Toward Multivalent Signaling across G Protein-Coupled Receptors

Stijn F. M. van Dongen , Hans-Peter M. de Hoog , Ruud J. R. W. Peters ... Kim , Béatrice Hechler , Zhan-Guo Gao , Christian Gachet and Kenneth A. Jac...
0 downloads 0 Views 274KB Size
406

Bioconjugate Chem. 2008, 19, 406–411

Toward Multivalent Signaling across G Protein-Coupled Receptors from Poly(amidoamine) Dendrimers Yoonkyung Kim,† Béatrice Hechler,‡ Athena M. Klutz,† Christian Gachet,‡ and Kenneth A. Jacobson*,† Molecular Recognition Section, Laboratory of Bioorganic Chemistry, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, Maryland 20892, INSERM U.311, EFS-Alsace, and Université Louis Pasteur, Strasbourg, France. Received August 28, 2007; Revised Manuscript Received November 3, 2007

Activation of the A2A receptor, a G protein-coupled receptor (GPCR), by extracellular adenosine, is antiaggregatory in platelets and anti-inflammatory. Multiple copies of an A2A agonist, the nucleoside CGS21680, were coupled covalently to PAMAM dendrimers and characterized spectroscopically. A fluorescent PAMAM-CGS21680 conjugate 5 inhibited aggregation of washed human platelets and was internalized. We envision that our multivalent dendrimer conjugates may improve overall pharmacological profiles compared to the monovalent GPCR ligands.

Small-molecule ligands for a G protein-coupled receptor (GPCR) generally bind within its heptahelical transmembrane domain to trigger intracellular signaling pathways (1, 2). Accordingly, the surface contact of a distal structural unit of a bound ligand, away from the pharmacophore region, with a GPCR at/near its extracellular domain (i.e., a secondary interaction) may impart additional selectivity and affinity. An approach to designing ligands for GPCRs by applying structural modification at their permissive distal sites has led to the concept of functionalized congeners (3, 4). For instance, ligands for GPCRs may require attachment to a carrier molecule for delivery, and the chemical functional groups and geometry of a linker moiety can be varied, as guided by receptor affinity. Dendrimers (5–7) are treelike macromolecules which possess favorable characteristics: structural integrity, control of the component functional groups and their corresponding physical properties by chemical synthesis, feasibility to conjugate multiple functional units at the periphery and the interior, and a low enzymatic degradation rate. For the past 15 years, dendrimers have been used for biomedical research (8–12) including gene transfection (polycationic nature) (13–16), drug delivery (targeted/controlled release, encapsulation, or covalent/ electrostatic attachment) (17–21), protein-carbohydrate interactions (multivalent effect) (22–27), medical diagnostics (signal amplification) (12, 28), and tissue engineering (29–31). Poly(amidoamine) (PAMAM) dendrimer is composed of aliphatic amino and amido groups, and its relatively biocompatible nature has found many applications in biomedicine (32–35). Although the heterogeneity of commercial PAMAM dendrimers may hinder the structural characterization at the monomolecular level (36), usage of PAMAM dendrimers may provide quick and easy access to gauge the applicability of dendrimers as advanced therapeutic agents before progressing to tailor-made dendrimers. Here, we report the first example of dendrimer applications to induce the multivalent intracellular signal transduction across GPCRs. Adenosine receptors (ARs) are members of the Group A GPCR family that are involved in various disease states such as inflammation, cancer, cardiovascular damage, and nervous system disorders (37–39). Four subtypes of ARs, A1, A2A, A2B, and A3, have been identified to date, with each manifesting a * Author to whom correspondence should be addressed. Phone: 301496-9024, fax: 301-480-8422; e-mail: [email protected]. † National Institutes of Health. ‡ INSERM U.311, EFS-Alsace, and Université Louis Pasteur.

unique pharmacological profile, tissue distribution, and effector coupling. To test the feasibility of our approach, the carboxylic acid derivative CGS216801(1 (40–42)), an A2A AR agonist, was selected as a ligand for the attachment to the peripheral amino groups of a third generation (G3) PAMAM dendrimer with an ethylenediamine core (2). Here, the PAMAM dendrimer linked to CGS21680 through its aliphatic amino end group can be considered a part of the distal modification under the functionalized congener concept to modulate the biological activity of the original ligand (3, 4). Our plan for the conjugation of CGS21680 to the PAMAM dendrimer was to utilize its distal carboxylic acid group extending from the C-2 position of the adenine base, which was shown to be a site suitable for attachment that is tolerated in receptor binding (3). The synthetic feasibility to make PAMAMCGS21680 conjugates through a peptide coupling method was first examined at a monomeric level using CGS21680 in its unprotected form. For peptide coupling, PyBOP1 was adopted as a coupling reagent. PyBOP generally allows rapid condensation in high yield unless a significant steric challenge is involved. Unlike a somewhat sluggish and unselective carbodiimidemediated coupling, which acts essentially through dehydration (i.e., forming urea as a byproduct), PyBOP does not generate any esterified products between a free hydroxyl and a carboxylic acid. Thus, CGS21680 1 was treated with an excess of base, followed by an equimolar amount of N-Boc-ethylenediamine1 in DMF1 (Scheme 1). PyBOP (0.8 equiv) was added to this mixture to form an activated intermediate in situ and to achieve the coupling exclusively at the primary aliphatic amine. Thus, the preactivation step of the carboxylic acid group commonly applied before the addition of amine was avoided in order to minimize the formation of intra- or intermolecular cross-coupled products involving the adenine N6amine. Indeed, neither the 1H NMR nor the mass spectrum of the crude mixture showed the formation of any unwanted side products, and this reaction produced the desired compound 3 in 90% isolated yield. 1 Abbreviations: CGS21680, 2-[4-(2-carboxylethyl)phenylethylamino]5′-N-ethylcarboxamidoadenosine; PyBOP, benzotriazol-1-yl-oxytripyrrolidinophosphonium hexafluorophosphate; Boc, tert-butoxycarbonyl; DIEA, N,N-diisopropylethylamine; Et3N, triethylamine; DMF, N,Ndimethylformamide; DMSO, dimethyl sulfoxide; DHB, 2,5-dihydroxybenzoic acid; THAP, 2,4,6-trihydroxyacetophene; ADP, adenosine 5′diphosphate.

10.1021/bc700327u CCC: $40.75  2008 American Chemical Society Published on Web 01/05/2008

Communications

Bioconjugate Chem., Vol. 19, No. 2, 2008 407

Scheme 1

Scheme 2

Next, this PyBOP coupling strategy was applied to the PAMAM dendrimer. We began by preparing the dendrimer conjugate 4, which was fully amide-coupled with CGS21680 (Scheme 2). A slight excess of CGS21680 was added to a basic solution of PAMAM 2, which was then treated with 32 equiv of PyBOP as a DMSO-d61 solution. Deuterated DMSO was used as the reaction solvent in order to avoid later complications in the NMR analysis based on integration in DMSO-d6. The reaction continued for 22 h, and the product was purified by size exclusion chromatography (SEC) in DMF to give the desired compound 4 in quantitative yield as confirmed by 1H NMR. Here, the peak corresponding to the methylene R- to the carbonyl at the CGS21680 linker terminus was well-resolved at 2.33 ppm and exhibited the maximal shift upon coupling, similar to the result observed for the monomeric analogue 3. The integration value of a methyl peak from CGS21680 at 0.95 ppm (96.3 H) relative to a PAMAM methylene peak “c” at 2.18 ppm (120 H)san internal standard of PAMAM G3 for characterization by integrationsindicated that the reaction proceeded completely to produce the desired product 4 bearing 32 CGS21680 groups. To aid in the microscopic visualization in biological assays, a fluorescent dendrimer derivative 5 was then prepared. We chose Alexa Fluor 488 as a fluorophore, which has similar

absorption/emission maxima to those of fluorescein, but is reported to have a higher photostability, pH insensitivity, and good water solubility (43, 44). Thus, an equimolar amount of Alexa Fluor 488 was added to PAMAM dendrimer 2 as a 5-carboxy tetrafluorophenyl ester 6 in DMSO-d6 (Scheme 2). A small amount of red precipitate was observed almost instantly upon addition of 6. However, removal of this precipitate by filtration did not affect the recovery of an anticipated mass balance when calculated after the full conjugation of the ligands in the next step to prepare 5. Analysis of the fluorescent intermediate 7 by 1H NMR integration was difficult, and thus, the conjugation of 1 equiv of Alexa Fluor 488 was assumed on the basis of the stoichiometry of addition. The filtered crude mixture 7 was used for the next step without any purification. Subsequent attachment of CGS21680 to 7 was carried out using the same PyBOP coupling strategy as for 4, but by adding 31 equiv of PyBOP instead. The fluorescent PAMAM-GS21680 conjugate 5 was purified by SEC and was similarly characterized by 1H NMR integration to confirm that, on average, 31 CGS21680 groups were attached to the dendrimer (Supporting Information). PAMAM-CGS21680 conjugates 4 and 5 were then characterized by matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) mass spectrometry (see Supporting Information).

408 Bioconjugate Chem., Vol. 19, No. 2, 2008

Communications

Table 1. Comparison of Molecular Weights and the Number of CGS21680 Groups Attached Determined by NMR and MALDI MS MALDI NMR

DHB matrix b

c

cmpd

no. of CGS21680

MW

Mn

2 4 5a

0 32 31

6909 22317 22352

5772 18839 19050

Mw

d

5909 19391 19502

e

THAP matrix f

c

PDI

no. of CGS21680

Mn

1.02 1.03 1.02

0 28 27

5956 19789 19576

Mw

d

6085 20449 20044

PDIe

no. of CGS21680f

1.02 1.03 1.02

0 30 28

a

Counterions of sulfate groups for Alexa Fluor 488 were assumed to be protons. b PAMAM dendrimer 2 used for calculation here was assumed to have 32 peripheral amino groups and no structural defects. c Number-average molar mass. d Weight-average molar mass. e Polydispersity index. f No matrix adduct was considered for the estimation.

Figure 1. Platelet aggregation studies at 37 °C in the presence or absence of dendrimer 5 (Supporting Information). A suspension of human washed platelets was incubated with DMSO (top), dendrimer 5 (middle), or CGS21680 1 (bottom) (each added as a DMSO solution), to which ADP was then added to induce the platelet aggregation. An increase in light transmission indicates increased aggregation. The final concentration of dendrimer 5 or CGS21680 1 in the media was 2 µM, and the total content of DMSO in the media was 0.4% (v/v) in all cases.

Various MALDI matrices were attempted; however, in all cases, the intensity of the desired peak was generally weak, and the peak shape was overly broad due to the heterogeneity and possible fragmentation of PAMAM under the applied MALDI condition. Among those attempted matrices, DHB1 and THAP1 gave the best results for analysis. A broad peak corresponding to the half-size of the desired compound 4 or 5 was detected, which may have originated from either a doubly charged species or the fragmentation of the desired compound at the middle of the ethylenediamine core as suggested previously (45). In comparison to the 1H NMR integration-based characterization, MALDI underestimated the average molecular weights (MWs)

for both 4 and 5 in either matrix when the broad peak region above the MW of a half-size dendrimer was considered for calculations (Table 1). MALDI of a commercial PAMAM dendrimer 2 obtained under similar conditions also underestimated the average MW by ca. 1000 Da. When the MALDIestimated average MW of PAMAM 2 was subtracted from those of 4 and 5, dendrimers 4 and 5 were found to have 28–30 and 27–28 CGS21680 attachments, respectively, thus a slight underestimation when compared to those calculated by NMR. The deviations in MALDI-estimated MWs from those determined by NMR integration may have resulted from the difference in the tendency to form multiple matrix-adducts

Communications

between 2, with many free amino groups, and the fully conjugated dendrimer 4 or 5. In either matrix, the polydispersity of PAMAM-CGS21680 conjugate 4 or 5 was relatively low, ca. 1.02–1.03, and was similar to the value obtained for PAMAM dendrimer 2. This result further supports that the conjugation of CGS21680 was nearly complete for both 4 and 5, by reacting at most of the available peripheral amino groups of the PAMAM dendrimer. Preliminary conformational studies of PAMAM-CGS21680 conjugate 4 were performed by nuclear Overhauser enhancement (NOE) experiments and computer modeling (see Supporting Information). PAMAM interiors were predicted to maintain a relatively compact structure by forming exchangeable multiple hydrogen bonds in water or methanol (32). Unlike water molecules, which may interact mostly with the charged ammonium ions at the surface of PAMAM, polar aprotic DMSO may be more likely to penetrate into the PAMAM interior to disrupt the hydrogen-bonded network. Indeed, NOE cross-peaks were observed between methylene groups of PAMAM (“b” and “c”) and the protons from the ribose moiety (i.e., methyl, H-1′, H-2′, H-3′, and H-4′) of CGS21680 in DMSO-d6 (Supporting Information), suggesting a possible internalization of the ligand into the PAMAM interior. In addition, the energy-minimized structure of 4 (HyperChem7.5.2, Amber force field) exhibited a relatively large volume in the interior, which is sufficiently spacious to accommodate several internalized ligands simultaneously (46). Eighteen hydrogen bonds were detected in the interior of the PAMAM-CGS21680 conjugate 4. The overall shape of 4 was somewhat ellipsoidal (32), with the diameter ranging ca. 65–85 Å. Next, our fluorescent PAMAM-CGS21680 conjugate 5 was subjected to in vitro functional assays. We specifically chose the platelet aggregatory system (47–49) to examine the potential pharmacological value of our dendrimer-based multivalent ligand carriers for GPCR signaling. CGS21680, an A2A AR agonist, is already known to display a potent antiaggregatory effect in human platelet preparations (48, 49). The degree of platelet aggregation was determined by aggregometry after addition of ADP1 to a washed human platelet suspension (50) containing either the PAMAM-CGS21680 dendrimer conjugate 5 or CGS21680 monomer 1 as a DMSO solution (2 µM) or the same volume of pure DMSO as a control (Figure 1) (Supporting Information). Our preliminary findings indicated that ADPinduced platelet aggregation, which occurs through the nucleotide binding to P2Y receptors on the platelet surface, was inhibited in the platelet suspension containing either dendrimer 5 or monomer 1 to a similar degree (i.e., 61% inhibition by 5; 57% inhibition by 1, both at 2 µM), whereas platelets treated with DMSO alone aggregated in response to ADP. The inhibitory effect was the same whether the dendrimer conjugate was incubated for 5 or 60 min. Interestingly, the platelet aggregation response in the presence of dendrimer 5 exhibited a slower onset in comparison to that of control or the platelet suspension treated with monovalent CGS21680. Moreover, the fluorescent dendrimer conjugate 5 was clearly observed inside the platelets under a fluorescent confocal microscope (Figure 2), when the platelet suspension was incubated with dendrimer 5 and washed to remove any free dendrimer 5 in the extracellular medium. Thus, the extracellularly administered dendrimer conjugate is effectively internalized in platelets. The internalization of intravenously injected macromolecules into platelet granules has been described (53). Future studies will be directed toward determining the structural features of the dendrimer conjugates that affect internalization. In summary, in an effort to develop highly potent agonists acting through multivalent signaling across GPCRs, PAMAM dendrimer conjugates 4 and 5 loaded with multiple copies of

Bioconjugate Chem., Vol. 19, No. 2, 2008 409

Figure 2. Fluorescence confocal microscopy of human washed platelets incubated with either (A) DMSO or (B-E) dendrimer 5 (Supporting Information). Incubation time: (B) 5 min, (C) 15 min, (D) 30 min, and (E) 60 min. The final concentration of dendrimer 5 in the media was 2 µM, and the total content of DMSO in the media was 0.4% (v/v) in all cases.

CGS21680 were synthesized using a peptide coupling method and purified by SEC. The average MWs and the degree of ligand attachments for dendrimers 4 and 5 were determined by 1H NMR integration and MALDI. Both analyses support the formation of nearly fully loaded dendrimers. Our preliminary functional assay clearly demonstrated a potent antiaggregatory effect of the PAMAM-CGS21680 dendrimer 5 on the platelet aggregation induced by ADP. In addition, the fluorescent dendrimer 5 was found to be internalized into the platelet as shown by the fluorescent confocal microscopy studies. The approach described here to develop multivalent ligands for GPCRs can be extended to create a dendrimer containing a set of heterogeneous agonists/antagonists, in order to induce combined synergistic biological effects via binding to different receptor subtypes simultaneously on a cellular membrane (51, 52). A detailed investigation on the pharmacological effects of our dendrimers on platelet aggregation will be reported in due course.

ACKNOWLEDGMENT This research was supported in part by the Intramural Research Program of the NIH, NIDDK. We thank Dr. Herman Yeh for the helpful advice on the NMR experiments. We are grateful to Drs. Ratna Dutta and Haijun Yao at the Mass Spectrometry Laboratory of the University of Illinois, for numerous attempts to obtain MALDI spectra of our PAMAM

410 Bioconjugate Chem., Vol. 19, No. 2, 2008

dendrimer derivatives. Y.K. thanks the Can-Fite Biopharma for financial support. Supporting Information Available: Experimental procedures, 1H NMR, NOESY, and MALDI spectra, and the energyminimized structure. This material is available free of charge via the Internet at http://pubs.acs.org/BC.

LITERATURE CITED (1) Landry, Y., Niederhoffer, N., Sick, E., and Gies, J. P. (2006) Heptahelical and other G-protein-coupled receptors (GPCRs) signaling. Curr. Med. Chem. 13, 51–63. (2) Ratnala, V. R., Kiihne, S. R., Buda, F., Leurs, R., de Groot, H. J., and DeGrip, W. J. (2007) Solid-state NMR evidence for a protonation switch in the binding pocket of the H1 receptor upon binding of the agonist histamine. J. Am. Chem. Soc. 129, 867–872. (3) Jacobson, K. A., and Daly, J. W. (1991) Purine functionalized congeners as molecular probes for adenosine receptors. Nucleosides Nucleotides 10, 1029–1038. (4) Jacobson, K. A., Kirk, K. L., Padgett, W. L., and Daly, J. W. (1985) Functionalized congeners of adenosine: preparation of analogues with high affinity for A1-adenosine receptors. J. Med. Chem. 28, 1341–1346. (5) Tomalia, D. A. (2005) Birth of a new macromolecular architecture: dendrimers as quantized building blocks for nanoscale synthetic polymer chemistry. Prog. Polym. Sci. 30, 294– 324. (6) Grayson, S. M., and Fréchet, J. M. J. (2001) Convergent dendrons and dendrimers: from synthesis to applications. Chem. ReV. 101, 3819–3867. (7) Zeng, F., and Zimmerman, S. C. (1997) Dendrimers in supramolecular chemistry: from molecular recognition to selfassembly. Chem. ReV. 97, 1681–l712. (8) Lee, C. C., MacKay, J. A., Fréchet, J. M. J., and Szoka, F. C. (2005) Designing dendrimers for biological applications. Nat. Biotechnol. 23, 1517–1526. (9) Svenson, S., and Tomalia, D. A. (2005) Dendrimers in biomedical applications-reflections on the field. AdV. Drug DeliVery ReV. 57, 2106–2129. (10) Shabat, D., Amir, R. J., Gopin, A., Pessah, N., and Shamis, M. (2004) Chemical adaptor systems. Chem. Eur. J. 10, 2626– 2634. (11) Boas, U., and Heegaard, P. M. H. (2004) Dendrimers in drug research. Chem. Soc. ReV. 33, 43–63. (12) Stiriba, S.-E., Frey, H., and Haag, R. (2002) Dendritic polymers in biomedical applications: from potential to clinical use in diagnostics and therapy. Angew. Chem., Int. Ed. 41, 1329– 1334. (13) Zhou, J., Wu, J., Hafdi, N., Behr, J.-P., Erbacher, P., and Peng, L. (2006) PAMAM dendrimers for efficient siRNA delivery and potent gene silencing. Chem. Commun. 2362–2364. (14) Luo, D., Haverstick, K., Belcheva, N., Han, E., and Saltzman, W. M. (2002) Poly(ethylene glycol)-conjugated PAMAM dendrimer for biocompatible, high-efficiency DNA delivery. Macromolecules 35, 3456–3462. (15) Tang, M. X., Redemann, C. T., and Szoka, F. C., Jr. (1996) In vitro gene delivery by degraded polyamidoamine dendrimers. Bioconjugate Chem. 7, 703–714. (16) Haensler, J., and Szoka, F. C., Jr. (1993) Polyamidoamine cascade polymers mediate efficient transfection of cells in culture. Bioconjugate Chem. 4, 372–379. (17) Gillies, E. R., and Fréchet, J. M. J. (2005) Dendrimers and dendritic polymers in drug delivery. Drug DiscoVery Today 10, 35–43. (18) D’Emanuele, A., and Attwood, D. (2005) Dendrimer-drug interactions. AdV. Drug DeliVery ReV. 57, 2147–2162. (19) Haag, R. (2004) Supramolecular drug delivery systems based on polymeric core-shell architectures. Angew. Chem., Int. Ed. 43, 278–282.

Communications (20) Patri, A. K., Majoros, I. J., and Baker, J. R., Jr. (2002) Dendritic polymer macromolecular carriers for drug delivery. Curr. Opin. Chem. Biol. 6, 466–471. (21) Liu, M., and Fréchet, J. M. J. (1999) Designing dendrimers for drug delivery. Pharm. Sci. Tech. Today 2, 393–401. (22) Woller, E. K., Walter, E. D., Morgan, J. R., Singel, D. J., and Cloninger, M. J. (2003) Altering the strength of lectin binding interactions and controlling the amount of lectin clustering using mannose/hydroxyl-functionalized dendrimers. J. Am. Chem. Soc. 125, 8820–8826. (23) Woller, E. K., and Cloninger, M. J. (2002) The lectin-binding properties of six generations of mannose-functionalized dendrimers. Org. Lett. 4, 7–10. (24) Jayaraman, N., Nepogodiev, S. A., and Stoddart, J. F. (1997) Synthetic carbohydrate-containing dendrimers. Chem. Eur. J. 3, 1193–1199. (25) Page, D., and Roy, R. (1997) Synthesis and biological properties of mannosylated starburst poly(amidoamine) dendrimers. Bioconjugate Chem. 8, 714–723. (26) Kieburg, C., and Lindhorst, T. K. (1997) Glycodendrimer synthesis without using protecting groups. Tetrahedron Lett. 38, 3885–3888. (27) Mammen, M., Choi, S.-K., and Whitesides, G. M. (1998) Polyvalent interactions in biological systems: implications for design and use of multivalent ligands and inhibitors. Angew. Chem., Int. Ed. 37, 2754–2794. (28) Krause, W., Hackmann-Schlichter, N., Maier, F. K., and Müller, R. (2000) Dendrimers in diagnostics. Top. Curr. Chem. 210, 261–308. (29) Wathier, M., Jung, P. J., Carnahan, M. A., Kim, T., and Grinstaff, M. W. (2004) Dendritic macromers as in situ polymerizing biomaterials for securing cataract incisions. J. Am. Chem. Soc. 126, 12744–12745. (30) Grinstaff, M. W. (2002) Biodendrimers: new polymeric biomaterials for tissue engineering. Chem. Eur. J. 8, 2838–2846. (31) Carnahan, M. A., Middleton, C., Kim, J., Kim, T., and Grinstaff, M. W. (2002) Hybrid dendritic-linear polyester-ethers for in situ photopolymerization. J. Am. Chem. Soc. 124, 5291–5293. (32) Lee, H., Baker, J. R., Jr., and Larson, R. G. (2006) Molecular dynamics studies of the size, shape, and internal structure of 0% and 90% acetylated fifth-generation polyamidoamine dendrimers in water and methanol. J. Phys. Chem. B 110, 4014–4019. (33) Maiti, P. K., Çag˘ın, T., Wang, G., and Goddard, W. A. (2004) Structure of PAMAM dendrimers: generations 1 through 11. Macromolecules 37, 6236–6254. (34) Lee, I., Athey, B. D., Wetzel, A. W., Meixner, W., and Baker, J. R., Jr. (2002) Structural molecular dynamics studies on polyamidoamine dendrimers for a therapeutic application: effects of pH and generation. Macromolecules 35, 4510–4520. (35) Esfand, R., and Tomalia, D. A. (2001) Poly(amidoamine) (PAMAM) dendrimers: from biomimicry to drug delivery and biomedical applications. Drug DiscoVery Today 6, 427–436. (36) Shi, X., Majoros, I. J., Patri, A. K., Bi, X., Islam, M. T., Desai, A., Ganser, T. R., and Baker, J. R., Jr. (2006) Molecular heterogeneity analysis of poly(amidoamine) dendrimer-based mono- and multifunctional nanodevices by capillary electrophoresis. Analyst 131, 374–381. (37) Jacobson, K. A., and Gao, Z.-G. (2006) Adenosine receptors as therapeutic targets. Nat. ReV. Drug DiscoVery 5, 247–264. (38) Yan, L., Burbiel, J. C., Maass, A., and Müller, C. E. (2003) Adenosine receptor agonists: from basic medicinal chemistry to clinical development. Exp. Opin. Emerging Drugs 8, 537–576. (39) Fredholm, B. B., IJzerman, A. P., Jacobson, K. A., Klotz, K.N., and Linden, J. (2001) International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacol. ReV. 53, 527–552. (40) Hutchison, A. J., Williams, M., de Jesus, R., Yokoyama, R., Oei, H. H., Ghai, G. R., Webb, R. L., Zoganas, H. C., Stone, G. A., and Jarvis, M. F. (1990) 2-(Arylalkylamino)adenosinuronamides: A new class of highly selective adenosine A2 receptor ligands. J. Med. Chem. 33, 1919–1924.

Communications (41) Jarvis, M. F., Schulz, R., Hutchison, A. J., Do, U. H., Sills, M. A., and Williams, M. (1989) [3H]CGS 21680, A selective A2 adenosine receptor agonist directly labels A2 receptors in rat brain. J. Pharm. Exp. Therap. 251, 888–893. (42) Hutchison, A. J., Webb, R. L., Oei, H. H., Ghai, G. R., Zimmerman, M. B., and Williams, M. (1989) CGS 21680C an A2 selective adenosine receptor agonist with preferential hypotensive activity. J. Pharm. Exp. Therap. 251, 47–55. (43) Haugland, R. P. (2002) Handbook of Fluorescent Probes and Research Products, 9th ed., Molecular Probes, Invitrogen, pp 20–35. (44) Panchuk-Voloshina, N., Haugland, R. P., Bishop-Stewart, J., Bhalgat, M. K., Millard, P. J., Mao, F., Leung, W.-Y., and Haugland, R. P. (1999) Alexa dyes, a series of new fluorescent dyes that yield exceptionally bright, photostable conjugates. J. Histochem. Cytochem. 47, 1179–1188. (45) Peterson, J., Allikmaa, V., Subbi, J., Pehk, T., and Lopp, M. (2003) Structural deviations in poly(amidoamine) dendrimers: a MALDI-TOF MS analysis. Eur. Polym. J. 39, 33–42. (46) Petkova, V., Parvanova, V., Tomaliab, D., Swansonb, D., Bergstromb, D., and Vogt, T. (2005) 3D structure of dendritic and hyper-branched macromolecules by X-ray diffraction. Solid State Commun. 134, 671–675. (47) Gachet, C. (2006) Regulation of platelet functions by P2 receptors. Annu. ReV. Pharmacol. Toxicol. 46, 277–300.

Bioconjugate Chem., Vol. 19, No. 2, 2008 411 (48) Gessi, S., Varani, K., Merighi, S., Ongini, E., and Borea, P. A. (2000) A2A adenosine receptors in human peripheral blood cells. Br. J. Pharmacol. 129, 2–11. (49) Ledent, C., Vaugeois, J.-M., Schiffmann, S. N., Pedrazzini, T., Yacoubi, M. E., Vanderhaeghen, J.-J., Costentin, J., Heath, J. K., Vassart, G., and Parmentier, M. (1997) Aggressiveness, hypoalgesia and high blood pressure in mice lacking the adenosine A2a receptor. Nature 388, 674–678. (50) Gurden, M. F., Coates, J., Ellis, F., Evans, B., Foster, M., Hornby, E., Kennedy, I., Martin, D. P., Strong, P., and Vardey, C. J. (1993) Functional characterization of three adenosine receptor types. Br. J. Pharmacol. 109, 693–698. (51) Jacobson, K. A., Xie, R., Young, L., Chang, L., and Liang, B. T. (2000) A novel pharmacological approach to treating cardiac ischemia: binary conjugates of A1 and A3 adenosine receptor agonists. J. Biol. Chem. 275, 30272–30279. (52) Jacobson, K. A., Lipkowski, A. W., Moody, T. W., Padgett, W., Pijl, E., Kirk, K. L., and Daly, J. W. (1987) Binary drugs: conjugates of purines and a peptide that bind to both adenosine and substance P receptors. J. Med. Chem. 30, 1529–1532. (53) Handagama, P. J., Shuman, M. A., and Bainton, D. F. (1989) Incorporation of intravenously injected albumin, immunoglobulin G, and fibrinogen in guinea pig megakaryocyte granules. J. Clin. InVest. 84, 73–82. BC700327U