Abstract

1,2,3-Triazole scaffolds are not obtained in nature, but they are still intensely investigated by synthetic chemists in various fields due to their excellent properties and green synthetic routes. This review will provide a library of all synthetic routes used in the past 21 years to synthesize 1,2,3-triazoles and their derivatives using various metal catalysts (such as Cu, Ni, Ru, Ir, Rh, Pd, Au, Ag, Zn, and Sm), organocatalysts, metal-free as well as solvent- and catalyst-free neat syntheses, along with their mechanistic cycles, recyclability studies, solvent systems, and reaction condition effects on regioselectivity. Constant developments indicate that 1,2,3-triazoles will help lead to future organic synthesis and are useful for creating molecular libraries of various functionalized 1,2,3-triazoles.
1. Introduction
Heterocyclic chemistry is wide, important and most studied discipline of medicinal chemistry. Heterocyclic bioactive molecules have various heteroatoms such as nitrogen,1,2 oxygen,3,4 sulfur,5−8 and other9 atoms which have significant biological applications. The study of heterocyclic bioactive molecules containing nitrogen atoms is one of the most important disciplines of medicinal chemistry.10,11 Among nitrogen-containing heterocycles, azoles are five-membered heterocyclic moieties which are essential structural parts in diverse biologically active natural products.12 Azoles and their derivatives display various biological effects including capable antibacterial activity.13 A 1,2,3-triazole scaffold is a potent nitrogen-bearing heterocyclic scaffold which has found wide applications.14,15 1,2,3-triazole is an unsaturated, π-excessive, five-membered heterocycle with a 6π delocalized electron ring system which gives it an aromatic character. 1,2,3-triazole is made up of three nitrogens and two carbons. All five atoms are sp2-hybridized. One N atom is pyrrole kind, and the other two atoms are pyridine kind. Monocyclic 1,2,3-triazole, 1,2,3-triazolium salt, and benzotriazoles are primary classes of 1,2,3-triazoles. Depending on the location of the NH proton, monocyclic 1,2,3-triazoles are further categorized into three subclasses (Figure 1).
Figure 1.

Isomers of monocyclic 1,2,3-triazole.
The 1H- and 2H-1,2,3-triazoles are aromatic and in equilibrium with each other in solution as well as a gas phase, while 4H-1,2,3-triazole is nonaromatic in nature.16 Among all possible isomers of monocyclic 1,2,3-triazoles, 1H-1,2,3-triazole is a powerful scaffold and widely present in therapeutic agents and has gained more interest due to its medicinal chemistry applications, agrochemicals and material science.17−21 Some synthetic biologically active drugs are mentioned below (Figure 2).
Figure 2.
Synthetic bioactive compound containing a 1,2,3-triazole ring.
Rufinamide (1-[2,6-difluorobenzyl]-1H-1,2,3-triazole-4-carboxamide) IV is a novel 1,2,3-triazole bearing antiepileptic drug known by the trade name Inovelon and Xilep, which was discovered by Novartis Pharmaceutical.22 1,2,3-Triazole thioether V (IC50: 10 nM) is more potent than the original lead molecule RN-18 (IC50: 6 μM) against HIV-1 strain.23,24 1,2,3-Triazole-tethered sulfonamide–berberine hybrid compounds VI showed an antimalarial activity with IC50 = 0.142–28.006 μM in which the R = p-chlorophenylamino substituent was the most potent (IC50 = 0.142 μM).25 Imidazole-bearing triazole VII showed antibacterial activity comparable to or better than that of linezolid and vancomycin against Enterococcus faecium (MIC = 0.5 μg/mL).26 Cefatrizine showed β-lactam antibiotic activity against Serratia marcescens (MIC: >400 μg/mL).27 Cyclotetrapeptide analogues IX showed a 3-fold increase in tyrosinase activity (IC50 = 0.5 mM) compared to that of naturally occurring cyclo-[Pro-Tyr-Pro-Val] (IC50 = 1.5 mM).28 Macrotriazoles (analogous to Migrastatin) X are active against the MDA-MB-361 cell line.29 Tazobactam is also a 1,2,3-triazole-containing β-lactam antibiotic that shows MIC values of 1.56, 0.1, and 3.13 mg/L against Escherichia coli, Staphylococcus aureus, and Klebsiella pneumoniae, respectively.30 1,2,3-Triazole is used as an isostere of carboxylic acid, amide, and ester in the synthesis of many drugs.31 Several review articles covered the synthetic strategy of 1,2,3-triazole derivatives using only transition-metal-catalyzed cycloaddition. Ding and co-workers32 covered the literature on metal-catalyzed cycloaddition of azide and internal alkyne only, whereas da Silva Júnior and co-workers33 presented a review beyond the use of copper-based catalysts due to selective access to 1,5-disubstituted triazole, which involves other transition metal scopes except for copper. In 2020, Ramachary and co-workers34 reviewed the synthesis of vinyl-functionalized 1,2,3-triazole molecules. Sahu and co-workers35 reported an advanced synthetic approach for synthesizing a 1,2,3-triazole scaffold. In 2013, Schubert and co-workers36 reported the coordination and supramolecular chemistry of triazole. Moses and Moorhouse37 gave a detailed application of the click approach. Kappe and Van der Eycken38 discussed nonclassical methods (such as microwave, heating, or continuous flow processing) for the click reaction. Jozwiak and co-workers39 reviewed the role of click reaction in drug development. As copper is cytotoxic, Koo and co-workers40 reported a copper-free click approach as a valuable tool in the biomedical field. So, there are relatively few systematic and recent reviews on 1,2,3-triazole and its derivatives. The present review focuses on the systematic, readable, and researched synthesis of monocyclic 1,2,3-triazoles and their derivatives with their mechanism. The main goal of this review is to present and analyze various methodologies for the synthesis of titled compounds via metal catalysts, organocatalysts, solvent-promoted, catalyst- and solvent-free neat routes which are useful in the future advancement in the synthesis of heterocyclic compounds incorporated with 1,2,3-triazoles (triazole) and their derivatives. Some privileged triazole-bearing scaffolds are shown in Figure 3.
Figure 3.
Some privileged triazole-bearing scaffolds covered in this review.
2. Synthesis of 1,2,3-Triazole via Metal-Catalyzed Azide–Alkyne Cycloaddition
1,2,3-Triazole derivatives were first synthesized in 1960 through a Huisgen 1,3-dipolar cycloaddition (Scheme 1) using azide 2 and alkyne 1 under thermal conditions to produce a mixture of 1,4- and 1,5-disubstituted triazoles 3 and 4.41
Scheme 1. Synthesis of 3 and 4 through Huisgen 1,3-Dipolar Cycloaddition.

The regioselectivity of the approach is controlled by the electronic as well as steric factors of substrates. This approach has not been widely used in synthetic chemistry due to its elevated temperature, low yield, and generation of two regioisomers when using asymmetric alkynes. Later, various strategies were developed to control regioselectivity. With the introduction of a click reaction, traditional 1,3-dipolar cycloaddition failed. Sharpless and co-workers42 gave criteria for “click chemistry” in 2001. Using a copper catalyst, the click approach generates only 1,4-disubstituted 1,2,3-triazole 3 at ambient temperature in an aqueous medium.43 It was later found that the ruthenium catalyst produced only 1,5-disubstituted products 4.44,45 Metal-catalyzed azide–alkyne cycloaddition (MAAC) reactions are discussed below.
2.1. Copper-Catalyzed Azide–Alkyne Cycloaddition (CuAAC)
The CuAAC reaction was discovered independently by two research groups of Sharpless–Folkin46 and Meldal and co-workers47 in 2002. The CuAAC reaction is often quite slow because of the kinetic stability of starting materials, azide 2 and alkynes 1. In addition, the CuAAC reaction required high pressure or temperature, as well as a long time to complete. The Huisgen reaction produces a combination of 1,4- and 1,5-products 3 and 4, whereas with a CuI catalyst or precursors of the CuI catalyst, the reaction of terminal alkynes is selective in the synthesis of 1,4-disubstituted triazoles 3(47,48) (Scheme 2). Using this catalyst, the reaction rate increased up to 107 times in comparison to the catalyst-free reaction. The CuAAC reaction becomes an excellent reaction for organic synthesis. Many click reactions have been reported recently, but none of them can meet all of the standards of the ideal click reaction given by Sharpless and co-workers.48
Scheme 2. Cu Catalyzed Synthesis of 3.

However, the CuAAC reaction has been regarded as the most suitable reaction for click reactions since it fits with the majority of the criteria for a perfect click reaction. Copper-catalyzed reaction of alkynes with organic azides represents the classic example of click chemistry.49
Fokin and co-workers proposed the mononuclear mechanism of the CuAAC catalytical cycle first. The generation of the alkynes Cu(I) complex A initiates the reaction process, and the formation of copper(I) acetylide B occurs. The formation of Cu(I) with a terminal alkyne reduces the pKa of the acetylene proton up to pH 9.8,46,51 providing deprotonation in the aqueous system without the use of the strong base. By coordinating to the metal center, azide is activated. This results in an intermediate between two species—azide and acetylene. The next stage is the generation of the C–N bond, which goes through a strained six-membered copper-containing intermediate. Second, C–N bond formation along with reductive elimination results in the formation of the desired 1,4-disubstituted 1,2,3-triazole 3 byproduct isolation, and catalyst regeneration takes place51 (Figure 4). Fokin and co-workers52 presented a common binuclear mechanism for the CuAAC reaction in which copper(I) acetylide is activated through a second copper center via a fragile interaction to form the binuclear copper intermediate D.50
Figure 4.

Fokin and co-workers proposed a catalytic cycle for the CuAAC reaction. Reprinted from ref (50). Copyright 2005 American Chemical Society.
2.1.1. Cu(I)-Catalyzed Azide–Alkyne Cycloaddition
2.1.1.1. CuI-Catalyzed Synthesis
Orita and co-workers53 reported a process-controlled regiodivergent copper-catalyzed process for the synthesis of 4-bromo- and 5-bromotriazoles 6 and 7 through cycloaddition between bromo(phosphoryl)ethyne 5 and azide 2 using CuI as a catalyst and MeOH/H2O as a proton source for regeneration of the copper catalyst (Scheme 3). KOH-promoted one-shot dephosphorylation of 5 with azide resulted in 4-bromo-1,2,3-triazole 6. CuAAC reaction of 5 with azide 2 formed 5-bromo-1,2,3-triazole 7 instead of 6 because the 4-position was occupied by the sterically bulky Ph2P(O) group. Pd-catalyzed and MeOK-promoted nucleophilic substitution converted products 6 and 7 to 4- and 5-functionalized triazoles 8 and 9, respectively.
Scheme 3. Cu-Catalyzed Synthesis of 8 and 9 without and with the Phosphorylation Approach, Respectively.

Wu and co-workers54 reported a one-pot synthesis of 5-trifluoromethylthiotriazole 10 from 1 and 2 with sulfur powder (S8) and trifluoromethyltrimethylsilane (TMSCF3) using a CuI catalyst (Scheme 4). A terminal alkyne with the electron-donating group (EDG) on aryl alkyne favored this multicomponent reaction, while the electron-withdrawing group (EWG) led to poor yield. In addition, a moderate yield was obtained for alkyl alkyne. An electron-withdrawing group on azide also decreases the yield; heterocyclic azide is also a suitable substrate for this reaction.
Scheme 4. Synthesis of 10 from Alkyne, Azide, S8, and TMSCF3 Using CuI as the Catalyst.

Xu and co-workers55 reported a CuAAC/persulfuration reaction with a wide substrate scope, complete regioselectivity, and excellent functional group tolerability for the synthesis of asymmetric triazole disulfides 12 from 1, 2, and electrophilic persulfur reagent 11 (SS-t-butyl p-toluenesulfono) in the presence of the CuI catalyst and LiOtBu as a base (Scheme 5). They evaluated a wide range of 5-persulfur-functionalized triazoles in moderate to good yield and produced single regioisomers. The scope of the reaction was investigated with various azides and alkynes. Under the standard circumstances, all of the aliphatic azides and alkynes reacted easily, giving the appropriate disulfides in good to outstanding yields.
Scheme 5. Synthesis of Derivatives of 10 Using a Persulfuration Reaction via CuI and tBuOLi.
Under the standard conditions, functionalized alkynes obtained from various carbohydrates, vitamins, and proteins can be simply converted to the corresponding triazole disulfides in good to excellent yields. These reactions indicate that this modification has great potential for generating novel physiologically active compounds.
To access a wide range of 5-functionalized triazoles 17, Xu and co-workers56 reported a copper-catalyzed interrupted click reaction from 1, 2, and different heteroatom electrophiles 16 using much less catalytic quantities of CuI catalyst and LiOtBu (act as a base) under relatively mild conditions (Scheme 6). This approach has a very broad scope and no required ligands for alkynes, azide, and other heteroatom electrophiles. This intramolecular reaction can be used to construct a bicyclic triazole with various ring sizes.
Scheme 6. Use of Click Reaction for the Synthesis of Derivatives of 17 Using CuI as a Catalyst.

Zhang and co-workers57 developed a direct route to synthesize bifunctional trimethylsilyl-5-iodotriazoles (TMSIT) 20 from TMS-alkyne 18, 2, and iodide followed by copper/palladium-catalyzed nucleophile coupling reaction and desilylation, resulting in a novel approach for the synthesis of structurally diverse 1,5-disubstituted triazoles 21, 22, 23, and 24 via a one-pot reaction, respectively, in high to excellent yields (Scheme 7). This route gave a great alternative for the synthesis of structurally diverse 1,5-disubstituted triazoles due to its short synthetic route and stability of the key precursors 5-iodo-4-TMS-triazoles.
Scheme 7. Synthesis of Key Precursor 20 and Further Synthesis of Triazole Analogues.

Renslo and co-workers58 reported a one-pot two-step procedure of copper-catalyzed cycloaddition of terminal alkyne 1, diazide 29/30, and CuI in DMSO for the synthesis of disulfide-bearing triazole 25/26 (Scheme 8).
Scheme 8. Synthesis of 27/28 via Cu-Catalyzed Cycloaddition Reaction of Terminal Alkyne and Diazide.
They found that heating the triazoles to 60 °C during the exchange action maintained the homogeneity of the reaction mixtures. The exchange reaction gave a mixture of thermodynamic products which favored the asymmetrical disulfide products 27/28 and decreased the reversible reaction which formed the symmetrical starting material 25/26.
Imperiali and co-workers59 suggested a synthesis of red-shifted 5(4-substituted-1H-1,2,3-triazol-1-yl)quinoline-8-ol 33 from 5-azidoquinoline-8-ol 32 and 1 in the presence of a catalytic amount of CuI, ascorbic acid, DMF/4-methylpiperidine (8:2) at RT. The intended product was obtained with 100% yield within 12 h (Scheme 9). Because of the electron-rich azido group’s quenching action, 32 displayed no fluorescence. Due to the removal of azide quenching following synthesis of the triazole, 33 showed fluorescence in the presence of 10 mM MgCl2 (λex = 371 nm and λem = 522 nm). Because of these derivative fluorescent properties, they employed them as chelation-sensitive fluorophores to make peptide-based probes for the MAPK-activated protein kinase-2 (MK2) and sarcoma kinase (Src), which are both MAPK-activated protein kinases.
Scheme 9. Synthesis of Analogues of 33 Using CuI Catalyst and Ascorbic Acid at RT.

Ley and co-workers60 reported a flow-based process for the synthesis of 3 from 1 and 2 via the CuI catalyst and DCM as a base. The starting material is introduced into the flow stream and pumped through the column containing amberlyst A-21 (PS-NMe2) 34, quadrapure TU (QP-TU) 35, and phosphine resin (PS-PPh2) 36 placed in a series. The process required 3 h to give the product in 70–93% yield with >95% purity (Scheme 10).
Scheme 10. Triazole 3 Synthesized from 1 and 2 in the Presence of CuI Catalyst Using a Flow Reactor.
Suresh and co-workers61 reported CuI(I)-catalyzed and Ru(II)-catalyzed 1,3-dipolar cycloaddition of aryl azide 1-azido-4-(methanesulfonyl)benzene 37 and various phenyl acetylenes 38 for the synthesis of 1,4- and 1,5-diaryl-substituted triazoles 39 and 40, respectively (Scheme 11).
Scheme 11. Synthesis of 39 and 40 through CuI and Ru(II) Catalysts.

Wu and co-workers62 reported the synthesis of 5-thio- or 5-selenotriazole 42 from 1, 2, organohalides/arylboronic acids/tosylates 41, and elemental sulfur or selenium using CuI as a catalyst and K2CO3 under mild conditions with good subtract scope and excellent yields (Scheme 12).
Scheme 12. CuI promoted synthesis of 42 and analogous under mild reaction condition.

Noroozi Pesyan and co-workers63 developed a mild, clean, and effective process for the synthesis of 1,4-disubstituted triazole 47 from 2-amino-6-(azidomethyl)-8-oxo-4-aryl-4,8-dihydropyrano[3,2-b]pyran-3-carbonitrile 45 and phenylacetylene 46 in the presence of copper iodide (CuI) and green solvent EtOH/H2O (2:1) at 80 °C for 2 h (Scheme 13). 45 was synthesized from Knoevenagel adducts and 43 using NaN344, SOCl2, dry DMF, and EtOH. A wide range of azides gave a reaction with phenylacetylene and gave excellent yields of 47.
Scheme 13. Synthesis of 47 and Derivatives from 45 Using CuI as the Promoter.
Belokon and co-workers64 reported an efficient reaction for the asymmetric synthesis of amino-acid-containing 1,4-disubstituted triazoles 51 and 52 from azides 48 and Schiff bases of (S)- or (R)-BPB chiral Ni(II) complexes 49 and 50, respectively, using copper(I) iodide catalyst, Et3N, and DMSO at 70 °C with excellent enantioselectivities (>99% ee) (Scheme 14). Complex 50 was prepared using (S)-BPB Ni-Gly followed by propargyl bromide alkylation.65 For the substrate scope of azides, they found that most of all azides gave good to excellent yields (68–93%) except 1-(azidomethyl)-3-fluorobenzene, 4-(azidomethyl)benzonitrile, and 9-(azidomethyl)anthracene, which gave comparatively fewer yields of 53, 38, and 42%, respectively.
Scheme 14. CuI-Promoted Synthesis of 51 and 52.
Zheng and co-workers66 reported novel, one-pot, efficient multicomponent reactions (MCRs) of OBoc-alkynes 53, azides 2, amines 54, and 2H-azirines 55 for the synthesis of fully substituted 1,2,3-triazole 56 using copper iodide, DIPEA, and MeOH (Scheme 15). Here, 2H-azirines are readily disclosed for the C–N bond formation. Electron-donating and electron-withdrawing alkynes had a good transformation, but alkyl-substituted OBoc-alkyne does not transform into the desired product. For 55, alkyl 2H-azirines did not provide the desired product. This approach is very useful for polyfunctionalized triazole with a wide substrate scope, mild conditions, and good yield.
Scheme 15. Synthesis of Polyfunctionalized Triazole 56.

2.1.1.2. CuCl-Catalyzed Synthesis
Xu and co-workers67 reported a three-component interrupted click approach for a bench-stable 5-stannyl triazole 58 from easily available 1, 2, and Bu3SnOMe using CuCl as a catalyst (Scheme 16). Through a Sn/Cu transmetalation, 5-stanyl triazole 58 was produced which can be used as a potent nucleophilic reagent. This mutual Sn/Cu transmetalation is crucial for future progress in tin chemistry and useful for the synthesis of fully functionalized triazole 59, trifluoromethylthiolated triazole 10, and trifluoromethylated triazole 60.
Scheme 16. Synthesis of 59, 60, and 10 Using Divergent Routes.
Zhou and co-workers in 201368 reported the first highly enantioselective asymmetric desymmetrization for the synthesis of quaternary oxindoles bearing 1,2,3-triazole 64 from oxindole-based 1,6-heptadiynes 61 and azide 2 using CuCl (18 mol %), PyBOX ligand (15 mol %) 63, and 2,5-hexanedione at 0 °C for 96 h (Scheme 17a). The N-protecting group of oxindole favors most of the alkyl and acetyl groups but is less favorable for the electron-withdrawing acetyl protecting group, which leads to the formation of undesired ditriazole 65.
Scheme 17a. Asymmetric Desymmetrization of Oxindole Derivatives 61.
Here, 64 with the alkyne group uses as a versatile synthetic tool for further modification such as [3 + 2] cycloaddition (a), full or partial hydrogenation (b,d), and Sonogashira reaction (c) (Scheme 17b).
Scheme 17b. Synthetic Elaboration of 64.
Zhou and co-workers69 first reported a highly enantioselective desymmetric CuCl-catalyzed reaction of terminal alkynes 1 or 1-iodoalkynes 71 with diazides 70 using PyBOX ligands 72 and DCM at −50 to −40 °C for the synthesis of tertiary alcohols containing 1,2,3-triazoles 73 and 74 with excellent yield (Scheme 18). For the PyBOX ligand study, they found that different C4 shielding groups showed a different result. The OBn group increased the % ee ratio of 73 but led to poor yield. Increasing the steric hindrance of the shielding group using −1-naphthyl and −2-MeO-3,5-tBu2-C6H2 also gave poor results. When the phenyl group (−Ar) group was replaced by the 4-fluorophenyl group, the reaction led to a higher % ee ratio and good yield with DCM solvent. Irrespective of the electronic effect of substituents meta-substituted diazo alcohol gave a slightly higher % ee in comparison to that at the para-position. Based on the study of the reaction mechanism, this approach showed that % ee increased with time and decreased the chiral/achiral ratio, which indicates that enantioselectivity increased by the formation of an achiral ditriazole.
Scheme 18. Highly Enantioselective Synthesis of Alcohol-Containing Triazoles 73 and 74.
Zhou and co-workers in 202170 reported a highly enantioselective asymmetric CuAAC synthesis of a tertiary alcohol functional group bearing 1,2,3-triazole 78 from azides 2 and highly functionalized tertiary alcohol containing ethynyl 75 using CuCl, 3-pentyl-containing PyBOX-phosphonate ligand L79, AgTFA additive, and tBuOAc at °0 C for 3–4 days (Scheme 19a). The 3-pentyl group at the C4 position increase the s-factor. Kinetic resolution (KR) of racemic α-ethynyl alcohol 75 gave a good yield with 87–99% ee and 39–50% yield of 77. α-Alkyl alcohol such as α-isopropyl alcohol did not give good results (60% ee and 34% yield). The E-CuAAC reaction of this tertiary alcohol containing ethynyl 75 with azide 2 gave chiral α-1,2,3-triazole bearing tertiary alcohol 79 in 70–99% yield with 48–96% ee. This approach is useful for the synthesis of bioactive compounds such as indomethacin, fenbufen, and celecobix.
Scheme 19a. Synthesis of 77 and 78.
This approach is also suitable for the kinetic resolution of α-monofluorinated α-ethynyl alcohol 80 and benzyl azide 76 for the synthesis of monofluorinated α-triazole-substituted alcohol 82 in 40–50% yield with 20–96% ee and gave chiral alcohol 81 (Scheme 19b).
Scheme 19b. Kinetic Resolution of 80.
2.1.1.3. Cu(MeCN)4PF6-Catalyzed Synthesis
Song and co-workers71 reported an efficient synthesis of 4-heterofunctionalized triazolyl–organosulfur 84 with high regioselectivity and wide substrate scope (33 examples) such as 85–88 from internal thiocynatoalkynes 83 and 2; they used (CuOTf)2PhMe and Cu(MeCN)4PF6 (tetrakisacetonitrile)copper(I)hexafluorophosphate) as a catalyst (Scheme 20). The electronic effect was not influenced by electron-rich and electron-deficient substituents as they gave the same yield, but for the azide scope of the reaction, the yield for electron-rich azide was slightly lower than that of the electron-deficient azide. The reaction could also occur for a secondary azide.
Scheme 20. Synthesis of Analogues Promoted by Cu(I) Catalyst with High Regioselectivity.
Goldup and co-workers72 reported an active template Cu-catalyzed alkyne–azide cycloaddition (AT-CuAAC) reaction of propargyl cytosine 89 and azido thymine 90 for the preparation of rotaxane 92 in the presence of macrocycle 91 using [Cu(MeCN)4]PF6 as a catalyst, NiPr2Et, and THF at 50 °C for 18 h (Scheme 21). At 32 °C, no amplification was observed when oligonucleotide rotaxane was utilized as the forward primer in PCR amplification. According to the findings, the mechanical link in rotaxanes efficiently limits the interlocked oligonucleotide’s capacity to act as a primer for the T7 polymerase. This method utilized in click DNA ligation may be easily extended to the active template manifold for the synthesis of biocompatible triazole-linked oligonucleotides based on rotaxanes.
Scheme 21. [Cu(MeCN)4]PF6-Promoted Synthesis of Rotaxane 92.
Zheng and co-workers73 reported a 3-CAP (three-component asymmetric polymerization) synthesis of chiral polytriazole-methanamines 96 and 97 from OBoc-alkynes 93, azides 2, and various types of amines 94 and 95 using Cu(CH3CN)4PF6, ligand (L) 63, and DIPEA at room temperature (Scheme 22a). For 3-CAP, the secondary and primary amine gave good yield and gave Mn up to 46 700 g/mol.
Scheme 22a. Synthesis of 96 and 97 via 3-CAP.
This approach is also useful for the synthesis of structurally diverse polymers 101 and 102. With diazide 98, secondary amines 54 and 100 and OBoc-alkynes 93 and 99 were used as a substrate and yielded excellent Mn up to 10 200 g/mol (Scheme 22b).
Scheme 22b. Synthesis of Diverse Polymers.
Zheng and co-workers74 reported MCP (multicomponent polymerization) and interrupted click synthesis of 1,4,5-polytriazoles (1,4,5-PTAs) 106 from diynes 103, diazides 104, and electrophile 105 using Cu(MeCN)4BF4 catalyst, DIPEA (N,N-diisopropylethylamine) base, and DMF as a solvent at room temperature with high yields, Mn values, and excellent modification efficiency (Scheme 23). Poor solubility of inorganic CuI does not catalyze this polymerization.
Scheme 23. MCP Reaction of Diynes 103, Diazides 104, and Electrophiles 105.

2.1.1.4. CuOTf-Catalyzed Synthesis
Fokin and co-workers75 reported the first asymmetric CuAAC reaction via kinetic resolution of α-chiral azide and desymmetrization of gem-diazide. They also discussed the importance and the role of this new PyBOX ligand in the enantioselective synthesis of 1,2,3-triazole.
Uozumi and co-workers76 reported a highly enantioposition-selective CuAAC approach for the synthesis of axially chiral biaryl groups bearing 1,2,3-triazoles 108 and 109 from prochiral biaryl-bearing dialkynes 107 and benzyl azide 76 using CuOTf·(C6H6)0.5 (10 mol %), a tert-butyl(dimethyl)silyl (TBS) group containing PyBOX ligand L (20 mol %) 110, and 1,2-DCE at 60 °C for 24 h (Scheme 24). Increasing the load of azide up to 1.5 equiv enhanced the % ee to 91% with 64% yield. X-ray crystallographic analysis proved that the absolute configuration of triazole was R.
Scheme 24. Synthesis of Chiral Biaryl-Bearing Triazoles 108 and 109.
Fokin and co-workers77 reported the synthesis of 5-bismuth(III) triazolides 113 from a bench-stable, readily available, as well as a nontoxic group of σ-acetylides, 1-bismuth(III) acetylides 111 and 2, in the presence of CuOTf (Tf = trifluoromethanesulfonyl) and THF as a solvent for 3 h at RT (Scheme 25a).
Scheme 25a. CuOTf-Promoted Synthesis of 113.

Further reaction of 113 with various electrophiles enabled the synthesis of fully substituted 1,2,3-triazoles 59, and 114–117 (Scheme 25b).
Scheme 25b. Electrophilic Substitution Reactions of 113.

2.1.1.5. β-Cyclodextrin-TSC@Cu-Catalyzed Synthesis
Naimi-Jamal and co-workers78 reported a synthesis of 1,4-disubstituted triazoles 120 and 121 from 1, azides 44/119, and alkyl halides 118 using novel, stable, water-soluble, as well as homogeneous catalyst β-CD-TSC@Cu (copper(I) ions supported on functionalized β-cyclodextrin) as a supramolecular moiety in an aqueous medium with excellent regioselectivity and high yield (Scheme 26). The catalyst was utilized up to 7 times without a remarkable loss of its activity. After the third cycle, very low (≤2 ppm) copper leaching was observed by ICP-OES.
Scheme 26. Synthesis of 120 and 121 Catalyzed by a Nanocatalyst β-CD-TSC@Cu.

The substrate scope of this reaction reported that 2-bromoacetophenone derivatives minutely decreased the product yield and expanded the reaction time. The hydrophobic internal cavity of β-CD provided a favorable environment for copper ions and organic substrates to interact more effectively.
2.1.1.6. Phenylethynylcopper(I)-Catalyzed Synthesis
Varela-Palma and co-workers79 developed a synthesis of 1-(1-benzyl-1,2,3-triazol-4-yl)cyclohexanol 124 from 1-ethynylcyclohexanol 123 and benzyl azide 76 using phenylethynylcopper(I) 122 catalyst and CH2Cl2 as a solvent at RT for 24 h. Adding a Fehling A and B solution to the glucose–phenylacetylene 46 mixture resulted in yellow particles such as phenylethynylcopper(I). The 1,3-bis(1,2,3-triazol-1-yl)-propan-2-ol core containing a triazole (10 examples) was synthesized by this protocol with 80–92% yields (Scheme 27).
Scheme 27. Phenylethynylcopper(I)-Promoted Synthesis of 124.

2.1.1.7. Miscellaneous CuAAC-Catalyzed Reaction
Fossey and co-workers80 reported a short review on asymmetric CuAAC′ (chiral Click) reaction (Scheme 28). Taking advantage of the difference in the rate of reaction kinetic resolution through CuAAC gave the formation of new enantioenriched material or product. A racemic mixture (50:50) also converts into either R or S and becomes enantioenriched. Using this approach, kinetically resolved triazoles 131 and 132 are produced from an α-benzylic azide and phenylacetylene.75 Desymmetrization leads to the loss of one or more symmetry elements and results in a prochiral molecule from the chiral molecule. This approach is useful for producing a single enantiomer from a nonchiral reactant.81 Desymmetrization is very useful for the synthesis of some biologically active chiral compounds82 such as (−)-podophyllotoxin and (−)-picropodophyllin.83
Scheme 28. CuAAC Desymmetrization and CuAAC Kinetic Resolution.
Zhou and co-workers84 reported the synthesis of 1,2,3-triazoles 137 and 138 using desymmetrization and an enantioselective CuAAC approach. The newly synthesized chiral PyBOX ligand shows excellent enantioselectivity in 1,2,3-triazole. Enantioselective CuBr-catalyzed cycloaddition of azide 2 and alkynylphosphine oxide (P-chiral and P-substituted) 135 using 1-naphthyl-bearing PyBOX ligand L139 and MeCN at −20 °C for 4 days yielded P-chiral phosphine oxides bearing a 1,2,3-triazole scaffold 137 in excellent yield and high enantioselectivity (Scheme 29a). Desymmetrization of dialkynylphosphine 136 with azide 2 under the CuAAC approach yielded monotriazole 138 with excellent results. This ethynyl-containing triazole is useful for further functionalization.
Scheme 29a. Synthesis of 137 and 138 via Kinetic Resolution and Desymmetrization.
For the desymmetric CuAAC approach for phosphole diynes 140, the reaction of 140 with azide 2 yielded chiral monotriazole 141 with good enantioselectivity (Scheme 29b). This versatility of monoethynylphosphine oxide 141 was used as a chiral building block for various reactions.
Scheme 29b. Synthesis of Monoethynylphosphine Oxide 141.
2.1.2. Cu(II)-Catalyzed Azide–Alkyne Cycloaddition
2.1.2.1. CuCl2-Catalyzed Synthesis
Xu and co-workers85 reported a copper-catalyzed synthesis of indole–triazole 146 from 1-(prop-2-ynyl)-1H-indole 146 and 2 using CuCl2/Zn powder in water at RT. At RT, the H atom linked to the N atom of indole 143 was replaced by pro-2-ynyl 144 using Bu4N+·Br–, NaOH, and toluene, resulting in 145 (Scheme 30).
Scheme 30. Synthesis of Derivatives of 146 Using CuCl2/Zn Powder in Water at RT.
Most compounds show moderate to good antifungal activity against Colletotrichum capsica and cotton Physalospora pathogens, using flutriafol and hexaconazole as positive controls. The activity of compounds XII and XIII was unaffected by the distance between the phenyl ring and the triazole moiety.
The findings revealed that the ortho-substituents of the phenyl ring are helpful toward the activity; however, when R is an electron-withdrawing group or hydrogen at the meta-position is replaced, the impact is unfavorable. Compound XIV has no substituted phenyl ring and was shown to be the most efficient against Colletotrichum capsica, with an inhibition ratio of up to 83.33% (Figure 5).
Figure 5.
Typical triazole derivatives that show antifungal activity.
2.1.2.2. CuSO4-Catalyzed Synthesis
Cunha Lima and co-workers86 described a synthesis of 1,4-disubstituted triazoles 151 from ethyl 2-azidoacetate 149 and terminal alkyne 150 in the presence of CuSO4·5H2O, sodium ascorbate (NaASc), and EtOH/H2O (1:1) at RT (Scheme 31). 149 was synthesized by azidation of ethyl 2-bromoacetates 147 and 150 synthesized from a natural product by a propargylation reaction. All synthesized compounds showed moderate antioxidant activity with an EC50 value above 75.5 μg/mL in the DPPH assay and 101.1 μg/mL in the ABTS assay. In the CuAAC reaction, the formation of copper(I)acetylide is an important step from the deprotonation of a terminal alkyne. Electron-withdrawing groups raised the reactivity and acidity of the terminal alkyne hydrogen, so reaction times were decreased and yields were increased.
Scheme 31. CuSO4·5H2O-Promoted Synthesis of Derivative of 151.
Huang and co-workers87 developed a step-growth polymerization of diazide 153 and phthalimide dialkyne 152 using the copper catalyst for the synthesis of polytriazoleimide 154 under mild conditions (Scheme 32). IR, NMR, wide-angle XRD, differential scanning calorimetry, thermogravimetric analysis, and intrinsic viscosity are used for characterization. The polytriazoleimides have high solubility in polar solvents and could be readily cast into clear, robust, and flexible films with intrinsic viscosities of 0.39–0.58 dL g–1. The thermal stability and mechanical qualities of these new polytriazoleimide films were impressive.
Scheme 32. Synthesis of 154 Using Cu(I) Catalyst.
Cuevas-Yañez and co-workers88 reported a copper-catalyzed synthesis of 2-aryl-1-(1,2,4-triazolyl)-3-(1,2,3-triazolyl)propan-2-ol derivatives 157 and 158 from 2-(2,4-difluorophenyl)-1-[1,2,4]triazol-1-yl-pent-4-yn-2-ol 155 with various azides 2 and 1-azido-2-(2,4-difluorophenyl)-3-[1,2,4]triazol-1-yl-propan-2-ol 156 with different alkynes 1 using CuSO4 and sodium ascorbate, respectively (Scheme 33). These fluconazole analogues showed good antifungal activity.
Scheme 33. Synthesis of 157 and 158 Using CuSO4·5H2O.

Joshi and co-workers89 reported a facile, greener, and ultrasound-promoted synthesis of triazole 163 in the presence of CuSO4, sodium ascorbate, and DMF/t-BuOH/H2O (1:1:1) by reaction between 2-(azidomethyl)-1H-benzo[d]imidazole 162 and 1 with good yields. The enhanced product formation was observed due to the greater solubility of copper sulfate in this aqueous medium (Scheme 34). O-Phenylenediamine 160 reacted with chloroacetic acid in HCl and resulted in 2-(chloromethyl)-1H-benzo[d]imidazole 161, which was treated with sodium azide 44 in DMF at RT to yield 162. An in vitro antimicrobial study reported that all of the synthesized molecules showed moderate activity compared to that of standard drugs. According to the findings, all synthesized molecules were more active against Gram –ve bacteria in comparison to Gram +ve bacteria.
Scheme 34. CuSO4-Assisted Synthesis of 163 and Its Derivatives.

Crowley and co-workers90 performed a safe, one-pot CuAAC approach of an exofunctionalized pyridyl-triazole 166 from dialkyne 164, dibromide 165, and sodium azide 44 using CuSO4·5H2O and DMF/H2O (4:1) for 48 h, resultin in a 44% isolated yield (Scheme 35).
Scheme 35. Synthesis of 166 Using CuSO4·5H2O Catalyst in DMF/Water Solvent.
Weinreb and co-workers91 reported a synthesis of N-protected 4-substituted triazole 169 from β-tosylethylazide 168 and 1 using sodium ascorbate/CuSO4 as a catalyst in aqueous tert-butanol at RT. With a further reaction of 169 with KOt-Bu in THF at −78 °C, the mixture was gently warmed to 0 °C to produce unprotected 4-substituted 1,2,3-triazole 170. 168 was synthesized from p-tolyl vinyl sulfone 167, sodium azide 44, and sulfuric acid in MeOH at 0 °C to RT with good yield (Scheme 36). This cycloaddition was also favored for acetylene dicarboxylic acid and dimethyl acetylenecarboxylate with azide. They also reported an efficient ruthenium-catalyzed cycloaddition of alkyl azide with terminal and asymmetric internal alkynes using Cp*Ru(PPh3)2 and PhH to produce 1,2,3-triazole.
Scheme 36. Synthesis of 169 and Its Derivative 170.
Crowley and co-workers92 reported a synthesis of various bi- and tridentate pyridyl-triazole ligands with Cu(II) and Ag(I) 173 and 174 from halides (2-(bromomethyl)pyridine 171 and 2,6-(bis-9-bromomethyl)pyridine 172) with sodium azide 44 and 1 using CuSO4·5H2O, ascorbic acid, DMF/H2O at RT for 20 h, respectively (Scheme 37).
Scheme 37. CuSO4·5H2O-Catalyzed Synthesis of 173 and 174 and Their Derivatives.
Thibonnet and co-workers93 performed the one-pot method of 1,2,3-triazole-containing morpholine scaffold 176 in the presence of 18-crown-6, NaASc, CuSO4·5H2O, and CHCl3/H2O from 1, iodomorpholinone 175, and sodium azide 44 at RT for 24 h using the CuAAC approach (Scheme 38). This synthetic approach was carried out under an argon atmosphere.
Scheme 38. Synthesizing 176 via Utilization of 18-Crown-6, Sodium Ascorbate, and CuSO4 in Chloroform/Water.

Tan and co-workers94 reported a synthesis of novel hybrid phthalimide analogues containing a triazole 181 and 182 via CuAAC reaction of 2-ethynylisoindoline-1,3-dione 178 with various organic azides (2-azido-N,N-disubstituted acetamide 179 and 2-azido-N-substituted acetamide 180) in the presence of CuSO4·5H2O, sodium l-ascorbate, and MeOH/H2O (1:1) at RT (Scheme 39). Compound 178 was synthesized from phthalimide 177 and propargyl bromide using K2CO3 in acetonitrile at 75 °C.
Scheme 39. CuSO4·5H2O and Sodium-Ascorbate-Promoted Synthesis of 181 and 182 and Their Derivatives in Aqueous Methanol.
The molecular docking of compounds XV, XVI, and XVII (Figure 6) with molecular dynamic simulation studies showed interaction of PPI with ACE2-S1, the best binding energy of XV with a value of −9.70 kcal/mol, and the best approximated Ki value of XV (0.077 μM). In the interaction with the Mpro protein, compound XVI exhibited the best binding energy (−8.76 kcal/mol) and approximate Ki value (0.315 μM), while in the interaction with the PLpro, compound XVII showed the best binding energy (−8.87 kcal/mol) and estimated Ki value (0.315 μM). Based on an in silico study, these derivatives may block the entry of SARS-CoV-2 into the host cell.
Figure 6.
Compounds XV, XVI, and XVII.
Sierra and co-workers95 developed a synthesis for 1,2,3-triazole-BODIPY scaffold 186 from anilines 184 and alkynyl sulfoxides 183 using NfN3185, NaHCO3, and CuSO4·5H2O/sodium ascorbate in MeOH/H2O/Et2O with 41–80% yield (Scheme 40). Further substitution on this moiety led to chiral-at-metal BODIPY-based iridium(III) complexes 194–196, which showed excellent luminescence properties.
Scheme 40. Synthesis of Ir-Containing Complexes 194–196 Containing Luminescence Properties.
Thordarson and co-workers96 reported a step process for the synthesis of iridium(III)azides 199 and further reaction of 199 with 1-(prop-2-ynyl)-1H-pyrrole-2,5-dione 198 using DMF, CuSO4, and NaASc for the synthesis of Ir(III)-containing triazole-bisterpyridine 199 for 3 days (Scheme 41). This complex showed photophysical activity which is important for the modification of the polymer and the surface.
Scheme 41. Synthesis of 199 in the Presence of Aqueous CuSO4 with NaASc in DMF.

2.1.3. Heterogeneous Cu-Catalyzed Azide–Alkyne Cycloaddition
2.1.3.1. Cu2O/HTNT-7-Nanocatalyzed Synthesis
Peddiahgari and co-workers97 developed a novel approach for the CuAAC reaction from organic halides 118, sodium azide 44, and 1 using Cu2O/HTNT-7 nanoparticles (Cu2O nanoparticles supported on hydrogen trititanate nanotubes) as a catalyst and 4-MeO-C6H5I/4-MeO-C6H5NH2 as an azide precursor in water mediated at RT with excellent yield for the synthesis of 1,4-disubstituted triazole 120 (Scheme 42). Using different halides and alkynes, they investigated the extent of the CuAAC reaction, in which they found 3-(bromomethyl)thiophene, 2-(bromomethyl)thiophene, and 2-(bromomethyl)pyridine immediately produced a complex mixture of byproducts, turning the reaction mass dark and viscous. The reusability of the catalyst was tested under optimal conditions and recovered using centrifugation. The retrieved catalyst was reused seven times with no discernible difference in yield. ICP-MS investigation revealed that very low (≤1 ppm) copper leaching was observed.
Scheme 42. Utilizing Cu2O/HTNT-7 as a Catalyst to Synthesize 120 at RT.

2.1.3.2. Cu@N–C-Catalyzed Synthesis
Xie and co-workers98 reported a one-pot synthesis of 3 from 1, aryl halides 41, and sodium azide 44 using a novel, efficient, recyclable heterogeneous catalyst Cu@N–C (copper supported on nitrogen-doped carbon) and t-BuOH/H2O (3:1) at 50 °C with high yield and broad substrate scope (Scheme 43). Under an argon flow, the powder Cu(im)2 (copper(II) bisimidazolate) was deposited in a tube furnace and calcined to 600 °C at a heating rate of 5 °C·min–1 for 5 h. The resultant solid was cooled to RT to afford the Cu@N–C(600). The retrieved catalyst was used up to 4 times without a discernible change under the standard conditions. Benzyl chloride required a longer time for completion because of its low activity for nucleophilic substitution reaction with sodium azide. Aliphatic alkynes required high temperatures and long reaction times due to their lower reactivity compared to aryl alkynes. The para-substituted compounds containing electron-deficient groups resulted in lower yield compared to that with ortho- and meta-substituted compounds. This approach was also used for the synthesis of ethisterone and zidovudine.
Scheme 43. Cu@N–C(600)-Promoted Synthesis of 3 in Aqueous tBuOH.

2.1.3.3. CuI Nanoparticle Synthesis
Maurya and co-workers99 developed an effective copper-catalyzed protocol using heterogeneous CuI nanoparticles to synthesize 1,4,5-trisubstituted triazole 202 derivatives by reacting alkynes 200 and benzyl azides 201 in a water medium at 70 °C for 3 h (Scheme 44). CuI NPs displayed a significant activity and recyclability (up to 6 cycles) for this reaction, and the intended product was formed in 81–96% yields.
Scheme 44. CuI NP-Promoted Synthesis of 202 and Its Derivatives.

2.2. Nickel-Catalyzed Azide–Alkyne Cycloaddition (NiAAC)
A possible catalytic path for NiAAC reaction is shown in Figure 7. First, intermediate A is produced from Ni(0) and internal alkyne. In the following cycloaddition process, intermediate A acts as a Nu, which allows electron-rich C to attack the terminal N of azide, resulting in nickelacycle intermediate B, which further leads to intermediate C. After removal of a metal atom, the desired triazole product is obtained and regeneration of the catalyst occurs.
Figure 7.

Conceivable mechanism for NiAAC. Reprinted with permission from ref (32). Copyright 2020 John Wiley and Sons.
2.2.1. Ni(COD)2-Catalyzed Synthesis
Topczewski and Liu100 reported the synthesis of N-chiral triazole derivative 205 that proceeds via the NiAAC reaction by dynamic kinetic resolution (DKR) between allylic azide 203 and alkynes 204 using Ni(COD)2 as a catalyst and (R)-1-(2-(diphenylphosphanyl)phenyl)ethan-1-amine 206 as a ligand in PhMe at 50 °C for 24 h (Scheme 45). This DKR is enabled by a spontaneous [3,3]-sigmatropic rearrangement of the allylic azide. For the alkyne scope of E-NiAAC (enantioselective NiAAC), the reaction with a neutral and electron-deficient substituent on the aryl group had less effect on enantioselectivity, while a strong EWG decreased regioselectivity. The azide scope of the E-NiAAC reaction reported that the scope of allylic azide was not much affected by electron-rich and electron-deficient groups, but when the cyclohexyl part was changed by acyclic allylic azide, both (E)- and (Z)-alkene isomers were observed (7:1 E/Z). However, the desired derivative was separated as a single isomer.
Scheme 45. Preparation of 205 and Its Derivatives Using Ni(COD)2 as a Promoter.
2.2.2. Cp2Ni/Xantphos-Catalyzed Synthesis
Hong and co-workers101 reported an efficient and simple nickel-catalyzed cycloaddition reaction of 1 with 2 in water or organic solvent (toluene) using Cp2Ni/xantphos as a catalyst. Product 4 was obtained in 42–95% yield with broad substrate scope and high regioselectivity (Scheme 46). They also reported that ortho-OMe-substituted alkyne gave no reaction due to a steric effect, whereas meta- and para-OMe-substituted alkynes gave 93 and 91% yield, respectively. So, a lower steric hindrance of substituted alkynes strongly favored the NiAAC reaction. This approach was further explored for biomolecules such as carbohydrates and amino acids. The xantphos ligand and Cp2Ni precatalyst were essential in achieving the catalytic manifold, which was insensitive to molecular oxygen and water.
Scheme 46. Synthesis of Derivatives of 4 at RT Using the NiCp2/Xantphos Catalyst.

Bosc and co-workers102 reported an effective synthesis of various 1,5-disubstituted-triazole 207 via nickel catalyzed (nickelocene) cycloaddition reaction of 1, bromide 118 and sodium azide 44 using xantphos, Cs2CO3 and DMF. They heated the reaction mixture for 4 h at 50 °C in a microwave which showed excellent yield and regioselectivity (Scheme 47). This approach is very useful when the low-molecular-weight azide is added as a reactant for the synthesis of triazoles. Due to steric hindrance, the ortho-substituted phenyl ring containing bromide provided less yield compared to that with the meta-/para-substituted bromide.
Scheme 47. Regioselective Synthesis of 207 and 120 by Utilizing Nickelocene-Catalyzed Cycloaddition via Microwave Irradiation.
Hong and co-workers103 reported a synthesis of various fully substituted triazoles 42, 211, and 212 from 2 and various alkynes 209, 208, and 210, respectively. They used Cp2Ni as a catalyst, Xanthos 213/dppbz 214 as a ligand, and Cs2CO3 and DCM or toluene as a solvent at RT (Scheme 48).
Scheme 48. Synthesis of 42, 211, and 212 Using a Ni Catalyst.

2.2.3. Raney-Nickel-Catalyzed Synthesis
Surya Prakash Rao and co-workers104 developed a regioselective method of 1,4-disubstituted triazole 216 at 45 °C using a Raney nickel catalyst and toluene solvent from 2 and propargylic ether 215 with excellent yield (Scheme 49). An additional reducing agent was not needed in this reaction with Raney nickel. Here, alkyl and aryl acetylene gave a good yield. However, the NiAAC reaction of propargyl alcohol and phenyl azide yielded the 1,4- and 1,5-regioisomers in a 3:2 ratio with a 93% yield. Without any catalyst, this reaction also produced the 1,4- and 1,5-regioisomeric adducts in a 3:2 ratio; however, it needed 2 min of microwave heating to 140 °C in PEG-200 (polyethylene glycol-200). Per the mechanistic study, Raney Ni serves as a coordinating species in azide and alkyne cycloaddition reactions, and the method does not proceed through Ni acetylides.
Scheme 49. Raney-Nickel-Promoted Synthesis of 216 and Its Derivatives.

2.2.4. Ni-rGO-Zeolite Nanocatalyzed Synthesis
Basu and co-workers105 performed a one-pot method of 1,4-disubstituted triazole 120 with good regioselectivity in water from 1, halides 118, and NaN344 in the presence of the stable catalyst ternary nanocomposite substance, Ni-rGO-zeolite, for 4–6 h at 90 °C (Scheme 50). For the synthesis of the nanocatalyst, they prepared hybrid GO-zeolite from 2D GO (graphene oxide) and NaY zeolite. Synthesis of the GO-zeolite complex involved the addition of NaY zeolite to an aqueous suspension of GO and increasing its pH near 7 due to protonation. The negative charge on zeolite balanced due to protonation, resulting in Al–OH–Si bridging. The reaction mass was heated at 60 °C for 16 h under gentle stirring with subsequent evaporation of H2O and drying under vacuum. This GO-zeolite was then treated with Ni(OAc)2·4H2O in the presence of NaBH4 under hydrothermal conditions, and the ternary nanocomposite material, Ni-rGO-zeolite, resulted. The catalyst performed up to four cycles without loss of activity.
Scheme 50. Ni-rGO-Zeolite NP-Promoted Synthesis of 120 and Its Derivatives.

2.3. Ruthenium-Catalyzed Azide–Alkyne Cycloaddition (RuAAC)
Based on density functional theory (DFT), Boren, Lin, and Fokin described the RuAAC cycle (Figure 8) based on the computational study.106
Figure 8.

Proposed catalytic cycle of the RuAAC reaction.106−108 Reprinted from ref (106). Copyright 2008 American Chemical Society.
2.3.1. Cp*Ru(COD)Cl-Catalyzed Synthesis
Goswami and co-workers109 reported a one-pot [3 + 2]/[2 + 2 + 2], facile, and atom-economical approach for the synthesis of 2-triazolyl thio-/selenopyridines 218 and 219 from 1-alkynyl thio-/selenocyanates 217 and alkyl/aryl azides 2 using Cp*Ru(COD)Cl catalyst in THF/toluene as a solvent at 70 °C/RT conditions. Further reaction of 218 with diyne 220 in the presence of Cp*Ru(COD)Cl in EtOH yielded 222 with 26–91% yield, whereas the reaction of 217 and 2 with tetraynes 221 using Cp*Ru(COD)Cl in ethanol gave 223 with moderate to good yield (Scheme 51).
Scheme 51. Synthesis of 218 and 219 and Their Derivatives Using Ru(Cp*)(COD)Cl Catalyst.
Song and co-workers110 reported a regioselective synthesis of fully substituted 5-thiocyanatotriazoles 225 from thiocyanato alkynes 224 and 2 using Cp*Ru(COD)Cl catalyst in THF at RT for 12–24 h (Scheme 52). Further functionalization of 225 using TFAA (trifluoroacetic acid) and H2O2 at 80 °C for 24 h yielded 226. The AAC/nucleophilic substitution cascade three-component reaction of 224 and 2 and alkyl/aryl magnesium bromide using Ru(Cp*)(COD)Cl gave 42 with moderate to good yield. This route is also fit for the synthesis of non-natural carbohydrates.
Scheme 52. Utilizing (Cp*)Ru(COD)Cl as a Promoter for the Synthesis of 225 and Its Derivatives 42 and 226.

Mascareñas and co-workers45 reported an orthogonal reaction for the synthesis of thio-functionalized triazole 42 using thioalkynes 209 and 2 using (Cp*)Ru(COD)Cl as a catalyst in water at RT (Scheme 53). Aliphatic azides are also suitable for this approach and gave a good yield. Also, water is replaced by some biomolecular additives, which also gave good results with this methodology.
Scheme 53. Synthesis of 42 and Derivatives Using Ru(Cp*)(COD)Cl Catalyst in Water at RT.

Guo and co-workers44 reported a mild and convenient method for the synthesis of CF3CH2- and CF3S-containing 1,4,5-trisubstituted triazoles 230–232 from (trifluoromethyl)thiolated alkynes 227 and 228 and trifluoromethylated alkynes 229 with various alkyl 2 and aryl azides 119 and 76 using (Cp*)Ru(COD)Cl catalyst at RT in benzene or toluene (Scheme 54). Also, this methodology is suitable for aryl halides, hydroxyl groups, ester, and N-, S-, and O-containing heterocyclic molecules. Regioselectivity of the reaction affords CF3CH2 or CF3S at the 5-position of triazole.
Scheme 54. Synthesis of 230–232 and Their Respective Derivatives by Utilizing a Ru Catalyst.

Blixt and co-workers111 developed solid-phase peptide synthesis (SPPS) of triazole peptide 234 from peptide-terminated (S)-(−)-4-tert-butyl 2-azidopeptide 233 and internal alkynes 200 using (Cp*)Ru(COD)Cl and HFIP (1,1,1,3,3,3-hexafluoroisopropanol)-DCM (1:4) at RT for 15 min. For the alkyne scope, dimethyl acetylene dicarboxylate gave <2% yield (Scheme 55).
Scheme 55. Synthesis of Derivatives of 234 Using the RuCl(COD)(Cp*) Catalyst at RT.
2.3.2. [Cp*RuCl]4-Catalyzed Synthesis
Liskamp and co-workers112 reported a first Ru-catalyzed macrocyclization for the production of a bicyclic 1,5-triazole-bridged vancomycin CDE ring 242 from amino-acid-containing azides 235 and pentapeptide-containing alkynes 237 using [RuCl(Cp*)]4 catalysts (Scheme 56). In this method, the two ether bridges of bicyclic vancomycin CDE ring were replaced by 1,5-disubstituted triazole.
Scheme 56. [RuCl(Cp*)]4-Promoted Synthesis of 242.
2.3.3. [RuClCp*(PPh3)2]-Catalyzed Synthesis
Johansson and co-workers113 developed a one-pot two-step methodology for the synthesis of 1,5-disubstituted-1H-1,2,3-triazole 207 from 1, alkyl halide 118, and NaN344 using [RhClCp*(PPh3)2] under MW heating at 100 °C for 30 min (Scheme 57). The secondary alkyl halide did not give any reaction even at a high temperature. An acidic-group-containing alkyl halide is not a suitable substrate for this methodology.
Scheme 57. [RuCl(PPh3)2(Cp*) Impulsive Synthesis of 207.

2.4. Iridium-Catalyzed Azide–Alkyne Cycloaddition (IrAAC)
A proposed mechanism for IrAAC reaction of alkyne 196 and azide 2 shown in Figure 9. First, Ir(I) complex A triggered the alkyne, which led to the generation of the alkynyl complex B as the CuAAC mechanism.115 Then the terminal nitrogen of azide had an electrophilic center after coordination with metal, which led to the formation of stabilized Ir carbenoids C and C′. Subsequent Ir [3 + 2] cycloaddition underwent “Cope-like” cyclization, and D and E yielded metallacycle F,114 which finally led to the production of the titled product 197.
Figure 9.
Postulated mechanism of IrAAC. Reprinted from ref (114). Copyright 2013 American Chemical Society.
2.4.1. [Ir(COD)Cl]2-Catalyzed Synthesis
Ding and co-workers116 reported a novel, simple, efficient, and sequence-defined polytriazole 246. First, 1 converted into internal thoialkynes 244 by reacting with 243. 244 reacted with 2 in the presence of an Ir catalyst to give 5-thio-functionalized 1,2,3-triazole 245. 245 further underwent an iterative sequential growth (ISG) method for the formation of polytriazole 246 (Scheme 58). Various functional groups were incorporated at the C4 position of the triazole scaffold. The C–S bond is easily detected by tandem mass spectrometry analysis due to the cleavage of a relatively weak C–S bond.117−119
Scheme 58. [Ir(COD)Cl]2-Catalyzed Synthesis of 246.

Gao and co-workers120 reported a synthesis of fully substituted 5-thiotriazoles 248 from alkynyl thioether 247 and benzyl azide 76 using [Ir(COD)Cl]2 catalyst and CH2Cl2 overnight in a N2 atmosphere at RT to give 81% yield (Scheme 59). 247 can be easily synthesized from bench-stable, rapidly prepared, and easily activated N-alkynylthiophthalimide and 4-methoxyphenylmagnesium bromide via thioester, which under subsequent acid hydrolysis yielded the desired compound 247.
Scheme 59. Iridium-Catalyzed Synthesis of 248 in DCM at RT under a N2 Atmosphere.

Sun and co-workers121 reported a first electron-rich internal alkyne for the synthesis of 5-thiotriazole 42 and fully substituted triazoles 59 using 2 and internal thoialkynes 209 as well as internal alkynes 200 using [Ir(COD)Cl]2 catalysts and dichloromethane at RT, respectively (Scheme 60). Diverse aryl and alkyl azides are suitable for the approach as well as for enhancing steric hindrance on internal thoialkynes 209, which also did not affect the reaction efficiency. Due to mild reaction conditions, alcohol, ester, aryl halides, ethers, THP-, and silyl-protected alcohols and phthalimide- and Boc-protected amines were also tolerated in this approach. For internal alkynes, they reported that electron-donating and neutral alkynes gave poor yield with greater regioselectivity. Also, prop-1-yn-1-ylbenzene, 1,2-diphenylethyne, and 1,2-diphenylethyne gave a trace amount of yield (<5%), whereas hex-3-yne showed very low conversion.
Scheme 60. Synthesis of 42 and 59 and Their Derivatives through an Ir Catalyst.

Song and co-workers122 developed a highly regioselective, mild, bioorthogonal strategy for the synthesis of fully substituted 5-amidotriazole 250 from ynamides 249 and 2 using [Ir(COD)Cl]2 catalysts with DCM or aqueous conditions at RT (Scheme 61). Cyclic and acyclic ynamide work well with the approach; the electron-rich cyclic ynamide gave a slightly higher yield, while the electron-deficient group gave a slightly low yield. If the adjacent position of ynamide (−R2) was substituted by a bulky group, the reaction yield decreases.
Scheme 61. Preparation of 250 and Its Derivatives Using [Ir(COD)Cl]2 as a Catalyst.

Cui and co-workers123 reported a mild, efficient, and hydroxyl group synthesis of fully substituted triazole 252 from alkynes 251 and 2 using [Ir(COD)Cl]2 catalysts in DCM at RT (Scheme 62). With the achievement of the cycloaddition of 252 with 2, they predicted that a hydroxyl group of internal alkynes would work as a directing group.124−128 For the substrate scope of alkynes, they found that alkyl alkynes gave a yield comparatively less than that of aryl alkynes. Aliphatic alcohol-containing alkynes did not give any reaction,128 whereas the meta-substituted phenolic group gave a very low yield.
Scheme 62. Synthesis of 252 and Its Derivatives Using Ir-Based Azide–Alkyne Cycloaddition.

Song and co-workers129 reported excellent regio- and chemoselective synthesis of 5-ether triazoles 254 and 256 from internal alkynes 253 and 255 with azides 2 and 76 using [Ir(COD)Cl2]2 catalysts in chloroform at 60 °C for 12–24 h (Scheme 63). Upon further investigation, they found that the yield dramatically decreases with an increase in the steric hindrance.
Scheme 63. Synthesis of 254 and 256 by Utilizing [Ir(COD)Cl]2 as a Catalyst.

Xu and co-workers130 reported a first atroposelective approach for the synthesis of axially chiral aryl triazole 258 from internal alkynes 257 and azides 2 using Ir(I)/squaramide (O) 259 and DCE/EtOH (5:1) at 25 °C (Scheme 64). The optimum asymmetric induction was achieved with quinidine squaramide (O) and Ir(I). Other metals such as Cu(I) and Pd(II) are not suitable metal catalysts for this reaction, and organocatalysts (O) by themselves did not generate any product. The result suggested that both organocatalysts and metal catalysts are a necessity for the reaction. When the epimer of O (quinine-derived squaramide O′) was applied with the Ir(I)/squaramide (O) catalyst, the configuration-reversed product (aR) was formed in excellent yield with 3:97 er in standard conditions. Sterically hindered mesityl azide is also a suitable substrate for this approach that results in great enantioselectivity. The reaction was completely stopped when the −OH group of alkynes was protected, suggesting that the VQM (vinylidene ortho-quinone methide) intermediate was probably formed by the organocatalyst. Here, the VQM bifunctional ligands provide hydrogen bonding interactions with ketone. Therefore, the C–N bond determines the stereoselectivity of the approach.131
Scheme 64. Synthesis of 258 Using an Ir Catalyst.
2.4.2. [Ir(COD)OMe]2-Catalyzed Synthesis
Taran and co-workers114 reported a synthesis of 4-bromo-1,5-substituted triazoles 261 using bromoalkynes 260 and 2 using [Ir(COD)OMe]2 catalysts in dichloromethane at −25 °C for 15 h (Scheme 65). For the substrate scope of the reaction, they found that electron-donating aryl alkynes gave a good yield compared to that with an electron-withdrawing group, and this approach is also suitable for sulfur-containing scaffolds but is affected by the steric effect.
Scheme 65. Synthesis of Analogues of 261 Using [Ir(COD)OMe]2 Catalysts in DCM.

2.5. Rhodium-Catalyzed Azide–Alkyne Cycloaddition (RhAAC)
2.5.1. [Rh(CO)2Cl]2-Catalyzed Synthesis
Zheng and co-workers132 reported a one-pot and regiodivergent synthesis of fully substituted 4-sulfonyl-1,2,3-triazole 263 and 5-sulfonyl-1,2,3-triazole 116 using the RhAAC reaction. Here, the regioselectivity of the reaction is controlled by nonmetallic sulfur(II) and sulfur(VI), giving 4- and 5-sulfonyl-functionalized products, respectively (Scheme 66).106 For the synthesis of 263, internal thoialkynes 209 oxidized using m-CPBA gave internal sulfonyl alkyne 262, which further reacts with azide using [Rh(CO)2Cl]2 and DCE at 40 °C for 12–24 h. The electronic effect of aryl substituents does not affect the regioselectivity and yield, but p-nitrophenyl-substituted derivatives gave a slightly lower yield (73%). For the synthesis of 116, the first internal thoialkyne 116 undergoes the RhAAC reaction to yield 5-sulfur-1,2,3-triazole 42, which further oxidized into 5-sulfonyl-1,2,3-triazole 116. Here, 263 followed the nonchelation path, while 116 followed the chelation path (Figure 10).
Scheme 66. Regiodivergent Synthesis of 116 and 263.
Figure 10.
Mechanistic cycle of the Rh catalyst.32,132 Reprinted with permission from ref (32). Copyright 2020 John Wiley and Sons.
Song and co-workers133 reported a regioselective approach for the synthesis of fully substituted 5-thiotriazoles 266 and fully substituted 5-trifluoromethylthiotriazoles 10 from internal thoialkynes 265 and internal alkynyl trifluoromethyl sulfides 264 with 2 at 40 °C and RT, respectively (Scheme 67). They used [Rh(CO)2Cl]2 as a catalyst and CHCl3 as a solvent. They investigated the substrate scope of the reaction between 264 and 2 in which they found that various fully substituted 5-trifluoromethylthiotriazoles 10 were synthesized at RT without an inert gas medium in good yield and excellent regioselectivity. Yields were similar, and no significant electronic effects were observed. However, the yield of the corresponding compound was lower when p-nitrophenylethynyl(trifluoromethyl)sulfane was used as the reactant, but the regioselectivity was excellent. The RhAAC reaction could also be carried out with good yields using ortho- and meta-substituted phenylacetylenes. Due to a combination of unfavorable electronic as well as steric effects with a prolonged reaction time and higher temperature, the reaction for alkyl-substituted 264 did not occur. The expected compound could be produced with good yield and high regioselectivities when alkyl or aryl azides were used as substrates. This approach tolerated a broad range of functional groups, including halogen, ester, and carbonyl groups. The yield was drastically reduced when aryl azides were employed instead of alkyl azides.
Scheme 67. [Rh(CO)2Cl]2-Promoted Synthesis of 10 and 266 and Their Derivatives.

Li and co-workers134 reported the first rhodium(I)-catalyzed open flask approach for the synthesis of 5-aminotriazoles 268 from internal ynamides 267 and 2 using [Rh(CO)2Cl]2 catalysts and MeCN as a solvent at RT for 0.5–22 h (Scheme 68). They also reported that exclusion of air and moisture was not necessary because the reaction maintains similar efficiency with an open flask. This approach was also carried out with a variety of solvents without any discernible change in the reaction yield. The reaction outcomes were unaffected by ynamides with different protective groups. Interestingly, the reaction of ynamides with the nosyl group proceeded successfully, producing a triazole with an 84% isolated yield. The prolonged reaction time is most likely due to the nosyl group’s strong electron-withdrawing group properties. Regardless of the electronic nature, the reaction of para-methyl, methoxy, trifluoromethyl, and chloro-substituted ynamides produced the appropriate triazoles in excellent yields when compared to those with the standard ynamide. The meta- and ortho-substitution on the phenyl ring was also tolerated. The ynamides with a heteroaromatic ring or an extra alkenyl moiety were found to be good substrates for the RhAAC reaction.
Scheme 68. Open Flask Approach to Synthesize 268 and Its Derivatives Using RhAAC.

Moses and co-workers135 reported 2-substituted alkynyl-1-sulfonyl fluorides (SASFs) 269 as a new class of connective hubs. Stereoselective DOC (diversity-oriented clicking) of SASFs 269 with azide 270 showed various 1,5-substituted 1H-1,2,3-triazole-4-sulfonyl fluorides 271 in the presence of [Rh(CO)2Cl]2 catalysts and DCE solvent for 16 h at 40 °C (Scheme 69). Method A was effective with the electron-poor substrate but ineffective for electron-rich substrate caused by separation problems, while method B is effective for electron-rich reactants.
Scheme 69. Use of [Rh(CO)2Cl]2 as a Promoted Synthesis of 271.
2.5.2. [Rh(COD)Cl]2-Catalyzed Synthesis
Li and co-workers136 reported an unprecedented and challenging approach for five-membered atropisomerism of 1,2,3-triazole 272 as chiral units. Axially chiral 1,4,5-substituted 1,2,3-triazole 272 was synthesized from internal alkyne 257 and azides 270 using enantioselective Rh catalyst, ligand (L) 273, and 4 Å MS (molecular sieves) in toluene at 10 °C for 10 h (Scheme 70). Here, employment of 4 Å MS enhanced the enantioselectivity and efficiency of E-RhAAC. This approach showed excellent functional group tolerance, as well as electronic properties of the substituent of the aryl rings, but does not affect the enantioselectivity and efficiency of E-RhAAC. Based on a DFT study and experimental observation, it was found that the hydroxy group plays an important role in regioselectivity.
Scheme 70. Synthesis of Axially Chiral Triazole 272.
2.6. Palladium-Catalyzed Azide–Alkyne Cycloaddition (PdAAC)
Mechanistically, there are two paths A and B for this PdAAC reaction. Both paths A and B are expected to lead to the title product. Surface-bound Pd-alkyne complex A collapses to produce complex B, which might lead through either pathway A or pathway B. These pathways are characterized by the earlier (path A) and later (path B) substitution reaction step using organohalide. Meanwhile, treatment of commercially available 4-phenyl-1H-1,2,3-triazole with 1,2-dichloroethane under the same condition did not yield 1-(2-chloroethyl)-4-phenyl-1H-1,2,3-triazole, ruling out pathway A (Figure 11).137,138
Figure 11.

Mechanistic cycle for the Pd catalyst.137,138 Reprinted with permission from ref (137). Copyright 2015 Royal Society of Chemistry.
2.6.1. Pd(PPh3)4-Catalyzed Synthesis
Kar and co-workers139 reported the synthesis of 1,2,3-triazolopolyhydroarenes/cycloalkenes 203 class using tandem Sonogashira coupling–CuAAC reaction. 275 was synthesized by Pd(0)–Cu(I)-catalyzed intramolecular heteroannulation of 2-/1-azidomethyl-1-/2-bromodihydronaphthalenes/arene/cycloalkenes 274 and 1 using catalyst Pd(PPh3)4, cocatalyst CuI, and Et3N as the base at 60 °C using DMF as a solvent in an inert atmosphere (Scheme 71). This efficient method gave moderate to good yields of up to 59% and required 8–10 h for completion. This approach is very helpful for the synthesis of potential bioactive fused triazoloarenes.
Scheme 71. Synthesis of 275 with Its Derivatives through Sonogashira Coupling Using Pd(PPh3)4.

2.6.2. Pd@PR Nanocatalyzed Synthesis
Das and co-workers137 developed an efficient method for synthesizing 4-aryl-1-alkyl-1H-1,2,3-triazoles 280, 281 and 282 from selective monoazidation of 1,2-dihaloethane 278/279, 2-bromoethanol 277 and benzyl bromide 276 with sodium azide 44 and terminal aryl alkynes 46 using Pd@PR (polystyrene resin-supported palladium[0]) nanocomposite and DMF as a solvent for 8–12 h at 100 °C, respectively. 282 further reacted with K2CO3 and DMF at 110 °C for 3 h, giving the corresponding N-vinyl derivatives 283. Pd@PR-catalyzed Heck coupling reaction of 283 with aryl iodide 284 in the presence of K2CO3 and DMF at 120 °C for 20 h gave 4-aryl-1-(2-arylalkenyl)-1H-1,2,3-triazoles 285. In addition, Pd@PR-catalyzed MW-assisted dehydrohalogenative Heck coupling expanded the scope of N-vinyl-1H-1,2,3-triazole 283 (Scheme 72). The aryl alkynes with electron-releasing and electron-withdrawing functional groups produced the same yields comparatively. Both heteroaromatic and polyaromatic alkynes were suitable for this phenomenon. Under identical reaction conditions, 1-bromo-2-chloroethane in place of DCE generated an indivisible mixture of corresponding Cl and Br derivatives in moderate yield. Due to more prevalent SN2 at the C–Br center, a higher percentage of chloro derivatives was produced. This novel phenomenon can provide outstanding examples of Pd@PR nanocomposite catalytic efficiency and selectivity as well as industrial interest.
Scheme 72. Synthesis of 280–283 and Their Derivatives Using Polystyrene Resin-Supported Pd(0) Nanocomposites.
2.7. Gold-Catalyzed Azide–Alkyne Cycloaddition (AuAAC)
Au cycloaddition followed a stepwise mechanism as in the previous report141 in which due to Au(I) active metal ion A, the electron density of alkynes is decreased. So, azide undergoes nucleophilic attack on B,142,143 which leads to the formation of six-membered intermediate C. Finally, 3 was derived by the removal of gold from D (Figure 12).141
Figure 12.

Au/TiO2-catalyzed cycloaddition.140,141 Reprinted from ref (140). Copyright 2013 American Chemical Society.
2.7.1. Au/TiO2 Nanocatalyzed Synthesis
Muthusubramanian and co-workers140 reported the green, efficient, and regioselective synthesis of 1,4-disubstituted triazoles 287 and 288 from substituted phenacyl azide 286 with terminal/internal alkynes 1/200 using porous Au/TiO2 nanoparticles in water at RT for 20–30/45 min, respectively (Scheme 73). They also reported that solvents such as THF, DMSO, and ethanol showed moderate yields, and water or tBuOH/water resulted in remarkable yields as a single regioisomer, whereas p-xylene, 1,2-dichloroethane, acetonitrile, and toluene offered poor yields. Researchers further investigated that a stepwise technique for the synthesis of triazoles performed better than a one-pot MCR using azide 286, phenacyl/alkyl bromide, and 1 in 75% yield for 30 min. The catalyst was synthesized using the deposition–precipitation method and used up to five cycles without remarkable loss in the yield. For the substrate scope of alkynes, they found that aliphatic and aromatic alkynes with electron-withdrawing and electron-donating groups were suitable for this approach.
Scheme 73. Utilizing Gold-Titania Nanocomposites to Prepare 287 and 288.

2.8. Silver-Catalyzed Azide–Alkyne Cycloaddition (AgAAC)
A mechanistic cycle of the AgAAC reaction is discussed below (Figure 13). The active catalyst A is formed by losing electrons from 18e– molecules, which generates ligated silver(I) acetylide B. Nucleophilic interaction of azide 2 with B produces intermediate C, which is indicated as a 16e– system, in which nitrogen migrates to carbon, carrying both electrons to produce metalated triazole D,145 which leads to targeted molecule 3 on protonation and regenerates the active catalyst.
Figure 13.

Mechanistic cycle of AgAAC. Reprinted with permission from ref (144). Copyright 2011 John Wiley and Sons.
2.8.1. Ag(I)-Complex-Catalyzed Synthesis
McNulty and co-workers144 reported a novel, mild, regioselective, and the first purely silver-catalyzed azide–alkynes cycloaddition (AgAAC) reaction for the synthesis of 1,4-disubstituted triazoles 292 from azides 291 and alkynes 41 using a 2-diphenylphosphino-N,N-diisopropylcarboxamide-ligated silver(I)acetate complex as catalyst 290, caprylic acid, and PhMe at RT for 48 h (Scheme 74). Catalyst 290 was synthesized using silver(I)acetate with the amide of the tunable 2-diphenylphosphinobenzoic acid ligand 289 and CH2Cl2. They found that the hemilabile nature of the P,O-type ligand may play an important role in cycloaddition because Ag(I) salts alone did not give cyclization products. This hemilabile nature of the ligand may contribute to opening coordination sites for azide complex formation and provide electron density to the alkyl bond via metal to achieve cyclization.
Scheme 74. Silver-Catalyzed Azide–Alkyne Cycloaddition to Form 292.

McNulty and co-workers145 developed a well-defined, chemically stable, highly effective homogeneous silver(I) 293 catalyst for the regioselective synthesis of 1,4-disubstituted triazole 3 from 2 and 1 using caprylic acid and PhMe at 90 °C. The reaction required 24 h for completion (Scheme 75). This approach provides a broad substrate scope as well as a catalyst that can be reused three times without noteworthy loss in product.
Scheme 75. Formation of 3 Using the Homogeneous Silver(I) Catalyst 293.

Sarma and co-workers146 reported a highly efficient, robust, as well as novel approach for the regioselective synthesis of 3 from 2 and 1 catalyzed by highly efficient silver dicyanamide/DIPEA using H2O/ethylene glycol at RT for 2–6 h (Scheme 76). For the substrate scope of alkynes, they found that aromatic alkyne gave slightly more yield (90–97%) compared to aliphatic alkyne (90–94%). Various functional groups containing alkynes such as alcohol and ester are also suitable substrates for this approach.
Scheme 76. AgN(CN)2, DIPEA-Catalyzed Synthesis of 3 in Aqueous Glycol at RT.

2.9. Zinc-Catalyzed Azide–Alkyne Cycloaddition (ZnAAC)
Based on a deuterated study148 and according to the proposed mechanistic study of the ZnAAC reaction, transformation initially metalizes the alkyne C–H due to amine base, and formation of zinc acetylide B occurs.149,150 Azide 2 attack on acetylide accelerated the generation of a six-membered intermediate C. After a subsequent reaction of D with an electrophile, desired product 4 was obtained (Figure 14).147
Figure 14.

Proposed mechanistic for ZnAAC. Reprinted from ref (147). Copyright 2013 American Chemical Society.
2.9.1. ZnEt2-Catalyzed Synthesis
Greaney and co-workers147 reported a regioselective synthesis of 4 from 1 and 2 using ZnEt2 as a catalyst and NH4Cl at RT for 72 h (Scheme 77). At high temperatures, the reaction gave less yield due to the decomposition of the reactant. For the alkyne substrate scope, they found that ester, thioether, propargylic ether, 1,2-diphenyl acetylene, and (iodoethynyl)benzene were suitable, whereas for the azide scope, tosyl azide and alkyl azide were not suitable for this approach. However, substitution at the 4-position was also observed by replacing NH4Cl with D2O/D3CCO2D through a 3-MCR coupling reaction to afford 1,4,5-trisubstituted triazole. The above approach takes 18 h for completion at ambient temperature.
Scheme 77. Regioselective Synthesis of 4 and 294 Using ZnEt2.

2.9.2. Zn/C-Catalyzed Synthesis
To overcome the problem of separation of homogeneous catalysts from products and relatively few advancements done on heterogeneous solid catalysts, Chen and co-workers151 develop a novel, mild, and heterogeneous Zn/C (zinc on charcoal)-catalyzed approach for the synthesis of 4 and 59 from 2 with aryl alkynes 1 and 200, respectively (Scheme 78). They also found that aprotic and polar solvent DMF was a suitable solvent, and 50 °C was the suitable temperature for this approach. Zn dust typically contains Cu(II) impurities, whereas charcoal might be reduced from Cu(II) to Cu(I). They thought that a trace amount of CuI produced in situ catalyzed cycloaddition instead of Zn/C, but no cycloaddition products were observed. It is suggested that Zn/C could catalyze the reaction. They also reported that electron-deficient aryl alkynes gave less yield than electron-rich substituents, and that aliphatic alkynes did not cause a reaction. This approach was also performed under a one-pot, three-component reaction that produced less yield. The reaction required 15–20 h, and the efficiency of the catalyst decreased after the fifth cycle.
Scheme 78. Use of Zn/C as a Heterogeneous Catalyst to Synthesize 4 and 59.

2.9.3. Zn(OTf)2-Catalyzed Synthesis
Eppakayala and co-workers152 developed a simple and efficient synthesis of novel benzimidazole-linked triazoles 298 from 2-(4-azidophenyl)1H-benzo[d]imidazole 297 and alkynes 1 using t-BuOH/H2O, Zn(OTf)2, and NaASc at RT. For the synthesis of compound 297, o-phenylenediamine 159 and 4-aminobenzoic acid were reacted in the presence of PPA at 250 °C for 4 h, yielding compound 296, which further reacted with sodium azide 44 to obtain desired 297 azides (Scheme 79).
Scheme 79. Synthesis of 298 Using Zn(OTf)2 Catalyst.

2.10. Lanthanide-Catalyzed Azide–Alkyne Cycloaddition (LnAAC)
A possible chemical path for the Ln[N(SiMe3)2]3-catalyzed cycloaddition of 1 with azide is shown in Figure 15.153 Activation of the C–H bond of 1 proceeds through the generation of lanthanide acetylide A and release of HN(SiMe3)2. The important intermediates (C and C′) are formed by the coordination and subsequent 1,1-insertion of azide into the Ln–C bond of A.155 Following that, the distant N atom’s intramolecular anti-nucleophilic interaction on a p-coordinated alkyne moiety would result in the generation of triazole complex (D). Then protonation of D with another alkyne results in triazole 4 and the regeneration of lanthanide acetylide A. On the other hand, path b might be used to complete the catalytic cycle.
Figure 15.

Proposed mechanism for SmAAC reaction153,154 Reprinted with permission from ref (153). Copyright 2013 Royal Society of Chemistry.
2.10.1. Sm[N(SiMe3)2]3-Catalyzed Synthesis
Zhou and co-workers153 reported a first organolanthanide-/rare-earth-metal-catalyzed cycloaddition with broad substrate scope, mild conditions, and easily available catalyst. They synthesized 4 from 2 and 1 using Sm[N(SiMe3)2]3, nBuNH2, and toluene at 50 °C for 24 h with good to excellent yield (Scheme 80). This catalytic approach differentiates internal and terminal alkynes. It also exhibits novel mechanistic characteristics such as a tandem anionic cascade cyclization reaction and antiaddition across the C≡C triple bond. For the substrate scope, they found that aryl azides are more reactive than alkyl azides because the aryl ring promotes the delocalization of negative charge to its neighboring N atom via conjugation, leading to an increased nucleophilicity of the N atom during the cyclization process. Noncoordinated substituents at the para-position of aromatic azides had little effect on the reaction. o-Methoxyphenyl azide formed in moderate yield, which might be due to the chelating coordination capable of increasing the bond between the Ln–N kinetic inertness and thermodynamic stability. Furthermore, the competitive coordination of nitro to the metal caused a decrease in yield. Moreover, the reactivity of alkyne is lowered by the strong chelating coordination of the pyridyl group. They also observed that 1-azido-4-(azidomethyl)benzene formed mono- and dicyclized compounds, depending on the stoichiometric amount of alkyne, while hexa-1,5-diyne gave only monocyclization product.
Scheme 80. [Sm[N(SiMe3)2]3-Promoted Synthesis of 4.

3. Organocatalytic Synthesis of 1,2,3-Triazole
Heavy metals in biological systems have been linked to cellular toxicity, oxidative damage, and metabolic instability.156 To date, several metal-free (3 + 2) cycloaddition methods have been reported to synthesize numerous functionalized 1,2,3-triazoles. As a result, considerable scientific efforts have been put toward the evolution of a metal-free approach for the synthesis of triazoles in mild conditions. Organocatalysts have recently received a lot of interest in comparison with metal catalysts.157,158 To accelerate the chemical reaction, organocatalysis employs a small organic moiety mostly consisting of C, P, N, O, H, and S. So, if we compared it with transition metal catalysts, the benefits of organocatalysis include their lack of sensitivity toward moisture and oxygen, their inexpensive, lower toxicity, and ready availability.159 In 2008, Ramachary and co-workers167 reported the first organo-catalyzed method for the synthesis of triazole. Later on, Bressy and co-workers160 reported a synthesis of triazole using unactivated ketones, which becomes a solution for avoiding even trace amounts of metal residues in the final product. Here, we covered DBU, l-proline, diethylamine, pyrrolidine, and prolinamide-catalyzed approaches for the synthesis of 1,2,3-triazole.
Organocatalyzed synthesis is useful for metal-free as well as a one-pot combination of MCRs. Here, ketones are used as a surrogate of alkyne for enamine formation and various fully functionalized 1,2,3-triazoles, which are only obtained intermediately (Figure 16).160,162,163 According to the catalyst screening done by Wang and co-workers,161 in addition to a secondary amine, primary and tertiary amines do not show significant catalytic activity and gave very low yields of 26 and <5%, respectively.
Figure 16.

Organocatalytic synthesis of 1,2,3-triazole.160,161 Reprinted with permission from ref (160). Copyright 2011 John Wiley and Sons.
3.1. 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU)-Catalyzed Synthesis
Dehaen and co-workers164 describe a novel approach for the synthesis of nonsymmetrical 5,5′-C,C-linked bi-1,2,3-triazoles 305 and 306 from 5-formyl-1,2,3-triazole 302. 5-Methoxybenzyl analogue 301 was easily synthesized from methyl 4-methoxyacetoacetate 299 and phenyl azide 300 using DBU and DMSO at ambient temperature for 3 h (Scheme 81). A subsequent photochemical conversion of 301 with bromine yielded 302 with a 65% yield. Aldehyde 302 was further transformed into nitroalkene derivatives 304 via oxidative [3 + 2] cycloaddition reaction. This aldehyde 302 and nitroalkene 304 are used as a reactant for the synthesis of axially chiral unsymmetrical tetra-ortho-substituted 5,5′-bi-1,2,3-triazoles 305 and 306, respectively. More steric hindered ortho-substituents prevent atropisomerism. Both approaches showed a broad scope for alkyl and aryl azide with various functional groups.
Scheme 81. DBU-Catalyzed Synthesis of Triazoles 305 and 306.
Mandal and co-workers165 developed a regioselective, efficient, straightforward, as well as environmentally friendly protocol for the synthesis of sulfonyl-1,2,3-triazolyl glycoconjugates 311 and completely substituted 1,2,3-triazolyl glycoconjugates 312 from glycosyl azides 307, β-keto sulfone 308, and substituted phenyl aldehyde 309 using DMSO in an open flask at 50 °C for 1 h and DBU and DMSO for 1 h at RT, respectively (Scheme 82). For the DBU-promoted reaction, they found that aprotic polar solvents such as DMSO and DMF gave a good result compared to that with MeOH and CH3CN (30% yield). They also reported that DMF with DBU gave 85% yield, whereas DMSO with DBU resulted in 92% yield. These results clearly showed that the formation of reactive enolates occurs, and DABCO, benzylamine-like less basic tertiary amine, gave moderate yield compared to that with DBU, while Et3N in DMSO did not have a reaction after 24 h at RT.
Scheme 82. DBU and Pyrrolidine-Catalyzed Synthesis of 311 and 312.

Ramachary and co-workers166 reported a regiospecific and mild DBU-catalyzed synthesis of 314 and its derivative 315. Ketones 313 and azides 2 reacted with DBU and DMSO at RT for 0.75–1.5 h to synthesize 314, which further reacted with Raney Ni and EtOH at RT for 1–3 h to obtain 315 (Scheme 83).
Scheme 83. DBU-Promoted Synthesis of 314 and 315.
3.2. l-Proline-Catalyzed Synthesis
Bressy and co-workers160 reported a highly substituted triazole 318 from unactivated ketone 316 and aryl azides 119 using l-proline 317 as an organocatalyst and dichloromethane at 80 °C under thermal conditions (Scheme 84). All cyclic ketones gave good yields. Among all the cyclic ketones cyclooctane gave a higher yield, and the dissymmetrical cyclic ketone showed excellent regioselectivity. They also reported that 4-(NO2)PhN3 as well as acetophenone did not react through this approach. Here, the reaction’s regioselectivity depends on a combination of two factors. First, regioselectivity results from the preferred form from two enamine intermediates, and second, the regioselectivity is caused by the addition of azide to the enamine.
Scheme 84. l-Proline-Catalyzed Synthesis of 318.

Ramachary and co-workers167 reported the first NH-1,2,3-triazoles 321 and 322 from Hagemann’s ester 319 and p-toluenesulfonyl azide 320 using amino acid (l-proline) catalyst 317 and DMSO as a solvent at RT for 0.75–24 h (Scheme 85). This approach gave 322 as a major product. If the reaction proceeds with BnNH2 and DMSO, it yields diazo compound 323 as a major product.
Scheme 85. Synthesis of 322 via l-Proline.

3.3. Diethylamine-Catalyzed Synthesis
Wang and co-workers162 reported the first metal-free, regioselective, and organocatalyst-promoted enamide–azide cycloaddition reaction using diethylamine 325 as a catalyst. 1,4,5-Trisubstituted 1,2,3-triazole 59 was synthesized from β-ketoester 324 and azides 2 using DMSO as a solvent at 70 °C with 80–96% yields (Scheme 86). Also, this approach is useful for various functional groups like ester, ketone, and nitrile for further synthetic applications. Without affecting the yields, the electron-donating group required a longer time compared to the electron-withdrawing group. However, alkyl azide is also a suitable substrate for this approach but required 10 mol % catalyst loading.
Scheme 86. Synthesis of 59 and Its Derivatives Using Et3N Catalyst in DMSO.

Alves and co-workers168 reported a mild, efficient, and novel approach for a wide range syntheses of (arylselanyl)-phenyl-1H-1,2,3-triazole-4-carboxamides 328 from β-oxo-amides 326 and aryl azidophenyl selenides 327 using diethylamine 325 catalyst and DMSO at RT for 2–5 h (Scheme 87). For the substrate scope of β-oxo-amides, they reported that the electron-deficient group nearer to the oxo position (−R) decreased the yield, while the EWG or EDG group on the amide aromatic ring (−R1) gave a good isolated product. However, when they performed a reaction with a strong electron-withdrawing group such as R1 = −NO2, only 59% desired product was obtained. For the aryl azidophenyl selenide scope of this reaction, they found that the electron-EWG-containing aromatic ring (−R3) gave good yields compared to those with the EDG. This approach may be useful in the future for the synthesis of novel selenium-containing triazole scaffolds.
Scheme 87. Development of Analogues of 328 Using Diethyl Amine Catalyst in DMSO at RT.
3.4. Miscellaneous Organocatalysts
3.4.1. Pyrrolidine-Catalyzed Synthesis
Wang and co-workers161 reported a Huisgen [3 + 2] cycloaddition-promoted enamine-catalyzed method for the synthesis of highly substituted triazoles 59/330 using 2 and acyclic/cyclic carbonyl compounds 324/329 using pyrrolidine 310 as a catalyst and DMSO at up to 80 °C with complete regioselectivity (Scheme 88). This approach gave moderate to good yield with a very vast substrate scope (28 examples) and was useful for further sophisticated heterocyclic scaffolds. Both cyclic and acyclic ketone compounds gave a good yield with this approach; however, the acyclic ketone gave a good yield compared to that with the cyclic ketone. They also reported that the reaction of TsN3 with cyclohexanone in the presence of l-proline did not react, while the pyrrolidine-catalyzed reaction gave an 85% yield.
Scheme 88. Pyrrolidine-Mediated Synthesis of 59/330.

Ramachary and co-workers163 developed a one-pot combination of MCR pyrrolidine 310 catalyzed functionalized bicyclic N-aryltriazole 332 from activated cyclic enone 319 and aryl azides 119 using DMSO as a solvent at RT for 1–6 h (Scheme 89). For the azide scope of the method, they found that the electron-withdrawing group and neutral azides resulted in good to excellent yields. However, 4-MeOC6H4N3 did not react even at a higher temperature. The predicted bicyclic N-aryltriazole was obtained in good to excellent yields from aliphatic and aromatic unmodified cyclic enones. The aryl substitution at the C6 position of cyclic enone enhanced the yield, while aliphatic substitution lowered the yield and increased the reaction time.
Scheme 89. Synthesis of Derivatives of 332 through the Use of Pyrrolidine Catalyst.
3.4.2. Prolinamide-Catalyzed Synthesis
Wang and co-workers169 reported a green, enamine-catalyzed, and more feasible method for the synthesis of triazole 59 from cycloaddition of ketones 324 and 2 using water as a solvent. To access a vast library of 1,2,3-triazoles, a longer aliphatic chain-tolerated prolinamide 333 was used as an effective organocatalyst to fully drive the Huisgen 1,3-dipolar cycloaddition (Scheme 90a). This catalyst is effective as a broad substrate scope for the reaction.
Scheme 90a. Use of Proline Amide as a Catalyst to Synthesize 59 and Its Derivatives.

Cyclic ketones with six- to eight-membered rings are suitable substrates for this approach, and naphthalene is a suitable substrate for the azide scope of this reaction.
They also reported that dissymmetrical cyclic ketone has a high level of regioselectivity (Scheme 90b). For example, 3,3-dimethylcyclohexanone 335 resulted in a single regioisomer 337 because the heterocycle is sterically separated from the gem-dimethyl group, while the isomer of 335, which is 4,4-dimethylcyclohexanone 336, yielded 338, which is also an isomer of 337 because cycloaddition occurs with the most stable enamine to explain the regioselectivity.
Scheme 90b. Prolinamide-Catalyzed Synthesis of 337 and 338.

4. Metal-Free Synthesis of 1,2,3-Triazole
Apart from this organocatalysis method, some simple, green, general, efficient, and convenient metal-free approaches are developed for the various types of functionalized triazole derivatives. Some of the reactions are important for the synthesis of thiolated and sulfone-functionalized-1,2,3-triazoles. For example, Wan and co-workers170 reported a metal-free, organic solvent-free, and easy synthesis of 5-thiolated 1,2,3-triazole. Mohanan and co-workers171 reported mild as well as metal-free synthesis of a sulfone-functionalized triazole scaffold. These methods involve some base, solvent, ionic liquid, and molecular iodine.
Harmata and co-workers172 reported an experimentally convenient synthesis of N-TBS-S-alkynyl sulfoximines 341 from sulfoximines 339 and methyl ester (RCO2Me) using LDA as a base, THF, Tf2O, and pyridine with good to excellent yield. Further utilization of 341 with 334 in H2O at reflux for 18 h gave a mixture of regioisomers of 1,4,5-trisubstituted triazoles 342 and 343 with a good yield (Scheme 91). They also reported that the steric influence controls the regioselectivity of the reaction.
Scheme 91. Regioselective Synthesis of Triazoles 341 and 343 Using LDA Base.
Wan and co-workers170 reported a metal-free synthesis of sulfur-containing triazole 272 from easily available tosyl azide 320 and β-thiolated enaminones 344 using TMEDA as a base and water in a sealed tube at 120 °C (Scheme 92). This approach has broad substrate scope with excellent yield. The use of water as the only medium173−176 provided a greener route without interfering with any trace amount of metal toxicity.
Scheme 92. Green Synthesis of 345 Using TMEDA.

Mohanan and co-workers171 reported a metal-free, mild reaction for the synthesis of fully substituted sulfonyl triazole 349 from benzaldehyde 346, primary amine 347, and α-diazo-β-keto sulfone 348 using K2CO3, I2, and EtOH at RT for 48 h (Scheme 93). For the substrate scope of the approach, they determined that aryl aldehyde containing and electron-deficient group did not react because of rapid hydrolysis of imine, and heteroaryl and bi- and trisubstituted aryl aldehyde were also suitable for this approach. Aliphatic, aromatic, as well as sterically crowded primary amine are suitable for this reaction and gave moderate to good yield.
Scheme 93. Synthesis of 349 under Metal-Free Conditions.

Pokhodylo and co-workers177 reported a base-catalyzed cycloaddition of β-keto sulfone/β-nitrile sulfones 350 and aryl azides 119 using MeONa and MeOH at RT to synthesize 1H-1,2,3-triazol-4-yl sulfones 351 with moderate to good yield for 2–5 h (Scheme 94). These molecular libraries are very useful for biological activity testing. The internal electrophile in β-keto sulfone or β-nitrile sulfone might be a carbonyl or cyano group. Also, sulfonyl azides were employed for diazo transfer to the CH-acid moiety, specifically for activated β-keto ester and β-keto sulfones.178,179 In this approach, enolate or enol attacks the 119, and the formation of triazene occurs, which combines with the diazo moiety and sulfonamide after tautomerization.
Scheme 94. Synthesis of 351 Using a Base Catalyst.

Back and co-workers180 reported a cycloaddition reaction of acetylenic sulfones 352 and benzyl azide 76 using toluene, LiOH, and THF for 3–7 days for the synthesis of 1,4,5-trisubstituted triazole isomers 353 and 354 (Scheme 95).
Scheme 95. Synthesis of 353 and 354 from Acetylenic Sulfones.
Shafii and co-workers181 reported the synthesis of 1-(4-aminosulfonylphenyl)-5-aryltriazoles and 1-(4-methylsulfonylphenyl)-5-aryltriazoles 359 as COX-2 inhibitors. First, they prepared an imino compound 357 by the reported method,182 but it did not work. So they performed this reaction in a mixture of ethanol/THF from 4-methylsulfonylaniline 355 and substituted benzaldehyde 356 to afford compound 357 with a good yield (Scheme 96). According to the literature, subsequent 1,3-cycloaddition of 357 with diazomethane in ether also did not provide a cyclization product.183 Ultimately, cycloaddition was achieved with diazomethane in dioxane/water184 to obtain 1-(4-methylsulfonylphenyl)-5-aryl-4,5-dihydro-1H-1,2,3-triazoles 358. Subsequent oxidation of 358 with potassium permanganate in acetone gave a low yield185 of product of 359 but achieved moderate yield (40–52%) with a phase transfer catalyst (tetrabutylammonium chloride).186 For the synthesis of 1-(4-aminosulfonylphenyl)-5-aryltriazoles, the −NH2 bearing imino compound did not undergo cycloaddition with diazomethane in dioxane. Therefore, 4-acetamidobenzenesulfonyl chloride was reacted with dibenzyl amine in THF followed by acid hydrolysis to obtain compound 362, which further reacted with substituted benzaldehyde in p-toluenesulfonic acid to give imino compound 363. Further reaction of 363 gave 364, which further oxidized in the presence of KMnO4 to give compound 365. Deprotonation of compound 365 yielded the desired product 1-(4-aminosulfonylphenyl)-5-aryltriazoles 366 with 20–49% yield.
Scheme 96. Synthesis of 359 and 366.
Fokin and co-workers187 reported a mild, transition-metal-free approach for the synthesis of 1,5-disubstituted triazole 367 from terminal aryl alkynes 41 and aryl azides 119 using tetraalkylammonium hydroxide in DMSO at RT (Scheme 97). For the substrate scope, it was found that aryl, heteroaryl, terminal alkynes, and base-labile functional groups are suitable for the reaction. Due to the low acidity, alkyl acetylene did not provide any yield under this condition.188 Electron-deficient azides and alkyne both gave lower yields due to ineffective triazenide cyclization and a decrease in nucleophilicity.
Scheme 97. Regioselective Synthesis of 367 by NMe4OH and DMSO.

The aryl acetylide A formed by reversible deprotonation of the 41 reacts as a Nu– for an attack on the terminal nitrogen of aryl azide 119 and forms triazenide intermediate B, which undergoes 6π-electrocyclization or 5-endo-dig cyclization to generate the 1,5-disubstituted triazolyl anion C and achieves the catalytic cycle by deprotonation of DMSO, water, or a terminal alkyne to form 367.
Wan and co-workers189 developed a novel, regioselective, metal- and azide-free 3-MCR approach for the synthesis of 1,5-disubstituted triazole 371 using enaminones 368, tosyl hydrazine 369, and primary amines 370 using molecular iodine and DMSO at 110 °C (Scheme 98). For enaminones, they found that the electron-donating group containing the aryl ring gave a good yield compared to that with the electron-withdrawing group, and heteroaryl-based (thiophene) enaminones are also a suitable substrate for this reaction. Primary alkyl amine is not a suitable substrate for this reaction due to the inactive key intermediate.190
Scheme 98. Synthesis of 371 Using Molecular Iodine.

4.1. Strain-Promoted Azide–Alkyne Cycloaddition (SPAAC)
Bioorthogonal synthesis of such chemically modified biomolecules is an emerging field of chemistry as well as biology. Besides these, the CuAAC reaction is important for such synthesis, but the cytotoxicity of the copper catalyst is still a disadvantage for bioorthogonal chemistry. To reduce this restriction, Bertozzi and co-workers191,192 performed the bioorthogonal, copper-free, and strain-promoted azide–alkyne cycloaddition (SPAAC).
Wills and co-workers193 reported a catalyst-free and strain-promoted azide–alkyne cycloaddition of strained alkynes 372 with azide 2 using MeCN at room temperature to 60 °C for 1–14 days or using PhC(Cl)NOH and DMF at room temperature for 3 days for the synthesis of functionalized triazole 373 with 81–96% yield (Scheme 99). These strained alkynes 372 were synthesized from biaryl diols and 1,4-ditosylbut-2-yne.194
Scheme 99. Strain-Promoted Synthesis of 1,2,3-Triazole 373.

For the versatile use of the SPAAC reaction, Kii and co-workers195 reported a catalyst-free strain-promoted “double-click” (SPDC) reaction of Sondheimer (sym-dibenzo-1,5-cyclooctadiene-3,7-diyne) 374, which has two highly strained alkyne bonds with two azido molecules, 375 and 376, which gave the desired 1,2,3-triazole product 377 (Scheme 100). Here, they use these azido groups which are azido biomolecules (such as proteins, sugars, lipids, and nucleotides) 375 and small azido compounds (fluorescent dyes, photoreactive group, chemical ligands) 376.
Scheme 100. Synthesis of Functionalized Triazole 377 via SPDC.
5. Solvent- and Catalyst-Free Neat Synthesis of 1,2,3-Triazole
For utilization of greener and more simple routes for the synthesis of the titled compound, some metal-/metal-additive-free, catalyst-free, and solvent-free reactions are important because these routes do not require any purification step.
Gouin and co-workers196 reported a metal-free, green, and regioselective generation of 4-sulfonyl-functionalized triazoles 379 and 380 from p-toluene sulfonyl alkyne 378 and azides 2 at ambient temperature (Scheme 101). The strong withdrawing action of the sulfonyl group lowers the activation energy barrier for the [3 + 2] cycloaddition and allows for direct incorporation of a chemical group onto the aromatic ring for structuring the titled compound. First, the reaction mixture was dissolved in a minimum amount of dichloromethane, and then the mixture was evaporated at 16 °C under reduced pressure for 5 min and then took 2 h under MW conditions to complete the reaction with high regioselectivity. Steric hindrance does not affect the reaction time and product yield.
Scheme 101. Solvent-Free and Catalyst-Free Synthesis of 379 and 380.
Surya Prakash and co-workers197 reported a solvent- and catalyst-free microwave-assisted synthesis of 4,5-disubstituted triazoles 382 using alkynes 200 and trimethylsilyl azide (TMSN3) 381 at a constant temperature of 200 °C for 1–7 h under inert conditions (Ar/N2). After MW irradiation, the reaction vessels were introduced to air for 30 min (Scheme 102). For the alkyne substrate scope, they found that electron-deficient aromatic, electron-rich aromatic, halogenated, allylic, as well as symmetric and asymmetric substrates are also suitable for this approach. Because the polar moiety selectively absorbs MW irradiation, and nonpolar moieties are inert to it, the presence of polar moieties is necessary to attain more effective heating across the reaction, and as a result, a shorter reaction time can be expected as several polar moieties on the substrate increase. Because of the presence of the trimethylsilyl group, steric hindrance and electron-rich moieties also interfere with the reaction. This approach also provides a safe, atom-economical, as well as a more convenient route for the synthesis of the title compound.
Scheme 102. Microwave-Assisted Synthesis of 382.

6. Conclusion
This review paper analyzed the synthesis recorded in the past 21 years (2002–2022) for 1,2,3-triazole. The CuAAC reaction is the most efficient and often used bioorthogonal reaction because it possesses the majority of criteria for the click reaction and has huge applications in various fields. However, cytotoxicity is also a severe problem that has been detected in various types of cells. Ruthenium is a good choice for 1,5-regioisomers and has a broad substrate scope. This approach is influenced by the steric and electronic factors of the substrate. Zn and Sm are also suitable catalysts for the synthesis of 1,5-disubstituted triazole. IrAAC and RhAAC reactions have broad substrate scope as well as the same steric and electronic effect as that in the RuAAC reaction. Some other MAAC reactions have not been explored yet. Recently, organocatalysis synthesis is dominant for the title compound. Nowadays, some green and click-chemistry-based metal- and solvent-free reactions are developed. The majority of this research has been done using simple substances. A few efficient approaches for the manufacture of triazole-scaffold-containing pharmaceuticals have been recently discovered. A constant invention of new developments indicates that 1,2,3-triazoles will help lead to future organic synthesis and are useful for creating molecular libraries of various functionalized 1,2,3-triazoles.
Acknowledgments
We are thankful to the Department of Chemistry, Sardar Patel University, Vallabh Vidyanagar for proving the necessary research facilities.
Glossary
Abbreviations
- MAAC
metal-catalyzed azide–alkyne cycloaddition
- CuAAC
copper-catalyzed azide–alkyne cycloaddition
- NiAAC
nickel-catalyzed azide–alkyne cycloaddition
- RuAAC
ruthenium-catalyzed azide–alkyne cycloaddition
- IrAAC
iridium-catalyzed azide–alkyne cycloaddition
- RhAAC
rhodium-catalyzed azide–alkyne cycloaddition
- PdAAC
palladium-catalyzed azide–alkyne cycloaddition
- AuAAC
gold-catalyzed azide–alkyne cycloaddition
- AgAAC
silver-catalyzed azide–alkyne cycloaddition
- ZnAAC
zinc-catalyzed azide–alkyne cycloaddition
- SmAAC
samarium-catalyzed azide–alkyne cycloaddition
- SPAAC
strain-promoted azide–alkyne cycloaddition
- TMSCF3
trifluoromethyltrimethylsilane
- NPs
nanoparticles
- TMSIT
trimethylsilyl-5-iodo-triazoles
- TMS
trimethylsilyl
- DMSO
dimethyl sulfoxide
- DMF
dimethylformamide
- RT
room temperature
- DCM
dichloromethane
- BPB
N-benzyl proline benzophenone
- AT-CuAAC
active template Cu-catalyzed alkyne–azide cycloaddition
- PCR
polymerase chain reaction
- Tf
trifluoromethanesulfonyl
- THF
tetrahydrofuran
- β-CD-TSC@Cu
copper(I)-ion-supported thiosemicarbazide-functionalized β-cyclodextrin
- ICP-OES
inductively coupled plasma optical emission spectroscopy
- PPI
protein–protein interactions
- NfN3
nonafluorobutanesulfonyl or nonaflyl azide
- NaASc
sodium ascorbate
- Cu2O/HTNT-7
Cu2O nanoparticles supported on hydrogen trititanate nanotubes
- ICP-MS
inductively coupled plasma mass spectroscopy
- DKR
dynamic kinetic resolution
- PEG-200
polyethylene glycol-200
- GO
graphene oxide
- DFT
density functional theory
- COD
1,5-cyclooctadiene
- TFAA
trifluoroacetic acid
- SPPS
solid-phase peptide synthesis
- HFIP
1,1,1,3,3,3-hexafluoroisopropanol
- MW
microwave
- ISG
iterative sequential growth
- THP
tetrahydropyranyl
- Boc
tert-butyloxycarbonyl
- SASFs
2-substituted alkynyl-1-sulfonyl fluorides
- DOC
diversity-oriented clicking
- ET3N
triethyl amine
- Pd@PR
polystyrene resin supported palladium[0]
- DIPEA
N,N-diisopropylethylamine
- Zn(OTf)2
bis(trifluoromethanesulfonato)zinc
- PPA
polyphosphoric acid
- Ln
lanthanum
- DBU
1,8-diazabicyclo[5.4.0]undec-7-ene
- DABCO
1,4-diazabicyclo[2.2.2]octane
- TsN3
tosyl azide
- LDA
lithium diisopropylamide
- N-TBS
N-tert-butyldimethylsilyl
- Tf2O
triflic anhydride
- TMEDA
tetramethylethylenediamine
- COX-2
cyclooxygenase-2
- MCRs
multicomponent reactions
- TMSN3
trimethylsilyl azide
Author Contributions
D.P.V. and H.M.P. designed a review article collection. All authors contributed equally to the drafting.
The authors declare no competing financial interest.
This paper was published on October 10, 2022. Text related to references 160 and 167 was revised, and the corrected paper was reposted on October 13, 2022.
References
- Parker W. B. Enzymology of purine and pyrimidine antimetabolites used in the treatment of cancer. Chem. Rev. 2009, 109 (7), 2880–93. 10.1021/cr900028p. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henary M.; Kananda C.; Rotolo L.; Savino B.; Owens E. A.; Cravotto G. Benefits and applications of microwave-assisted synthesis of nitrogen containing heterocycles in medicinal chemistry. RSC Adv. 2020, 10 (24), 14170–14197. 10.1039/D0RA01378A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khanam H.; Shamsuzzaman Bioactive Benzofuran derivatives: A review. Eur. J. Med. Chem. 2015, 97, 483–504. 10.1016/j.ejmech.2014.11.039. [DOI] [PubMed] [Google Scholar]
- Naik R.; Harmalkar D. S.; Xu X.; Jang K.; Lee K. Bioactive Benzofuran Derivatives: Moracins A—Z in Medicinal Chemistry. Eur. J. Med. Chem. 2015, 90, 379–393. 10.1016/j.ejmech.2014.11.047. [DOI] [PubMed] [Google Scholar]
- Bozorov K.; Nie L. F.; Zhao J.; Aisa H. A. 2-Aminothiophene scaffolds: Diverse biological and pharmacological attributes in medicinal chemistry. Eur. J. Med. Chem. 2017, 140, 465–493. 10.1016/j.ejmech.2017.09.039. [DOI] [PubMed] [Google Scholar]
- Zheng Y.; Wu X.; Xue B.; Li M.; Ji M. Design, synthesis, docking and antitumor activity of quinazolino [3, 4-a] thieno [3, 2-d] pyrimidin-8-one derivatives. Chem. Biol. Drug Des. 2010, 76 (3), 285–290. 10.1111/j.1747-0285.2010.01008.x. [DOI] [PubMed] [Google Scholar]
- Bozorov K.; Ma H. R.; Zhao J. Y.; Zhao H. Q.; Chen H.; Bobakulov K.; Xin X. L.; Elmuradov B.; Shakhidoyatov K.; Aisa H. A. Discovery of diethyl 2,5-diaminothiophene-3,4-dicarboxylate derivatives as potent anticancer and antimicrobial agents and screening of anti-diabetic activity: synthesis and in vitro biological evaluation. Part 1. Eur. J. Med. Chem. 2014, 84, 739–45. 10.1016/j.ejmech.2014.07.065. [DOI] [PubMed] [Google Scholar]
- Bozorov K.; Zhao J. y.; Nie L. F.; Ma H.-R.; Bobakulov K.; Hu R.; Rustamova N.; Huang G.; Efferth T.; Aisa H. A. Synthesis and in vitro biological evaluation of novel diaminothiophene scaffolds as antitumor and anti-influenza virus agents. Part 2. RSC Adv. 2017, 7 (50), 31417–31427. 10.1039/C7RA04808D. [DOI] [Google Scholar]
- Majumdar P.; Pati A.; Patra M.; Behera R. K.; Behera A. K. Acid Hydrazides, Potent Reagents for Synthesis of Oxygen-, Nitrogen-, and/or Sulfur-Containing Heterocyclic Rings. Chem. Rev. 2014, 114 (5), 2942–2977. 10.1021/cr300122t. [DOI] [PubMed] [Google Scholar]
- Ahmad S.; Alam O.; Naim M. J.; Shaquiquzzaman M.; Alam M. M.; Iqbal M. Pyrrole: An insight into recent pharmacological advances with structure activity relationship. Eur. J. Med. Chem. 2018, 157, 527–561. 10.1016/j.ejmech.2018.08.002. [DOI] [PubMed] [Google Scholar]
- Phukan P.; Sarma D. Synthesis of Medicinally Relevant Scaffolds-Triazoles and Pyrazoles in Green Solvent Ionic Liquids. Curr. Org. Chem. 2021, 25, 1523. 10.2174/1385272825666210412160142. [DOI] [Google Scholar]
- Emami S.; Tavangar P.; Keighobadi M. An overview of azoles targeting sterol 14α-demethylase for antileishmanial therapy. Eur. J. Med. Chem. 2017, 135, 241–259. 10.1016/j.ejmech.2017.04.044. [DOI] [PubMed] [Google Scholar]
- Li S.; Chen J.-X.; Xiang Q.-X.; Zhang L.-Q.; Zhou C.-H.; Xie J.-Q.; Yu L.; Li F.-Z. The synthesis and activities of novel mononuclear or dinuclear cyclen complexes bearing azole pendants as antibacterial and antifungal agents. Eur. J. Med. Chem. 2014, 84, 677–686. 10.1016/j.ejmech.2014.07.075. [DOI] [PubMed] [Google Scholar]
- Juríček M.; Kouwer P. H.; Rowan A. E. Triazole: a unique building block for the construction of functional materials. Chem. Commun. 2011, 47 (31), 8740–8749. 10.1039/c1cc10685f. [DOI] [PubMed] [Google Scholar]
- Liang L.; Astruc D. J. C. C. R. The copper (I)-catalyzed alkyne-azide cycloaddition (CuAAC)“click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255 (23–24), 2933–2945. 10.1016/j.ccr.2011.06.028. [DOI] [Google Scholar]
- Ji Ram V., Sethi A., Nath M., Pratap R.. Five-Membered Heterocycles. The Chemistry of Heterocycles; Elsevier, 2019; pp 149–478. [Google Scholar]
- Rani A.; Singh G.; Singh A.; Maqbool U.; Kaur G.; Singh J. CuAAC-ensembled 1,2,3-triazole-linked isosteres as pharmacophores in drug discovery: review. RSC Adv. 2020, 10 (10), 5610–5635. 10.1039/C9RA09510A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ince T.; Serttas R.; Demir B.; Atabey H.; Seferoglu N.; Erdogan S.; Sahin E.; Erat S.; Nural Y. Polysubstituted pyrrolidines linked to 1,2,3-triazoles: Synthesis, crystal structure, DFT studies, acid dissociation constant, drug-likeness, and anti-proliferative activity. J. Mol. Struct. 2020, 1217, 128400. 10.1016/j.molstruc.2020.128400. [DOI] [Google Scholar]
- Jiang X.; Hao X.; Jing L.; Wu G.; Kang D.; Liu X.; Zhan P. Recent applications of click chemistry in drug discovery. Expert Opin. Drug Discovery 2019, 14 (8), 779–789. 10.1080/17460441.2019.1614910. [DOI] [PubMed] [Google Scholar]
- Singh N.; Pandey S. K.; Tripathi R. P. Regioselective [3 + 2] cycloaddition of chalcones with a sugar azide: easy access to 1-(5-deoxy-d-xylofuranos-5-yl)-4,5-disubstituted-1H-1,2,3-triazoles. Carbohydr. Res. 2010, 345 (12), 1641–1648. 10.1016/j.carres.2010.04.019. [DOI] [PubMed] [Google Scholar]
- Bozorov K.; Zhao J.; Aisa H. A. 1,2,3-Triazole-containing hybrids as leads in medicinal chemistry: A recent overview. Bioorg. Med. Chem. 2019, 27 (16), 3511–3531. 10.1016/j.bmc.2019.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hakimian S.; Cheng-Hakimian A.; Anderson G. D.; Miller J. W. Rufinamide: a new anti-epileptic medication. Expert Opin. Pharmacother. 2007, 8 (12), 1931–1940. 10.1517/14656566.8.12.1931. [DOI] [PubMed] [Google Scholar]
- Sharkey M.; Sharova N.; Mohammed I.; Huff S. E.; Kummetha I. R.; Singh G.; Rana T. M.; Stevenson M. HIV-1 Escape from Small-Molecule Antagonism of Vif. mBio 2019, 10 (1), e00144–19. 10.1128/mBio.00144-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng L.-S.; Zheng M.-J.; Zhao F.; Liu D. 1,2,3-Triazole hybrids with anti-HIV-1 activity. Arch. Pharm. (Weinheim, Ger.) 2021, 354 (1), 2000163. 10.1002/ardp.202000163. [DOI] [PubMed] [Google Scholar]
- Batra N.; Rajendran V.; Agarwal D.; Wadi I.; Ghosh P. C.; Gupta R. D.; Nath M. Synthesis and Antimalarial Evaluation of [1, 2,3]-Triazole-Tethered Sulfonamide-Berberine Hybrids. ChemistrySelect 2018, 3 (34), 9790–9793. 10.1002/slct.201801905. [DOI] [Google Scholar]
- Phillips O. A.; Udo E. E.; Ali A. A. M.; Al-Hassawi N. Synthesis and antibacterial activity of 5-substituted oxazolidinones. Bioorg. Med. Chem. 2003, 11 (1), 35–41. 10.1016/S0968-0896(02)00423-6. [DOI] [PubMed] [Google Scholar]
- Neu H.; Fu K. Cefatrizine Activity compared with that of other Cephalosporins. Antimicrob. Agents Chemother. 1979, 15 (2), 209–212. 10.1128/AAC.15.2.209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bock V. D.; Speijer D.; Hiemstra H.; van Maarseveen J. H. 1,2,3-Triazoles as peptide bond isosteres: synthesis and biological evaluation of cyclotetrapeptide mimics. Org. Biomol Chem. 2007, 5 (6), 971–5. 10.1039/b616751a. [DOI] [PubMed] [Google Scholar]
- Gabba A.; Robakiewicz S.; Taciak B.; Ulewicz K.; Broggini G.; Rastelli G.; Krol M.; Murphy P. V.; Passarella D. Synthesis and Biological Evaluation of Migrastatin Macrotriazoles. Eur. J. Org. Chem. 2017, 2017 (1), 60–69. 10.1002/ejoc.201600988. [DOI] [Google Scholar]
- Higashitani F.; Hyodo A.; Ishida N.; Inoue M.; Mitsuhashi S. Inhibition of β-lactamases by tazobactam and in-vitro antibacterial activity of tazobactam combined with piperacillin. J. Antimicrob. Chemother. 1990, 25 (4), 567–574. 10.1093/jac/25.4.567. [DOI] [PubMed] [Google Scholar]
- Bonandi E.; Christodoulou M. S.; Fumagalli G.; Perdicchia D.; Rastelli G.; Passarella D. The 1, 2, 3-triazole ring as a bioisostere in medicinal chemistry. Drug Discovery Today 2017, 22 (10), 1572–1581. 10.1016/j.drudis.2017.05.014. [DOI] [PubMed] [Google Scholar]
- Ma J.; Ding S. Transition Metal-Catalyzed Cycloaddition of Azides with Internal Alkynes. Asian J. Org. Chem. 2020, 9 (12), 1872–1888. 10.1002/ajoc.202000486. [DOI] [Google Scholar]
- Gomes R. S.; Jardim G. A. M.; de Carvalho R. L.; Araujo M. H.; da Silva Júnior E. N. Beyond copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition: Synthesis and mechanism insights. Tetrahedron 2019, 75 (27), 3697–3712. 10.1016/j.tet.2019.05.046. [DOI] [Google Scholar]
- Surendra Reddy G.; Anebouselvy K.; Ramachary D. B. [3 + 2]-Cycloaddition for Fully Decorated Vinyl-1,2,3-Triazoles: Design, Synthesis and Applications. Asian J. Org. Chem. 2020, 15 (19), 2960–2983. 10.1002/asia.202000731. [DOI] [PubMed] [Google Scholar]
- Sahu A. Advance Synthetic Approaches to 1, 2, 3-triazole Derived Compounds: State of the Art 2004–2020. Curr. Organocatal. 2021, 8 (3), 271–288. 10.2174/2213337208666210428123033. [DOI] [Google Scholar]
- Schulze B.; Schubert U. S. Beyond click chemistry – supramolecular interactions of 1,2,3-triazoles. Chem. Soc. Rev. 2014, 43 (8), 2522–2571. 10.1039/c3cs60386e. [DOI] [PubMed] [Google Scholar]
- Moses J. E.; Moorhouse A. D. The growing applications of click chemistry. Chem. Soc. Rev. 2007, 36 (8), 1249–1262. 10.1039/B613014N. [DOI] [PubMed] [Google Scholar]
- Kappe C. O.; Van der Eycken E. Click chemistry under non-classical reaction conditions. Chem. Soc. Rev. 2010, 39 (4), 1280–1290. 10.1039/B901973C. [DOI] [PubMed] [Google Scholar]
- Thirumurugan P.; Matosiuk D.; Jozwiak K. Click Chemistry for Drug Development and Diverse Chemical–Biology Applications. Chem. Rev. 2013, 113 (7), 4905–4979. 10.1021/cr200409f. [DOI] [PubMed] [Google Scholar]
- Kim E.; Koo H. Biomedical applications of copper-free click chemistry: in vitro, in vivo, and ex vivo. Chem. Sci. 2019, 10 (34), 7835–7851. 10.1039/C9SC03368H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breugst M.; Reissig H.-U. The Huisgen Reaction: Milestones of the 1,3-Dipolar Cycloaddition. Angew. Chem., Int. Ed. 2020, 59 (30), 12293–12307. 10.1002/anie.202003115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 (11), 2004–2021. . [DOI] [PubMed] [Google Scholar]
- Kolb H. C.; Sharpless K. B. The growing impact of click chemistry on drug discovery. Drug Discovery Today 2003, 8 (24), 1128–1137. 10.1016/S1359-6446(03)02933-7. [DOI] [PubMed] [Google Scholar]
- Shen Q.; Han E.-j.; Huang Y.-g.; Chen Q.-Y.; Guo Y. Synthesis of Fluorinated 1,4,5-Substituted 1,2,3-Triazoles by RuAAC Reaction. Synthesis 2015, 47 (24), 3936–3946. 10.1055/s-0035-1560352. [DOI] [Google Scholar]
- Destito P.; Couceiro J. R.; Faustino H.; López F.; Mascareñas J. L. Ruthenium-Catalyzed Azide–Thioalkyne Cycloadditions in Aqueous Media: A Mild, Orthogonal, and Biocompatible Chemical Ligation. Angew. Chem., Int. Ed. 2017, 56 (36), 10766–10770. 10.1002/anie.201705006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. 2002, 114 (14), 2708–2711. . [DOI] [PubMed] [Google Scholar]
- Tornøe C. W.; Christensen C.; Meldal M. Peptidotriazoles on Solid Phase: [1,2,3]-Triazoles by Regiospecific Copper(I)-Catalyzed 1,3-Dipolar Cycloadditions of Terminal Alkynes to Azides. J. Org. Chem. 2002, 67 (9), 3057–3064. 10.1021/jo011148j. [DOI] [PubMed] [Google Scholar]
- 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 (14), 2596–2599. . [DOI] [PubMed] [Google Scholar]
- Haldón E.; Nicasio M. C.; Pérez P. J. Copper-catalysed azide–alkyne cycloadditions (CuAAC): an update. Org. Biomol. Chem. 2015, 13 (37), 9528–9550. 10.1039/C5OB01457C. [DOI] [PubMed] [Google Scholar]
- Himo F.; Lovell T.; Hilgraf R.; Rostovtsev V. V.; Noodleman L.; Sharpless K. B.; Fokin V. V. Copper(I)-catalyzed synthesis of azoles. DFT study predicts unprecedented reactivity and intermediates. J. Am. Chem. Soc. 2005, 127 (1), 210–6. 10.1021/ja0471525. [DOI] [PubMed] [Google Scholar]
- Himo F.; Lovell T.; Hilgraf R.; Rostovtsev V. V.; Noodleman L.; Sharpless K. B.; Fokin V. V. Copper(I)-Catalyzed Synthesis of Azoles. DFT Study Predicts Unprecedented Reactivity and Intermediates. J. Am. Chem. Soc. 2005, 127 (1), 210–216. 10.1021/ja0471525. [DOI] [PubMed] [Google Scholar]
- Worrell B.; Malik J.; Fokin V. Direct evidence of a dinuclear copper intermediate in Cu (I)-catalyzed azide-alkyne cycloadditions. Science 2013, 340 (6131), 457–460. 10.1126/science.1229506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Okuda Y.; Imafuku K.; Tsuchida Y.; Seo T.; Akashi H.; Orita A. Process-Controlled Regiodivergent Copper-Catalyzed Azide–Alkyne Cycloadditions: Tailor-made Syntheses of 4- and 5-Bromotriazoles from Bromo(phosphoryl)ethyne. Org. Lett. 2020, 22 (13), 5099–5103. 10.1021/acs.orglett.0c01681. [DOI] [PubMed] [Google Scholar]
- Zhang L.-L.; Li M.-T.; Shen L.-L.; Wu Q.-P. Efficient Synthesis of 5-Trifluoromethylthio-1,2,3-Triazoles: One-Pot Multicomponent Reaction from Elemental Sulfur and TMSCF3. Synthesis 2020, 52 (02), 304–310. 10.1055/s-0039-1690716. [DOI] [Google Scholar]
- Wang W.; Lin Y.; Ma Y.; Tung C.-H.; Xu Z. Copper(I)-Catalyzed Three-Component Click/Persulfuration Cascade: Regioselective Synthesis of Triazole Disulfides. Org. Lett. 2018, 20 (10), 2956–2959. 10.1021/acs.orglett.8b01002. [DOI] [PubMed] [Google Scholar]
- Wang W.; Peng X.; Wei F.; Tung C.-H.; Xu Z. Copper(I)-Catalyzed Interrupted Click Reaction: Synthesis of Diverse 5-Hetero-Functionalized Triazoles. Angew. Chem., Int. Ed. 2016, 55 (2), 649–653. 10.1002/anie.201509124. [DOI] [PubMed] [Google Scholar]
- Li L.; Shang T.; Ma X.; Guo H.; Zhu A.; Zhang G. 4-Trimethylsilyl-5-iodo-1,2,3-triazole: A Key Precursor for the Divergent Syntheses of 1,5-Disubstituted 1,2,3-Triazoles. Synlett 2015, 26 (05), 695–699. 10.1055/s-0034-1379970. [DOI] [Google Scholar]
- Turner D. M.; Tom C. T.; Renslo A. R. Simple plate-based, parallel synthesis of disulfide fragments using the CuAAC click reaction. ACS Comb. Sci. 2014, 16 (12), 661–4. 10.1021/co500132q. [DOI] [PubMed] [Google Scholar]
- González-Vera J. A.; Luković E.; Imperiali B. Synthesis of Red-Shifted 8-Hydroxyquinoline Derivatives Using Click Chemistry and Their Incorporation into Phosphorylation Chemosensors. J. Org. Chem. 2009, 74 (19), 7309–7314. 10.1021/jo901369k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith C. D.; Baxendale I. R.; Lanners S.; Hayward J. J.; Smith S. C.; Ley S. V. [3 + 2] Cycloaddition of acetylenes with azides to give 1,4-disubstituted 1,2,3-triazoles in a modular flow reactor. Org. Biomol. Chem. 2007, 5 (10), 1559–1561. 10.1039/b702995k. [DOI] [PubMed] [Google Scholar]
- Wuest F.; Tang X.; Kniess T.; Pietzsch J.; Suresh M. Synthesis and cyclooxygenase inhibition of various (aryl-1,2,3-triazole-1-yl)-methanesulfonylphenyl derivatives. Bioorg. Med. Chem. 2009, 17 (3), 1146–1151. 10.1016/j.bmc.2008.12.032. [DOI] [PubMed] [Google Scholar]
- Zhang L.-L.; Li Y.-T.; Gao T.; Guo S.-S.; Yang B.; Meng Z.-H.; Dai Q.-P.; Xu Z.-B.; Wu Q.-P. Efficient Synthesis of Diverse 5-Thio- or 5-Selenotriazoles: One-Pot Multicomponent Reaction from Elemental Sulfur or Selenium. Synthesis 2019, 51 (22), 4170–4182. 10.1055/s-0039-1690618. [DOI] [Google Scholar]
- Ojaghi Aghbash K.; Noroozi Pesyan N.; Şahin E. Cu(I)-catalyzed alkyne–azide ‘click’ cycloaddition (CuAAC): a clean, efficient, and mild synthesis of new 1,4-disubstituted 1H-1,2,3-triazole-linked 2-amino-4,8-dihydropyrano[3,2-b]pyran-3-carbonitrile–crystal structure. Res. Chem. Intermed. 2019, 45 (4), 2079–2094. 10.1007/s11164-018-03723-x. [DOI] [Google Scholar]
- Larionov V. A.; Adonts H. V.; Gugkaeva Z. T.; Smol’yakov A. F.; Saghyan A. S.; Miftakhov M. S.; Kuznetsova S. A.; Maleev V. I.; Belokon Y. N. J. C. The Elaboration of a General Approach to the Asymmetric Synthesis of 1, 4-Substituted 1, 2, 3-Triazole Containing Amino Acids via Ni (II) Complexes. ChemistrySelect 2018, 3 (11), 3107–3110. 10.1002/slct.201800228. [DOI] [Google Scholar]
- Collet S.; Bauchat P.; Danion-Bougot R.; Danion D. Stereoselective, nonracemic synthesis of ω-borono-α-amino acids. Tetrahedron: Asymmetry 1998, 9 (12), 2121–2131. 10.1016/S0957-4166(98)00186-4. [DOI] [Google Scholar]
- Duan X.; Zheng N.; Liu G.; Li M.; Wu Q.; Sun X.; Song W. Copper-Catalyzed One-Step Formation of Four C–N Bonds toward Polyfunctionalized Triazoles via Multicomponent Reaction. Org. Lett. 2022, 24 (32), 6006–6012. 10.1021/acs.orglett.2c02273. [DOI] [PubMed] [Google Scholar]
- Wei F.; Zhou T.; Ma Y.; Tung C.-H.; Xu Z. Bench-Stable 5-Stannyl Triazoles by a Copper(I)-Catalyzed Interrupted Click Reaction: Bridge to Trifluoromethyltriazoles and Trifluoromethylthiotriazoles. Org. Lett. 2017, 19 (8), 2098–2101. 10.1021/acs.orglett.7b00701. [DOI] [PubMed] [Google Scholar]
- Zhou F.; Tan C.; Tang J.; Zhang Y.-Y.; Gao W.-M.; Wu H.-H.; Yu Y.-H.; Zhou J. Asymmetric Copper(I)-Catalyzed Azide–Alkyne Cycloaddition to Quaternary Oxindoles. J. Am. Chem. Soc. 2013, 135 (30), 10994–10997. 10.1021/ja4066656. [DOI] [PubMed] [Google Scholar]
- Wang C.; Zhu R.-Y.; Liao K.; Zhou F.; Zhou J. Enantioselective Cu(I)-Catalyzed Cycloaddition of Prochiral Diazides with Terminal or 1-Iodoalkynes. Org. Lett. 2020, 22 (4), 1270–1274. 10.1021/acs.orglett.9b04522. [DOI] [PubMed] [Google Scholar]
- Liao K.; Gong Y.; Zhu R.-Y.; Wang C.; Zhou F.; Zhou J. Highly Enantioselective CuAAC of Functional Tertiary Alcohols Featuring an Ethynyl Group and Their Kinetic Resolution. Angew. Chem., Int. Ed. 2021, 60 (15), 8488–8493. 10.1002/anie.202016286. [DOI] [PubMed] [Google Scholar]
- Li M.; Dong K.; Zheng Y.; Song W. Copper-catalyzed cascade click/nucleophilic substitution reaction to access fully substituted triazolyl-organosulfurs. Org. Biomol. Chem. 2019, 17 (46), 9933–9941. 10.1039/C9OB02081K. [DOI] [PubMed] [Google Scholar]
- Acevedo-Jake A.; Ball A. T.; Galli M.; Kukwikila M.; Denis M.; Singleton D. G.; Tavassoli A.; Goldup S. M. AT-CuAAC Synthesis of Mechanically Interlocked Oligonucleotides. J. Am. Chem. Soc. 2020, 142 (13), 5985–5990. 10.1021/jacs.0c01670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li M.; Duan X.; Jiang Y.; Sun X.; Xu X.; He J.; Zheng Y.; Song W.; Zheng N. Three-Component Asymmetric Polymerization toward Chiral Polymer. CCS Chem. 2022, 4 (10), 3402–3415. 10.31635/ccschem.021.202101655. [DOI] [Google Scholar]
- Li M.; Duan X.; Jiang Y.; Sun X.; Xu X.; Zheng Y.; Song W.; Zheng N. Multicomponent Polymerization of Azides, Alkynes, and Electrophiles toward 1,4,5-Trisubstituted Polytriazoles. Macromolecules 2022, 55 (16), 7240–7248. 10.1021/acs.macromol.2c00966. [DOI] [Google Scholar]
- Meng J.-c.; Fokin V. V.; Finn M. G. Kinetic resolution by copper-catalyzed azide–alkyne cycloaddition. Tetrahedron Lett. 2005, 46 (27), 4543–4546. 10.1016/j.tetlet.2005.05.019. [DOI] [Google Scholar]
- Osako T.; Uozumi Y. Enantioposition-Selective Copper-Catalyzed Azide–Alkyne Cycloaddition for Construction of Chiral Biaryl Derivatives. Org. Lett. 2014, 16 (22), 5866–5869. 10.1021/ol502778j. [DOI] [PubMed] [Google Scholar]
- Worrell B. T.; Ellery S. P.; Fokin V. V. Copper(I)-Catalyzed Cycloaddition of Bismuth(III) Acetylides with Organic Azides: Synthesis of Stable Triazole Anion Equivalents. Angew. Chem., Int. Ed. 2013, 52 (49), 13037–13041. 10.1002/anie.201306192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tajbakhsh M.; Naimi-Jamal M. R. Copper-doped functionalized β-cyclodextrin as an efficient green nanocatalyst for synthesis of 1,2,3-triazoles in water. Sci. Rep. 2022, 12 (1), 4948. 10.1038/s41598-022-08868-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. 10.3390/ecsoc-24-08384. [DOI] [Google Scholar]
- Brittain W. D. G.; Buckley B. R.; Fossey J. S. Asymmetric Copper-Catalyzed Azide–Alkyne Cycloadditions. ACS Catal. 2016, 6 (6), 3629–3636. 10.1021/acscatal.6b00996. [DOI] [Google Scholar]
- Etayo P.; Badorrey R.; Díaz-de-Villegas M. D.; Gálvez J. A. Asymmetric homologation of ketones. A new entry to orthogonally protected (2R,4R)-piperidine-2,4-dicarboxylic acid. J. Org. Chem. 2008, 73 (21), 8594–8597. 10.1021/jo801515k. [DOI] [PubMed] [Google Scholar]
- García-Urdiales E.; Alfonso I.; Gotor V. Enantioselective enzymatic desymmetrizations in organic synthesis. Chem. Rev. 2005, 105 (1), 313–54. 10.1021/cr040640a. [DOI] [PubMed] [Google Scholar]
- Berkowitz D. B.; Maeng J.-H.; Dantzig A. H.; Shepard R. L.; Norman B. H. Chemoenzymatic and Ring E-Modular Approach to the (−)-Podophyllotoxin Skeleton. Synthesis of 3′,4′,5′-Tridemethoxy-(−)-podophyllotoxin. J. Am. Chem. Soc. 1996, 118 (39), 9426–9427. 10.1021/ja961489s. [DOI] [Google Scholar]
- Zhu R.-Y.; Chen L.; Hu X.-S.; Zhou F.; Zhou J. Enantioselective synthesis of P-chiral tertiary phosphine oxides with an ethynyl group via Cu(i)-catalyzed azide–alkyne cycloaddition. Chem. Sci. 2020, 11 (1), 97–106. 10.1039/C9SC04938J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu G.; Zhao J.; Jiang Y.; Zhang P.; Li W. Design, synthesis and antifungal activity of novel indole derivatives linked with the 1, 2, 3-triazole moiety via the CuAAC click reaction. J. Chem. Res. 2016, 40 (5), 269–272. 10.3184/174751916X14597828245275. [DOI] [Google Scholar]
- Cunha Lima J. A.; de Farias Silva J.; Santos C. S.; Caiana R. R.; De Moraes M. M.; Da Camara C. A.; Freitas J. C. Synthesis of new 1,4-disubstituted 1,2,3-triazoles using the CuAAC reaction and determination of their antioxidant activities. An. Acad. Bras. Cienc. 2021, 93 (3), e20201672. 10.1590/0001-3765202120201672. [DOI] [PubMed] [Google Scholar]
- Zhou X. a.; Wan L.; Hu Y.; E Y.; Huang F.; Du L. Synthesis and characterization of novel polytriazoleimides by CuAAC step-growth polymerization. Polym. J. 2010, 42 (3), 216–222. 10.1038/pj.2009.337. [DOI] [Google Scholar]
- Valdés A. K. E.; Cuevas-Yañez E. Design and synthesis of antifungal compounds from 1, 2, 3-triazoles through the click chemistry approach. Organic Medicinal Chemistry International Journal 2019, 8 (2), 43–45. 10.19080/OMCIJ.2019.08.555734. [DOI] [Google Scholar]
- Kapupara V. H.; Kalavadiya P. L.; Gojiya D. G.; Parmar N. D.; Joshi H. S. Greener, facile and ultrasound assisted synthesis of 1, 2, 3-triazole via CuAAC of 2-(azidomethyl)-1H-benzo[d]imidazole and alkynes and their evaluation of antimicrobial activity. World Sci. News 2019, 119, 139–167. 10.1134/s1068162020050106. [DOI] [Google Scholar]
- Noor A.; Lewis J. E.; Cameron S. A.; Moratti S. C.; Crowley J. D. A multi-component CuAAC ‘click’approach to an exo functionalised pyridyl-1, 2, 3-triazole macrocycle: synthesis, characterisation, Cu (I) and Ag (I) complexes. Supramol. Chem. 2012, 24 (7), 492–498. 10.1080/10610278.2012.688126. [DOI] [Google Scholar]
- Yap A. H.; Weinreb S. M. β-Tosylethylazide: a useful synthon for preparation of N-protected 1,2,3-triazoles via click chemistry. Tetrahedron Lett. 2006, 47 (18), 3035–3038. 10.1016/j.tetlet.2006.03.020. [DOI] [Google Scholar]
- Crowley J. D.; Bandeen P. H.; Hanton L. R. A one pot multi-component CuAAC “click” approach to bidentate and tridentate pyridyl-1,2,3-triazole ligands: Synthesis, X-ray structures and copper(II) and silver(I) complexes. Polyhedron 2010, 29 (1), 70–83. 10.1016/j.poly.2009.06.010. [DOI] [Google Scholar]
- Mayooufi A.; Romdhani-Younes M.; Thibonnet J. Efficient One-Pot Synthesis of Triazole-Linked Morpholinone Scaffolds by CuAAC in the Presence of 18-Crown-6. SynOpen 2018, 02 (04), 0298–0305. 10.1055/s-0037-1610399. [DOI] [Google Scholar]
- Tan A. Synthesis, spectroscopic characterization of novel phthalimides derivatives bearing a 1,2,3-triazole unit and examination as potential SARS-CoV-2 inhibitors via in silico studies. J. Mol. Struct. 2022, 1261, 132915. 10.1016/j.molstruc.2022.132915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Avello M. G.; de la Torre M. C.; Guerrero-Martínez A.; Sierra M. A.; Gornitzka H.; Hemmert C. Chiral-at-Metal BODIPY-Based Iridium(III) Complexes: Synthesis and Luminescence Properties. Eur. J. Inorg. Chem. 2020, 2020 (42), 4045–4053. 10.1002/ejic.202000745. [DOI] [Google Scholar]
- Goldstein D. C.; Peterson J. R.; Cheng Y. Y.; Clady R. G. C.; Schmidt T. W.; Thordarson P. Synthesis and Luminescence Properties of Iridium(III) Azide- and Triazole-Bisterpyridine Complexes. Molecules 2013, 18 (8), 8959. 10.3390/molecules18088959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhoomireddy R. P. R.; Narla L. G. B.; Peddiahgari V. G. R. Green synthesis of 1,2,3-triazoles via Cu2O NPs on hydrogen trititanate nanotubes promoted 1,3-dipolar cycloadditions. Appl. Organomet. Chem. 2019, 33 (3), e4752. 10.1002/aoc.4752. [DOI] [Google Scholar]
- Wang Z.; Zhou X.; Gong S.; Xie J. MOF-Derived Cu@N-C Catalyst for 1,3-Dipolar Cycloaddition Reaction. Nanomaterials 2022, 12 (7), 1070. 10.3390/nano12071070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nayal O. S.; Thakur M. S.; Kumar M.; Shaifali; Upadhyay R.; Maurya S. K. Sustainable and Efficient CuI-NPs-Catalyzed Cross-Coupling Approach for the Synthesis of Tertiary 3-Aminopropenoates, Triazoles, and Ciprofloxacin. Asian J. Org. Chem. 2018, 7 (4), 776–780. 10.1002/ajoc.201700682. [DOI] [Google Scholar]
- Liu E.-C.; Topczewski J. J. Enantioselective Nickel-Catalyzed Alkyne–Azide Cycloaddition by Dynamic Kinetic Resolution. J. Am. Chem. Soc. 2021, 143 (14), 5308–5313. 10.1021/jacs.1c01354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim W. G.; Kang M. E.; Lee J. B.; Jeon M. H.; Lee S.; Lee J.; Choi B.; Cal P. M. S. D.; Kang S.; Kee J.-M.; Bernardes G. J. L.; Rohde J.-U.; Choe W.; Hong S. Y. Nickel-Catalyzed Azide–Alkyne Cycloaddition To Access 1,5-Disubstituted 1,2,3-Triazoles in Air and Water. J. Am. Chem. Soc. 2017, 139 (35), 12121–12124. 10.1021/jacs.7b06338. [DOI] [PubMed] [Google Scholar]
- Camberlein V.; Kraupner N.; Bou Karroum N.; Lipka E.; Deprez-Poulain R.; Deprez B.; Bosc D. Multi-component reaction for the preparation of 1,5-disubstituted 1,2,3-triazoles by in-situ generation of azides and nickel-catalyzed azide-alkyne cycloaddition. Tetrahedron Lett. 2021, 73, 153131. 10.1016/j.tetlet.2021.153131. [DOI] [Google Scholar]
- Kim W. G.; Baek S.-y.; Jeong S. Y.; Nam D.; Jeon J. H.; Choe W.; Baik M.-H.; Hong S. Y. Chemo- and regioselective click reactions through nickel-catalyzed azide–alkyne cycloaddition. Org. Biomol. Chem. 2020, 18 (17), 3374–3381. 10.1039/D0OB00579G. [DOI] [PubMed] [Google Scholar]
- Surya Prakash Rao H.; Chakibanda G. Raney Ni catalyzed azide-alkyne cycloaddition reaction. RSC Adv. 2014, 4 (86), 46040–46048. 10.1039/C4RA07057G. [DOI] [Google Scholar]
- Choudhury P.; Chattopadhyay S.; De G.; Basu B. Ni–rGO–zeolite nanocomposite: an efficient heterogeneous catalyst for one-pot synthesis of triazoles in water. Mater. Adv. 2021, 2 (9), 3042–3050. 10.1039/D1MA00143D. [DOI] [Google Scholar]
- Boren B. C.; Narayan S.; Rasmussen L. K.; Zhang L.; Zhao H.; Lin Z.; Jia G.; Fokin V. V. Ruthenium-catalyzed azide-alkyne cycloaddition: scope and mechanism. J. Am. Chem. Soc. 2008, 130 (28), 8923–30. 10.1021/ja0749993. [DOI] [PubMed] [Google Scholar]
- Yamamoto Y.; Kinpara K.; Saigoku T.; Takagishi H.; Okuda S.; Nishiyama H.; Itoh K. Cp*RuCl-Catalyzed [2 + 2 + 2] Cycloadditions of α,ω-Diynes with Electron-Deficient Carbon–Heteroatom Multiple Bonds Leading to Heterocycles. J. Am. Chem. Soc. 2005, 127 (2), 605–613. 10.1021/ja045694g. [DOI] [PubMed] [Google Scholar]
- Ura Y.; Sato Y.; Shiotsuki M.; Kondo T.; Mitsudo T.-a. Ruthenium-catalysed synthesis of o-phthalates by highly chemoselective intermolecular [2 + 2 + 2] cycloaddition of terminal alkynes and dimethyl acetylenedicarboxylate. J. Mol. Catal. A: Chem. 2004, 209 (1), 35–39. 10.1016/j.molcata.2003.08.007. [DOI] [Google Scholar]
- Bhatt D.; Singh P. R.; Kalaramna P.; Kumar K.; Goswami A. An Atom-Economical Approach to 2-Triazolyl Thio-/Seleno Pyridines via Ruthenium-Catalyzed One-pot [3 + 2]/[2 + 2+2] Cycloadditions. Adv. Synth. Catal. 2019, 361 (23), 5483–5489. 10.1002/adsc.201900791. [DOI] [Google Scholar]
- Song W.; Li M.; Dong K.; Zheng Y. Ruthenium-Catalyzed Highly Regioselective Azide-Internal Thiocyanatoalkyne Cycloaddition under Mild Conditions: Experimental and Theoretical Studies. Adv. Synth. Catal. 2019, 361 (22), 5258–5263. 10.1002/adsc.201901014. [DOI] [Google Scholar]
- Engholm E.; Stuhr-Hansen N.; Blixt O. Facile solid-phase ruthenium assisted azide-alkyne cycloaddition (RuAAC) utilizing the Cp*RuCl(COD)-catalyst. Tetrahedron Lett. 2017, 58 (23), 2272–2275. 10.1016/j.tetlet.2017.04.095. [DOI] [Google Scholar]
- Zhang J.; Kemmink J.; Rijkers D. T.; Liskamp R. M. Synthesis of 1,5-triazole bridged vancomycin CDE-ring bicyclic mimics using RuAAC macrocyclization. Chem. Commun. 2013, 49 (40), 4498–500. 10.1039/c3cc40628h. [DOI] [PubMed] [Google Scholar]
- Johansson J. R.; Lincoln P.; Nordén B.; Kann N. Sequential One-Pot Ruthenium-Catalyzed Azide–Alkyne Cycloaddition from Primary Alkyl Halides and Sodium Azide. J. Org. Chem. 2011, 76 (7), 2355–2359. 10.1021/jo200134a. [DOI] [PubMed] [Google Scholar]
- Rasolofonjatovo E.; Theeramunkong S.; Bouriaud A.; Kolodych S.; Chaumontet M.; Taran F. Iridium-Catalyzed Cycloaddition of Azides and 1-Bromoalkynes at Room Temperature. Org. Lett. 2013, 15 (18), 4698–4701. 10.1021/ol402008u. [DOI] [PubMed] [Google Scholar]
- Genin E.; Antoniotti S.; Michelet V.; Genêt J. P. An IrI-Catalyzed exo-Selective Tandem Cycloisomerization/Hydroalkoxylation of Bis-Homopropargylic Alcohols at Room Temperature. Angew. Chem. 2005, 117 (31), 5029–5033. 10.1002/ange.200501150. [DOI] [PubMed] [Google Scholar]
- Zhang X.; Gou F.; Wang X.; Wang Y.; Ding S. Easily Functionalized and Readable Sequence-Defined Polytriazoles. ACS Macro Lett. 2021, 10 (5), 551–557. 10.1021/acsmacrolett.1c00145. [DOI] [PubMed] [Google Scholar]
- Porel M.; Alabi C. A. Sequence-Defined Polymers via Orthogonal Allyl Acrylamide Building Blocks. J. Am. Chem. Soc. 2014, 136 (38), 13162–13165. 10.1021/ja507262t. [DOI] [PubMed] [Google Scholar]
- Solleder S. C.; Meier M. A. R. Sequence Control in Polymer Chemistry through the Passerini Three-Component Reaction. Angew. Chem., Int. Ed. 2014, 53 (3), 711–714. 10.1002/anie.201308960. [DOI] [PubMed] [Google Scholar]
- Porel M.; Thornlow D. N.; Phan N. N.; Alabi C. A. Sequence-defined bioactive macrocycles via an acid-catalysed cascade reaction. Nat. Chem. 2016, 8 (6), 590–596. 10.1038/nchem.2508. [DOI] [PubMed] [Google Scholar]
- Gao W. C.; Shang Y. Z.; Chang H. H.; Li X.; Wei W. L.; Yu X. Z.; Zhou R. N-Alkynylthio Phthalimide: A Shelf-Stable Alkynylthio Transfer Reagent for the Synthesis of Alkynyl Thioethers. Org. Lett. 2019, 21 (15), 6021–6024. 10.1021/acs.orglett.9b02174. [DOI] [PubMed] [Google Scholar]
- Ding S.; Jia G.; Sun J. Iridium-Catalyzed Intermolecular Azide–Alkyne Cycloaddition of Internal Thioalkynes under Mild Conditions. Angew. Chem., Int. Ed. 2014, 53 (7), 1877–1880. 10.1002/anie.201309855. [DOI] [PubMed] [Google Scholar]
- Song W.; Zheng N. Iridium-Catalyzed Highly Regioselective Azide–Ynamide Cycloaddition to Access 5-Amido Fully Substituted 1,2,3-Triazoles under Mild, Air, Aqueous, and Bioorthogonal Conditions. Org. Lett. 2017, 19 (22), 6200–6203. 10.1021/acs.orglett.7b03123. [DOI] [PubMed] [Google Scholar]
- Chen R.; Zeng L.; Lai Z.; Cui S. Iridium-Catalyzed Hydroxyl-Enabled Cycloaddition of Azides and Alkynes. Adv. Synth. Catal. 2019, 361 (5), 989–994. 10.1002/adsc.201801410. [DOI] [Google Scholar]
- Sasano K.; Takaya J.; Iwasawa N. Palladium(II)-Catalyzed Direct Carboxylation of Alkenyl C–H Bonds with CO2. J. Am. Chem. Soc. 2013, 135 (30), 10954–10957. 10.1021/ja405503y. [DOI] [PubMed] [Google Scholar]
- Huang C.; Ghavtadze N.; Chattopadhyay B.; Gevorgyan V. Synthesis of Catechols from Phenols via Pd-Catalyzed Silanol-Directed C–H Oxygenation. J. Am. Chem. Soc. 2011, 133 (44), 17630–17633. 10.1021/ja208572v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seoane A.; Casanova N.; Quiñones N.; Mascareñas J. L.; Gulías M. Straightforward Assembly of Benzoxepines by Means of a Rhodium(III)-Catalyzed C–H Functionalization of o-Vinylphenols. J. Am. Chem. Soc. 2014, 136 (3), 834–837. 10.1021/ja410538w. [DOI] [PubMed] [Google Scholar]
- Seoane A.; Casanova N.; Quiñones N.; Mascareñas J. L.; Gulías M. Rhodium(III)-Catalyzed Dearomatizing (3 + 2) Annulation of 2-Alkenylphenols and Alkynes. J. Am. Chem. Soc. 2014, 136 (21), 7607–7610. 10.1021/ja5034952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thirunavukkarasu V. S.; Donati M.; Ackermann L. Hydroxyl-Directed Ruthenium-Catalyzed C–H Bond Functionalization: Versatile Access to Fluorescent Pyrans. Org. Lett. 2012, 14 (13), 3416–3419. 10.1021/ol301387t. [DOI] [PubMed] [Google Scholar]
- Duan X.; Zheng N.; Li M.; Sun X.; Lin Z.; Qiu P.; Song W. Remote ether groups-directed regioselective and chemoselective cycloaddition of azides and alkynes. Chin. Chem. Lett. 2021, 32 (12), 4019–4023. 10.1016/j.cclet.2021.05.037. [DOI] [Google Scholar]
- Zhang X.; Li S.; Yu W.; Xie Y.; Tung C.-H.; Xu Z. Asymmetric Azide–Alkyne Cycloaddition with Ir(I)/Squaramide Cooperative Catalysis: Atroposelective Synthesis of Axially Chiral Aryltriazoles. J. Am. Chem. Soc. 2022, 144 (14), 6200–6207. 10.1021/jacs.2c02563. [DOI] [PubMed] [Google Scholar]
- Krautwald S.; Sarlah D.; Schafroth M. A.; Carreira E. M. Enantio- and Diastereodivergent Dual Catalysis: α-Allylation of Branched Aldehydes. Science 2013, 340 (6136), 1065–1068. 10.1126/science.1237068. [DOI] [PubMed] [Google Scholar]
- Song W.; Zheng N.; Li M.; Dong K.; Li J.; Ullah K.; Zheng Y. Regiodivergent Rhodium(I)-Catalyzed Azide-Alkyne Cycloaddition (RhAAC) To Access Either Fully Substituted Sulfonyl-1,2,3-triazoles under Mild Conditions. Org. Lett. 2018, 20 (21), 6705–6709. 10.1021/acs.orglett.8b02794. [DOI] [PubMed] [Google Scholar]
- Song W.; Zheng N.; Li M.; He J.; Li J.; Dong K.; Ullah K.; Zheng Y. Rhodium(I)-Catalyzed Regioselective Azide-internal Alkynyl Trifluoromethyl Sulfide Cycloaddition and Azide-internal Thioalkyne Cycloaddition under Mild Conditions. Adv. Synth. Catal. 2019, 361 (3), 469–475. 10.1002/adsc.201801216. [DOI] [Google Scholar]
- Liao Y.; Lu Q.; Chen G.; Yu Y.; Li C.; Huang X. Rhodium-Catalyzed Azide–Alkyne Cycloaddition of Internal Ynamides: Regioselective Assembly of 5-Amino-Triazoles under Mild Conditions. ACS Catal. 2017, 7 (11), 7529–7534. 10.1021/acscatal.7b02558. [DOI] [Google Scholar]
- Smedley C. J.; Li G.; Barrow A. S.; Gialelis T. L.; Giel M.-C.; Ottonello A.; Cheng Y.; Kitamura S.; Wolan D. W.; Sharpless K. B.; Moses J. E. Diversity Oriented Clicking (DOC): Divergent Synthesis of SuFExable Pharmacophores from 2-Substituted-Alkynyl-1-Sulfonyl Fluoride (SASF) Hubs. Angew. Chem., Int. Ed. 2020, 59 (30), 12460–12469. 10.1002/anie.202003219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo W.-T.; Zhu B.-H.; Chen Y.; Yang J.; Qian P.-C.; Deng C.; Ye L.-W.; Li L. Enantioselective Rh-Catalyzed Azide-Internal-Alkyne Cycloaddition for the Construction of Axially Chiral 1,2,3-Triazoles. J. Am. Chem. Soc. 2022, 144 (15), 6981–6991. 10.1021/jacs.2c01985. [DOI] [PubMed] [Google Scholar]
- Shil A. K.; Kumar S.; Sharma S.; Chaudhary A.; Das P. Polystyrene resin supported palladium (0)(Pd@ PR) nanocomposite mediated regioselective synthesis of 4-aryl-1-alkyl/(2-haloalkyl)-1 H-1, 2, 3-triazoles and their N-vinyl triazole derivatives from terminal alkynes. RSC Adv. 2015, 5 (15), 11506–11514. 10.1039/C4RA15133J. [DOI] [Google Scholar]
- Ackermann L.; Potukuchi H. K. Regioselective syntheses of fully-substituted 1,2,3-triazoles: the CuAAC/C–H bond functionalization nexus. Org. Biomol. Chem. 2010, 8 (20), 4503–4513. 10.1039/c0ob00212g. [DOI] [PubMed] [Google Scholar]
- Rakshit M.; Kar G. K.; Chakrabarty M. A new general synthesis of annulated 1,2,3-triazoles using tandem Sonogashira-CuAAC reaction. Monatsh. Chem. 2015, 146 (10), 1681–1688. 10.1007/s00706-015-1430-y. [DOI] [Google Scholar]
- Boominathan M.; Pugazhenthiran N.; Nagaraj M.; Muthusubramanian S.; Murugesan S.; Bhuvanesh N. Nanoporous Titania-Supported Gold Nanoparticle-Catalyzed Green Synthesis of 1,2,3-Triazoles in Aqueous Medium. ACS Sustainable Chem. Eng. 2013, 1 (11), 1405–1411. 10.1021/sc400147r. [DOI] [Google Scholar]
- Kidwai M.; Bansal V.; Kumar A.; Mozumdar S. The first Au-nanoparticles catalyzed green synthesis of propargylamines via a three-component coupling reaction of aldehyde, alkyne and amine. Green Chem. 2007, 9 (7), 742–745. 10.1039/b702287e. [DOI] [Google Scholar]
- Asao N.; Nogami T.; Lee S.; Yamamoto Y. Lewis Acid-Catalyzed Benzannulation via Unprecedented [4 + 2] Cycloaddition of o-Alkynyl(oxo)benzenes and Enynals with Alkynes. J. Am. Chem. Soc. 2003, 125 (36), 10921–10925. 10.1021/ja036927r. [DOI] [PubMed] [Google Scholar]
- Asao N.; Takahashi K.; Lee S.; Kasahara T.; Yamamoto Y. AuCl3-Catalyzed Benzannulation: Synthesis of Naphthyl Ketone Derivatives from o-Alkynylbenzaldehydes with Alkynes. J. Am. Chem. Soc. 2002, 124 (43), 12650–12651. 10.1021/ja028128z. [DOI] [PubMed] [Google Scholar]
- McNulty J.; Keskar K.; Vemula R. The First Well-Defined Silver(I)-Complex-Catalyzed Cycloaddition of Azides onto Terminal Alkynes at Room Temperature. Chem. - Eur. J. 2011, 17 (52), 14727–14730. 10.1002/chem.201103244. [DOI] [PubMed] [Google Scholar]
- McNulty J.; Keskar K. Discovery of a Robust and Efficient Homogeneous Silver(I) Catalyst for the Cycloaddition of Azides onto Terminal Alkynes. Eur. J. Org. Chem. 2012, 2012 (28), 5462–5470. 10.1002/ejoc.201200930. [DOI] [Google Scholar]
- Ali A. A.; Chetia M.; Saikia B.; Saikia P. J.; Sarma D. AgN(CN)2/DIPEA/H2O-EG: a highly efficient catalytic system for synthesis of 1,4-disubstituted-1,2,3 triazoles at room temperature. Tetrahedron Lett. 2015, 56 (43), 5892–5895. 10.1016/j.tetlet.2015.09.025. [DOI] [Google Scholar]
- Smith C. D.; Greaney M. F. Zinc Mediated Azide–Alkyne Ligation to 1,5- and 1,4,5-Substituted 1,2,3-Triazoles. Org. Lett. 2013, 15 (18), 4826–4829. 10.1021/ol402225d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akula H. K.; Lakshman M. K. Synthesis of Deuterated 1,2,3-Triazoles. J. Org. Chem. 2012, 77 (20), 8896–8904. 10.1021/jo301146j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Turlington M.; Pu L. Asymmetric Alkyne Addition to Aldehydes Catalyzed by BINOL and Its Derivatives. Synlett 2012, 23 (05), 649–684. 10.1055/s-0031-1290530. [DOI] [Google Scholar]
- Clegg W.; García-Álvarez J.; García-Álvarez P.; Graham D. V.; Harrington R. W.; Hevia E.; Kennedy A. R.; Mulvey R. E.; Russo L. Synthesis, Structural Authentication, and Structurally Defined Metalation Reactions of Lithium and Sodium DA-Zincate Bases (DA = diisopropylamide) with Phenylacetylene. Organometallics 2008, 27 (11), 2654–2663. 10.1021/om8001813. [DOI] [Google Scholar]
- Meng X.; Xu X.; Gao T.; Chen B. Zn/C-Catalyzed Cycloaddition of Azides and Aryl Alkynes. Eur. J. Org. Chem. 2010, 2010 (28), 5409–5414. 10.1002/ejoc.201000610. [DOI] [Google Scholar]
- Harkala K. J.; Eppakayala L.; Maringanti T. C. Synthesis and biological evaluation of benzimidazole-linked 1,2,3-triazole congeners as agents. Org. Med. Chem. Lett. 2014, 4 (1), 14. 10.1186/s13588-014-0014-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong L.; Lin W.; Zhang F.; Liu R.; Zhou X. Ln[N(SiMe3)2]3-catalyzed cycloaddition of terminal alkynes to azides leading to 1,5-disubstituted 1,2,3-triazoles: new mechanistic features. Chem. Commun. (Cambridge, U. K.) 2013, 49 (49), 5589–91. 10.1039/c3cc42534g. [DOI] [PubMed] [Google Scholar]
- Wang J.-M.; Yu S.-B.; Li Z.-M.; Wang Q.-R.; Li Z.-T. Mechanism of Samarium-Catalyzed 1,5-Regioselective Azide–Alkyne [3 + 2]-Cycloaddition: A Quantum Mechanical Investigation. J. Phys. Chem. A 2015, 119 (8), 1359–1368. 10.1021/jp5104615. [DOI] [PubMed] [Google Scholar]
- Evans W. J.; Mueller T. J.; Ziller J. W. Reactivity of (C5Me5)3LaLx Complexes: Synthesis of a Tris(pentamethylcyclopentadienyl) Complex with Two Additional Ligands, (C5Me5)3La(NCCMe3)2. J. Am. Chem. Soc. 2009, 131 (7), 2678–2686. 10.1021/ja808763y. [DOI] [PubMed] [Google Scholar]
- Tchounwou P. B.; Yedjou C. G.; Patlolla A. K.; Sutton D. J.. Heavy Metal Toxicity and the Environment. In Molecular, Clinical and Environmental Toxicology: Vol. 3: Environmental Toxicology; Luch A., Ed.; Springer: Basel, 2012; pp 133–164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hui R.; Zhao M.; Chen M.; Ren Z.; Guan Z. One-Pot Synthesis of 4-Aryl-NH-1,2,3-Triazoles through Three-Component Reaction of Aldehydes, Nitroalkanes and NaN3. Chin. J. Chem. 2017, 35 (12), 1808–1812. 10.1002/cjoc.201700367. [DOI] [Google Scholar]
- Liu L.; Ai Y.; Li D.; Qi L.; Zhou J.; Tang Z.; Shao Z.; Liang Q.; Sun H.-B. Recyclable Acid–Base Bifunctional Core–Shell–Shell Nanosphere Catalyzed Synthesis of 5-Aryl-1H-1,2,3-triazoles through the “One-Pot” Cyclization of Aldehydes, Nitromethane, and Sodium Azide. ChemCatChem. 2017, 9 (16), 3131–3137. 10.1002/cctc.201700401. [DOI] [Google Scholar]
- Mondal D. S.Unit IV Catalysis (Transition-metal and Organo-catalysis in organic synthesis). Lecture Notes 2018, 10.13140/RG.2.2.29115.52001. [DOI] [Google Scholar]
- Belkheira M.; El Abed D.; Pons J.-M.; Bressy C. Organocatalytic Synthesis of 1,2,3-Triazoles from Unactivated Ketones and Arylazides. Chem. - Eur. J. 2011, 17 (46), 12917–12921. 10.1002/chem.201102046. [DOI] [PubMed] [Google Scholar]
- Wang L.; Peng S.; Danence L. J. T.; Gao Y.; Wang J. Amine-Catalyzed [3 + 2] Huisgen Cycloaddition Strategy for the Efficient Assembly of Highly Substituted 1,2,3-Triazoles. Chem. - Eur. J. 2012, 18 (19), 6088–6093. 10.1002/chem.201103393. [DOI] [PubMed] [Google Scholar]
- Danence L. J. T.; Gao Y.; Li M.; Huang Y.; Wang J. Organocatalytic Enamide–Azide Cycloaddition Reactions: Regiospecific Synthesis of 1,4,5-Trisubstituted-1,2,3-Triazoles. Chem. - Eur. J. 2011, 17 (13), 3584–3587. 10.1002/chem.201002775. [DOI] [PubMed] [Google Scholar]
- Ramachary D. B.; Shashank A. B. Organocatalytic Triazole Formation, Followed by Oxidative Aromatization: Regioselective Metal-Free Synthesis of Benzotriazoles. Chem. - Eur. J. 2013, 19 (39), 13175–13181. 10.1002/chem.201301412. [DOI] [PubMed] [Google Scholar]
- Vroemans R.; Horsten T.; Van Espen M.; Dehaen W. 5-Formyltriazoles as Valuable Starting Materials for Unsymmetrically Substituted Bi-1,2,3-Triazoles. Front. Chem. 2020, 8, 271. 10.3389/fchem.2020.00271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sangwan R.; Javed; Dubey A.; Mandal P. K. Organocatalytic [3 + 2] Cycloadditions: Toward Facile Synthesis of Sulfonyl-1,2,3-Triazolyl and Fully Substituted 1,2,3-Triazolyl Glycoconjugates. ChemistrySelect 2017, 2 (17), 4733–4743. 10.1002/slct.201700805. [DOI] [Google Scholar]
- Ramachary D. B.; Krishna P. M.; Gujral J.; Reddy G. S. An Organocatalytic Regiospecific Synthesis of 1,5-Disubstituted 4-Thio-1,2,3-triazoles and 1,5-Disubstituted 1,2,3-Triazoles. Chem. - Eur. J. 2015, 21 (47), 16775–16780. 10.1002/chem.201503302. [DOI] [PubMed] [Google Scholar]
- Ramachary D. B.; Ramakumar K.; Narayana V. V. Amino Acid-Catalyzed Cascade [3 + 2]-Cycloaddition/Hydrolysis Reactions Based on the Push–Pull Dienamine Platform: Synthesis of Highly Functionalized NH-1,2,3-Triazoles. Chem. - Eur. J. 2008, 14 (30), 9143–9147. 10.1002/chem.200801325. [DOI] [PubMed] [Google Scholar]
- Seus N.; Goldani B.; Lenardão E. J.; Savegnago L.; Paixão M. W.; Alves D. Organocatalytic Synthesis of (Arylselanyl)phenyl-1H-1,2,3-triazole-4-carboxamides by Cycloaddition between Azidophenyl Arylselenides and β-Oxo-amides. Eur. J. Org. Chem. 2014, 2014 (5), 1059–1065. 10.1002/ejoc.201301547. [DOI] [Google Scholar]
- Yeung D. K. J.; Gao T.; Huang J.; Sun S.; Guo H.; Wang J. Organocatalytic 1,3-dipolar cycloaddition reactions of ketones and azides with water as a solvent. Green Chem. 2013, 15 (9), 2384–2388. 10.1039/c3gc41126e. [DOI] [Google Scholar]
- Deng L.; Cao X.; Liu Y.; Wan J.-P. In-Water Synthesis of 5-Thiolated 1,2,3-Triazoles from β-Thioenaminones by Diazo Transfer Reaction. J. Org. Chem. 2019, 84 (21), 14179–14186. 10.1021/acs.joc.9b01817. [DOI] [PubMed] [Google Scholar]
- Ahamad S.; Kumar A.; Kant R.; Mohanan K. Metal-Free Three-Component Assembly of Fully Substituted 1,2,3-Triazoles. Asian J. Org. Chem. 2018, 7 (8), 1698–1703. 10.1002/ajoc.201800340. [DOI] [Google Scholar]
- Harmata M.; Huang C.; Chen Y.; Zheng P.; Gao X.; Ying W. The Synthesis of N-TBS-S-Alkynyl Sulfoximines. Synlett 2008, 2008 (13), 2051–2055. 10.1055/s-2008-1077956. [DOI] [Google Scholar]
- Yang J.; Mei F.; Fu S.; Gu Y. Facile synthesis of 1,4-diketones via three-component reactions of α-ketoaldehyde, 1,3-dicarbonyl compound, and a nucleophile in water. Green Chem. 2018, 20 (6), 1367–1374. 10.1039/C7GC03644B. [DOI] [Google Scholar]
- Wu C.; Lu L.-H.; Peng A.-Z.; Jia G.-K.; Peng C.; Cao Z.; Tang Z.; He W.-M.; Xu X. Ultrasound-promoted Brønsted acid ionic liquid-catalyzed hydrothiocyanation of activated alkynes under minimal solvent conditions. Green Chem. 2018, 20 (16), 3683–3688. 10.1039/C8GC00491A. [DOI] [Google Scholar]
- Guo Y.; Wang G.; Wei L.; Wan J.-P. Domino C-H Sulfonylation and Pyrazole Annulation for Fully Substituted Pyrazole Synthesis in Water Using Hydrophilic Enaminones. J. Org. Chem. 2019, 84 (5), 2984–2990. 10.1021/acs.joc.8b02897. [DOI] [PubMed] [Google Scholar]
- Liu L.; Wang Z. Metal-free intramolecular amino-acyloxylation of 2-aminostyrene with carboxylic acid for the synthesis of 3-acyloxyl indolines in water. Green Chem. 2017, 19 (9), 2076–2079. 10.1039/C7GC00212B. [DOI] [Google Scholar]
- Pokhodylo N. T.; Matiychuk V. S.; Obushak M. D. (Arylsulfonyl)acetones and -acetonitriles: New Activated Methylenic Building Blocks for Synthesis of 1,2,3-Triazoles. Synthesis 2009, 2009 (14), 2321–2323. 10.1055/s-0029-1216850. [DOI] [Google Scholar]
- Davies H. M. L.; Cantrell W. R. Jr; Romines K. R.; Baum J. S. Synthesis of Furans via Rhodium(II) Acetate-Catalyzed Reaction of Acetylenes with α-Diazocarbonyls: Ethyl 2-Methyl-5-Phenyl-3-Furancarboxylate. Org. Synth. 2003, 93. 10.1002/0471264180.os070.11. [DOI] [Google Scholar]
- Hughes C. C.; Kennedy-Smith J. J.; Trauner D. Synthetic Studies toward the Guanacastepenes. Org. Lett. 2003, 5 (22), 4113–4115. 10.1021/ol035559t. [DOI] [PubMed] [Google Scholar]
- Gao D.; Zhai H.; Parvez M.; Back T. G. 1,3-Dipolar Cycloadditions of Acetylenic Sulfones in Solution and on Solid Supports. J. Org. Chem. 2008, 73 (20), 8057–8068. 10.1021/jo801621d. [DOI] [PubMed] [Google Scholar]
- Daha F.; Matloubi H.; Tabatabi S.; Shafii B.; Shafiee A. Syntheses of 1-(4-methylsulfonylphenyl)-5-aryl-1,2,3-triazoles and -(4-aminosulfonylphenyl)-5-aryl-1,2,3-triazoles. J. Heterocycl. Chem. 2005, 42, 33–37. 10.1002/jhet.5570420106. [DOI] [Google Scholar]
- Kadaba P. K. Triazolines—II: Solvent effects on the 1,3-cycloaddition of diazomethane to schiff bases and the synthesis of 1,5-diaryl-1,2,3-triazolines. Tetrahedron 1966, 22 (8), 2453–2460. 10.1016/S0040-4020(01)99034-9. [DOI] [Google Scholar]
- Buckley G. D. Reaction of diazomethane with arylideneanilines. J. Chem. Soc. 1954, 0, 1850–1851. 10.1039/jr9540001850. [DOI] [Google Scholar]
- Kadaba P. K. Role of Protic and Dipolar Aprotic Solvents in Heterocyclic Syntheses via 1,3-Dipolar Cycloaddition Reactions. Synthesis 1973, 1973 (02), 71–84. 10.1055/s-1973-22136. [DOI] [Google Scholar]
- Fehlhammer W. P.; Beck W. Azide Chemistry – An Inorganic Perspective, Part II[‡] [3 + 2]-Cycloaddition Reactions of Metal Azides and Related Systems. Z. Anorg. Allg. Chem. 2015, 641 (10), 1599–1678. 10.1002/zaac.201500165. [DOI] [Google Scholar]
- Kadaba P. Triazoline IX. A new ring transformation reaction of a 4,5-acylsubstituted-1,2,3-triazoline. a new route to synthesis of enamine and pyrrolidine ketones. J. Heterocycl. Chem. 1976, 13 (5), 1153–1154. 10.1002/jhet.5570130551. [DOI] [Google Scholar]
- Kwok S. W.; Fotsing J. R.; Fraser R. J.; Rodionov V. O.; Fokin V. V. Transition-Metal-Free Catalytic Synthesis of 1,5-Diaryl-1,2,3-triazoles. Org. Lett. 2010, 12 (19), 4217–4219. 10.1021/ol101568d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chrisment J.; Delpuech J.-J. Thermodynamic and kinetic acidities in dimethyl sulfoxide. Part 2. Acetylenic compounds. J. Chem. Soc., Perkin Trans. 2 1977, 4, 407–412. 10.1039/p29770000407. [DOI] [Google Scholar]
- Wan J.-P.; Cao S.; Liu Y. A metal-and azide-free multicomponent assembly toward regioselective construction of 1, 5-disubstituted 1, 2, 3-triazoles. J. Org. Chem. 2015, 80 (18), 9028–9033. 10.1021/acs.joc.5b01121. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Zhou R.; Wan J.-P. Water-Promoted Synthesis of Enaminones: Mechanism Investigation and Application in Multicomponent Reactions. Synth. Commun. 2013, 43 (18), 2475–2483. 10.1080/00397911.2012.715712. [DOI] [Google Scholar]
- Agard N. J.; Prescher J. A.; Bertozzi C. R. A Strain-Promoted [3 + 2] Azide–Alkyne Cycloaddition for Covalent Modification of Biomolecules in Living Systems. J. Am. Chem. Soc. 2004, 126 (46), 15046–15047. 10.1021/ja044996f. [DOI] [PubMed] [Google Scholar]
- Agard N. J.; Baskin J. M.; Prescher J. A.; Lo A.; Bertozzi C. R. A Comparative Study of Bioorthogonal Reactions with Azides. ACS Chem. Biol. 2006, 1 (10), 644–648. 10.1021/cb6003228. [DOI] [PubMed] [Google Scholar]
- Del Grosso A.; Galanopoulos L.-D.; Chiu C. K. C.; Clarkson G. J.; O’Connor P. B.; Wills M. Strained alkynes derived from 2,2′-dihydroxy-1,1′-biaryls; synthesis and copper-free cycloaddition with azides. Org. Biomol. Chem. 2017, 15 (21), 4517–4521. 10.1039/C7OB00991G. [DOI] [PubMed] [Google Scholar]
- Koch-Pomeranz U.; Hansen H.-J.; Schmid H. Die durch Silberionen katalysierte Umlagerung von Propargyl-phenyläther. Helv. Chim. Acta 1973, 56 (8), 2981–3004. 10.1002/hlca.19730560838. [DOI] [Google Scholar]
- Kii I.; Shiraishi A.; Hiramatsu T.; Matsushita T.; Uekusa H.; Yoshida S.; Yamamoto M.; Kudo A.; Hagiwara M.; Hosoya T. Strain-promoted double-click reaction for chemical modification of azido-biomolecules. Org. Biomol. Chem. 2010, 8 (18), 4051–4055. 10.1039/c0ob00003e. [DOI] [PubMed] [Google Scholar]
- Gouin S. G.; Kovensky J. A Procedure for Fast and Regioselective Copper-Free Click Chemistry at Room Temperature with p-Toluenesulfonyl Alkyne. Synlett 2009, 2009 (09), 1409–1412. 10.1055/s-0029-1216744. [DOI] [Google Scholar]
- Roshandel S.; Suri S. C.; Marcischak J. C.; Rasul G.; Surya Prakash G. K. Catalyst and solvent free microwave-assisted synthesis of substituted 1,2,3-triazoles. Green Chem. 2018, 20 (16), 3700–3704. 10.1039/C8GC01516C. [DOI] [Google Scholar]




















































