suggested to provide a constraint that stabilized the inactive state of GPCRs, leading to this interaction being termed the ionic lock (25). In the case of rhodopsin, the R3.50 − E6.30 salt bridge was proposed to be one of the major determinants for the stability of the inactive state. Hormone receptors have a far more complex pharmacology than does rhodopsin. Many hormone-binding GPCRs can partially activate their respective G proteins in the absence of agonist ligand (26). The extent of this activity varies among GPCRs and their subtypes, reflecting different needs for a basal level of signaling in a physiological context. Receptors with no basal activity, Author contributions: R.M., T.W., C.G.T., and G.F.X.S. designed research; R.M., T.W., P.C.E., and M.J.S.-V. performed research; R.M. and A.G.W.L. analyzed data; and R.M., A.G.W.L., C.G.T., and G.F.X.S. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. M.J.L. is a guest editor invited by the Editorial Board. Freely available online through the PNAS open access option. Data deposition: The atomic coordinates and structure factors have been deposited in the Protein Data Bank, www.pdb.org [PDB ID codes 2YCW (t1118), 2YCX (t148), 2YCY (t468), 2YCZ (t756)]. 1

To whom correspondence may be addressed. E-mail: [email protected] and cgt@ mrc-lmb.cam.ac.uk.

2

Present address: Paul Scherrer Institut, Laboratory of Biomolecular Research, OFLC 109, CH-5232 Villigen PSI, Switzerland.

This article contains supporting information online at www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1100185108/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1100185108

such as rhodopsin, seem appropriate for mediating sensory inputs that control how well we can see in the dark, where any basal activity would add undesirable noise. In contrast, increased levels of basal activity may be more useful for the controlled modulation of an activity such as heart rate that varies over a range of only two- to fivefold. It is thought that basal activity is caused by a subset of receptor conformations, which resemble the fully activated agonist-bound state despite the absence of bound agonist, and that these can bind and activate their respective G proteins. Recent structures of β1 AR and β2 AR have greatly increased our understanding of how ligands are recognized and the overall architecture of the receptors (8–11, 27). However, there are also many questions that remain unresolved. One feature of all of the β-receptor structures is that they appear to represent inactive states (i.e., they are most similar to the structure of dark-state rhodopsin), yet the ionic lock between helices 3 and 6 is invariably broken. This observation raises the question of whether the β-receptors can adopt a conformation analogous to dark-state rhodopsin and, if so, how this affects their basal activity. Existing β-receptor crystal structures cannot address this issue, because either the cytoplasmic end of H6 is disordered and unresolved or it is potentially perturbed by the interaction with an antibody fragment or a T4 lysozyme fusion. Here we present eight structures of the β1 AR, which show that the receptor can indeed form an inactive state with an ionic lock analogous to the dark-state of rhodopsin, in addition to an inactive state where the ionic lock is broken. Crystallization and Structure Determination. The turkey β1 AR construct that was crystallized contains amino acid deletions in the flexible regions at the N and C termini and in cytoplasmic loop 3 (CL3), and is called β36 (28). Six thermostabilizing mutations were introduced to make the receptor more tolerant to shortchain detergents (29), resulting in the β36–m23 construct whose structure was determined to 2.7-Å resolution with cyanopindolol bound (11). During the extensive screening required to obtain these crystals (29), lower resolution datasets were collected from a variety of crystals containing either bound cyanopindolol (crystals t148, t468, resolution limits 3.25 and 3.15 Å) or iodocyanopindolol (crystal t756, resolution limit 3.65 Å). Cocrystallization of β36–m23 with the high-affinity antagonist carazolol also led to crystals suitable for structure determination (crystal t1118, resolution limit 3.00 Å). All structures were solved by molecular replacement using the structure of β36–m23 (PDB ID code 2VT4) as a model. There are two monomers (A and B) in the crystallographic asymmetric unit of each crystal, resulting in a total of eight structures for β36–m23. Data collection and refinement statistics are shown in Table S1. Crystal Packing Interactions. Crystals of β36–m23 containing bound carazolol grew in the monoclinic space group P21 , whereas crystals of β36–m23 with cyanopindolol or iodocyanopindolol bound grew in the monoclinic space group C2 and displayed significant variations in lattice parameters (Table S1) as a result of different cryoprotection protocols (Methods; ref. 28). The packing interactions between receptors in the four crystals are quite different, although in each case the crystals appear to be built from similar nonphysiological antiparallel dimers with the dimer interface comprising H4 and H5 (Fig. S1). Analysis of the crystal packing interactions (Table S2) shows that the third CL3 region linking transmembrane helices 5 and 6 is always involved in intermolecular contacts (although to different extents) and suggests that one of the two distinct conformations of this region (see following section) is selected by the ability to provide stabilizing crystal contacts (Fig. 1). Both CL3 conformations are observed (in different monomers) in the carazolol-bound and in the iodocyanopindololbound structures, demonstrating that the different conformations Moukhametzianov et al.

BIOCHEMISTRY

Results

Fig. 1. Crystal contacts involving CL3. Lattice contacts between different monomers of the antiparallel crystallographic dimer are compared for different crystal forms. The thickness of the strand representing the Cα backbone reflects the temperature factor (increasing thickness reflects higher B values). Residues making crystal contacts are shown explicitly. In the case of t148, t468, t756 C2 monoclinic crystals, CL3 is in contact with CL3 of the symmetry related molecule.

are not the result of different crystal growth conditions. In addition, the two conformations show no correlation to the nature of antagonist ligand. The importance of the CL3 region in forming lattice contacts probably explains the success of the deletions made to enhance crystallization (residues 244–271 and 277–278). The full length CL3 is likely to be conformationally mobile, hindering formation of a stable crystal lattice. Overall Comparison of the β1 AR Structures The eight receptor structures, regardless of the type of antagonist bound, are very similar to the previously reported cyanopindololbound structure with respect to the hydrophobic transmembrane regions (residues 54–67, 78–101, 114–143, 158–179, 208–234, 289–306, 323–345) with rmsds for Cα atoms ranging from 0.03 to 0.53 Å (Table S3). There are also only minor differences in the previously defined extracellular and intracellular loops (rmsds in Cα atoms 0.6–0.7 Å). However, in the case of CL3, there were substantial differences between some of the monomers. CL3 comprises the cytoplasmic extensions to transmembrane helices H5 and H6 and a relatively short loop in an extended conformaPNAS ∣

May 17, 2011 ∣

vol. 108 ∣

no. 20 ∣

8229

tion between them. In the previous structure of β1 AR (PDB ID code 2VT4), 15 residues in the CL3 region (239–283, but note that residues 244–271 and 277–278 are not in the construct) were disordered and therefore not modeled (11). Amongst the eight structures reported here, two distinct conformations of the CL3 region were observed (Fig. 2). Both conformations are clearly in the inactive R state as neither show the characteristic conformation changes observed in opsin. Structural Differences in CL3 and H6 Transmembrane helix H6 in the published structure of cyanopindolol-bound β1 AR–M23 (11) extends from residue Glu2856.30 to Phe3156.60 , with a single kink formed by Pro3056.50 . The cytoplasmic membrane boundary is estimated to be in the region of Ala288 to Leu289, so the cytoplasmic helical extension of H6 is relatively short (5–6 residues). In one of the two distinct conformations of CL3 observed in the new structures, the conformation of H6 is unchanged but the cytoplasmic extension is longer by approximately 10 Å, starting at residue Lys276. In the second conformation, there is a marked kink in the cytoplasmic extension of H6 at residue Ala2886.33 (t1118, t148) or Lys2876.32 (t756). Residues 274–286 are still α-helical and the helix is just as long, but the helix axis is bent by about 35° toward CL2 and rotated about 10° around the helix axis clockwise if viewed from the cytoplasmic side. The two conformations are therefore referred to as “straight” or “bent” accordingly (Fig. 3). The folding of the

Fig. 3. Interhelical polar contacts at the cytoplasmic face of β1 AR–M23. (A) The structure of the bent conformation is depicted, with the intact ionic lock between Arg139 and Glu285. (B) The structure of the straight conformation is shown, with the ionic lock broken. Structures are shown as cartoons in rainbow coloration (N terminus in blue, C terminus in red), except for CL2 and the cytoplasmic extension of H6 which are in gray and residues predicted to form interhelical hydrogen bonds shown as sticks.

Fig. 2. Two conformations of CL3. (A) Superposition of the straight and bent H6 conformations of the carazolol-bound structures of β1 AR–M23 (crystal t1118) colored according to the B factor (increasing values from blue to red). (B) Close up of the CL3 region from panel A. (C and D) Omit densities of the CL3 for the bent and straight conformations, respectively (t1118 crystal). Densities are contoured either at (C) 0.8σ or (D) 0.7σ (gray mesh), along with the Cα trace (red). 8230 ∣

www.pnas.org/cgi/doi/10.1073/pnas.1100185108

CL3 region is highly conserved within these two conformations (rmsds in Cα atoms of 0.28–0.39 Å for four structures in the straight conformation and 0.97–1.21 Å for four structures in the bent conformation; Fig. 2, Tables S3 and S4). In structures with H6 in the bent conformation, H5 is extended by one helical turn relative to H6 in the straight conformation. The remaining residues Lys240–Ser273 (bent conformation), Gln237–Arg275 (straight conformation) form an extended loop. The omit densities for the CL3 region in the carazolol structure are shown in Fig. 2. The electron density for the bent conformation is systematically better defined in all structures. The straight conformation has breaks in the density that correspond to the residues with very high B factors in this region. The bent conformation of H6 is partially stabilized by van der Waals interactions primarily involving Met281 and Leu282 with side chains on H5. The salt bridge between the highly conserved Arg1393.50 of the E(D)RY motif on H3 and Glu2856.30 on H6 (the ionic lock) will provide additional stabilization. However, the minimum distance between a nitrogen atom of the guanidinium group of Arg1393.50 and a carboxylate oxygen of Glu2856.30 Moukhametzianov et al.

Discussion Crystallization of β1 AR–M23 in different crystal forms with three different ligands has resulted in receptor structures that are virtually identical except for the different conformations of the cytoplasmic end of H6. The fact that only two conformations of H6 are observed, either bent or straight, despite the range of different crystallization conditions and crystal packing interactions, suggests that these reflect unbiased conformations that could be adopted in the native cell membranes. This idea is consistent with current thinking that GPCRs are in a finely balanced dynamic equilibrium between multiple inactive states and multiple activated states, even in the absence of ligands (18, 19). Recent molecular dynamics simulations performed on the β2 AR also indicate the coexistence of two different inactive conformations with the ionic lock either intact or disrupted (30). The conformation corresponding to an intact ionic lock in the β2 AR simulation is broadly similar to that observed in the crystal structures, except that H6 was described as kinking slightly at Gly2766.38 compared with the more marked kink observed at Lys2876.32 ∕Ala2886.33 in β1 AR. Given that the bent conformation is very similar to that observed in rhodopsin (Figs. S2 and S3), this structure probably represents the most inactive state of receptor —i.e., the conformation that would be expected if an inverse agonist were bound (Rinv ). In this conformation, the cytoplasmic extension of H6 blocks the G-protein peptide binding site observed in the low-pH structure of opsin (7). In the straight conformation, there is still insufficient room for the C-terminal peptide of the G protein to insert into the receptor, so further conformational changes upon agonist binding are required for productive G-protein coupling. Thus transformation of H6 by the conformational dynamics of the H2866.31 KAL2896.34 region is one of the essential steps required for G-protein signaling. Rhodopsin is unusual amongst GPCRs because its basal activity is virtually zero, which has been ascribed in part to the ionic lock between R3.50 and E6.30 with the distance between the N-O of the side chains being 2.8–3.2 Å. In contrast, the basal activity of

β1 AR is much higher than rhodopsin and there is a significantly longer distance between the equivalent side chains of 3.7–3.8 Å. The difference in the length of the ionic interaction will undoubtedly affect its strength and hence the conformational flexibility of the region, which could contribute to the higher basal activity of β1 AR. However, clearly this is not the only structural feature in GPCRs that affects basal activity. For example, the naturally occurring polymorphism in human β2 AR, T164I4.56 , is at the H4–H5 interface and significantly reduces both basal activity and the agonist-induced activation of the receptor (31). In addition, the recent structure of the dopamine (D3) receptor (D3R) shows that it contains an intact ionic lock, with the N-O distance of 2.5 Å (13). There is also no perceptible kink in the cytoplasmic end of H6 of D3R equivalent to the kink observed in the β1 AR structures presented here. Thus, although the region of the ionic lock is conserved, there are clearly subtle variations in the structure between different receptors. The ability of the HKAL sequence to have two structurally different conformations constitutes a ligand-independent mechanism that could affect the basal activity of β1 AR and other GPCRs, depending upon where the equilibrium between the different conformations of the cytoplasmic extension of H6 lies. If the equilibrium in one GPCR favors the bent conformation with an intact ionic lock (Rinv ), as is found in rhodopsin, then the basal activity of the receptor is likely to be low; a strong salt bridge between R3.50 −D∕E6.30 will produce this situation. If the salt bridge is weak or absent, the equilibrium will favor formation of the straight conformation where the ionic lock is open. Once this state is attained, there are no strong interactions between the cytoplasmic ends of H3 and H6 that will hinder the movement of H6 to adopt the conformation (R*) capable of coupling to G proteins; i.e., the probability of R* formation increases once the ionic lock is open. Thus the dynamics of the HKAL region can potentially have an effect on the level of basal activity of a receptor and in this regard it is noteworthy that many constitutively activating mutations are found here, particularly in position 6.34 (Fig. 4). Thus the HKAL region has probably evolved to maintain a level of basal activity that is compatible with the physiological function of the receptor, with some receptors having higher levels of basal activity than others. However, there are many sites that can affect the basal activity of receptors, and the importance of the HKAL sequence might not be general for all rhodopsin-like GPCRs. For example, the M5 muscarinic receptor, when mutated in the same 6.34 site, was devoid of constitutive activity although this site was still found to be the determinant of M5 receptor affinity to the G protein (32). In contrast, all 19 amino acids substitutions in this site in the α1B adrenergic

Fig. 4. Sites of constitutively activating mutations (CAMs) in the CL3 region for different GPCRs. The secondary structure of the bent conformation with the ionic lock present is shown (Left) along with the CL3 region including the ends of TM5 and TM6 enlarged (Middle). The corresponding amino sequence of β36–m23 is shown for H6, with residues corresponding to CAMs in the β2 adrenergic receptor shown in yellow and the thermostabilizing mutation A282L shown in red. GPCRs with CAMs at the 6.34 site are shown on the right with reference to their evolutionary relationships. Amino acid residues corresponding to Leu289 (6.34) in the turkey β1 AR are shown in parentheses. Moukhametzianov et al.

PNAS ∣

May 17, 2011 ∣

vol. 108 ∣

no. 20 ∣

8231

BIOCHEMISTRY

(the N-O distance) is consistently between 3.7–3.9 Å in the bent conformation, which is significantly longer than seen in dark-state rhodopsin (2.8–3.2 Å; Table S4). An additional potential hydrogen bond is formed between His286 and Gln237 in H5. All these polar interactions are absent in the straight conformation of H6, with a minimum N-O distance for the ionic lock of 6.2 Å. In the straight conformation of H6, there are no hydrogen bonds apparent between the cytoplasmic extension of H6 and the neighboring H5, although there are van der Waals contacts between Leu282 in H6 and Ile241 in H5.

receptor have shown different degrees of constitutive activity (33). In the rat μ-opioid receptor, different substitutions of Thr6.34 had different effects; the lysine mutant was constitutively active, whereas the aspartate mutant had lower basal activity and decreased ability to activate G proteins (34). Several human diseases have also been caused by mutations in this region. Mutation of this site in the luteinizing hormone receptor gene causes male gonadotropin-independent precocious puberty (35) and the analogous mutation in the thyrotropin receptor causes hyperfunctioning thyroid adenomas (36). Increased activity of these receptors could be due to the perturbing effect of the mutations on the structure and stabilization of the inactive state. Methods Purification and Crystallization. For the preparation of β1 AR36–M23 with either cyanopindolol or iodocyanopindolol bound (crystals t148, t468, and t756), baculovirus-expressed receptor was solubilized in β-decylmaltoside and purified by Ni2þ -affinity chromatography and alprenolol sepharose ligand affinity chromatography, with detergent exchange to octylthioglucoside (0.35%) on the ligand affinity column (28, 37). Crystallization was by vapor diffusion with addition of receptor (5.5 mg∕mL) to an equal volume (0.5–1.0 μL) 0.1 M N-(2-acetamido)iminodiacetate-NaOH, pH 6.9–7.3, 29–33% PEG 600. For crystallization of β1 AR36–M23 with carazolol bound (crystal t1118), detergent was exchanged to Hega-10 (0.35%) and receptor was eluted with 50 μM carazolol. Prior to crystallization, ligand and detergent concentrations were increased to 0.5 mM and 0.5%, respectively. Crystallization was by vapor diffusion with addition of 200 nL receptor (9 mg∕mL) to an equal volume of precipitant solution (0.1 M Tris • HCl, pH 8.1, 40% PEG 400). Crystals were cryoprotected by a very short soak in either 60% PEG600 (t148), 40% PEG400 1. Fredriksson R, Schioth HB (2005) The repertoire of G-protein-coupled receptors in fully sequenced genomes. Mol Pharmacol 67:1414–1425. 2. Foord SM, et al. (2005) International Union of Pharmacology. XLVI. G protein-coupled receptor list. Pharmacol Rev 57:279–288. 3. Li J, Edwards PC, Burghammer M, Villa C, Schertler GF (2004) Structure of bovine rhodopsin in a trigonal crystal form. J Mol Biol 343:1409–1438. 4. Okada T, et al. (2004) The retinal conformation and its environment in rhodopsin in light of a new 2.2 Å crystal structure. J Mol Biol 342:571–583. 5. Palczewski K, et al. (2000) Crystal structure of rhodopsin: A G protein-coupled receptor. Science 289:739–745. 6. Park JH, Scheerer P, Hofmann KP, Choe HW, Ernst OP (2008) Crystal structure of the ligand-free G-protein-coupled receptor opsin. Nature 454:183–187. 7. Scheerer P, et al. (2008) Crystal structure of opsin in its G-protein-interacting conformation. Nature 455:497–502. 8. Cherezov V, et al. (2007) High-resolution crystal structure of an engineered Human β2 -adrenergic G protein coupled receptor. Science 318:1258–1265. 9. Rasmussen SG, et al. (2007) Crystal structure of the human β2 adrenergic G-protein-coupled receptor. Nature 450:383–387. 10. Hanson MA, et al. (2008) A specific cholesterol binding site is established by the 2.8 A structure of the human β2 -adrenergic receptor. Structure 16:897–905. 11. Warne T, et al. (2008) Structure of a β1 -adrenergic G-protein-coupled receptor. Nature 454:486–491. 12. Jaakola VP, et al. (2008) The 2.6 angstrom crystal structure of a human A2A adenosine receptor bound to an antagonist. Science 322:1211–1217. 13. Chien EY, et al. (2010) Structure of the human dopamine D3 receptor in complex with a D2/D3 selective antagonist. Science 330:1091–1095. 14. Wu B, et al. (2010) Structures of the CXCR4 chemokine GPCR with small-molecule and cyclic peptide antagonists. Science 330:1066–1071. 15. Kobilka B, Schertler GF (2008) New G-protein-coupled receptor crystal structures: insights and limitations. Trends Pharmacol Sci 29:79–83. 16. Schwartz TW, Hubbell WL (2008) Structural biology: A moving story of receptors. Nature 455:473–474. 17. Tate CG, Schertler GF (2009) Engineering G protein-coupled receptors to facilitate their structure determination. Curr Opin Struct Biol 19:386–395. 18. Deupi X, Kobilka B (2007) Activation of G protein-coupled receptors. Adv Protein Chem 74:137–166. 19. Kobilka BK, Deupi X (2007) Conformational complexity of G-protein-coupled receptors. Trends Pharmacol Sci 28:397–406. 20. Bond RA, Ijzerman AP (2006) Recent developments in constitutive receptor activity and inverse agonism, and their potential for GPCR drug discovery. Trends Pharmacol Sci 27:92–96. 21. Costa T, Herz A (1989) Antagonists with negative intrinsic activity at delta opioid receptors coupled to GTP-binding proteins. Proc Natl Acad Sci USA 86:7321–7325. 22. Hofmann KP, et al. (2009) A G protein-coupled receptor at work: The rhodopsin model. Trends Biochem Sci 34:540–552. 23. Altenbach C, Kusnetzow AK, Ernst OP, Hofmann KP, Hubbell WL (2008) High-resolution distance mapping in rhodopsin reveals the pattern of helix movement due to activation. Proc Natl Acad Sci USA 105:7439–7444.

8232 ∣

www.pnas.org/cgi/doi/10.1073/pnas.1100185108

(t1118), or 35% PEG600 and 5% glycerol (t468). No additional cryoprotectant was added to crystal t756 (iodocyanopindolol). Crystallography. Diffraction data for crystals t148 and t756 were collected at 100 K with a MarCCD detector on beamline ID23-2 at the European Synchrotron Radiation Facility (λ ¼ 0.873 Å). Data for crystals t468 and t1118 were collected at 100 K with a Pilatus detector on beamline X06SA at the Swiss Light Source (λ ¼ 0.98 and 1.00 Å, respectively). Each dataset was collected from multiple positions on a single crystal, following extensive screening. Diffraction data were processed with MOSFLM and SCALA (38). Because the diffraction was anisotropic, different wedges of data were processed to different resolutions and consequently the completeness of the highest resolution shell is low for crystals t148, t468, and t1118 (Table S1). All structures were obtained with molecular replacement using PHASER (39) starting from the 2VT4 model (11). Refinement was carried out in PHENIX (40) and REFMAC (41) with application of strict noncrystallographic symmetry averaging between appropriate regions of the two independent monomers in the asymmetric unit. Structure validation was performed using the MolProbity Web server. ACKNOWLEDGMENTS. We thank F. Gorrec for his help with crystallization robotics and F. Marshall, M. Weir, and R. Henderson for helpful comments on the manuscript. This work was supported by core funding from the Medical Research Council (MRC) and a grant jointly funded by Pfizer Global Research and Development and MRC Technology. Additional financial support was from a Human Frontier Science Project Program Grant (RG/0052), a European Commission FP6 specific targeted research project (LSH-20031.1.0-1), a European Synchrotron Radiation Facility (ESRF) long-term proposal and a Biotechnology and Biological Sciences Research Council Grant (BB/ G003653/1). We would like to thank beamline staff at the ESRF beamline ID23-2 and at the Swiss Light Source beamline X06SA. 24. Ballesteros JA, Weinstein H (1995) 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. 25. Ballesteros JA, et al. (2001) Activation of the β2 -adrenergic receptor involves disruption of an ionic lock between the cytoplasmic ends of transmembrane segments 3 and 6. J Biol Chem 276:29171–29177. 26. Costa T, Cotecchia S (2005) Historical review: Negative efficacy and the constitutive activity of G-protein-coupled receptors. Trends Pharmacol Sci 26:618–624. 27. Wacker D, et al. (2010) Conserved binding mode of human β2 adrenergic receptor inverse agonists and antagonist revealed by X-ray crystallography. J Am Chem Soc 132:11443–11445. 28. Warne T, Serrano-Vega MJ, Tate CG, Schertler GF (2009) Development and crystallization of a minimal thermostabilised G protein-coupled receptor. Protein Expr Purif 65:204–213. 29. Serrano-Vega MJ, Magnani F, Shibata Y, Tate CG (2008) Conformational thermostabilization of the β1 -adrenergic receptor in a detergent-resistant form. Proc Natl Acad Sci USA 105:877–882. 30. Dror RO, et al. (2009) Identification of two distinct inactive conformations of the β2 -adrenergic receptor reconciles structural and biochemical observations. Proc Natl Acad Sci USA 106:4689–4694. 31. Green SA, Rathz DA, Schuster AJ, Liggett SB (2001) The Ile164 β2 -adrenoceptor polymorphism alters salmeterol exosite binding and conventional agonist coupling to G(s). Eur J Pharmacol 421:141–147. 32. Spalding TA, Burstein ES (2006) Constitutive activity of muscarinic acetylcholine receptors. J Recept Signal Transduct Res 26:61–85. 33. Kjelsberg MA, Cotecchia S, Ostrowski J, Caron MG, Lefkowitz RJ (1992) Constitutive activation of the alpha 1B-adrenergic receptor by all amino acid substitutions at a single site. Evidence for a region which constrains receptor activation. J Biol Chem 267:1430–1433. 34. Huang P, et al. (2001) Functional role of a conserved motif in TM6 of the rat mu opioid receptor: Constitutively active and inactive receptors result from substitutions of Thr6.34(279) with Lys and Asp. Biochemistry 40:13501–13509. 35. Schulz A, Schoneberg T, Paschke R, Schultz G, Gudermann T (1999) Role of the third intracellular loop for the activation of gonadotropin receptors. Mol Endocrinol 13:181–190. 36. Parma J, et al. (1993) Somatic mutations in the thyrotropin receptor gene cause hyperfunctioning thyroid adenomas. Nature 365:649–651. 37. Warne T, Chirnside J, Schertler GF (2003) Expression and purification of truncated, non-glycosylated turkey β-adrenergic receptors for crystallization. Biochim Biophys Acta 1610:133–140. 38. Evans P (2006) Scaling and assessment of data quality. Acta Crystallogr D Biol Crystallogr 62:72–82. 39. McCoy AJ, et al. (2007) Phaser crystallographic software. J Appl Crystallogr 40:658–674. 40. Adams PD, et al. (2002) PHENIX: Building new software for automated crystallographic structure determination. Acta Crystallogr D Biol Crystallogr 58:1948–1954. 41. Murshudov GN, Vagin AA, Dodson EJ (1997) Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr D Biol Crystallogr 53:240–255.

Moukhametzianov et al.