Molecular Recognition Section, Division of Intramural Research : NIDDK : National Institutes of Health
NavigationINDEX

Computational Chemistry

An Introduction to Modelling G Protein-Coupled Receptors

This page describes the current research in our lab to better understand the molecular mechanisms involved in receptor-ligand interactions.

For this purpose we use

This particular page is an excerpt from a publication:

  • van Rhee, A.M.; Fischer, B.; Jacobson, K.A.
    Modelling the P2Y purinoceptor using rhodopsin as template.
    Drug Design Discovery, 1995, 13: 133-154.

which describes how we approached the problem of visualizing the molecular architecture of the G protein-coupled P2Y1 purinoceptor.

It is the intention to add more features (both classical medicinal chemistry and computational chemistry) as we go along, but time is the limiting factor. You could make a bookmark now, and check in every now and then.

A very characteristic feature of G protein-coupled receptors (GPCRs) is their overall topology:

They all have similar hydrophobicity profiles with seven stretches of increased hydrophobicity, most probably corresponding to alpha-helical regions spanning the cell membrane.

Unfortunately, isolation, purification and crystallization of P2 purinoceptors, as for other GPCRs, remains to be demonstrated.

Not only is a crystallographic structure as yet unobtainable, but the very limited quantities of receptor protein present in the membrane, and isolation problems, prohibit the use of modern techniques, such as nuclear magnetic resonance spectroscopy, towards structure elucidation.

Chemical modification 26 of the ligand binding site and site-directed mutagenesis have been shown to be useful techniques in understanding receptor-ligand interactions, but is particularly in the case of P2 purinoceptors restricted by the lack of suitable radioligands.

A possible further approach is to construct models of the receptors and their ligand complexes using molecular modelling, as was first proposed by Hibert et al.

On these pages we present models of the chick P2Y purinoceptor, and the human A2A receptor based on the electron density map of rhodopsin 28 , rather than on the bacteriorhodopsin template 29 used by Hibert et al.

The models are verified by incorporation of site-directed mutagenesis 41, 50, 51 and ligand affinity data. 14, 15


Computational Methods.
Software and Hardware.
Calculations and manipulations were performed using the

Model building.
Kyte-Doolittle hydrophobicity and Emini surface probability parameters were calculated using a 7 amino acid window. With these parameters, the putative transmembrane domains (TMs) were identified.

Sequences of the putative transmembrane domains were aligned manually using the alignment procedure described on our "GPCR page".

Helices were built from the full length of the TM, using the Biopolymer module in the Insight II program. The N terminus was capped with an acetamido group, and the C terminus was capped with a carboxamido group. The secondary structure was assumed to be a right-handed alpha helix (default phi and psi angles from the Biopolymer module). After generation of the Amber atom type parameters and template charges, the helices were energy minimized as described in the Energy Minimization section.

The energy minimized helices were converted to Protein Data Bank (PDB) format files and imported in the Quanta program. Coils (an alpha carbon backbone representation of the helix) of the minimized helices were drawn, that resemble the helical wheel representation to aid in orienting the individual helices towards each other.

The following constraints were used constructing the helical bundle:

  • 1. The helical axes of the first, third, fifth and seventh helices were quasi-antiparallel to those of the second, fourth and sixth helices.
  • 2. The hydrophobic side of each helix was facing the lipid phase and the hydrophilic side of each helix was facing either another helix or the pore formed by the helical bundle.
  • 3. Conserved residues, either identical throughout a subset or highly homologous, determined the orientation of each helix relative to the other helices.
  • 4. Distances in the electron density map of rhodopsin 28 correlated to atomic coordinates in the model. This was obtained through triangulation of the electron density map and conversion of the coordinates to a grid on the screen.
  • 5. The assembly of helices maintained a clockwise order, when seen from the intracellular side, as argued by Baldwin. 31
  • 6. None of the helices were intersecting.

Energy Minimization.
The individual helices generated with Insight II's Biopolymer module were energy minimized in Discover using a stepwise process. Initially, 200 steps of Steepest Descent were performed, followed by minimization using the Conjugate Gradient (CG) method until the gradient reached a value below 0.1 kcal/mol/Å. Use of the Amber forcefield in Discover required that in all calculations the 1-4 nonbonded interactions were scaled by a factor 0.5, and the dielectric constant was assumed to be distance independent with a magnitude of 4. The bond, theta, phi, out-of-plane, nonbonded and Coulombic interactions were all used to obtain the final energy, but were not expressly scaled.

The helical bundles generated with Quanta were energy minimized in Discover in a stepwise process. Initially, 500 steps CG were performed with the backbone of the helices tethered with a force constant of 100 kcal/Å. In consecutive runs (500 steps CG each), the force constant was reduced to 50 kcal/Å, then 25 kcal/Å and, finally, to 0 (no tethering).

Ligand Docking.
Selective and/or specific ligands for the receptors were fully minimized using the MNDO hamiltonian of MOPAC. All ligands were rigidly docked into the helical bundle using graphical manipulation coupled to continuous energy monitoring, i.e. the ligand was manually docked into the binding site without relaxing the atomic coordinates of either ligand or protein while continuously calculating the energy of the whole. When a final position was reached, consistent with a low local energy and known pharmacological data, the complex of receptor and ligand was subjected to a minimization run of 4000 steps CG (or until the gradient was < 0.1 kcal/mol/Å). Charges for the ligands were imported from the MOPAC output files.


DISCUSSION.

We will now briefly discuss the results obtained through this approach, but the information specific to either the P2Y purinoceptor or the adenosine A2A receptor is kept on separate pages.

Alignment of G protein-coupled receptors and dendrogram.
In TM1, the motif GXXXN (P2Y) or GXXGN (P2U) occurs in P2 purinoceptors, rather than the GN motif present in biogenic amine or adenosine receptors. The alternate motifs, however, are not exclusively used by P2 receptors, but are shared with, e.g., the PGE3-II (GXXGN) and the PAF (GXXXN) receptors. In all sequences the last 5 C-terminal residues of this helix are frequently occupied by lysine or arginine residues, indicating the end of the transmembrane domain. Such basic residues may serve as 'membrane anchors', 34, 42 and are useful in determining the position of the helix in the lipid bilayer.

In TM2, the residue preceding the conserved leucine in the LXXXD motif is a conserved serine for the biogenic amine receptors, but is either an asparagine or a histidine in the P2Y and P2U receptors, respectively. The asparagine is conserved among the P2Y, AT1A, CKR1, IL-8A, PAF, NK2 receptors, the opsins and rhodopsin. The histidine is shared between P2U and thrombin receptors.

Another difference between biogenic amine receptors and P2 receptors is the position of the conserved proline relative to the conserved aspartate.

Although absent in the muscarinic receptor, the proline is consistently spaced by 8 residues from the aspartate in the other biogenic amine receptors, whereas in many other receptors, the P2 receptors included, only 7 residues separate these two residues.

This last difference is possibly significant since the proline is located near the luminal side of the receptor protein, which is thought to be important for ligand binding.

The only exceptions, observed so far, to the conservation of the aspartate occur in the substance P receptor where it is substituted with a glutamate, and in the gonadotropin-releasing hormone receptor where an asparagine replaces this residue.

The conserved aspartate in TM2 was shown by Horstman et al to be a sodium-dependent allosteric regulatory site in the alpha2-adrenergic receptor.

The DRY motif, characteristic of the third transmembrane domain in biogenic amine receptors, is replaced by a HRY motif in the P2Y1 receptor (HRC in the P2U receptor). This motif has not been found in other receptors. The only other substitution for the conserved aspartate, seems to be the D to E in, e.g., the TXA2, PGE3-II, opsin and rhodopsin sequences. The significance of this deviation is not clear, but is supposedly important for coupling of the P2 receptors to G proteins and not for ligand binding, the subject of this study.

One or more of the first 6 positions in the N-terminal sequence of TM4 are generally (37 out of 40 aligned sequences of family A GPCRs ) occupied by lysine or arginine residues. Again these residues could well serve as 'membrane anchors'. They seem to be occurring more frequently towards the cytosolic side of helices than on the luminal side, thus reflecting the polarity of the membrane. These 'membrane anchors' also occur much less frequently in the C-terminal sequences of helices, than in the N-terminal sequences, as is demonstrated by the alignment for TM5. The reason for this disparity is believed to be the direction in which the side-chains are pointing, i.e. opposite to the propagation direction of the helix, allowing more rotational freedom for lysine and arginine residues near the cytosolic N-terminus than for the cytosolic C-terminus.
There are no other marked differences between the various receptor subfamilies for either TM4 or TM5.

The CXXP motif used for the alignment of TM6 could be substituted by WXP for most receptors, but the P2 receptors deviate at this position. The P2Y receptor has a tyrosine at the position of the conserved tryptophan, and this seems to be rather unique. The phenylalanine in the P2U receptor at the same position occurs more frequently, such as in the thrombin, and various orphan receptors. Characteristic of the P2 receptors is the presence of a lysine (P2Y: K269) or an arginine (P2U: R265) at an otherwise non-conserved position. It shares this feature only with the endothelin receptors (both ET-A and ET-B subtypes) and the orphan receptor RSC338. The proposed 'membrane anchors' occur quite frequently at the first position in the alignment of this transmembrane domain.

TM7 is best aligned by means of the NPXXY motif. However, the P2 receptors, the gonadotropin-releasing hormone receptor, the TXA2, PGE3-II, and several orphan receptors constitute an exception to this rule. The conserved asparagine is replaced by an aspartate residue in the latter cases. A non-conserved aspartate (D352) in the 5HT1B receptor aligns perfectly with aspartates in the IL-8A receptor (D288), and in the D1B receptor (D342), a glutamine in the P2Y receptor (Q296) and a lysine in the P2U receptor (K289). Arginines 299 (P2Y) and 292 (P2U) align with Y530 in the m3, E291 in the IL-8A and E287 in the CKR1 receptor, respectively. Both positions are near the luminal side of the receptor and are likely involved in ligand binding. The K and R residues occurring near the C-terminus are probably better characterized as 'membrane anchors'.

If you are interested in the alignment of over 150 GPCRs, or in mutagenesis studies of GPCRs, just follow these links. You might want to read our introductory page first, though.

Bacteriorhodopsin, included in this study to facilitate comparison with other modelling studies, [e.g., 27, 35] is clearly not related to any of the GPCR subfamilies. The degree of relatedness between bacteriorhodopsin and GPCRs shown in the dendrogram is probably an overestimate of the actual distance, caused by the residue alignment procedure of the GCG program.

Kyte-Doolittle hydrophobicity and Emini surface probability analysis.
At the portion of the sequence where proteins are supposed to cross the lipid bilayer membrane, the hydrophobicity of that segment is usually increased relative to the cytosolic and luminal portions. The procedure developed by Kyte and Doolittle uses this phenomenon to identify transmembrane domains in protein sequences with unknown structure. In the case of GPCRs, these transmembrane domains are thought to be alpha helical, and more importantly amphiphilic. The amphiphilicity of the TMs supposedly reflects the way a GPCR is built out of seven of these helical transmembrane domains. Consequentially, the hydrophobicity profile of these GPCRs is not as clear as one would wish.

An example of this effect is the poor separation of the sixth (VI) and seventh (VII) TM in both the P2Y1 receptor profile (below) and the bacteriorhodopsin profile (not shown).


To facilitate identification of TMs, the Kyte-Doolittle hydrophobicity method is often supplemented with methods describing other sequence dependent vectorial parameters, such as the 'conservation moment' and 'hydrophilic moment', 36 for each helix or other computational methods 43 . Use of the 'conservation moment' method requires a particularly well defined alignment of highly homologous sequences, and can not be applied to the P2 GPCRs because of the low sequence similarity of these receptors with any other GPCR subfamily. The 'hydrophilic moment' method relies on a database of partial hydrophilic factors for any given amino acid and is dependent on the method with which these factors were measured. Furthermore, the high incidence of basic residues in the P2Y receptor sequence greatly influences the results obtained with this method.

In contrast, the Emini surface probability can be calculated from a given sequence, indicating the propensity of a stretch of amino acids (the window) to be at the surface of a protein. The combination of an increase in the Kyte-Doolittle hydrophobicity index and a decrease in the Emini surface probability index was successfully applied to predict the TMs in the P2Y1 receptor (see figure above) and in the reference protein bacteriorhodopsin. The start and end residue of each helix was usually predicted within 3 residues of the TMs used by Henderson et al. The only major deviation in the P2Y1 receptor prediction occurred at the N-terminus of TM7 (VII). This particular sequence contains 3 hydrophilic residues (K, Q and R) that are possibly involved in ligand binding, thus delaying the start of the predicted TM.

Building the receptor model.
Modelling of G protein-coupled receptors has become an important tool in understanding drug-receptor interactions and in the development of new ligands for these receptors.

The first widely accepted method was the homology modeling method by Hibert et al. This method involved the alignment of the receptor sequence with the sequence of bacteriorhodopsin, and the subsequent mapping of the sequence onto the structure of bacteriorhodopsin that was determined by Henderson et al.

Bacteriorhodopsin is a proton pump in the outer membrane of Halobacterium halobium, and lacks any functional or sequence homology with GPCRs. Nevertheless, the procedure was based on the assumption that there would be considerable structural homology. This structural homology was inferred by the extraordinary similarity in the hydrophobicity plots, or Kyte-Doolittle plots, of the biogenic amine subfamily of GPCRs and bacteriorhodopsin.

Recently, a low resolution electron density map of rhodopsin, a true member of the GPCR superfamily, was published. 28 The low sequence homology with bacteriorhodopsin, the structural differences that must arise from the different placement of proline residues (causing bends in helices) in bacteriorhodopsin- and GPCR-sequences, and the availability of an electron density map of a true member of the GPCR superfamily prompted us to adapt a new method to build models of GPCRs. 34 This novel method is based on a computational approach rather than strict compliance with the atomic coordinates of a distantly related protein, albeit with higher resolution.

To ascertain the viability of modelling transmembrane proteins (based on the structure of rhodopsin 34 ) by the methods described in the paper, we built models of bacteriorhodopsin (data not shown) based on the electron density map of bacteriorhodopsin, 29 and rhodopsin (data not shown) based on the electron density map of bovine rhodopsin as published by Schertler et al.

Root mean square (r.m.s.) distance calculations on superimposed structures were performed to establish how well the models fitted the experimental data. The bacteriorhodopsin model that was constructed, compared with the one deposited in the PDB, 29 had an r.m.s. value of 2.15 Å when measured on all backbone atoms and 2.06 Å on all Calpha atoms. Both values are lower than the resolution of the model, i.e. 3.5 Å. When the rhodopsin model was compared with the bacteriorhodopsin model by Henderson et al, however, the r.m.s. value increased to 16.96 Å when measured on all backbone atoms and 16.86 Å on all Calpha atoms. These rather high values indicate that there are more structural differences between bacteriorhodopsin and rhodopsin than have been assumed thus far. 44

In our opinion, the position of the proline residues in the helices is a major determinant in this matter. As can be seen from the alignment for the helices, proline residues occur less frequently and at different positions in the sequence of bacteriorhodopsin than in the sequence of rhodopsin. The influence of proline residues in alpha-helices was extensively studied by Sankararamakrishnan and Vishveshwara and von Heijne.

We have found that applying the Amber forcefield to our calculations yields results in agreement with Sankararamakrishnan and Vishveshwara and von Heijne, and that those results are consistent with data obtained from a crystallographic study of the transmembrane protein photosynthetic reaction center (PDB:1prc), and globular proteins such as phosphoglycerate kinase (PDB:3pgk), lysozyme (PDB:1l27) and alcohol dehydrogenase (PDB:5adh). The differences between the structures derived from the electron density maps of bacteriorhodopsin and rhodopsin also illustrate why helical wheel models, widely used by molecular biologists, are highly imprecise when applied to GPCRs. 31

The general picture looks like this:


For a detailed view of either our


REFERENCES.

  • 1. Hoyle, C.H.V. and Burnstock, G. (1991) ATP receptors and their physiological roles. In Adenosine in the nervous system. Stone, T.W., Ed., pp. 43-76. London: Academic Press Ltd.
  • 2. Hoyle, C.H.V. (1992) Transmission: purines. In Autonomic neuroeffector mechanisms. Burnstock, G. and Hoyle, C.H.V., Eds., pp. 367-407. Chur: Harwood Academic Press.
  • 3. Edwards, F.A., Gibb, A.J., Colquhoun, D. (1992) ATP receptor mediated synaptic currents in the central nervous system. Nature , 359, 144-147.
  • 4. Burnstock, G. (1978) A basis for distinguishing two types of purinergic receptor. In Cell Membrane Receptors for Drugs and Hormones: a Multidisciplinary Approach. Straub, P.W. and Bolis, L., Eds., pp. 106-118. New York: Raven Press.
  • 5. Cusack, N.J. (1993) P2 receptors: subclassification and structure-activity relationships. Drug Dev. Res. 28, 244-252.
  • 6. Gordon, J.L. (1986) Extracellular ATP: effects, sources and fate. J. Biochem. 233, 309- 319.
  • 7. Fredholm, B.B., Abbracchio, M.P., Burnstock, G., Daly, J.W., Harden, T.K., Jacobson, K.A., Leff, P., Williams, M. (1994) Nomenclature and classification of purinoceptors. Pharmacol. Rev. 46, 143-156.
  • 8. Abbracchio, M.P., Burnstock, G. (1994) Purinoceptors: are there families of P2X and P2Y purinoceptors? Pharmac. Ther. 64, 445-475.
  • 9. Dubyak, G.R. (1991) Signal transduction by P2-purinergic receptors for extracellular ATP. Am. J. Respir. Cell. Mol. Biol. 4, 295-300.
  • 10. Harden, T.K., Hawkins, P.T., Stephens, L., Boyer, J.L., Downes, P. (1988) Phosphoinositide hydrolysis by guanosine 5'-([g-thio]triphosphate)-activated phospholipase C of turkey erythrocyte membranes. Biochem. J. 252, 583-593.
  • 11. Pirotton, S., Boeynaems, J.M. (1991) Transduction mechanism of P2 purinergic receptors: Role of phospholipase C and calcium. Nucleos. Nucleot. 10, 1009-1017.
  • 12. Benham, C.D., Tsien, R.W. (1987) A novel receptor-operated Ca2+-permeable channel activated by ATP in smooth muscle. Nature 328, 275-278.
  • 13. Bean, B.P. (1992) Pharmacology and electrophysiology of ATP activated ion channels. Trends Pharmacol. Sci. 13, 87-90.
  • 14. Fischer, B., Boyer, J.L., Hoyle, C.H.V., Ziganshin, A.U., Brizzolara, A.L., Knight, G.E., Zimmet, J., Burnstock, G., Harden, T.K., Jacobson, K.A. (1994) Identification of potent, selective P2Y-purinoceptor agonist structure-activity relationship for 2-thioether derivatives of adenosine 5'-triphosphate. J. Med. Chem. 36, 3937-3946.
  • 15. Burnstock, G., Fischer, B., Hoyle, C.H.V., Maillard, M., Ziganshin, A.U., Brizzolara, A.L., von Isakovics, A., Boyer, J.L., Harden, T.K., Jacobson, K.A. (1994) Structure activity relationship for derivatives of adenosine 5'-triphosphate as agonists at P2 purinoceptors: heterogeneity within P2X and P2Y subtypes. Drug Dev. Res. 31, 206-219.
  • 16. Windscheif, U., Ralevic, V., Bäumert, H.G., Mutschler, E., Lambrecht, G., Burnstock, G. (1994) Vasoconstrictor and vasodilator responses to various agonists in the rat perfused mesenteric arterial bed - Selective-inhibition by PPADS of contractions mediated via P2X- purinoceptors. Br. J. Pharmacol. 113, 1015-1021.
  • 17. Humphries, R.G., Tomlinson, W., Ingall, A.H., Cage, P.A., Leff, P. (1994) FPL-66096 - A novel, highly potent and selective antagonist at human platelet P2T- purinoceptors. Br. J. Pharmacol. 113, 1057-1063.
  • 18. van Rhee, A.M., van der Heijden, M.P.A., Beukers, M.W., IJzerman, A.P., Soudijn, W., Nickel, P. (1994) Novel competitive antagonists for P2 purinoceptors. Eur. J. Pharmacol. Mol. Pharmacol. Sect. 268, 1-7.
  • 19. Kyte, J., Doolittle, R.F. (1982) A simple method for displaying the hydrophobic character of a protein. J. Mol. Biol. 157, 105-132.
  • 20. Kubo, T., Fukuda, K., Mikama, A., Maeda, A., Takahashi, H., Mishina, M., Haga, Y., Ichiyama, A., Kangawa, K., Kajima, M., Matsuo, H. Hirose, T., Numa, T. (1986) Cloning, sequencing and expressing of complementary DNA coding the muscarinic acetylcholine receptor. Nature 323, 411-416.
  • 21. Webb, T.E., Simon, J., Krishek, B.J., Bateson, A.N., Smart, T.G., King, B.F., Burnstock, G., Barnard, E.A. (1993) Cloning and functional expression of a brain G- protein-coupled ATP receptor. FEBS Lett. 324, 219-225.
  • 22. Filtz, T.M., Li, Q., Boyer, J.L., Nicholas, R.A., Harden, T.K. (1994) Expression of a P2Y purinergic receptor that couples to phospholipase C. Mol. Pharmacol. 40, 8-14.
  • 23. Lustig, K.D., Shiau, A.K., Brake, A.J., Julius, D. (1993) Expression cloning of an ATP receptor from mouse neuroblastoma cells. Proc. Natl. Acad. Sci. USA 90, 5113-5117.
  • 24. Parr, C.E., Sullivan, D.M., Paradiso, A.M., Lazarowski, E.R., Burch, L.H., Olsen, J.C., Erb, L., Weisman, G.A., Boucher, R.C., Turner, J.T. (1994) Cloning and expression of a human P2U nucleotide receptor, a target for cystic fibrosis pharmacotherapy. Proc. Natl. Acad. Sci. USA 91, 3275-3279.
  • 25. Rice, W.R., Burton, F.M., Fiedeldey, D.T. (1995) Cloning and expression of the alveolar type II cell P2U-purinergic receptor. Am. J. Respir. Cell Mol. Biol. 12, 27-32.
  • 26. Jacobson, K.A., Stiles, G.L., Ji, X.D. (1992) Chemical modification and irreversible inhibition of striatal A2a adenosine receptors. Mol. Pharmacol. 42, 123-133.
  • 27. Hibert, M.F., Trumpp-Kallmeyer, S., Bruinvels, A., Hoflack, J. (1991) Three- dimensional models of neurotransmitter G-binding protein-coupled receptors. Mol. Pharmacol. 40, 8-15.
  • 28. Schertler, G.F., Villa, C., Henderson, R. (1993) Projection structure of rhodopsin. Nature 362, 770-772.
  • 29. Henderson, R., Baldwin, J.M., Ceska, T.A., Zemlin, F., Beckmann, E., Downing, K.H. (1990) Model for the structure of bacteriorhodopsin based on high-resolution electron cry- microscopy. J. Mol. Biol. 213, 899-929.
  • 30. Oliveira, L., Paiva, A.C.M., Vriend, G. (1993) A common motif in G-protein-coupled seven transmembrane helix receptors. J. Comp.-Aided Mol. Design 7, 649-658.
  • 31. Baldwin, J.M. (1993) The probable arrangement of the helices in G protein-coupled receptors., EMBO J. 12, 1693-1703.
  • 32. Kovács, T. and Ötvös, L. (1988) Simple synthesis of 5-vinyl and 5-ethynyl-2'- deoxyuridine-5'-triphosphates. Tetrahedron Lett. 29, 4525-4528.
  • 33. Moffat, J.G. (1964) A general synthesis of nucleoside-5'-triphosphates. Can. J. Chem. 42, 599.
  • 34. Ballesteros, J.A., 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.
  • 35. IJzerman, A.P., van Galen, P.J.M., Jacobson, K.A. (1992) Molecular modeling of adenosine receptors. I. The ligand binding site on the A1 receptor. Drug Des. Discov. 9, 49-67.
  • 36. Zhang, D., Weinstein, H. (1994) Polarity conserved positions in transmembrane domains of G-protein coupled receptors and bacteriorhodopsin. FEBS Lett. 337, 207-212.
  • 37. Sankararamakrishnan, R., Vishveshwara, S. (1992) Geometry of proline-containing alpha- helices in proteins. Int. J. Peptide Protein Res. 39, 356-363.
  • 38. Noel, J.P., Hamm, H.E., Sigler, P.B. (1993) The 2.2 Å crystal structure of transducin-a complexed with GTPgS. Nature 366, 654.
  • 39. Brake, A. J., Wagenbach, M. J., Julius, D. (1994) New structural motif for ligand-gated ion channels defined by an ionotropic ATP receptor. Nature 371, 519-523.
  • 40. Valera, S., Hussy, N., Evans, R. J., Adami, N., North, R. A., Surprenant, A., Buell, G. (1994) New class of ligand-gated ion-channel defined by P2X receptor for extracellular ATP. Nature 371, 516-519.
  • 41. Erb, L., Garrad, R., Wang, Y., Quinn, T., Turner, J.T., Weisman, G.A. (1995) Site- directed mutagenesis of P2U purinoceptors. J. Biol. Chem. 270, 4185-4188.
  • 42. von Heijne, G. and Manoil, C. (1990) Membrane proteins: from sequence to structure. Protein Eng. 4, 109-112.
  • 43. Sander, C. and Schneider, R. (1991) Database of homology-derived protein structures. Proteins 9, 56-68.
  • 44. Hoflack, J., Trumpp-Kallmeyer, S., Hibert, M. (1994) Re-evaluation of bacteriorhodopsin as a model for G protein-coupled receptors. Trends Pharmacol. Sci. 43, 348-350.
  • 45. von Heijne, G. (1991) Proline kinks in transmembrane alpha-helices. J. Mol. Biol. 218, 499- 503.
  • 46. Wess, J., Gdula, D., Brann, M.R. (1991) Site-directed mutagenesis of the m3 muscarinic receptor: identification of a series of threonine and tyrosine residues involved in agonist but not antagonist binding. EMBO J. 10, 3729-3734.
  • 47. Suryanarayana, S., Daunt, D.A., von Zastrow, M., Kobilka, B.K. (1991) A point mutation in the seventh hydrophobic domain of the a2 adrenergic receptor increases its affinity for a family of beta receptor antagonists. J. Biol. Chem. 266, 15488-15492.
  • 48. Oksenberg, D., Marsters, S.A., O'Dowd, B.F., Jin, H., Havlik, S., Peroutka, S.J., Ashkenazi, A. (1992) A single amino-acid difference confers major pharmacological variation between human and rodent 5-HT1B receptors. Nature 360, 161-163.
  • 49. Fraser, C.M., Wang, C.D., Robinson, D.A., Gocayne, J.D., Venter, J.C. (1989) Site- directed mutagenesis of m1 muscarinic acetylcholine receptors: conserved aspartic acids play important roles in receptor function. Mol. Pharmacol. 36, 840-847.
  • 50. Elling, C.E., Møller Nielsen, S., Schwartz, T.W. (1995) Conversion of antagonist binding site to metal-ion site in the tachykinin NK-1 receptor. Nature 374, 74-77.
  • 51. Kim, J., Wess, J., van Rhee, A.M., Schöneberg, T., Jacobson, K.A. (1995) Site-directed mutagenesis identifies residues involved in ligand recognition in the human A2a adenosine receptor. J. Biol. Chem. 270, 13987-13997.
  • 52. Zhou, W., Flanagan, C., Ballesteros, J.A., Konvicka, K., Davidson, J.S., Weinstein, H., Millar, R.P., Sealfon, S.C. (1994) A reciprocal mutation supports helix 2 and helix 7 proximity in the gonadotropin-releasing hormone receptor. Mol. Pharmacol. 2, 165-170.
  • 53. Horstman, D.A., Brandon, S., Wilson, A.L., Guyer, C.A., Cragoe, E.J., Limbird, L.E. (1990) An aspartate conserved among G-protein receptors confers allosteric regulation of a2-adrenergic receptors by sodium. J. Biol. Chem. 265, 21590-21595.
  • 54. IJzerman, A. P., van der Wenden, E. M., van Galen, P.J.M., Jacobson, K. A. (1994) Molecular modeling of adenosine receptors - the ligand-binding site on the rat adenosine A2a receptor. Eur. J. Pharmacol. Mol. Pharmacol. Sect. 268, 95-104.
  • 55.Jacobson, K.A., Fischer, B., van Rhee, A.M. (1995) Molecular probes for muscarinic receptors: functionalized congeners of selective muscarinic antagonists. Life Sci. 56, 823- 830.
  • MRS Home Page

    Staff
    Selected Publications
    Adenosine Receptors
    P2 Receptors
    Medicinal Chemistry
    Computational Chemistry
    The Purine Club
    GPCRs
    Links

    Chief: Dr. Kenneth Jacobson
    The Molecular Recognition Section (MRS) is in the Laboratory of Bioorganic Chemistry at the National Institute of Diabetes and Digestive and Kidney Diseases which is part of the National Institutes of Health, Bethesda, MD, USA. General inquiries may be addressed to NIDDK, NIH, Building 31, Room 9A04, 31 Center Drive, MSC 2560, Bethesda, MD 20892-2560, USA.
    Privacy || Copyright, Disclaimers & Access Restrictions || Accessibility

    National Institute of Diabetes, Digestive & Kidney Diseases U. S. National Institutes of Health U.S. Department of Health & Human Services USA.gov Logo - link to the U.S. government's official web portal
    National Institutes of Health