Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Aug 13.
Published in final edited form as: Mol Pharmacol. 1996 Sep;50(3):512–521.

Hydrophilic Side Chains in the Third and Seventh Transmembrane Helical Domains of Human A2A Adenosine Receptors Are Required for Ligand Recognition

QIAOLING JIANG 1, A MICHIEL VAN RHEE 1, JEONGHO KIM 1, SUSAN YEHLE 1, JüRGEN WESS 1, KENNETH A JACOBSON 1
PMCID: PMC3418326  NIHMSID: NIHMS397603  PMID: 8794889

SUMMARY

Hydrophilic residues of the G protein-coupled human A2A adenosine receptor that are potentially involved in the binding of the ribose moiety of adenosine were targeted for mutagenesis. Residues in a T88QSS91 sequence in the third transmembrane helical domain (TM3) were individually replaced with alanine and other amino acids. Two additional serine residues in TM7 that were previously shown to be involved in ligand binding were mutated to other uncharged, hydrophilic amino acids. The binding affinity of agonists at T88 mutant receptors was greatly diminished, although the receptors were well expressed and bound antagonists similar to the wild-type receptor. Thus, mutations that are specific for diminishing the affinity of ribose-containing ligands (i.e., adenosine agonists) have been identified in both TM3 and TM7. The T88A and T88S mutant receptor fully stimulated adenylyl cyclase, with the dose-response curves to CGS 21680 highly shifted to the right. A Q89A mutant gained affinity for all agonist and antagonist Iigands examined in binding and functional assays. Q89 likely plays an indirect role in ligand binding. S9OA, S91A, and S277C mutant receptors displayed only moderate changes in ligand affinity. A 5281 N mutant gained affinity for all adenosine denyatives (agonists), but antagonist affinity was generally diminished, with the exception of a novel tetrahydnobenzothiophenone derivative.


Adenosine acts as a neuromodulator in the central and peripheral nervous systems and as a homeostatic regulator in a variety of other systems, including the cardiovascular, renal, and immune systems (1). Four pharmacologically distinct adenosine receptor subtypes, A1, A2A, A2B, and A3, have been cloned (2, 3). Activation of adenosine A2A receptors, in general, increases the energy supply in various organs. The regulation of blood pressure by centrally (4) and peripherally (5) mediated mechanisms involves A2A receptors. Activation of A2A receptors results in vasodilatation, and this effect has been examined as a potential antihypertensive therapy using selective A2A agonists such as CGS 21680 {2-[4-[(2-carboxyethyl)phenyl]ethyl-amino]-5′-N-ethylcarboxamidoadenosine) (6). A2A receptors, present in platelets, where they inhibit aggregation, and in the liver, have also been investigated for therapeutic applications. In the brain, A2A receptors occur primarily in the striatum, where they are colocalized with D2 dopamine receptors (7). Adenosine acts in a manner opposite to dopamine and thus elicits locomotor depression (8). Diseases in which the dopaminergic system is hyperactive [e.g., schizophrenia (9) and Huntington’s disease (8)] may be mitigated by A2A receptor activation. Parkinson’s disease, in which the dopaminergic system is hyporesponsive, may be amenable to treatment with selective A2A receptor antagonists (10, 11).

We characterized A adenosine receptors through the design and use of novel ligand probes, including radioligands and affinity labels (1, 12), and biotinylated probes and fluorescent labels (13, 14). In addition, site-directed mutagenesis (15) and molecular modeling (16) were used for A2A receptor characterization. A rhodopsin-based model (15) of the human A2A-receptor has been proposed. This model is highly consistent with mutagenesis results regarding orientation of individual amino acid residues within the central ligand-binding cavity, thus implicating residues in TM5, TM6, and TM7 of the A2A receptor in ligand recognition (15). In particular, two histidine residues (H250 and H278) and S277 are essential for ligand binding. The adenine moiety likely interacts with a pocket of aromatic amino acids in TM6, and the ribose moiety likely interacts with a hydrophilic region in TM7. Mutagenesis studies of A1 adenosine receptors have identified some of the corresponding residues as necessary for ligand binding. The corresponding histidine residues (17), a hexapeptide region of TM5 (18), and 1274 and 5277 in TM7 of bovine A1 receptors (19, 20) are involved in ligand binding. Thus, the position of the receptor-bound adenosine is likely to be similar in these subtypes.

In this study, we introduced the novel finding that residues of TM3 are essential for the binding of ligands to the human A2A receptor and further explore the role of the previously modified serine residues in TM7 (15). A long-range goal of this investigation is the design of more selective pharmacological agents based on structural differences in receptors.

Experimental Procedures

Materials

Human A2A adenosme receptor cDNA (pSVLA2A) was provided by Dr. Marlene A. Jacobson (Merck Research Labs, West Point, PA). Taq polymerase for the PCR was purchased from PerkinElmer Cetus (Norwalk, CT). All enzymes used in this study were ohtamed from New England Biolabs (Beverly, MA). The agonists CGS 21680, NECA, R-PIA, CADO, and DPMA and the antagonists XAC and CGS 15943 {9-chloro-2-(fiiryl)[1,2,4}triazolo[1,5-c]quinazolin-5-amine) were from RBI (Natick, MA). [3H]CGS 21680 (41.2 Ci/mmol) and [3HIXAC (118 Ci/mmol) were obtained from DuPont-New England Nuclear (Boston, MA), and [3H]adenine (15 Ci/mmol) was purchased from American Research Chemicals Inc. (St. Louis, MO). IB-MECA was prepared as described previously (21). BTH4 was obtained from Maybridge Chemicals (Trevillet, UK). Chemical structures of the agonist ligands used in this study may be found in Kim et al. (15). FBS and o-phenylenediamine dihydrochloride were purchased from Sigma Chemical Co. (St. Louis, MO). The Sequenase Kit, ATP, and cAMP were from United States Biochemical (Cleveland, OH). All oligonucleotides used were synthesized by Bioserve Biotechnologies (Laurel, MD). A monoclonal antibody (12CA5) against an HA epitope was purchased from Boehringer-Mannheim Biochemicals (Indianapolis, IN), and goat anti-mouse IgG (γ-chain specific) antibody conjugated with horseradish peroxidase was purchased from Sigma. DEAE-dextran was obtained from Pharmacia LKB (Piscataway, NJ). Rolipram was a gift from Schering AG (Berlin, Germany). SCH 58261 {5-amino-7(phenylethyl)-2-(2-furyl)-pyrazolo[4,3-e]1,2,4-triazolo[1,5-c]pyrimidine) was a gift from Dr. E. Ongini (Schering, Milan, Italy) and Prof. P. Baraldi (Urnversity of Ferrara, Italy). ZM 241385 {4-(2-[7-amino-2-(2-furyl)[1,2,4}triazolo[2,3a][1,3,5]triazinyl-amino]ethyl)-phenol) was a gift from Dr. S. Poucher (Zeneca, Macclesfield, UK).

Plasmid construction and site-directed mutagenesis

The coding region of pSVLA2A was subcloned into the pcD cDNA expression vector (22), yielding pcDA2A. All mutations were introduced into pcDA2A by using standard PCR mutagenesis techniques (23). The accuracy of all PCR-derived sequences was confirmed by dideoxy sequencing of the mutant plasmids (24).

Epitope tagging

A 9-amino acid sequence derived from the influenza virus HA protein (TAC CCC TAC GAC GTC CCC GAC TAC GCC; peptide sequence: YPYDVPDYA) was inserted after the second methionine residue at the extracellular amino terminus of the A2A adenosine receptor gene (15). Oligonucleotides containing the HA-tag sequence were designed and used to generate PCR fragments, which were then used to replace the homologous wild-type pcDA2A sequences.

Transient expression of mutant receptors in COS-7 cells

COS-7 cells (2 × 106) were seeded into 100-mm culture dishes contaming 10 ml of Dulbecco’s modified Eagle’s medium supplemented with 10% FBS. Cells were transfected with plasmid DNA (4 μg of DNA/dish) according to the DEAE-dextran method (25) ~24 hr later and grown for an additional 72 hr at 37°.

Membrane preparation and radioligand binding assay

Cells were scraped into ice-cold lysis buffer (4 ml of 50 mM Tris, pH 6.8, at room temperature, containing 10 m i MgCl2). Harvested cells were homogenized using a Polytron homogenizer (Brinkmann, Westbury, NY) and then spun at 27,000 × g for 15 mm. Cell membranes (pellet) were resuspended in the same buffer.

For saturation and competition binding experiments, each tube contained 100 μl of membrane suspension (containing 2 units/ml adenosine deaminase (Boehringer-Mannheim Biochemicals), 50 μl of radioligand, and either 50 μl of buffer/competitor (50 mM Tris, pH 6.8, 10 mM MgCl2) or 50 μl of 80 μm CADO in buffer (to determine nonspecific binding). The mixtures were incubated at 25° for 120 mm, filtered, and washed three times with ~5 ml of ice-cold buffer/wash using a Brandel cell harvester (Gaithersburg, MD). Pharmacological parameters were analyzed using the KaleidaGraph program (version 3.01; Abelbeck/Synergy Software, Reading, PA). Statistical analysis was performed using the alternate t test (InStat version 2.04; GraphPad, San Diego, CA).

cAMP determination

cAMP levels were determined by measurement of the conversion of [3H]ATP to [3H]cAMP. One day after transfection, cells were transferred from 100-mm dishes into six-well dishes (~3 × 105 cells/well) and incubated with culture media contaming 2 μCi/ml [3H]adenine. After 24 hr, the cultures were washed and incubated with 1 ml/well Hanks’ balanced salt solution containing 0.1 mM rolipram for 15 mm at 37°. The cells were incubated with different concentrations ofthe agonist CGS 21680 (in culture media) for 30 mm at 37°. The reaction was terminated by aspiration of the media and the addition of 1 ml of ice-cold 5% trichloroacetic acid containing 1 mM ATP and 1 mM cAMP. After a 30-mm incubation at 4°, cell lysates were eluted through sequential chromatography on Dowex and alumina columns (26). cAMP formation is expressed as percentage of maximal stimulation of conversion of [3H]ATP to [3H]cAMP (27). At agonist concentrations of > 100 μm, a stimulation was observed in nontransfected COS-7 cells (15), and these values were subtracted from values obtained in the transfected mutant receptor cells.

ELISA

For indirect cellular ELISA measurements, cells were transferred to 96-well dishes (4–5 × 104 cells/well) 1 day after transfection. At ~48 hr after splitting, cells were fixed in 4% formaldehyde in phosphate-buffered saline for 30 mm at room temperature. After washing with phosphate-buffered saline three times and blocking with Dulbecco’s modified Eagle’s medium (containing 10% FBS), cells were incubated with HA-specific monoclonal antibody (12CA5; 20 μg/ml) for 3 hr at 37°. Plates were washed and incubated with a 1:2000 dilution of a peroxidase-conjugated goat anti-mouse IgG antibody (Sigma) for 1 hr at 37°. Hydrogen peroxide and o-phenylenediamine (each 2.5 mm in 0.1 m phosphate/citrate buffer, pH 5.0) served as substrate and chromogen, respectively. The enzymatic reaction was stopped after 30 mm at room temperature with 1 m H2SO4 solution containing 0.05 m Na2SO3, and the color development was measured bichromatically in the BioKinetics reader (EL 312; Bio Tek Instruments, Winooski, VT) at 490 nm and at 630 nm (base-line).

Results

Sequence alignments for selected transmembrane regions of adenosine receptors and other GPCRs are shown in Fig. 1. The residues of the human A2A receptor selected as targets for site-directed mutagenesis are shown in bold; they include hydrophilic residues potentially involved in the binding of the ribose moiety, which was suggested in our previous studies (15, 16) to occur at TM7 (from mutagenesis results) and at TM3 (predicted by molecular modeling). Mutated residues include 591 and 5281, which are highly conserved among GPCRs. Another mutated residue, T88, is conserved among all known adenosine receptor sequences. Other mutated residues include those that are conserved between A1 and A2 receptors (Q89, 590) or between A2 and A3 receptors (5277). Each of these amino acid residues was individually replaced with alanine and/or other amino acids (see below). In addition, each mutant contained an epitope-tag sequence included at the amino terminus for immunological detection (see below). The pharmacological properties were compared with those of the wild-type receptor that was similarly modified.

Fig. 1.

Fig. 1

Location of mutations carried out in this study, illustrated through an alignment of the TM3 and TM7 of selected receptor subtypes. Bold, residues mutated in A2A receptors (present study) and in a chimeric bovine A1/rat A3 construct (18). Accession numbers are h A2a (human) P29274, rA2a (rat) P30543, cA2a (dog) P11617, gpA2a (guinea pig) U04201 , hA2b (human) P29275, hAl (human) P30542, hA3 (human) P33765, rA3 (rat) P28647, m3 (rat) P08483, β2 (hamster) P04274, and NK1 (human) P25103. In TM3, the canine and guinea pig A2A sequences are identical to the rat A2A sequence.

Ligand binding properties of mutant human A adenosine receptors

Radioligand saturation studies and competition binding experiments using a fixed concentration of either the agonist [3H]CGS 21680 at a concentration of 15 nm (6) or the antagonist [3H]XAC at a concentration of 2 nm (28) were carried out on the wild-type and mutant receptors. The agonists selected for competition (Table 1) included adenosine derivatives modified at the 2 position (CADO), the N6 position (DPMA, R-PIA, and ADAC), the 5′ position (NECA), the 5′ and 2 positions (CGS 21680), and the 5′ and N6 positions (IB-MECA). A diverse set of adenosine antagonists (Fig. 2), including XAC, a potent and nonselective nonxanthine (CGS 15943), several recently reported (29, 30) potent and highly selective A2A antagonists (SCH 58261 and ZM 241385), and a novel structure (BTH4) that does not contain any nitrogen atoms (31), were studied in competition for [3H]CGS 21680 binding.

TABLE 1. Binding characteristics of wild-type and mutant human A2A-adenosine receptors using the antagonist radioligand [3H]XAC.

Data are presented as mean ± standard deviation of two or three independent experiments, each performed in duplicate. Each sample contained 7–11 μg of membrane protein/tube. Agonist and antagonist binding affinities [Ki values, structures in Fig. 2 and Jacobson et al., (35)] were determined in 3H]XAC (1.0 nM) competition binding studies using membrane homogenates prepared from transiently transfected COS-7 cells, as described in Experimental Procedures. Ki values were calculated from IC50 values by using the KaleidaGraph program. All constructs contain an HA-tag sequence at the amino terminus (15).

Construct
Compound Wild-type T88A T88S 188R
Bmax (pmol/mg) [3H]XAC 14.9 ± 1.2 6.54 ± 1.61b 6.26 ± 070b 5.54 ± 0.47b
Kd (nM) [3H]XAC 8.83 ± 1.46 10.8 ± 18d 5.76 ± l.82d 4.21 ± 1.17c
Ki (nM) Agonists
 CADO 118 ± 22 16,400 ± 3,200c 776 ± 313c 9,140 ± 2,590c
 DPMA 65.5 ± 0.3 5,710 ± 2,520c 655 ± 134c 4,090 ± 760c
 NECA 19.1 ± 1.7 1,590 ± 280c 778 ± 224c 2,660 ± 1,390c
 CGS 21680 26.7 ± 6.5 22,100 ± 2,900b 422 ± 180c 4,620 ± 620b
Antagonists
 CGS 15943 1.20 ± 0.30 7.98 ± 1.71c 6.28 ± 1.68c 5.00 ± 052b
 ZM 241385 0.758 ± 0.135 3.47 ± 2.41d 1.16 ± 0.20d 2.00 ± 0.71d
a

p <0.001.

b

p <0.01.

c

p <0.05.

d

Not significant.

FIg. 2.

FIg. 2

Structures of agonist (A) and antagonist (B) ligands used in this study.

Among TM3 mutations, amino acid substitution of T88 distinguished between agonist and antagonist ligands (Table 1). The specific binding of [3H]CGS 21680 was greatly diminished (i.e., <2% of the specific binding of 15 nm [3H]CGS 21680 observed with the wild-type receptor) in the T88A, T88S, and T88R mutant receptors, whereas binding of [3H]XAC was similar to that ofthe wild-type receptor. The Kd determined for [3H]XAC at the wild-type receptor (8.83 ± 1.46 nM) was not significantly different from the Kd at the T88 mutant receptors (4.2–10.8 nM). The receptor densities, Bmax, in membranes of COS-7 cells transfected with the T88 mutant receptors (6–7 pmollmg protein) were approximately half those of the COS-7 cells expressing wild-type receptors (14.9 ± 1.2 μmol/mg of protein). Ki values for competition of binding of [3H]XAC by agonists in T88A and T88R mutant receptors indicated a 60–830-fold lower affinity than in wildtype receptors. Even the T88S mutant receptor, which differs only in the absence of a methyl group, bound agonists with 6.6-fold (for CADO) to 41-fold (for NECA) lower affinity. Affinities of the antagonists XAC and ZM 241385 were unchanged in the T88S mutant receptor, whereas CGS 15943 was 5.2-fold weaker than in wild-type receptors. In a comparison of T88S and T88A mutants, the largest contrast in ligand affinity occurred for the agonist CGS 21680. The removal of the serine hydroxyl group of the T88S mutant, already impaired in its ability to bind agonists, diminished by 52-fold the affinity of this adenosine derivative. Thus, the requirements for agonist binding at position 88 are highly specific, whereas the binding of antagonists was less dramatically affected by mutation at this site.

The T88 mutant receptors were found to be properly expressed on the cell surface using an ELISA (15). To estimate approximate levels of receptor protein present in the plasma membrane, a standard curve was constructed from different batches oftransfected COS-7 cells expressing different levels of HA-tagged A2A wild-type receptors (see Experimental Procedures and Ref. 15). This ELISA procedure does not interfere with the intactness of the plasma membrane barrier; thus, the assay is specific for receptor molecules having the proper orientation (extracellular amino terminus). Expression levels for the various mutants (HA-tagged wild-type receptor = 100%) determined according to this method were as follows (six experiments): T88A, 77.3 ± 12.5%; T88S, 89.8 ± 43.0%; and T88R, 134 ± 38.3%. The combination of ELISA and radioligand binding results indicates that the T88 residue is important, either directly or indirectly, for the high affinity binding of agonist ligands.

In contrast, replacement of other amino acids, located carboxyl-terminally to T88 in TM3, did not have as detrimental an effect on binding of the agonist radioligand [3H]CGS 21680 (Table 2). The S9OA mutant receptor displayed a slight trend toward higher affinity of both agonists and antagonists relative to the wild-type receptor (Kd of [3H]CGS 21680 was 2-fold lower). In competition binding studies at the S91A mutant receptor, only slight changes were noted in Ki values relative to the wild-type receptor. CA.DO, CGS 21680, and ADAC (N6-substituted) were approximately half as potent in displacing radioligand binding as the wild-type receptor.

TABLE 2. Binding characteristics of wild type and mutant human Aı-adenosine receptors using the agonist radioligand [3H]CGS 21680.

Data are presented as mean ± standard deviation of two or three independent experiments, each performed in duplicate. Each sample contained 7-11 μg of membrane protein/tube. Agonist and antagonist binding affinities [Ki values, structures in Fig. 2 and Jacobson et al. (35)] were determined in [3H]CGS 21 680 (15 nM) competition binding studies using membrane homogenates prepared from transiently transfected COS-7 cells, as described in Experimental Procedures. Ki values were calculated from IC50 values by using the KaleidaGraph program. All constructs contain an HA-tag sequence at the amino terminus (15).

Construct
Compound Wild-type Q89A S9OA S91A S277C S281N
Bmax (pmol/mg) [3H]CGS 21680 11.2 ± 0.3 13.2 ± 1.5d 12.1 ± 1.2d 14.3 ± 2.1d 1.20 ± 0.01a 6.26 ± 0.86c
Kd (nM) [3H]CGS 21680 31.0 ± 1.0 6.70 ± 0.86a 16.1 ± 35c 59.7 ± 77c 37.2 ± 1.2b 12.3 ± 1.88
Ki (nM) Agonists
 CADO 100 ± 14 19.1 ± 6.2b 19.4 ± 4.0c 224 ± 7b 384 ± 22a 7.43 ± l.80b
 DPMA 77.3 ± 4.8 16.6 ± 3.0a 11.1 ± 1.7b 88.4 ± 28.6d 355 ± 113d 5.92 ± 2.00b
 NECA 27.6 ± 0.9 1.90 ± 0.22a 5.90 ± 0.40a 35.5 ± 1.6b 129 ± 12b 4.50 ± 0.50a
R-PIA 299 ± 83 28.6 ± 10.0c 76.1 ± 20.0c 314 ± 13d 1590 ± 570d 17.6 ± 5.4c
 IB-MECA 435 ± 100 3.21 ± 0.07c 78.5 ± 13.6c 614 ± 212d 1755 ± 400c 53.6 ± 15.1c
 ADAC 1330 ± 50 41.2 ± 2.6b 675 ± 83c 2530 ± 560d 3580 ± 168b 156 ± 88a
Antagonists
 SCH 58261 1.85 ± 0.15 0.246 ± 0.084a 1.50 ± 0.50d 2.30 ± 1.12d 4.70 ± 0.50c 7.85 ± 0.55b
 CGS 15943 1.95 ± 0.45 0.455 ± 0.165b 0.925 ± 0.475d 1.95 ± 0.56d 1.53 ± 11d 5.10 ± 0.90c
 XAC 8.71 ± 1.60 2.01 ± 0.72c 16.6 ± 16b 18.1 ± 2.3c 10.3 ± 4.6d 82.3 ± 2.1a
 BTH4 859 ± 14 133 ± 15a 508 ± 101c 1880 ± 30a 279 ± 17a 245 ± 23a
 BTH4 859 ± 14 133 ± 15a 508 ± 101c 1880 ± 30a 279 ± 17a 245 ± 23a
 ZM 241385 0.525 ± 0.035 0.197 ± 0.041a 0.360 ± 0.080d 1.09 ± 0.11c 1.35 ± 0.35d 1.93 ± 0.23b
a

p <0.001.

b

p <0.01.

c

p <0.05.

d

Not significant.

The mutation of Q89 gave an unanticipated enhancement of affinity for both agonist and antagonist ligands (Tables 2 and 3). From saturation binding, it was determined that the Q89A mutant receptor had a 4.6-fold greater affinity for [3H]CGS 21680 (6.70 ± 0.86 nm) than the wild-type receptor (31.0 ± 1.0 nm). The affinities ofthe competing ligands at the Q89A mutant receptor (Table 2) were dramatically higher than in the wild-type receptor. IB-MECA had the greatest ratio of affinities (136-fold versus the wild-type receptor; Fig. 3A), whereas the potencies of ADAC (Fig. 3B) and NECA were enhanced 32- and 15-fold, respectively. Other N6-modified analogues, such as R-PIA and DPMA, and the C2-modified analogue CADO were 5–12-fold more potent in binding at the Q89A mutant receptor. The antagonists generally displayed a 4–5-fold enhancement of affinity at the Q89A mutant receptor versus the wild-type receptor.

TABLE 3. Radioligand binding characteristics of Q89 mutant human A2A-denosine receptors using an agonist radioligand.

Data are presented as mean ± standard deviation of two or three independent experiments, each performed in duplicate. Each sample contained 7-11 μg of membrane protein/tube. Saturation of binding of [3H]CGS 21680 using membrane homogenates prepared from transiently transfected COS-7 cells, as described in Experimental Procedures. All constructs contain the HA-epitope tag sequence at the amino terminus (15).

Constructa Kd B max
nM pmol/mg
Wild-type 31.0 ± 1.0 11.2 ± 0.3
Q89A 6.70 ± 0.86b 13.2 ± 1.5e
Q89N 98.7 ± 23.6d 8.71 ± 0.69c
Q89S 93.3 ± 11.5d 14.6 ± 34e
Q89L 35.6 ± 0.7c 8.68 ± 0.32c
Q89H 46.8 ± 0.4c 11.4 ± 0.2e
Q89R 50.8 ± 10.8e 4.68 ± 0.00b
a

Specific binding versus [3H]XAc (5.5 nM) showed levels comparable to HA-tagged wild-type receptors for Q89A, Q89N, Q89S, Q89L, and Q89H mutant receptors. For the Q89R mutant, <8% of the specific binding found for wild-type receptors was observed for Q89R.

***

p <0.001;

**

p<0.01;

*

p <0.05: n.s. not significant.

b

p <0.001.

c

p <0.01.

d

p <0.05.

e

Not significant.

Fig. 3.

Fig. 3

Displacement of binding of the agonist radioligand [3H]CGS 21680 from HA-tagged A2A wildtype (WT) and Q89A mutant receptors expressed in COS-7 cells. Competitors used were IB-MECA (A) and ADAC (B). Competition binding studies were carried out using membrane homogenates prepared from transfected COS-7 cells, as described in Experimental Procedures. Data are from a representative experiment performed in duplicate.

Other amino acids were substituted at position 89 to probe the contribution to ligand recognition of factors such as size, polarity, aromaticity, or charge (Table 3). Substitution of Q89 by residues, charged or uncharged, larger than alanine (Q89S, Q89N, Q89L, Q89H, and Q89R) did not preclude the high affinity binding of[3H]CGS 21680. Histidine was chosen because it occurs at this position in A3 receptors; however, no selective enhancement of the affinity of the A3 receptor-selective agonist IB-MECA was observed. The mutant receptors Q89S and Q89N had a lower affinity (3-fold) for the radioligand than wild-type receptors. The mutant receptor Q89A was unique in having increased affinity for [3HICGS 21680. Thus, for binding of the agonist radioligand, there existed a great deal of tolerance for substitution at position 89.

Competition binding was studied for three ligands (NECA, IB-MECA, and XAC) at this set of mutant receptors at position 89 (Table 3). The affinities showed an approximate dependence on the size of the amino acid side chain. Solvent-accessible surface, a theoretical value reported for each amino acid (36) that was used as a relative steric indicator for ordering the amino acids (Fig. 4), fell within the range of 224 Å2 for A to 355 Å2 for R, and affinity varied over several orders of magnitude. For each competitor, the Ki value tended to increase as a function of the size of residue 89. The effects on affinity relative to wild-type receptors ranged from a large enhancement (for the two agonists NECA and IB-MECA at alanine, serine, asparagine, and leucine mutants) to substantial reduction (for the antagonist XAC at H and R mutant receptors).

Fig. 4.

Fig. 4

Plot of affinity (Ki in nA2A) in competition binding experiments of three adenosine receptor ligands at Q89 mutant receptors as a function of the amino acid residue. The amino acids were arranged in order of increasing solvent-accessible surface area, using theoretical values for each isolated amino acid as calculated by Hubbard et al. (36). Ki values for the agonist (NECA and lB-MECA) and antagonist (XAC) ligands (structures in Fig. 2) were determined in [3H]CGS 21680 (15 nM) competition binding studies using membrane homogenates prepared from transiently transfected COS-7 cells, as described in Experimental Procedures. Ki values were calculated from IC50 values by using the KaleidaGraph program. All constructs contain the HA epitope tag sequence at the amino terminus (15). ***, p < 0.001; **, p < 0.01; *, p < 0.05; n.s., not significant.

The TM7 mutant receptors S277C and S281N, both small polar substitutions capable of hydrogen bonding, were also constructed (Table 2). In a previous study (15), these serine residues were mutated to alanine and found to be required for ligand recognition (5277 for agonists and 5281 for all ligands). Changes in the affinities of agonists at the S277C mutant receptor were relatively modest (i.e., ≤5-fold lower than at the wild-type receptor). Some antagonists (XAC and CGS 15943) were approximately equipotent at S277C mutant receptors and at wild-type receptors. The nonxanthine A2A-receptor antagonists SCH 58261 and ZM 241385 were moderately diminished in affinity, whereas the novel tetrahydrobenzothiophenone derivative BTH4 displayed 3-fold enhanced affinity at S277C mutant receptors. At S281N mutant receptors, the affinities of agonists, including those substituted at N6, C2, or C5′ positions, were 6–17-fold higher than at wild-type receptors. The agonist order (decreasing) for enhancement of affinity produced by the S281N mutation was R-PIA (17.0-fold) > CADO (13.5-fold) = DPMA (13.1-fold) > ADAC (8.5-fold) = IB-MECA (8.1-fold) > NECA (6.1-fold) > CGS 21680 (2.5-fold). Antagonist affinity at the S281N mutant receptor was generally diminished (3–10-fold) compared with wild-type receptors, except for BTH4 (3.5-fold enhancement).

Functional assay

To determine whether T88A mutant receptors that lacked high affinity radioligand binding were still functional at high agonist concentrations, their ability to mediate increases in intracellular cAMP levels in transfected COS-7 cells was studied. Rolipram was used as an inhibitor of phosphodiesterases. The T88A and T88S mutant receptors showed a dose-dependent stimulation of cAMP production after treatment with CGS 21680, with EC50 values of 14.4 ± 0.423 and 0.323 ± 0.048 μm, respectively. Thus, functional potency of CGS 21680 was far less at the T88A and T88S mutant receptors than at wild-type receptors (EC50 = 0.915 ± 0.213 nM) by factors of 15,700 and 353, respectively. A maximal response of these mutant receptors was reached at ~10−4 M. The T88R mutant receptor however, showed minimal stimulation of adenylyl cyclase even at 1 mm CGS 21680 (Fig. 5).

Fig. 5.

Fig. 5

Stimulation of adenylyl cyclase in COS-7 cells transiently expressing HA-tagged A2A wild-type (WT) or mutant A2A-adenosine receptors in the presence of 2 units/mI adenosine deaminase and 0.1 m rolipram. The following receptors were studied: wild-type, T88A, T88R, and T88S mutant receptors. Transfected COS-7 cells were incubated for 30 mm at 37° (for details, see Experimental Procedures) with increasing concentrations of CGS 21680. Data are presented as percentage of maximal increase in cAMP above basal levels in the absence of CGS 21680 for a representative experiment. For each curve, the maximal stimulation represents a 4–5-fold stimulation over basal levels. At agonist concentrations of ≥100 μm, the stimulation observed in nontransfected COS-7 cells (15) was subtracted. EC50 values (average of three independent experiments, each carried out in duplicate) were wild-type receptor, 0.915 ± 0.213 nm; T88A, 14.4 ± 0.423 μm; and T88S, 0.323 ± 0.048 μm.

The 2- and 5′-disubstituted adenosine agonist CGS 21680 acting at the Q89A mutant receptor elicited a dose-dependent stimulation of cAMP production (Fig. 6, top), with an EC50 value of 0.223 ± 0.054 nm [i.e., 4-fold more potent (p < 0.05) than at wild-type receptors (EC50 = 0.915 ± 0.213 nm)]. A similar gain of potency was observed for the N6-substituted compound DPMA (EC50 = 10.8 ± 1.4 nm at the wild-type and 2.73 ± 0.59 nm at the Q89A mutant receptor; a 4-fold increase, p < 0.05; Fig. 6, bottom). This effect was even more pronounced for the 5′-substituted compound NECA (EC50 = 44.5 ± 6.2 nm at the wild-type and 2.91 ± 0.54 n t at the Q89A mutant receptor), whereas the Q89A mutation induced a 15.3-fold increase in potency (p < 0.01, Fig. 6, middle). Thus, there was a parallel between the enhanced binding affinities of agonists at the Q89A mutant receptor (Table 2) and their functional potencies. No increase in basal adenylyl cyclase activity was observed for this mutant receptor.

Fig. 6.

Fig. 6

Stimulation of adenylyl cyclase in COS-7 cells transiently expressing HA-tagged A2A wild-type (WT) or mutant A2A-adenosine receptors in the presence of 2 units/mI adenosine deaminase and 0.1 m rolipram. The following receptors were studied: wild-type and Q89A mutant receptors with CGS 21680 (top), NECA (middle), or DPMA (bottom). Transfected COS-7 cells were incubated for 30 mm at 37° (for details, see Experimental Procedures) with increasing concentrations of agonist. Data are presented as percentage of maximal increase in cAMP above basal levels in the absence of agonist for a representative experiment. For each curve, the maximal stimulation represents a 4-5-fold stimulation over basal levels. EC50 values were calculated averaged over three independent experiments, each carried out in duplicate.

Discussion

Ligand binding and stimulation of adenylyl cyclase in TM3 mutant human A adenosine receptors

The current study demonstrates clearly that hydrophilic residues of TM3 of the human A2A receptor are involved in ligand binding. Several TM3 mutants prepared in this study are highly unnatural in their ligand binding properties, having either enhanced (Q89A) or greatly diminished (T88A and T88R) affinity for various ligands. Alanine scanning mutagenesis showed that T88 is essential for high affinity agonist binding, whereas only moderate changes in ligand binding affinity occur on replacement of 590 and S91. T88A and T88R mutations had similar detrimental effects on agonist binding, whereas high affinity antagonist binding was still observed in these mutant receptors. Affinity in the T88S receptor mutant was decreased as well but to a lesser extent. Thus, a section of TM3 seems to be involved in ligand (especially agonist) recognition. This is consistent with a molecular model based on a rhodopsin template (15), which predicted that T88 and Q89 are in proximity to the ribose moiety of adenosine.

At high agonist concentrations, the T88A and T88S mutant receptors were active functionally in the stimulation of adenylyl cyclase (Fig. 5), with the dose-response curves right-shifted by 15,700- and 350-fold, respectively (p < 0.01 ). This change in potency reflects the relative agonist affinity shifts in the binding assays. The T88R mutant receptor seemed to be functionally impaired (Fig. 5), even at agonist concentrations that clearly displaced bound [3H]XAC (Table 1).

For single amino acid replacements of Q89, the ligand affinity varied from dramatic increases (e.g., IB-MECA at Q89A) to decreases for the larger amino acid substitutions. The trend of inverse dependence of ligand affinity on steric bulk ofthe side chain at position 89 (Fig. 4) was shown for a number of structurally divergent amino acids substituted at this position through mutagenesis. Because electronic factors of the amino acids seemed to be not as important as steric factors and because three selected competing ligands were similarly affected, it is hypothesized that residue Q89 affects the size of the crevice that constitutes the binding site. Diminishment ofthe size ofa sterically limiting side chain (i.e., in the Q89A mutant receptor) would increase the accessibility of the binding site, thus lowering the Ki values. This mechanism either might involve the side chain acting as an energy barrier to a ligand occupying the binding site or may be more indirect (e.g., by changing interhelical distances). Several other residues of the human A2A receptor (i.e., Y271) have been postulated to affect ligand binding through interhelical contacts (15), although alanine substitution decreased agonist affinity in that case.

The finding of enhanced affinity for all ligands, both agonists and antagonists, on single amino acid replacement, as in the Q89A mutant receptor, is a rather rare finding. An increase in agonist affinity, but not antagonist affinity, is often observed in constitutively active mutants (32, 33). The Q89A mutation, however, did not affect basal adenylyl cyclase levels. Even though affinity and potency of agonists were both increased in the Q89A mutant receptor versus the wild-type receptor, this mutant receptor was not constitutively active.

In the S90A mutant A2A receptor, both agonist and antagonist affinities were only slightly affected. Mutation ofS9l to alanine had virtually no effect on ligand affinity.

Structurally related differences in ligand binding in TM7 mutant human A2A adenosine receptors

In our previous study (15), we reported that residue S277 is importa.nt for agonist recognition. We now present data regarding a new mutant, S277C, in which ligand binding affinities are only slightly affected (agonist affinity was unaffected for CGS 21680 and decreased by 3–5-fold for other agonists). This mutant to a certain extent rescues ligand binding compared with the earlier S277A mutant, in which agonist affinity decreased 43–1070-fold relative to the wild-type receptor.

This study has revealed differences in affinity shifts between agonists and antagonists [e.g., S281N (agonists become more potent and antagonists less potent) and mutations of T88 (agonists selectively become much less potent)]. Therefore, a partially different set of amino acid residues in the receptor is involved in agonist versus antagonist binding. The results of chemical modification and conformational analysis (34, 35) as well as mutagenesis studies of adenosine receptors (15, 17) have suggested that the ribose moiety is coordinated to the histidine residue of TM7, common to all adenosine receptors. Furthermore, changes in affinity of the ribose-modified agonist NECA at A1 receptors are associated with mutation of an adjacent threonine residue (19, 20). Mutations that are specific for diminishing the affinity of ribose-containing ligands (i.e., adenosine agonists) have been identified in both TM3 and TM7, consistent with the previously reported molecular models of human A2A receptors that predicted that the ribose moiety of adenosine may bridge these domains (15, 16).

It was proposed (15) that 5281 probably is not directly involved in ligand binding, yet in the current study the S281N mutant receptor distinguishes between agonists and antagonists. All agonists examined displayed an increased affinity (2.5-fold for CGS 21680 to 17-fold for R-PIA) toward the S281N mutant receptor, whereas affinity for most antagonists was moderately decreased (3–9-fold). The A1-selective antagonist XAC showed the largest decrease in affinity. The A2-selective antagonists SCH 58261 and ZM 241385 exhibited an intermediate decrease, and the nonselective antagonist CGS 15943 had the smallest decrease. Interestingly, the novel tetrahydrobenzothiophenone derivative BTH4 showed a moderate increase in affinity for the S281N mutant receptor. Unlike other antagonists, BTH4 displayed enhanced affinity at both TM7 mutant receptors examined in this study (Table 2). Thus, the binding mode of this antagonist seems to be unique, as is its chemical structure. It is among the least purine-like of known adenosine antagonists, which are almost exclusively nitrogen-containing heterocycles.

Although there are several models for mapping the xanthine binding site onto the partially overlapping adenosine binding site (34, 37), no such exercises have been performed for the nonxanthine antagonists SCH 58261, ZM 241385, and CGS 15943. The current data suggest that the mode of nonxanthine antagonist binding to the receptor may differ significantly from the modes proposed for xanthine-based antagonists.

Structural homology of human A2A adenosine receptors to other GPCRs

In the biogenic amine GPCRs, there is an essential, conserved aspartate residue in TM3 that coordinates the positively charged secondary nitrogen group of the endogenous ligand. This has been demonstrated for adrenoceptors (38, 39) and receptors for dopamine (40, 41), histamine (42), serotonin (43, 44), and acetylcholine (45, 46). Adenosine, being uncharged at physiological pH, would not benefit from a strong electrostatic interaction with the receptor at this position. In all adenosine receptors, the residue homologous to the above-mentioned aspartate corresponds to a valine (V84 in the human A2A adenosine receptor). The current study has identified important residues for ligand recognition at positions in the human A2A adenosine receptor deeper in the membrane than V84.

The essential T88 of A2A receptors would be expected to be facing approximately the same direction as that of the essential aspartate residue involved in recognition of the secondary nitrogen of biogenic amines at their receptors, being four residues (i.e., one helical turn) closer to the cytoplasmic side. Residue C118 of the human D2 receptor, equivalent to T88, reacts with a thiol reagent, indicating that this residue is solvent exposed and thus faces the central binding cavity (41).

The S505R mutant of the human thyrotropin receptor (position equivalent to T88) was constitutively active (47), whereas none of the T88 mutants in the current study were constitutively active. Residue T88 of the A2A receptor aligns with Vi 16 of the NK1 receptor (Fig. 1). Mutation of this residue to leucine reversed the selectivity of antagonists (48). There is no direct evidence that Q89 is pointing directly into the binding cavity. By analogy with rhodopsin, residue E122, the equivalent to Q89, is facing the binding cavity. This residue was found to coordinate and neutralize the Schiff base of retinal (formed with K296), based on a blue shift in the absorption spectrum of the E122Q mutant (49). Residue A120 of the human D2 receptor (41), the position equivalent to 590, was found not to be exposed to the ligand binding cavity because the A12OC mutant receptor did not react with thiol reagents.

Almost one full helical turn down from T88 is 591. This residue is conserved as a serine in 98 of 156 GPCR sequences probed (e.g., m3 muscarinic receptor, hamster β2-adrenoceptor: Fig. 1) and therefore is not likely to confer ligand binding specificity. Mutations at this site in the hamster β2-adrenoceptor led to decreased receptor expression and improper post-transcriptional processing (S12OA; Ref. 50), although the 591A mutant A2A adenosine receptor had a Bmax value for [3H]CGS 21680 binding similar to wild-type receptors. This site was shown to be facing the binding cavity in the human D2 receptor (S121C; Ref. 41).

Also in TM7, homology to amino acids known to be involved in ligand binding was found. In the rat ml muscarinic receptor, the mutant receptor C407S, corresponding to residue 5277 ofthe A2A receptor, had decreased agonist affinity (51). Residue S281 corresponds to the essential S319 in the hamster β2-adrenergic receptor (50). In the human luteinizing hormone receptor, the naturally occurring mutation at the same site, S616Y, results in hypogonadism and decreased agonist affinity (52).

Conclusions

Results of the current study strongly imply that residues of TM3 are involved in ligand recognition and extend the findings of Kim et al. (15) concerning TM7. Thus, it seems that hydrophilic residues in both TM3 and TM7 are important for recognition of the ribose moiety. The binding affinity and functional activity of agonists at T88 mutant receptors were greatly diminished. A Q89A mutant gained affinity for all agonist and antagonist ligands examined. Q89 likely plays an indirect role in ligand binding. Furthermore, the fact that all of the residues examined in this study are conserved among most adenosine receptors (species and subtypes) suggests that the proposed mode of binding of agonists is common or very similar in these receptors. Divergent effects of mutagenesis within TM3 and TM7 on the binding of adenosine derivatives, xanthines, and nonxanthine antagonists suggest nonidentical mechanisms of binding.

Acknowledgments

We thank Prof. Gary Stiles and Dr. Mark Olah (Duke University, Durham, NC) for helpful discussions and Dr. Marlene Jacobson (Merck, West Point, PA) for providing the human A A plasmid. We thank Marc Glashofer for assisting with binding experiments. We thank Dr. Simon Poucher of Zeneca Pharmaceuticals (Alderley Park, UK) and Dr. Ennio Ongini of Schering-Plough (Milan, Italy) for the gifts of ZM 241385 and SCH 58261, respectively.

ABBREVIATIONS

TM

(helical) transmembrane domain

ADAC

N6-[4-[[[[4-[[[(2-aminoethyl)amino]carbonyl]methyl]anilino]carbonyl]methyl]-phenyl]adenosine

BTH4

ethyl 3-benzylthio-4,5,6,7-tetrahydrobenzo[c]thiophen-4-one-1-carboxylate

CADO

2-chloroadenosine

DPMA

N6-[2-(3,5-dimethoxyphenyl)-2-(2-methylphenyl)ethyl]adenosine

ELISA

enzyme-linked immunosorbent assay

FBS

fetal bovine serum

GPCR

G protein-coupled receptor

HA

hemagglutinin

IB-MECA

N6-(3-iodobenzyl)adenosine-5′-N-methyluronamide

NECA

5′-N-ethylcarboxamidoadenosine

PCR

polymerase chain reaction

R-PIA

(R)-N6-phenylisopropyladenosine

XAC

(8-[4-[[[[(2-aminoethyl)-amino]carbonyl]methyl]oxy]phenyl]-1,3-dipropylxanthine)

References

  • 1.Jacobson KA, van Galen PJM, Williams M. Adenosine receptors: pharmacology, structure activity relationships, and therapeutic potential. J. Med. Chem. 1992;35:407–422. doi: 10.1021/jm00081a001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Libert F, Parmentier M, Lefort A, Dinsart C, van Sande J, Maenhaut C, Simons MJ, Dumont JE, Vassart G. Selective amplification and cloning of four new members of the G protein-coupled receptor family. Science (Washington D. C.) 1989;244:569–572. doi: 10.1126/science.2541503. [DOI] [PubMed] [Google Scholar]
  • 3.Jacobson M. Molecular biology of adenosine receptors. In: Belardinelli L, Pelleg A, editors. Ad.snosine and Adenine Nucleotides: From Molecular Biology to Integrative Physiology. Kluver; Norwell, MA: 1995. pp. 5–14. [Google Scholar]
  • 4.Barraco RA, Martens K, Parizon M, Normile HJ. Role of adenosine A2a receptors in the nucleus-accumbens. Prog. Neuropsychopharmacol. Biol. Psychiatry. 1994;18:545–553. doi: 10.1016/0278-5846(94)90011-6. [DOI] [PubMed] [Google Scholar]
  • 5.Olsson RA, Pearson JD. Cardiovascular purinoceptors. Pharmacol. Rev. 1990;3:761–845. doi: 10.1152/physrev.1990.70.3.761. [DOI] [PubMed] [Google Scholar]
  • 6.Jarvis MF, Schulz R, Hutchison AJ, Do UH, Sills MA, Williams M. [3H]CGS 21680, a selective A2 adenosine receptor agonist directly labels A2 receptors in rat brain. J. Pharmacol. Exp. Ther. 1989;251:888–893. [PubMed] [Google Scholar]
  • 7.Ferr#{233} S, O’Connor WT, Snaprud P, Ungerstedt U, Fuxe K. Antagonistic interaction between adenosine A2a receptors and dopamine D2 receptors in the ventral striopallidal system: implications for the treatment of schizophrenia. Neuroscience. 1994;03:765–773. doi: 10.1016/0306-4522(94)90521-5. [DOI] [PubMed] [Google Scholar]
  • 8.Nikodijevi#{233} 0, Jacobson KA, Daly JW. Acute treatment of mice with high-doses of caffeine: an animal-model for choreiform movement. Drug Dev. Res. 1993;30:121–128. [Google Scholar]
  • 9.Martin GE, Rossi DJ, Jarvis MF. Adenosine agonists reduce conditioned avoidance responding in the rat. Pharm. Biochem. Behav. 1993;45:951–958. doi: 10.1016/0091-3057(93)90146-k. [DOI] [PubMed] [Google Scholar]
  • 10.Schiffman SN, Vanderhaegen J-J. Adenosine A2 receptor regulation of striatal gene expression. In: Belardinelli L, Pelleg A, editors. Adenosine, and Adenine Nucleotides: From Molecular Biology to Integrative Physiology. lOuver; Norwell, MA: 1995. pp. 71–76. [Google Scholar]
  • 11.Kanda T, Shiozaki S, Shimada J, Suzuki F, Nakamura J. KF1783: a novel selective adenosine A2a receptor antagonist with anticataleptic activity. Eur. J. Pharmacol. 1994;256:263–268. doi: 10.1016/0014-2999(94)90551-7. [DOI] [PubMed] [Google Scholar]
  • 12.Barrington WW, Jacobson KA, Hutchison AJ, Williams M, Stiles GL. Identification of the A2 adenosine receptor binding subunit by photoaffinity crosslinking. Proc. Nati. Acad. Sci. USA. 1989;86:6572–6576. doi: 10.1073/pnas.86.17.6572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Jacobson KA, Pannell LK, Ji XD, Jarvis MF, Williams M, Hutchison AJ, Barrington WW, Stiles GL. Agonist-derived molecular probes for A2-adenosine receptors. J. Mol. Recognit. 1989;2:170–178. doi: 10.1002/jmr.300020406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.McCabe RT, Skolnick P, Jacobson KA. FITC-APEC: a fluorescent ligand for A2-adenosine receptors. J. Fluoresc. 1992;2:217–223. doi: 10.1007/BF00865279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kim JH, Wess J, van Rhee AM, Schoneberg T, Jacobson KA. Site-directed mutagenesis identifies residues involved in ligand recognition in the human A2, adenosine receptor. J. Biol. Chem. 1995;270:13987–13997. doi: 10.1074/jbc.270.23.13987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.IJzerman APP, van Galen JM, Jacobson KA. Molecular modeling of adenosine receptors: the ligand-binding site on the rat adenosine A2a receptor. Eur. J. Pharmacol. 1994;268:95–104. doi: 10.1016/0922-4106(94)90124-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Olah ME, Ren HZ, Ostrowski J, Jacobson KA, Stiles GL. Cloning, expression, and characterization of the unique bovine-A1 adenosine receptor: studies on the ligand binding site by site-directed mutagenesis. J. Biol. Chem. 1992;267:10764–10770. [PMC free article] [PubMed] [Google Scholar]
  • 18.Olah ME, Jacobson KA, Stiles GL. Identification ofan adenosine receptor domain specifically involved in binding of 5-substituted adenosine agonists. J. Biol. Chem. 1994;269:18016–18020. [PMC free article] [PubMed] [Google Scholar]
  • 19.Townsend-Nicholson A, Schofield PR. A threonine residue in the 7th transmembrane domain ofthe human Al-adenosine receptor mediates specific agonist binding. J. Biol. Chem. 1994;269:2373–2376. [PubMed] [Google Scholar]
  • 20.Tucker A, Robeva AS, Taylor HE, Holeton D, Bockner M, Lynch KR, Linden J. A, adenosine receptors: 2 ammo-acids are responsible for species-differences in ligand recognition. J. Biol. Chem. 1994;269:27900–27906. [PubMed] [Google Scholar]
  • 21.Gallo-Rodriguez C, Ji XD, Melman N, Siegman BD, Sanders LH, Orlina J, Fischer B, Pu QL, Olah ME, van Galen PJM, Stiles GL, Jacobson KA. Structure-activity-relationships of N6-benzyladenosine-5′-uronamides as A3-selective adenosine agonists. J. Med. Chem. 1994;37:636–646. doi: 10.1021/jm00031a014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Okayama H, Berg PA. A cDNA cloning vector that permits expression of cDNA inserts in mammalian cells. Mol. Cell. Bwl. 1983;3:280–289. doi: 10.1128/mcb.3.2.280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Higuchi R. Using PCR to engineer DNA. In: Ehrlich HA, editor. PCR Technology. Stockton Press; New York: 1989. pp. 61–70. [Google Scholar]
  • 24.Sanger R, Nicklen S, Coulson AR. DNA sequencing with chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA. 1977;74:5463–5467. doi: 10.1073/pnas.74.12.5463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Cullen BR. Use of eukaryotic expression technology in the functional analysis of cloned genes. Methods Enzymol. 1987;152:684–704. doi: 10.1016/0076-6879(87)52074-2. [DOI] [PubMed] [Google Scholar]
  • 26.Enjalbert A, Bockaert J. Pharmacological characterization of the D2 dopamine receptor negatively coupled with adenylate cyclase in rat antenor pituitary. Mol. Pharmacol. 1983;23:576–584. [PubMed] [Google Scholar]
  • 27.Weiss S, Sebben M, Garcia-Sainz JA, Bockaert J. D2-dopamine receptor-mediated inhibition of cyclic AMP formation in striate! neurons in primary culture. Mol. Pharmacol. 1985;27:595–599. [PubMed] [Google Scholar]
  • 28.Ji X-D, Lubitz DKJE, Olah ME, Stiles GL, Jacobson KA. Species differences in ligand affinity at central A3-adenosine receptors. Drug Dev. Res. 1994;33:51–59. doi: 10.1002/ddr.430330109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Baraldi PG, Manfredini S, Simoni D, Zappaterra L, Zocchi C, Dionisotti S, Ongini E. Synthesis of new pyrazolo[4,3-ell,2,4-triazolo{1,5-cJ pyrimidine and 1,2,3-triazolo[4,5-e]1,2,4-triazolo[1,5-c]pyrimidine displaying potent and selective activity as A2a adenosine receptor antagonists. Bioorg. Med. Chem. Lett. 1994;4:2539–2544. [Google Scholar]
  • 30.Poucher SM, Keddie JR, Singh P, Stoggall SM, Caulkett PWR, Jones G, Collis MG. The in-vitro pharmacology of ZM241385, a potent, nonxanthine, A2a selective adenosine receptor antagonist. Br. J. Pharmacol. 1995;115:1096–1102. doi: 10.1111/j.1476-5381.1995.tb15923.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.van Rhee AM, Siddiqi SM, Melman N, Shi D, Padgett WL, Daly JW, Jacobson KA. Tetrahydrobenzothiophenone derivatives as a novel class of adenosine receptor antagonists. J. Med. Chem. 1996;39:398–406. doi: 10.1021/jm9504823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kjelsberg MA, Cotecchia S, Ostrowski J, Caron MG, Lefkowitz RJ. Constitutive activation of the a) -adrenergic receptor by all amino acid substitution at a single site. J. Biol. Chem. 1992;267:1430–1433. [PubMed] [Google Scholar]
  • 33.Ren Q, Kurose H, Lefkowitz RJ, Cotecchia S. Constitutively active mutants of the a2-adrenergic receptor. J. Biol. Chem. 1993;268:16483–16487. [PubMed] [Google Scholar]
  • 34.van Galen PJM, Stiles GL, Michaels G, Jacobson KA. Adenosine A1 and A2 receptors: structure-function relationships. Med. Res. Rev. 1992;12:423–471. doi: 10.1002/med.2610120502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Jacobson KA, Stiles GL, Ji XD. Chemical modification and irreversible inhibition of striatal A2a-adenosine receptors. Mol. Pharmacol. 1992;42:123–133. [PMC free article] [PubMed] [Google Scholar]
  • 36.Hubbard SJ, Campbell SF, Thornton JM. Molecular recognition: conformational analysis of limited proteolysis sites and serine protease inhibitors. J. Mol. Biol. 1991;220:507–530. doi: 10.1016/0022-2836(91)90027-4. [DOI] [PubMed] [Google Scholar]
  • 37.van der Wenden EM, Price SL, Apaya RP, IJzerman AP, Soudijn W. Relative binding orientations of adenosine A, receptor ligands: a test case for Distributed Multipole Analysis in medicinal chemistry. J. Comput. Aided Mol. Design. 1995;9:44–54. doi: 10.1007/BF00117277. [DOI] [PubMed] [Google Scholar]
  • 38.Strader CD, Sigal IS, Candelore MR, Rends E, Hill WS, Dixon RA. Conserved aspartic acid residues 79 and 1 13 of the f3-adrenergic receptor have different roles in receptor function. J. Biol. Chem. 1988;263:10267–10271. [PubMed] [Google Scholar]
  • 39.Strader CD, Gaffney T, Sugg EE, Candelore MR, Keys R, Patchett AA, Dixon RA. Allele-specific activation of genetically engineered receptors. J. Biol. Chem. 1991;266:5–S. [PubMed] [Google Scholar]
  • 40.Mansour A, Meng F, Meador WJ, Taylor LP, Civelli 0, Akil H. Site-directed mutagenesis of the human dopamine D2 receptor. Eur. J. Pharmacol. 1992;227:205–214. doi: 10.1016/0922-4106(92)90129-j. [DOI] [PubMed] [Google Scholar]
  • 41.Javitch JA, Fu D, Chen J, Karlin A. Mapping the binding-site crevice of the dopaniine D2 receptor by the substituted-cysteine accessibility method. Neuron. 1995;14:825–831. doi: 10.1016/0896-6273(95)90226-0. [DOI] [PubMed] [Google Scholar]
  • 42.Gantz I, Del Valle J, Wang LD, Tashiro T, Munzert G, Guo YJ, Konda Y, Yamada T. Molecular basis for the interaction of histamine with the histamine H2 receptor. J. Biol. Chem. 1992;267:20840–20843. [PubMed] [Google Scholar]
  • 43.Ho BY, Karschin A, Branchek T, Davidson N, Lester HA. The role of conserved aspartate and serine residues in ligand binding and in function ofthe 5-HTIA receptor: a site-directed mutation study. FEBS Lett. 1992;312:259–262. doi: 10.1016/0014-5793(92)80948-g. [DOI] [PubMed] [Google Scholar]
  • 44.Wang CD, Gallaher TK, Shih JC. Site-directed mutagenesis of the serotonin 5-hydroxytrypamine2 receptor: identification of amino acids necessary for ligand binding and receptor activation. Mol. Pharmacol. 1993;43:931–940. [PubMed] [Google Scholar]
  • 45.Fraser CM, Wang CD, Robinson DA, Gocayne JD, Venter JC. Site-directed mutagenesis of ml muscarinic acetylcholine receptors: conserved aspartic acids play important roles in receptor function. Mol. Pharmacol. 1989;36:840–847. [PubMed] [Google Scholar]
  • 46.Kurtenbach E, Curtis CA, Pedder EK, Aitken A, Harris AC, Hulme EC. Muscarinic acetylcholine receptors: peptide sequencing identifies residues involved in antagonist binding and disulfide bond formation. J. Biol. Chem. 1990;265:13702–13708. [PubMed] [Google Scholar]
  • 47.Vassart G, Parma J, Van Sande J, Dumont JE. The thyrotropin receptor and the regulation ofthyrocyte function and growth: update 1994. In: Braverman LE, Refetoff S, editors. Endocrine Reviews Monographs 3. Clinical and Molecular Aspects of Diseases of the Thyroid. The Endocrine Society Press; Bethesda, MD: 1994. pp. 77–80. [Google Scholar]
  • 48.Fong TM, Yu H, Strader CD. Molecular basis for the species selectivity of the neurokinin-1 receptor antagonists CP-96,345 and RP67580. J. Biol. Chem. 1992;267:25668–25671. [PubMed] [Google Scholar]
  • 49.Zvyaga TA, Mm KC, Beck M, Sakmar TP. Movement of the retinylidene Schiffbase counterion in rhodopsin by one helix turn reverses the pH dependence ofthe metarhodopsin I to metarhodopsin II transition. J. Biol. Chem. 1993;268:4661–4667. [PubMed] [Google Scholar]
  • 50.Strader CD, Candelore MR, Hill WS, Sigal IS, Dixon RAF. Identification of two serine residues involved in agonist activation of the 13-adrenergic receptor. J. Biol. Chem. 1989;264:13572–13578. [PubMed] [Google Scholar]
  • 51.Savarese TM, Wang C-D, Fraser CM. Site-directed mutagenesis of the rat ml muscarinic receptor: role of conserved cysteines in receptor function. J. Biol. Chem. 1992;267:11439–11448. [PubMed] [Google Scholar]
  • 52.Latronico AC, Anasti J, Arnhold IJP, Rapaport R, Mendonca BB, Bloise W, Castro M, Tsigos C, Chrousos GP. Testicular and ovarian resistance to luteinizing-hormone caused by inactivating mutations of the luteinizing hormone-receptor gene. N. EngI. J. Med. 1996;334:507–512. doi: 10.1056/NEJM199602223340805. [DOI] [PubMed] [Google Scholar]

RESOURCES