Next Article in Journal
Interactions between Isoniazid and α-Hydroxycarboxylic Acids
Previous Article in Journal
Direct Arylation-Based Synthesis of Carbazoles Using an Efficient Palladium Nanocatalyst under Microwave Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis of 1,2,3-Triazoles from Alkyne-Azide Cycloaddition Catalyzed by a Bio-Reduced Alkynylcopper (I) Complex †

1
Centro Conjunto de Investigación en Química Sustentable UAEM-UNAM, Carretera Toluca-Atlacomulco Km. 14.5, Toluca 50200, Estado de México, Mexico
2
Facultad de Química, Universidad Autónoma del Estado de México. Paseo Colón esq. Paseo Tollocan, Toluca 50120, Mexico
*
Author to whom correspondence should be addressed.
Presented at the 24th International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2020; Available online: https://ecsoc-24.sciforum.net/.
Published: 14 November 2020

Abstract

:
A small library of 1,2,3-triazoles was synthesized from diverse alkynes and azides using catalytic amounts of an alkynylcopper (I) complex, which in turn was prepared from direct treatment of phenylacetylene with a Fehling reagent in the presence of glucose as a reducing agent. The results suggest that copper-catalyzed azide alkyne cycloaddition (CuAAC) reactions require only 0.5 mg/mmol copper (I) phenylacetylide without any further additives.

1. Introduction

For more than 15 years, copper-catalyzed azide alkyne cycloaddition (CuAAC) has been the most extended “click” reaction providing a successful method for molecular assembly as well as a growing approach for drug design and materials discovery [1,2,3,4]. In this reaction, the copper catalyst plays an essential role in driving the formation of 1,2,3-triazole, recognized as the main CuAAC product serving as a molecular scaffold in these processes [5,6].
In this regard, diverse copper catalysts have been designed, prepared, and used in CuAAC reactions for several reaction conditions; among these catalysts, alkynylcopper (I) complexes, namely copper acetylides, have been identified as efficient catalysts in these reactions [7]. Although first reports described that metal acetylides are explosives [8], recent reports show some stable copper acetylides with excellent catalytic properties in CuAAC reactions [9,10,11,12].
In conjunction with other research, we observed the in situ formation of 1,2,3-triazole copper complexes by a straightforward mixing of alkyne, azide, and a copper (I) salt [13]. From these studies, we detected the formation of a copper acetylide and we decided to investigate in detail this process, aiming to prepare efficient catalysts for the synthesis of 1,2,3-triazoles through CuAAC reactions. Herein is disclosed a summary of our recent findings about this challenge.

2. Results and Discussion

A particular chemical process that has drawn our attention is the bio-reduction of copper (II) sulfate promoted by glucose as a reducing agent, which has been used as a catalytic source for CuAAC reactions [14]. Inspired by these features, we proceeded to adapt the methodologies developed by our group to obtain exclusively copper acetylides. Thus, the successively addition of Fehling A and B solutions to a mixture of glucose–phenylacetylene produced a characteristic yellow precipitate, which was identified as phenylethynylcopper (I) 1, presenting the same physical data described in the literature (Scheme 1) [12,13]. In addition, XPS analysis of compound 1 displays a signal 2p 3/2 at 935 eV corresponding to a copper species with an oxidation number (I) in agreement with previous reports (Figure 1) [13,14,15,16,17], whereas a signal at 533.8 eV (spectrum b, Figure 1) suggests a Cu–C interaction [18,19]; moreover, the C 1s bond energy at 285.8 eV assigned to aromatic carbons confirmed the formation of copper (I) phenylacetylide 1, which was used in the following steps.
The reaction between 1-ethynylcyclohexanol 2 and benzylazide 3 was used as a model to evaluate the catalytic activity of phenylacetylide 1 (Scheme 2). In all cases, 1-(1-benzyl-1,2,3-triazol-4-yl)cyclohexanol 4 was obtained as the only reaction product. The results showed in Table 1 indicate that the best conditions were obtained using a 0.5 mg/mmol catalyst after 24 h of using CH2Cl2 as a solvent. The conditions found were extended to a series of diverse alkynes and both benzyl azide 3 and 1,3-diazidopropan-2-ol, affording the corresponding 1,2,3-triazoles in 70–90% yields (Table 2) with a broad functional group tolerance and without other kinds of additives.
A particular group of triazoles synthesized by this protocol contains a 1,3-bis-(1,2,3-triazol-1-yl)-propan-2-ol core, which has been recognized as a potential building block to antifungal compound development due to its similarity to the fluconazole structure [20]. Thus, compounds 1113 were obtained in 80–92% through this simple and mild protocol, opening possibilities to develop new potential antifungal drugs; hence, future studies are glimpsed to determine biological properties for these compounds.

3. Experimental

The starting materials were purchased from Aldrich Chemical Co. and were used without further purification. The solvents were distilled before use. Silica plates of 0.20 mm thickness were used for thin layer chromatography. Melting points were determined with a Krüss Optronic melting point apparatus, and they were uncorrected. 1H and 13C NMR spectra were recorded using a Bruker Avance 300-MHz; the chemical shifts (δ) are given in ppm relative to TMS as an internal standard (0.00). For analytical purposes, the mass spectra were recorded on a Shimadzu GCMS-QP2010 Plus in the EI mode, 70 eV, and 200 °C via direct inlet probe. Only the molecular and parent ions (m/z) are reported. IR spectra were recorded on a Bruker Tensor 27 equipment.
The XPS wide and narrow spectra were acquired with a JEOL JPS-9200, equipped with a Mg X-ray source (1253.6 eV) at 250 W over an analysis area of 3 mm2 under vacuum on the order of 1 × 10−8 Torr for all samples. The spectra were analyzed using the specsurf™ software included with the instrument, and all spectra were charge-corrected by means of the carbon signal (C1s) at 284 eV. The Shirley method was used for background subtraction, whereas curve fitting was performed with the Gauss–Lorentz method. Samples were directly deposited on the sample holder and analyzed without any further preparation.

3.1. Copper (I) Phenylacetylide (1)

Phenylacetylene (0.109 mL, 0.102 g, 1 mmol) was added to a solution of glucose (0.45 g, 2.5 mmol) in H2O (5 mL) and MeOH (15 mL). The mixture was treated successively with tartrate–NaOH solution (Fehling B solution, 1.0 mL) and CuSO4 solution (Fehling A solution, 2.5 mL, 1.0 mmol). The resulting mixture was stirred for 24 h at room temperature. The solid was filtered and washed with cold diethyl ether (10 mL), MeOH (20 mL), and H2O (20 mL). The product was dried under reduced pressure. Yield: 0.156 g (95%), m.p. 290 °C. IR (ATR) νmax 3048, 2100 cm−1. Elemental analysis calculated: C, 58.35; H, 3.06; found: C, 58.91; H, 3.19.

3.2. General Procedure for the Synthesis of 1,2,3-Triazoles Catalyzed by Copper Phenylacetylide

Copper (I) phenylacetylide 1 (0.05 g, 0.025 mmol) was added to a stirred solution containing the corresponding alkyne (1.0mmol) and the appropriate azide (1.0 mmol) in CH2Cl2 (10 mL). The resulting reaction mixture was stirred at room temperature for 24 h. The mixture was filtered through celite. The solvent was removed under reduced pressure, and the final product was purified by crystallization.

3.2.1. 1-(1-Benzyl-1,2,3-triazol-4-yl)cyclohexanol 4

1-Ethynylcyclohexanol and benzyl azide afforded 1-(1-benzyl-1,2,3-triazol-4-yl)-cyclohexanol 4 as a white solid, m.p. 150°C. Yield: 0.198 g (77%). IR (ATR) νmax 3386, 3291, 2930, 2855, 1604 cm−1. 1H NMR (300 MHz, DMSO-d6) δ 7.98 (s, 1H), 7.39 (m, 5H), 5.60 (s, 1H), 4.92 (s, 1H), 1.92–1.46 (m, 10H); 13C NMR (75 MHz, DMSO-d6) δ 136.2, 128.7, 128.5, 128.0, 121.1, 68.0, 52.7, 37.8, 25.224, 21.6. MS [EI+] m/z (%): 257 [M]+ (20), 91 [C6H5CH2]+ (100).

3.2.2. 1-Benzyl-4-phenyl-1,2,3-triazole 5

Phenylacetylene and benzyl azide afforded 1-Benzyl-4-phenyl-1,2,3-triazole 5 as a white solid, m.p. 131 °C. Yield: 0.188 g (80%). IR (ATR) νmax 3250, 2850, 1650, 1600 cm−1. 1H NMR (300 MHz, CDCl3) δ 7.82 (m, 2H), 7.68 (s, 1H), 7.41 (m, 4H), 7.33 (m, 1H), 5.59 (s, 2H); 13C NMR (75 MHz, CDCl3) δ 148.2, 134.6, 130.5, 129.1, 128.8, (2 X CH), 128.7, 127.9, 125.6, 119.5, 54.2. MS (EI+) m/z (%): 235[M]+ (21), 206 [M − HN2]+ (74), 116 [M − C6H5N3]+ (100).

3.2.3. 1-Benzyl-4-(4-chlorophenoxymethyl)-1,2,3-triazole 6

1-Chloro-4-prop-2-ynyloxybenzene and benzyl azide afforded 1-benzyl-4-(4-chlorophenoxymethyl)-1,2,3-triazole 6 as a white solid, m.p. 103 °C. Yield: 0.207 g (70%). IR (ATR) νmax 1650, 1600 cm−1. 1H NMR (300 MHz, CDCl3) δ 7.38 (m, 3H), 7.27 (m, 2H), 7.22 (dd, 2H, J = 3Hz, J = 9Hz), 6.89 (dd, 2H, J = 2Hz, J = 9Hz), 5.54 (s, 2H), 5.16 (s, 2H); 13C NMR (75 MHz, CDCl3) δ 54.2 (CH2), 62.2(CH2), 116.0 (2 × CH), 122.6 (CH), 126.1 (C), 128.0 (2 × CH), 128.8, 129.4 (2 × CH) 129.7 (2 × CH), 134.3 (C), 144.1 (C), 156.7 (C). MS (EI+) m/z (%): 299 [M]+ (15), 91 [C6H5CH2]+ (100).

3.2.4. 1-Benzyl-4-(4-nitrophenoxymethyl)-1,2,3-triazole 7

1-Nitro-4-prop-2-ynyloxybenzene and benzyl azide afforded 1-benzyl-4-(4-nitrophenoxymethyl)-1,2,3-triazole 7 as a white solid, m.p. 95 °C. Yield: 0.269 g (87%). IR (ATR) νmax 3260, 3109, 3084, 2923, 2853, 2129, 1608, 1586, 1383, 1247 (m), 1105 cm−1. 1H NMR (300 MHz, CDCl3) δ 8.24 (d, J = 9Hz, 2H) 7.38 (s, 1H) 7.25 (m, 2H), 7.16 (m, 4 H), 5.16 (s, 2H), 4.53; 13C NMR (75 MHz, CDCl3) δ 164.5, 144.2, 140.1, 134.9, 129.7, 129.4, 128.8, 126.1, 116.0, 63.6, 53.7. MS (EI+) m/z (%): 310 [M]+ (5), 91 [C6H5CH2]+ (100).

3.2.5. 1-Benzyl-4-(4-bromophenoxymethyl)-1,2,3-triazole 8

1-Bromo-4-prop-2-ynyloxybenzene and benzyl azide afforded 1-benzyl-4-(4-bromophenoxymethyl)-1,2,3-triazole 8 as a white solid, m.p. 110 °C. Yield: 0.246 g (72%). IR (ATR) νmax 3260, 3040, 2954, 2926, 2873, 1581, 1487 1105 cm−1. 1H NMR (300 MHz, CDCl3) δ 7.51 (s, 1H), 7.38 (m, 3H), 7.30 (m, 2H), 7.25 (d, 2H, 𝐽 = 8.3Hz), 6.81 (d, 2H, 𝐽 = 8.2 Hz), 5.52 (s, 2H), 5.14 (s, 2H); 13C NMR (75 MHz, CDCl3) δ 157.3, 144.2, 134.4, 132.3, 129.2, 128.9, 128.1, 122.7, 116.6, 113.5, 62.2, 54.3. MS (EI+) m/z (%): 343 [M]+ (5), 91 [C7H7]+ (100).

3.2.6. 1-Benzyl-4-p-tolyloxymethyl-1,2,3-triazole 9

1-Methyl-4-prop-2-ynyloxybenzene and benzyl azide afforded 1-benzyl-4-p-tolyloxymethyl-1,2,3-triazole as a white solid, m.p. 111 °C. Yield: 0.228 g (82%). IR (ATR) νmax 3212, 2954, 2919, 2869, 1607, 1287 cm−1. 1H NMR (300 MHz, CDCl3) δ 7.50 (s, 1H), 7.34 (m, 3H), 7.24 (m, 2H), 7.04 (d, 2H, 𝐽 = 8.2Hz), 6.83 (d, 2H, 𝐽 = 8.2Hz), 5.50 (s, 2H), 5.54 (s, 2H), 5.15 (s, 2H), 2.27 (s, 3H); 13C NMR (75 MHz, CDCl3) δ 156.1, 144.9, 134.5, 130.1, 129.9, 129.1, 128.7, 128.1, 122.5, 114.7, 62.3, 54.2, 23.5. MS (EI+) m/z (%): 279 [M]+ (20), 91 [C7H7]+ (100).

3.2.7. 1-Benzyl-4-(naphthalen-1-yloxymethyl)-1,2,3-riazole 10

1-Prop-2-ynyloxynaphthalene and benzyl azide afforded 1-benzyl-4-(naphthalen-1-yloxymethyl)-1,2,3-riazole 10 as a white solid, m.p. 76 °C. Yield: 0.252 g (80%). IR (ATR) νmax 3126, 3084, 3066, 3040, 2956, 2924, 2874, 2854, 1579, 1460, 1378, 1267, 1240 cm−1. 1H NMR (300 MHz, CDCl3) δ 8.20 (d, 1H), 7.82 (m 2H), 7.65 (m, 2H), 7.59 (s, 1H), 7.38 (m, 3H), 7.30 (m, 3H), 6.95 (d, 1H), 5.55 (s, 2H), 5.39 (s, 2H); 13C NMR (75 MHz, CDCl3) δ 152.2, 143.1, 132.8, 127.4, 127.0, 126.3, 125.7, 124.1, 124.0, 123.5, 120.7, 120.2, 119.1, 103.7, 60.8, 52.5. MS (EI+) m/z (%): 315 [M]+ (10), 91 [C7H7]+ (100).

3.2.8. 1,3-Bis-[4-(1-hydroxy)cyclohexyl-1,2,3-triazol-1-yl]propan-2-ol 11

1-Ethynylcyclohexanol 2 and 1,3-diazidopropan-2-ol afforded 1,3-Bis-[4-(1-hydroxy)cyclohexyl-1,2,3-triazol-1-yl]propan-2-ol 11 as a white solid, m.p. 100 °C. Yield: 0.319 g (82%). IR (ATR) νmax 3270, 3143, 3094, 2915, 2874, 1605 cm−1. 1H NMR (300 MHz, CDCl3) δ 8.21 (s, 2H), 5.04 (s, 1H), 4.60–4.51 (m), 4.20–3.89 (m), 2.22–2.00 (m), 1.87–1.71 (m), 1.60–1.44 (m). 13C NMR (75 MHz, CDCl3) δ 149.1, 120.3, 72.8, 71.9, 52.9, 36.6, 25.5, 22.6. MS [EI+] m/z (%): 390 [M]+ (5), 210 [C10H16N3O]+ (100).

3.2.9. 1,3-Bis-(4-phenyl-1,2,3-triazol-1-yl)-propan-2-ol 12

Penylacetylene and 1,3-diazidopropan-2-ol afforded 1,3-bis-(4-phenyl-1,2,3-triazol-1-yl)-propan-2-ol 12 as a white solid, m.p. 200 °C. Yield: 0.318 g (92%). IR (ATR) νmax 3368, 3123, 3094, 2936, 1610, 1579, 1557, 1147, 1126 cm−1. 1H NMR (300 MHz, CDCl3) δ 8.53 (s, 2H), 7.85 (d, J = 7.5 Hz, 4H), 7.43 (t, J = 8 Hz, 4H), 7.31 (t, J = 7.5 Hz, 2H), 5.78 (d, J = 5 Hz, 1H), 4.60 (d, J = 10.5 Hz, 2H), 4.41 (d, J = 10.5 Hz, 2H). 13C NMR (75 MHz, CDCl3) δ 146.1, 130.8, 128.8, 127.7, 125.0, 122.4, 68.2, 53.2. MS [EI+] m/z (%): 346 [M]+ (100).

3.2.10. 1,3-Bis-[4-(4-chlorophenoxymethyl)-1,2,3-triazol-1-yl]-propan-2-ol 13

1-Chloro-4-prop-2-ynyloxybenzene and 1,3-diazidopropan-2-ol afforded 1,3-bis-[4-(4-chlorophenoxymethyl)-1,2,3-triazol-1-yl]-propan-2-ol 13 as a white solid, m.p. 119 °C. Yield: 0.421 g (89%). IR (ATR) νmax 3397, 3128, 3095, 2923, 2853, 1578, 1487, 1386 cm−1. 1H NMR (300 MHz, CDCl3) δ 7.75 (s, 1H), 7.38 (d, J = 8.3 Hz, 3H), 7.26 (d, J = 1.4 Hz, 3H), 6.87 (d, J = 8.0 Hz, 3H), 5.18 (s, 2H), 4.53 (d, J = 14.9 Hz, 1H), 4.38 (s, 1H); 13C NMR (75 MHz, CDCl3) δ 158., 134.9, 130.3, 124.8, 121, 6, 68.5, 61.9, 53.1. MS [EI+] m/z (%) 474 [M]+ (100).

4. Conclusions

1,3-Bis-1,2,3-triazol-1-yl-propan-2-ol-based compounds are easily available from CuAAC reactions using phenylacetylide 1 as an inexpensive catalyst obtained from glucose-promoted bio-reduction of a Fehling reagent in the presence of phenylacetylene through a mild synthetic protocol that does not requires other additives with high functional group tolerance. The simplicity of this synthetic method suggests that this route to 1,2,3-triazoles will enjoy widespread application.

Author Contributions

Conceptualization, E.C.-Y.; methodology, J.V.-P. and J.G.; formal analysis, G.L.-T.; investigation, J.V.-P., J.G., E.C.-Y., G.L.-T., M.A.G.-E., and M.V.B.U.; resources, E.C.-Y.; data curation, G.L.-T., M.A.G.-E., and M.V.B.U.; writing—original draft preparation, E.C.-Y.; writing—review and editing, E.C.-Y.; supervision, E.C.-Y., M.A.G.-E., and M.V.B.U.; project administration, E.C.-Y.; funding acquisition, E.C.-Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CONACYT-Mexico, project No. A1-S-18230.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Financial support from CONACYT is gratefully acknowledged. The authors would like to thank N. Zavala, A. Nuñez, L. Triana, and M. C. Martínez for the technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Y.; Tong, Z.R. Click Chemistry: Approaches, Applications and Chellenges; Nova Science Publishers: New York, NY, USA, 2017. [Google Scholar]
  2. Chandrasekaran, S. Click Reactions in Organic Synthesis; Wiler-VCH: Wienheim, Germany, 2016. [Google Scholar]
  3. El-Azab, A.S.; Abdel-Aziz, A.A.M. Click Chemistry and Applications; LAP Lambert Academic Publishing: Saarbrücken, Germany, 2014. [Google Scholar]
  4. Lahann, J. Click Chemistry for Biotechnology and Materials Science; JohnWiley & Sons: Chichester, UK, 2009. [Google Scholar]
  5. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  6. Bock, V.D.; Hiemstra, H.; van Maarseveen, J.H. CuI-Catalyzed Alkyne–Azide “Click” Cycloadditions from a Mechanistic and Synthetic Perspective. Eur. J. Org. Chem. Soc. 2006, 2006, 51–68. [Google Scholar] [CrossRef]
  7. Buckley, B.R.; Dann, S.E.; Heaney, H. Experimental Evidence for the Involvement of Dinuclear Alkynylcopper(I) Complexes in Alkyne–Azide Chemistry. Chem. Eur. J. 2010, 16, 6278–6284. [Google Scholar] [CrossRef] [PubMed]
  8. Sladkov, A.M.; Ukhin, L.Y. Copper and Silver Acetylides in Organic Synthesis. Russ. Chem. Rev. 1968, 37, 748–763. [Google Scholar] [CrossRef]
  9. Buckley, B.R.; Dann, S.E.; Heaney, H.; Stubbs, E.C. Heterogeneous Catalytic Reactions “On Water” by Using Stable Polymeric Alkynylcopper(I) Pre-Catalysts: Alkyne/Azide Cycloaddition Reactions. Eur. J. Org. Chem. 2011, 2011, 770–776. [Google Scholar] [CrossRef]
  10. Buckley, B.R.; Dann, S.E.; Harris, D.P.; Heaney, H.; Stubbs, E.C. Alkynylcopper(I) polymers and their use in a mechanistic study of alkyne–azide click reactions. Chem. Commun. 2010, 46, 2274–2276. [Google Scholar] [CrossRef] [PubMed]
  11. Evano, G.; Jouvin, K.; Theunissen, C.; Guissart, C.; Laouiti, A.; Tresse, C.; Heimburger, J.; Bouhoute, Y.; Veillard, R.; Lecomte, M.; et al. Turning unreactive copper acetylides into remarkably powerful and mild alkyne transfer reagents by oxidative umpolung. Chem. Commun. 2014, 50, 10008–10018. [Google Scholar] [CrossRef] [PubMed]
  12. Díez-González, S. Copper(I)–Acetylides: Access, Structure, and Relevance in Catalysis. Adv. Organomet. Chem. 2016, 66, 93–141. [Google Scholar]
  13. Ramírez-Palma, M.T.; Segura-Arzate, J.; López-Téllez, G.; Cuevas-Yañez, E. Ligand Synthesis Catalyst and Complex Metal Ion: Multicomponent Synthesis of 1,3-Bis(4-phenyl-[1,2,3]triazol-1-yl)-propan-2-ol Copper(I) Complex and Application in Copper-Catalyzed Alkyne-Azide Cycloaddition. J. Chem. 2016, 2016, 6432492. [Google Scholar] [CrossRef]
  14. García, M.A.; Ríos, Z.G.; González, J.; Pérez, V.M.; Lara, N.; Fuentes, A.; González, C.; Corona, D.; Cuevas-Yañez, E. The Use of Glucose as Alternative Reducing Agent in Copper-Catalyzed Alkyne-Azide Cycloaddition. Lett. Org. Chem. 2011, 8, 701–706. [Google Scholar] [CrossRef]
  15. Theunissen, C.; Lecomte, M.; Jouvin, K.; Laouiti, A.; Guissart, C.; Heimburger, J.; Loire, E.; Evano, G. Convenient and Practical Alkynylation of Heteronucleophiles with Copper Acetylides. Synthesis 2014, 46, 1157–1166. [Google Scholar]
  16. Okamoto, Y.; Kundu, S.K. Photoconductive Properties of Arylethynylcopper Polymers. Effects of Structure and Oxygen. J. Phys. Chem. 1973, 77, 2677–2680. [Google Scholar] [CrossRef]
  17. Velasco, B.E.; López-Téllez, G.; González-Rivas, N.; García-Orozco, I.; Cuevas-Yañez, E. Catalytic Activity of Dithioic Acid Copper Complexes in the Alkyne-Azide Cycloaddition. Can. J. Chem. 2013, 91, 292–299. [Google Scholar] [CrossRef]
  18. Alonso, F.; Moglie, Y.; Radivoy, G.; Yus, M. Click chemistry from organic halides, diazonium salts and anilines in water catalysed by copper nanoparticles on activated carbon. Org. Biomol. Chem. 2011, 9, 6385–6395. [Google Scholar] [CrossRef] [PubMed]
  19. Carley, A.F.; Dollard, L.A.; Norman, P.R.; Pottage, C.; Roberts, M.W. The reactivity of copper clusters supported on carbon studied by XPS. J. Electron Spectrosc. Relat. Phenom. 1999, 98–99, 223–233. [Google Scholar] [CrossRef]
  20. Zambrano-Huerta, A.; Cifuentes-Castañeda, D.D.; Bautista-Renedo, J.; Mendieta-Zerón, H.; Melgar-Fernández, R.C.; Pavón-Romero, S.; Morales-Rodríguez, M.; Frontana-Uribe, B.A.; González-Rivas, N.; Cuevas-Yañez, E. Synthesis and in vitro biological evaluation of 1,3-bis-(1,2,3-triazol-1-yl)-propan-2-ol derivatives as antifungal compounds fluconazole analogues. Med. Chem. Res. 2019, 28, 571–579. [Google Scholar] [CrossRef]
Scheme 1. Preparation of copper phenylacetylide 1.
Scheme 1. Preparation of copper phenylacetylide 1.
Chemproc 03 00054 sch001
Figure 1. XPS narrow spectra of phenylacetylide 1 for (a) Cu 2p 3/2, (b) Cu–C at the O (1s) level, and (c) C 1s regions.
Figure 1. XPS narrow spectra of phenylacetylide 1 for (a) Cu 2p 3/2, (b) Cu–C at the O (1s) level, and (c) C 1s regions.
Chemproc 03 00054 g001
Scheme 2. Synthesis of 1,2,3-triazole 4 catalyzed by copper phenylacetylide 1.
Scheme 2. Synthesis of 1,2,3-triazole 4 catalyzed by copper phenylacetylide 1.
Chemproc 03 00054 sch002
Table 1. Synthesis of triazole 4 catalyzed by phenylacetylide 1.
Table 1. Synthesis of triazole 4 catalyzed by phenylacetylide 1.
EntryCatalyst Ratio (mg/mmol)SolventReaction Time (h)%Yield
10.25CH3OH2450
20.25Acetone2458
30.25CH2Cl22460
40.25CH3OH4851
50.25Acetone4855
60.25CH2Cl24863
70.5CH3OH2472
80.5Acetone2470
90.5CH2Cl22475
100.5CH3OH4871
110.5Acetone4871
120.5CH2Cl24876
131CH3OH2474
141Acetone2472
151CH2Cl22477
161.5CH3OH2474
171.5Acetone2474
181.5CH2Cl22476
Table 2. 1,2,3-triazole yields catalyzed by copper phenylacetylide 1.
Table 2. 1,2,3-triazole yields catalyzed by copper phenylacetylide 1.
CompoundAlkyneAzide% Yield
4CH2(CH2CH2)2C(OH)C≡CHPhCH2N377
5PhC≡CHPhCH2N380
64-ClC6H4OCH2C≡CHPhCH2N370
74-NO2C6H4OCH2C≡CHPhCH2N387
84-BrC6H4OCH2C≡CHPhCH2N372
94-CH3C6H4OCH2C≡CHPhCH2N382
10C10H7OCH2C≡CHPhCH2N380
11CH2(CH2CH2)2C(OH)C≡CHN3CH2CH(OH)CH2N383
12PhC≡CHN3CH2CH(OH)CH2N392
134-ClC6H4OCH2C≡CH N3CH2CH(OH)CH2N389
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Varela-Palma, J.; González, J.; Lopez-Téllez, G.; Unnamatla, M.V.B.; García-Eleno, M.A.; Cuevas-Yañez, E. Synthesis of 1,2,3-Triazoles from Alkyne-Azide Cycloaddition Catalyzed by a Bio-Reduced Alkynylcopper (I) Complex. Chem. Proc. 2021, 3, 54. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08384

AMA Style

Varela-Palma J, González J, Lopez-Téllez G, Unnamatla MVB, García-Eleno MA, Cuevas-Yañez E. Synthesis of 1,2,3-Triazoles from Alkyne-Azide Cycloaddition Catalyzed by a Bio-Reduced Alkynylcopper (I) Complex. Chemistry Proceedings. 2021; 3(1):54. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08384

Chicago/Turabian Style

Varela-Palma, Josué, Jaime González, Gustavo Lopez-Téllez, M. V. Basavanag Unnamatla, Marco A. García-Eleno, and Erick Cuevas-Yañez. 2021. "Synthesis of 1,2,3-Triazoles from Alkyne-Azide Cycloaddition Catalyzed by a Bio-Reduced Alkynylcopper (I) Complex" Chemistry Proceedings 3, no. 1: 54. https://0-doi-org.brum.beds.ac.uk/10.3390/ecsoc-24-08384

Article Metrics

Back to TopTop