Next Article in Journal
High Endogenously Synthesized N-3 Polyunsaturated Fatty Acids in Fat-1 Mice Attenuate High-Fat Diet-Induced Insulin Resistance by Inhibiting NLRP3 Inflammasome Activation via Akt/GSK-3β/TXNIP Pathway
Next Article in Special Issue
“One-Pot” CuCl2-Mediated Condensation/C–S Bond Coupling Reactions to Synthesize Dibenzothiazepines by Bi-Functional-Reagent N, N′-Dimethylethane-1,2-Diamine
Previous Article in Journal
Different Effects of Quercetin Glycosides and Quercetin on Kidney Mitochondrial Function—Uncoupling, Cytochrome C Reducing and Antioxidant Activity
Previous Article in Special Issue
A Study on the Direct Esterification of Monoalkylphosphates and Dialkylphosphates; The Conversion of the Latter Species to Trialkylphosphates by Alkylating Esterification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Green Chemical Approach for Iodination of Pyrimidine Derivatives by Mechanical Grinding under Solvent-Free Conditions

1
Department of Chemistry and RINS, Gyeongsang National University, Jinju 52828, Korea
2
Department of Chemistry Education and RINS, Gyeongsang National University, Jinju 52828, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(19), 6386; https://doi.org/10.3390/molecules27196386
Submission received: 16 August 2022 / Revised: 23 September 2022 / Accepted: 24 September 2022 / Published: 27 September 2022
(This article belongs to the Special Issue Green Approaches in Organic Chemistry)

Abstract

:
The iodination of pyrimidines is usually carried out by using toxic reagents under acidic conditions, such as with sulfuric acid and nitric acid. To avoid toxic reagents, we developed a simple and eco-friendly approach for the iodination of pyrimidine derivatives under solvent-free conditions using solid iodine and AgNO3 as an electrophilic iodinating reagent. The advantages of this method are the relatively short reaction time (20–30 min), simple set-up procedure, high yields (70–98%), and environmentally friendly reaction conditions. Our novel approach for the iodination of pyrimidines, as well as a variety of their derivatives, will contribute to the development of nucleobase-related drug candidates.

1. Introduction

Many studies on modified nucleobases have reported that C5-halogeneted (iodo and bromo) pyrimidine is highly active in medicinal usage, and is widely utilized for therapeutic purposes. In particular, iodinated uridine (C5-iodo-2′-deoxyuridine) is widely used as an antiviral drug [1,2]. Iodinated nucleotides are essential precursors for various functional group transformations [3,4]. Therefore, developing a convenient procedure for the iodination of pyrimidines is an important field in synthetic bioorganic chemistry.
In the production of a wide variety of pharmaceutical and bioactive materials, aromatic iodides, including iodinated pyrimidines, play an essential role as intermediates [5,6]. Electrophilic aromatic substitutions are one of the most widely used synthetic methods to create C-I bonds in aromatic iodides [7,8]. In the case of the electron-deficient arenes and heterocycles, they can be iodinated through the combination of molecular iodine (I2) and electrophilic iodinating reagents, such as silver methylsulfonate (AgOMs), as follows (Scheme 1) [9]:
Ar-H and Ar-I indicate the aromatic molecules and aromatic iodide, respectively. Pyrimidines, which are one type of heterocycles, and their derivatives are also able to be iodinated through similar methods.
The electrophilic iodination of aromatic substrates is quite difficult because iodine has weak reactivity. Thus, this reaction can only occur under harsh reaction conditions such as acidic conditions using nitric acid, acetic acid, or sulfuric acid, or using strong oxidative agents as iodination sources. Environmentally friendly methods for the iodination of pyrimidines have not yet been reported.
We developed solvent-free mechanochemistry methods for the iodination of pyrimidines, namely uracil and cytosine, as well as their derivatives, which have advantages in terms of generating less pollution and lower costs. We developed a procedure for iodination at the C5 position of pyrimidines via a solid-state reaction, performed by grinding all of the reaction mixtures (Figure 1). We performed the reactions using molecular iodine (I2) and various nitrate/nitrite salts (AgNO3, NaNO3, and NaNO2) as follows (Scheme 2):
In this study, direct iodination at the C5 position of the pyrimidines using nitrate salts and iodine was performed under solvent-free conditions at room temperature (Figure 1). The combination of iodine and nitrate salts was found to be the best reagent for the regioselective iodination reactions shown in Figure 1c. The C5-iodo pyrimidine derivatives were confirmed and identified by 1H and 13C NMR and ESI mass spectroscopy. This study suggests an environmentally friendly approach for the iodination of pyrimidine derivatives that will contribute toward the development of nucleobase-related drug candidates.

2. Results and Discussion

2.1. Optimization of Iodination Reactions of Uracil and Cytosine

A previous study reported that AgNO3 was found to give better results than other nitrate salts for the simple aromatic substrates [7]. Here, AgNO3 acted as a Lewis acid and generated nitryl iodide in situ by reaction of silver nitrate with iodine (Scheme 2). Although the grinding reactions were carried out in a solvent-free environment, a few drops of acetonitrile were used for easier grinding. Interestingly, this gentle grinding method was effective for C5 iodination under solid iodine and silver nitrate conditions. Furthermore, the common reaction byproduct was silver iodide (Scheme 2) [7,10].
Our initial optimization of the iodination reaction used uracil (U) as a model system. The optimization was tested with AgNO3 (0.5 Equiv.), molecular iodine (I2) (1.2 Equiv.), and 0.1 mmol uracil, based on a previous study [11], under several different reaction times. During optimization, the relative molar ratio of AgNO3 compared to uracil was changed in the range of 0.5~2.5 (Table 1). At 0.5 Equiv. of AgNO3, the reaction at 25 °C produced 5-iodo-uracil (5I-U) in 38% yield (10 min) and 63% yield (30 min) (Table 1). When the amount of AgNO3 was increased to 2.0 Equiv., 90% of 5I-U was detected after 30 min (Table 1). However, the reaction yield did not increase, even though AgNO3 was increased up to 2.5 Equiv. (Table 1). Thus, the optimal amount of AgNO3 in this reaction is 2.0 Equiv. compared to uracil (Table 1).
Similarly, the optimization of iodination of cytosine was also performed. As shown in Table 2, 2.0 Equiv. of AgNO3 was required for optimal reaction in the iodination of cytosine.

2.2. Iodination Reaction Using AgNO3

2.2.1. Iodination of Uracil and Uridine Derivatives

The iodination of uridine (rU) and 2′-deoxyuridine (dU) was performed with 2.0 Equiv. of AgNO3 for 25 min (optimized conditions). The reaction mixture was filtered, washed with methanol, and purified with silica-gel column chromatography. The 5-iodo-uridine derivative products, 5-iodo-uridine (5I-rU) and 5-iodo-2′-deoxyuridine (5I-dU), were identified by 1H NMR spectra, in which the H5 signals disappeared (Supplementary Figures S2 and S3). The reaction yields for the 5I-rU and 5I-dU were 83% and 86%, respectively (Table 3).
We also performed the iodination reaction on the sugar-modified 2′-OMe-uridine (mU). Interestingly, the reaction to 5-iodo-2′-OMe-uridine (5I-mU) was completed within 15 min with a 98% yield (Table 3).

2.2.2. Iodination of Cytosine and Cytidine Derivatives

Next, the iodination of cytidine (rC), 2′-deoxycytidine (dC), and 2′-OMe-cytidine (mC) was also performed with 2.0 Equiv. of AgNO3 for 30 min (optimized conditions) to produce 5-iodo-cytidine (5I-rC), 5-iodo-2′-deoxycytidine (5I-dC), and 5-iodo-2′-OMe-cytidine (5I-mC), respectively. The iodination products, 5I-rC, 5I-dC, and 5I-mC, were identified by the 1H and 13C -NMR spectra (Supplementary Figures S6–S8). Interestingly, the 5I-rC was produced with a yield of only 59%, which was significantly lower than the other cytidine derivatives (Table 3).

2.3. Iodination Reaction Using Ag2SO4, NaNO3, and NaNO2

In order to clarify the salt effect on the solvent-free iodination of pyrimidine derivatives, we also performed the same reaction with Ag2SO4, NaNO3, and NaNO2 as the Lewis acid to generate nitryl iodide.
In the case of uracil, the reaction yield to 5-iodo-uracil was 73% when Ag2SO4 was used (Table 3). However, the reaction with NaNO3 and NaNO2 led to reaction yields of only 33% and 38%, respectively (Table 3). A similar result was observed for the iodination of uridine (Table 3). Surprisingly, when Ag2SO4, NaNO3, and NaNO2 were used, only <50% of dU and mU could be converted to 5I-dU and 5I-mU, respectively (Table 3).
In the case of cytidine derivatives, all iodination reactions with Ag2SO4, NaNO3, and NaNO2 showed reaction yields lower than 35%, except the iodination of mU with Ag2SO4 (Table 3). The iodination of mU with Ag2SO4 had 80% yield (Table 3).
The synthetic strategy for the electrophilic iodination of pyrimidine was an electrophilic reaction of the reactive iodine species produced by nitrate salt. Table 3 summarizes the direct iodination and the oxidative iodination required for I+ species, which were generated by an electrophilic iodinating reagent (Ag2SO4), in comparison to the other metal catalysts (NaNO3 and NaNO2) [12,13,14]. The reactivity was different depending on each pyrimidine substrate. Based on the obtained results, we can conclude that the order of the reactivity of the nitrate salts is AgNO3 >> Ag2SO4 >> NaNO3 >> NaNO2.
It has been reported that electrophilic reactions can be catalyzed by metal-based salts [15,16]. Silver-containing salts can more efficiently catalyze this reaction [17]. Our results revealed that AgNO3 is a better catalyst for the direct iodination reaction compared to Ag2SO4, even though both contain the Ag+ ion. In the presence of oxygen, NaNO3 and NaNO2 can also promote this reaction, although their efficiencies were slightly lower than the Ag+-based catalysts. In addition, NaNO2 showed higher efficiency than NaNO3, because it can be converted to nitrogen oxide [12,18].

3. Materials and Methods

3.1. General Information for Experiments

All the reagents and solvents were purchased from local commercial suppliers. The elemental iodine and nitrate salts were purchased from Dungsun and used without further purification. The 1H NMR spectra were recorded on a Bruker 300-MHz instrument with DMSO as the solvent and TMS as the internal standard (TMS, δ = 0.00 ppm). The 13C NMR spectra were recorded on a Bruker 500 MHz instrument with DMSO as the solvent. All the products were identified by comparison of their spectral and physical data with those of the known sample. The progress of the reaction and the purity of the products were checked by thin-layer chromatography (TLC) on silica gel 60 F254-coated TLC plates (Merck KGaA, Darmstadt, Germany) and visualized by short-UV light at 254 nm. The coupling constants (J) were reported in Hz. Peak multiplicity was indicated as follows: s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet; and dd, doublet of the doublet. Mass spectra were obtained by an LCQ Fleet ion trap mass spectrometer using positive-ion (ESI+) and negative-ion (ESI–) mode electrospray ionizations (ThermoScientific; Qual Browser software, version 2.0.7, Thermo Fischer Scientific, Waltham, MA, USA). An amount of 20 μL of the sample was diluted with HPLC grade methanol solution for analysis. HPLC separation with UV was performed on a 1290 infinity UHPLC system (Agilent technology, Waldbroen, Germany) with a 1μL detection cell and a 20 μL sample loop. The chromatographic column was carried out using a Waters µBondapak C18 Column, 10 µm, 7.8 mm × 300 mm.

3.2. Synthesis of 5-Iodo Pyrimidine Derivatives

Uracil and cytosine (4.4 mmol), iodine in the solid state (1.12 mmol), nitrate salts (9 mmol), and 2–4 drops of acetonitrile were mixed in a mortar. The reaction mixture was ground together in a mortar using a pestle to generate a violet-colored tacky solid within 20–30 min. The reaction proceeded exothermically (indicated by an increase in temperature of 20–35 °C). After the reaction was completed (as confirmed by TLC), a saturated solution of sodium thiosulfate (5 mL) was added to remove the unreacted iodine, and the remaining solid was separated out. The resulting solid was washed with a saturated solution of sodium thiosulfate (5 mL), filtered, and then finally washed with cold water. The crude products were purified by silica gel column chromatography (solvent: CH2Cl2: MeOH).

3.3. Spectroscopic Details of C5-Iodo Pyrimidine Derivatives

3.3.1. 5-Iodo-Uracil (5I-U)

5I-U was obtained according to the general procedure as an off-white solid (0.96 g, 90% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.54 (s, 1H), 7.97 (s, 4H). 13C NMR (500 MHz, DMSO-d6) δ 161.96, 151.69, 147.42, 68.05. MP (259 °C) [15]. HRMS (ESI) calcd. for C4H3IN2O2 [M + H]+ 237.9845, found at 237.9167.

3.3.2. 5-Iodo-Uridine (5I-rU)

5I-rU was obtained according to the general procedure as an off-white solid (0.63 g, 83% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.68 (s, 1H), 8.48 (s, 1H), 5.72 (d, J = 4.6 Hz, 1H), 4.01 (dt, J = 16.7, 4.8 Hz, 2H), 3.86 (dt, J = 4.8, 2.8 Hz, 1H), 3.68 (dd, J = 12.1, 2.9 Hz, 1H), 3.57 (dd, J = 12.0, 2.7 Hz, 1H), 3.17 (s, 0H). 2.5.7 (s, 0H). 13C NMR (500 MHz, DMSO-d6) δ 160.96, 150.84, 145.61, 88.76, 85.19, 74.40, 69.85, 69.76, 60.66, 49.08. MP (206–208 °C) [19]. HRMS (ESI) calcd. for C9H11IN2O6 [M + H]+ 370.0995, found at 369.1667.

3.3.3. 5-Iodo-2′-Deoxyuridine (5I-dU)

5I-dU was obtained according to the general procedure as an off-white solid (0.67 g, 86% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.66 (s, 1H), 8.40 (s, 1H), 6.10 (t, J = 6.5 Hz, 1H), 4.29–4.19 (m, 1H), 3.79 (q, J = 3.2 Hz, 1H), 3.65–3.51 (m, 2H), 2.20–2.04 (m, 2H). 13C NMR (500 MHz, DMSO-d6) δ 160.97, 150.58, 145.53, 88.00, 85.13, 70.48, 69.73, 61.30. MP (180–183 °C) [19]. HRMS (ESI) calcd. for C9H11IN2O5 [M + H]+ 354.1005, found at 353.9707.

3.3.4. 5-Iodo-2′-OMe-Uridine (5I-mU)

5I-mU was obtained according to the general procedure as an off-white solid (0.73 g, 98% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.69 (s, 1H), 8.53 (s, 1H), 5.78 (d, J = 3.9 Hz, 1H), 4.11 (t, J = 5.2 Hz, 1H), 3.85 (dd, J = 5.6, 2.7 Hz, 1H), 3.79 (t, J = 4.5 Hz, 1H), 3.68 (d, J = 2.7 Hz, 1H), 3.57 (dd, J = 12.1, 2.5 Hz, 1H), 3.38 (s, 4H). 13C NMR (500 MHz, DMSO-d6) δ 160.97, 150.60, 145.33, 87.02, 85.27, 83.51, 69.78, 68.28, 60.14, 58.05. MP (149–151 °C) [19]. HRMS (ESI) calcd. for C10H13IN2O6 [M + H]+ 384.1265, found at 383.9813.

3.3.5. 5-Iodo-Cytosine (5I-C)

5I-C was obtained according to the general procedure as an off-white solid (0.95 g, 90% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.55 (s, 4H), 7.97 (s, 5H), 5.87 (d, J = 7.2 Hz, 1H). 13C NMR (500 MHz, DMSO) δ 164.66, 154.87, 150.94, 145.94, 93.47, 56.13. MP (149–151 °C) [19]. HRMS (ESI) calcd. for C4H4 IN3O [M + H]+ 237.0005, found at 236.9394.

3.3.6. 5-Iodo-Cytidine (5I-rC)

5I-rC was obtained according to the general procedure as an off-white solid (0.45 g, 59% yield). 1H NMR (500 MHz, DMSO-d6) δ 9.53 (s, 1H), 8.53–8.05 (m, 2H), 6.10 (d, J = 7.9 Hz, 1H), 5.72 (d, J = 3.6 Hz, 1H), 4.06 (t, J = 4.2 Hz, 1H), 3.98 (s, 0H), 3.96–3.89 (m, 1H), 3.72 (dd, J = 12.3, 2.9 Hz, 1H), 3.60 (dd, J = 12.3, 3.0 Hz, 1H). 13C NMR (500 MHz, DMSO-d6) δ 159.52, 147.86, 144.95, 94.27, 89.98, 85.20, 74.59, 69.32, 60.43. HRMS (ESI) calcd. for C9H12IN3O5 [M + H]+ 369.1155, found at 369.1667. MP (225 °C) [19].

3.3.7. 5-Iodo-2′-Deoxycytidine (5I-dC)

5I-dC was obtained according to the general procedure as an off-white solid (0.61 g, 79% yield). 1H NMR (300 MHz, DMSO-d6) δ 9.42 (s, 30H), 8.23 (s, 2H), 4.22 (s, 1H), 3.86 (s, 1H), 2.14 (d, J = 6.4 Hz, 1H). 13C NMR (500 MHz, DMSO-d6) δ 159.54, 147.70, 144.96, 94.22, 88.50, 86.30, 70.22, 61.26. MP (175–177 °C) [19]. HRMS (ESI) calcd. for C9H12IN3O4 [M + H]+ 353.1165, found at 352.9867.

3.3.8. 5-Iodo 2′-OMe-Cytidine (5I-mC)

5I-mC was obtained according to the general procedure as an off-white solid (0.71 g, 95% yield). 1H NMR (300 MHz, DMSO-d6) δ 11.48 (s, 1H), 8.54 (s, 1H), 5.79 (dd, J = 7.2, 3.0 Hz, 1H), 4.07 (dd, J = 6.8, 4.9 Hz, 1H), 3.57 (dd, J = 12.1, 2.3 Hz, 1H). 13C NMR (500 MHz, DMSO-d6) δ 167.57, 159.99, 150.48, 148.16, 88.23, 83.41, 69.03, 67.38, 61.70, 55.63, 49.08. HRMS (ESI) calcd. for C10H14IN3O5 [M + H]+ 383.1425, found at 383.0833.

4. Conclusions

In conclusion, this study identified a convenient and highly effective method for the solid-state iodination of C5-pyrimidines using iodine (the grinding method). The advantages of this method are the relatively short reaction time (20–30 min), simple set-up procedure, high yields (70–98%), and environmentally friendly reaction conditions that do not require the use of any additional reagents or solvents. Detailed information on the solid-state iodination reaction at the C5 position in various pyrimidine derivatives was also reported. We believe that this method will be helpful in future nucleobase-related drug design and related research.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196386/s1, Figures S1–S8: 1H, 13C, and ESI-Ms spectra of the 5-iodo pyrimidine derivatives and Figure S9: TLC Rf values of the 5-iodo pyrimidine derivatives.

Author Contributions

Conceptualization, T.B., B.-S.K., S.K.K. and J.-H.L.; methodology, T.B. and J.-H.L.; software, T.B.; validation, T.B., B.-S.K. and J.-H.L.; formal analysis, T.B. and B.P.; investigation, T.B. and B.P.; resources, T.B. and H.-S.J.; data curation, T.B. and H.-S.J.; writing—original draft preparation, T.B. and J.-H.L.; writing—review and editing, T.B., B.-S.K., S.K.K. and J.-H.L.; visualization, T.B.; supervision, S.K.K. and J.-H.L.; project administration, J.-H.L.; funding acquisition, B.-S.K., S.K.K. and J.-H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (2020R1A2C1006909, 2020R1C1C1013785, 2021R1A2C1012140, and 2022R1A4A1021817) and the Samsung Science and Technology Foundation (SSTF-BA1701-10).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of all compounds are available from the corresponding authors.

References

  1. Prusoff, W.H.; Chen, M.S.; Fischer, P.H.; Lin, T.S.; Shiau, G.T.; Schinazi, R.F.; Walker, J. Antiviral iodinated pyrimidine deoxyribonucleosides: 5-iodo-2′-deoxyuridine; 5′-iodo-2′-deoxycytidine; 5-iodo-5′-amino-2′,5′-didoxyuridine. Pharmacol. Ther. 1979, 7, 1–34. [Google Scholar] [CrossRef]
  2. Markham, A.F.; Newton, C.R.; Porter, R.A.; Sim, I.S. Synthesis of some 5′-amino-2′,5′-dideoxy-5-iodouridine derivatives and their antiviral properties against herpes simplex virus. Antivir. Res. 1982, 2, 319–330. [Google Scholar] [CrossRef]
  3. Liang, Y.; Gloudeman, J.; Wnuk, S.F. Palladium-Catalyzed Direct Arylation of 5-Halouracils and 5-Halouracil Nucleosides with Arenes and Heteroarenes Promoted by TBAF. J. Org. Chem. 2014, 79, 4094–4103. [Google Scholar] [CrossRef] [PubMed]
  4. Kodape, M.M.; Aswar, A.S.; Gawhale, N.D.; Humne, V.T.; Mir, B.A. Facile aromatization of 3,4-dihydropyrimidin-2(1H)-ones (DHPMs) by iodine. Chin. Chem. Lett. 2012, 23, 1339–1342. [Google Scholar] [CrossRef]
  5. Sharma, V.; Srivastava, P.; Bhardwaj, S.K.; Agarwal, D. Iodination of industrially important aromatic compounds using N-iodosuccinimide by grinding method. Green Process. Synth. 2018, 7, 477–486. [Google Scholar] [CrossRef]
  6. Wu, Y.; Xu, S.; Wang, H.; Shao, D.; Qi, Q.; Lu, Y.; Ma, L.; Zhou, J.; Hu, W.; Gao, W.; et al. Directing Group Enables Electrochemical Selectively Meta-Bromination of Pyridines under Mild Conditions. J. Org. Chem. 2021, 86, 16144–16150. [Google Scholar] [CrossRef] [PubMed]
  7. Sy, W.-W.; Lodge, B.A. Iodination of alkylbenzenes with iodine and silver nitrite. Tetrahedron Lett. 1989, 30, 3769–3772. [Google Scholar] [CrossRef]
  8. Zeng, Y.; Zhang, L.; Zhao, Y.; Ni, C.; Zhao, J.; Hu, J. Silver-mediated trifluoromethylation-iodination of arynes. J. Am. Chem. Soc. 2013, 135, 2955–2958. [Google Scholar] [CrossRef] [PubMed]
  9. Tanwar, L.; Börgel, J.; Lehmann, J.; Ritter, T. Selective C–H Iodination of (Hetero)arenes. Org. Lett. 2021, 23, 5024–5027. [Google Scholar] [CrossRef] [PubMed]
  10. Patil, K.; Rao, C.; Lacksonen, J.; Dryden, C. The silver nitrate-iodine reaction: Iodine nitrate as the reaction intermediate. J. Inorg. Nucl. Chem. 1967, 29, 407–412. [Google Scholar] [CrossRef]
  11. Yusubov, M.S.; Tveryakova, E.N.; Krasnokutskaya, E.A.; Perederyna, I.A.; Zhdankin, V.V. Solvent-Free Iodination of Arenes using Iodine–Silver Nitrate Combination. Synth. Commun. 2007, 37, 1259–1265. [Google Scholar] [CrossRef]
  12. Iskra, J.; Stavber, S.; Zupan, M. Aerobic oxidative iodination of organic molecules activated by sodium nitrite. Tetrahedron Lett. 2008, 49, 893–895. [Google Scholar] [CrossRef]
  13. Aso, M.; Kaneko, T.; Nakamura, M.; Koga, N.; Suemune, H. A simple and efficient method for synthesis of 5-substituted 2′-deoxyuridine nucleosides using metal-halogen exchange reaction of 5-iodo-2′-deoxyuridine sodium salt. Chem. Commun. 2003, 1094–1095. [Google Scholar] [CrossRef] [PubMed]
  14. Kumar, R.; Wiebe, L.I.; Knaus, E.E. A mild and efficient methodology for the synthesis of 5-halogeno uracil nucleosides that occurs via a 5-halogeno-6-azido-5,6-dihydro intermediate. Can. J. Chem. 1994, 72, 2005–2010. [Google Scholar] [CrossRef]
  15. Sheppard, T.D. Metal-catalysed halogen exchange reactions of aryl halides. Org. Biomol. Chem. 2009, 7, 1043–1052. [Google Scholar] [CrossRef]
  16. Klapars, A.; Antilla, J.C.; Huang, X.; Buchwald, S.L. A General and Efficient Copper Catalyst for the Amidation of Aryl Halides and the N-Arylation of Nitrogen Heterocycles. J. Am. Chem. Soc. 2001, 123, 7727–7729. [Google Scholar] [CrossRef] [PubMed]
  17. Racys, D.T.; Sharif, S.A.I.; Pimlott, S.L.; Sutherland, A. Silver(I)-Catalyzed Iodination of Arenes: Tuning the Lewis Acidity of N-Iodosuccinimide Activation. J. Org. Chem. 2016, 81, 772–780. [Google Scholar] [CrossRef] [PubMed]
  18. Ajvazi, N.; Stavber, S. Electrophilic Iodination of Organic Compounds Using Elemental Iodine or Iodides: Recent Advances 2008–2021: Part I. Compounds 2022, 2, 3–24. [Google Scholar] [CrossRef]
  19. Paolini, L.; Petricci, E.; Corelli, F.; Botta, M. Microwave-Assisted C-5 Iodination of Substituted Pyrimidinones and Pyrimidine Nucleosides. Synthesis 2003, 2003, 1039–1042. [Google Scholar] [CrossRef]
Scheme 1. Chemical reaction scheme for the synthesis of aromatic iodides.
Scheme 1. Chemical reaction scheme for the synthesis of aromatic iodides.
Molecules 27 06386 sch001
Scheme 2. Chemical reaction scheme for the synthesis of 5-iodo-pyrimidine derivatives.
Scheme 2. Chemical reaction scheme for the synthesis of 5-iodo-pyrimidine derivatives.
Molecules 27 06386 sch002
Figure 1. Schematic representations of the mechanical grinding reaction. (a) Grinding with a mortar and pestle (“grindstone” activation); (b) high-speed vibration milling in a mixer mill (activation of the compound); (c) C5 iodination of pyrimidine derivatives under solvent-free conditions.
Figure 1. Schematic representations of the mechanical grinding reaction. (a) Grinding with a mortar and pestle (“grindstone” activation); (b) high-speed vibration milling in a mixer mill (activation of the compound); (c) C5 iodination of pyrimidine derivatives under solvent-free conditions.
Molecules 27 06386 g001
Table 1. Optimization of the iodination of uracil at room temperature 1.
Table 1. Optimization of the iodination of uracil at room temperature 1.
Molar Ratio of AgNO3 2 Yield (%) 3
10 min20 min30 min
0.5385063
1.0386368
1.5506477
2.0658690 4
2.5688390
1 The relative molar ratio of I2 compared to the substrate was 1.2. 2 The relative molar ratio of AgNO3 compared to the substrate. 3 Reaction yield was determined from the weight of the final product. 4 Optimized condition.
Table 2. Optimization of the iodination of cytosine at room temperature 1.
Table 2. Optimization of the iodination of cytosine at room temperature 1.
Molar Ratio of AgNO3 2 Yield (%) 3
10 min20 min30 min
0.5305560
1.0406065
1.5556370
2.0738090 4
2.5708190
1 The relative molar ratio of I2 compared to the substrate was 1.2. 2 The relative molar ratio of AgNO3 compared to the substrate. 3 Reaction yield was determined from the weight of the final product. 4 Optimized condition.
Table 3. Summary of the iodination at the C5 position of pyrimidine derivatives with metal catalysts at room temperature 1.
Table 3. Summary of the iodination at the C5 position of pyrimidine derivatives with metal catalysts at room temperature 1.
SubstrateProductStructure of ProductMolar Ratio 2Yield (%) 3
AgNO3Ag2SO4NaNO3NaNO2
Uracil
(U)
5-iodo-uracil
(5I-U)
Molecules 27 06386 i0012.090733338
Uridine
(rU)
5-iodo-uridine
(5I-rU)
Molecules 27 06386 i0022.083705034
Deoxyuridine
(dU)
5-iodo-2′-deoxyuridine
(5I-dU)
Molecules 27 06386 i0032.08691814
2′-OMe-uridine
(mU)
5-iodo-2′-OMe-uridine
(5I-mU)
Molecules 27 06386 i0042.09836648
Cytosine
(C)
5-iodo-cytosine
(5I-C)
Molecules 27 06386 i0052.090282918
Cytidine
(rC)
5-iodo-cytidine
(5I-rC)
Molecules 27 06386 i0062.0595419
Deoxycytidine
(dC)
5-iodo-2′-deoxycytidine
(5I-dC)
Molecules 27 06386 i0072.07910762
2′-OMe-cytidine5-iodo-2′-OMe-cytidine
(5I-mC)
Molecules 27 06386 i0082.095803210
1 The relative molar ratio of I2 compared to the substrate was 1.2. 2 The relative molar ratio of the salt compared to the substrate. 3 Reaction yield was determined from the weight of the final product.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Balasubramaniyam, T.; Kim, B.-S.; Pallavi, B.; Jin, H.-S.; Kim, S.K.; Lee, J.-H. A Green Chemical Approach for Iodination of Pyrimidine Derivatives by Mechanical Grinding under Solvent-Free Conditions. Molecules 2022, 27, 6386. https://doi.org/10.3390/molecules27196386

AMA Style

Balasubramaniyam T, Kim B-S, Pallavi B, Jin H-S, Kim SK, Lee J-H. A Green Chemical Approach for Iodination of Pyrimidine Derivatives by Mechanical Grinding under Solvent-Free Conditions. Molecules. 2022; 27(19):6386. https://doi.org/10.3390/molecules27196386

Chicago/Turabian Style

Balasubramaniyam, Thananjeyan, Byeong-Seon Kim, Badvel Pallavi, Ho-Seong Jin, Sung Kuk Kim, and Joon-Hwa Lee. 2022. "A Green Chemical Approach for Iodination of Pyrimidine Derivatives by Mechanical Grinding under Solvent-Free Conditions" Molecules 27, no. 19: 6386. https://doi.org/10.3390/molecules27196386

Article Metrics

Back to TopTop