Next Article in Journal
Antidiabetic Drugs Can Reduce the Harmful Impact of Chronic Smoking on Post-Traumatic Brain Injuries
Next Article in Special Issue
The Effect of 8,5′-Cyclo 2′-deoxyadenosine on the Activity of 10-23 DNAzyme: Experimental and Theoretical Study
Previous Article in Journal
ABC Transporters in Bacterial Nanomachineries
Previous Article in Special Issue
A Key Role in Catalysis and Enzyme Thermostability of a Conserved Helix H5 Motif of Human Glutathione Transferase A1-1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enzymatic Synthesis of 2-Chloropurine Arabinonucleosides with Chiral Amino Acid Amides at the C6 Position and an Evaluation of Antiproliferative Activity In Vitro

by
Barbara Z. Eletskaya
1,*,
Maria Ya. Berzina
1,
Ilya V. Fateev
1,
Alexei L. Kayushin
1,
Elena V. Dorofeeva
1,
Olga I. Lutonina
1,
Ekaterina A. Zorina
1,
Konstantin V. Antonov
1,
Alexander S. Paramonov
1,
Inessa S. Muzyka
1,
Olga S. Zhukova
2,
Mikhail V. Kiselevskiy
2,
Anatoly I. Miroshnikov
1,
Roman S. Esipov
1 and
Irina D. Konstantinova
1,*
1
Shemyakin and Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences, Miklukho-Maklaya St. 16/10, 117997 Moscow, Russia
2
State N.N. Blokhin Russian Cancer Research Center, Kashirsky Highway, 24, 115478 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6223; https://doi.org/10.3390/ijms24076223
Submission received: 27 February 2023 / Revised: 13 March 2023 / Accepted: 22 March 2023 / Published: 25 March 2023
(This article belongs to the Special Issue Latest Advances in Enzymology)

Abstract

:
A number of purine arabinosides containing chiral amino acid amides at the C6 position of the purine were synthesized using a transglycosylation reaction with recombinant E. coli nucleoside phosphorylases. Arsenolysis of 2-chloropurine ribosides with chiral amino acid amides at C6 was used for the enzymatic synthesis, and the reaction equilibrium shifted towards the synthesis of arabinonucleosides. The synthesized nucleosides were shown to be resistant to the action of E. coli adenosine deaminase. The antiproliferative activity of the synthesized nucleosides was studied on human acute myeloid leukemia cell line U937. Among all the compounds, the serine derivative exhibited an activity level (IC50 = 16 μM) close to that of Nelarabine (IC50 = 3 μM) and was evaluated as active.

1. Introduction

The involvement of adenosine and its nucleotides into numerous biological processes has led to many works on the synthesis of structural analogues of adenosine. Agonists and antagonists of adenosine receptors [1,2] represent the broadest group of adenosine analogs obtained. Compounds with antitumor activity have been found among A3-type adenosine receptor agonists [3,4]. N6-Allyl, N6-isopropyl, and N6-propargyl analogs have shown to significantly increase the lifespans of mice experiencing mammary carcinoma. The short-chain adenosine analogues are more active in the treatment of animal carcinomas than in leukemia or sarcoma tumor cell systems [5].
Establishment of the structure of the adenylosuccinate (Figure 1) involved in the purine nucleotide cycle prompted researchers to synthesize a series of purine bases substituted at C6 by chiral amino acid residues [6,7,8,9]. Conjugates were synthesized both with amino acids [10,11] and with corresponding methyl esters [10,12]. Conjugates with amino acids have antimycobacterial activity, but no antitumor activity [6,7]. Interestingly, the arabinoside analogue of natural N6-isopentenyladenosine did not show anti-proliferative capacity on T24 human bladder carcinoma cells [13], while the N6-isopentenyladenosine riboside did.
Adenosine analogs lose their biological activity due to rapid degradation by intracellular adenosine deaminase (ADA). For example, the antiviral D-arabinofuranosyladenine (vidarabine or Ara-A) is rapidly deaminated in vivo to the less active D-arabinofuranosylhypoxantine (Ara-H) [14]. Cordycepin (3′-deoxyadenosine) was found to be ineffective as an antibacterial agent due to metabolic degradation by ADA [15].
To enhance the biological activity of the compounds, it was necessary to increase the lifetime of the compound in the cell. Attempts have been made to synthesize ADA-resistant purine nucleosides. The introduction of a halogen atom in the C2 position of purine inhibits the action of intracellular adenosine deaminase or makes the nucleoside completely resistant to deamination [16]. The antitumor drugs Cladribine, Fludarabine, and Clofarabine (Figure 1) are examples of such compounds. They are used in the treatment of oncohematological diseases [17,18,19].
Another way to make nucleosides resistant to the catabolic action of ADA is introducing various substituents at the C6 position of adenosine. However, no unambiguous relationship between the substituent structure and ADA resistance was found. The rate of hydrolysis of adenosines decreased in a series of C6 substituents: hydroxylamino, chlorine, bromine, iodine, methylamino, and ethylamino (Figure 2). Deoxyadenosines with mercapto-, benzylamino-, and p-nitrobenzyl substituents at C6 were resistant to ADA [20,21,22].
On the other hand, ADA is a key intracellular enzyme that modifies the anticancer drug Nelarabine (9-β-D-arabinofuranosyl-6-O-methylguanine, Figure 1). A necessary step in the metabolism of Nelarabine is the demethoxylation of the purine base by ADA with the formation of 9-β-D-arabinofuranosyl guanine (Ara-G), which suppresses cell proliferation [23].
It is difficult to find a clear correlation between the antimetabolic activity of nucleosides and the modification of the C2 and C6 purine positions or the carbohydrate residue. We decided to synthesize a number of 2-chloro-6-substituted arabinonucleosides using enzymatic synthesis. The next step was to study both their resistance to ADA and their antiproliferative activity, in order to draw conclusions about the relationship between the structure of the nucleoside, its activity against ADA, and its antiproliferative activity.

2. Results and Discussion

2.1. Synthesis of Arabinonucleosides

Chemical modifications of purine are known to be more convenient to carry out in a nucleoside (where the N9 position of purine is protected by a carbohydrate residue) instead of a heterocyclic base [24,25]. The classical method for obtaining arabinonucleosides is the glycosylation of a modified purine base with a protected carbohydrate derivative [26,27,28,29] or complex ribose modifications [30,31]. Enzymatic methods for the synthesis of arabinosides [32,33,34] are very convenient due to the high selectivity and stereospecificity of enzymatic reactions.
We recently published an article describing the synthesis of purine ribosides 1a12a (Scheme 1) modified with chiral amino acid amides at the C6 position of the purine [35]. We showed that derivatives with tyrosine, valine, and serine residues exhibit the properties of A1 adenosine receptor partial agonists. The ribosides 1a12a became the starting compounds for the enzymatic synthesis of arabinonucleosides (Scheme 1).
The synthesis of arabinosides was carried out using enzymatic transglycosylation with recombinant E. coli nucleoside phosphorylases [34,36]. Nucleosides substituted at the C6 position of purine are known to be good substrates for E. coli purine nucleoside phosphorylase (PNP) [37].
The transglycosylation reaction can be carried out in two ways: using synthetic arabinose 1-phosphate (1-P-Ara) (Scheme 1) [34] or 1-b-D-arabinofuranosyluracil (Ara-U) as an arabinose donor (Scheme 1, Table 1). In the first case, PNP is used (Scheme 1, black route). In the second case, PNP and uridine phosphorylase (UP) are used [38] (Scheme 1, azure route). Ara-U can be easily synthesized from uridine via 2,2′-anhydrouridine according to the method of I. Wempen [39].
Figure 3 shows the conversion of riboside 1a to arabinoside 1b. Ara-U turned out to be the best arabinose donor.
According to HPLC data, the concentration of the product in the reaction mixture reached 82%, with a fivefold excess of Ara-U (green trend line). The conversion of riboside to arabinoside was 90% in 5 days. Syntheses of nucleosides 1b12b were carried out using Ara-U at its fivefold molar excess.
As a result of the optimization of enzymatic synthesis, the following conditions for obtaining arabinosides were chosen: the ratio of riboside substrates to Ara-U was 1:5; 0.80 units PNP per 1 mmol of substrate and 0.18 units UP per 1 µmol Ara-U; pH 7.0, 52 °C.
During the experiments, we found that the retention times of ribosides and arabinosides were very close (the HPLC profile of the reaction mixture for the synthesis of 6-N-[L-alanylamido]-2-chloro-9-β-D-arabinofuranosylpurine 2b is shown on Figure 4A, Table 2). In addition to the target arabinoside, the reaction mixture contains up to 8% of the starting riboside and a heterocyclic purine base. This made it difficult to isolate the target product.
To decrease the number of components in the reaction mixture, we decided to employ the arsenolysis of ribosides. It is known that PNP catalyzes the reversible riboside phosphorolysis with the formation of α-D-ribose-1-phosphate.
In the presence of arsenates in the active site of PNP, the intermediate α-D-ribose-1-arsenate is formed from the riboside. Being extremely unstable, it rapidly hydrolyzes to ribose and inorganic arsenate, making the reaction irreversible (Scheme 2) [40,41,42].
According to the HPLC data, the number of components in the reaction mixture decreases from six to four when using arsenolysis (Figure 4B, Table 2), and the isolation of target products is facilitated to a significant degree. We used the usual column chromatography instead of preparative HPLC to isolate the target nucleosides.
The efficiency of arabinoside synthesis does not change depending on whether the arsenate was added at the beginning of the synthesis or upon reaching the equilibrium state. For example, when arsenate was added at the beginning of the synthesis of methionine arabinoside 10b in the first 10 h, the equilibrium shifted towards the formation of arabinoside (Figure 5). However, over time (24 h), both reactions achieved the same equilibrium state. The use of arsenate helps to obtain a higher conversion of the riboside in a shorter period of time.
Previously, we showed that the competitive arsenolysis reaction removes ribose from the reaction sphere [43]. When a catalytic amount of sodium arsenate (up to 0.5 mM) is added, arabinose arsenate is essentially not formed in the reaction mixture. The modified arabinose nucleosides hardly undergo arsenolysis (<0.5% in 24 h at 50 °C). As a result, the formation of arabinosides in the reaction mixture prevails.
The conditions of enzymatic reactions for the production of nucleosides 1b12b are shown in Table 3. The physicochemical characteristics of all the synthesized arabinosides are given in Table 4. NMR spectra and HPLC-profiles are provided in the Supporting Information (Figures S1–S61).
To establish the structure–activity relationship in the biological experiments, we synthesized six nucleosides: 9-β-D-arabinofuranosyl-2-amino-6-(Nα-glycinyl)-purine (13), 9-β-D-arabinofuranosyl-2-amino-6-(Nα-glycinylamido)-purine (14), 2-chloradenosine (15), 2-chloro-arabinoadenosine (16), 2-chloro-6-O-methyl-(9-β-D-arabinofuranosyl)guanine (Cl-Nelarabine, 17), and 2-fluoro-6-O-methyl-(9-β-D-arabinofuranosyl)guanine (F-Nelarabine, 18).
Detailed procedures for the synthesis of 13 and 14 are provided in the Supporting Information. Detailed procedures for the synthesis of 2-aminopurine arabinosides 13 and 14 are provided in the Supporting Information. Nucleosides 15 [25] and 16 [44] (Scheme SI-1 in Supporting Information.), 17, and 18 [45] were synthesized according to previously developed procedures. NMR spectra and HPLC-profiles are provided in the Supporting Information (Figures S62–S74).

2.2. The ADA Substrate Specificity

The resistance of the obtained compounds to the action of E. coli adenosine deaminase was evaluated (Scheme 3). All arabinonucleosides 1–18 were resistant to bacterial ADA.
The inhibition of adenosine deaminase by the obtained compounds was studied. Adenosine was used as a control compound for ADA testing. Then, adenosine was deaminated in the presence of synthesized nucleosides. The synthesized compounds at 0.1 mM concentration had little effect on the enzymatic activity of ADA. These reduced the rate of adenosine deamination by less than 10%. Nucleosides 4b, 15, and 16 reduced the rate of adenosine deamination under the action of ADA by 23%, 63%, and 54%, respectively. The effect of 2-chloradenosine (15), 2-chloro-9-β-D-arabinofuranosyladenine (16), and L-seryl arabinofuranoside 4b on ADA activity is shown in Figure 6.
2-Cloroadenosine appeared to be the best inhibitor among the tested compounds. When ribose was replaced by arabinose (compound 16), Ki increased by 1.7 times (Table 5). The introduction of the serine fragment at the C6 position of purine (compound 4b) increased Ki 9.2 times. The compounds 15, 16, and 4b are competitive inhibitors. Lineweaver–Burk plots for the adenosine deamination at various concentrations of 15, 16, and compound 4b are presented in Figure S75.

2.3. The Influence of the Synthesized Nucleosides on U937 Cell Survival

The activity of the synthesized compounds was tested on the U937 cell line. Atrians (9-β-D-arabinofuranosyl-6-O-methylguanine, Nelarabine, GlaxoSmithKline) was used as a reference drug. Atrians is used to treat recurrent and stable T-cell acute lymphoblastic leukemia.
The activity of the compounds was evaluated using IC50. The IC50 is commonly used to compare the effects of cytotoxic compounds on cell lines. This index determines both the effect on cell death and proliferation during incubation.
9-β-D-Arabinofuranosyl-2-chloro-6-(Nα-L-serinylamido)-purine 4b has shown inhibitory activity comparable to Nelarabine. Dose–response curves for Nelarabine (IC50 = 3 μM) and compound 4b (IC50 = 16 μM) are shown in Figure 7. For both compounds, the effect depended on the concentration in the range of 0.1–50 μM.
Among all the studied compounds, only nucleoside 4b and, to a lesser extent, 2-chloroadenosine 15 exhibited antiproliferative activity. The mechanism of antiproliferative activity is not entirely clear, since compound 4b is not a substrate for bacterial ADA, while compound 15 is a competitive ADA inhibitor.
In addition, the chlorine/fluorine atom at the C2 position of purine prevents nucleosides from binding to the ADA active site. Testing of Cl-Nelarabine and F-Nelarabine showed the absence of both the ADA-substrate and antitumor properties (IC50 > 50 μM).
Both glycine arabinofuranosides of 2-chloroadenine 1b and 2-aminoadenine 14 are non-substrates of bacterial ADA. The introduction of an amino acid amide at the C6 position of purine makes nucleosides resistant to adenosine deaminase.
A number of purine arabinosides containing chiral amino acid amides at the C6 position of the purine 1b12b did not exhibit antiproliferative activity (IC50 over 50 μM) (Table 6). A correlation was found between the lack of ADA substrate properties of the arabinosides synthesized and low anti-tumor activity against human T-lymphoblastic leukemia cells. ADA cannot replace the amino acid residue at position C6 with a hydroxyl to form an analog of the active Ara-G molecule. It can be concluded that the antiproliferative effect of compound 4b does not depend on the action of intracellular ADA. Thus, the mechanism of antiproliferative activity of the serine analog 4b differs from that of Nelarabine.
The antiproliferative activity of the compounds synthesized was tested on other tumor cell lines: LS174T human Caucasian colon adenocarcinoma, SKOV3 ovarian cancer cells, MCF7 breast cancer cells, and A549 non-small-cell lung cancer cells. None of the compounds tested affected cell viability: survival was >80% (within concentrations up to 50.00 μM).

3. Materials and Methods

3.1. General Procedures

All chemicals and solvents were of laboratory grade, were obtained from commercial suppliers, and were used without further purification. The method of producing of recombinant E. coli phosphorylases was described above. The following recombinant E. coli [36] enzymes were used in the present study: UP with specific activity 100 units per mg of protein, 9 mg per mL; PNP 52 units per mg, 15 mg per mL. Ara-U was synthesized as described in [34]. 1-P-Ara was synthesized as described in [46]. Analytical HPLC was performed using the Waters system (Waters 1525, Waters 2489, Breeze 2, (Waters Inc., Milford, MA, USA); (a) Nova Pak C18, 4.6 × 150 mm, eluent A—0.1% TFA/H2O, eluent B—70% acetonitrile in 0.1% TFA/H2O, flow rate 1 mL/min, detection at 280 nm. Gradient 0–100% B, 20 min; (b) Ascentis® Express C18, 2.7 μm 7.5 × 3.0 mm, eluent A—0.1% TFA/H2O, eluent B—70% acetonitrile in 0.1% TFA/H2O, flow rate 0.5 mL/min, detection at 280 nm. Gradient 0–100% B, 20 min.
NMR spectra were recorded using a Bruker Avance II 700 spectrometer (Bruker BioSpin, Rheinstetten, Germany) in DMSO-d6 at 30 °C. Chemical shifts in ppm (δ) were measured relative to the residual solvent signals as internal standards (2.50). Coupling constants (J) were measured in Hz.
Liquid chromatography mass spectrometry was performed using an Agilent 6210 TOF LC/MS system (Agilent Technologies, Santa Clara, CA, USA). UV-Spectra were recorded using a Hitachi U-2900 spectrophotometer (Tokyo, Japan).
The optical rotations were measured for purified samples using the digital polarimeter JASCO model DIP-370 (cylindrical glass cell 3.5 mm D × 50 mm) (Tokyo, Japan).

3.2. Enzymatic Reactions

The reaction conditions of arabinosides synthesis (ratio of reagents, PNP amount, reaction time, and conversion of the base into nucleoside) are provided in Table 1. Riboside (1a12a), KH2PO4, and Ara-U were dissolved in water at 40–50 °C, the pH was adjusted up to 7.0, and Na2HAsO4 was added to the reaction mixture.
The enzymes (PNP, UP) were added and the reaction mixtures were incubated at 50 °C. The reaction progress was monitored by HPLC. When conversion reached the highest value, the reaction was terminated by addition of ethanol (50%, v/v). The reaction mixtures were evaporated up to 5 mL, and the desired products 1b12b were isolated by reversed-phase column chromatography (silica gel C18, Merck), column 100 × 20 mm (the elution conditions are shown in Table 1). The physicochemical and spectral properties of nucleosides are reported in Table 3. The NMR spectra and HPLC of nucleosides are provided in the Supporting Information.

3.3. The ADA Substrate Specificity

The substrate specificity of E. coli adenosine deaminase toward arabinosides 1b12b was determined using a previously published method [47]. According to the HPLC data, the obtained nucleosides do not undergo deamination in vitro.

3.4. Inhibition of E. coli ADA

The reaction mixtures contained 0.04 mM adenosine (Serva), 5 mM potassium phosphate (pH 7.0), and 0.1 mM inhibitor: 6-Amino-2-chloropurine riboside (2-chloroadenosine), 6-Amino-2-chloro-9-beta-D-arabinofuranosyl-9H-purine (2-Cl-AraA), or compound 4b. ADA (0.008 units) was added and the reaction mixtures were incubated at 25 °C for 20 min. The conversion of adenosine into inosine was monitored every 5 min by HPLC.

3.5. Biological Assay

For the experiments, all compounds were dissolved in DMSO and then brought to the desired concentration with an RPMI 1640 nutrient medium. The final concentration of DMSO in the samples did not exceed 0.1% and did not affect cell growth.
The antiproliferative activity of the compounds synthesized was tested on tumor cell lines: U937, LS174T, SKOV3, MCF7, and A549 cancer cells. The cells lines were kindly provided by the Tumor Strain Bank of the N. N. Blokhin Cancer Research Center. The cells were grown in RPMI1640 medium containing 2 mM glutamine and 10% fetal calf (FCS) serum at 37 °C and 5% CO2.
Influence on cell growth was evaluated according to the cell survival rate in the presence of compounds in the studied concentrations. Cells were seeded in 96-well flat-bottom microplates at a seeding density of 8000 cells/well. The test compounds were added to the wells in a volume of 20 µL. The total incubation volume was 200 μL. The incubation time with nucleosides was 48 h (96 h for U937). At the end of the incubation, the MTT reagent was added to the cells and the cells were incubated for 2 h.
The formed formazan crystals were dissolved in 100 µL of DMSO at 37 °C for 20 min. The absorbance of DMSO solutions was measured on an optical counter for multiwell plates at λ = 540 nm.
Cell survival for the corresponding concentration was calculated using the following formula: Ti/C × 100%, where Ti is the OD after incubation with the test agents and C is the OD of parallel control samples (without test compounds). A compound was considered active if the concentration of it that caused growth inhibition by 50% (IC50 calculated from the dose–response curve) was equal to or less than IC50 for Nelarabine. The measurement error did not exceed 5%.
The results were expressed as averages for 4 parallel measurements in 2–3 experiments as cell survival in % (experiment/control) ×100.
Statistical processing of the obtained results was carried out using the Microsoft Excel 2010 program with the method of variational statistics with the determination of the arithmetic mean and the error of the mean (M ± m). The differences between values were considered significant at p < 0.05.

4. Conclusions

A series of purine arabinosides (12 compounds) containing chiral amino acid amides at the C6 position of the purine was synthesized using a transglycosylation reaction. To establish the structure–activity relationship in the biological experiments, six nucleosides were additionally synthesized: 9-β-D-ribofuranosyl-2-amino-6-(Nα-glycinyl)-purine, 9-β-D-ribofuranosyl-2-amino-6-(Nα-glycinylamido)-purine, 2-chloradenosine, 2-chloro-arabinoadenosine, 2-chloro-6-O-methyl-(9-β-D-arabinofuranosyl)guanine (Cl-Nelarabine), and 2-fluoro-6-O-methyl-(9-β-D-arabinofuranosyl)guanine (F-Nelarabine).
Arsenolysis of 2-chloropurine ribosides with chiral amino acid amides at C6 was used in the enzymatic synthesis to simplify the composition of the reaction mixtures. In the presence of arsenates in the active site of PNP, the unstable intermediate α-D-ribose-1-arsenate was formed from the riboside, but arabinose arsenate was essentially not formed in the reaction mixture. The modified arabinose nucleosides hardly undergo arsenolysis, and the formation of arabinosides in the reaction mixture prevails.
The synthesized nucleosides were shown to be resistant to the action of E. coli adenosine deaminase. The antiproliferative activity of synthesized nucleosides was studied on human acute myeloid leukemia cell line U937. Among all compounds, the serine derivative exhibited an activity level (IC50 = 16 μM) close to that of Nelarabine (IC50 = 3 μM) and was evaluated as active.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24076223/s1.

Author Contributions

Conceptualization, B.Z.E., I.V.F. and I.D.K.; methodology, B.Z.E., I.V.F. and O.S.Z.; formal analysis, A.L.K., E.V.D., O.I.L., E.A.Z., A.S.P., O.S.Z. and K.V.A.; validation, I.V.F. and O.S.Z.; investigation, B.Z.E., M.Y.B., I.V.F., O.I.L., E.V.D. and K.V.A.; resources, O.I.L., E.V.D., K.V.A. and R.S.E.; writing—original draft preparation, B.Z.E.; M.Y.B. and I.D.K., writing—review and editing, B.Z.E.; M.Y.B. and I.D.K.; supervision, I.S.M., M.V.K., A.I.M., R.S.E. and I.D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation (Project No. 21-13-00429).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aurelio, L.; Baltos, J.-A.; Ford, L.; Nguyen, A.T.N.; Jörg, M.; Devine, S.M.; Valant, C.; White, P.J.; Christopoulos, A.; May, L.T.; et al. A Structure–Activity Relationship Study of Bitopic N6-Substituted Adenosine Derivatives as Biased Adenosine A1 Receptor Agonists. J. Med. Chem. 2018, 61, 2087–2103. [Google Scholar] [CrossRef]
  2. Jacobson, K.A.; Gao, Z.-G.; Paoletta, S.; Kiselev, E.; Chakraborty, S.; Jayasekara, P.S.; Balasubramanian, R.; Tosh, D.K. John daly lecture: Structure-guided drug design for adenosine and P2Y receptors. Comput. Struct. Biotechnol. J. 2015, 13, 286–298. [Google Scholar] [CrossRef]
  3. 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. 2004, 10, 4472–4479. [Google Scholar] [CrossRef]
  4. Gessi, S.; Merighi, S.; Varani, K.; Cattabriga, E.; Benini, A.; Mirandola, P.; Leung, E.; Mac Lennan, S.; Feo, C.; Baraldi, S.; et al. Adenosine receptors in colon carcinoma tissues and colon tumoral cell lines: Focus on the A3 adenosine subtype. J. Cell. Physiol. 2007, 211, 826–836. [Google Scholar] [CrossRef] [PubMed]
  5. Fleysher, M.H.; Bernacki, R.J.; Bullard, G.A. Some short-chain N6-substituted adenosine analogs with antitumor properties. J. Med. Chem. 1980, 23, 1448–1452. [Google Scholar] [CrossRef] [PubMed]
  6. Krasnov, V.P.; Vigorov, A.Y.; Musiyak, V.V.; Nizova, I.A.; Gruzdev, D.A.; Matveeva, T.V.; Levit, G.L.; Kravchenko, M.A.; Skornyakov, S.N.; Bekker, O.B.; et al. Synthesis and antimycobacterial activity of N -(2-aminopurin-6-yl) and N -(purin-6-yl) amino acids and dipeptides. Bioorg. Med. Chem. Lett. 2016, 26, 2645–2648. [Google Scholar] [CrossRef]
  7. Musiyak, V.V.; Gruzdev, D.A.; Kravchenko, M.A.; Vakhrusheva, D.V.; Levit, G.L.; Krasnov, V.P.; Charushin, V.N. Synthesis and antimycobacterial activity of purine conjugates with (S)-lysine and (S)-ornithine. Mendeleev Commun. 2019, 29, 11–13. [Google Scholar] [CrossRef]
  8. Krasnov, V.P.; Vigorov, A.Y.; Gruzdev, D.A.; Levit, G.L.; Demin, A.M.; Nizova, I.A.; Tumashov, A.A.; Sadretdinova, L.S.; Gorbunov, E.B.; Charushin, V.N. Synthesis of enantiomers of N-(2-aminopurin-6-yl)amino acids. Russ. Chem. Bull. 2015, 64, 2106–2113. [Google Scholar] [CrossRef]
  9. Ward, D.N.; Wade, J.; Walborg, E.F.; Osdene, T.S. The synthesis of N-(6-Purinyl)amino acids. amino acids with a single reactive amino group1a. J. Org. Chem. 1961, 26, 5000–5005. [Google Scholar] [CrossRef]
  10. Letham, D.S.; Young, H. The synthesis and cytokinin activities of N-(purin-6-yl)amino acids. Phytochemistry 1971, 10, 23–28. [Google Scholar] [CrossRef]
  11. Matsubara, S.; Fujii, T.; Nishitani, T. Cytokinin Activities of N-(Purin-6-yl) amino Acids, N-(Purin-6-yl) peptides and Related Compounds (A. NATURAL SCIENCE). Sci. Rep Kyoto Prefect. Univ. Nat. Sci. Living Sci. 1988, 39, 1–6. [Google Scholar]
  12. Iwamura, H.; Yada, M.; Koshimizu, K.; Matsubara, S. Synthesis and comparative cytokinin activities of N-(Purin-6-yl)-d- and -l-amino acid methyl esters. Chem. Biol. Technol. Agric. 1978, 42, 1009–1014. [Google Scholar] [CrossRef]
  13. Ottria, R.; Casati, S.; Manzocchi, A.; Baldoli, E.; Mariotti, M.; Maier, J.A.; Ciuffreda, P. Synthesis and evaluation of in vitro anticancer activity of some novel isopentenyladenosine derivatives. Bioorg. Med. Chem. 2010, 18, 4249–4254. [Google Scholar] [CrossRef] [PubMed]
  14. Lapponi, M.J.; Rivero, C.W.; Zinni, M.A.; Britos, C.N.; Trelles, J.A. New developments in nucleoside analogues biosynthesis: A review. J. Mol. Catal. B Enzym. 2016, 133, 218–233. [Google Scholar] [CrossRef]
  15. Rottenberg, M.E.; Masocha, W.; Ferella, M.; Petitto-Assis, F.; Goto, H.; Kristensson, K.; McCaffrey, R.; Wigzell, H. Treatment of African trypanosomiasis with cordycepin and adenosine deaminase inhibitors in a mouse model. J. Infect. Dis. 2005, 192, 1658–1665. [Google Scholar] [CrossRef] [PubMed]
  16. Cristalli, G.; Vittori, S.; Eleuteri, A.; Grifantini, M.; Volpini, R.; Lupidi, G.; Capolongo, L.; Pesenti, E. Purine and 1-deazapurine ribonucleosides and deoxyribonucleosides: Synthesis and biological activity. J. Med. Chem. 1991, 34, 2226–2230. [Google Scholar] [CrossRef]
  17. Robak, P.; Robak, T. Older and new purine nucleoside analogs for patients with acute leukemias. Cancer Treat. Rev. 2013, 39, 851–861. [Google Scholar] [CrossRef]
  18. Robak, T.; Lech-Maranda, E.; Korycka, A.; Robak, E. Purine nucleoside analogs as immunosuppressive and antineoplastic agents: Mechanism of action and clinical activity. Curr. Med. Chem. 2006, 13, 3165–3189. [Google Scholar] [CrossRef]
  19. Vodnala, S.K.; Lundbäck, T.; Yeheskieli, E.; Sjöberg, B.; Gustavsson, A.L.; Svensson, R.; Olivera, G.C.; Eze, A.A.; de Koning, H.P.; Hammarström, L.G.; et al. Structure-activity relationships of synthetic cordycepin analogues as experimental therapeutics for African trypanosomiasis. J. Med. Chem. 2013, 56, 9861–9873. [Google Scholar] [CrossRef]
  20. Chassy, B.M.; Suhadolnik, R.J. Adenosine Aminohydrolase: Binding and hydrolysis of 2- and 6-substituted purine ribonucleosides and 9-substituted adenine nucleosides. J. Biol. Chem. 1967, 242, 3655–3658. [Google Scholar] [CrossRef]
  21. Robins, M.J.; Basom, G.L. Nucleic acid related compounds. 8. direct conversion of 2′-Deoxyinosine to 6-Chloropurine 2′-Deoxyriboside and selected 6-substituted deoxynucleosides and their evaluation as substrates of adenosine deaminase. Can. J. Chem. 1973, 51, 3161–3169. [Google Scholar] [CrossRef]
  22. Pospísilová, H.; Sebela, M.; Novák, O.; Frébort, I. Hydrolytic cleavage of N6-substituted adenine derivatives by eukaryotic adenine and adenosine deaminases. Biosci. Rep. 2008, 28, 335–347. [Google Scholar] [CrossRef]
  23. Gandhi, V.; Keating, M.J.; Bate, G.; Kirkpatrick, P. Nelarabine. Nat. Rev. Drug Discov. 2006, 5, 17–18. [Google Scholar] [CrossRef]
  24. Berzin, V.B.; Dorofeeva, E.V.; Leonov, V.N.; Miroshnikov, A.I. The preparative method for 2-fluoroadenosine synthesis. Russ. J. Bioorg. Chem. 2009, 35, 210–214. [Google Scholar] [CrossRef] [PubMed]
  25. Berzin, V.B.; Dorofeeva, E.V.; Leonov, V.N.; Lutonina, O.L.; Miroshnikov, A.I. Method of production of 2-chloroadenosine. Patent RU № 23,246,98C1, 20 May 2008. [Google Scholar]
  26. Montgomery, J.; Hewson, K. Nucleosides of 2-Fluoroadenine. J. Med. Chem. 1969, 12, 498–504. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, X.-J.; Li, G.-X.; Qi, X.-X.; Deng, Y.-Q. Stereoselective synthesis of 9-β-d-arabianofuranosyl guanine and 2-amino-9-(β-d-arabianofuranosyl)purine. Bioorg. Med. Chem. Lett. 2005, 15, 683–685. [Google Scholar] [CrossRef]
  28. Glaudemans, C.P.J.; Fletcher, H.G. Syntheses with partially benzylated sugars. III.1 A simple pathway to a “cis- Nucleoside”, 9-β-D-Arabinofuranosyladenine (Spongoadenosine). J. Org. Chem. 1963, 28, 3004–3006. [Google Scholar] [CrossRef]
  29. Tuncbilek, M.; Kucukdumlu, A.; Guven, E.B.; Altiparmak, D.; Cetin-Atalay, R. Synthesis of novel 6-substituted amino-9-(β-d-ribofuranosyl)purine analogs and their bioactivities on human epithelial cancer cells. Bioorg. Med. Chem. Lett. 2018, 28, 235–239. [Google Scholar] [CrossRef] [PubMed]
  30. Utley, L.M.; Maldonado, J.; Awad, A.M. A practical synthesis of xylo- and arabinofuranoside precursors by diastereoselective reduction using Corey-Bakshi-Shibata catalyst. Nucleosides Nucleotides Nucleic Acids 2018, 37, 20–34. [Google Scholar] [CrossRef]
  31. Chattopadhyaya, J.B.; Reese, C.B. A synthesis of purine arabinosides. Nucleic Acids Res. 1978, 5, s67–s72. [Google Scholar] [CrossRef]
  32. Koszalka, G.W.; Averett, D.R.; Fyfe, J.A.; Roberts, G.B.; Spector, T.; Biron, K.; Krenitsky, T.A. 6-N-substituted derivatives of adenine arabinoside as selective inhibitors of varicella-zoster virus. Antimicrob. Agents Chemother. 1991, 35, 1437–1443. [Google Scholar] [CrossRef]
  33. Hanrahan, J.R.; Hutchinson, D.W. The enzymatic synthesis of antiviral agents. J. Biotechnol. 1992, 23, 193–210. [Google Scholar] [CrossRef] [PubMed]
  34. Konstantinova, I.D.; Antonov, K.V.; Fateev, I.V.; Miroshnikov, A.I.; Stepchenko, V.A.; Baranovsky, A.V.; Mikhailopulo, I.A. A Chemo-enzymatic synthesis of β-d-Arabinofuranosyl purine nucleosides. Synthesis 2011, 2011, 1555–1560. [Google Scholar] [CrossRef]
  35. Berzina, M.Y.; Eletskaya, B.Z.; Kayushin, A.L.; Dorofeeva, E.V.; Lutonina, O.I.; Fateev, I.V.; Paramonov, A.S.; Kostromina, M.A.; Zayats, E.A.; Abramchik, Y.A.; et al. Synthesis of 2-chloropurine ribosides with chiral amino acid amides at C6 and their evaluation as A1 adenosine receptor agonists. Bioorg. Chem. 2022, 126, 105878. [Google Scholar] [CrossRef]
  36. Esipov, R.S.; Gurevich, A.I.; Chuvikovsky, D.V.; Chupova, L.A.; Muravyova, T.I.; Miroshnikov, A.I. Overexpression of Escherichia coli genes encoding nucleoside phosphorylases in the pET/Bl21(DE3) system yields active recombinant enzymes. Protein Expr. Purif. 2002, 24, 56–60. [Google Scholar] [CrossRef] [PubMed]
  37. Hassan, A.E.; Abou-Elkhair, R.A.; Riordan, J.M.; Allan, P.W.; Parker, W.B.; Khare, R.; Waud, W.R.; Montgomery, J.A.; Secrist, J.A. Synthesis and evaluation of the substrate activity of C-6 substituted purine ribosides with E. coli purine nucleoside phosphorylase: Palladium mediated cross-coupling of organozinc halides with 6-chloropurine nucleosides. Eur. J. Med. Chem. 2012, 47, 167–174. [Google Scholar] [CrossRef] [PubMed]
  38. Mikhailopulo, I.A.; Miroshnikov, A.I. New trends in nucleoside biotechnology. Acta Nat. 2010, 2, 36–59. [Google Scholar] [CrossRef] [PubMed]
  39. Wempen, I.; Fox, J.J. [11] Synthesis of nucleoside derivatives by conversion from preformed nucleosides. Methods Enzymol. 1967, 12, 76–93. [Google Scholar] [CrossRef]
  40. Schramm, V.L. [13] Enzymatic transition-state analysis and transition-state analogs. Methods Enzym. 1999, 308, 301–355. [Google Scholar] [CrossRef]
  41. Kline, P.C.; Schramm, V.L. Purine nucleoside phosphorylase. Catalytic mechanism and transition-state analysis of the arsenolysis reaction. Biochemistry 1993, 32, 13212–13219. [Google Scholar] [CrossRef]
  42. Schramm, V.L. Enzymatic transition state theory and transition state analogue design. J. Biol. Chem. 2007, 282, 28297–28300. [Google Scholar] [CrossRef]
  43. Konstantinova, I.D.; Fateev, I.V.; Miroshnikov, A.I. The arsenolysis reaction in the biotechnological method of synthesis of modified purine β-D-arabinonucleosides. Russ. J. Bioorg. Chem. 2016, 42, 372–380. [Google Scholar] [CrossRef]
  44. Konstantinova, I.D.; Fateev, I.V.; Miroshnikov, A.I. Method for Production of Purine Nucleosides of β-D-Arabinofuranose Series. Patent RU № 262,402,3C2, 30 June 2017. [Google Scholar]
  45. Fateev, I.V.; Kostromina, M.A.; Abramchik, Y.A.; Eletskaya, B.Z.; Mikheeva, O.O.; Lukoshin, D.D.; Zayats, E.A.; Berzina, M.Y.; Dorofeeva, E.V.; Paramonov, A.S.; et al. Multi-enzymatic cascades in the synthesis of modified nucleosides: Comparison of the thermophilic and mesophilic pathways. Biomolecules 2021, 11, 586. [Google Scholar] [CrossRef] [PubMed]
  46. Fateev, I.V.; Antonov, K.V.; Konstantinova, I.D.; Muravyova, T.I.; Seela, F.; Esipov, R.S.; Miroshnikov, A.I.; Mikhailopulo, I.A. The chemoenzymatic synthesis of clofarabine and related 2’-deoxyfluoroarabinosyl nucleosides: The electronic and stereochemical factors determining substrate recognition by E. coli nucleoside phosphorylases. Beilstein J. Org. Chem. 2014, 10, 1657–1669. [Google Scholar] [CrossRef] [PubMed]
  47. Eletskaya, B.Z.; Gruzdev, D.A.; Krasnov, V.P.; Levit, G.L.; Kostromina, M.A.; Paramonov, A.S.; Kayushin, A.L.; Muzyka, I.S.; Muravyova, T.I.; Esipov, R.S.; et al. Enzymatic synthesis of novel purine nucleosides bearing a chiral benzoxazine fragment. Chem. Biol. Drug Des. 2019, 93, 605–616. [Google Scholar] [CrossRef]
Figure 1. Biologically active purine nucleosides.
Figure 1. Biologically active purine nucleosides.
Ijms 24 06223 g001
Figure 2. ADA substrates (green frame) and non-substrates (red frame). Data from ref. [20,21].
Figure 2. ADA substrates (green frame) and non-substrates (red frame). Data from ref. [20,21].
Ijms 24 06223 g002
Scheme 1. Possible pathways for the synthesis of arabinosides 1b12b. Two approaches to carrying out the transglycosylation reaction: using synthetic arabinose 1-phosphate (1-P-Ara) (black route) [34] or 1-b-D-arabinofuranosyluracil (Ara-U) as an arabinose donor (azure route).
Scheme 1. Possible pathways for the synthesis of arabinosides 1b12b. Two approaches to carrying out the transglycosylation reaction: using synthetic arabinose 1-phosphate (1-P-Ara) (black route) [34] or 1-b-D-arabinofuranosyluracil (Ara-U) as an arabinose donor (azure route).
Ijms 24 06223 sch001
Figure 3. The dependence of the conversion of riboside 1a to arabinoside 1b on the type and amount of arabinose donor. The 1 mL reaction mixtures contained 1 mM modified riboside 1a; 3 or 6 mM 1-P-Ara or 2, 3, or 5 mM Ara-U, 3.9 units PNP, 4.5 units UP in 2 mM potassium-phosphate buffer (pH 7.0).
Figure 3. The dependence of the conversion of riboside 1a to arabinoside 1b on the type and amount of arabinose donor. The 1 mL reaction mixtures contained 1 mM modified riboside 1a; 3 or 6 mM 1-P-Ara or 2, 3, or 5 mM Ara-U, 3.9 units PNP, 4.5 units UP in 2 mM potassium-phosphate buffer (pH 7.0).
Ijms 24 06223 g003
Scheme 2. Arsenolysis in the synthesis of arabinosides.
Scheme 2. Arsenolysis in the synthesis of arabinosides.
Ijms 24 06223 sch002
Figure 4. HPLC-profiles of the reaction mixture for the synthesis of an alanine derivative 2b without arsenate (A) and with sodium arsenate (B).
Figure 4. HPLC-profiles of the reaction mixture for the synthesis of an alanine derivative 2b without arsenate (A) and with sodium arsenate (B).
Ijms 24 06223 g004
Figure 5. The dependence of the arabinoside 10b synthesis rate on the moment of arsenate addition. Red line: arsenate was added to the reaction at the start of synthesis. Violet line: arsenate was added to the reaction 24 h after the start of synthesis.
Figure 5. The dependence of the arabinoside 10b synthesis rate on the moment of arsenate addition. Red line: arsenate was added to the reaction at the start of synthesis. Violet line: arsenate was added to the reaction 24 h after the start of synthesis.
Ijms 24 06223 g005
Scheme 3. Testing the substrate specificity of E. coli ADA towards the synthesized nucleosides.
Scheme 3. Testing the substrate specificity of E. coli ADA towards the synthesized nucleosides.
Ijms 24 06223 sch003
Figure 6. The rate of adenosine deamination in the presence of the nucleosides 15, 16, and 4b (control—adenosine). All reactions were performed in 1 mL 20 mM potassium phosphate (pH 7.0) at 25 °C. The reaction progress was monitored by HPLC. Reaction mixtures contained adenosine (0.26 mg, 0.1 mmol), 15, 16 (0.3 mg, 0.1 mmol), or 4b (0.38 mg, 0.1 mmol) and 0.23 units of recombinant E. coli ADA. The measurement error did not exceed 5%.
Figure 6. The rate of adenosine deamination in the presence of the nucleosides 15, 16, and 4b (control—adenosine). All reactions were performed in 1 mL 20 mM potassium phosphate (pH 7.0) at 25 °C. The reaction progress was monitored by HPLC. Reaction mixtures contained adenosine (0.26 mg, 0.1 mmol), 15, 16 (0.3 mg, 0.1 mmol), or 4b (0.38 mg, 0.1 mmol) and 0.23 units of recombinant E. coli ADA. The measurement error did not exceed 5%.
Ijms 24 06223 g006
Figure 7. Dependence of U937 cell survival on the concentration of Nelarabine and 4b.
Figure 7. Dependence of U937 cell survival on the concentration of Nelarabine and 4b.
Ijms 24 06223 g007
Table 1. Overall conversion and yield.
Table 1. Overall conversion and yield.
CompoundAmino Acid
Amide Residue
RConversion into
Nucleoside, % (HPLC),
Azure Route
Nucleoside Yield (%), Azure Route
1bGly-CH-CONH282.8863
2bL-Ala-CH(CH3)-CONH281.4450
3bL-Val-CH(CH3)2-CONH269.1546
4bL-Ser-CH(CH2OH)-CONH290.6081
5bD-Ser-CH(CH2OH)-CONH278.1077
6bL-Thr-CH(CH(OH)CH3)-CONH280.5760
7bL-Met-CH(CH2-CH2-S-CH3)-CONH291.5188
8bL-S-Me-Cys-CH(CH2-S-CH3)-CONH280.3975
9bL-Tyr-CH(CH2-C6H4-OH)-CONH294.8368
10bL-Lys-CH(CH2-CH2-CH2-CH2-NH2)-CONH296.8592
11bNε-Lys-CH2-CH2-CH2-CH2-CH(NH2)-CONH271.7652
12bβ-Ala-CH2-CH2-CONH270.0263
Table 2. Retention time (RT) and peak area of compounds.
Table 2. Retention time (RT) and peak area of compounds.
Peak NameUraUrdAra-Ubase2a2b
RT (min)2.3123.0623.4546.2857.0147.248
% Area before arsenate adding13.252.8444.723.002.8733.32
% Area after arsenate adding15.09-44.003.57-37.34
Table 3. Experimental data for the enzymatic synthesis of arabinosides 1b12b a.
Table 3. Experimental data for the enzymatic synthesis of arabinosides 1b12b a.
CompoundAcceptor
mol wt.
Donor
mol wt.
SubstratesReaction
Volume, mL
PNP,
Units
(A) b
UP,
Units
(B) b
Volume, ml
(20 mM Na2HAsO4)
Reaction Time,
h
Conversion of Base into Nucleoside (HPLC Data), %Eluent c
%
Acceptor
mg (mmol)
Donor
mg (mmol)
1b1a
359.09
Ara-U
244.2
75
(0.21)
260
(1.06)
104162
(776)
94
(88)
0.20819682.8850
2b2a
373.10
80
(0.21)
260
(1.06)
106165
(770)
95
(89)
0.21219681.4450
3b3a
401.13
80
(0.20)
244
(1.00)
100156
(782)
90
(90)
0.20033669.1570
4b4a
389.10
80
(0.21)
251
(1.03)
103161
(783)
93
(90)
0.20621690.6050
5b5a
389.10
100
(0.26)
313
(1.28)
128200
(778)
115
(90)
0.25650478.1050
6b6a
403.11
100
(0.25)
300
(1.23)
75115
(464)
70
(57)
0.15019680.5770
7b7a
433.11
34
(0.08)
90
(0.37)
20133
(1694)
90
(244)
0.04016891.5170
8b8a
419.09
40
(0.10)
120
(0.49)
4061
(639)
41
(83)
0.08019680.3970
9b9a
465.13
100
(0.21)
260
(1.06)
6090
(419)
54
(51)
0.12019694.8370
10b10a
430.16
15
(0.03)
41
(0.17)
5076
(2179)
45
(268)
0.10016896.8570
11b11a
430.16
100
(0.23)
286
(1.17)
11791
(391)
166
(142)
0.23419671.7670
12b12a
373.10
80
(0.21)
261
(1.07)
107167
(779)
96
(90)
0.21416870.0250
a All enzymatic reactions were performed in 5 mM potassium phosphate buffer (pH 7.0) at 50 °C. Ratio Acceptor:Donor—1:5. The reaction progress was monitored by HPLC. The reactions were stopped when the riboside was transformed into a nucleoside, as in the test reactions. Solutions of recombinant E. coli UP and PNP in 5 mM potassium phosphate buffer (pH 7.0) with activities of 1700 and 1400 units per mL, respectively, were used [36]. Target products were isolated by column chromatography (C18, 20 × 100 mm). b A—ratio of UP (units) per Ara-U (1 mmol). B—ratio of PNP (units) per nucleoside 1b12b (1 mmol). c Nucleosides were eluted from the C18 column with a water–methanol mixture. Eluent—% (vol).
Table 4. Physicochemical properties of arabinonucleosides 1b–12b.
Table 4. Physicochemical properties of arabinonucleosides 1b–12b.
CompoundYield
(%), (mg)
Purity
(%)
RT, min[α]D25UV
λmax, nm (ε)
HRMS: m/z Calcd
[M+H]+
HRMS: m/z Found
[M+H]+
1b63 (43)98.699.50 a16.8
(c 0.25, H2O/DMSO 1:1)
268 (14,900),
212 (18,800)
227.0442 (base)
359.0865
227.0458
359.0894
2b50 (40)99.727.47 a40.4
(c 0.5, H2O)
270 (15,200)
213 (18,600)
241.0599 (base)
373.1021
241.0592
373.1015
3b46 (37)96.639.24 a20.8
(c 0.5, H2O)
270 (16,800)
213 (18,700)
269.0912 (base)
401.1334
269.0889
401.1335
4b81 (58)99.276.15 a43.6
(c 0.5, H2O)
270 (17,900)
212 (23,600)
257.0548 (base)
389.0971
257.0541
389.1007
5b77 (77)95.485.76 a−20.4
(c 0.5, H2O)
268 (16,600)
212 (20,400)
257.0548 (base)
389.0971
257.0541
389.1007
6b60 (60)95.755.68 b76.0
(c 1.0, H2O/DMSO)
269 (21,000)
213 (24,800)
271.0704 (base)
403.1127
271.0695
403.1107
7b88 (30)98.207.35 b54.0
(c 1.0, H2O/DMSO)
268 (19,400)
212 (23,200)
301.0632 (base)
433.1055
301.0636
433.1061
8b75 (21)91.207.01 b80.0
(c 1.0, H2O/DMSO)
270 (16,700)
212 (18,700)
287.0476 (base)
419.0898
287.0457
419.08703
9b68 (68)94.386.91 b46.0
(c 1.0, H2O/DMSO)
271 (19,200)
213 (25,400)
333.0861 (base)
465.1284
333.0851
465.1269
10b92 (10)98.426.63 b-270 (15,300)
213 (17,800)
298.1177 (base)
430.1600
298.1161
430.1581
11b52 (52)95.605.35 b4.8
(c 1.0, H2O/DMSO)
270 (15,400)
213 (17,800)
298.1177 (base)
430.1600
298.1161
430.1581
12b63 (50)99.137.24 b−4.8
(c 0.25, H2O/DMSO 1:1)
270 (16,500)
213 (19,400)
241.0599 base
373.1021
241.0603
373.1043
All nucleosides were obtained as a white amorphous powder. The UV spectra were recorded using a Hitachi U-2900 spectrophotometer. The mass spectra were recorded using an Agilent 6224, ESI-TOF, LC/MS. The optical rotations [α]D25 were measured for purified samples using the digital polarimeter JASCO model DIP-370 (cylindrical glass cell 3.5 mm D × 50 mm). HPLC was performed using the Waters system (Waters, Breeze 2); HPLC methods for each compound are given in the Supporting Information. a Nova Pak C18, 4.6 × 150 mm column, eluent A—0.1% TFA/ H2O, eluent B—70% acetonitrile in 0.1% TFA/H2O, flow rate 1 mL/min, detection at 280 nm. Gradient 0–100% B, 20 min; b Ascentis® Express C18. 2.7 μm 7.5 × 3.0 mm, eluent A—0.1% TFA/H2O, eluent B—70% acetonitrile in 0.1% TFA/H2O, flow rate 0.5 mL/min, detection at 280 nm. Gradient 0–100% B, 20 min.
Table 5. Inhibition of E. coli adenosine deaminase.
Table 5. Inhibition of E. coli adenosine deaminase.
InhibitorKi, mM
2-chloradenosine 150.078 ± 0.012
2-chloro-arabinoadenosine 160.13 ± 0.02
4b1.2 ± 0.2
Table 6. Effect of nucleosides on U937 cell survival.
Table 6. Effect of nucleosides on U937 cell survival.
CompoundU937 Cell Survival, %IC50, µM
1.00 µM5.00 µM25.00 µM50.00 µM
1b79 ± 2.383 ± 2.464 ± 3.260 ± 3.0>50
2b73 ± 3.690 ± 3.671 ± 2.868 ± 2.0>50
3b81 ± 3.285 ± 2.569 ± 2.765 ± 1.9>50
4b78 ± 3.181 ± 4.027 ± 1.19 ± 0.216.0
5b80 ± 4.076 ± 3.858 ± 2.348 ± 1.450
6b98 ± 4.990 ± 2.785 ± 2.579 ± 2.3>50
7b100 ± 4.0105 ± 5.290 ± 3.676 ± 2.2>50
8b85 ± 3.483 ± 2.470 ± 2.156 ± 1.7≈50
9b88 ± 4.482 ± 4.176 ± 2.370 ± 2.8>50
10b74 ± 2.271 ± 2.169 ± 3.461 ± 1.8>50
11b84 ± 4.289 ± 4.484 ± 4.279 ± 2.3>50
12b100 ± 4.082 ± 3.271 ± 2.169 ± 2.0>50
1398 ± 4.087 ± 2.675 ± 3.063 ± 1.8>50
1494 ± 2.877 ± 2.369 ± 3.450 ± 2.050
1583 ± 3.380 ± 2.447 ± 1.927 ± 1.122.0
1792 ± 3.688 ± 4.480 ± 4.073 ± 2.1>50
1898 ± 3.993 ± 3.784 ± 2.576 ± 3.8>50
Nelarabine85 ± 2.513 ± 0.67 ± 0.36 ± 0.13.3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Eletskaya, B.Z.; Berzina, M.Y.; Fateev, I.V.; Kayushin, A.L.; Dorofeeva, E.V.; Lutonina, O.I.; Zorina, E.A.; Antonov, K.V.; Paramonov, A.S.; Muzyka, I.S.; et al. Enzymatic Synthesis of 2-Chloropurine Arabinonucleosides with Chiral Amino Acid Amides at the C6 Position and an Evaluation of Antiproliferative Activity In Vitro. Int. J. Mol. Sci. 2023, 24, 6223. https://doi.org/10.3390/ijms24076223

AMA Style

Eletskaya BZ, Berzina MY, Fateev IV, Kayushin AL, Dorofeeva EV, Lutonina OI, Zorina EA, Antonov KV, Paramonov AS, Muzyka IS, et al. Enzymatic Synthesis of 2-Chloropurine Arabinonucleosides with Chiral Amino Acid Amides at the C6 Position and an Evaluation of Antiproliferative Activity In Vitro. International Journal of Molecular Sciences. 2023; 24(7):6223. https://doi.org/10.3390/ijms24076223

Chicago/Turabian Style

Eletskaya, Barbara Z., Maria Ya. Berzina, Ilya V. Fateev, Alexei L. Kayushin, Elena V. Dorofeeva, Olga I. Lutonina, Ekaterina A. Zorina, Konstantin V. Antonov, Alexander S. Paramonov, Inessa S. Muzyka, and et al. 2023. "Enzymatic Synthesis of 2-Chloropurine Arabinonucleosides with Chiral Amino Acid Amides at the C6 Position and an Evaluation of Antiproliferative Activity In Vitro" International Journal of Molecular Sciences 24, no. 7: 6223. https://doi.org/10.3390/ijms24076223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop