Advancements in transition metal-catalyzed 1,2,3-triazole synthesis via azide–alkyne cycloaddition
-
Arif Ullah
, Abdul Wahab
, Somnath Mukherjee
Abstract
1,2,3-triazoles have emerged as important structural motifs in chemical biology, attracting increasing focus in recent years because of their wide range of uses. These compounds can be efficiently synthesized using click reactions. Its versatility makes it valuable in drug discovery and materials science. Significant advancements have been made in recent years in the process of making 1,2,3-triazoles, reflecting the growing curiosity and investment in the area of research. The catalytic performance of transition metals Cu, Ir, Rh, Ru, Ni, Pd, Au, Ag, and Zn, which were used as ligands and salts in the azide–alkyne cycloaddition method of 1,2,3-triazole synthesis, has been covered in this review. Cu-complexes and salts were found to be more effective for selective synthesis among all transition metals. Furthermore, it is determined that some azide–alkyne reactions are entirely catalytic in nature and cannot be carried out as such by switching transition metals.
1 Introduction
1,2,3-triazole is an important structure synthesis that has undergone a profound transformation, 1 , 2 , 3 , 4 , 5 , 6 principally motivated by the CuAAC’s adaptability, or cycle of azide–alkyne addition, reaction. 7 , 8 , 9 This synthetic route or transition-metal-catalyzed cross-coupling reactions has become essential to the synthesis of these heterocyclic molecules, owing to its wide substrate applicability, regioselectivity, and broad utility across multidisciplinary scientific realms, including materials science, chemical biology, and medicinal chemistry. Over the past several years, a concerted research effort has been dedicated to advancing transition metal-catalyzed CuAAC reactions, with a specific focus on expanding the repertoire of catalytic systems, refining reaction conditions, and integrating sustainable practices. 10 , 11 , 12 , 13 , 14 The exploration of catalytic systems beyond the traditional copper-based catalysts has been a transformative endeavor in recent research, encompassing the utilization of ruthenium, palladium, and iron. 15 , 16 , 17 This catalytic chemical process not only broadens the scope of applicable substrates but also offers the promise of milder reaction conditions, thus enhancing the overall effectiveness and usefulness of the synthetic procedure. The strategic design of ligands and the optimization of reaction parameters have been crucial in achieving not only heightened yields but also improved regioselectivity in diverse CuAAC transformations. 18 , 19 , 20
Prominent initiatives include the exploration of alternative reaction media, such as water, and the design of recyclable catalytic systems. These strides toward sustainability not only address the imperative of reducing environmental impact but also respond to the global call for more resource-efficient and ecologically mindful chemical processes. 20 , 21 , 22 Furthermore, the significant applications of CuAAC in the synthesis of bioactive molecules, pharmaceuticals, and advanced materials underscores the method’s versatility and its indispensable role in contemporary research. 7 , 23 , 24 This study also examines the synergistic integration of computational methods, providing a predictive dimension to the field by offering insights into reactivity patterns and elucidating complex reaction mechanisms. The marriage of experimental and computational approaches facilitates a deeper understanding of the CuAAC process, guiding the rational design of catalysts and the fine-tuning of reaction conditions for heightened precision and efficiency.
In organic catalytic synthesis, transition metals like palladium, platinum, and nickel take part through their partially filled d-orbitals in the preparation of products. These orbitals enable the metals to form complex interactions with organic molecules, stabilizing reaction intermediates and lowering activation energies, which significantly accelerate reaction rates. In processes such as hydrogenation and cross-coupling reactions (e.g., Suzuki and Heck), the d-orbitals facilitate the formation and breaking of bonds by alternating between oxidation states, allowing smooth electron transfer. By stabilizing transition states, these d-orbitals reduce the energy required for chemical transformations, making transition metals indispensable in catalytic reactions, including industrial applications like the Haber process for nitrogen fixation. 25
Here, we examine the developments in the synthesis of 1,2,3-triazoles using catalysts made of transition metals. Our aim extends beyond mere cataloging of recent findings; we aspire to offer an accessible and comprehensive resource for scientists at all levels, from seasoned researchers to those just entering the field. Through insightful analyses of the current state of the discipline and by charting a course for future investigations, this review contributes to the ongoing narrative of synthetic chemistry’s evolution (Figure 1).

A few significant 1, 2, and 3 triazoles that transition metals catalyze.
2 The synthesis of significant 1,2,3-triazoles is accomplished through the cycloaddition of azides and alkynes.
Researching physiologically active heterocyclic molecules with nitrogen atoms is one of the most significant areas of medical chemistry. 26 , 27 Azole is a nitrogen-containing heterocycle with five members. It is a structural component of several naturally occurring compounds that are biologically active. 28 A range of biological activities, including antibacterial activity, are displayed by azoles and their derivatives. 29 With several uses, the 1,2,3-triazole scaffold is a useful scaffold for heterocycles containing nitrogen. 30 The aromatic feature of 1,2,3-triazole comes from its 6π delocalized electron ring structure, which is not soaked, π-excess heterocycle with five members. Two carbons and three nitrogens’ make up 1,2,3-triazoles. Every atom in the five is sp2 hybridized. N atoms come in two varieties: pyridine and pyrrole. Monocyclic 1,2,3-triazoles are further subdivided into three subclasses according to the placement of the NH proton. 1H- and 2H-1,2,3-triazoles are aromatic and equilibrate each other in solution and gas phase, however, 4H-1,2,3-triazole is non-aromatic by nature 31 among the monocyclic 1,2,3-triazoles in all possible isomers, 1H-1,2,3-triazoles are powerful scaffolds that are widely found in therapeutic, agrochemicals, and materials science have received more attention. 31 , 32 , 33
Triazole rings may function as crucial transitions in a variety of significant organic reactions. 34 , 35 These five-membered N-heterocycles’ many uses stem from their simple synthesis and ring functionalization processes. 36 , 37 , 38 Significant emphasis has also been given in the last few years to the identification of synthetic procedures centered upon ecological sustainability. Numerous strategies in this respect make use of cutting-edge Lewis catalysts, unusual solvents, and eco-friendly components. 39 , 40 Much emphasis has been paid to 1,3-dipolar cycloaddition (1,3-DPCA), the most important step in the synthesis of heterocycles. 41 This method has proven to be especially helpful in adding chemical features of a species, like nitrone, which has been demonstrated to function as a radical scavenger and is an essential property of medications, that the relevant precursors usually already carry (i.e., 1,3-dipoles). 42 They serve as helpful dipole partners in the synthesis of new bioactive substances. 42 Moreover, it has been demonstrated that the 1,3-DPCA reaction is especially compatible with the application of cutting-edge green chemistry methods, like microwave-assisted reactions, 43 , 44 Ultrasound-assisted responses 44 as well as non-traditional reaction solvents. 45
The steric and electronic properties of the substrate significantly influence the regioselectivity of this method. However, due to the elevated reaction temperatures, low yields, and the formation of two regioisomers when using unsymmetrical alkynes, this technique has seen limited application in synthetic chemistry. In response, several strategies have been developed to improve control over regioselectivity. Among these, ruthenium catalysts were found to selectively produce 1,5-disubstituted compounds, thereby they enhance significance of this method (Scheme 1). 46 , 47

Synthesis of 3 and 4 using Huisgen 1,3-dipolar cycloaddition method.

Cu-catalyzed reaction of 7.
3 Copper catalyzes the azide–alkyne cycloaddition (CuAAC)
In 2002, Sharpless and Folkin, 48 along with Meldal and colleagues, independently discovered the CuAAC reaction, where alkyne 6 and azide 5 serve as primary constituents. In organic synthesis, the CuAAC reaction stands out, although recent click chemistry approaches have been reported, none fully meet the criteria set by Sharpless and colleagues for the ideal click response. The copper-catalyzed reaction between alkynes and organic azides exemplifies click chemistry (Scheme 2). 49
In the 1970s, Huisgen et al. reported in the 1,3-dipolar cycloaddition procedure, organic azides, and alkynes as well as the successfully generated 1,2,3-triazole compounds. 50 The Sharpless and Fokin group reported a Cu(I)-catalyzed reaction in 2002. [3 + 2] benzylic azides and alkynes undergo cycloaddition to produce 1,2,3-triazoles. 48 (Scheme 3). Using this method, CuSO4(II) was reduced to Cu(I), which shows significant regioselectivity in a 2:1 mixture of water and tert-butanol. Finally, rather than 1,5-disubstituted triazole compound 11, a series of 1,4-disubstituted 1,2,3-triazole were generated. The same Sharpless and Fokin group reported the cycloaddition reaction of organic azides and 1-iodoalkynes catalyzed by Cu(I) in 2009 20 (Scheme 3). CuI catalyzed the reaction between organic azide and iodoalkyne in the presence of an amine ligand (TTTA) (14) to successfully synthesize 5-iodo-1,4,5-trisubstituted-1,2,3-triazole (15), use a cycloaddition procedure. Furthermore, the pace and chemical selectivity of the reaction is mostly determined by the presence of amine ligands.

3 + 2 cycloaddition process catalyzed by Cu(I).
The trio of components 51 cascade response, which is catalyzed by copper, is the sophisticated synthetic strategy employed by the authors to create nitrogen poly heterocycles. They integrate direct C–H arylation, azide–acetylene cycloaddition, and Ullmann-type coupling reactions into a single-pot process. The main breakthrough is the ability to quickly assemble intricate polyheterocyclic systems by using 2-azidoacetamides as flexible building blocks. With this method, the versatility of copper catalysis in forming multiple bonds and rings within a single cascade reaction is demonstrated, highlighting its potential to coordinate a variety of transformations. The study underscores the efficacy and synthetic utility of the proposed methodology through a comprehensive evaluation of its diverse applications (Scheme 4).

Nitrogen poly heterocycles can be synthesized in three steps using copper as a catalyst via cyclization, C–N coupling, domino cycloaddition, or (C–H arylation).
Using 1-azido-2-isocyanoarenes 52 in combination with terminal acetylenes or substituted acetaldehydes, a highly effective procedure has been developed for synthesizing 1,2,3-triazolo[1,5-a] scaffolds made of quinoxaline. The four chemical linkages are made up of two C–C and two C–N bonds, which can be formed sequentially without isolating the triazole intermediate, thanks to the trifluoromethylation-triggered cyclization process. Moreover, rhodium-catalyzed carbenoid insertion reactions can easily convert these triazo-fused products into a variety of quinoxaline derivatives (Scheme 5).
![Scheme 5:
With good bond-forming efficiency, 1-azido-2-isocyanoarenes are synthesized in a single pot to yield triazolo[1,5-a] quinoxalines [1,2,3].](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_005.jpg)
With good bond-forming efficiency, 1-azido-2-isocyanoarenes are synthesized in a single pot to yield triazolo[1,5-a] quinoxalines [1,2,3].
This work establishes a 53 simultaneous response involving intramolecular direct arylation, Ullmann-type C–N coupling and Azide–alkyne cycloaddition (CuAAC), that is catalyzed by an unbound ligand and copper. The developed approach makes it possible to synthesize a novel azaheterocycle framework with triazole fused in one step with efficiency. A noteworthy outcome of the reaction is the production of Triazole-1,2,3-fused imidazole [1,2-a] pyridines, with favorable yields between 59 % and 77 %. This method uses a series of copper-catalyzed transformations to create a logical and efficient pathway to complex azaheterocyclic structures. The copper catalyst’s ligand-free state makes the synthetic protocol more straightforward and effective (Scheme 6).

Intramolecular direct arylation, tandem azide–alkyne cycloaddition catalyzed by copper and ullmann type C–N coupling.
To produce 54 the medicinal significant Tetrahydro-[1,2,3] hexagonal triazole pyrazine pattern, a tandem synthetic approach has been developed in this study. The technique is applied in mixed aqueous–organic media and involves a concise tandem sequence that combines a “click” reaction with a 6-exo-dig cyclization. This approach shows its adaptability by taking into account a range of substrates that are based on amino acids. The pharmaceutical significance of synthesizing the tetrahydro-[1,2,3]triazolopyrazine scaffold makes it an especially significant process. For the tandem process, the mixed aqueous–organic conditions offer a comfortable and effective environment. The paper provides a thorough discussion of the method’s applicability and limitations, providing insights into the range of substrates that can be used and potential difficulties that may arise (Scheme 7).
![Scheme 7:
A highly effective method for creating triazole fused pyrazines is the tandem cycloaddition/6-exo-cyclization [3 + 2] “click”.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_007.jpg)
A highly effective method for creating triazole fused pyrazines is the tandem cycloaddition/6-exo-cyclization [3 + 2] “click”.
In this study, overlooked 55 potential cyclic diaryl iodoniums, in contrast to their widely utilized linear counterparts, have been explored for their ability to start two arylations, offering additional atoms and a step-by-step synthesis pathway. The researchers successfully conducted a cascade reaction with three components that uses cyclic diaryliodoniums, Alkynes, and sodium azide in mild circumstances, employing cost-effective copper species as catalysts. With unsymmetrical iodonium in particular, the regioselectivity of the reaction was increased by carefully adding two methyl groups, ortho, and para, to the iodine core. This reaction produces a wide variety of intricate compounds quickly, specifically triazole phenanthridine derivatives. This work highlights the efficiency and versatility of cyclic diaryl iodoniums in accessing intricate molecular architectures through a streamlined cascade process (Scheme 8).

A Cu(I)-catalyzed cascade reaction with calciodoniums, alkynes, and sodium azide efficiently produces triazolophenanthridines.
4 Cycloaddition of azide–alkyne catalyzed by Nikal
Apart from Cu, other metals can also be used to catalyze the reaction of 1,3-dipolar cycloaddition between alkynes and organic azides, such as Ru, Rh, Ni, and Pd. 56 , 57 , 58 , 59 , 60 , 61 The E-NiAAC reaction’s azide scope revealed that while electron-rich and electron-poor groups do not significantly impact the scope of allyl azides, acyclic allyl azide can change the cyclohexyl moiety, resulting in the observation of (E) – and (Z) – olefin isomers (7:1 E/Z). But just one isomer of the intended derivative was isolated. Hong and colleagues 62 revealed that 38 and 39 could be easily and efficiently cycloaddition by nickel utilizing Cp2Ni/xanthous as a catalyst in either water or an organic solvent (toluene). With a wide range of substrates and strong regioselectivity, product 40 was produced in 42–95 % yield (Scheme 9b).

Nickel-catalyzed NiAAC, or azide–alkyne cycloaddition.
Basu et al. developed a one-pot 63 method (working at 90 °C for four to 6 h) to synthesis 1,4-disubstituted triazoles 45 with high regioselectivity from halides 44, NaN3 43, and 39. Ni-rGO-zeolite, a stable ternary nanocomposite catalyst, was present to aid in this reaction (Scheme 9d). To create effective nanocatalysts for this procedure, they combined 2D GO (graphene oxide) with NaY zeolites to create hybrid GO zeolites. The GO-zeolite composite is made by mixing an aqueous suspension of NaY zeolite with protons to acidify the mixture’s pH to roughly 7. By lowering the negative charges on the zeolite composite corresponding to the transition point, this process aids in the formation of Al–OH–Si bridges. After giving the mixture a very gentle stir, it is heated for about 16 h at 60 °C. The water is then evaporated, and the entire mass is emptied in order to dry it out. After hydrothermally reducing this GO/zeolite composite with Ni(OAc)2·4H2O and NaBH4, ternary Ni-rGO-zeolite nanocomposite is produced. The catalyst becomes more active during the final step.
To obtain disubstituted 1,4- or 1,5-azide–alkyne 64 cycloadditions with transition metal catalysis or those without can be used to 1,2,3-triazoles. In aqueous and ambient conditions, azaide-alkyne cycloaddition fails to provide entry to the complementary 1,5-regioisomers due to its narrow substrate scope or the use of catalysts that are insensitive to moisture or air. Nevertheless, copper-catalyzed cycloaddition is effective in obtaining 1,4-disubstituted products in biomolecular reaction systems. We present a catalytic system employing Xantphos and Cp2Ni for synthesizing 1,5-disubstituted 1,2,3-triazoles. The reaction occurs in both water and organic solvents at room temperature. This method is straightforward and scalable, accommodating a wide range of substrates, including aromatic and aliphatic compounds. Moreover, it enables the preparation of triazoles linked to carbohydrates or amino acids via cycloaddition (Scheme 10).

To get access to air and water-based 1,5-disubstituted 1,2,3-triazoles, azide–alkyne cycloaddition catalyzed by nickel.
Production of 1,5-disubstituted 65 1,2,3-triazoles can be accomplished efficiently in a single pot using an effective method that combines the conversion of bromide into azide and NiAAC. The use of isolated azides, which are not easily found on the market and may be unstable and challenging to handle, is prohibited by this method. Alcohol, carbamate, and ester are just a few of the functional moieties that this one-pot method can tolerate. Utilizing functionalized aryl and alkyl alkynes along with bromides, a range of 1,5-disubstituted 1,2,3-triazoles can be synthesized with moderate to good yields and regioselectivities. This approach facilitates the generation of diverse libraries of functionalizable 1,5-disubstituted 1,2,3-triazoles, offering valuable prospects for medicinal chemistry and chemical biology applications, among others (Scheme 11).

A multi-component reaction that generates azides in situ and produces 1,5-disubstituted 1,2,3-triazoles through azide–alkyne cycloaddition mediated by nickel.
Heterocyclic functionalized frameworks 66 can be synthesized quickly via cycloaddition catalyzed by metals. Conducting azide–alkyne cycloadditions involving competing internal alkynes under reactivity-controlled and metal-catalyzed conditions has posed significant challenges. Here, we introduce a [3 + 2] cycloaddition approach using asymmetrical alkynes and organic azides. Employing nickel as a catalyst leads to the formation of functionalized 1,2,3-triazoles with enhanced control over chemoselectivity and regioselectivity. Specifically, cyano alkynes bearing a 4-cyano substituent and terminal alkynes yield 1,4,5-trisubstituted and 1,5-disubstituted triazoles, respectively. Inverse regioselectivity is seen when the alkynes and ynamides are compared to terminal alkynes and cyano alkynes, respectively. As a result, 1,4,5-trisubstituted triazoles with substituents of 5 amides and 5 thiols are produced. Through calculations using density functional theory, the reaction mechanism was made clear. The alkyne substrate is added oxidatively to the Ni(0)-Xantphos catalyst in order to form the nickel lacyclopropene intermediate, as per the suggested mechanism (Scheme 12).

Click reactions with chemo- and regioselectivity via nickel-catalyzed azide–alkyne cycloaddition.
5 Azide–alkyne cycloaddition catalyzed by iridium (IrAAc)
Regarding the complete synthesis 67 in place of 5-aminotriazole 58, Song and associates devised a highly selective, mild, and bioorthogonal approach (Scheme 14). For this technique, cyclic ynamides as well as acyclic ynamides work well; cyclic ynamides yields that are marginally higher than those of electron-poor groupings. If the nearby position of ynamide were to be replaced with a bulky group, the reaction’s yield would be decreased (−R2) (Scheme 13).

Azide–alkyne cycloaddition catalyzed by iridium (IrAAC).
A technique of obtaining 68 5-amido completely replaced highly selective 1,2,3-triazoles is described. It involves the gentle, airy, watery, and bioorthogonal azide-ynamide cycloaddition process, which is catalyzed by iridium. Ynamide’s strong regioselectivities may be related to the strong interaction between π-acidic iridium and carbonyl oxygen. In chemical and biological contexts, the iridium ion can efficiently catalyze this reaction due to its low cytotoxicity and insensitivity to oxygen and water. The gram-scale production of carbohydrates using this method (Scheme 14).

Under mild circumstances, iridium catalyzes the complete substitution of 5-amido in 1,2,3-triazoles via azide-ynamide cycloaddition.
It has been possible to create 69 a novel cycloaddition of azides and alkynes with hydroxyl support, employing iridium as the catalyst. The 2-alkynyl phenols would rapidly access a range of triazoles through a regioselective cycloaddition with distinct azides when 2 mol% of [Ir(cod)Cl]2 is present. This process is aided by the hydroxyl group, which acts as a directing group. The process yields succinctness, diversity, and mild, biocompatible reaction conditions. Moreover, the likelihood of the products undergoing late-stage divergent transformations is increased by the presence of the hydroxyl group (Scheme 15).

Iridium-catalyzed hydroxyl-enabled cycloaddition of azides and alkynes.
6 Ruthenium-catalyzed azide–alkyne cycloaddition
The first one was done by Jia and co-workers in 2005 where they described the selective fashion manner preparation of terminal and internal alkynes and the second was Ru(II) catalyzed transformation of 1,4,5-trisubstituted and 1,5-disubstituted organic azides. 70 These reactions coupled with CuAAC make it possible to obtain both regioisomers of 1,2,3-triazole which is a heterocycle that has been recently reviewed in medicinal chemistry/organic synthesis and material science. 71 These reactions coupled with CuAAC make it possible to obtain both regioisomers of 1,2,3-triazole which is a heterocycle that has been recently reviewed in medicinal chemistry/organic synthesis and material science. Sometime later in 2007, Fokin created a new catalytic system for the Ru(II) mediated cycloaddition of azides and alkynes; the presented process is useful for the formation of 1-arylated 1,2,3-triazoles from aryl azides. Regarding the substrate scope much freedom is possible in this reaction as both aryl and aliphatic azides are compatible with alkynes. When the described reaction is conducted under the microwave condition, it takes only 20 min, and the yield of new 1,2,3-triazole compounds is much higher (Scheme 16).

Azide–alkyne cycloaddition catalyzed by ruthenium (RuAAC).
This reaction occurs efficiently under mild conditions and typically occurs at room temperature. 71 Through the reduction of activation energy, the catalyst facilitates the formation of products. DMF ensures the good solubility of the reactants by acting as a suitable solvent. Microwave irradiation further improves the reaction by raising the yield of the product and accelerating it. This is due to the fact that rapid and uniform heating minimizes adverse effects. Taking everything into account, this method offers a useful way to synthesize 1,5-disubstituted 1,2,3-triazoles, with the reaction products being maximized by microwave irradiation (Scheme 17).

Cycloaddition of alkynes and aryl azides is catalyzed by ruthenium.
Starting with Ru(PPh3)3Cl2 complex, 72 a complex of [Ru(dppp)2(CH3CN)Cl][BPh4] can thus be obtained. According to structural analysis, this complex primarily uses silver salts to act as a catalyst in the homocoupling of alkynes. Furthermore, it mediates the subsequent alkyne-azide click reaction as well as the alkyne homocoupling. The structural details of two catalytically active species, a new complex [Ru(dip)2Cl][BPh4] (2) and an intermediate complex [Ru(dip)2(CCPh)2] (6), have been obtained through synthesis and characterization (Scheme 18).
![Scheme 18:
[Ru(dip)₂Cl][BPh₄] and [Ru(dip)₂(CH₃CN)Cl][BPh₄] are synthesized through ruthenium-catalyzed alkyne reactions.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_018.jpg)
[Ru(dip)₂Cl][BPh₄] and [Ru(dip)₂(CH₃CN)Cl][BPh₄] are synthesized through ruthenium-catalyzed alkyne reactions.
Huisgen cycloaddition, a regioselective 73 intramolecular process, generated macrocyclic triazole rings from several azido alkyne substrates. With the aid of catalyst control, two structurally different macrocycles with a 1,5- or 1,4-triazole and a n or n + 1 ring size have been produced from a common intermediate. This is the first instance of ruthenium catalyzing an intramolecular Huisgen cycloaddition (Scheme 19).

Using catalyst control to access skeletal diversity: macrocyclic triazole ring.
Formation for Numbers n and n + 1.The Huisgen [3 + 2] cycloaddition, 74 an important synthetic process that produces 1-protected, Azide–ynamide cycloaddition process catalyzed by ruthenium is known as 5-amido 1,2,3-triazoles. During this reaction, ruthenium acts as a catalyst to combine ynamides – which are compounds with an amide group and a triple bond – with azides. First, by coordination, the catalyst activates the azide and the ynamide. A five-membered ring intermediate is formed when the azide nitrogen atom nucleophilically attacks the activated ynamide. After that, this intermediate goes through tautomerization and rearrangement, which causes the intended 1,2,3-triazole product to develop. A common modification to the triazole product is 1 protection, which is achieved by employing different protecting groups such as tert-butyl or benzyl to shield the nitrogen atom at position 1 (Scheme 20).
![Scheme 20:
Ruthenium was used for [3 + 2] catalysis in order to protect 1, protected 5-amido 1,2,3-triazoles.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_020.jpg)
Ruthenium was used for [3 + 2] catalysis in order to protect 1, protected 5-amido 1,2,3-triazoles.
The ease of preparation of the title compounds by the catalytic action of Ruthenium(II) azide (cyclooctadiene) chloride and 75 zynamide cyclopentadienyl involving 4-haloisoxazoles is achievable by the reaction of electron-deficient 1-chloro-, 1-bromo-, and 1-iodoalkynes with nitrile oxides. Organic azides are good 1,3-dipoles that create 5-halo-1,2,3-triazoles using air-stable transformations under the direction of alkyne-substrate 1. Twenty five equivalents of the corresponding 1,3-dipole compared to the alkyne partner at 30 °C. Reactive 1-haloalkynes are propionic amides and ketones, esters and phosphonates. More so, the halogenated azole products can be further subjected to palladium catalyzed cross coupling reactions and reactive amide group transformations. Given that [CpRuCl(cod)] has no catalytic activity, the steric hindrance of the Cp(η5-C5Me5) fragment in this complex is proposed to be greater than that which is found with the Cp (η5-C5H5) moiety (Scheme 21).

2. Cycloaddition catalyzes haloalkynes to generate nitrile oxides and organic azides halotriazoles, fourth and fifth.
7 Azide–alkyne cycloaddition catalyzed by rhodium (RhAAC)
Li and colleagues 76 described the first open flask technique utilizing rhodium(I) catalyst to synthesize 5-aminotriazoles from lactams 85 and 86 at room temperature. The catalyst used was [Rh(CO)2Cl]2, and the solvent used was MeCN (Scheme 22). They further state that since the reaction was still as effective in the open flask, there was no need to exclude air or moisture. Numerous solvents could be used with this technique without significantly altering the reaction yield. Different protective group ynamides did not affect the reaction outcomes. It’s interesting to note that ynamides and noisy groups had a successful reaction that produced an isolated yield of 84 % triazole. Most likely, the strong electron-withdrawing group characteristics of nasal acyl groups are the cause of the extended reaction time. The reactions to para-methyl, methoxy, trifluoromethyl, and chloroynamides produce the necessary triazoles in excellent yields in contrast to ordinary ynamides, regardless of their electronic nature. Additionally accepted were ortho and meta-alterations on the benzene ring. It was discovered that alkyne amides or ynamides with extra alkenyl moieties or heteroaryl rings are suitable RhAAC reaction substrates (Scheme 22).

Azide alkyne cycloaddition catalyzed by rhodium (RhAAC).
8 Cycloaddition of azide and alkyne catalyzed by palladium
A two-step Sonogashira coupling-CuAAC reaction was developed by Kar et al. to produce cycloolefins and 1,2,3-triazolopolyhydroaromatics. 77 The catalyst for the reaction was Pd(PPh3)4, the base was Et3N, and the co-catalyst was CuI. The starting materials for the reaction were 2- or 1-azidomethyl-1- or 2-bromodihydronaphthalenes, arenes, or cycloalkenes. The desired compound was obtained as the product of this process, which also aided in intramolecular hetero-annulation. The method yielded moderate to good amounts, up to 59 % yields when the reaction was conducted at 60 °C in a N2 atmosphere with DMF solvent, and an 8–10 h reaction time (Scheme 23).

Azide–alkyne cycloaddition catalyzed by palladium (PdAAC).
A sophisticated synthetic technique 78 involving palladium–copper catalyzed reactions in a single pot has yielded a mixture of 1,2,3-triazoles fused with benzoheterocycles of varying ring sizes, including five, six, seven, and eight-membered rings such as isoindoline, tetrahydroisoquinoline, benzoazepine, and benzoazocine. This method showcases the synthesis of biologically significant uracil derivatives, highlights the impact of bis-hetero-annulations, and demonstrates the versatility of utilizing acetylene gas as a cost-effective substrate. By employing readily available substrates, these reactions offer a broad range of applications and are amenable to experimental manipulation (Scheme 24).
![Scheme 24:
Benzoheterocycles are efficiently synthesized [1, 2, 3] through Pd–Cu catalyzed heterocyclization.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_024.jpg)
Benzoheterocycles are efficiently synthesized [1, 2, 3] through Pd–Cu catalyzed heterocyclization.
Here, we present an efficient one-pot synthesis 79 method for producing 2-(2-bromophenyl)imidazo[1,2-a]pyridines, incorporating sodium azide and alkynes to yield 1,2,3-triazole/quinoline-fused imidazo[1,2-a]pyridines. This innovative technique combines intramolecular cross dehydrogenative C–C coupling between 1,2,3–triazole and imidazo[1,2-a]pyridine, azide–alkyne cycloaddition, and C–N coupling between 1,2,3–triazole and aryl bromide. Catalyzed by a one-pot bimetallic relay, this cascade process stands out for its simplicity in starting materials, minimal synthetic steps, high efficiency, and utilization of renewable oxidants, rendering it a unique method (Scheme 25).
![Scheme 25:
Through single-pot cascade reactions, bimetallic relay catalysis yields [1,2,3]triazolo[1,5-a]quinolines.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_025.jpg)
Through single-pot cascade reactions, bimetallic relay catalysis yields [1,2,3]triazolo[1,5-a]quinolines.
Fused benzoxazino[1,2,3]Quinolinone 80 4a–4k triazolyl [4,5–c] derivatives 49 can be made in a single pot using the 5-iodo-4-(prop-2–yn–1–yl) reaction-2H-benzo[b][1,4] Oxazin-3(4H)-one (3) is synthesized in ligand-free conditions using a range of aryl azides and copper and palladium catalysis. The intended fused triazoles were produced in good to exceptional yields by the process. The synthetic derivatives’ in vitro cytotoxic potential against the cancer cell lines MCF-7, A-549, and HeLa was assessed using the MTT assay. The outcomes demonstrated that compound 4d′s cytotoxic activity against MCF-7 and A-549 demonstrated a wide range of activity. The IC50 values of cisplatin, the recommended drug, are similar at 11.18 ± 0.8 and 17.81 ± 0.6 μM (Scheme 26).
![Scheme 26:
Anticancer characteristics of fused benzoxazino [1, 2, 3] triazolyl [4, 5-c] quinolinone derivatives synthesized in a single pot.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_026.jpg)
Anticancer characteristics of fused benzoxazino [1, 2, 3] triazolyl [4, 5-c] quinolinone derivatives synthesized in a single pot.
A quick method for 81 Palladium catalysis has been successfully employed to create a variety of fused triazoles with different functionalities under ligand-free conditions. The palladium-catalyzed cyclization that yields the hitherto unattainable fused triazolo[4,5-d]quinoline/chromene/thiochromene derivatives is significantly aided by the addition of TBAI, it was discovered. This technique offers a quick and easy way to get good to outstanding yields of fused triazoles. Compounds 2a and 2f exhibit inhibitory effects against a certain non-small cell lung cancer cell line, as demonstrated by the screening of the produced compounds against several cancer cell lines (Scheme 27).
![Scheme 27:
Fusion of fused triazolo[4,5-d]quinoline derivatives is catalyzed by Pd and requires tetrabutylammonium iodide.](/document/doi/10.1515/revic-2024-0088/asset/graphic/j_revic-2024-0088_scheme_027.jpg)
Fusion of fused triazolo[4,5-d]quinoline derivatives is catalyzed by Pd and requires tetrabutylammonium iodide.
9 Azide–alkyne cycloaddition catalyzed by gold (AuAAC)
Muthusubramanian and associates in 2013 82 revealed the synthesis of 1,4-disubstituted triazoles 109 and 110 from substituted benzoyl azides 108 and terminal/internal alkynes 1/200 RT in water utilizing porous Au/TiO2 nanoparticles for 20–30/45 min, respectively. This process was green, efficient, and regioselective (Scheme 28). They also stated that while p-xylene, 1,2-dichloroethane, acetonitrile, and toluene offered extremely low yields, solvents like THF, DMSO, and ethanol gave intermediate yields while water or tBuOH/water gave substantial yields as single regioisomers. After more research, the scientists discovered that a stepwise method utilizing azide 108, benzoyl/alkyl bromide, and 107 produced triazoles with a 75 % yield in 30 min, outperforming one-pot MCR. With no appreciable yield loss, the catalyst was created utilizing a deposition-precipitation technique in up to five cycles. They discovered that aliphatic and aromatic alkynes with electron-donating and electron-withdrawing groups were appropriate for this technique, given the range of alkynes used as substrates (Scheme 28).

Azide–alkyne cycloaddition catalyzed by gold (AuAAC).
10 Azide–alkyne cycloaddition catalyzed by silver (AgAAC)
Sarma and associates 83 revealed a unique, robust, and extremely effective method for the regioselective synthesis of 113 from 112 and 11 utilizing water and ethylene glycol at room temperature for 2–6 h, which was catalyzed by very efficient silver dicyanamide/DIPEA (Scheme 29). In terms of alkyne substrate scope, they discovered that aromatic alkynes produced a yield that was somewhat higher (90–97 %) than that of aliphatic alkynes (90–94 %). Alcohol and ester, among other functional groups with alkynes, were also appropriate substrates for this methodology (Scheme 29).

Azide–alkyne cycloaddition catalyzed by silver (AgAAC).
Even though they have 84 Vinyl-1,2,3-triazole synthesis and polymerization have long been known, but only recently have they drawn attention due to the remarkable advancements in the most common application of the “click” chemical theory, the copper-catalyzed azide–alkyne cycloaddition. It is true that up to this point, a substantial library of monomers and polymers based on vinyl-1,2,3-triazoles with aromatic, aliphatic, heterocyclic, or functional groups has been described. Their polymerization hasn’t, however, received much media attention, though this is anticipated to change soon. This review discusses the synthesis and polymerization of vinyl-1,2,3-triazole regioisomers, the properties of the resultant materials, and their possible uses in the creation of macromolecular objects that are intelligent and stimuli-responsive (Scheme 30).

Polymerization and N- and C-vinyl-1,2,3-triazole synthesis.
A novel, reliable, and effective technique 83 for silver-mediated azide–alkyne cycloaddition (AgAAC) reactions in H2O/Ethylene Glycol at room temperature has been created. The method states that under mild conditions, 1,4-disubstituted-1,2,3 triazoles can be formed regioselectively using silver dicyanamide/DIPEA, a highly effective catalyst system (Scheme 31).

A room temperature catalytic method that is very successful for synthesizing 1,4-disubstituted-1,2,3 triazoles.
The term “click reaction” describes the dipolar cycloaddition reaction that is catalyzed by copper and involves azides and alkynes. This reaction forms a triazole ring by integrating nitrogens into a cyclic hydrocarbon scaffold. This reaction and its byproducts, triazole-containing heterocycles, are highly important in medicinal chemistry due to their efficiency and versatility. The only metal that can catalyze this process is copper, though its exact mechanism is still a mystery. Here we show that silver halides can also catalyze “click reactions” in non-aqueous media. It serves as a substitute for the widely recognized CuAAC click reaction. The kind of counter ion in the silver salt determines the reaction’s yield. Numerous notable characteristics of this reaction include strong regioselectivity, mild reaction conditions, easy availability of substrate, and reasonably high yields. This correspondence aims to shed light on the reaction’s still-mysterious mechanism by showcasing the effects of solvent and counter ions and presenting the outcomes of a novel catalyst (Scheme 32).

Azide–alkyne cycloaddition (AgAAC) catalyzed by silver: evaluating the mechanism through calculations using density functional theory.
There is a report on an all-purpose 85 hydroazidation cascade and the alkyne-azide 1,3-dipolar cycloaddition of diynes using silver catalysis. Matching 1,5-fused-1,2,3-triazoles were produced in good to exceptional yields when trimethylsilyl azide (TMS-N3) was reacted with a variety of diynes in the presence of H2O. This novel technique provides convenient access to a range of fused 1,2,3-triazoles and is easy to operate, thanks to its broad substrate scope, good functional group tolerance, and high reaction efficiency (Scheme 33).

Silver-catalyzed diynes 1,5-fused 1,2,3-triazole frameworks are produced with TMS-N3.
The cycloaddition catalyzed by silver 86 Density functional theory (DFT) simulations were used to study the synthesis of 1,4-disubstituted-1,2,3-triazoles. We present computational investigations of the primary reaction mechanism, two side reaction processes that have been postulated, and the contribution of silver(I) to the understanding of the key characteristics of the transformation in this paper. The investigation of the mechanism suggests that the [3 + 2]-cycloaddition is more likely to be the step that determines the primary reaction’s rate, whereas the two competitive side reaction paths are kinetically unfavorable and have comparatively high barriers. Frontier Molecular Orbital (FMO) study suggests that the properties of the isocyanide FMO and how FMO overlap occurs in the cycloaddition are altered when silver(I) is coordinated with isocyanides. According to Natural Population Analysis (NPA) charge analysis, the electron-withdrawing characteristic of silver(I) causes the –NC group’s charge density to drop, improving the substituent tolerance and reactivity of isocyanides. Furthermore, in comparison to the substantial electron-withdrawing impact of silver(I) on isocyanides, the effect of substituents on the aryl rings on –NC groups is essentially insignificant. Therefore, it is believed that the superior aryl-substituted tolerance and strong regioselectivity of the technique are due to the catalyst, silver(I) (Scheme 34).

To provide mechanistic insight into the cycloaddition synthesis of 1,4-disubstituted-1,2,3-triazoles, silver is an essential component of the process.
11 Azide–alkyne cycloaddition catalyzed by zinc (ZnAAC)
Greaney and coworkers 87 revealed how to create 131 regioselectively from 129 and 130 by utilizing NH4Cl and ZnEt2 as catalysts for 72 h at room temperature (Scheme 35). Because of the reactants’ breakdown at high temperatures, the reaction’s yield is reduced. Esters, thioethers, propargyl ethers, 1,2-diphenylacetylene, and (iodoethynyl)benzene were found to be suitable for the alkyne substrate range, however, tosyl azide compounds and alkyl azides were found to be unsuitable for this approach for the azide range. Nevertheless, 1,4,5-trisubstituted triazoles were also produced by substituting D2O/D3CCO2D for NH4Cl via a 3-MCR coupling process, demonstrating substitution at the 4-position. The aforementioned procedure takes 18 h to finish at room temperature (Scheme 35).

Azide–alkyne cycloaddition catalyzed by zinc (ZnAAC).
Chen and coworkers 70 developed a novel, mild, heterogeneous method for Zn/C (zinc on carbon) catalysis to synthesize 135 and 136 from 134 with aryl alkynes 133 and 137, respectively. This enabled them to overcome the problem of homogeneous catalyst-product separation and the fact that relatively little progress has been made on heterogeneous solid catalysts (Scheme 36). Additionally, they discovered that 50 °C was the ideal temperature for the procedure and that the aprotic and polar solvent DMF was the appropriate solvent. While charcoal may convert Cu(II) to Cu(I), zinc dust frequently contains Cu(II) impurities. Despite the lack of cycloaddition products, they thought that a tiny amount of CuI, rather than Zn/C, caused an in situ catalyzed cycloaddition reaction. This suggests that the reaction might be catalyzed by Zn/C. Additionally, they discovered that aliphatic alkynes did not initiate a reaction and that aryl alkynes with low electron counts were formed in lower yields than those with high electron counts. Lower yields were obtained when the same procedure was used in a one-pot, three-component reaction. The catalyst’s efficiency drops after the fifth cycle, and the reaction takes 15–20 h to complete (Scheme 36).

Intramolecular Schmidt reaction to prepare amides.
12 Synthesis of 1,4-disubstituted 1,2,3-trizole without metal
Furthermore, there has been a lot of interest lately in the organocatalysis-based synthesis of triazoles from azides. 88 , 89 , 90 , 91 Organocatalysis does not yield as many difficult-to-handle intractable metal salt by-products as metal catalysis does. A three-component one-pot technique for the synthesis of 1,4,5-trisubstituted 1,2,3-triazoles was created in 2012 by Cao’s research group (Scheme 37a) 92 The activated primary alcohol 137 reacts with N-(p-toluenesulfonyl)imidazole (Tslm) to generate alkyl tosylate (R1OTs) in this method, which does not require the use of a metal catalyst. The azide ion then combines with the R1OTs through a nucleophilic substitution reaction to produce an azido alkyl group (R1N3). A series of 1,2,3-triazole compounds 140 were produced by dehydration and the cycloaddition process between azido alkyl and enolate. A metal-free three-component process for the synthesis of 1,4,5-trisubstituted 1,2,3-triazoles was also described by Professor Dehaen in 2014 (Scheme 37b) 93 l-proline catalyzes the following reactions: 1,3-dipolar cycloaddition of azide and activated alkenes yields trisubstituted 1,2,3-triazole 147 this reaction is then followed by the Knoevenagel condensation reaction of formyl group with nitro compound 145. A number of triazole moieties and fused triazole heterocycles can be synthesized using this process, which also offers the benefits of good regioselectivity, high yield, and compatibility with functional groups.

Organocatalyzed azide to triazole.
13 Conclusions
Advances in the synthesis of 1,2,3-triazoles via azide–alkyne cycloaddition, catalyzed by transition metals, have increased the importance of these molecules in chemical biology, drug development, and materials science. Click chemistry offers a strong platform for their efficient synthesis, underscoring their versatility and potential in various applications. In this review, we have discussed catalytic performance of transition metals Cu, Ir, Rh, Ru, Ni, Pd, Au, Ag, and Zn, which were applied in the form ligands and salts in 1,2,3-triazole synthesis via azide–alkyne cycloaddition. Among all transition metals the Cu-complexes and salts were found more efficient for selective synthesis. It is also concluded that there are certain azide–alkyne reactions which completely catalytic specific, and cannot be perform as such by changing transition metal. Through this review, we have outlined the growing interest and recent advancements in this area with particular focus on reaction conditions, and mechanistic insights.
Correction note
Correction added [November 22, 2024] after online publication November 13, 2024: The e-mail address of Shahab Khan – shahabkhan262@gmail.com – was misspelled as shahabkhan262@gamil.com. In the section heading Cycloaddition of azide–alkyne catalyzed by Nickel the name of the chemical element was misspelled as Nikal.
Funding source: The Deanship of Scientific Research at King Khalid University, Saudi Arabia
Award Identifier / Grant number: RGP-2/611/45
-
Research ethics: Not applicable.
-
Informed consent: All the authors are agreed with this publication.
-
Author contributions: The literature survey and collection of data were performed by K.S.B. and R.I. along with validation. S.M. and A.W. validated, organized the data and improved the schemes. While S.A. and B.B.K.A. improved the manuscript quality along with language and grammatical improvement. M.A.I. improved the schemes and checked condition of provided reactions. The manuscript initial draft was prepared by S.K. while A.U. presented the data, along with editing, validation, and supervision.
-
Use of Large Language Models, AI and Machine Learning Tools: Not applicable.
-
Conflict of interest: Not applicable.
-
Research funding: The study was funded by the Deanship of Scientific Research at King Khalid University, Saudi Arabia, through a research group program under grant number RGP-2/611/45.
-
Data availability: Not applicable.
References
1. Wahab, A.; Gao, Z.; Gou, J.; Yu, B. The Construction of Phosphonate Triazolyl by Copper (II)-Catalyzed Furan Dearomatized [3+2] Cycloaddition. Org. Biomol. Chem. 2022, 20, 6319–6323; https://doi.org/10.1039/d2ob01060g.Search in Google Scholar PubMed
2. Singh, P.; Sharma, P.; Anand, A.; Bedi, P.; Kaur, T.; Saxena, A.; Kumar, V. Azide-Alkyne Cycloaddition En Route to Novel 1H-1, 2, 3-Triazole Tethered Isatin Conjugates With In Vitro Cytotoxic Evaluation. Eur. J. Med. Chem. 2012, 55, 455–461; https://doi.org/10.1016/j.ejmech.2012.06.057.Search in Google Scholar PubMed
3. Xu, Z.; Song, X.-F.; Hu, Y.-Q.; Qiang, M.; Lv, Z.-S. Azide-Alkyne Cycloaddition Towards 1H-1, 2, 3-Triazole-Tethered Gatifloxacin and Isatin Conjugates: Design, Synthesis and In Vitro Anti-mycobacterial Evaluation. Eur. J. Med. Chem. 2017, 138, 66–71; https://doi.org/10.1016/j.ejmech.2017.05.057.Search in Google Scholar PubMed
4. Wahab, A.; Cheng, G.; Su, H.; Yang, L.; Gao, Z.; Yu, B. Furan Ring Opening Reaction for the Synthesis of 2, 5-Dicarbonyl-3-Ene-Phosphates. Org. Biomol. Chem. 2023, 21, 7219–7223; https://doi.org/10.1039/d3ob00933e.Search in Google Scholar PubMed
5. Guo, J.; Yu, B.; Wang, Y.-N.; Duan, D.; Ren, L.-L.; Gao, Z.; Gou, J. Conversion of 2-Furylcarbinols With Alkyl or Aryl Azides to Highly Functionalized 1, 2, 3-Triazoles via Cascade Formal [3+ 2] Cycloaddition/Ring-Opening. Org. Lett. 2014, 16, 5088–5091; https://doi.org/10.1021/ol502437y.Search in Google Scholar PubMed
6. Hao, Z.; Zhao, P.; Xing, Q.; Wahab, A.; Gao, Z.; Gou, J.; Yu, B. Dual Roles of Azide: Dearomative Dimerization of Furfuryl Azides. J. Org. Chem. 2022, 87, 10185–10198; https://doi.org/10.1021/acs.joc.2c01118.Search in Google Scholar PubMed
7. El Malah, T.; Mageid, R. E. A.; Awad, H. M.; Nour, H. F. Copper (I)-Catalysed Azide–Alkyne Cycloaddition and Antiproliferative Activity of Mono-And Bis-1, 2, 3-Triazole Derivatives. New J. Chem. 2020, 44, 18256–18263; https://doi.org/10.1039/d0nj04308g.Search in Google Scholar
8. de Souza, M. V.; da Costa, C. F.; Facchinetti, V.; Gomes, C. R. B.; Pacheco, P. M. Advances in Triazole Synthesis From Copper-Catalyzed Azide-Alkyne Cycloadditions (Cuaac) Based on Eco-Friendly Procedures. Curr. Org. Synth. 2019, 16, 244–257; https://doi.org/10.2174/1570179416666190104141454.Search in Google Scholar PubMed
9. 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 1 H-1, 2, 3-Triazole-Linked 2-Amino-4, 8-Dihydropyrano [3, 2-b] Pyran-3-Carbonitrile–Crystal Structure. Res. Chem. Intermed. 2019, 45, 2079–2094; https://doi.org/10.1007/s11164-018-03723-x.Search in Google Scholar
10. Yang, M.; Yang, Y.; Chen, P. R. Transition-Metal-Catalyzed Bioorthogonal Cycloaddition Reactions. In Cycloadditions in Bioorthogonal Chemistry, 1st ed.; Springer, Cham.: Switzerland, 2016; pp 27–55.10.1007/978-3-319-29686-9_3Search in Google Scholar
11. Wang, C.; Ikhlef, D.; Kahlal, S.; Saillard, J.-Y.; Astruc, D. Metal-catalyzed Azide-Alkyne “Click” Reactions: Mechanistic Overview and Recent Trends. Coord. Chem. Rev. 2016, 316, 1–20; https://doi.org/10.1016/j.ccr.2016.02.010.Search in Google Scholar
12. Zhou, W.; Zhang, M.; Li, H.; Chen, W. One-Pot Three-Component Synthesis of Enamine-Functionalized 1, 2, 3-triazoles via Cu-Catalytic Azide–Alkyne Click (CuAAC) and Cu-Catalyzed Vinyl Nitrene Transfer Sequence. Org. Lett. 2017, 19, 10–13; https://doi.org/10.1021/acs.orglett.6b02850.Search in Google Scholar PubMed
13. Ikhlef, D.; Wang, C.; Kahlal, S.; Maouche, B.; Astruc, D.; Saillard, J.-Y. Reaction Mechanisms of Transition-Metal-Catalyzed Azide–Alkyne Cycloaddition “Click” Reactions: A DFT Investigation. Comput. Theor. Chem. 2015, 1073, 131–138; https://doi.org/10.1016/j.comptc.2015.09.020.Search in Google Scholar
14. Cornier, P. G.; Delpiccolo, C. M.; Martiren, N. L.; Mata, E. G.; Mendez, L.; Squizatto, C. P.; Pizzio, M. G. Transition Metal-Catalyzed Reactions and Solid-Phase Synthesis: A Convenient Blend. Eur. J. Org Chem. 2022, 2022, e202200526; https://doi.org/10.1002/ejoc.202200526.Search in Google Scholar
15. Johansson, J. R.; Beke-Somfai, T.; Said Stålsmeden, A.; Kann, N. Ruthenium-Catalyzed Azide Alkyne Cycloaddition Reaction: Scope, Mechanism, and Applications. Chem. Rev. 2016, 116, 14726–14768; https://doi.org/10.1021/acs.chemrev.6b00466.Search in Google Scholar PubMed
16. De Nino, A.; Maiuolo, L.; Costanzo, P.; Algieri, V.; Jiritano, A.; Olivito, F.; Tallarida, M. A. Recent Progress in Catalytic Synthesis of 1, 2, 3-Triazoles. Catalysts 2021, 11, 1120; https://doi.org/10.3390/catal11091120.Search in Google Scholar
17. Struthers, H.; Mindt, T. L.; Schibli, R. Metal Chelating Systems Synthesized Using the Copper (I) Catalyzed Azide-Alkyne Cycloaddition. Dalton Trans. 2010, 39, 675–696; https://doi.org/10.1039/b912608b.Search in Google Scholar PubMed
18. Mendoza-Espinosa, D.; Negrón-Silva, G. E.; Ángeles-Beltrán, D.; Álvarez-Hernández, A.; Suárez-Castillo, O. R.; Santillán, R. Copper (II) Complexes Supported by Click Generated Mixed NN, NO, and NS 1, 2, 3-Triazole Based Ligands and Their Catalytic Activity in Azide–Alkyne Cycloaddition. Dalton Trans. 2014, 43, 7069–7077; https://doi.org/10.1039/c4dt00323c.Search in Google Scholar PubMed
19. Jiang, L.; Wang, Z.; Bai, S.-Q.; Hor, T. A. “Click-and-Click”–Hybridised 1, 2, 3-Triazoles Supported Cu (I) Coordination Polymers for Azide–Alkyne Cycloaddition. Dalton Trans. 2013, 42, 9437–9443; https://doi.org/10.1039/c3dt50987g.Search in Google Scholar PubMed
20. Hein, J. E.; Tripp, J. C.; Krasnova, L. B.; Sharpless, K. B.; Fokin, V. V. Copper (I)-Catalyzed Cycloaddition of Organic Azides and 1-Iodoalkynes. Angew. Chem. 2009, 121, 8162–8165; https://doi.org/10.1002/ange.200903558.Search in Google Scholar
21. Yadav, D.; Singh, N.; Kim, T. W.; Kim, J. Y.; Park, N.-J.; Baeg, J.-O. Highly Regioselective and Sustainable Solar Click Reaction: A New Post-synthetic Modified Triazole Organic Polymer as a Recyclable Photocatalyst for Regioselective Azide–Alkyne Cycloaddition Reaction. Green Chem. 2019, 21, 2677–2685; https://doi.org/10.1039/c9gc00894b.Search in Google Scholar
22. Ali, A. A.; Konwar, M.; Chetia, M.; Sarma, D. [Bmim] OH Mediated Cu-Catalyzed Azide–Alkyne Cycloaddition Reaction: A Potential Green Route to 1, 4-Disubstituted 1, 2, 3-Triazoles. Tetrahedron Lett. 2016, 57, 5661–5665; https://doi.org/10.1016/j.tetlet.2016.11.014.Search in Google Scholar
23. Kouznetsov, V. V.; Vargas-Mendez, L. Y.; Zubkov, F. I. Recent Advances in the Synthesis of Bioactive Quinoline-Based 1, 2, 3-Triazoles via Cu-Catalyzed Huisgen 1, 3-Dipolar Cycloaddition (“Click Reaction”). Mini-Reviews Org. Chem. 2016, 13, 488–503.10.2174/1570193X13666161102150625Search in Google Scholar
24. Lauria, A.; Delisi, R.; Mingoia, F.; Terenzi, A.; Martorana, A.; Barone, G.; Almerico, A. M. 1, 2, 3-Triazole in Heterocyclic Compounds, Endowed with Biological Activity, Through 1, 3-Dipolar Cycloadditions. Eur. J. Org Chem. 2014, 2014, 3289–3306; https://doi.org/10.1002/ejoc.201301695.Search in Google Scholar
25. Mbugua, S. N. Thesis on the subject “Synthesis, Structure Elucidation and Reactivity of Palladium”. In Synthesis, Structure Elucidation and Reactivity of Palladium (II) and Platinum (II) Complexes for Anticancer Applications; University of Nairobi: Nairobi, 2020.Search in Google Scholar
26. 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; https://doi.org/10.1016/j.ejmech.2018.08.002.Search in Google Scholar PubMed
27. Phukan, P.; Sarma, D. Synthesis of Medicinally Relevant Scaffolds-Triazoles and Pyrazoles in Green Solvent Ionic Liquids. Curr. Org. Chem. 2021, 25, 1523–1538; https://doi.org/10.2174/1385272825666210412160142.Search in Google Scholar
28. 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; https://doi.org/10.1016/j.ejmech.2017.04.044.Search in Google Scholar PubMed
29. 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; https://doi.org/10.1016/j.ejmech.2014.07.075.Search in Google Scholar PubMed
30. Juríček, M.; Kouwer, P. H.; Rowan, A. E. Triazole: A Unique Building Block for the Construction of Functional Materials. Chem. Commun. 2011, 47, 8740–8749; https://doi.org/10.1039/c1cc10685f.Search in Google Scholar PubMed
31. 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. RSC Adv. 2020, 10, 5610–5635; https://doi.org/10.1039/c9ra09510a.Search in Google Scholar PubMed PubMed Central
32. Jiang, X.; Hao, X.; Jing, L.; Wu, G.; Kang, D.; Liu, X.; Zhan, P. Recent Applications of Click Chemistry in Drug Discovery. Expet Opin. Drug Discov. 2019, 14, 779–789; https://doi.org/10.1080/17460441.2019.1614910.Search in Google Scholar PubMed
33. Singh, N.; Pandey, S.; 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, 1641–1648; https://doi.org/10.1016/j.carres.2010.04.019.Search in Google Scholar PubMed
34. 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, 60–69; https://doi.org/10.1002/ejoc.201600988.Search in Google Scholar
35. 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, 567–574; https://doi.org/10.1093/jac/25.4.567.Search in Google Scholar PubMed
36. 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 Discov. Today 2017, 22, 1572–1581; https://doi.org/10.1016/j.drudis.2017.05.014.Search in Google Scholar PubMed
37. Ma, J.; Ding, S. Transition Metal-Catalyzed Cycloaddition of Azides with Internal Alkynes. Asian J. Org. Chem. 2020, 9, 1872–1888; https://doi.org/10.1002/ajoc.202000486.Search in Google Scholar
38. Maiuolo, L.; Algieri, V.; Olivito, F.; De Nino, A. Recent Developments on 1, 3-Dipolar Cycloaddition Reactions by Catalysis in Green Solvents. Catalysts 2020, 10, 65; https://doi.org/10.3390/catal10010065.Search in Google Scholar
39. Maiuolo, L.; Algieri, V.; Olivito, F.; Tallarida, M. A.; Costanzo, P.; Jiritano, A.; De Nino, A. Chronicle of Nanocelluloses (NCs) for Catalytic Applications: Key Advances. Catalysts 2021, 11, 96; https://doi.org/10.3390/catal11010096.Search in Google Scholar
40. Chetia, M.; Ali, A. A.; Bordoloi, A.; Sarma, D. Facile Route for the Regioselective Synthesis of 1, 4-Disubstituted 1, 2, 3-Triazole Using Copper Nanoparticles Supported on Nanocellulose as Recyclable Heterogeneous Catalyst. J. Chem. Sci. 2017, 129, 1211–1217; https://doi.org/10.1007/s12039-017-1318-y.Search in Google Scholar
41. Maiuolo, L.; Feriotto, G.; Algieri, V.; Nardi, M.; Russo, B.; Di Gioia, M. L.; Furia, E.; Tallarida, M. A.; Mischiati, C.; De Nino, A. Antiproliferative Activity of Novel Isatinyl/indanyl Nitrones (INs) as Potential Spin Trapping Agents of Free Radical Intermediates. MedChemComm 2018, 9, 299–304; https://doi.org/10.1039/c7md00537g.Search in Google Scholar PubMed PubMed Central
42. Merino, P.; Maiuolo, L.; Delso, I.; Algieri, V.; De Nino, A.; Tejero, T. Chemical Approaches to Inhibitors of Isoprenoid Biosynthesis: Targeting Farnesyl and Geranylgeranyl Pyrophosphate Synthases. RSC Adv. 2017, 7, 10947–10967; https://doi.org/10.1039/c6ra28316k.Search in Google Scholar
43. Maiuolo, L.; De Nino, A.; Algieri, V.; Nardi, M. Microwave-Assisted 1, 3-Dipolar Cyclo-Addition: Recent Advances in Synthesis of Isoxazolidines. Mini-Reviews Org. Chem. 2017, 14, 136–142; https://doi.org/10.2174/1570193x14666170206123513.Search in Google Scholar
44. Martina, K.; Tagliapietra, S.; Veselov, V. V.; Cravotto, G. Green Protocols in Heterocycle Syntheses via 1, 3-Dipolar Cycloadditions. Front. Chem. 2019, 7, 95; https://doi.org/10.3389/fchem.2019.00095.Search in Google Scholar PubMed PubMed Central
45. Calandra, P.; Szerb, E. I.; Lombardo, D.; Algieri, V.; De Nino, A.; Maiuolo, L. A Presentation of Ionic Liquids as Lubricants: Some Critical Comments. Appl. Sci. 2021, 11, 5677; https://doi.org/10.3390/app11125677.Search in Google Scholar
46. 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, 3936–3946; https://doi.org/10.1055/s-0035-1560352.Search in Google Scholar
47. 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, 10766–10770; https://doi.org/10.1002/anie.201705006.Search in Google Scholar PubMed PubMed Central
48. 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, 2708–2711; https://doi.org/10.1002/1521-3757(20020715)114:14<2708::aid-ange2708>3.0.co;2-0.10.1002/1521-3757(20020715)114:14<2708::AID-ANGE2708>3.0.CO;2-0Search in Google Scholar
49. Haldón, E.; Nicasio, M. C.; Pérez, P. J. Copper-Catalysed Azide–Alkyne Cycloadditions (CuAAC): An Update. Org. Biomol. Chem. 2015, 13, 9528–9550; https://doi.org/10.1039/c5ob01457c.Search in Google Scholar
50. Huisgen, R. Kinetics and Mechanism of 1, 3-Dipolar Cycloadditions. Angew Chem. Int. Ed. Engl. 1963, 2, 633–645; https://doi.org/10.1002/anie.196306331.Search in Google Scholar
51. Qian, W.; Wang, H.; Allen, J. Copper-Catalyzed Domino Cycloaddition/C-N Coupling/Cyclization/(C-H Arylation): An Efficient Three-Component Synthesis of Nitrogen Polyheterocycles. Angew. Chem. Int. Ed. 2013, 52, 10992–10996; https://doi.org/10.1002/anie.201305970.Search in Google Scholar
52. Li, D.; Mao, T.; Huang, J.; Zhu, Q. A One-Pot Synthesis of [1,2,3]Triazolo[1,5-A]Quinoxalines from 1-Azido-2-Isocyanoarenes With High Bond-Forming Efficiency. Chem. Commun. 2017, 53, 1305–1308; https://doi.org/10.1039/c6cc08543a.Search in Google Scholar PubMed
53. Pericherla, K.; Jha, A.; Khungar, B.; Kumar, A. Copper-Catalyzed Tandem Azide–Alkyne Cycloaddition, Ullmann Type C–N Coupling, and Intramolecular Direct Arylation. Org. Lett. 2013, 15, 4304–4307; https://doi.org/10.1021/ol401655r.Search in Google Scholar PubMed
54. Roy, B.; Mondal, D.; Hatai, J.; Bandyopadhyay, S. A Highly Efficient Tandem [3 + 2] “Click” Cycloaddition/6-Exo-Cyclization Strategy for the Construction of Triazole Fused Pyrazines. RSC Adv. 2014, 4, 56952–56956; https://doi.org/10.1039/c4ra12489h.Search in Google Scholar
55. Liu, Z.; Zhu, D.; Luo, B.; Zhang, N.; Liu, Q.; Hu, Y.; Pi, R.; Huang, P.; Wen, S. Mild Cu(I)-Catalyzed Cascade Reaction of Cyclic Diaryliodoniums, Sodium Azide, and Alkynes: Efficient Synthesis of Triazolophenanthridines. Org. Lett. 2014, 16, 5600–5603; https://doi.org/10.1021/ol502654a.Search in Google Scholar PubMed
56. 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, 8923–8930; https://doi.org/10.1021/ja805956e.Search in Google Scholar
57. Qureshi, Z.; Kim, J. Y.; Bruun, T.; Lam, H.; Lautens, M. Cu/Pd-Catalyzed Synthesis of Fully Decorated Polycyclic Triazoles: Introducing C–H Functionalization to Multicomponent Multicatalytic Reactions ((MC) 2R). ACS Catal. 2016, 6, 4946–4952; https://doi.org/10.1021/acscatal.6b00858.Search in Google Scholar
58. Wang, Y.-C.; Xie, Y.-Y.; Qu, H.-E.; Wang, H.-S.; Pan, Y.-M.; Huang, F.-P. Ce (OTf) 3-Catalyzed [3+ 2] Cycloaddition of Azides with Nitroolefins: Regioselective Synthesis of 1, 5-Disubstituted 1, 2, 3-Triazoles. J. Org. Chem. 2014, 79, 4463–4469; https://doi.org/10.1021/jo5004339.Search in Google Scholar PubMed
59. Zhou, P.; Lin, L.; Chen, L.; Zhong, X.; Liu, X.; Feng, X. Iron-Catalyzed Asymmetric Haloazidation of α, β-unsaturated Ketones: Construction of Organic Azides with Two Vicinal Stereocenters. J. Am. Chem. Soc. 2017, 139, 13414–13419; https://doi.org/10.1021/jacs.7b06029.Search in Google Scholar PubMed
60. Lamberti, M.; Fortman, G. C.; Poater, A.; Broggi, J.; Slawin, A. M.; Cavallo, L.; Nolan, S. P. Coordinatively Unsaturated Ruthenium Complexes as Efficient Alkyne–Azide Cycloaddition Catalysts. Organometallics 2012, 31, 756–767; https://doi.org/10.1021/om2012425.Search in Google Scholar
61. Liu, P.; Clark, R. J.; Zhu, L. Synthesis of 1-Cyanoalkynes and Their Ruthenium (II)-Catalyzed Cycloaddition with Organic Azides to Afford 4-Cyano-1, 2, 3-Triazoles. J. Org. Chem. 2018, 83, 5092–5103; https://doi.org/10.1021/acs.joc.8b00424.Search in Google Scholar PubMed
62. 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, 12121–12124; https://doi.org/10.1021/jacs.7b06338.Search in Google Scholar PubMed
63. 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, 3042–3050; https://doi.org/10.1039/d1ma00143d.Search in Google Scholar
64. 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, 12121–12124; https://doi.org/10.1021/jacs.7b06338.Search in Google Scholar
65. 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; https://doi.org/10.1016/j.tetlet.2021.153131.Search in Google Scholar
66. GyumáKim, W.; YeongáJeong, S.; HwanáJeon, J.; YouáHong, S.; 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, 3374–3381; https://doi.org/10.1039/d0ob00579g.Search in Google Scholar PubMed
67. 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, 6200–6203; https://doi.org/10.1021/acs.orglett.7b03123.Search in Google Scholar PubMed
68. 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, 6200–6203; https://doi.org/10.1021/acs.orglett.7b03123.Search in Google Scholar
69. Chen, R.; Zeng, L.; Lai, Z.; Cui, S. Iridium-Catalyzed Hydroxyl-Enabled Cycloaddition of Azides and Alkynes. Adv. Synth. Catal. 2019, 361, 989–994; https://doi.org/10.1002/adsc.201801410.Search in Google Scholar
70. Zhang, L.; Chen, X.; Xue, P.; Sun, H. H.; Williams, I. D.; Sharpless, K. B.; Fokin, V. V.; Jia, G. Ruthenium-Catalyzed Cycloaddition of Alkynes and Organic Azides. J. Am. Chem. Soc. 2005, 127, 15998–15999; https://doi.org/10.1021/ja054114s.Search in Google Scholar PubMed
71. Rasmussen, L. K.; Boren, B. C.; Fokin, V. V. Ruthenium-Catalyzed Cycloaddition of Aryl Azides and Alkynes. Org. Lett. 2007, 9, 5337–5339; https://doi.org/10.1021/ol701912s.Search in Google Scholar PubMed
72. Das, U. K.; Jena, R. K.; Bhattacharjee, M. Synthesis, Structure and Catalytic Properties of [Ru(dppp)2(CH3CN)Cl] [BPh4] and Isolation of Catalytically Active [Ru(dppp)2Cl] [BPh4]: Ruthenium Catalysed Alkyne Homocoupling and Tandem Alkyne–Azide Cycloaddition. RSC Adv. 2014, 4, 21964–21970; https://doi.org/10.1039/c4ra01411a.Search in Google Scholar
73. Kelly, A. R.; Wei, J.; Kesavan, S.; Marié, J.-C.; Windmon, N.; Young, D. W.; Marcaurelle, L. A. Accessing Skeletal Diversity Using Catalyst Control: Formation of N and N + 1 Macrocyclic Triazole Rings. Org. Lett. 2009, 11, 2257–2260; https://doi.org/10.1021/ol900562u.Search in Google Scholar PubMed PubMed Central
74. Oppilliart, S.; Mousseau, G.; Zhang, L.; Jia, G.; Thuéry, P.; Rousseau, B.; Cintrat, J. C. 1-Protected 5-Amido 1,2,3-Triazoles via Ruthenium-Catalyzed [3+2] Cycloaddition of Azides and Ynamides. Tetrahedron 2007, 63, 8094–8098; https://doi.org/10.1016/j.tet.2007.06.008.Search in Google Scholar
75. Oakdale, J. S.; Sit, R. K.; Fokin, V. V. Ruthenium-Catalyzed Cycloadditions of 1-Haloalkynes with Nitrile Oxides and Organic Azides: Synthesis of 4-Haloisoxazoles and 5-Halotriazoles. Chem. Eur J. 2014, 20, 11101–11110; https://doi.org/10.1002/chem.201402559.Search in Google Scholar PubMed PubMed Central
76. 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, 7529–7534; https://doi.org/10.1021/acscatal.7b02558.Search in Google Scholar
77. 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, 1681–1688; https://doi.org/10.1007/s00706-015-1430-y.Search in Google Scholar
78. Brahma, K.; Achari, B.; Chowdhury, C. Facile Synthesis of [1, 2, 3]-Triazole-Fused Isoindolines, Tetrahydroisoquinolines, Benzoazepines and Benzoazocines by Palladium-Copper Catalysed Heterocyclisation. Synthesis 2013, 45, 545–555; https://doi.org/10.1055/s-0032-1316839.Search in Google Scholar
79. Wang, Z.; Li, B.; Zhang, X.; Fan, X. One-Pot Cascade Reactions Leading to Pyrido[2′,1′:2,3]Imidazo[4,5-c] [1,2,3]Triazolo[1,5-a]Quinolines Under Bimetallic Relay Catalysis with Air as the Oxidant. J. Org. Chem. 2016, 81, 6357–6363; https://doi.org/10.1021/acs.joc.6b00996.Search in Google Scholar PubMed
80. Narsimha, S.; Battula, K. S.; Nukala, S. K.; Gondru, R.; Reddy, Y. N.; Nagavelli, V. R. One-pot Synthesis of Fused Benzoxazino [1, 2, 3] Triazolyl [4, 5-c] Quinolinone Derivatives and Their Anticancer Activity. RSC Adv. 2016, 6, 74332–74339; https://doi.org/10.1039/c6ra12285j.Search in Google Scholar
81. Chen, C.-Y.; Yang, C.-H.; Hu, W.-P.; Kishore Vandavasi, J.; Chung, M.-I.; Wang, J.-J. Synthesis of Fused Triazolo[4,5-D]quinoline/chromene/thiochromene Derivatives via Palladium Catalysis Mediated by Tetrabutylammonium Iodide. RSC Adv. 2013, 3, 2710–2719; https://doi.org/10.1039/c2ra22799a.Search in Google Scholar
82. 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 Sustain. Chem. Eng. 2013, 1, 1405–1411; https://doi.org/10.1021/sc400147r.Search in Google Scholar
83. 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, 5892–5895; https://doi.org/10.1016/j.tetlet.2015.09.025.Search in Google Scholar
84. Beghdadi, S.; Miladi, I. A.; Addis, D.; Romdhane, H. B.; Bernard, J.; Drockenmuller, E. Synthesis and Polymerization of C-Vinyl- and N-Vinyl-1,2,3-Triazoles. Polym. Chem. 2012, 3, 1680–1692; https://doi.org/10.1039/c1py00446h.Search in Google Scholar
85. Ning, Y.; Wu, N.; Yu, H.; Liao, P.; Li, X.; Bi, X. Silver-Catalyzed Tandem Hydroazidation/Alkyne–Azide Cycloaddition of Diynes with TMS-N3: An Easy Access to 1,5-Fused 1,2,3-Triazole Frameworks. Org. Lett. 2015, 17, 2198–2201; https://doi.org/10.1021/acs.orglett.5b00784.Search in Google Scholar PubMed
86. Jia, F.; Zhang, B. Mechanistic Insight into the Silver-Catalyzed Cycloaddition Synthesis of 1,4-Disubstituted-1,2,3-Triazoles: The Key Role of Silver. New J. Chem. 2019, 43, 8634–8643; https://doi.org/10.1039/c9nj01700c.Search in Google Scholar
87. 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, 4826–4829; https://doi.org/10.1021/ol402225d.Search in Google Scholar PubMed PubMed Central
88. Gil de Montes, E.; Martı́nez-Bailén, M.; Carmona, A. T.; Robina, I.; Moreno-Vargas, A. J. Regioselectivity of the 1, 3-Dipolar Cycloaddition of Organic Azides to 7-heteronorbornadienes. Synthesis of β-Substituted Furans/Pyrroles. J. Org. Chem. 2020, 85, 8923–8932; https://doi.org/10.1021/acs.joc.0c00810.Search in Google Scholar PubMed
89. Chakraborty, A.; Dey, S.; Sawoo, S.; Adarsh, N.; Sarkar, A. Regioselective 1, 3-Dipolar Cycloaddition Reaction of Azides with Alkoxy Alkynyl Fischer Carbene Complexes. Organometallics 2010, 29, 6619–6622; https://doi.org/10.1021/om1009523.Search in Google Scholar
90. Verbelen, B.; Dehaen, W. Two-Step Synthesis of Fluorescent 3-Arylated 1, 3a, 6a-Triazapentalenes via a Three-Component Triazolization Reaction. Org. Lett. 2016, 18, 6412–6415; https://doi.org/10.1021/acs.orglett.6b03309.Search in Google Scholar PubMed
91. Thomas, J.; Goyvaerts, V.; Liekens, S.; Dehaen, W. Metal-Free Route for the Synthesis of 4-Acyl-1, 2, 3-Triazoles from Readily Available Building Blocks. Chem.--Eur. J. 2016, 22, 9966–9970; https://doi.org/10.1002/chem.201601928.Search in Google Scholar PubMed
92. John, J.; Thomas, J.; Parekh, N.; Dehaen, W. Tandem Organocatalyzed Knoevenagel Condensation/1, 3-Dipolar Cycloaddition Towards Highly Functionalized Fused 1, 2, 3-Triazoles. Eur. J. Org Chem. 2015, 2015, 4922–4930; https://doi.org/10.1002/ejoc.201500459.Search in Google Scholar
93. Jin, G.; Zhang, J.; Fu, D.; Wu, J.; Cao, S. One-Pot, Three-Component Synthesis of 1, 4, 5-Trisubstituted 1, 2, 3-Triazoles Starting from Primary Alcohols. Eur. J. Org Chem. 2012, 2012, 5446–5449; https://doi.org/10.1002/ejoc.201200830.Search in Google Scholar
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Advancements in CNT-based materials for optimized pharmaceutical removal via adsorption and photocatalysis
- Review on removal of heavy metals from industrial effluents by adsorption
- Cutting-edge techniques in low-temperature electrochemical water splitting: advancements in hydrogen production
- Biological activities of metal complexes with Schiff base
- Potential of organometallic complexes in medicinal chemistry
- Advancement in schiff base complexes for treatment of colon cancer
- Magnetic nanoparticles for efficient heavy metal removal: synthesis, adsorption capacity, and key experimental parameters
- Coal-based carbon/graphene quantum dots: formation mechanisms and applications
- Advancements in transition metal-catalyzed 1,2,3-triazole synthesis via azide–alkyne cycloaddition
- Gold complexes: a new frontier in the battle against lung cancer
Articles in the same Issue
- Frontmatter
- Advancements in CNT-based materials for optimized pharmaceutical removal via adsorption and photocatalysis
- Review on removal of heavy metals from industrial effluents by adsorption
- Cutting-edge techniques in low-temperature electrochemical water splitting: advancements in hydrogen production
- Biological activities of metal complexes with Schiff base
- Potential of organometallic complexes in medicinal chemistry
- Advancement in schiff base complexes for treatment of colon cancer
- Magnetic nanoparticles for efficient heavy metal removal: synthesis, adsorption capacity, and key experimental parameters
- Coal-based carbon/graphene quantum dots: formation mechanisms and applications
- Advancements in transition metal-catalyzed 1,2,3-triazole synthesis via azide–alkyne cycloaddition
- Gold complexes: a new frontier in the battle against lung cancer