Next Article in Journal
Cytotoxic Activity of Ursolic Acid Derivatives Obtained by Isolation and Oxidative Derivatization
Next Article in Special Issue
Synthesis and Luminescence Properties of Iridium(III) Azide- and Triazole-Bisterpyridine Complexes
Previous Article in Journal
NMR Solution Structure of a Chymotrypsin Inhibitor from the Taiwan Cobra Naja naja atra
Previous Article in Special Issue
Recent Trends in Bioorthogonal Click-Radiolabeling Reactions Using Fluorine-18
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Well-Defined Diimine Copper(I) Complexes as Catalysts in Click Azide-Alkyne Cycloaddition Reactions

by
Jordi Markalain Barta
and
Silvia Díez-González
*
Department of Chemistry, Imperial College London, Exhibition Road, South Kensington, SW7 2AZ London, UK
*
Author to whom correspondence should be addressed.
Erasmus Student from the Universidad de Barcelona, Barcelona 08130, Spain.
Molecules 2013, 18(8), 8919-8928; https://doi.org/10.3390/molecules18088919
Submission received: 29 June 2013 / Revised: 21 July 2013 / Accepted: 23 July 2013 / Published: 26 July 2013
(This article belongs to the Collection Advances in Click Chemistry)

Abstract

:
A series of 1,4-disubstituted 1,2,3-triazoles have been prepared in high yields while respecting the stringent Click criteria. In these reactions, highly stable pre-formed complexes bearing diimine ligands were used.

Graphical Abstract

1. Introduction

The introduction of copper(I) catalysts in the cycloaddition of azides and terminal alkynes represents one of the latest success stories of organometallic catalysis [1,2,3,4]. Not only is this transformation high yielding and completely regioselective, but also it exemplifies the utility and importance of Click chemistry [5]. Whereas this cycloaddition reaction has been applied in a plethora of fields, the efforts to develop efficient catalytic systems, respectful of the Click criteria, have been significantly fewer. Nevertheless, the use of ligands in the cycloaddition of alkynes and azides has been shown to stabilize the copper(I) centre, increase the catalytic activity, and even modulate it [6]. Among the different families of ligands applied in this reaction, nitrogen-based ones are arguably the most widely used. Whereas Meldal et al. used diisopropylethylamine in their groundbreaking report [1], N,N',N''-pentametyletylentriamine (PMDETA) and tris-triazoles are also very popular choices [6]. However, examples of well-defined catalysts containing nitrogen based remain scarce [7,8,9,10,11]. Most reported examples employ polydentate ligands such as polytriazoles or tren ligands (Figure 1). Particularly relevant to this work are complexes bearing bis(aryl)acenaphthenenquinonediimine (Ar-BIAN) ligands C. These have been shown to lead to the formation of triazoles in conversion ranging from modest to good in THF at 50 °C [11].
Figure 1. Reported preformed copper(I) catalysts with N-ligands.
Figure 1. Reported preformed copper(I) catalysts with N-ligands.
Molecules 18 08919 g001
Herein, we report the catalytic activity of pre-formed copper(I) complexes bearing one or two α-diimine ligands in the preparation of 1,2,3-triazoles from azides and terminal alkynes. The structures of the screened complexes are shown in Figure 2. These complexes are all highly stable and easy to prepare from simple diazabutadiene compounds [12].
Figure 2. Copper(I) catalysts used in this study.
Figure 2. Copper(I) catalysts used in this study.
Molecules 18 08919 g002

2. Results and Discussion

We started our optimization studies with the screening of different solvents. We chose cationic complex 1 and benzyl azide (9a) and phenylacetylene as the model reaction. Using 5 mol% of the copper catalyst, virtually no conversion of the starting materials was observed with MeCN, cyclohexane, or ethyl acetate, whereas poor results were obtained in MeOH or under neat conditions (Table 1). Satisfyingly, complete conversions into triazole 10a were obtained on water and in acetone (Table 1, entries 8 and 9). Similar results were observed in THF or in a mixture water/t-BuOH, however, reactions in these solvents were not always reproducible and they were not studied further (Table 1, entries 6 and 7).
In order to determine the best reaction medium for this reaction, the copper loading was next reduced to 2 mol% with acetone and water (Table 1, entries 10 and 11). A higher conversion of 86% was observed in acetone and hence it was kept as solvent for the rest of the study. It is important to note that all tested solvents were technical grade, and in particular, the acetone employed here was the one normally used for cleaning the glassware in the laboratory.
Table 1. Solvent screening. Molecules 18 08919 i001
Table 1. Solvent screening. Molecules 18 08919 i001
EntrySolvent[Cu] (mol%)Conv. (%) a
1MeCN5<5
2Cyclohexane5<5
3Ethyl Acetate5<5
4MeOH552
5Neat513
6THF584 to <5
7Water/t-BuOH5>95 to 9
8Water5>95
9Acetone5>95
10Water252
11Acetone286
a 1H-NMR conversions are the average of at least two independent reactions.
We next performed a catalyst screening in acetone at room temperature using 2 mol% of different diimine copper complexes (Scheme 1). All the tested complexes performed well in the model reactions with conversion ranging from 42% to complete. For complexes bearing aromatic groups on the ligands, steric hindrance proved to be more beneficial to the reaction outcome than the presence of electron donating groups. In general, cationic complexes performed better than their neutral analogues, except for complexes bearing adamantyl imine ligands 5 and 6. Additionally, the best performing complexes had aliphatic substituents on the diimine ligand(s). Hence, complexes 5, 7 and 8 led to total conversions and further experiments were run with these in order to determine the best performing one.
Scheme 1. Catalyst screening a.
Scheme 1. Catalyst screening a.
Molecules 18 08919 g003
a 1H-NMR conversions are the average of at least two independent reactions.
First, the metal loading was further reduced (Table 2). Similar results were obtained with 1 mol% [Cu], but at 0.5 mol% [Cu] it became apparent that complexes bearing either one or two cyclohexyl diimine ligands 7 and 8 displayed a better activity than complex 5, with adamantyl substituents. Gratifyingly, complex 8 could still achieve total conversion in a more concentrated reaction (1 M rather than 0.5 M), and it was chosen as the optimal catalyst within the tested series.
With an optimized catalytic system in hand, the scope of the reaction was then explored. All reactions were carried out in technical grade acetone, in air and in no cases oxidation to copper(II) or disproportionation to copper(0) and copper(II) were observed. Also, no other by-products were formed and the prepared triazoles could be easily isolated as pure products in high yields after a simple extraction (Scheme 2).
A number of functional groups were well tolerated, such as alcohols, amines, alkenes and nitriles. Benzyl, alkyl and aryl azides well suitable cycloaddition partners, as well as electron rich or electron poor alkynes. In some cases, total conversions were not reached under the optimized conditions. However, a slight increase in the reaction temperature (from RT to 40 °C) or the metal loading was enough to ensure a high isolated yield.
Table 2. Final optimization reactions. Molecules 18 08919 i002
Table 2. Final optimization reactions. Molecules 18 08919 i002
Entry[Cu]Concentration (M)Conv. (%) a
15, 1 mol%0.5<95
27, 1 mol%0.5<95
38, 1 mol%0.5<95
45, 0.5 mol%0.563
57, 0.5 mol%0.592
68, 0.5 mol%0.5<95
77, 0.5 mol%1.093
88, 0.5 mol%1.0<95
a 1H-NMR conversions are the average of at least two independent reactions.
Scheme 2. Preparation of 1,2,3-triazoles catalyzed by complex 8 a.
Scheme 2. Preparation of 1,2,3-triazoles catalyzed by complex 8 a.
Molecules 18 08919 g004
a Isolated yields are the average of at least two independent reactions; b 2 mol% 8; c 40 °C; d 1 mol% 8.

3. Experimental Section

3.1. General

All reagents were commercially available and used as received. 1H-NMR (400 MHz) and 13C-NMR (100 MHz) spectra were recorded in CDCl3 on a Bruker AVANCE400 spectrometer at room temperature. Chemical shifts (δ) are reported in ppm and referenced to tetramethylsilane (1H) and deuterated chloroform (13C), respectively. All reactions were carried out in air and using technical solvents with no particular precautions to exclude oxygen or moisture. All reported yields are isolated yields and are the average of at least two independent reactions. Benzyl azide (9a) [13] 2-azidoethyl-diisopropylamine (9b) [14], 2-(2-azidoethyl)-1,3-dioxolane (9c) [15], (2-azidoethyl)benzene (9d) [13], 3-(azidoprop-1-en-yl)benzene (9e) [16], phenyl azide (9f) [17], 6-azidohexanenitrile (9g) [18] and (1-azidoethy)benzene (9h) [13] are known in the literature and the corresponding spectroscopic data for all these compounds were in good agreement with the reported data.

3.2. Catalytic Results

General Procedure for the [3+2] Cycloaddition of Azides and Terminal Alkynes
In a vial fitted with a screw cap, azide 9 (1 mmol), alkyne (1 mmol), acetone (1 mL) and 8 (3 mg, 0.5 mol%) were loaded. The reaction was allowed to proceed at room temperature overnight (20 h). The reaction mixture was then hydrolyzed with a saturated aqueous solution of NH4Cl (1 h) and extracted with EtOAc. In all examples, the crude products were estimated to be greater than 95% pure by 1H-NMRs.
1-Benzyl-4-phenyl-1H-1,2,3-triazole (10a). Using the general procedure 223 mg (95%) of the title compound were prepared from benzyl azide (9a, 125 μL) and phenylacetylene (112 μL). Spectroscopic data for 10a were consistent with previously reported data for this compound [19]. 1H-NMR (CDCl3): δ 7.80 (d, J = 7.0 Hz, 2H, HAr), 7.66 (s, 1H, NCH = C), 7.42–7.37 (m, 5H, HAr), 7.34–7.31 (m, 3H, HAr), 5.59 (s, 2H, PhCH2). 13C-NMR (CDCl3): δ 148.2 (C, =C–Ph), 134.7 (C, CAr), 130.5 (C, CAr), 129.1 (CH, CHAr), 128.8 (CH, CHAr), 128.8 (CH, CHAr), 128.1 (CH, CHAr), 128.0 (CH, CHAr), 125.7 (CH, CHAr), 119.5 (CH, NCH=), 54.2 (CH2).
1-Benzyl-4-butyl-1H-1,2,3-triazole (10b). Using the general procedure 199 mg (92%) of the title compound were prepared from benzyl azide (9a, 125 μL), 1-hexyne (115 μL), and 8 (12 mg, 2 mol%) at 40 °C, Spectroscopic data for 10b were consistent with previously reported data for this compound [10]. 1H-NMR (CDCl3): δ 7.40–7.34 (m, 3H, HAr), 7.26–7.24 (m, 2H, HAr), 7.18 (s, 1H, NCH=C), 5.50 (s, 2H, PhCH2N), 2.69 (t, J = 7.5 Hz, 2H, C=CNCH2), 1.62 (quintet, J = 7.5 Hz, 2H, CH2CH2CH2), 1.37 (sextet, J = 7.3 Hz, 2H, CH2CH3), 0.91 (t, J = 7.3 Hz, 3H, CH3). 13C-NMR (CDCl3): δ 149.8 (C, NC=C-butyl), 135.0 (C, CAr), 129.0 (CH, CHAr), 128.6 (CH, CHAr), 127.9 (CH, CHAr), 120.9 (CH, NCH=), 54.0 (CH2, PhCH2), 31.4 (CH2), 25.4 (CH2), 22.3 (CH2), 13.8 (CH3).
1-Benzyl-4-(2-hydroxypropan-2-yl)-1H-1,2,3-triazole (10c). Using the general procedure 183 mg (84%) of the title compound were prepared from benzyl azide (9a, 125 μL), 2-methylbut-3-yn-2-ol (100 μL), and 8 (6 mg, 1 mol%) at 40 °C. Spectroscopic data for 10c were consistent with previously reported data for this compound [19]. 1H-NMR (CDCl3): 7.39–7.35 (m, 3H, HAr), 7.34 (s, 1H, NCH=C), 7.31–7.27 (m, 2H, HAr), 5.51 (s, 2H, CH2), 2.34 (br s, 1H, OH), 1.61 (s, 6H, CH3). 13C-NMR (CDCl3): δ 156.0 (C, NC=C–), 134.6 (C, CAr), 129.2 (CH, CHAr), 128.6, (CH, CHAr), 128.2 (CH, CHAr), 119.1 (CH, NCH=), 68.3 (C, C(OH)Me2), 54.2 (CH2, PhCH2), 30.5 (CH3).
1-(2-Diisopropylaminoethyl)-4-phenyl-1H-1,2,3-triazole (10d). Using the general procedure 232 mg (85%) of the title compound were prepared from azide 9b (170 mg) and phenylacetylene (112 μL). 1H-NMR (CDCl3): δ 7.84–7.81 (m, 3H, HAr), 7.45 (t, J = 6.4 Hz, 2H, HAr), 7.35 (t, J = 6.4 Hz, 1H, HAr), 4.38 (t, J = 6.5 Hz, 2H, CH2), 3.04 (septet, J = 6.5 Hz, 2H, CH(CH3)2), 2.95 (t, J = 6.5 Hz, 2H, CH2), 0.99 (d, J = 6.5 Hz, 12H, CH3); 13C-NMR (CDCl3): δ 147.1 (NCH=C), 130.9 (C, CAr), 128.8 (CH, CHAr), 127.9 (CH, CHAr), 125.7 (CH, CHAr), 120.8 (CH, NCH=), 51.2 (CH2, CH2N), 48.7 (CH, CHN), 45.7 (CH2, CH2N), 20.8 (CH3). HRMS calculated for C16H25N4: 273.2079; found 273.2086 [(M+H)+].
Ethyl 1-(2-(1,3-dioxolan-2-yl)ethyl)-1H-1,2,3-triazole-4-carboxylate (10e). Using the general procedure 203 mg (84%) of the title compound were prepared from azide 9c (143 mg), ethyl propiolate (102 μL), and 8 (6 mg, 1 mol%). Spectroscopic data for 10e were consistent with previously reported data for this compound [20]. 1H-NMR (CDCl3): δ 8.12 (s, 1H, NCH=C), 4.93 (t, J = 4.0 Hz, 1H, OCHO), 4.58 (t, J = 7.0 Hz, 2H, CH2CH2N), 4.43 (q, J = 7.0 Hz, OCH2CH3), 4.03–3.93 (m, 2H, OCH2CH2O), 3.93–3.83 (m, 2H, OCH2CH2O), 2.33 (dt, J = 4.0; 7.0 Hz, CHCH2CH2), 1.41 (t, J = 7.0 Hz, 3H, OCH2CH3). 13C-NMR (CDCl3): δ 160.2 (C, C=O), 139.3 (C, =C–CO2Et), 127.5 (CH, CH=C), 127.5 (CH, O–CH–O), 64.55 (CH2, O–CH2–CH2–O), 64.48 (CH2, O–CH2–CH2–O), 60.6 (CH2, OCH2–CH3), 45.0 (CH2, CH2CH2–N), 33.2 (CH2, CH2–CH2–N), 13.7 (CH3).
N,N-Dimethyl-1-(1-phenethyl-1H-1,2,3-triazol-4-yl)methanamine (10f). Using the general procedure 202 mg (88%) of the title compound were prepared from (2-azidoethyl)benzene (9d, 147 mg), dimethylprop-2-ynylamine (111 μL) and 8 (6 mg, 1 mol%). Spectroscopic data for 10f were consistent with previously reported data for this compound [20]. 1H-NMR (CDCl3): δ 7.31–7.21 (m, 4H, HAr + NCH=C), 7.09 (d, J = 7.0 Hz, 2H, HAr), 4.58 (t, J = 7.0 Hz, 2H, PhCH2CH2), 3.57 (s, 2H, CH2NMe2), 3.21 (t, J = 7.0 Hz, 2H, PhCH2CH2), 2.23 (s, 6H, NMe2). 13C-NMR (CDCl3): δ 144.7 (C, NCH=C), 137.0 (C, CAr), 128.7 (CH, CHAr), 128.6 (CH, CHAr), 126.9 (CH, CHAr), 122.6 (CH, NCH=), 54.2 (CH2, PhCH2CH2N), 51.4 (CH2, CH2NMe2), 44.9 (CH2, PhCH2), 36.6 (CH3).
4-Phenyl-1-(3-phenyl-2-propenyl)-1H-1,2,3-triazole (10g). Using the general procedure 250 mg (96%) of the title compound were prepared from azide 9e (159 mg) and phenylacetylene (112 μL). Spectroscopic data for 10g were consistent with previously reported data for this compound [21]. 1H-NMR (CDCl3): δ 7.84–7.82 (m, 3H, HAr + NCH=C), 7.43–7.39 (m, 4H, HAr), 7.36–7.32 (m, 4H, HAr), 6.72 (d, J = 16.0 Hz, 1H, PhCH=CH), 6.40 (dt, J = 16.0, 6.5 Hz, 1H, PhCH=CH), 5.20 (d, J = 6.5 Hz, 2H, CH2N). 13C-NMR (CDCl3): δ 148.0 (C, NCH=C), 135.4 (C, CAr), 135.2 (CH, CHAr), 130.5 (C, CAr), 128.7 (CH, CHAr), 128.6 (CH, CHAr), 128.4 (CH, CHAr), 128.0 (CH, CHAr), 126.6 (CH, PhCH=), 125.6 (CH, PhCH=CH), 121.8 (CH, CHAr), 119.3 (CH, NCH=), 52.3 (CH2, CH2N).
Dimethyl(1-phenyl-1H-1,2,3-triazol-4-ylmethyl)amine (10h). Using the general procedure 196 mg (97%) of the title compound were prepared from phenyl azide (9f, 107 μL) and dimethylprop-2-ynylamine (111 μL). Spectroscopic data for 10h were consistent with previously reported data for this compound [22]. 1H-NMR (CDCl3): δ 7.95 (s, 1H, NCH=C), 7.74 (d, J = 8.0 Hz, 2H, HAr), 7.53 (t, J = 8.0, 2H, HAr), 7.43 (t, J = 8.0 Hz, 1H, HAr), 3.71 (s, 2H, CH2NMe2), 2.34 (s, 6H, NMe2). 13C-NMR (CDCl3): δ 146.0 (C, NCH=C), 137.0 (C, CAr), 129.6 (CH, CHAr), 128.5 (CH, CHAr), 120.4 (CH, NCH=), 120.3 (CH, CHAr), 54.3 (CH2, CH2N), 45.2 (CH3).
6-(4-Phenyl-1,2,3-triazol-1-yl)hexanenitrile (10i). Using the general procedure 227 mg (94%) of the title compound were prepared from 6-azidohexanenitrile (9g, 138 mg) and phenylacetylene (112 μL). Spectroscopic data for 10i were consistent with previously reported data for this compound [23]. 1H-NMR (CDCl3): δ 7.85–7.82 (m, 2H, HAr), 7.76 (s, 1H, NCH=), 7.45–7.42 (m, 2H, HAr), 7.36–7.31 (m, 1H, HAr), 4.44 (t, J = 7.0 Hz, 2H, CH2N), 2.36 (t, J = 7.0 Hz, 2H, CH2CN), 2.08–1.97 (m, 2H,CH2), 1.77–1.67 (m, 2H,CH2), 1.59–1.48 (m, 2H, CH2). 13C-NMR (CDCl3): δ 147.8 (C, C=C–Ph), 130.5 (C, CAr), 128.8 (CH, CAr), 128.1 (CH, CAr), 125.6 (CH, CAr), 119.5 (CH, CH=C–Ph), 119.3 (C, CN), 49.8 (CH2, CH2–N), 29.5 (CH2), 25.5 (CH2), 24.7 (CH2), 17.0 (CH2).
4-Cyclopropyl-1-(1-phenylethyl)-1H-1,2,3-triazole (10j). Using the general procedure 183 mg (85%) of the title compound were prepared from azide 9h (147 mg), ethynylcyclopropane (87 μL) and 8 (6 mg, 1 mol%) at 40 °C. Spectroscopic data for 10j were consistent with previously reported data for this compound [23]. 1H-NMR (CDCl3): δ 7.40–7.28 (m, 3H, HAr), 7.27–7.24 (m, 2H, HAr), 7.11 (s, 1H, NCH=C), 5.76 (q, 1H, J = 7.0 Hz, PhCH), 1.95 (d, 3H, J = 7.0 Hz, CH3), 1.93–1.87 (m, 1H, CHcyclopropyl), 0.94–0.87 (m, 2H, CH2), 0.84–0.79 (m, 2H, CH2). 13C-NMR (CDCl3): δ 150.2 (C=Ccyclopropyl), 140.1 (C, CAr), 128.9 (CH, CHAr), 128.3 (CH, CHAr), 126.4 (CH, CHAr), 118.3 (CH, N–CH=C), 58.9 (CH, PhCH), 21.2 (CH3), 7.6 (CH, CHcyclopropyl), 6.7 (CH2, CH2).

4. Conclusions

Diimine copper(I) complexes are highly efficient catalysts for the [3+2] cycloaddition of azides and terminal alkynes. Low copper loadings were enough to ensure high isolated yields of 1,2,3-triazoles at room temperature, in air and in technical grade acetone. Furthermore, purification by column chromatography was not necessary, and a simple extraction was enough to isolate pure triazoles. The reported catalysts are noticeably more active than related Ar-BIAN complexes. This might be attributed to two factors: (1) the presence of alkyl groups on the diimine ligands; which in these series systematically outperformed aryl diimines; or (2) the increased chemical stability of simple diimine scaffolds, in particular towards oxidation and reduction reactions.

Acknowledgments

Imperial College London is acknowledged for financial support. J.M.B. thanks the Universidad de Barcelona for an Erasmus Scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tornøe, C.W.; Christensen, C.; Meldal, M. Peptidotriazoles on solid phase: [1,2,3]-Triazoles by regiospecific copper(I)-catalyzed 1,3-dipolar cycloaddition of terminal alkynes to azides. J. Org. Chem. 2002, 67, 3057–3064. [Google Scholar] [CrossRef]
  2. Rostovtsev, V.V.; Green, L.G.; Fokin, V.V.; Sharpless, K.B. A stepwise huisgen cycloaddition process: Copper(I)-catalyzed regioselective “ligation” of azides and terminal alkynes. Angew. Chem. Int. Ed. 2002, 41, 2596–2599. [Google Scholar] [CrossRef]
  3. L’abbe, G. Are azidocumulenes accessible? Bull. Soc. Chem. Belg. 1984, 93, 579–592. [Google Scholar] [CrossRef]
  4. Special Issue on Click Chemistry. Chem. Soc. Rev. 2010, 39, 1221–1408. [CrossRef]
  5. Kolb, H.C.; Finn, M.G.; Sharpless, K.B. Click chemistry: Diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  6. Díez-González, S. Well-defined copper(I) complexes for Click azide-alkyne cycloaddition reactions: One Click beyond. Catal. Sci. Tech. 2011, 1, 166–178. [Google Scholar] [CrossRef]
  7. Donnelly, P.S.; Zanatta, S.D.; Zammit, S.C.; White, J.M.; Williams, S.J. “Click” cycloaddition catalysts: Copper(I) and copper(II) tris(triazolylmethyl)amine complexes. Chem. Commun. 2008, 2008, 2459–2461. [Google Scholar]
  8. Chan, T.R.; Fokin, V.V. Polymer-supported copper(i) catalysts for the experimentally simplified azide-alkyne cycloaddition. QSAR Comb. Sci. 2007, 26, 1274–1279. [Google Scholar] [CrossRef]
  9. Özçubukçu, S.; Ozkal, E.; Jimeno, C.; Pericàs, M.A. A highly active catalyst for Huisgen 1,3-dipolar cycloadditions based on the tris(triazolyl)methanol-Cu(I) structure. Org. Lett. 2009, 11, 4680–4683. [Google Scholar] [CrossRef]
  10. Candelon, N.; Lastécouères, D.; Diallo, A.K.; Aranzaes, J.R.; Astruc, D.; Vincent, J.M. A highly active and reusable copper(I)-tran catalyst for the “click” 1,3-dipolar cycloaddition of azides and alkynes. Chem. Commun. 2008, 2008, 741–743. [Google Scholar]
  11. Li, L.; Lopez, P.S.; Rosa, V.; Figueira, C.A.; Lemos, M.A.N.D.A.; Duarte, M.T.; Avilés, T.; Gomes, P.T. Synthesis and structural characterisation of (aryl-BIAN)copper(I) complexes and their application as catalysts for the cycloaddition of azides and alkynes. Dalton Trans. 2012, 41, 5144–5154. [Google Scholar] [CrossRef]
  12. Frutos Pedreño, R.; Markalain-Barta, J.; Vega Isa, E.; Díez-González, S. Manuscript in preparation. 2013. [Google Scholar]
  13. Alvarez, S.G.; Alvarez, M.T. A practical procedure for the synthesis of alkyl azides at ambient temperature in demethylsulfoxide in high purity and yield. Synthesis 1997, 1997, 413–414. [Google Scholar] [CrossRef]
  14. Engler, A.C.; Bonner, D.K.; Buss, H.G.; Cheung, E.Y.; Hammond, P.T. The synthetic tuning of clickable pH responsive cationic polypeptides and block copolypeptides. Soft Matter 2011, 7, 5627–5637. [Google Scholar] [CrossRef]
  15. Carboni, B.; Vaultier, M.; Carrié, R. Etude de la chimioselectivite de la reaction des dichloroboranes avec les azides fonctionnels: une synthese efficace d'amines secondaires fonctionnalisees. Tetrahedron 1987, 43, 1799–1810. [Google Scholar] [CrossRef]
  16. Van Kalkeren, H.A.; Bruins, J.J.; Rutjes, F.P.J.T.; van Delft, F.L. Organophosphorus-catalysed staudinger reduction. Adv. Synth. Catal. 2012, 354, 1417–1421. [Google Scholar] [CrossRef]
  17. Rehse, K.; Cwiklicki, A. Antiaggregating and antithrombotic activities of new 1,2,3-triazolecarboxamides. Arch. Pharm. Pharm. Med. Chem. 2004, 337, 156–163. [Google Scholar] [CrossRef]
  18. Luo, L.; Wilhelm, C.; Sun, A.; Grey, C.P.; Lauher, J.W.; Goroff, N.S. Poly(diiododiacetylene): Preparation, isolation, and full characterization of a very simple poly(diacetylene). J. Am. Chem. Soc. 2008, 130, 7702–7709. [Google Scholar]
  19. Appukkuttan, P.; Dehaen, W.; Fokin, V.V.; van der Eycken, E. A microwave-assisted Click chemistry synthesis of 1,4-disubstituted 1,2,3-triazoles via a copper(I)-catalyzed three-component reaction. Org. Lett. 2004, 6, 4223–4225. [Google Scholar] [CrossRef]
  20. Díez-González, S.; Nolan, S.P. [(NHC)2Cu]X complexes as efficient catalysts for azide-alkyne Click chemistry at low catalyst loadings. Angew. Chem. Int. Ed. 2008, 47, 8881–8884. [Google Scholar] [CrossRef]
  21. Kamijo, S.; Jin, T.; Huo, Z.; Yamamoto, Y. A one-pot procedure for the regioconrolled synthesis of allyltriazoles via the Pd-Cu bimetallic catalysed three-component coupling reaction of nonactivated terminal alkynes, allyl carbonate, and trimethylsilyl azide. J. Org. Chem. 2004, 69, 2386–2393. [Google Scholar] [CrossRef]
  22. Yan, Z.Y.; Zhao, Y.B.; Fan, M.J.; Liu, W.M.; Liang, Y.M. General synthesis of (1-substituted-1H-1,2,3-triazol-4-ylmethyl)-dialkylamines via a copper(I)-catalyzed three-component reaction in water. Tetrahedron 2005, 61, 9331–9337. [Google Scholar] [CrossRef]
  23. Lal, S.; Díez-González, S. [CuBr(PPh3)3] for azide-alkyne cycloaddition reactions under strict Click conditions. J. Org. Chem. 2011, 76, 2367–2373. [Google Scholar] [CrossRef]
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Barta, J.M.; Díez-González, S. Well-Defined Diimine Copper(I) Complexes as Catalysts in Click Azide-Alkyne Cycloaddition Reactions. Molecules 2013, 18, 8919-8928. https://doi.org/10.3390/molecules18088919

AMA Style

Barta JM, Díez-González S. Well-Defined Diimine Copper(I) Complexes as Catalysts in Click Azide-Alkyne Cycloaddition Reactions. Molecules. 2013; 18(8):8919-8928. https://doi.org/10.3390/molecules18088919

Chicago/Turabian Style

Barta, Jordi Markalain, and Silvia Díez-González. 2013. "Well-Defined Diimine Copper(I) Complexes as Catalysts in Click Azide-Alkyne Cycloaddition Reactions" Molecules 18, no. 8: 8919-8928. https://doi.org/10.3390/molecules18088919

Article Metrics

Back to TopTop