Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2006 Feb 1.
Published in final edited form as: Bioorg Med Chem Lett. 2005 Nov 10;16(3):596–601. doi: 10.1016/j.bmcl.2005.10.054

Conversion of A3 Adenosine Receptor Agonists into Selective Antagonists by Modification of the 5′-Ribofuran-uronamide Moiety

Zhan-Guo Gao a, Bhalchandra V Joshi a, Athena Klutz a, Soo-Kyung Kim a, Hyuk Woo Lee b, Hea Ok Kim b, Lak Shin Jeong b, Kenneth A Jacobson a,*
PMCID: PMC1351218  NIHMSID: NIHMS7067  PMID: 16289820

Abstract

The highly selective agonists of the A3 adenosine receptor (AR), Cl-IB-MECA (2-chloro-N6-(3-iodobenzyl)-5′-N-methylcarboxamidoadenosine) and its 4′-thio analogue, were successfully converted into selective antagonists simply by appending a second N-methyl group on the 5′-uronamide position. The 2-chloro-5′-(N,N-dimethyl)uronamido analogues bound to, but did not activate the human A3AR, with Ki values of 29 nM (4′-O) and 15 (4′-S) nM, showing >100-fold selectivity over A1, A2A, and A2BARs. Competitive antagonism was demonstrated by Schild analysis. The 2-(dimethylamino)-5′-(N,N-dimethyl)uronamido substitution also retained A3AR selectivity but lowered affinity.

Keywords: nucleoside, G protein-coupled receptor, adenylyl cyclase, molecular modeling, radioligand binding

ABBREVIATIONS: AR, adenosine receptor; CGS21680, 2-[p-(2-carboxyethyl)phenylethylamino]-5′-N-ethylcarboxamido-adenosine; CHO, Chinese hamster ovary; Cl-IB-MECA, 2-chloro-N6-(3-iodobenzyl)-5′-N-methylcarboxamidoadenosine; CPA, N6-cyclopentyladenosine; DMEM, Dulbecco’s modified Eagle’s medium; I-AB-MECA, N6-(4-amino-3-iodobenzyl)-5′-N-methylcarboxamidoadenosine; NECA, 5′-N-ethylcarboxamidoadenosine; PIA, N6-(phenylisopropyl)adenosine; PTLC, preparative thin layer chromatography


Antagonists of the A3 adenosine receptor (AR) are of potential use for clinical targets, including the treatment of glaucoma, allergic conditions, and inflammation.1 Potent and selective antagonists for the human A3AR have recently been synthesized.26 These human A3AR antagonists were found to be weak or ineffective at the rat A3AR7,8 and were unsuitable for evaluation in small animal models or for further development as drugs. Thus, A3AR antagonists of which the affinity and selectivity are independent of species are sought as drug candidates. In previous studies, it was found that antagonists derived from adenosine analogs, in contrast to nonpurine heterocyclic antagonists, could be species-independent, potent, and selective A3AR antagonists.9

Among the four subtypes of receptors for adenosine 1, the intrinsic efficacy of nucleosides acting at the A3 subtype is known to be especially sensitive to structural changes. For example, substitution by a benzyl or a 3-iodobenzyl group at the N6 position of adenosine was demonstrated to increase affinity but decrease efficacy at the human A3AR, in the inhibition of forskolin-stimulated adenylyl cyclase.911 Additional substitution by a 2-Cl substituent further increased affinity and decreased efficacy.9,10 As a result, N6-(3-iodobenzyl)adenosine 2 was a partial agonist achieving only 46% of the maximal effect, while 2-chloro-N6-(3-iodobenzyl)adenosine 3 was a potent antagonist for the A3AR, albeit nonselective. A 5′-methyluronamide substitution of 3 restored its efficacy and rendered it selective for the A3AR. Thus, Cl-IB-MECA 4 and its 4′-thio analogue LJ568 5 are selective agonists for the A3AR.9,12,13 Here, we report that 4 and 5 were successfully converted into selective antagonists by appending an additional methyl group on the 5′-uronamide nitrogen. The finding that removing H-binding ability in the region of the 5′-uronamide lowers efficacy is consistent with expectations derived from rhodopsin-based, dynamic molecular modeling and ligand docking.14

Synthetic routes to N,N-dimethylamide derivatives 68 are shown in Scheme 1. The 5′-ester group was aminolyzed with dimethylamine, which also led to a side product substituted at the 2-position.15 The 4′-S analogue was prepared from an acetoxy intermediate,16 similar to a reported series of derivatives.12

Scheme 1a.

Scheme 1a

aReagents & conditions: a) K2CO3, MeOH, r.t.; b) HOAc; c) 40% aq. Me2NH, r.t.; d) Me2NH, EDC, HOBT, DIPEA; e) TBAF, THF.

Binding assays were carried out using standard radioligands in Chinese hamster ovary (CHO) cells stably expressing a human AR subtype.17 The binding affinity at the human A3AR of the 2-Cl antagonist derivatives 6 and 7 was shown to be 29 and 15 nM, respectively (Table 1). Thus, the compounds were demonstrated to be over 100-fold selective in binding to the human A3AR in comparison to other AR subtypes. In addition, the 2-(dimethylamino) substitution in 8 resulted in A3AR selectivity but with lower affinity.

Table 1.

Potency of a series of adenosine derivatives for four subtypes of human ARs and the efficacy at the A3AR.

Potency (Ki or EC50, nM) Efficacyc
Compound A1d A2Ad A2Be A3d A3
1a 310e 700e 24,000 290e 100
2b 7.4 ± 1.7 132 ± 22 ~10,000 5.8 ± 0.4 46 ± 8
3b 16.8 ± 2.2 197 ± 34 >10,000 1.8 ± 0.1 0
4b 222 ± 22 5360 ± 2470 >10,000 1.4 ± 0.3 100
5f 193 ± 46 223 ± 36 ND 0.38± 0.07 114 ± 9
6g 5870 ± 930 >10,000 >10,000 29.0 ± 4.9 0
7g 6220 ± 640 >10,000 >10,000 15.5 ± 3.1 0
8 >10,000 >10,000 >10,000 315 ± 19 0

All experiments were done on CHO cells stably expressing one of four subtypes of human ARs. The binding affinity for A1, A2A and A3ARs was expressed as Ki values and was determined by using agonist radioligands ([3H]R-PIA), ([3H]CGS21680), [125I]I-AB-MECA, respectively. The potency at the A2BAR was expressed as EC50 values and was determined by stimulation of cAMP production in AR-transfected CHO cells. The efficacy at A3ARs was determined by inhibition of forskolin-stimulated cAMP production in AR-transfected CHO cells, as described in the text. Data are expressed as mean ± standard error.

a

Values from reference 21.

b

Values from references 9, 21.

c

At a concentration of 10 μM, in comparison to the maximal effect of a full agonist NECA at 10 μM.

d

Ki in binding, unless noted.

e

cAMP assay.

f

Ki values from reference 13.

g

6, MRS3771; 7, LJ-1256.

ND not determined.

To probe species differences, the affinity of 6 and 7 was also measured at the rat A3AR. Although the affinity decreased with respect to the affinity at the human A3AR, these two compounds showed moderate affinity at the rat A3AR. The Ki values were 286 ± 31 and 321 ± 74 nM for 6 and 7, respectively.

In functional assays consisting of measuring inhibition of forskolin-stimulated production of 3′,5′-cyclic-adenosine monophosphate (cAMP) in intact CHO cells heterologously expressing ARs,18 single concentration determinations (Table 1) indicated that A3AR agonism was absent in compounds 68. In contrast, the concentration-response curve for compound 4 indicated full agonism, as previously reported, with an EC50 of 1.2 nM at the human A3AR (Figure 1A). A Schild analysis19 indicated that 6 concentration-dependently antagonized the A3 agonist 4 to inhibit forskolin-stimulated cAMP accumulation in CHO cells stably expressing the human A3AR with a KB value of 48 nM (Figure 1B,C). In contrast, 6, at a concentration of 1 μM, had no significant effect on cAMP accumulation inhibited by the A1 agonist CPA (N6-cyclopentyladenosine) in CHO cells expressing the human A1AR under the similar conditions (data not shown).

Figure 1.

Figure 1

A) Inhibition by nucleoside derivatives of forskolin-stimulated cAMP accumulation in CHO cells stably expressing the human A3AR. Cl-IB-MECA 4 concentration-dependently inhibited with an EC750 of 1.2±0.3 nM (n=3), while 6 and 7 alone had no effect. B) Right-shift of the concentration-response curves and C) Schild analysis for the concentration-dependent inhibition by 6 of the effect of Cl-IB-MECA 4 on forskolin-stimulated cAMP accumulation in CHO cells stably expressing the human A3AR. D) A3AR-induced changes in intracellular [Ca2+] in A3AR-expressing CHO cells (proportional to relative fluorescence units, RFU) upon activation by nucleosides. In control CHO cells there was no response to NECA.

Similar results were obtained in an assay of A3AR-induced changes in intracellular [Ca2+].20 The nonselective AR agonist NECA (5′-N-ethyl-carboxamidoadenosine) activated the human A3AR expressed in CHO cells to induce a concentration-dependent rise in intracellular [Ca2+] with a potency corresponding to an EC50 of 58 ± 16 nM (n=4), while 6 and 7 alone did not induce calcium transients (Figure 1D). Compound 8 also did not induce a Ca2+ response (data not shown). Increasing concentrations of 6 right-shifted the activation curves for NECA (not shown), and a KB value for 6 of 20 nM was determined. The A3AR antagonist 7 similarly shifted the NECA-induced Ca2+ response to the right (not shown).

In this study, two selective A3AR agonists, Cl-IB-MECA and its 4′-thio analogue, have been successfully transformed into antagonists selective for the A3AR by appending an additional N-methyl group on the 5′-uronamide position. Other related N,N-dialkyl substitution in the 4′-thio nucleoside series of adenosine agonists had more complex changes in affinity and efficacy.12 Thus, it appears that the 5′-(N,N-dimethyl)uronamido group especially tends to preserve affinity and selectivity in N6-3-iodobenzyladenosine derivatives, while entirely abolishing activation of the human A3AR. There was also an interdependence of the effects of an N,N-dimethyl moiety and the N6-substituent, since the N6-methyl analogue corresponding to 7 was a partial agonist.12 A further advantage of the N6-benzyl substitution of 6 and 7 over many other smaller or larger groups is that the high A3AR affinity tends to be retained in murine species,21 which was confirmed in the present study.

Molecular modeling of the A3AR indicated that flexibility of the adenosine derivative in the 5′-region correlated with putative conformational changes of the receptor associated with activation. Although there is no global conformational model of the activated state of the receptor, local conformational changes have been proposed. One such change is the rotation, anticlockwise from the extracellular perspective of the receptor, of the conserved Trp243 in TM6.9,21 We have proposed that both flexibility of the 5′-uronamide and its ability to make and break multiple H-bonds as this conformational change occurs are needed for receptor activation. The low efficacy of 5′-thioether derivatives is also consistent with the need for H-bonding in this region in order to activate the A3AR.22

The present findings are consistent, in that we have removed the H-bond-donating ability of the 5′-uronamide with a relatively subtle structural alteration, resulting in the loss of ability to activate the A3AR. From a comparison of the A2A and A3AR models,23 the main difference in the docking complex was the preference of the χ angle; the A3AR preferred nucleosides bound in an anti-form, while the A2AAR preferred a high-anti conformation (approximately −70°) about the glycosidic bond. Another difference was the binding at the 5′-position. In the A2AAR, the 5′-NH formed a H-bond with an important T88 residue in TM3, but in the model of the A3AR complex, this was replaced by a stronger interaction proposed between the 5′-carbonyl group and S7.42. Accordingly, the 5′-dimethylamides displayed dramatically diminished binding affinity at the A2AAR.

A3AR antagonists are potentially useful therapeutically for a number of disorders.1,24,25 However, the A3 antagonists have not been widely used in animal models due to their extremely weak potency in murine species.8 Thus, A3AR antagonists independent of species are of high priority to be developed. For this reason, a nucleoside analog IB-MECA (2-H analogue of 4) was converted to an antagonist by cyclization of the 5′-uronamide moiety into a spirolactam, resulting in a selective A3AR antagonist of moderate affinity for both human and rat A3ARs.9 Another advantage of nucleoside A3AR antagonists over other heterocycles7 is increased aqueous compatibility. For example, the clog p values of 6 and 7 are 1.69 and 1.73, respectively, in comparison to 6.86 for MRS1191, a dihydropyridine antagonist of the A3AR.7 Here, two additional selective antagonists with reasonable affinity for both human and rat A3ARs were introduced, which increased the diversity of rat A3AR antagonists.

It is clear that more potent and selective antagonists are needed for eventual therapeutic purposes. The newly synthesized A3AR antagonists could be evaluated in models of a number of disorders related to the A3AR, such as glaucoma and inflammation. Currently, existing nucleoside analogs should be good templates for further modification and development of potent and selective antagonists for the A3ARs in diverse species.

Chart 1.

Chart 1

Acknowledgments

B.V. Joshi thanks Gilead Sciences (Foster City, CA) for financial support. This research was supported in part by the Intramural Research Program of the NIH, National Institute of Diabetes and Digestive and Kidney Diseases and by the Korea Health R & D Project, Ministry of Health & Welfare, Korea (03-PJ2-PG4-BD02-0001). We thank Prof. Mortimer Civan (University of Pennsylvania) for helpful discussion.

References

  • 1.Fredholm BB, IJzerman AP, Jacobson KA, Klotz KN, Linden J. Pharmacol Rev. 2001;53:527. [PMC free article] [PubMed] [Google Scholar]
  • 2.Kim YC, Ji XD, Jacobson KA. J Med Chem. 1996;39:4142. doi: 10.1021/jm960482i. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.van Muijlwijk-Koezen JE, Timmerman H, Link R, van der Goot H, IJzerman AP. J Med Chem. 1998;41:3994. doi: 10.1021/jm980037i. [DOI] [PubMed] [Google Scholar]
  • 4.Baraldi PG, Cacciari B, Moro S, Spalluto G, Pastorin G, Da Ros T, Klotz KN, Varani K, Gessi S, Borea PA. J Med Chem. 2002;45:770. doi: 10.1021/jm0109614. [DOI] [PubMed] [Google Scholar]
  • 5.Maconi A, Pastorin G, Da Ros T, Spalluto G, Gao ZG, Jacobson KA, Baraldi PG, Cacciari B, Varani K, Moro S, Borea PA. J Med Chem. 2002;45:3579. doi: 10.1021/jm020974x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Okamura T, Kurogi Y, Hashimoto K, Sato S, Nishikawa H, Kiryu K, Nagao Y. Bioorg Med Chem Lett. 2004;14:3775. doi: 10.1016/j.bmcl.2004.04.099. [DOI] [PubMed] [Google Scholar]
  • 7.Moro, S.; Spalluto, G.; Gao, Z.G.; Jacobson, K.A. Med. Res. Rev. in press. [DOI] [PMC free article] [PubMed]
  • 8.Yang H, Avila MY, Peterson-Yantorno K, Coca-Prados M, Stone RA, Jacobson KA, Civan MM. Current Eye Res. 2005;30:747–754. doi: 10.1080/02713680590953147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Gao ZG, Kim SK, Biadatti T, Chen W, Lee K, Barak D, Kim SG, Johnson CR, Jacobson KA. J Med Chem. 2002;45:4471. doi: 10.1021/jm020211+. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Gao ZG, Blaustein JB, Gross AS, Melman N, Jacobson KA. Biochem Pharmacol. 2003;65:1675. doi: 10.1016/s0006-2952(03)00153-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Gao ZG, Jeong LS, Moon HR, Kim HO, Choi WJ, Shin DH, Elhalem E, Comin MJ, Melman N, Mamedova L, Gross AS, Rodriguez JB, Jacobson KA. Biochem Pharmacol. 2004;67:893. doi: 10.1016/j.bcp.2003.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Jeong, L.S.; Lee, H.W.; Jacobson, K.A.; Kim, H.O.; Shin, D.H.; Lee, J.A.; Gao, Z.G.; Lu, C.; Duong, H.T.; Gunaga, P.; Lee, S.K.; Jin, D.Z.; Chun, M.W.; Moon, H.R. J. Med. Chem. in press.
  • 13.Jeong LS, Jin DZ, Kim HO, Shin DH, Moon HR, Gunaga P, Chun MW, Kim YC, Melman N, Gao ZG, Jacobson KA. J Med Chem. 2003;46:3775. doi: 10.1021/jm034098e. [DOI] [PubMed] [Google Scholar]
  • 14.Jacobson KA, Kim SK, Costanzi S, Gao ZG. Molecular Interventions. 2004;4:337. doi: 10.1124/mi.4.6.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.(2S,3S,4R,5R)-5-[2-Chloro-6-(3-iodo-benzylamino)-purin-9-yl]-3,4-dihydroxy-tetrahydrofuran-2-carboxylic acid dimethylamide (6). The methyl ester 10 (0.031 g, 0.05 mmol) was dissolved in MeOH (5 mL) and potassium carbonate (0.014 g, 0.1 mmol) was added and the mixture stirred at room temperature for 10 min. Acetic acid (0.2 mL) was added to neutralize the base and the resulting diol was treated in situ with aqueous dimethylamine (0.5 mL, 40%) and further stirred for 1 hr. The reaction mixture was concentrated under reduced pressure and subjected to preparative thin layer chromatography by using chloroform/methanol (9:1) as solvent to afford the dimethylamide 6 as a colorless solid (0.0072 g, 26%).1H NMR (CDCl3) δ2.97 (s, 3 H, N-CH3), 3.14 (s, 3 H, N-CH3), 3.52-3.80 (m, 4 H), 4.45-4.85 (m, 3 H), 5.05 (d, 1 H, J = 5.5 Hz), 6.06 (d, 1 H, J = 5.1 Hz), 7.17 (t, 1 H, J = 7.8 Hz, 5′-H), 7.65-7.82 (m, 3 H), 8.21 (s, 1 H); TOFMS m/z 559.0353 (M + H+) (calculated for C19H21N6O4ClI+) 559.0358. (2S,3S,4R,5R)-5-[2-dimethylamino-6-(3-iodo-benzylamino)-purin-9-yl]-3,4-dihydroxy-tetrahydrofuran-2-carboxylic acid dimethylamide (8). Aqueous dimethylamine (0.5 mL, 40%) was added to methyl ester 10 (0.016 g, 0.025 mmol) and the resulting reaction mixture was stirred at room temperature for 6 h. The mixture was concentrated under reduced pressure and purified by preparative thin layer chromatography by using chloroform/methanol (9:1) as solvent to afford the dimethylamide 8 as a colorless solid (0.0058 g, 42%). 1H NMR (CDCl3) δ2.90-3.23 (m, 12 H), 3.79 (bs, 3 H), 4.42-4.95 (m, 4 H), 6.11 (d, 1 H, J = 5.1 Hz), 6.22 (bs, 1 H) 7.14 (t, 1 H, J = 7.5 Hz, 5′-H), 7.40 (d, 1 H, J = 7.6 Hz, 6′-H), 7.65 (d, 1 H, J = 7.8 Hz, 4′-H), 7.78 (s, 1 H, 2′-H), 8.22 (s, 1 H, H-8); TOFMS m/z 568.1162 (M + H+) (calculated for C21H27N7O4I+) 568.1169.
  • 16.(2S,3S,4R,5R)-5-[2-Chloro-6-(3-iodo-benzylamino)-purin-9-yl]-3,4-dihydroxy-tetrahydro-thiophene-2-carboxylic acid dimethylamide (7). To a solution of 12 (483.0 mg, 0.622 mmol), N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC, 179 mg, 0.933mmol), 1-hydroxybenzotriazole (HOBt, 126 mg, 0.933 mmol), and dimethylamine.HCl (76 mg, 0.933 mmol) in CH2Cl2 (20 mL) was added N,N-diisopropylethylamine (DIPEA, 0.325 mL, 1.87 mmol) and the mixture was stirred at room temperature for 12 h. The reaction mixture was evaporated and the residue was purified by a silica gel column chromatography (hexane/EtOAc = 10:1-5:1) to give the silyl-protected amide intermediate as a white foam. To a stirred solution of the silyl amide (414 mg, 0.516 mmol) in THF (10 mL) was added tetrabutylammonium fluoride (1.29 mL, 1.29 mmol, 1 M THF solution) and the reaction mixture was stirred at room temperature for 1 h. The solvent was evaporated and the resulting residue was purified by silica gel column chromatography (CH2Cl2/MeOH = 10:1) to give 7 (214 mg, 64%): white solid; mp 186.1-186.3 °C; [α]20D -12.4 (c 0.10, MeOH); UV (MeOH) λmax 274 nm (pH 7); 1H NMR (DMSO-d6) δ 2.95 (s, 3 H, N-CH3), 3.04(s, 3H, N-CH3), 4.30 (d, 1 H, J = 4.6 Hz, 2-H), 4.52 (br dd, 1 H, J = 4.6, 8.4 Hz, 3-H), 4.58 (m, 1 H, 4-H), 4.65 (d, 2 H, J = 5.7 Hz, N-CH2), 5.59 (d, 1 H, J = 5.5 Hz, exchangeable with D2O, OH), 5.86 (d, 1 H, J = 5.1 Hz, exchangeable with D2O, OH), 5.91 (d, 1 H, J = 5.4 Hz, 5-H), 7.17 (t, 1 H, J = 7.8 Hz, 5′-H), 7.40 (d, 1 H, J = 7.6 Hz, 6′-H), 7.65 (d, 1 H, J = 7.8 Hz, 4′-H), 7.78 (s, 1 H, 2′-H), 8.52 (s, 1 H, H-8), 9.00 (br t, 1 H, J = 6.1 Hz, exchangeable with D2O, NH); 13C NMR (CD3OD) δ36.6, 38.2, 44.5, 64.6, 77.2, 80.7, 95.0, 119.7, 128.3, 131.5, 137.5, 138.0, 141.5, 142.8, 145.2, 151.7, 155.8, 156.5, 172.7; FAB-MS m/z 575 (M++1); Anal. (C19H20ClIN6O3S) C, H, N, S.
  • 17.The CHO cells stably expressing recombinant ARs were cultured in DMEM and F12 (1:1) supplemented with 10% fetal bovine serum, 100 units/ml penicillin, 100 μg/mL streptomycin, 2 μmol/ml glutamine and 800 μg/ml geneticin. After harvest and homogenization, cells were centrifuged at 500 g for 10 min, and the pellet was re-suspended in 50 mM Tris·HCl buffer (pH 7.4) containing 10 mM MgCl2, 1 mM EDTA. The suspension was homogenized with an electric homogenizer for 10 sec, and was then re-centrifuged at 20,000 g for 20 min at 4 °C. The resultant pellets were resuspended in buffer in the presence of 3 units/ml adenosine deaminase, and the suspension was stored at −80°C until the binding experiments. The protein concentration was measured as described [Bradford, M.M. Anal. Biochem. 1976, 72, 248]. For A3AR binding assays, each tube contained 100 μl of membrane suspension, 50 μl of [125I]I-AB-MECA (final concentration 0.5 nM), and 50 μl of increasing concentrations of compounds in Tris·HCl buffer (50 mM, pH 7.4) containing 10 mM MgCl2 Nonspecific binding was determined using 10 μM NECA. The mixtures were incubated at 25°C for 60 min. Binding reactions were terminated by filtration through Whatman GF/B filters under reduced pressure using a MT-24 cell harvester (Brandell, Gaithersburg, MD). Filters were washed three times with ice-cold buffer. Radioactivity was determined in a Beckman 5500B γ-counter. The binding of [3H]R-PIA to A1ARs and the binding of [3H]CGS21680 to A2AARs were as previously described.10 IC50 values were converted to Ki values as described [Cheng Y-C, Prusoff WH. Biochem. Pharmacol. 1973, 22, 3099.]. [125I]N6-(4-amino-3-iodo-benzyl)adenosine-5′-N-methyluronamide ([125I]I-AB-MECA; 2000 Ci/mmol), [3H]R-PIA (R-N6-[phenylisopropyl]adenosine, 34 Ci/mmol), [3H]CGS21680 (2-[p-(2-carboxyethyl)phenylethylamino]-5′-N-ethylcarboxamido-adenosine, 47 Ci/mmol) and [3H]cAMP (40 Ci/mmol) were from Amersham Pharmacia Biotech (Buckinghamshire, UK). NECA, CGS21680, CPA, and R-PIA were purchased from Sigma-RBI (St. Louis, MO). Other chemicals were from standard commercial sources and of analytical grade.
  • 18.Intracellular cAMP levels were measured with a competitive protein binding method [Nordstedt, C.; Fredholm, B.B., Anal. Biochem. 1990, 189, 231]. CHO cells expressing one of four subtypes of recombinant ARs were harvested by trypsinization. After resupension in medium, cells were planted in 24-well plates in 0.5 ml medium. After 24 hr, the medium was removed and cells were washed three times with 0.5 ml DMEM, containing 50 mM HEPES, pH 7.4. Cells were then treated with agonists and/or test compounds in the presence of rolipram (10 μM) and adenosine deaminase (3 units/mL). After 45 min forskolin (10 μM) was added to the medium, and incubation was continued an additional 15 min. The reaction was terminated by removing the medium, and cells were lysed upon the addition of 200 μL of 0.1 M ice-cold HCl. The cell lysate was resuspended and stored at −20°C. For determination of cAMP production, protein kinase A was incubated with [3H]cAMP (2 nM) in K2HPO4/EDTA buffer (K2HPO4, 150 mM; EDTA, 10 mM), 20 μL of the cell lysate, and 30 μL 0.1 M HCl or 50 μL of cAMP solution (0-16 pmol/200 μL for standard curve). Bound radioactivity was separated by rapid filtration through Whatman GF/C filters and washed once with cold buffer. Bound radioactivity was measured by liquid scintillation spectrometry.
  • 19.Arunlakshana O, Schild HO. Br J Pharmacol Chemother. 1959;14:48. doi: 10.1111/j.1476-5381.1959.tb00928.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.CHO cells stably expressing human A3ARs were grown overnight in 100 μl of media in 96 well flat bottom plates at 37°C at 5% CO2 to reach approx. 90% confluency. The calcium assay kit (Molecular Devices) was used as directed with no washing of cells, and with probenecid added to the loading dye at a final concentration of 2.5 mM to increase dye retention. Cells were loaded with 50 μl of dye with probenecid to each well and incubated for 60 minutes at room temperature. The compound plate was prepared using dilutions of various compounds in Hanks Buffer. For antagonist studies, both agonist and antagonist were added to the sample plate. Samples were run in duplicate using a Molecular Devices Flexstation I at room temperature. Cell fluorescence (Excitation = 485 nm; Emission = 525 nm) was monitored following exposure to compound. Increases in intracellular calcium are reported as the maximum fluorescence value after exposure minus the basal fluorescence value before exposure.
  • 21.Tchilibon S, Joshi BV, Kim SK, Duong HT, Gao ZG, Jacobson KA. J Med Chem. 2005;48:1745. doi: 10.1021/jm049580r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.van Tilburg EW, von Frijtag Drabbe Künzel J, de Groote M, IJzerman AP. J Med Chem. 2002;45:420. doi: 10.1021/jm010952v. [DOI] [PubMed] [Google Scholar]
  • 23.Kim SK, Gao ZG, Van Rompaey P, Gross AS, Chen A, Van Calenbergh S, Jacobson KA. J Med Chem. 2003;46:4847. doi: 10.1021/jm0300431. [DOI] [PubMed] [Google Scholar]
  • 24.Fredholm BB, Irenius E, Kull B, Schulte G. Biochem Pharmacol. 2001;61:443. doi: 10.1016/s0006-2952(00)00570-0. [DOI] [PubMed] [Google Scholar]
  • 25.Kim HO, Hawes C, Towers P, Jacobson KA. J Labelled Comp Radiopharm. 1996;38:547. doi: 10.1002/(SICI)1099-1344(199606)38:6<547::AID-JLCR870>3.0.CO;2-Y. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES