Next Article in Journal
Synthesis of 4-(tert-Butyldimethylsilyl)-4,5,6,7-tetrahydro-[1,2,3]triazolo[1,5-a]pyridine
Previous Article in Journal
4-Amino-3-(1-{[amino(3-methyl-2-oxido-1,2,5-oxadiazol-4-yl)methylene]hydrazinylidene}ethyl)-1,2,5-oxadiazole 2-Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

SNAr Reactions on 2-Amino-4,6-dichloropyrimidine-5-carbaldehyde

by
Jorge Trilleras
1,*,
Alfredo Pérez-Gamboa
1 and
Jairo Quiroga
2
1
Grupo de Investigación en Compuestos Heterocíclicos, Universidad del Atlántico, Puerto Colombia 081007, Colombia
2
Heterocyclic Compounds Research Group, Department of Chemistry, Universidad del Valle, Cali 760032, Colombia
*
Author to whom correspondence should be addressed.
Molbank 2022, 2022(3), M1426; https://doi.org/10.3390/M1426
Submission received: 30 June 2022 / Revised: 5 August 2022 / Accepted: 10 August 2022 / Published: 12 August 2022
(This article belongs to the Section Organic Synthesis)

Abstract

:
We report the experimental results of unexpected aromatic nucleophilic substitution reaction products on 2-amino-4,6-dichloropyrimidine-5-carbaldehyde. The isolated compounds are products of amination, solvolysis, and condensation processes under mild and environmentally friendly conditions, due to the influence of structural factors of the starting pyrimidine and a high concentration of alkoxide ions. This method allows the building of pyrimidine-based compound precursors of N-heterocyclic systems.

Graphical Abstract

1. Introduction

After the pyridine scaffold, the pyrimidine core is regarded as the second heteroaromatic ring present in pharmaceutically active compounds [1,2]. The main justification for this work is researching and exploring synthetic methods to produce this significant class of heterocyclic systems. In particular, functionalized pyrimidines with amino and halogen groups become suitable precursors for a number of structural alterations in the synthesis of pyrimidine-based compounds [3,4,5,6,7]. Halopyrimidines can be made by the Vilsmeier–Haack reaction, which efficiently converts 4,6-dihydroxypyrimidines to 4,6-dichloropyrimidines [8,9]. Another regioselective synthetic strategy is to obtain 2-chloropyrimidines from available 2,4-dichloropyrimidines through regioselective dechlorination [10]. These synthetic protocols are complementary to the conventional synthesis of pyrimidines and analogues through the reaction between β-dicarbonyl compounds (and its synthons) and components with an N–C–N fragment (urea, thiourea, amidine, or guanidine) [11,12]. The halogenated pyrimidines can incorporate nucleophiles regioselectivity via SNAr reaction, unlike other nitrogen heterocycles such as pyridine and imidazole [4]. Substitution reactions that include Grignard agents [13], cross-coupling reactions to build aryl-heteroaryl or heteroaryl-heteroaryl bonds [4,14,15], and alkoxy and amino groups [16,17,18,19]. In this work, we report a sequence of reactions of SNAr, solvolysis, and Claisen–Schmidt condensation on symmetrically substituted 2-amino-4,6-dichloropyrimidine-5-carbaldehyde.

2. Results

Chemistry

Unlike the reactions of SNAr on 2,4-dichloropyrimidine derivatives—in which Lewis acid atoms, functional groups, or catalysts are required to enhance the ratio of isomers and favor substitutions at the C-2 position when weak nucleophilic amines are used—SNAr reactions on symmetrically substituted 4,6-dichloropyrimidine derivatives under conventional conditions and stoichiometric control of reactants work well with different types of nucleophilic amines [19]. However, by increasing the alkalinity on the medium in which the reaction takes place under stoichiometric control of reactants’ reactions, both SNAr amination and solvolysis reactions are observed (Scheme 1).
Similar behavior was observed in derivatization reactions to produce heterocyclic analogs of chalcones by the conventional Claisen–Schmidt condensation method between pyrimidines mono-aminated (prepared following the reported methodology [19]) and acetophenone (Scheme 2).

3. Discussion

Conventional synthetic protocols without requiring regio- and chemoselective control using 2-amino-4,6-dichloropyrimidine-5-carbaldehyde as starting material should generate the expected products. In our study, exploring the conditions for incorporating complexity and molecular diversity on 2-amino-4,6-dichloropyrimidine-5-carbaldehyde, several conditions were tested—including changes in solvents, bases, and heating sources— to find the best reaction conditions for the synthesis of pyrimidine derivatives. We found that using triethylamine (TEA) in refluxing ethanol provides suitable conditions for SNAr amination reactions with different amines (aliphatics, cyclic and alicyclic amines, aromatic, and heteroaromatic and benzylic amines) and to provide the mono-substituted compounds [19]. The increase in alkalinity on the reaction medium influences the isolated products (IIV) (Figure 1), unlike previously reported results [19]. The alcohols used as solvents favor the SNAr reaction and at the same time, with NaOH in the medium, the formation of alkoxide ions.
The 2-amino-4,6-dichloropyrimidine-5-carbaldehyde is a symmetric compound in which all positions, except for the position at C–5, are α and/or γ to a nitrogen inside the ring. Consequently, when undergoing SNAr amination reactions, the second nitrogen atom of the ring decreases electron density and contributes to the stabilization of anionic intermediates, increasing reactivity.
Based on these results, the structural characteristics of the compounds obtained moderate yield and incorporation of alkoxide ions, the formation of compounds (IIV) results from competition between soft (amine) and hard (alkoxide) nucleophiles present. Equimolar addition of the amine is expected to control the mono-substitution on the pyrimidine ring, but the formation of alkoxide ions—which are stronger bases and better nucleophiles—compete towards substitution even if they are not in equal proportions.
Compounds IIV were isolated, purified by recrystallization from ethanol, and their structures were ascertained by IR, MS, and NMR (1H and 13C) analysis. Copies of the spectra are included in supporting information. In the 1H-NMR spectrum of the isolated compounds IIV, the inclusion of the respective alkoxide ions is evidenced. Low field signals, with displacement and multiplicity characteristics for groups –OCH3 and –OCH2CH3, were observed. In compound I, the signal for the methoxy group appear as a singlet at 3.92 ppm; for compound II, the signals at 1.34 and 4.37 ppm associated with protons the ethoxy group.
In the 1H-NMR spectrum for the pyrimidine-based chalcones, the signals for the ethoxy groups appear at 1.13 and 4.37 ppm for compound III, and at 1.38 and 4.39 ppm for compound IV. Additionally, the doublets associated with the α,β-vinyl protons in E configuration of the new C=C bond formed at 7.17 ppm (J = 15 Hz, Hα) and 7.31 ppm (J = 15.5 Hz, Hβ) for compound III; and at 7.15 (J = 15.5 Hz, Hα) and 7.35 (J = 15.5 Hz, Hβ) for compound IV.
Compounds VX were not isolated under these reaction conditions. The excess of NaOH added increases the alkoxide ions and the solvolysis reactions, making it difficult to direct towards the exclusive obtention of the SNAr amination product. No study was conducted with equimolar, lesser, or catalytic quantities of NaOH. Pyrimidine-based chalcones IIIIV were prepared under the conventional method of cross-aldol condensation. The Claisen–Schmidt condensation reaction uses excess solid basic catalyst and polar solvent, generating conditions in which the 2-amino-4,6-dichloropyrimidine-5-carbaldehyde undergoes solvolysis and condensation via the ‘one pot’ process. These reaction conditions can lead to byproduct formation due to Michael addition; although they were not identified, it could be a factor that affects the low yield of the isolated products [20].

4. Materials and Methods

4.1. General Remarks

The 2-amino-4,6-dichloropyrimidine-5-carboxaldehyde was prepared according to Vilsmeier–Haack formylation methodology reported in the literature, starting from the commercially acquired 2-amino-4,6-dihydroxypyrimidine [19]. The reagents and solvents used in the SNAr and Claisen–Schmidt reactions were commercially purchased (Sigma-Aldrich (St. Louis, MO, USA) or Merck (Kenilworth, NJ, USA)), without further purification.
Monitoring of the reactions was carried out with silica gel TLC plates (Merck (Kenilworth, NJ, USA) silica 60 F254). Spots were visualized with UV light at 254 and 365 nm. Melting points (uncorrected) were determined with a Thermo Scientific melting point apparatus, model IA 9100/Capillary. Infrared spectra (FT–IR) were collected at resolution of 2 cm−1 and 16 scans, (transmission mode 4000–500 cm−1) using a Shimadzu FTIR 8400 spectrophotometer (Scientific Instruments Inc., Seattle, WA, USA) and KBr disks. NMR spectra 1H and 13C (DEPT—135) were recorded on a Bruker Advance spectrophotometer operating at 400 and 100 MHz, respectively, using TMS as internal standard (d, 0.0 ppm) and DMSO–d6 as solvents. The NMR signals are reported in ppm and coupling constants (J) are reported in Hertz. Mass spectra were recorded in a Thermo Fisher Scientific (Waltham, MA, USA) GC–MS spectrometer model DSQII (Thermo Fisher Scientific Inc., Waltham, MA, USA) using a direct insertion probe and the electron impact ionization technique (70 eV). HRMS was recorded in an Agilent Technologies QTOF 6520B spectrometer coupled to a HPLC Agilent–1200 equipped with an Agilent Zorbax extend C18 (2.1 × 50 × 1.8 mm) PN 727700-902 column, via electrospray ionization (ESI) and analyzed in a positive mode. HPLC method: 0,4 mL/min, gradient from acetonitrile/water (10%, with 0.1% of formic acid) to acetonitrile (with 0.1% of formic acid). Microanalysis was performed on an Agilent CHNS elemental analyzer (Thermo Fischer Scientific Inc., Madison, WI, USA) and the values are within ±0.4% of the calculated values.

4.2. Synthesis and Characterization

Synthesis of 2-amino-4-(indolin-1-yl)-6-alkoxypyrimidine-5-carbaldehydes (III). To a mixture between 2-amino-4,6-dichloropyrimidin-5-carbaldehyde and indoline—in equimolar amounts (1 mmol)—in ethanol or methanol (5.0 mL), NaOH (0.2 g, 5 mmol) was added. The reaction mixture was stirred at room temperature for 1 h. TLC was used to monitor the progress of the reaction. The solid obtained was isolated by filtration and recrystallized from ethanol. The purity of the product was confirmed by TLC test.
2-amino-4-(indolin-1-yl)-6-methoxypyrimidine-5-carbaldehyde I. White solid. M.p. 173–174 °C, 60% yield. 1H-NMR (400 MHz DMSO–d6) δ (ppm): 3.02 (t, 2H, CH2), 3.92 (m, 5H, NCH2, OCH3), 6.91 (t, 1H, CH, J = 7.4 Hz), 7.07 (t, 1H, CH, J = 7.5 Hz), 7.21 (d, 1H, CH, J = 7.5 Hz), 7.27 (s, 2H, NH2), 7.40 (d, 1H, CH, J = 8.1 Hz), 9.84 (s, 1H, CHO). 13C-NMR δ (ppm): 28.2 (CH2), 52.8 (CH2), 53.8 (CH3), 95.6 (C5), 116.8 (CH), 122.3 (CH), 124.6 (CH), 126.1 (CH), 132.8, 143.8, 160.1 (C6), 162.6 (C2), 172.7 (C4), 181.9 (CHO). IR, KBr (cm−1): 3486–3461 (NH2, st), 1663 (C=O, st). MS (70 eV) m/z (%): 270 (M+, 100), 252 (30), 253 (30), 227 (23), 118 (63), 117 (36), 91 (40), 77 (26), 68 (30), 43 (59). Anal. calcd. for C14H14N4O2 (270.11): C, 62.21; H, 5.22; N, 20.73. Found: C, 62.45; H, 5.12; N, 20.86.
2-amino-4-ethoxy-6-(indolin-1-yl)pyrimidine-5-carbaldehyde II. Yellow solid. M.p. 234–236 °C, 60% yield. 1H-NMR (400 MHz DMSO–d6 RT) δ (ppm): 1.34 (t, 3H, CH3), 3.03 (t, 2H, CH2), 3.91 (t, 2H, CH2), 4.37 (q, 2H OCH2), 6.91 (t, 1H, CH, J = 7.5 Hz), 7.08 (t, 1H, CH, J = 7.5 Hz), 7.22 (s, 3H, CH, NH2), 7.42 (d, 1H, CH, J = 8.06 Hz), 9.86 (s, 1H, CHO). 13C-NMR δ(ppm): 14.4 (CH3), 28.2 (CH2), 52.8 (CH2), 62.2 (OCH2), 95.5 (C5), 116.9 (CH), 122.3 (CH), 124.6 (CH), 126.1 (CH), 132.8, 143.7, 160.1 (C4), 162.7 (C2), 172.4 (C6), 182.1 (CHO). IR, KBr (cm−1): 3429–3380 (NH2, st), 1632 (C=O, st). HR–MS calculated for C15H16N4O2: 284.1273, found: 284.1263. Anal. calcd. for C15H16N4O2 (284.13): C, 63.37; H, 5.67; N, 19.71. Found: C, 63.52; H, 5.52; N, 19.87.
Synthesis of (E)-3-(2-amino-4-ethoxy-6-(amino)pyrimidin-5-yl)-1-phenylprop-2-en-1-one IIIIV. (i) SNAr amination reactions: 2-amino-4,6-dichloropyrimidin-5-carbaldehyde (1 mmol), amine (1 mmol), and triethylamine (1 mmol) in EtOH (5.0 mL) was heated under reflux for 3 h [19]. (ii) Pyrimidine-based chalcones are obtained under classical Claisen–Schmidt conditions. To an equimolar mixture of 2-amino-4-chloro-6-(ethyl(phenyl)aminopyrimidine-5-carbaldehyde (or 2-amino-4-ethoxy-6-((4-methoxy-phenyl)(methyl)amino)pyrimidin-5-carbaldehyde) and acetophenone in ethanol (8,0 mL), NaOH (0.2 g, 5 mmol) was added. Then, the reaction mixture was refluxed for 3 h. TLC was used to monitor the progress of the reaction. The precipitate obtained was isolated by filtration and recrystallized from ethanol to produce a yellow solid. The purity of the product was confirmed by TLC test.
(E)-3-(2-amino-4-ethoxy-6-(ethyl(phenyl)amino)pyrimidin-5-yl)-1-phenylprop-2-en-1-one III. Yellow solid. M.p. 157–159 °C, 20% yield. 1H—NMR (400 MHz DMSO–d6 RT) δ (ppm): 1.13 (t, 3H, CH3), 1.39 (t, 3H, CH3), 3.93 (q, 2H, NCH2), 4.37 (q, 2H, OCH2), 6.91–7.00 (m, 5H, CH), 7.17–7.23 (m, 3H, Ho, Hα J = 15 Hz), 7.31 (d, 1H, Hβ, J = 15.5 Hz), 7.45 (t, 2H, Hm, J = 7.2 Hz), 7.54 (t, 1H, Hp, J = 7.2 Hz), 7.62 (s, 2H, NH2). 13C–NMR δ (ppm): 13.3 (CH3), 14.5 (CH3), 46.8 (NCH2), 62.2 (OCH2), 93.0 (C5), 118.0 (Cα), 123.1 (Co), 123.2 (Cp), 127.6 (Cm), 128.5 (Co), 129.3 (Cm), 131.9 (Cp), 137.7 (Cβ), 138.5 (Ci), 146.9 (Ci), 161.9 (C6), 166.1 (C2), 169.6 (C4), 189.4 (CHO). IR, KBr (cm−1): 3457–3407 (NH2, st), 1641 (C=O, st). MS (70 eV) m/z (%): 388 (M+, 7), 283 (49), 270 (13), 269 (64), 255 (21), 241 (28), 105 (79), 77 (100), 43 (30). Anal. calcd. for C23H24N4O2 (388.19): C, 71.11; H, 6.23; N, 14.42. Found: C, 70.92; H, 6.11; N, 14.57.
(E)-3-(2-amino-4-ethoxy-6-((4-methoxyphenyl)(methyl)amino)pyrimidin-5-yl)-1-phenylprop-2-en-1-one IV. Yellow solid. M.p. 154–156 °C, 20% yield. 1H-NMR (400 MHz DMSO–d6 RT) δ (ppm): 1.38 (t, 3H, CH3), 3.31 (s, 3H, NCH3), 3.64 (s, 3H, OCH3), 4.39 (q, 2H, CH2), 6.77 (d, 2H, Hm, N-aryl, J = 8.9 Hz), 6.96–6.99 (m, 4H, Hp, NH2), 7.15 (d, 1H, Hα, J = 15.5 Hz), 7.35 (d, 1H, Hβ, J = 15.5 Hz), 7.46 (t, 2H, Hm, J = 7.2 Hz), 7.55 (t, 1H, Hp, J = 7.2 Hz), 7.64 (d, 2H, Ho, J = 8.1 Hz). 13C—NMR δ (ppm): 14.5 (CH3), 41.6 (NCH3), 55.2 (OCH3), 62.1 (CH2), 92.0 (C5), 114.6 (Cm, N-aryl), 117.9 (Cα), 124.8 (Co, N-aryl), 127.6 (Co), 128.5 (Cm), 131.8 (Cp), 137.8 (Cβ), 138.6 (Ci), 142.2 (Ci, N-aryl), 155.7 (Cp, N-aryl), 161.6 (C4), 166.7 (C6), 169.5 (C2), 189.5 (CHO). IR, KBr (cm−1): 3432–3322 (NH2, st), 1625 (C=O, st). HR–MS calculated for C23H24N4O3: 404.1848, found: 404.1832. Anal. calcd. for C23H24N4O3 (404.18): C, 68.30; H, 5.98; N, 13.85. Found: C, 68.10; H, 5.81; N, 14.05.

5. Conclusions

The compounds obtained, despite being unexpected, follow the parameters of the SNAr reactions and demonstrate how changes in any of the factors that govern these reactions affect the expected result or product. These mild and environmentally friendly SNAr reaction conditions on 2-amino-4,6-dichloropyrimidine-5-carbaldehyde represent an alternative protocol to functionalize them through SNAr amination with a variety of amines, an important intermediate for the investigation of N-heterocyclic compounds and potential applications.

Supplementary Materials

Spectroscopic data of the isolated compounds IIV are available online.

Author Contributions

Experimental execution and spectroscopic analysis J.T and A.P.-G.; Technical support, data analysis and interpretation, J.Q.; Writing—original draft preparation, J.T.; Writing—review and editing, A.P.-G. and J.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors express their gratitude to the Universidad del Atlántico and Universidad del Valle.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Frank, É.; Szőllősi, G. Nitrogen-Containing Heterocycles as Significant Molecular Scaffolds for Medicinal and Other Applications. Molecules 2021, 26, 4617. [Google Scholar] [CrossRef] [PubMed]
  2. Kerru, N.; Gummidi, L.; Maddila, S.; Kumar Gangu, K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [PubMed]
  3. Madia, V.N.; Nicolai, A.; Messore, A.; De Leo, A.; Ialongo, D.; Tudino, V.; Saccoliti, F.; De Vita, D.; Scipione, L.; Artico, M.; et al. Design, Synthesis and Biological Evaluation of New Pyrimidine Derivatives as Anticancer Agents. Molecules 2021, 26, 771. [Google Scholar] [CrossRef] [PubMed]
  4. Dolšak, A.; Mrgole, K.; Sova, M. Microwave-Assisted Regioselective Suzuki Coupling of 2,4-Dichloropyrimidines with Aryl and Heteroaryl Boronic Acids. Catalysts 2021, 11, 439. [Google Scholar] [CrossRef]
  5. Kharlamova, A.D.; Abel, A.S.; Averin, A.D.; Maloshitskaya, O.A.; Roznyatovskiy, V.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Mono- and Diamination of 4,6-Dichloropyrimidine, 2,6-Dichloropyrazine and 1,3-Dichloroisoquinoline with Adamantane-Containing Amines. Molecules 2021, 26, 1910. [Google Scholar] [CrossRef] [PubMed]
  6. Jadhav, M.; Sankhe, K.; Bhandare, R.R.; Edis, Z.; Bloukh, S.H.; Khan, T.A. Synthetic Strategies of Pyrimidine-Based Scaffolds as Aurora Kinase and Polo-like Kinase Inhibitors. Molecules 2021, 26, 5170. [Google Scholar] [CrossRef]
  7. Drewry, D.H.; Annor-Gyamfi, J.K.; Wells, C.I.; Pickett, J.E.; Dederer, V.; Preuss, F.; Mathea, S.; Axtman, A.D. Identification of pyrimidine-based lead compounds for understudied kinases implicated in driving neurodegeneration. J. Med. Chem. 2022, 65, 1313–1328. [Google Scholar] [CrossRef] [PubMed]
  8. Jansa, P.; Holý, A.; Draĉínnský, M.; Kolman, V.; Janeba, Z.; Kostecká, P.; Kmoníĉková, E.; Zídek, Z. 5-Substituted 2-amino-4,6-dihydroxypyrimidines and 2-amino-4,6-dichloropyrimidines: Synthesis and inhibitory effects on immune-activated nitric oxide production. Med. Chem. Res. 2014, 23, 4482–4490. [Google Scholar] [CrossRef]
  9. Quiroga, J.; Trilleras, J.; Abonía, R.; Insuasty, B.; Nogueras, M.; Cobo, J.; de la Torre, J.M. 4-Aminopyrimidine-5-carbaldehydes as intermediates in a Friedländer type synthesis of 7-arylpyrido[2,3-d]pyrimidines. ARKIVOC Online J. Org. Chem. 2009, 9–27. [Google Scholar] [CrossRef]
  10. Zhang, Y.-M.; Gu, M.; Ma, H.; Tang, J.; Lu, W.; Nan, F.-J. An Efficient Synthesis of 2-Chloropyrimidines via Pd-catalyzed Regioselective Dechlorination of 2,4-Dichloropyrimidines in the Presence of NaHCO3. Chin. J. Chem. 2008, 26, 962–964. [Google Scholar] [CrossRef]
  11. Mittersteiner, M.; Farias, F.F.S.; Bonacorso, H.G.; Martins, M.A.P.; Zanatta, N. Ultrasound-assisted synthesis of pyrimidines and their fused derivatives: A review. Ultrason. Sonochem. 2021, 79, 105683. [Google Scholar] [CrossRef] [PubMed]
  12. Marinescu, M. Biginelli Reaction Mediated Synthesis of Antimicrobial Pyrimidine Derivatives and Their Therapeutic Properties. Molecules 2021, 26, 6022. [Google Scholar] [CrossRef] [PubMed]
  13. Thomas, J. Delia. Grignard Reactions Involving Halogenated Pyrimidines. J. Heterocycl. Chem. 2013, 50, 735–745. [Google Scholar] [CrossRef]
  14. Bhaskaran, S.; Padusha, M.S.A.; Sajith, A.M. Application of Palladium Based Precatalytic Systems in the Suzuki-Miyaura Cross-Coupling Reactions of Chloro- Heterocycles. ChemistrySelect 2020, 5, 9005–9016. [Google Scholar] [CrossRef]
  15. Malik, A.; Rasool, N.; Kanwal, I.; Hashmi, M.A.; Zahoor, A.F.; Ahmad, G.; Altaf, A.A.; Shah, S.A.A.; Sultan, S.; Zakaria, Z.A. Suzuki–miyaura reactions of (4-bromophenyl)-4,6-dichloropyrimidine through commercially available palladium catalyst: Synthesis, optimization and their structural aspects identification through computational studies. Processes 2020, 8, 1341. [Google Scholar] [CrossRef]
  16. Kim, J.; Cho, Y.S.; Min, S.-J. Facile Synthesis of 2-Amino-4-alkoxypyrimidines via Consecutive Nucleophilic Aromatic Substitution (SNAr) Reactions. Bull. Korean Chem. Soc. 2016, 37, 1998–2008. [Google Scholar] [CrossRef]
  17. Kim, J.; Kim, Y.J.; Londhe, A.M.; Pae, A.N.; Choo, H.; Kim, H.J.; Min, S.-J. Synthesis and Biological Evaluation of Disubstituted Pyrimidines as Selective 5-HT2C Agonists. Molecules 2019, 24, 3234. [Google Scholar] [CrossRef]
  18. Bruening, F.; Lovelle, L.E. Highly Regioselective Organocatalytic SNAr Amination of 2,4-Dichloropyrimidine and Related Heteroaryl Chlorides. Eur. J. Org. Chem. 2017, 2017, 3222–3228. [Google Scholar] [CrossRef]
  19. Quiroga, J.; Trilleras, J.; Insuasty, B.; Abonía, R.; Nogueras, M.; Marchal, A.; Cobo, J. Microwave-assisted synthesis of pyrazolo[3,4-d]pyrimidines from 2-amino-4,6-dichloropyrimidine-5-carbaldehyde under solvent-free conditions. Tetrahedron Lett. 2008, 49, 3257–3259. [Google Scholar] [CrossRef]
  20. Yadav, G.D.; Wagh, D.P. Claisen-Schmidt Condensation using Green Catalytic Processes: A Critical Review. ChemistrySelect 2020, 5, 9059–9085. [Google Scholar] [CrossRef]
Scheme 1. Amination and solvolysis reactions on 2-amino-4,6-dichloropyrimidin-5-carbaldehyde.
Scheme 1. Amination and solvolysis reactions on 2-amino-4,6-dichloropyrimidin-5-carbaldehyde.
Molbank 2022 m1426 sch001
Scheme 2. (i) Ethanol, TEA, HNR1R2, reflux for 3 h [19]; (ii) Ethanol, NaOH(s), reflux for 3 h.
Scheme 2. (i) Ethanol, TEA, HNR1R2, reflux for 3 h [19]; (ii) Ethanol, NaOH(s), reflux for 3 h.
Molbank 2022 m1426 sch002
Figure 1. Isolated (IIV) and expected (VX) compounds.
Figure 1. Isolated (IIV) and expected (VX) compounds.
Molbank 2022 m1426 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Trilleras, J.; Pérez-Gamboa, A.; Quiroga, J. SNAr Reactions on 2-Amino-4,6-dichloropyrimidine-5-carbaldehyde. Molbank 2022, 2022, M1426. https://doi.org/10.3390/M1426

AMA Style

Trilleras J, Pérez-Gamboa A, Quiroga J. SNAr Reactions on 2-Amino-4,6-dichloropyrimidine-5-carbaldehyde. Molbank. 2022; 2022(3):M1426. https://doi.org/10.3390/M1426

Chicago/Turabian Style

Trilleras, Jorge, Alfredo Pérez-Gamboa, and Jairo Quiroga. 2022. "SNAr Reactions on 2-Amino-4,6-dichloropyrimidine-5-carbaldehyde" Molbank 2022, no. 3: M1426. https://doi.org/10.3390/M1426

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