Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jan 15.
Published in final edited form as: Eur J Med Chem. 2019 Nov 22;186:111886. doi: 10.1016/j.ejmech.2019.111886

Conjugable A3 adenosine receptor antagonists for the development of functionalized ligands and their use in fluorescent probes

Stephanie Federico a,*, Enrico Margiotta b, Stefano Moro b, Eszter Kozma c, Zhan-Guo Gao c, Kenneth A Jacobson c, Giampiero Spalluto a
PMCID: PMC7199890  NIHMSID: NIHMS1562663  PMID: 31787357

Abstract

Compounds able to simultaneously bind a biological target and be conjugated to a second specific moiety are attractive tools for the development of multi-purpose ligands useful as multi-target ligands, receptor probes or drug delivery systems, with both therapeutic and diagnostic applications. The human A3 adenosine receptor is a G protein-coupled receptor involved in many physio-pathological conditions, e.g. cancer and inflammation, thus representing a promising research target. In this work, two series of conjugable hA3AR antagonists, based on the pyrazolo[4,3-e]-1,2,4-triazolo[1,5-c]pyrimidine nucleus, were developed. The introduction of an aromatic ring at the 5 position of the scaffold, before (phenylacetamido moiety) or after (1,2,3-triazole obtained by click chemistry) the conjugation is aimed to increase affinity and selectivity towards the hA3AR receptor. As expected, conjugable compounds showed good affinity towards the hA3AR. In order to prove their potential in the development of hA3AR ligands for different purposes, compounds were also functionalized with fluorescent probes. Unfortunately, conjugation decreased affinity and selectivity for the target as compared to the hA2AAR. Computational studies identified specific non-conserved residues of the extracellular loops which constitute a structural barrier able to discriminate between ligands, giving insights into the rational development of new highly selective ligands.

Keywords: G protein-coupled receptor, Adenosine receptors, Fluorescent ligands, Molecular modeling, Molecular probes

1. Introduction

Activation or blockade of adenosine receptors (ARs) are responsible for diverse pharmacological effects in various tissues and organs. Considering the wide distribution of endogenous adenosine, it is generally accepted that the regulation of ARs has great therapeutic potential [16].

ARs are members of the superfamily of G protein-coupled receptors (GPCRs), and four subtypes are currently known, named A1AR, A2AAR, A2BAR, and A3AR [7]. In recent decades, numerous medicinal chemistry groups have developed promising ligands, agonists and antagonists for these receptor subtypes [8]. In particular, the search for selective antagonists held greater appeal than selective agonists, not only for their potential therapeutic applications but also for their preferential use as molecular probes for pharmacological characterization of receptors [9].

In this field, our research group and others have extensively investigated the pyrazolo[4,3-e]-1,2,4-triazolo[1,5-c]pyrimidine nucleus (PTP) that, by appropriate modifications at the N7, N8, and 5 positions, led to very potent and selective human (h) A2A and A3 AR antagonists [1016]. These derivatives have previously been used for the study of the functional role of A2A and A3 ARs, in particular through the preparation of radiolabeled or irreversible derivatives [1721]. Nevertheless, another interesting approach for the investigation of AR subtypes could be considered the functionalization of the PTP scaffold with spacers able to link various groups such as fluorophores or nanoparticles for drug delivery [2232].

Using the PTP scaffold we have recently reported a new series of derivatives of general formulae 1, in which a diamino moiety of different lengths was introduced at the 5 position. Using this approach, a new series of derivatives bearing a fluorophore group linked to the free amino group was reported as potential fluorescent probes (Fig. 1) [33].

Fig. 1.

Fig. 1.

Structures of previously reported AR antagonists and reference compounds.

In this work, we explored the structure activity relationship (SAR) of conjugable AR ligands useful for the development of probes selective for the hA3AR. In addition to their usefulness to better define the specific role of the target protein in both physiological and pathological processes, availability of hA3AR selective probes could be used for diagnostic purposes. In fact, hA3AR was found to be overexpressed in certain cancers, such as melanoma and breast and colorectal cancer, and the overexpression level well correlates with the severity of the tumor, thus making it a possible biomarker of such tumors [3438]. Unfortunately, the previously synthesized PTP derivatives generally showed a poor affinity for the hA3AR and low A3AR selectivity versus the other receptor subtypes [33]. A possible explanation of lack of affinity and selectivity of these compounds could be attributed to the absence of an aromatic moiety at the 5 position, which previous studies indicated to be important for both affinity and selectivity at the hA3AR [10].

Taking into account these experimental observations and with the aim of improving selectivity versus the hA3AR subtype, we decided to design a new series of conjugable PTPs bearing an aromatic portion at the 5 position (3–18) and anchored to chains of different lengths. These chains act as spacers and at their terminal position bear a functional group for subsequent functionalization. On the basis of our modeling of the AR interactions of PTP ligands, the most reliable orientation of the linker is towards the extracellular loops (ELs) of the receptors [33]. This means that any difference in sequence and conformation of the ELs of the AR subtypes could affect either selectivity or affinity through direct ligand interaction. Indeed, our recent work about the selection of the best template for hA3AR homology modeling highlighted that EL region displays the highest sequence variability between the hA1, hA2A, and hA3AR subtypes [39], suggesting that the ligand partitioning between the ELs and the hydrophobic cavity is crucial for binding.

Thus, finally, we decided to conjugate the new functionalized AR ligands with a fluorescent moiety (e.g. fluorescein) that is able to protrude in the extracellular side, giving new leads for the development of hA3AR fluorescent probes (Fig. 2).

Fig. 2.

Fig. 2.

Structures of the synthesized PTP derivatives.

Alternatively, a parallel series of N5 alkynyl functionalized antagonists (19–22) was prepared as precursors for copper(I)-catalyzed azide-alkyne (click) cycloaddition. Based on previous studies performed on the triazoloquinazoline nucleus of the nonseletive antagonist CGS15943, which led to a potent and selective fluorescent hA3AR antagonists (MRS5449, 2, Fig. 1) [40], a fluorophore (e.g. Alexa Fluor-488) was introduced to afford potential A3AR fluorescent antagonist 23 (Fig. 2).

1.1. Chemistry

All the designed compounds 3–23 have been synthesized as summarized in Schemes 13 [41]. Saponification with lithium hydroxide of previously reported PTP ester derivative 24 [42] led to the corresponding carboxylic acid 3, which after condensation in presence of EDCI with the appropriate mono BOC-protected diamine 25–29 afforded the desired functionalized compounds 4–8. A common BOC deprotection with trifluoroacetic acid in dichloromethane provided the free amino compounds 9–13 as salts. Then, after coupling with fluorescein isothiocyanate (FITC) in methanol under an argon atmosphere, the final compounds 14–18 were obtained (Scheme 1).

Scheme 1.

Scheme 1.

Synthesis of fluorescent adenosine receptor ligands 14–18. Reagents and conditions: i: LiOH, THF, MeOH, water, rt. overnight; ii: NH2XNHBoc (25–29), EDCI, DMAP, DMF, rt, overnight; iii: TFA, DCM, rt, 2 h; iv: FITC, TEA, MeOH, rt, 72 h.

Scheme 3.

Scheme 3.

Synthesis of Alexa Fluor-488 fluorescent AR ligand 23. Reagents and conditions: i: Alexa Fluor 488 5-carboxamido-(6-azidohexanyl) bis(triethylammonium salt), aq. sodium ascorbate, aq. cupric sulfate pentahydrate solution, water, DMF, rt, overnight.

Alkynyl PTP derivatives 19–22 were instead prepared by reacting the free amino PTP derivative 30 with the appropriate acyl chlorides 35–38 which were obtained by treatment of the corresponding carboxylic acids 31–34 with oxalyl chloride (Scheme 2).

Scheme 2.

Scheme 2.

Synthesis of alkyne derivatives 19–22. Reagents and conditions: i: oxalyl chloride, DMF cat., CDCl3, 0 °C to rt, 3 h; ii: TEA. dioxane, reflux, overnight.

The final derivative 23 was prepared by reacting the alkynyl derivative 22 with Alexa Fluor-488 5-carboxamido (6-azidohexanyl)bis (triethylammonium salt) in DMF [40] (Scheme 3).

2. Results and discussion

Newly synthesized compounds (3–23) were tested at the hA1, hA2A and hA3 ARs expressed in CHO (hA1AR, hA3AR) and HEK293 (hA2AAR) cells: [3H]RPIA (hA1AR), [3H]CGS21680 (hA2AAR) and [125I]AB-MECA (hA3AR) were used as radioligands in binding assays (Table 1) [40,43].

Table 1.

Binding profile of synthesized compounds 3–23 at the hA1, hA2A and hA3 ARs (see Fig. 2 for compounds’ structures).

Compd X R hA1a hA2Ab hA3c hA1/hA3 hA2A/hA3 hA1/hA2A
3 4% 1010 128 >78 7.8 >10
4 (CH2)3 Boc 15% 150 2.75 >3636 54.5 >66
5 (CH2)4 Boc 2% 304 3.46 >2890 88 >33
6 (CH2)5 Boc 14% 376 4.47 >2237 84.1 >26.5
7 (CH2CH2O)2CH2CH2 Boc 7% 456 74.8 >133.6 6.1 >22
8 (CH2)3O(CH2CH2O)2(CH2)3 Boc 11% 327 82.4 >121.3 3.9 >30.5
9 (CH2)3 Hd 9% 175 8.06 >1240 21.7 >57
10 (CH2)4 Hd 4% 172 6.00 >1666 28.6 >58
11 (CH2)5 Hd 6% 203 2.66 >3759 76.3 >49
12 (CH2CH2O)2CH2CH2 Hd 17% 267 19.2 >520 14 >37.4
13 (CH2)3O(CH2CH2O)2(CH2)3 Hd 1% 328 10.3 >970 30.3 >30
14 (CH2)3 FITC 27% 110 104 >96 1 >96
15 (CH2)4 FITC 28% 60.4 73.6 >23.7 0.82 >129
16 (CH2)5 FITC 21% 106 96.7 >103 1.1 >94
17 (CH2CH2O)2CH2CH2 FITC 29% 95.7 137 >73 0.69 >94
18 (CH2)3O(CH2CH2O)2(CH2)3 FITC 32% 91.5 207 >48 0.44 >109
19 (CH2)2 55% 133 11.0 >909 16.5 >2.34
20 (CH2)3 n.d. 199 4.11 48.4
21 (CH2)4 891 151 7.26 122.7 20.7 5.9
22 (CH2)5 62% 482 12.8 >781 37.6 >20.7
23e n.d. 90 31.8 2.83
a

Displacement of specific [3H]R-PIA binding at hA1AR expressed in CHO cells, (n = 3–6).

b

Displacement of specific [3H]CGS21680 binding at hA2AAR expressed in CHO cells, (n = 3–6).

c

Displacement of specific [125I]-AB-MECA binding at hA3AR expressed in CHO cells, (n = 3–6). Data are expressed as Ki (nM) or as % of inhibition (in italics) at a 10 μM concentration of radioligand. n.d. Not determined. Data complete of SEM are reported in ESI (Table S1).

d

Compounds as TFA salts.

e

MRS5763.

As clearly summarized in Table 1, all the synthesized ureido (3–18) and amido (19–23) compounds showed affinities at the hA3AR ranging from high nanomolar to nanomolar concentrations, with different degrees of selectivity versus the other subtypes.

The presence of a carboxylic group at the 5 position of compound 3 led to affinity values of 128 nM at the hA3AR and 1.01 mM at the hA2AAR, while it was inactive at the hA1AR (4% of radioligand displacement at 10 μM). Among all the series, compound 3 is the only one to exhibit a micromolar affinity against the hA2AAR, suggesting that the carboxylic moiety is involved in interactions with different key residues in the orthosteric binding site of the AR subtypes, which are, thus, responsible for the observed selectivity.

Conjugation of the acid moiety with the mono-N-BOC protected diamino spacers led to compounds 4–8 which showed high affinity for the hA3AR (Table 1). It should be noted that the length of the diamino spacer could be a major determinant of the hA3AR interaction. In fact, while shorter spacers (e.g. compounds 4–6) gave good results in terms of both affinity (Ki hA3 = 2.75–4.47 nM) and selectivity (hA2A/hA3 = 54.5–88) for the hA3AR (e.g. compound 3, Ki hA3 = 2.75 nM, hA1/hA3 = > 3636; hA2A/hA3 = 54.5), longer diamino moieties (e.g compounds 7, 8) led to derivatives showing a significant loss of affinity at the hA3AR with a consequent reduction of selectivity, especially towards the hA2AAR subtype (e.g. compound 8, Ki hA3 = 82.4 nM, hA1/hA3 > 133; hA2A/hA3 = 6.1).

However, the corresponding N-BOC deprotected derivatives 9–13 showed good hA3AR affinity independently of the spacer length, albeit, in general, the selectivity vs the hA2AAR subtype was significantly reduced with respect to the N-BOC derivatives (4–8) (e.g. compound 10, Ki hA3 = 6 nM, hA1/hA3 = > 1666; hA2A/hA3 = 28 vs. compound 5, Ki hA3 = 3.46 nM, hA1/hA3 = > 2890; hA2A/hA3 = 88). Interestingly, compounds bearing PEG-like spacers (7, 8 and 12, 13) showed a different selectivity profile, denoting that also the nature of the spacer was important for the AR interactions, in particular for the hA2A and hA3 ARs. In fact, while in the BOC protected series a very low selectivity was observed (compound 7, hA2A/hA3 = 6.1 and compound 8, hA2A/hA3 = 3.9), in the free amino series, compounds gained affinity and selectivity towards the hA3AR subtype (compound 12, hA2A/hA3 = 14 and compound 13, hA2A/hA3 = 30.3). In order to assess if the nature, and in particular the polarity of the linker could be someway related to the affinity profile displayed by compounds, the logarithm of the octanol/water partition coefficient (logP(o/w)) was calculated for each compound (Fig. 3). As expected, Boc-free ligands showed lower values than Boc-functionalized ones, with polyethers being more polar than alkyls. Only for the Boc-free mini-series, the affinity at the hA3AR increased with the logP(o/w) value of the compound, while in the other homologous series a correlation between polarity and affinity at the three adenosine receptors was not observed.

Fig. 3.

Fig. 3.

Calculated octanol/water partition coefficient logarithm.

Concerning the fluorescent ligands, introduction of a fluorophore such as FITC on amino function gave compounds 14–18 which showed a significant reduced hA3AR affinity with a drastic reduction of selectivity vs hA2AAR, independently of the spacer length. (e.g. compound 16, Ki hA3 = 96.7 nM, hA1/hA3 = > 103; hA2A/hA3 = 1.1). Thus, they can be considered as dual hA2A/hA3AR fluorescent ligands.

In the amido alkynyl series (19–22), all the synthesized compounds proved to be quite potent at the hA3AR with affinity ranging from 4 to 12 nM and poor affinities versus the other AR subtypes, even if the hA2AAR selectivity was not pronounced (e.g. compound 21, Ki hA3 = 7.26 nM, hA1/hA3 = 122; hA2A/hA3 = 20.7). These results suggest that the triple bond likely formed interactions with nonconserved residues in the hA3AR binding site.

Finally, Alexa Fluor-488 conjugated derivative 23, a potential fluorescent ligand, retained a good affinity at the hA3AR (Ki hA3 = 31.8 nM) and selectivity towards the hA2AAR was improved (hA1/hA3 = 2.83) compared to FITC derivatives (19–22), even if further optimization is needed (Table 1).

It is quite clear that, in the present case, it was not possible to obtain a conjugable, selective ligand which could be differentially functionalized and that retained the selectivity and affinity profile of the parent compound. Even if the functional moiety (such as the fluorophore) was separated from the pharmacophore through a long spacer, it influenced the ligand-receptor recognition process.

Molecular modeling studies were performed with a view to rationalize affinity data and SAR of the tested ligands. Molecular docking simulations were performed on each AR subtype, using inactive states, in the presence of the sodium ion and its hydration shell [44].

Independently of the receptor subtype considered, the ligands retained the classical binding pattern of AR antagonists: hydrogen bonds with Asn (6.55) and π-π interactions with Phe (EL2) (Videos S13). Moreover, the linker was generally oriented towards the ELs. Despite the overall findings, in the case of the hA1AR (Video S1), several binding modes result energetically unfavorable or geometrically unreliable (e.g. pointing the furyl ring towards the ELs), consistent with the high A3AR selectivity observed vs A1. Interaction energy fingerprint (IEF) maps calculated for each docking pose on three subtypes, revealed that docking at the hA1AR (Fig. 4, A) was less favourable than at hA2A and hA3 ARs (Fig. 4B and C), because of unfavourable electrostatics at the EL2 and TMs 5–7. hA3AR ligand poses were highly favoured at these locations, more than at hA2A and hA1 ARs, also in terms of hydrophobic contacts. Furthermore, as explained below, in the hA3AR, TM1 contributed significantly to stabilization by electrostatics, differently from other subtypes.

Fig. 4.

Fig. 4.

IE map of the electrostatic and hydrophobic interactions between each compound (y-axis) and each residue (x-axis) of AR subtypes (hA1, A; hA2A, B; hA3, C) The strength of the electrostatic interaction is represented by a colorimetric scale going from blue to red, from negative to positive values. The strength of the hydrophobic interaction is represented by a colorimetric scale going from white to dark green, from low to high values. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

Supplementary video related to this article can be found at https://doi.org/10.1016/j.ejmech.2019.111886

Supplementary data related to this article can be found at https://doi.org/10.1016/j.ejmech.2019.111886.

Supplementary video related to this article can be found at https://doi.org/10.1016/j.ejmech.2019.111886.

Per-residue interaction energy profiles (electrostatic, van der Waals and hydrophobic contributions) were calculated for three representative antagonists: 3 (EA2), 16 (FITC ligand, EA8) and 22 (ethynyl ligand, EA20). Both conserved and non-conserved residues for each receptor subtype were chosen from TMs and ELs, in order to give a better understanding of how they affect binding affinity and selectivity, as shown in Figs. 57.

Fig. 5.

Fig. 5.

Interaction energy histograms of compound 3 (EA2) for conserved and non-conserved residues of A1, A2A and A3 subtypes.

Fig. 7.

Fig. 7.

Selected docking pose of ligand 16 (EA8) at the hA3AR subtype and visualization of the most relevant structural regions interacting with the linker and the fluorophore (FITC).

Antagonist 3 (EA2) is the parent compound of the series. Nonetheless, it is the only one to show affinity towards the hA2AAR in the micromolar range (Ki hA2A = 1.01 μM, Ki hA3 = 128 nM). Interestingly, it is characterized by a negatively charged carboxylic moiety. When docked to the hA1AR, the ligand pose is highly disfavoured by strong repulsion with Glu172 (EL2) and other polar residues (Fig. 5). Similarly, at the hA2AAR, Glu169 (EL2) disfavors the final state. The mutation of Glu into Val169 at the EL2, and the mutation of Val in Leu 90 (3.32) at the TM3 of the hA3AR are crucial in improving complex stability by hydrophobic interactions. These results predict that negatively charged moieties, exposed to the ELs of the hA1 and hA2A ARs, inhibit binding by electrostatic repulsion with Glu (EL2). Finally, a favourable electrostatic interaction is observed with the conserved residue Glu19 (1.39) at the TM1 of the hA3AR, while the same is absent in other subtypes, giving indications about the different shape and nature of the A3 binding site. The antagonist 16 (EA8, Figs. 6 and 7), bearing a longer linker and a fluorophore moiety (FITC), shows strong van der Waals repulsion with Glu172 (EL2) and Asn 254 (6.55) at the hA1AR, while such interactions are favourable at the hA2AR (Glu169, Asn 253) and accompanied by several hydrophobic contacts (Leu267 (6.70), Val 92 (3.32), Ile 74 (3.14)). The hA3AR shows a similar profile, but hydrophobic interactions are not significant, consistently with the absence of selectivity observed, as compared to A2A. However, nonconserved residues Gly 257, Glu258 and Gln261 (EL3) contribute favourably by electrostatic and van der Waals interactions with the phenyl-acetamido moiety at the 5 position of the PTP. The FITC moiety is also involved and interacts with several residues located at the ELs 2 and 3, participating in the overall recognition process. As shown in Fig. 7, the alkyl linker sits between the loops and adapts to their conformation thanks to its high rotational fredoom. Depending on the latter, the fluorophore is pointed between the ELs, facing several polar residues like Thr173, Arg 174 (EL2) and Asn 256 (EL3).

Fig. 6.

Fig. 6.

Interaction energy histograms of compound 16 (EA8) for conserved and non-conserved residues of A1, A2A and A3 subtypes.

Ethynyl derivative 22 (EA20) prefers the hA3AR cavity through several hydrophobic contacts with non-conserved residues and electrostatic interactions, mainly due to the terminal triple bond, which could explain the selectivity observed (Fig. 8). At the hA2AAR, in fact, such interactions are absent, and even some electrostatic repulsion is observed with Glu169 (EL2) and His 250 (6.52).

Fig. 8.

Fig. 8.

Interaction energy histograms of compound 22 (EA20) for conserved and non-conserved residues of A1, A2A and A3 subtypes.

Finally, three supervised molecular dynamics (SuMD) simulations were performed with a view to 1) simulate binding to the hA3AR, 2) inspect which structural domains at the extracellular side or the hydrophobic cavity may be determinants of the ligand preference. Ligand 3 (EA2) was chosen because it yields the lowest affinity towards the hA2AAR within the entire series while preserving affinity for the A3 subtype in the nanomolar range. Moreover, basing on per-residue interaction fingerprints (Figs. 5 and 6), its carboxylic moiety is clearly important for the A2A/A3 selectivity. The ligand completely reached the binding site only in simulation 1. As shown in Video S4, the ligand intercalates between the ELs 2–3 and finally binds the receptor assuming a pose closely resembling the one predicted by molecular docking (Video S1). Two residues (conserved and non-conserved, respectively) seem to play an important role in the recognition process: Glu19 (1.39) and Gln261 (EL3). Glu19, located at the TM1, interacts with the pyrazole ring favouring binding during the entire process. Gln261 (EL3) (Leu267 in the hA2A AR) makes hydrogen bonds with the ligand’s carboxylate exposed to the ELs, either in the initial phase of approach or in the phase of stabilization. Such interaction is detectable in the ligand-protein interaction energy profile for distances of about 10 Å. Interestingly, in both simulations 2 and 3, the ligand stops just at 10 Å, between the ELs 2–3, giving the same interaction energy peak observed in simulation 1 (Fig. 9). The interaction energy is weaker than the bound state 1, indicating that the ligand can effectively cross that “barrier”. This result indicates that the ELs’ structural region is crucial for binding, as ligands can cross it more (or less) efficiently depending on the specific substitution present at the 5 position of PTP. Indeed, the ligand 3 establishes highly favourable contacts even at long distances, corresponding to the EL3 or EL2 (20–30 Å), where several non-conserved residues are located, in agreement with per-residue interaction energy calculations. Some contacts are more stabilizing than the final state 1 (meta-binding sites), however, the great conformational freedom of the EL2 (see Video S4) could explain the progression of the ligand path.

Fig. 9.

Fig. 9.

Ligand 3 (EA2)-protein interaction energy vs distance from the binding site, for each SuMD simulation. VdW surfaces of the respective final states are reported with different colors, along with particular regions of the receptor. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

Supplementary data related to this article can be found at https://doi.org/10.1016/j.ejmech.2019.111886.

3. Experimental section

3.1. Chemistry

3.1.1. General

Reactions were routinely monitored by thin-layer chromatography (TLC) on silica gel (precoated Macherey-Nagel, 60FUV254). Flash chromatography was performed using Macherey-Nagel, silica 60, 230–400 mesh silica gel. Light petroleum ether refers to the fractions boiling at 40–60 °C. Melting points were determined on a Buchi-Tottoli instrument and are uncorrected. 1H NMR were determined in CDCl3, DMSO-d6 or CD3OD solutions with a Varian Gemini 200, a Bruker 400 or a Varian 500 spectrometers. Peaks positions are given in parts per million (δ) downfield relative to the central peak of the solvents, and J values are given in Hz. The following abbreviations were used: s, singlet; bs, broad singlet; d, doublet; dd, double doublet; bd, broad doublet; t, triplet; m, multiplet. Electrospray mass spectra were recorded on a ESI Bruker 4000 Esquire spectrometer and compounds were dissolved in methanol; accurate mass spectra were recorded on a micrOTOF-Q-Bruker and compounds were dissolved in methanol. Absorption experiments were performed on a Varian Cary 5000 spectrophotometer. Fluorescence measurements were recorded on Varian Cary Eclipse fluorescence spectrophotometer, with excitation filter at 361 nm, emission filter at 390–600 nm (all experiments were performed using a λex of 488 nm). Oct-7-ynoic acid was synthesized as reported in the literature.

3.1.2. Synthesis of 2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo [4,3-e] [1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetic acid (3)

Ethyl ester derivative 24 (1.95 mmol, 900 mg) was dissolved in a mixture of tetrahydrofuran/methanol/water (4:1:1, 12 mL) and lithium hydroxide monohydrate (5.9 mmol, 245 mg) was added. The suspension was stirred at room temperature overnight. Reaction was monitored by TLC (ethyl acetate 9.5/methanol 0.5). Water was added to the mixture, the solution was cooled to 0 °C and 1 M hydrochloric acid was added (pH 2) leading to the precipitation of a pale brown solid. Solid was filtered and washed with ethyl ether and identified as the desired compound in a quantitative yield. Mp > 300 °C; 1H NMR (200 MHz, DMSO-d6) δ 10.67 (s, 1H) 9.55 (bs, 1H), 8.78 (s, 1H), 7.93 (s, 1H) 7.78–7.36 (m, 3H), 7.38–6.96 (m, 3H), 6.76 (bs, 1H), 4.34 (s, 3H), 3.56 (s, 2H). ES-MS negative mode (methanol) m/z: 431.0 [M − H].

3.1.3. General procedure for the synthesis of amides 4–8

The carboxylic acid derivative 3 (0.58 mmol, 250 mg) was dissolved in DMF (3 mL) and the appropriate mono-BOC-protected diamine (25–29) was added (0.58 mmol). Reaction was stirred and cooled to 0 °C under an argon atmosphere and then 1-ethyl-3-(3-dimethylaminopropyl)carbodiimmide hydrochloride (EDCI·HCl, 1.16 mmol, 222 mg) and 4-dimethylaminopyridine (DMAP, 1.44 mmol, 177 mg) were added. Mixture was stirred at room temperature overnight and monitored by TLC (dichloromethane 9.3/methanol 0.7). The solvent was evaporated under reduced pressure. The residue was suspended in water and extracted with ethyl acetate (3 times). The organic layers were collected and dried over sodium sulfate anhydrous and the solvent removed under reduced pressure. The crude product was purified on flash silica column chromatography (dichloromethane 9.7/methanol 0.3) to afford the desired compound (4–8).

3.1.4. tert-Butyl (3-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo [4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamido)propyl)carbamate (4)

Yield 29.4%; white solid; mp 198–205 °C; 1H NMR (200 MHz, CDCl3) δ 11.19 (s, 1H), 8.60 (s, 1H), 8.23 (s, 1H), 7.80–7.46 (m, 3H), 7.30–7.26 (m, 3H), 6.62 (dd, J = 3.4, 1.8 Hz, 1H), 6.05 (bs, 1H), 4.94 (bs, 1H), 4.21 (s, 3H), 3.58 (s, 2H), 3.27 (q, J = 6 Hz, 2H), 3.15–3.00 (m, 2H), 1.56 (t, J = 6, 2H), 1.42 (s, 9H); ES-MS (methanol) m/z: 611.3 [M+Na]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C28H32N10O5 611.2449; Found 611.2440.

3.1.5. tert-Butyl (4-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo [4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamido)butyl)carbamate (5)

Yield 34.5%; white solid; mp 186–193 °C; 1H NMR (200 MHz, CDCl3) δ 11.19 (s, 1H), 8.60 (s, 1H), 8.22 (s, 1H), 7.90–7.48 (m, 3H), 7.38–7.09 (m, 3H), 6.62 (s, 1H), 5.63 (bs, 1H), 4.60 (bs, 1H), 4.21 (s, 3H), 3.57 (s, 2H), 3.22 (bs, 2H), 3.09 (bs, 2H), 1.75 (bs, 4H), 1.43 (s, 9H); ES-MS (methanol) m/z: 625.4 [M+Na]+, 641.3 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C29H34N10O5 625.2606; Found 625.2608.

3.1.6. tert-Butyl (5-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo [4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamido)pentyl)carbamate (6)

Yield 28%; white solid; mp 78–185 °C; 1H NMR (200 MHz, CDCl3) δ 11.20 (s, 1H), 8.61 (s, 1H), 8.23 (s, 1H), 7.66 (bs, 3H), 7.26 (bs, 3H), 6.62 (bs, 1H), 5.47 (bs, 1H), 4.56 (bs, 1H), 4.21 (s, 3H), 3.57 (s, 2H), 3.20 (bs, 2H), 3.07 (bs, 2H), 1.60–1.10 (m, 15H); ESMS (methanol) m/z: 639.4 [M+Na]+, 655.3 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C30H36N10O5 639.2762; Found 639.2764.

3.1.7. tert-Butyl (2-(2-(2-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamido)ethoxy)ethoxy)ethyl)carbamate (7)

Yield 54%; white solid; mp 157–165 °C; 1H NMR (200 MHz, CDCl3) δ 11.18 (s, 1H), 8.60 (s, 1H), 8.23 (s, 1H), 7.78–7.56 (m, 3H), 7.42–7.15 (m, 3H), 6.62 (dd, J = 3.2, 1.6 Hz, 1H), 5.97 (bs, 1H), 5.02 (bs, 1H), 4.22 (s, 3H), 3.73–3.35 (m, 12H), 3.29 (bs, 2H), 1.42 (s, 9H); ES-MS (methanol) m/z: 685.4 [M+Na]+, 701.4 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C31H38N10O7 685.2817; Found 685.2807.

3.1.8. tert-Butyl (1-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo [4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)-2-oxo-7,10,13-trioxa-3-azahexadecan-16-yl)carbamate (8)

Yield 21.6%; white solid; mp 151–158 °C; 1H NMR (270 MHz, CDCl3) δ 11.19 (s, 1H), 8.60 (s, 1H), 8.24 (s, 1H), 7.80–7.61 (m, 3H), 7.45–7.19 (m, 3H), 6.63 (bs, 1H), 6.20 (bs, 1H), 4.96 (bs, 1H), 4.22 (s, 3H), 3.71–3.49 (m, 14H), 3.35 (bs, 2H), 3.20 (bs, 2H), 1.73 (t, J = 3.2, 4H), 1.49 (s, 9H); ES-MS (methanol) m/z: 757.4 [M+Na]+. HRMS Found (ESI-TOF) m/z: [M+Na]+ Calcd for C35H46N10O8 757.3392; 757.3393.

3.1.9. General synthesis of compounds 9–13 by N-BOC deprotection

The N-BOC protected derivatives (4–8) were dissolved in a solution of TFA and dichloromethane (1:1) and the mixtures were stirred for 2 h at room temperature. Reactions were monitored by TLC (dichloromethane 9.3/methanol 0.7). The solvent was removed under reduced pressure and the solids were filtered to afford the desired compounds as trifluoroacetate salts (9–13).

3.1.10. N-(3-aminopropyl)-2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamide trifluoroacetate salt (9)

Yield 89.8%; white solid; mp 142–150 °C; 1H NMR (200 MHz, CD3OD) δ 11.28 (s, 1H), 8.48 (s, 1H), 8.31 (bs, 1H), 7.76 (s, 1H), 7.62 (d, J = 8.2 Hz, 2H), 7.30 (d, J = 8.2 Hz, 1H), 7.24 (d, J = 3.4 Hz, 1H) 6.66 (bs, 1H), 4.15 (s, 3H), 3.52 (s, 2H), 3.34–3.10 (m, 2H), 2.89 (t, J = 7.0 Hz, 2H), 1.85 (p, J = 7.0 Hz, 2H). HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C23H24N10O3 489.2106; Found 489.2104.

3.1.11. N-(4-aminobutyl)-2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamide trifluoroacetate salt (10)

Yield 95.1%; white solid; mp 167–173 °C; 1H NMR (200 MHz, CD3OD) δ 8.48 (s, 1H), 8.19 (bs, 1H), 7.76 (s, 1H), 7.63 (d, J = 8.2 Hz, 2H), 7.39–7.13 (m, 3H), 6.70–6.62 (m, J = 1H), 4.14 (s, 3H), 3.51 (s, 2H), 3.32–3.30 (m, 2H), 2.92 (t, J = 6.7 Hz, 2H), 1.78–1.49 (m, 4H). HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C24H26N10O3 503.2262; Found 503.2260.

3.1.12. N-(5-aminopentyl)-2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamide trifluoroacetate salt (11)

Yield 85.3%; white solid; mp 170–178 °C; 1H NMR (200 MHz, CD3OD) δ 8.47 (s, 1H), 8.13 (bs, 1H), 7.75 (s, 1H), 7.61 (d, J = 7.7 Hz, 2H), 7.29 (d, J = 7.7 Hz, 2H), 7.25 (s, 1H), 6.66 (s, 1H), 4.14 (s, 3H), 3.49 (s, 2H), 3.33–3.08 (m, 2H), 2.88 (t, J = 7.3 Hz, 2H), 1.77–1.47 (m, 4H), 1.47–1.25 (m, 2H). HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C25H28N10O3 517.2419; Found 517.2417.

3.1.13. 2-(2-(2-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetamido) ethoxy)ethoxy)ethanamine trifluoroacetate salt (12)

Yield 83.3%; white solid; mp 152–160 °C; 1H NMR (200 MHz, CD3OD) δ 8.49 (s, 1H), 8.15 (bs, 1H), 7.76 (bs, 1H), 7.62 (d, J = 7.3 Hz, 2H), 7.44–7.13 (m, 3H), 6.66 (bs, 1H), 4.15 (s, 3H), 3.85–3.38 (m, 12H), 3.09 (bs, 2H); ES-MS (methanol) m/z: 563.3 [M+H]+, 585.3 [M+Na]+. HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C26H30N10O5 563.2473; Found 563.2474.

3.1.14. N-(3-(2-(2-(3-aminopropoxy)ethoxy)ethoxy)propyl)-2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetamide trifluoroacetate salt (13)

Yield 94.2%; white solid; mp 133–140 °C; 1H NMR (270 MHz, CDCl3) δ 11.13 (s, 1H), 8.28 (s, 1H), 7.81 (bs, 2H), 7.66 (s, 1H), 7.60 (d, J = 7.7 Hz, 2H), 7.41–7.13 (m, 3H), 6.62 (s, 1H), 5.40 (bs, 2H), 4.21 (s, 3H), 3.75 (s, 2H), 3.66–3.54 (m, 12H), 3.36 (bs, 2H), 3.24 (bs, 2H), 2.02 (bs, 2H), 1.78 (bs, 2H). HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C38H38N10O6 635.3049; Found 635.3046.

3.1.15. General synthesis of fluorescein conjugates (14–18)

0.075 mmol of amino derivatives 9–13 were dissolved in 4 mL of dry methanol. 21.6 μL (0.155 mmol) of TEA were added to the solution which was stirred at room temperature for 30 min under an argon atmosphere. Then 30.2 mg (0.075 mmol) of FITC were added and the reaction was stirred for 72 h in the dark. The products were purified by column chromatography starting with dichloromethane:methanol (9:1) as eluent. The obtained solids were suspended in dichloromethane, filtered and washed several times with ethyl ether yielding to the desired compounds as orange solids (14–18).

3.1.16. 5-(3-(3-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetamido) propyl)thioureido)-2-(6-hydroxy-3-oxo-3H-xanthen-9-yl)benzoate (14)

Yield 54.7%; orange solid; mp 198–204 °C; 1H NMR (200 MHz, DMSO-d6) δ 10.69 (s, 1H), 10.09 (bs, 3H), 8.76 (s, 1H), 8.41–8.05 (m, 3H), 7.98 (s, 1H), 7.76 (d, J = 3.2 Hz, 1H), 7.50 (d, J = 7.4 Hz, 2H), 7.39–7.03 (m, 4H), 6.76 (s, 1H), 6.65–6.47 (m, 6H), 4.12 (s, 3H), 3.50 (bs, 2H), 3.35 (s, 2H), 3.15 (bs, 2H), 1.71 (bs, 2H); 13C NMR (50 MHz, DMSO-d6) δ 180.29, 170.29, 168.44, 160.06, 159.79, 154.77, 152.93, 151.98, 151.93, 151.91, 149.57, 149.52, 148.96, 148.89, 145.30, 144.99, 141.23, 140.14, 136.60, 131.29, 129.65, 129.59, 129.49, 129.23, 129.20, 129.16, 129.13, 129.01, 126.24, 124.15, 118.95, 112.68, 112.54, 112.18, 109.81, 102.17, 98.68, 45.69, 41.86, 41.44, 36.37, 28.73; UV-VIS λmax 523 nm; Fluorescence λem 537; ES-MS (methanol) m/z 878.3 [M+H]+, 900.3 [M+Na]+, 916.3 [M+K]+. HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C44H35N11O8S 878.2464; Found 878.2336.

3.1.17. 5-(3-(4-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetamido) butyl)thioureido)-2-(6-hydroxy-3-oxo-3H-xanthen-9-yl)benzoate (15)

Yield 53.8%; orange solid; mp 200–207 °C; 1H NMR (200 MHz, DMSO-d6)δ 10.72 (s, 1H), 10.06 (bs, 3H), 8.74 (s, 1H), 8.26 (bs, 2H), 8.07 (bs, 1H), 7.96 (s, 1H), 7.76 (d, J = 6.9 Hz, 1H), 7.50 (d, J = 8.0 Hz, 2H), 7.35–7.08 (m, 4H), 6.75 (s, 1H), 6.70–6.37 (m, 6H), 5.76(s, 1H), 4.11 (s, 3H) 3.36 (s, 2H), 3.08 (bs, 4H), 1.51 (bs, 4H); 13C NMR (50 MHz, DMSO-d6) δ 180.21, 169.96, 168.47, 160.33, 160.10, 154.46, 153.01, 152.94, 152.43, 152.19, 149.96, 149.86, 148.88, 148.81, 145.30, 145.03, 141.33, 140.40, 136.61, 131.40, 129.44, 129.10, 129.04, 129.02, 126.20, 124.23, 118.98, 112.94, 112.91, 112.45, 112.16, 109.88, 102.25, 98.62, 45.76, 43.43, 41.85, 38.10, 26.82, 25.84; UV-VIS λmax 524 nm; Fluorescence λem 535; ES-MS (methanol) m/z: 892.4 [M+H]+, 914.3 [M+Na]+, 930.3 [M+K]+. HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C45H37N11O8S 914.2439; Found 914.2437.

3.1.18. 5-(3-(5-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)acetamido) pentyl)thioureido)-2-(6-hydroxy-3-oxo-3H-xanthen-9-yl)benzoate (16)

Yield 28%; orange solid; mp 190–200 °C; 1H NMR (200 MHz, DMSO-d6) δ 10.73 (s, 1H), 10.13 (bs, 1H), 8.75 (s, 1H), 8.27 (bs, 2H), 8.11–7.87 (m, 2H), 7.76 (d, J = 5.6 Hz, 1H), 7.50 (d, J = 6.3 Hz, 2H), 7.39–7.02 (m, 4H), 6.75 (s, 1H), 6.71–6.39 (m, 6H), 4.12 (s, 3H), 3.38 (s, 2H), 3.07 (bs, 4H), 1.74–1.10 (m, 4H), 1.08–0.75 (m, 2H); 13C NMR (50 MHz, DMSO-d6) δ 180.17, 169.90, 168.45, 160.02, 159.91, 154.73, 152.99, 151.96, 149.80, 149.77, 148.92, 148.88, 145.24, 145.03, 141.35, 140.32, 136.63, 131.37, 129.49, 129.44, 129.27, 129.23, 129.02, 129.00, 126.21, 124.08, 119.27, 118.92, 112.94, 112.88, 112.85, 112.46, 112.18, 109.86, 102.17, 98.64, 43.74, 41.84, 40.61, 40.20, 28.79, 28.03, 23.88; UV-VIS λmax 523 nm; Fluorescence λem 537; ES-MS (methanol) m/z: 928.4 [M+Na]+. HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C46H39N11O8S 928.2596; Found 928.2596.

3.1.19. 5-(3-(2-(2-(2-(2-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl) acetamido)ethoxy)ethoxy)ethyl)thioureido)-2-(6-hydroxy-3-oxo-3H-xanthen-9-yl)benzoate (17)

Yield 54.3%; orange solid; mp 182–185 °C; 1H NMR 1H NMR (200 MHz, DMSO-d6) δ 10.72 (s, 1H), 10.14 (bs, 3H), 8.75 (s, 1H), 8.30 (s, 1H), 8.13 (d, J = 5.0 Hz, 2H), 7.95 (s, 1H), 7.76 (d, J = 7.7 Hz, 1H), 7.49 (d, J = 7.9 Hz, 2H), 7.37–7.07 (m, 4H), 6.74 (s,1H), 6.69–6.42 (m, 6H), 4.11 (s, 3H), 3.79–3.48 (m, 8H), 3.51–2.86 (m, 6H); 13C NMR (50 MHz, DMSO-d6) δ 180.47, 170.26, 168.54, 159.65, 159.61, 154.82, 154.78, 152.92, 151.96, 151.86, 149.68, 148.89, 145.26, 145.00, 141.27, 140.25, 136.60, 131.30, 129.58, 129.47, 128.97, 126.77, 126.27, 124.04, 118.95, 112.70, 112.67, 112.53, 112.19, 109.77, 102.21, 102.19, 98.68, 69.62, 69.52, 69.07, 68.46, 45.70, 43.69, 41.71, 41.03; UV-VIS λmax 524 nm; Fluorescence λem 537; ES-MS (methanol) m/z: 952.4 [M+H]+, 974.4 [M+Na]+, 990.4 [M+K]+. HRMS (ESI-TOF) m/z: [M+H]+ Calcd for C47H41N11O10S 952.2831; Found 952.2833.

3.1.20. 5-(3-(1-(4-(3-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e] [1,2,4]triazolo[1,5-c]pyrimidin-5-yl)ureido)phenyl)-2-oxo-7,10,13-trioxa-3-azahexadecan-16-yl)thioureido)-2-(6-hydroxy-3-oxo-3H-xanthen-9-yl)benzoate (18)

Yield 63%; orange solid; mp 143–149 °C; 1H NMR (200 MHz, DMSO-d6) δ 10.65 (bs, 1H), 10.36–9.87 (m, 3H), 9.51 (bs, 1H), 8.78 (s, 1H), 8.34–8.11 (m, 2H), 7.99 (s, 2H), 7.74 (d, J = 8.2 Hz, 1H), 7.49 (d, J = 7.8 Hz, 2H), 7.37–7.03 (m, 4H), 6.76 (s, 1H), 6.69–6.45 (m, 6H), 4.13 (s, 3H), 3.65–3.18 (m, 14H), 3.18–2.79 (m, 4H), 1.81 (bs, 2H), 1.63 (bs, 2H); 13C NMR (50 MHz, DMSO-d6) δ 180.23, 169.94, 168.39, 159.33, 159.27, 154.82, 151.74, 149.01, 148.96, 148.89, 145.34, 144.91, 141.28, 139.69, 136.40, 131.51, 129.47, 129.38, 129.27, 129.22, 128.92, 128.77, 126.34, 123.87, 118.94, 112.66, 112.49, 112.47, 112.39, 112.21, 112.15, 109.43, 102.16, 98.78, 69.72, 69.53, 69.51, 68.13, 68.01, 41.82, 41.35, 40.62, 35.97, 29.31, 28.58, 9.10; UV-VIS λmax 524 nm; Fluorescence λem 537; ES-MS (methanol) m/z: 1046.4 [M+Na]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C51H49N11O11S 1026.3226; Found 1026.3223.

3.1.21. General synthesis for the N-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)alkynamides (19–22)

Alkyne acid (1.175 mmol) (31–24) was dissolved in deuterated chloroform (2 mL) and a catalytic amount of DMF was added (200 μL). The solution was cooled to 0 °C and oxalyl dichloride was added dropwise (1.175 mmol, 99 μL) and the reaction was stirred at room temperature for 3 h. A little amount of reaction was diluted in deuterated chloroform and monitored by 1H NMR. The peaks of desired acyl chlorides (35–38) were shifted from those of the corresponding carboxylic acids. When the conversion was complete, to the reaction were added 1,4-dioxane (5 mL), triethylamine(1.959 mmol, 273 μL) and the 5-amino-pyrazolo-triazolo-pyrimidine 30 (0.392 mmol, 100 mg). The reaction was stirred at reflux overnight (TLC ethyl acetate 9.5/methanol 0.5) then the solvent was removed under reduced pressure. The residue was dissolved in water and extracted with dichloromethane (3 times), The organic layers were collected and dried over sodium sulfate anhydrous and the solvent removed under reduced pressure. The crude was purified on flash silica column chromatography (ethyl acetate 9.8/methanol 0.2) to afford the desired compound (19–22).

3.1.22. N-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4] triazolo[1,5-c]pyrimidin-5-yl)pent-4-ynamide (19)

Yield 40%; pale brown solid; mp 264–271 °C d; 1H NMR (200 MHz, CDCl3) δ 9.11 (s, 1H), 8.22 (s, 1H), 7.65 (s, 1H), 7.23 (s, 1H), 6.61 (s, 1H), 4.20 (s, 3H), 3.50 (t, J = 6.8 Hz, 2H), 2.70 (t, J = 6.8 Hz, 2H), 2.01 (s, 2H); 13C NMR (50 MHz, CDCl3) δ 171.47, 155.68, 153.84, 148.92, 145.09, 144.81, 138.67, 125.55, 111.13, 112.24, 100.04, 82.89, 74.39, 69.33, 41.06, 37.05, 13.86. ES-MS (methanol) m/z: 336.2 [M+H]+, 358.1 [M+Na]+, 374.1 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C16H13N7O2 358.1023; Found 358.1021.

3.1.23. N-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4] triazolo[1,5-c]pyrimidin-5-yl)hex-5-ynamide (20)

Yield 19%; pale yellow solid; mp 210–216 °C d; 1H NMR (200 MHz, CDCl3) δ 9.05 (s, 1H), 8.21 (s, 1H), 7.65 (s, 1H), 7.26 (s, 1H), 6.62 (s, 1H), 4.21 (s, 3H), 3.26 (t, J = 6.3 Hz, 2H), 2.40 (s, 2H), 2.18–1.85 (m, 3H); 13C NMR (50 MHz, CDCl3) δ 172.67, 156.07, 153.55, 149.05, 145.19, 144.92, 138.68, 124.95, 112.92, 112.16, 100.13, 83.49, 69.33, 40.80, 36.69, 23.34, 18.04. ES-MS (methanol) m/z: 372.2 [M+Na]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C17H15N7O2 372.1179; Found 372.1182.

3.1.24. N-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4] triazolo[1,5-c]pyrimidin-5-yl)hept-6-ynamide (21)

Yield 15%; white solid; mp 208–215 °C d; 1H NMR (200 MHz, CDCl3) δ 9.02 (s, 1H), 8.21 (s, 1H), 7.65 (s, 1H), 7.23 (d, J = 3.2 Hz, 1H), 6.60 (d, J = 3.2 Hz 1H), 4.20 (s, 3H), 3.36 (t, J = 7 Hz, 2H), 2.32–2.24 (m, 2H), 1.97–1.86 (m, 3H), 1.77–1.66 (m, 2H); 13C NMR (50 MHz, CDCl3) δ 173.09, 155.76, 153.92, 148.96, 144.96, 138.75, 125.38, 112.89, 112.19, 100.05, 84.24, 68.76, 41.03, 37.23, 28.01, 23.52, 18.56. ES-MS (methanol) m/z: 364.2 [M+H]+, 386.2 [M+Na]+, 402.1 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C18H17N7O2 386.1336; Found 386.1331.

3.1.25. N-(2-(furan-2-yl)-8-methyl-8H-pyrazolo[4,3-e][1,2,4] triazolo[1,5-c]pyrimidin-5-yl)oct-7-ynamide (22)

Yield 40%; white solid; mp 204–208 °C d; 1H NMR (200 MHz, CDCl3) δ 8.99 (s, 1H), 8.21 (s, 1H), 7.65 (d, J = 1.8 Hz, 1H), 7.25 (d, J = 3.2 Hz, 1H), 6.61 (dd, J = 1.8 Hz, J = 3.2 Hz 1H), 4.20 (s, 3H), 3.24 (t, J = 7.2 Hz, 2H), 2.23–2.04 (m, 2H), 1.95–1.73 (m, 3H), 1.61–1.56 (m, 4H); 13C NMR (50 MHz, CDCl3) δ 173.28, 155.74, 153.95, 148.95, 144.96, 138.76, 125.37, 112.88, 112.18, 100.06, 84.57, 68.4, 41.02, 37.62, 28.47, 28.38, 23.91, 18.48. ES-MS (methanol) m/z: 378.2 [M+H]+, 400.2 [M+Na]+, 416.1 [M+K]+. HRMS (ESI-TOF) m/z: [M+Na]+ Calcd for C19H19N7O2 400.1492; Found 400.1488.

3.1.26. Synthesis of 2-(6-amino-3-iminio-4,5-disulfonato-3H-zanthen-9-yl)-5-((6-(4-((8-methyl-2-(furan-2-yl)- 8H-pyrazolo [4,3-e][1,2,4]triazolo[1,5-c]pyrimidin-5-yl)amino)-4-oxobutyl)-1H-1,2,3-triazol-1-yl)hexyl)carbamoyl)benzoate (23)

A solution of Alexa Fluor-488 5-carboxamido-(6-azidohexanyl) bis (triethyammonium salt) (0.5 mg, 0.58 μmol, Invitrogen-Life Technologies, Grand Island, N.Y.) in water (200 μl) was added to a solution of alkyne derivative 20 (0.38 mg, 1.1 μmol) in DMF (100 μl). A freshly prepared solution of aqueous sodium ascorbate (0.1 M, 8 μl, 5.86 μmol) was added to the reaction mixture followed by the addition of aqueous copper sulfate pentahydrate solution (0.1 M, 2.94 μl, 0.29 μmol): The resulting mixture was stirred overnight at room temperature; 10 mM aqueous solution of triethylammonium acetate buffer (2 mL) was added with constant mixing. The resulting mixture was lyophilized and purified by HPLC with a Luna 5μ RP.C18 semipreparative column (250 × 10 mm: Phenomenex, Torrance, Calif.) under the following conditions: flow rate of 2 mL/min; H2O–MeCN from 100:0 (v/v) to 0:100 (v/v) in 30 min and isolated at 20.07 min to get the Alexa Fluor-488 conjugate 23 (0.16 mg, 28%). Analytical purity >99% by HPLC (retention time 10.13 min).

3.2. Biology

3.2.1. Radioligand binding to hA1, hA2A and hA3 ARs

[3H]R-PIA and [125I]AB-MECA were utilized in radioligand binding assays to membranes prepared from CHO cells expressing recombinant hA1and hA3 ARs, respectively, as previously described [40]. While, [3H]CGS21680 binding was used at hA2AAR expressed in HEK293 cells. ADA (3 units/mL) was present during the preparation of the membranes, in a preincubation of 30 min at 30 °C, and during the incubation with the radioligands. All non-radioactive compounds were initially dissolved in DMSO and diluted with buffer to the final concentration, where the amount of DMSO never exceeded 2%. Incubations were terminated by rapid filtration over Whatman GF/B filters, using a Brandell cell harvester (Brandell, Gaithersburg, MD). The tubes were rinsed three times with 3 mL of buffer each. At least six different concentrations of competitor, spanning 3 orders of magnitude adjusted appropriately for the IC50 of each compound, were used. IC50 values, calculated with the nonlinear regression method implemented in Graph-Pad (Prism, San Diego, CA), were converted to Ki values as described [40,43].

3.3. Computational studies

The MOE suite [45] was used for general molecular modeling operations, including logP(o/w) calculations as well. Preliminary computations were carried out on a 12 CPU (Intel® Xeon® CPU E5–1650 3.80 GHz) Linux workstation. SuMD simulations were performed with the ACEMD engine [46] on NVIDIA drivers: NVIDIA GTX 980Ti and NVIDIA GTX 980. The CHARMM36/CHARMM36 general force field (CGenFF) combination was adopted. The GOLD [47] suite (goldscore scoring function) was used as docking program. Ligands’ 3D structures were constructed by the MOE-builder tool. Ionization states were predicted using the MOE-protonate 3D tool. Tautomerization and atom hybridization were checked. Structures were minimized by the MMFF94x until the root mean square (RMS) gradient was below 0.05 kcal mol−1 A−1. For each ligand, 20 docking simulation runs were performed on each receptor subtype, searching on a sphere of 20 Å radius, centered on the backbone nitrogen of the conserved Asn (6.55). Ligand and protein partial charges were calculated by means of the MMFF94 and AMBER14:EHT [48] force fields, respectively. The hA2A and hA1AR coordinates were retrieved from the Protein Data Bank [49] using the crystal structures 3PWH [50] and 5UEN [51], respectively. For the hA3AR a homology model was used, basing on our docking/structural based model assessment [39]. All molecular docking studies have been carried out with a sodium ion and its first hydration shell in all adenosine receptor structures, according to our previous studies [33]. The Ballesteros–Weinstein [52] numbering system was used sometimes to indicate conserved residues. IE electrostatic and van der Waals energy contributions to the binding energy were calculated by MOE, along with per residue electrostatic and hydrophobic interactions. Per-residue information was reported in the “Interaction Energy Fingerprints”: they are heat maps reporting the strength of the interaction of each residue (x-axis) and each ligand (y-axis) according to a colorimetric scale going from blue to red for negative to positive values in the case of electrostatic contributions and from white to dark green for low to high values for hydrophobic contributions. Interactions of most relevant residues were also reported by histograms, whose height is proportional to the strength of the interaction. All the plots were generated using Gnuplot 4.6 [53], except Fig. 3. Molecular graphics were performed with the UCSF Chimera package [54], except Fig. 8, which was obtained using the VMD program (version 1.9.3) [55]. The in-house MMsDocking video maker tool was exploited to produce videos showing the docking poses, per residue hydrophobic and electrostatic contributions for selected residues, experimental binding data, and scoring values. Representations of docking poses were produced using the UCSF Chimera package, 2D depictions were constructed by the cheminformatics toolkit RDKit [56] and the heat maps were obtained by Gnuplot 4.6; in the end, videos were mounted using MEncoder. Ligand 3 force field parameters for MD simulations were initially retrieved from the Paramchem web service and then deeply optimized in concordance with CGenFF, at the MP2/6–31G [57] level of the theory by using Gaussian 09 [58] and RESP partial charges. Systems were embedded in a 1-palmitoyl-2oleyl-sn-glycerol-3-phospho-choline (POPC) lipid bilayer, according to the pre-orientation provided by the Orientations Proteins in Membrane (OPM) database [59] and by using the VMD membrane builder plugin. Lipids within 0.4 Å from the protein were removed and TIP3P [60] model water molecules were added to solvate the system by means of Solvate 1.0 [61]. Charge neutrality of the system was obtained by adding Na+/Cl counterions to a final concentration of 0.154 M. System was equilibrated through a three-step procedure. In the first step, 1500 conjugate-gradient minimization steps were applied in order to reduce the clashes between protein and lipids. Then, a 5 ns long MD simulation was performed in the NPT ensemble, with a positional constraint of 1 kcal mol−1 Å−2 on ligand, protein, and lipid phosphorus atoms. During the second stage, 10 ns MD simulation in the NPT ensemble was performed constraining all the protein and ligand atoms but leaving the POPC residues free to diffuse in the bilayer. In the last equilibration stage, positional constraints were applied only to the ligand and protein backbone alpha carbons for further 5 ns MD simulation. All the MD simulations were performed using: (1) an integration time step of 2 fs; (2) the Berendsen barostat [62] to maintain the system pressure at 1 atm; (3) the Langevin thermostat [63] to maintain the temperature at 310 K with a low dumping of 1 ps− 1; (4) the M-SHAKE algorithm [64] to constrain the bond lengths involving hydrogen atoms; and (5) a long-range cutoff of 10 Å. According to the SuMD approach [65,66], the timescale needed to reproduce binding is in the range of nanoseconds, instead of hundreds of nanoseconds or microseconds usually necessary with unsupervised MD. Sampling is performed by a tabu-like algorithm to monitor the distance between the centers of mass of the ligand and the binding site during short unbiased MD simulations. SuMD considers the ligand atoms and the atoms of user-defined protein residues to monitor the distance between the centers of mass of the binder and the binding site. A series of 600 ps unbiased MD simulations are performed and after each simulation, the distance points collected at regular intervals are fitted into a linear function. If the resulting slope is negative the next simulation step starts from the last set of coordinates and velocities produced, otherwise the simulation is restarted by randomly assigning the atomic velocities. Short simulations are perpetuated under the supervision until the distance between the ligand and the binding site goes below 5 Å, then, the supervision is disabled and a classical MD simulation is performed. For the orthosteric center of mass, we considered hA3AR residues Asn 250, Phe 168, His 272, and Ser 247. NAMD energy Plugin 1.4 [67] was used to calculate ligand-protein interaction energies.

4. Conclusions

In this work we reported a new series of hA3AR conjugable derivatives that exhibited a very good selectivity towards hA1AR and a discrete selectivity versus the hA2AAR. We have developed both amino compounds, which could be functionalized by reactions with carboxylic acids or isothiocyanates, and alkyne derivatives, which are useful precursors for the click reaction. Unfortunately, attempts to use these compounds to obtain selective functionalized hA3AR ligands led mainly to dual hA2A/hA3 AR ligands, revealing that the functional moiety had an active role in the establishment of ligand-receptor interactions. Moreover, molecular docking and SuMD simulations highlighted that non-conservative residues, located at the ELs 2 and 3 (Gln 167, Val169, Gln261, Glu258 of the hA3AR), constitute a specific structural region able to discriminate ligands depending on their chemical and partition properties. These data can, thus, pave the way for a new understanding of selectivity towards ARs, leading finally to the rational design of highly selective ligands. In any case, among the functionalized compounds, Alexa Fluor 488 derivative MRS5763 (23) showed the best hA3A affinity and selectivity profile, which, by further computer-guided optimization, could lead to a selective hA3AR probe useful as fluorescent tool for this receptor.

Supplementary Material

SI file
Video1
Download video file (6.6MB, mp4)
Video2
Download video file (6.7MB, mp4)
Video3
Download video file (7MB, mp4)
Video4
Download video file (13.7MB, mp4)

Acknowledgments

KAJ thanks the NIDDK Intramural Research Program (ZIADK31117) for support.

Abbreviations:

AR

Adenosine receptor

BOC

Tert-butyloxycarbonyl

EDCI

1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide

FITC

Fluorescein isothiocyanate

PTP

pyrazolo[4,3-e]-1,2,4-triazolo[1,5-c]pyrimidine

Footnotes

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ejmech.2019.111886.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  • [1].Sawynok J, Adenosine receptor targets for pain, Neuroscience 338 (2016) 1–18, 10.1016/j.neuroscience.2015.10.031. [DOI] [PubMed] [Google Scholar]
  • [2].Guerrero A, A2A adenosine receptor agonists and their potential therapeutic applications. An update, Curr. Med. Chem 25 (2018) 3597–3612, 10.2174/0929867325666180313110254. [DOI] [PubMed] [Google Scholar]
  • [3].Hocher B, Adenosine A1 receptor antagonists in clinical research and development, Kidney Int. 78 (2010) 438–445, 10.1038/ki.2010.204. [DOI] [PubMed] [Google Scholar]
  • [4].Cacciari B, Spalluto G, Federico S, A2A adenosine receptor antagonists as therapeutic candidates: are they still an interesting challenge? Mini Rev. Med. Chem 18 (2018) 1168–1174, 10.2174/1389557518666180423113051. [DOI] [PubMed] [Google Scholar]
  • [5].Haskó G, Csóka B, Németh ZH, Vizi ES, Pacher P, A2B adenosine receptors in immunity and inflammation, Trends Immunol. 30 (2009) 263–270, 10.1016/j.it.2009.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Cheong SL, Federico S, Venkatesan G, Mandel AL, Shao Y-M, Moro S, Spalluto G, Pastorin G, The A3 adenosine receptor as multifaceted therapeutic target: pharmacology, medicinal chemistry, and in silico approaches, Med. Res. Rev 33 (2013) 235–335, 10.1002/med.20254. [DOI] [PubMed] [Google Scholar]
  • [7].Fredholm BB, Arslan G, Halldner L, Kull B, Schulte G, Wasserman W, Structure and function of adenosine receptors and their genes, Naunyn . Schmiedebergs. Arch. Pharmacol 362 (2000) 364–374, 10.1007/s002100000313. [DOI] [PubMed] [Google Scholar]
  • [8].Borea PA, Varani K, Vincenzi F, Baraldi PG, Tabrizi MA, Merighi S, Gessi S, The A3 adenosine receptor: history and perspectives, Pharmacol. Rev 67 (2014) 74–102, 10.1124/pr.113.008540. [DOI] [PubMed] [Google Scholar]
  • [9].Moro S, Gao Z, Jacobson KA, Spalluto G, Progress in the pursuit of therapeutic adenosine receptor antagonists, Med. Res. Rev 26 (2006) 131–159, 10.1002/med.20048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Redenti S, Ciancetta A, Pastorin G, Cacciari B, Moro S, Spalluto G, Federico S, Pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidines and structurally simplified analogs. Chemistry and sar profile as adenosine receptor antagonists, Curr. Top. Med. Chem 16 (2016) 3224–3257, 10.2174/1568026616666160506145831. [DOI] [PubMed] [Google Scholar]
  • [11].Baraldi PG, Cacciari B, Romagnoli R, Spalluto G, Klotz K, Leung E, Varani K, Gessi S, Merighi S, Borea PA, Pyrazolo[4,3-e]-1,2,4-triazolo[1,5-c]pyrimidine Derivatives as highly potent and selective human A3 adenosine receptor antagonists, J. Med. Chem 42 (1999) 4473–4478, 10.1021/jm991114s. [DOI] [PubMed] [Google Scholar]
  • [12].Baraldi PG, Cacciari B, Romagnoli R, Spalluto G, Moro S, Klotz K, Leung E, Varani K, Gessi S, Merighi S, Borea PA, Pyrazolo[4,3-e]1,2,4-triazolo[1,5-c] pyrimidine derivatives as highly potent and selective human A3 adenosine receptor antagonists: influence of the chain at the N8 pyrazole nitrogen, J. Med. Chem 43 (2000) 4768–4780, 10.1021/jm001047y. [DOI] [PubMed] [Google Scholar]
  • [13].Baraldi PG, Cacciari B, Moro S, Spalluto G, Pastorin G, Da Ros T, Klotz K, Varani K, Gessi S, Borea PA, Synthesis, biological activity, and molecular modeling investigation of new pyrazolo[4,3-e]-1,2,4-triazolo[1,5-c]pyrimidine derivatives as human A3 adenosine receptor antagonists, J. Med. Chem 45 (2002) 770–780, 10.1021/jm0109614. [DOI] [PubMed] [Google Scholar]
  • [14].Maconi A, Moro S, Pastorin G, Da Ros T, Spalluto G, Gao Z-G, Jacobson KA, Baraldi PG, Cacciari B, Varani K, Andrea Borea P, Synthesis, biological properties, and molecular modeling investigation of the first potent, selective, and water-soluble human A3 adenosine receptor antagonist, J. Med. Chem 45 (2002) 3579–3582, 10.1021/jm020974x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Federico S, Ciancetta A, Sabbadin D, Paoletta S, Pastorin G, Cacciari B, Klotz KN, Moro S, Spalluto G, Exploring the directionality of 5-substitutions in a new series of 5-alkylaminopyrazolo[4,3-e]1,2,4-triazolo[1,5-c]pyrimidine as a strategy to design novel human A3 adenosine receptor antagonists, J. Med. Chem 55 (2012), 10.1021/jm300899q. [DOI] [PubMed] [Google Scholar]
  • [16].Pastorin G, Da Ros T, Bolcato C, Montopoli C, Moro S, Cacciari B, Baraldi PG, Varani K, Borea PA, Spalluto G, Synthesis and biological studies of a new series of 5-heteroarylcarbamoylaminopyrazolo[4,3-e]1,2,4-triazolo[1,5-c]pyrimidines as human A3 adenosine receptor antagonists. Influence of the heteroaryl substituent on binding affinity and molecular modeling in, J. Med. Chem 49 (2006) 1720–1729, 10.1021/jm051147+. [DOI] [PubMed] [Google Scholar]
  • [17].Baraldi PG, Cacciari B, Dionisotti S, Egan J, Spalluto G, Zocchi C, Synthesis of the tritium labeled SCH 58261, a new non-xanthine A2A adenosine receptor antagonist, J. Label. Comp. Radiopharm 38 (1996) 725–732, . [DOI] [Google Scholar]
  • [18].Shryock JC, Snowdy S, Baraldi PG, Cacciari B, Spalluto G, Monopoli A, Ongini E, Baker SP, Belardinelli L, A2A -adenosine receptor reserve for coronary vasodilation, Circulation 98 (1998) 711–718, 10.1161/01.CIR.98.7.711. [DOI] [PubMed] [Google Scholar]
  • [19].Todde S, Moresco RM, Simonelli P, Baraldi PG, Cacciari B, Spalluto G, Varani K, Monopoli A, Matarrese M, Carpinelli A, Magni F, Kienle MG, Fazio F, Design, Radiosynthesis, and biodistribution of a new potent and selective ligand for in vivo imaging of the adenosine A2A receptor system using positron emission tomography, J. Med. Chem 43 (2000) 4359–4362, 10.1021/jm0009843. [DOI] [PubMed] [Google Scholar]
  • [20].Baraldi PG, Cacciari B, Romagnoli R, Varani K, Merighi S, Gessi S, Borea PA, Leung E, Hickey SL, Spalluto G, Synthesis and preliminary biological evaluation of [3H]-MRE 3008-F20: the first high affinity radioligand antagonist for the human A3 adenosine receptors, Bioorg. Med. Chem. Lett 10 (2000) 209–211, 10.1016/S0960-894X(99)00674-5. [DOI] [PubMed] [Google Scholar]
  • [21].Baraldi PG, Cacciari B, Moro S, Romagnoli R, Ji X, Jacobson KA, Gessi S, Borea PA, Spalluto G, Fluorosulfonyl- and bis-(b-chloroethyl)amino-phenylamino functionalized pyrazolo[4,3-e]1,2,4-triazolo[1,5-c]pyrimidine derivatives: irreversible antagonists at the human A3 adenosine receptor and molecular modeling studies, J. Med. Chem 44 (2001) 2735–2742, 10.1021/jm010818a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Federico S, Spalluto G, Functionalized ligands targeting G protein-coupled adenosine receptors, Future Med. Chem 11 (2019) 1673–1677, 10.4155/fmc-2019-0133. [DOI] [PubMed] [Google Scholar]
  • [23].Stoddart LA, Kilpatrick LE, Briddon SJ, Hill SJ, Probing the pharmacology of G protein-coupled receptors with fluorescent ligands, Neuropharmacology 98 (2015) 48–57, 10.1016/j.neuropharm.2015.04.033. [DOI] [PubMed] [Google Scholar]
  • [24].Sosnovik DE, Weissleder R, Emerging concepts in molecular MRI, Curr. Opin. Biotechnol (2007), 10.1016/j.copbio.2006.11.001. [DOI] [PubMed] [Google Scholar]
  • [25].Nabi B, Rehman S, Khan S, Baboota S, Ali J, Ligand conjugation: an emerging platform for enhanced brain drug delivery, Brain Res. Bull 142 (2018) 384–393, 10.1016/j.brainresbull.2018.08.003. [DOI] [PubMed] [Google Scholar]
  • [26].Srinivasarao M, Low PS, Ligand-targeted drug delivery, Chem. Rev 117 (2017) 12133–12164, 10.1021/acs.chemrev.7b00013. [DOI] [PubMed] [Google Scholar]
  • [27].Haag R, Kratz F, Polymer therapeutics: concepts and applications, Angew. Chem. Int. Ed 45 (2006) 1198–1215, 10.1002/anie.200502113. [DOI] [PubMed] [Google Scholar]
  • [28].Morphy R, Rankovic Z, Designed multiple ligands. an emerging drug discovery paradigm, J. Med. Chem 48 (2005) 6523–6543, 10.1021/jm058225d. [DOI] [PubMed] [Google Scholar]
  • [29].Tosh DK, Yoo LS, Chinn M, Hong K, Kilbey SM, Barrett MO, Fricks IP, Harden TK, Gao ZG, Jacobson KA, Polyamidoamine (PAMAM) dendrimer conjugates of “clickable” agonists of the A3 adenosine receptor and coactivation of the P2Y14 receptor by a tethered nucleotide, Bioconjug. Chem 21 (2010) 372–384, 10.1021/bc900473v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Ciruela F, Fernández-Dueñas V, Jacobson KA, Lighting up G protein-coupled purinergic receptors with engineered fluorescent ligands, Neuropharmacology 98 (2015) 58–67, 10.1016/j.neuropharm.2015.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Markovic M, Ben-Shabat S, Keinan S, Aponick A, Zimmermann EM, Dahan A, Prospects and challenges of phospholipid-based prodrugs, Pharmaceutics 10 (2018) 1–17, 10.3390/pharmaceutics10040210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [32].Jacobson KA, Gpcrs PR, Jacobson KA, Functionalized congener approach to the design of ligands for G Protein-coupled receptors (GPCRs), Bioconjug. Chem 20 (2009) 1816–1835, 10.1021/bc9000596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Federico S, Margiotta E, Paoletta S, Kachler S, Klotz K-N, Jacobson KA, Pastorin G, Moro S, Spalluto G, Pyrazolo[4,3-e][1,2,4]triazolo[1,5-c]pyrimidines to develop functionalized ligands to target adenosine receptors: fluorescent ligands as an example, Medchemcomm 10 (2019) 1094–1108, 10.1039/C9MD00014C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Fishman P, Bar-Yehuda S, Ardon E, Rath-Wolfson L, Barrer F, Ochaion A, Madi L, Targeting the A3 adenosine receptor for cancer therapy: inhibition of prostate carcinoma cell growth by A3AR agonist, Anticancer Res. 23 (2003) 2077–2083. [PubMed] [Google Scholar]
  • [35].Ohana G, Bar-Yehuda S, Arich A, Madi L, Dreznick Z, Rath-Wolfson L, Silberman D, Slosman G, Fishman P, Inhibition of primary colon carcinoma growth and liver metastasis by the A3 adenosine receptor agonist CF101, Br. J. Canc 89 (2003) 1552–1558, 10.1038/sj.bjc.6601315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Madi L, Bar-Yehuda S, Barer F, Ardon E, Ochaion A, Fishman P, A3 adenosine receptor activation in melanoma cells, J. Biol. Chem 278 (2003) 42121–42130, 10.1074/jbc.M301243200. [DOI] [PubMed] [Google Scholar]
  • [37].Madi L, Ochaion A, Rath-Wolfson L, Bar-Yehuda S, Erlanger A, Ohana G, Harish A, Merimski O, Barer F, Fishman P, The A3 adenosine receptor is highly expressed in tumor versus normal cells: potential target for tumor growth inhibition, Clin. Cancer Res 10 (2004) 4472–4479, 10.1158/1078-0432.CCR-03-0651. [DOI] [PubMed] [Google Scholar]
  • [38].Gessi S, Cattabriga E, Avitabile A, Gafa R, Lanza G, Cavazzini L, Bianchi N, Gambari R, Feo C, Liboni A, Gullini S, Leung E, Mac-Lennan S, Borea PA, Elevated expression of A3 adenosine receptors in human colorectal cancer is reflected in peripheral blood cells, Clin. Cancer Res 10 (2004) 5895–5901, 10.1158/1078-0432.CCR-1134-03\10/17/5895 ([pii]). [DOI] [PubMed] [Google Scholar]
  • [39].Margiotta E, Moro S, A comparison in the use of the crystallographic structure of the human A1 or the A2A adenosine receptors as a template for the construction of a homology model of the A3 subtype, Appl. Sci 9 (2019) 821, 10.3390/app9050821. [DOI] [Google Scholar]
  • [40].Kozma E, Kumar TSS, Federico S, Phan K, Balasubramanian R, Gao Z-G, Paoletta S, Moro S, Spalluto G, Jacobson KA, Novel fluorescent antagonist as a molecular probe in A3 adenosine receptor binding assays using flow cytometry, Biochem. Pharmacol 83 (2012) 1552–1561, 10.1016/j.bcp.2012.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Jacobson KA, Thatikonda KS, Kozma EE, Spalluto G, Moro S, Federico S, Fluorescent antagonists of the A3 adenosine receptor, 9,227,979 B2. http://patft.uspto.gov/netacgi/nph-Parser?Sect1=PTO2&Sect2=HITOFF&p=&u=%2Fnetahtml%2FPTO%2Fsearch-adv.htm&r=1&f=1G&l=50&d=PTXT&S1=9,227,979&OS=+9,227,979&RS=9,227,979, 2016. (accessed March 28, 2019). [Google Scholar]
  • [42].Moro S, Braiuca P, Deflorian F, Ferrari C, Pastorin G, Cacciari B, Baraldi PG, Varani K, Borea PA, Spalluto G, Combined target-based and ligand-based drug design approach as a tool to define a novel 3D-pharmacophore model of human A3 adenosine receptor antagonists: pyrazolo[4,3-e]1,2,4-triazolo [1,5-c]pyrimidine derivatives as a key study, J. Med. Chem 48 (2005) 152–162, 10.1021/jm049662f. [DOI] [PubMed] [Google Scholar]
  • [43].Gao ZG, Blaustein JB, Gross AS, Melman N, Jacobson KA, N6-Substituted adenosine derivatives: selectivity, efficacy, and species differences at A3 adenosine receptors, Biochem. Pharmacol 65 (2003) 1675–1684, 10.1016/S0006-2952(03)00153-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Margiotta E, Deganutti G, Moro S, Could the presence of sodium ion influence the accuracy and precision of the ligand-posing in the human A2A adenosine receptor orthosteric binding site using a molecular docking approach? Insights from Dockbench, J. Comput. Aided Mol. Des 32 (2018) 1337–1346, 10.1007/s10822-018-0174-2. [DOI] [PubMed] [Google Scholar]
  • [45].Chemical Computing Group (CCG) Inc., Molecular Operating Environment (MOE), 2016. [Google Scholar]
  • [46].MacKerell AD, Banavali N, Foloppe N, Development and current status of the CHARMM force field for nucleic acids, Biopolymers 56 (2000) 257–265, . [DOI] [PubMed] [Google Scholar]
  • [47].The Cambridge Crystallographic Data Centre (CCDC), Genetic optimization for ligand docking (GOLD). https://www.ccdc.cam.ac.uk/solutions/csd-discovery/components/gold/. [Google Scholar]
  • [48].Wang J, Wolf RM, Caldwell JW, Kollman PA, Case DA, Development and testing of a general amber force field, J. Comput. Chem 25 (2004) 1157–1174, 10.1002/jcc.20035. [DOI] [PubMed] [Google Scholar]
  • [49].Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE, The protein Data Bank, Nucleic Acids Res. 28 (2000) 235–242, 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Doré AS, Robertson N, Errey JC, Ng I, Hollenstein K, Tehan B, Hurrell E, Bennett K, Congreve M, Magnani F, Tate CG, Weir M, Marshall FH, Structure of the adenosine A2A receptor in complex with ZM241385 and the xanthines XAC and caffeine, Structure 19 (2011) 1283–1293, 10.1016/j.str.2011.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Glukhova A, Thal DM, Nguyen AT, Vecchio EA, Jörg M, Scammells PJ,May LT, Sexton PM, Christopoulos A, Structure of the adenosine A1 receptor reveals the basis for subtype selectivity, Cell 168 (2017) 867–877, 10.1016/j.cell.2017.01.042, e13. [DOI] [PubMed] [Google Scholar]
  • [52].Ballesteros JA, Weinstein H, Integrated methods for the construction of three-dimensional models and computational probing of structure-function relations in G protein-coupled receptors, Methods Neurosci. 25 (1995) 366–428, 10.1016/S1043-9471(05)80049-7. [DOI] [Google Scholar]
  • [53].(n.d. Gnuplot http://www.gnuplot.info/download.html.
  • [54].Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin TE, UCSF Chimera - a visualization system for exploratory research and analysis, J. Comput. Chem 25 (2004) 1605–1612, 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
  • [55].Humphrey W, Dalke A, Schulten K, VMD: visual molecular dynamics, J. Mol. Graph 14 (27) (1996) 33–38, 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
  • [56].RDKit, Cheminformatics and machine learning software, n.d, http://www.rdkit.org/.
  • [57].Head-Gordon M, Pople JA, Frisch MJ, MP2 energy evaluation by direct methods, Chem. Phys. Lett 153 (1988) 503–506, 10.1016/0009-2614(88)85250-3. [DOI] [Google Scholar]
  • [58].Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Scalmani G, Barone V, Petersson GA, Nakatsuji H, Li X, Caricato M, Marenich AV, Bloino J, Janesko BG, Gomperts R, Mennucci B, Hratchian HP, Ortiz JV, Izmaylov AF, Sonnenberg JL, Williams-Young D, Ding F, Lipparini F, Egidi F, Goings J, Peng B, Petrone A, Henderson T, Ranasinghe D, Zakrzewski VG, Gao J, Rega N, Zheng G, Liang W, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Vreven T, Throssell K, Montgomery JAJ, Peralta JE, Ogliaro F, Bearpark MJ, Heyd JJ, Brothers EN, Kudin KN, Staroverov VN, Keith TA, Kobayashi R, Normand J, Raghavachari K, Rendell AP, Burant JC, Iyengar SS, Tomasi J, Cossi M, Millam JM, Klene M, Adamo C, Cammi R, Ochterski JW, Martin RL, Morokuma K, Farkas O, Foresman JB, Fox DJ, Gaussian 16 (2016). [Google Scholar]
  • [59].Lomize MA, Lomize AL, Pogozheva ID, Mosberg HI, OPM: orientations of proteins in membranes database, Bioinformatics 22 (2006) 623–625, 10.1093/bioinformatics/btk023. [DOI] [PubMed] [Google Scholar]
  • [60].Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML, Comparison of simple potential functions for simulating liquid water, J. Chem. Phys 79 (1983) 926–935, 10.1063/1.445869. [DOI] [Google Scholar]
  • [61].Grubmuller H, Groll V, Solvate | max planck institute for biophysical chemistry. https://www.mpibpc.mpg.de/grubmueller/solvate, 1996. accessed August 7, 2019. [Google Scholar]
  • [62].Berendsen HJC, Postma JPM, van Gunsteren WF, DiNola A, Haak JR, Molecular dynamics with coupling to an external bath, J. Chem. Phys 81 (1984) 3684–3690, 10.1063/1.448118. [DOI] [Google Scholar]
  • [63].Loncharich RJ, Brooks BR, Pastor RW, Langevin dynamics of peptides: the frictional dependence of isomerization rates of N-acetylalanyl-N-methyl-amide, Biopolymers 32 (1992) 523–535, 10.1002/bip.360320508. [DOI] [PubMed] [Google Scholar]
  • [64].Essmann U, Perera L, Berkowitz ML, Darden T, Lee H, Pedersen LG, A smooth particle mesh Ewald method, J. Chem. Phys 103 (1995) 8577–8593, 10.1063/1.470117. [DOI] [Google Scholar]
  • [65].Sabbadin D, Ciancetta A, Moro S, Bridging molecular docking to membrane molecular dynamics to investigate GPCReligand recognition: the human A2A adenosine receptor as a key study, J. Chem. Inf. Model 54 (2014) 169–183, 10.1021/ci400532b. [DOI] [PubMed] [Google Scholar]
  • [66].Cuzzolin A, Sturlese M, Deganutti G, Salmaso V, Sabbadin D, Ciancetta A,Moro S, Deciphering the complexity of ligandeprotein recognition pathways using supervised molecular dynamics (SuMD) simulations, J. Chem. Inf. Model 56 (2016) 687–705, 10.1021/acs.jcim.5b00702. [DOI] [PubMed] [Google Scholar]
  • [67].Phillips JC, Braun R, Wang W, Gumbart J, Tajkhorshid E, Villa E, Chipot C, Skeel RD, Kalé L, Schulten K, Scalable molecular dynamics with NAMD, J. Comput. Chem 26 (2005) 1781–1802, 10.1002/jcc.20289. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI file
Video1
Download video file (6.6MB, mp4)
Video2
Download video file (6.7MB, mp4)
Video3
Download video file (7MB, mp4)
Video4
Download video file (13.7MB, mp4)

RESOURCES