Next Article in Journal
Indoor Air Photocatalytic Decontamination by UV–Vis Activated CuS/SnO2/WO3 Heterostructure
Next Article in Special Issue
Polymer-Supported Dioxidovanadium(V) Complex-Based Heterogeneous Catalyst for Multicomponent Biginelli Reaction Producing Biologically Active 3,4-Dihydropyrimidin-2-(1H)-ones
Previous Article in Journal
Electrocatalysis of 2,6-Dinitrophenol Based on Wet-Chemically Synthesized PbO-ZnO Microstructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Multicomponent Reactions Catalysed under Operationally Heterogeneous Conditions

Green Synthetic Organic Chemistry Laboratory, Department of Chemistry, University of Malta, 2080 Msida, MSD, Malta
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(7), 725; https://doi.org/10.3390/catal12070725
Submission received: 24 May 2022 / Revised: 25 June 2022 / Accepted: 28 June 2022 / Published: 30 June 2022
(This article belongs to the Special Issue Multicomponent Catalytic Reactions under Green Conditions)

Abstract

:
Multicomponent reactions (MCRs) have been gaining significance and attention over the past decade because of their ability to furnish complex products by using readily available and simple starting materials while simultaneously eliminating the need to separate and purify any intermediates. More so, most of these products have been found to exhibit diverse biological activities. Another paradigm shift which has occurred contemporarily is the switch to heterogeneous catalysis, which results in additional benefits such as the reduction of waste and an increase in the safety of the process. More importantly, it allows the user to recover and reuse the catalyst for multiple runs. In summary, both methodologies adhere to the principles of green chemistry, a philosophy which needs to become overarchingly enshrined. The plethora of reactions and catalysts which have been developed gives hope that chemists are slowly changing their ideology. As a result, this review attempts to discuss multicomponent reactions catalysed by operationally heterogeneous catalysts in the past 10 years. In this review, a further distinction is made between the MCRs which lead to the formation of heterocycles and those which do not.

1. Introduction

An obvious question which may immediately spring to mind when considering the vitality of catalysing multicomponent reactions under heterogeneous catalysis in synthetic organic chemistry is, “Why should one attempt multicomponent reactions? And why should the latter be performed under heterogeneous conditions instead of the classic homogeneous single-step reactions?”. The answer to this is two-fold. First and foremost, heterogeneous reactions are environmentally benign. Nowadays, our planet is continuously giving red alerts to humanity due to the environmental negligence and harm exerted in the previous 100 years, hence it is high time that waste production in the chemical industry be controlled via the use of recyclable catalysts, amongst other methods. Secondly, the use of more than two reagents simultaneously in multicomponent reactions allows for the formation of products which can never be produced by the single-stage stepwise reactions carried out in the past. In view of the outlined points, this review aims to underline the importance of various multicomponent reactions that have been studied under operationally heterogeneous conditions and reported in the past 10 years.
Yet, before embarking on such a review, it is critical to outline that nowadays, the use of the terms homogeneous and heterogeneous is becoming problematic. One of the founding pillars of green chemistry centres on the use of catalytic instead of stoichiometric amounts of reagents, which speeds up chemical reactions, hence resulting in shorter reaction times and possibly also renders the reactions viable at lower temperatures [1]. From an environmental perspective, if the catalyst is recoverable and reusable it would be even more benign and sustainable. In scientific diction, heterogeneous catalysts are those which can be recovered by filtration at the end of the reaction, whilst those which remain dissolved in the reaction mixture and cannot be filtered off nor reused are homogeneous [2].
The classical Ostwald definition of homogeneous and heterogeneous catalysts distinguishes between the two by stating that the former are those which are in the same phase as the reactants and the latter are those which are not [3]. Recent advances in catalytic systems and the development of metal clusters and nonclusters, nanoparticles, and nanocomposites, have evidenced that Ostwald’s definition is imprecise, if not completely irrelevant. A nanoparticle is a particle which is in the nanometre size range of 1–100 nm. Although these particles exist in a different phase, they normally remain suspended in solutions and are usually stabilized by the addition of appropriate surfactants [4]. During the course of a reaction, nanoparticles may aggregate with time to form nanoclusters and herein lies the fulcrum of the problem. When, and until what size, is a nanoparticle considered to be in the same phase as the reactants?
At this juncture, it is therefore crucial to distinguish whether a catalyst may be operationally or mechanistically homogeneous or heterogeneous. A catalyst may be described as operationally heterogeneous or homogeneous depending on whether the catalytic activity gets retained on a filter during a hot filtration test or ends up in the filtrate. An operationally heterogeneous catalyst is therefore one which should be able to be retrieved at the end of the reaction, rendering it environmentally friendly. However, in a supposedly homogeneously catalysed reaction by, for example, a metal complex, the actual active catalyst could be a nanoparticle metal cluster which forms by the transformation of the dissolved catalyst. In such a scenario, the reaction would be mechanistically heterogeneous but operationally homogeneous. Alternatively, in a heterogeneously catalysed reaction by a metal, a small fraction of the metal could be, in truth, dissolving in the reaction mixture, catalysing the reaction and then precipitating out once more onto the solid catalyst.
Keeping in mind all of the above, it is second-to-none to come up with another definition that it is clear and unambiguous. For greater clarity, Crabtree came up with two better terms that can be used to diversify catalysts, which are: homotopic and heterotopic [4]. The former term describes catalysts which have only one type of active site to catalyse a reaction (such as a metal–ligand complex), whilst a heterotopic one would be that which has several identical sites (Figure 1). A prime example would be palladium–charcoal, which is used to catalyse the hydrogenation of alkenes.
Keeping this definition in mind, the following discussion will revolve around the use of both types of catalysts which can still be ultimately recovered at the end of the reaction and reused, irrespective of the intrinsic mechanistic pathways. That is to say that, whenever the term heterogeneous is used, it could be that the catalyst is operationally heterogeneous but mechanistically either homo- or heterotopic.
This review will focus on environmentally benign catalysis in multicomponent reactions which have become crucial tools in synthetic chemistry, and which have generated interest due to their economic and ecological advantages [5]. Multicomponent reactions are one-pot processes in which three or more diverse reactants combine together simultaneously or in a stepwise domino fashion [6]. Not all multicomponent reactions proceed in the same fashion: the pathways can be either divergent or convergent (Figure 2). During convergent synthesis, separate reactants combine together in independent reactions to form intermediates, which subsequently react and combine together to form the final product. Conversely, in divergent pathways, reactants add on one after the other.
Multicomponent reactions (MCRs) exhibit a wide range of benefits over their classic sequential counterparts which involved reacting and purifying products individually in a sequential manner. First and foremost, in MCRs, purification is shortened and, hence, less solvent is required for purification purposes. In addition, multicomponent reactions are usually rapid and may be regio- and stereospecific [6]. Most often, the reactants of MCRs are simple, cheap, and readily available (in some cases they are even naturally sourced), making the process overall cheaper and more sustainable. Because most of the reactants will be incorporated into the final product, less waste is generated, resulting in what is known as an atom-economic reaction with a small E factor (E factor is a term introduced by Sheldon and is defined as the ratio of the mass of waste to the mass of product) [7].
E F = m   r a w   m a t e r i a l s + m   r e a g e n t s m   p r o d u c t m   p r o d u c t
In truth, the above is the simple E factor because it does not take water or solvents into account. The complete E factor considers the solvents and water that are used but not recovered at the end of the entire process, including in purification. From a biological perspective, the products of MCRs have a diverse and broad applicability. Amongst the plethora of reported uses, there are MCRs for which the products demonstrate antileishmanial [8], anti-inflammatory [9], anticancer [10], or antifibrotic activity, amongst others [11]. Moreover, MCR products also find uses as cloth pigments, food additives, flavourings, colourants [12], etc. This review aims to outline the importance of various multicomponent reactions studied under operationally heterogeneous conditions in the past 10 years. The following discussion will attempt to classify MCRs into two main categories: those leading to acyclic product formation and those which result in heterocyclic products. In relation to this, the review is further subdivided on the basis of the type of heterocycles and the state-of-the-art in their synthesis. Although a similar review for multicomponent reactions does exist, it is more focused on the mechanistic aspects and not so much on whether or not they are heterogeneously catalysed, and refer to a shorter timeframe [13].

2. Heterocyclic Compound Synthesis

Heterocycles are crucial in everyday life due to their diversity and uses. In biological reactions, such as those taking place in human cells, molecules are usually required to attach to chemoreceptors in cell membranes in order to elicit harmful/beneficial responses or trigger a sequence of reactions. These chemoreceptors have specific shapes and polarities that only allow for pairings with specific compounds, such as heterocycles, which, owing to their cyclic nature and lack of rotational activity, have fixed shapes and can bind effectively.

2.1. Heterocycles Containing 1-Heteroatom

2.1.1. Pyrroles

The pyrrole moiety is the main component of heme and chlorophyll, which are essential in two interlinked living organisms: heme is the oxygen-absorbing pigment in red blood cells in humans and chlorophyll is the green pigment in plants that allows for photosynthesis to take place. Undoubtedly, pyrrole derivatives constitute an important class of compounds which have other uses and have been synthesized in a multitude of ways, one of which involves the four-component reaction of an amine, an aldehyde, nitromethane, and a β-dicarbonyl compound. Li et al. catalysed this reaction efficiently using a molybdenum complex supported on magnetic nanoparticles of cobalt ferrite (CoFe2O4) in the presence of excess nitromethane at 90 °C (Figure 3) [14]. Various aromatic and some aliphatic amines were used, with yields mainly ranging 70–90% in 4–8 h (with two exceptions in a broad substrate range of 43 products). Positively, the catalyst could be easily recovered and reused for up to five consecutive runs with no loss of activity.
The reaction was also mediated by graphite in a study by Nandeesh et al., which had a smaller substrate scope (16 products) in reactions taking 4–6 h to complete, with yields of 60–88% [15]. Nitromethane was once more used in excess to serve simultaneously as both a reactant and a solvent.

2.1.2. Furans

The furan ring is quintessential in nature and is a structural motif in terpenes and alkaloids that elicits medicinal and agricultural activities [16]. Furan derivatives have been synthesized in various ways, such as by the 3-component reaction (3CR) of an aldehyde, an amine, and a dialkyl acetylene dicarboxylate using tin(II) oxide nanoparticles in ethanol solvent at room temperature (Figure 4) [17].

2.1.3. Pyridines

Even if the Hantzsch synthesis was developed over 100 years ago by Arthur Hantzsch, nowadays, the reaction is still being developed as it leads to the formation of 1,4-dihydropyridines (1,4-DHPs), which may exhibit antihypertensive and antianginal activity by blocking heart calcium channels [18,19].
One of the latest studies of this reaction was carried out by Rahman et al., who utilized an iron–silica nanocomplex to catalyse the combination of ethyl acetoacetate, various aldehydes, and ammonium acetate in ethanol under reflux conditions (Figure 5) [20].
Sulphated tin oxide (STO) was established as an alternative catalyst for the Hantzsch reaction, involving acetoacetanilide as the β-dicarbonyl compound (Figure 6). Yields ranged 90–94% when the reactions were performed at 80 °C in acetonitrile solvent for 2.5 h. Positively, the catalyst could be reused for up to five cycles with a three percent reduction in yield between the first and fifth trial [21].
Although 1,4-DHPs are the usually reported products of the Hantzsch synthesis, the 1,2-isomers can also be synthesized, as was successfully communicated by Bosica et al. In this synthetic study (Figure 7) [22], phosphotungstic acid (PW) was heterogenized on alumina before utilising it at room temperature to result in 1,2-DHPs yields ranging between 62 and 94%, obtained in 3.5–6 h. The catalyst could be reused for up to eight consecutive cycles with very minor loss of activity.
Instead of DHPs, there are also various studies which have managed to synthesize aromatized pyridines. A case in point, Khalifeh et al. developed recyclable copper nanoparticles supported on activated charcoal in order to form 2-amino-3-cyanopyridine derivatives under reflux conditions in the presence of air in acetonitrile solvent (Figure 8) [23]. Yields varied between 71 and 94% after stirring for 6–8 h. Impressively, the catalyst could be recycled for up to eight consecutive runs (six percent overall drop in yield).
In truth, the above reaction has been studied to great lengths, and amongst the catalysts that have been tried, there are also those listed in Table 1. In some instances, cyclic aliphatic ketones such as cyclohexanone have been also employed.
Similar in structure to the just-mentioned pyridines, 2-amino-3,5-dicarbonitrile-6-sulfanylpyridines have also been explored in terms of their synthesis and biological applications. As summarized in Table 2, nano-tin(IV) oxide, nano-titania, natural dolomitic limestone, borax, and a magnetite-supported Schiff base have all been used.

2.1.4. Pyridinones

Quite similar in structure are the class of compounds, pyridinones, which have been synthesized via an MCR employing Meldrum’s acid, during which decarboxylation and acetone loss take place (Figure 9). Interestingly, these compounds have been found to have antiurease activity (ureases are enzymes found ubiquitously in nature and are implicated in conditions such as peptic ulceration, stomach cancer, formation of urinary stones, etc.) [34]. The recoverable catalyst, silica-propylsulfonic acid, was used in neat conditions at 140 °C to obtain the products in 25–60 min in excellent yields (78–90%) [35].

2.1.5. Coumarins

The spectrum of applications of coumarins in everyday life is exceptionally broad and their uses vary from anticlotting agents to dyes in laser technology, anti-HIV compounds, and fluorescent indicators [36]. A common approach to their synthesis involves the combination of 4-hydroxycoumarin, carbonyl compounds, and active methylenes. Khodabakhsi et al. synthesized titanium dioxide nanowires to produce novel pyranocoumarins, also known as pyrano [2,3-c] chromenes, in ethanol/water solvent under reflux conditions (Figure 10) [37]. Yields were excellent and reaction times were relatively short.
A similar approach was taken by Kerru et al., whose zinc oxide catalyst, supported on fluorapatite, yielded furan-coumarin derivatives (Figure 11) at room temperature in ethanol solvent in less than 20 min [38].

2.1.6. Chromenes

2-Amino-4H-chromenes, i.e., derivatives of benzopyran, possess a wide range of biological properties, chief amongst them being their diuretic and anticoagulant activities, as well as being cognitive enhancers for the treatment of neurodegenerative diseases such as Parkinson’s and Alzheimer’s [39]. Amongst the several methods developed for their synthesis, various research papers have investigated the reaction between aldehydes, an activated methylene compound, and a nucleophile such as a naphthol or a phenol (Figure 12). One such study employed a piperazine-functionalized nickel–ferrite nanocatalyst in a neat reaction at 110 °C to obtain yields ranging between 81 and 95% in 40–60 min. In addition to 2-naphthol, other nucleophiles were tried, including 1-naphthol and 3,5-dihydroxytoluene, with yields for the former being in the range of 78–87% and for the latter in the range of 80–85% [40].
A noteworthy catalyst that was developed for the same reaction was prolinamide-functionalized polyacrylonitrile. The reactions involving 2-naphthol gave excellent yields (91–98%) at a lower temperature (80 °C, 1 h) and in water as solvent [41].
Other recent advances in this synthetic reaction are summarized in Table 3.

2.1.7. Xanthenes

Produced via a similar mechanism to chromenes, symmetrical xanthene derivatives are synthesized by the reaction of 2-naphthol or a β-dicarbonyl compound with aldehydes (Figure 13). Bansal et al. studied this reaction using copper sulphide quantum dots to obtain the products in excellent yields (88–95%) in a maximum of 15 min at 80 °C in neat conditions. Positively, the catalyst could be reused for up to five cycles [46].
A high temperature was required when the same reaction scope was studied using graphene oxide supported on copper ferrite. In addition, excellent yields were obtained (84–96%) in a maximum of 15 min [47].

2.2. Heterocycles Containing 2 Heteroatoms

2.2.1. Pyrazoles

Pyrazole is a 5-membered heterocyclic aromatic compound that is present in several notable drugs, principal amongst them being celecoxib (Celebrex®, a nonsteroidal anti-inflammatory drug–NSAID, Figure 14) [48].
Meanwhile, pyranopyrazoles are molecules with a pyran ring fused to a pyrazole ring. They exhibit a wide range of activities because they can act as Chk1 kinase inhibitors as well as acting as anticancer and antimicrobial agents [49,50]. One of the contemporary methodologies to synthesize dihydropyrano [2,3-c]pyrazole derivatives involves a 4CR between a β-dicarbonyl, a malononitrile, an aldehyde, and a hydrazine hydrate (Figure 15). A case in point is silver–titania nanofilms being utilized in ethanol/water solvent at 70 °C to collect products in a maximum of 55 min at yields of 75–93% [51].
An alternative method requires the combination of 3-methyl-1-phenyl-2-pyrazoline-5-one, aldehydes, and malononitrile (Figure 16). A recently developed catalyst, cobalt Schiff base complex immobilized on magnetic iron oxide nanoparticles, gave impressive results in very short reaction times. Moreover, the catalyst could be reused for up to eight runs with the yield decreasing from 95 to 88% for the chosen reaction [52].

2.2.2. Imidazoles

Isomeric with pyrazoles, imidazoles are heterocycles that have a 1,3-nitrogen arrangement instead of 1,2. The imidazole moiety is present in histidine (an α-amino acid), histamine (a hormone, Figure 17), and a sedative (midazolam), apart from other natural products and drugs [53].
A novel catalyst, composed of iron oxide and copper(I) oxide nanoparticles suspended on guarana, was used successfully for the reaction of aldehydes with benzyl to produce imidazole derivatives, specifically, 2,4,5-triaryl-1H-imidazoles (Figure 18). It is pertinent to note that the reaction was carried out under ultrasonication, leading to very short reaction times (20 min) and excellent yields (75–97%) [54].

2.2.3. Thiazolines

Thiazolines (also known as dihydrothiazoles), are penta-membered heterocyclic compounds containing both a sulphur and a nitrogen atom in their ring. A notable life example of a molecule consisting of such rings is luciferin, the compound which allows for the emission of light in fireflies. Vitamin B1 (Figure 19) has been reported to be an efficient catalyst in the formation of trisubstituted 1,3-thiazoles after having been supported on magnetite in a research study by Shaterian et al. The reaction entailed a one-pot combination of arylglyoxals monohydrate, cyclic 1,3-dicarbonyls, and thiobenzamides in aqueous conditions at 80 °C (Figure 20) [55].
Yields were superb (78–93%) and reactions took less than 45 min. Lastly, the catalyst could be easily removed by a magnet and reused for up to five times [55].
In a related approach, a glycine–proline–glutamine tripeptide was supported on silica-encapsulated-γ-iron(III) oxide nanoparticles to act as a catalyst in aqueous conditions. Impressively, reaction times were even lower (maximum of 15 min) and yields were generally slightly better (91–98%). Even better, the catalyst was reused 10 times, with the yield dropping from 98 to 91% for the chosen model reaction [56].

2.2.4. Thiazolidinones

The ketone analogues of thiazoles are the thiazolidinones and these are synthesized in a completely different manner from the above. One method involves the Mannich-type attack of a thiol followed by amide formation. An approach by Kumar et al. involved the use of supported perchloric acid on silica for the reaction of various aromatic aldehydes and amines with thioglycolic acid at 100 °C (Figure 21). Impressively, the study encompassed 43 reactant combinations and all yields were within the range of 70–87%. Despite this, toluene, a nonenvironmentally benign solvent, was required [57].
Recently, an alternative approach was taken in which, instead of using benzaldehyde, benzyl alcohol was utilised. The catalyst, cobalt–aluminium hydrotalcite, allowed for the air oxidation of the benzyl alcohol to the aldehyde before inducing the same reaction described above (Figure 22). All trials were concluded within 2 h in methanol solvent, with yields being in the range of 75–85%. It is well worth mentioning that in this study, unlike the previous one, no aliphatic amines were trialled and only 12 products were ultimately obtained [58].

2.2.5. Pyrimidines

The magnitude of the importance of pyrimidines in everyday life cannot be underlined enough. It suffices to say that pyrimidines are structural motifs found in nucleic acids such as thymine, cytosine, and uracil. In addition, there are several established drugs which have a pyrimidine nucleus, including 5-fluoruracil—an anticancer drug (Figure 23) [59,60].
In a novel method by Nimkar et al., guanidine hydrochloride, aldehydes, and acetophenone (Figure 24) were reacted in the presence of caesium hydroxide supported on γ-alumina after mixing for 1 h at room temperature and then refluxing for 4 h. Ethanol solvent was used, and yields varied from 72 to 80% [59].

2.2.6. Pyrimidinones

Discovered in approximately the same period as the Hantzsch synthesis, the Biginelli reaction provides 3,4-dihydropyrimidin-2-(1H)-ones (Figure 25). The latter have been investigated as potential Adenosine A2b receptor antagonists, ergo, they can be potentially used for the treatment of diabetes, asthma, and other inflammations [61].
Amongst the most recent catalysts employed for the reaction, those listed in the Table 4. are heterogeneous and recoverable in nature and gave good to excellent yields.
Ketones have also been explored in the synthesis of pyrimidinones, a version which is often described as being Biginelli-like. In fact, they are synthesized by the reaction between aldehydes, acetophenones, and urea (Figure 26).
Because most contemporary synthetic studies are looking into the possibility of using nanomaterials as catalysts, Safari et al. utilised titania–multiwalled carbon nanotube composites to drive the reactions [66]. Yields were excellent (80–98%) and reaction times were not longer than 35 min. Positively, the catalyst could be reused for five cycles with only a two percent drop in activity.

2.3. Fused Heterocycles

2.3.1. Indolizine

Indolizines are N-fused heterocycles with a significant level of pharmacological relevance owing to their antimicrobial, antileishmanial, antihistaminic, antitubercular, and herbicidal activity [67]. One simple method which can be employed to synthesize them involves the A3 coupling–cyclization tandem reaction of an alkyne, an amine, and 2-pyridinecarboxaldehyde (Figure 27). In fact, copper(I) iodide, supported on glucose-derived and hydrothermally obtained carbon spheres, was used as a catalyst in ethylene glycol or in neat conditions at 60 °C to obtain a wide range of indolizine derivatives (yields: 75–95%) [68].
The Lewis acid of choice was also copper iodide when the latter was supported on porous cross-linked poly(ethyleneamine)polysulfonamide gel to serve as a recyclable nanocatalyst. Most importantly, the reaction was also carried out in neat conditions and yields were very good (70–95% in 1–3 h). On the other hand, the number of conducted trials was only nine [69].

2.3.2. Benzofurans

Although quite diverse from the just-mentioned class of compounds, benzo[b]furans can be synthesized via a very similar route, replacing pyridine-2-carboxaldehyde with salicylaldehyde (Figure 28). Their pharmacological properties are just as diverse with, for instance, some exhibiting antitumour capabilities and others being used in cosmetic formulations [70,71]. Copper iodide nanoparticles were employed by Safaei-Ghomi et al. in the aforementioned reaction, which required water as solvent and an equivalent of potassium carbonate as a base. The reaction had to take place at 100 °C but was ready in 1–2 h to give products in excellent yields (82–95%) [72].
In a separate study by Purohit et al., it was remarkably established that, in the absence of potassium carbonate as base and using hierarchically porous copper(II) oxide (HS-CuO) as a heterogeneous catalyst, salicylaldehydes with electron-withdrawing groups gave benzo[b]furans, whereas salicylaldehyde itself gave dihydrobenzo[b]furans (Figure 29) [73].

2.3.3. Pyrimidobenzoimidazoles

Amongst recently synthesized antitumour agents are the pyrimido [1,2-b]benzoimidazoles, which have been found to exhibit activity against breast MCF7 cancer, and for some of which the compounds are equipotent to other drugs such as Tamoxifen® [74]. One of the diverse synthetic pathways that one can follow to obtain these compounds involves the combination of aldehydes, β-dicarbonyl compounds, and 2-aminobenzoimidazole (Figure 30). A novel catalyst composed of hafnium metal–organic frameworks drove the formation of several pyrimido [1,2-b]benzoimidazole derivatives to high yields (60–96%) in 3 h at 100 °C [75].
An alternative catalyst that was used was a Brønsted acidic ionic liquid supported on nanoporous sodium montmorillonite (Na+-MMT-[pmim][HSO4]). Impressively, the 22 reaction trials conducted were finalised in less than 50 min and gave yields between 87 and 95% [76].

2.3.4. Chromenopyridines

Replacing the aldehyde with salicylaldehyde for the previously mentioned reaction of an aldehyde, malononitrile, and thiophenol (Table 2) furnishes a diverse class of pyridine-3-carbonitriles, specifically called chromeno [2,3-b]pyridines, which have a chromene ring fused to the pyridine ring (Figure 31).
In a study by Safaei-Ghomi et al., amongst various other trialled metal oxides, tin(II) oxide was found to be the ideal one [29]. In fact, at only 6 mol%, tin(II) oxide yielded the products in 54–63 min in very good yields (80–88%, six examples). More so, it could be recycled five times with a total yield drop of six percent. The same research group synthesized citric acid-functionalized chitosan polymeric beads as catalysts for the reaction and obtained excellent yields (84–92%) in 35–40 min whilst refluxing in ethanol [77]. Furthermore, the catalyst could be reused for five consecutive reaction cycles with only a minor two percent yield drop registered between the first and fifth run.

2.3.5. Chromenopyrimidines

Delving deeper into the possibilities that chromene products demonstrate, one would encounter another pseudo-four-component reaction involving salicylaldehyde (two equivalents), malononitrile, and amines or alcohols to yield chromeno [2,3-d]pyrimidine derivatives (Figure 32). A summary of some of the recent results is tabulated in Table 5.

2.4. Spiro Compounds

A spiro compound is one which has at least two rings joined by one common atom. In the literature, there are a plethora of reported syntheses of such compounds and amongst these there are methods involving the formation of spirooxindole derivatives. Interestingly, in a study by Liang et al., the enzyme papain was effectively used to catalyse the reaction between isatin, cyclic-1,3-diketones, and 3-methyl-5-aminopyrazole in ethanol solvent at 50 0C (Figure 33). Even if no recycling trials were performed, the enzyme did not dissolve and could be recovered by filtration. Although ultimately the obtained yields were poor to average (27–71%), and reaction times were comparatively long (48 h), no metals or toxic chemicals were needed in the creation of a catalyst, and papain itself is easily extracted from the raw fruit of the papaya plant [82].
Meanwhile, Arya et al. developed a ZSM-5 zeolite-supported Brønsted acid ionic liquid to catalyse the reaction of isatins, 2-mercaptopyridine-3-carboxylic acid, and various amines under ultrasonication (Figure 34) [83]. Reactions were very fast and yielded excellent results (80–92%). Meanwhile, conventional heating was also tried but yields were poor (0–35% in 46–50 h).
The nanocatalyst alum–silica was used to prepare spirooxindole derivatives via β-ketoesters, malononitrile, and isatins in ethanol solvent under reflux conditions (Figure 35) [84].
Recently, an attractive approach was undertaken by Sadeghi et al., who developed nano-sulfonic acid-derivatised coconut shell powder to catalyse the just-mentioned reaction to obtain products (12 examples, 87–95%) in ethanol solvent after refluxing for only 15 min [85].

3. Acyclic Multicomponent Reactions

3.1. Mannich-Type Reactions

The classic Mannich reaction involves a C-C bond formation after the reaction of an aldehyde and an amine (to form an imine) with a resonance-stabilised carbon nucleophile [86]. With time, several variants have been discovered and established by replacing the carbon nucleophiles with others.
Two related variants of such aforementioned reactions are the A3 and KA2 coupling reactions, which involve the combination of an aldehyde (or ketone for the KA2), a primary/secondary amine, and a terminal alkyne. The diversity of catalysts tried for this reaction is vast, to say the least. Amongst these, copper(I) has been employed in various forms and supported on numerous heterogeneous supports. For example, Bosica et al. used copper(I) iodide supported on Amberlyst A21 polymeric beads to catalyse the reaction between various secondary aliphatic amines, aromatic aldehydes, and different aromatic and silyl terminal alkynes in neat conditions at 100 °C in 2–3 h, with yields ranging from 70 to 98% (Figure 36). Positively, the catalyst could be reused for up to six cycles [87].
The same catalyst was used in a separate study to instead investigate its feasibility in reactions involving ketones (KA2). Yields were overall lower, but still acceptable considering the relative unreactivity of ketones. When the alkyne used was phenylacetylene, yields ranged between 62 and 98% in 1.5–48 h at 98 °C. The use of 1-octyne resulted in poorer yields (49–81% in 2–15 h) [88]. Metal–organic frameworks (MOFs), porous hybrid materials composed of metals ions linked together by organic molecules, are other versatile catalysts or supports being developed in the field of heterogeneous chemistry. In fact, a silver(I)-functionalised metal–organic framework (MIL-101-SO3Ag) was demonstrated to be an efficient and environmentally friendly catalyst in the A3 coupling reaction, giving the products in excellent yields (>92%) after 4–10 h, under solvent-free conditions at 100 °C [89].
If the alkyne is replaced with an electron-rich alcohol, i.e., 2-naphthol, the Mannich-type reaction is termed as the Betti synthesis, and the products are specifically 1-aminoalkyl-2-naphthols (Figure 37). For instance, nano-silica boric acid was utilised in a neat reaction environment at 40 °C to drive the reaction in 30–40 min, and to give the products in excellent yields (88–94%). Furthermore, the catalyst was recyclable for up to five runs, with yield decreasing by 5–10% between the first and fifth trial [90].
This reaction can also be easily catalysed by the commercially available Montmorillonite K30 clay catalyst. Reactions were conducted at 60 °C in neat conditions. The use of proline as an amine resulted in yields ranging from 50 to 91% in 2–6 h. Morpholine and piperidine also worked to give the products in 4–12 h, at yields of 86–90% [91]. It is also possible to use amides instead of amines, as Lei et al. did when they succeeded in making use of graphite-supported perchloric acid (Figure 38). The reaction required higher temperatures (125 °C) than the amine-analogues due to the relative unreactivity of amides. Just the same, yields were moderate to very good (68–88%) and products were collected after 2 h of reaction time [92].
The heteroaromatic compound, indole, behaves similar to 2-naphthol and carries out the so-called aza-Friedel–Crafts reactions to form 3-substituted indoles. Ganesan et al. reported the use of magnetite-sulfonic acid for the synthesis of aza-Friedel–Crafts products, using either secondary or tertiary amines (Figure 39). In the case of the former, the reactions were performed in neat conditions at room temperature for 3 h to collect products originating from a C-N bond forming in very good yields (78–90%, four examples only). On other hand, tertiary amines required a higher temperature (100 °C, 1 h) and gave products that resulted only from C-C bond formation (82–88%) [93]. Interestingly, as reported by Bosica et al., the products which were obtained from primary, secondary, and tertiary aromatic amines were all only via C-C bond formation. The catalyst of choice was silicotungstic acid supported on Amberlyst 15, sulfonic acid-functionalized polymeric beads. The reactions involving secondary aromatic amines (25 examples) took 4–24 h, with yields of 40–99%. Primary amines gave products in 6 h at yields of 41–84%, (18 examples). Lastly, tertiary amines performed poorly, and heating was needed (33–47%,18–24 h) [94].
Another member of the Mannich-type reaction is the aza-Henry, or nitro-Mannich reaction, which involves the combination of aldehydes, amines, and nitroalkanes (Figure 40) [95]. The use of copper(I) iodide supported on Amberlyst A21 was reported for such a reaction, and did not require the addition of other solvents (even if nitromethane was used in large excess to act as a reactant and solvent simultaneously). Reaction times varied between 2 and 10 h, with yields occupying the range of 38–98% (in truth, only one result was below 50% in 27 trials).
Even if not encountered in the final product itself, boron, or more precisely boronic acids, are involved reactions falling within the same reaction family. The borono–Mannich reaction (also known as the Petasis reaction) entails the reaction of an aldehyde and an amine to form an imine that gets attacked by an alkyl/aryl boronic acid which transfers an alkyl/aryl group to it (Figure 41). Recently, cobalt ferrite was selected to catalyse the reaction in acetonitrile solvent at 80 °C for 2.5–6 h. Yields were very good (72–90%) and it was observed that the catalyst could be reused for up to five times with no apparent decrease in activity [96].
One of the most important routes to synthesize α-aminophosphonates is the Kabachnik–Fields reaction which, similar to the above, requires the formation of an imine before the latter gets attacked by dialkyl phosphites [97]. In a novel approach by Sobhani et al., phosphoric acid was supported on magnetic nanoparticles coated with silica (γ-Fe3O4@SiO2-PA) to obtain products in 1–4 h (85–96% yields, Figure 42). All reactions were performed in water at 80 °C and at the end of each reaction, the catalyst could be easily separated by a magnet before being reused for up to a total of five times.
β-Aminocarbonyl compounds are useful building blocks in the synthesis of pharmaceuticals and natural products such as β-amino acids. They can also be synthesized via the classical Mannich reaction, in which an aldehyde and an amine react together to form the imine, which then gets attacked by acetophenone (Figure 43). Amongst the wide range of catalysts attempted and tested for this reaction, recently, nano-sulphated zirconia was also discovered to give very good results (71–88%) at room temperature in ethanol solvent in 4.5–5 h [98].
Liu et al. managed to obtain similar products using cyclohexanone instead of acetophenone via the use of silica-supported phenylphosphinic acid in ethanol solvent at room temperature. Reactions were much faster (30–70 min), and yields were similarly high (67–96%) [99].
Lastly, if the imine-attacking nucleophile is cyanide, the end products of the Mannich-type reaction are specifically called α-aminonitriles (Figure 44). Whereas, in the past, the latter were obtained by the use of ionic cyanides, nowadays, alternative cyanide sources, such as trimethylsilyl cyanide (TMSCN), are sought [100]. One novel and heterogeneous approach established in recent years requires the use of nano-copper ferrite [101]. The reactions were performed at room temperature in ethanol/water (50:1) solvent to give the end products in 50–81 min at yields of 72–98%.

3.2. Knoevenagel-Type Reactions

The classical Knoevenagel condensation reaction is the nucleophilic addition of an active methylene-containing compound to a carbonyl-containing compound followed by condensation to form an α,β-unsaturated derivative [102].
With time, several Knoevenagel-type reactions have been established, which involve first the formation of an α,β-unsaturated carbonyl compound that could then be attacked by a nucleophile. For example, zeoltic imidazolate frameworks were found to be worthy catalysts for the Knoevenagel–phospha-Michael reaction (as shown in Figure 45), whereby the dialkyl phosphite (HPO(OEt)2) attacks the Knoevenagel product formed earlier from the aldehyde and malononitrile. In the six trials carried out, the products were obtained after 5 h (at 50 °C in 1,4-dioxane solvent in air) at yields of 98–100%. When malononitrile was replaced by ethyl cyanoacetate, the reaction time increased to 10 h but the yield remained at 98% [103].
3-Aminopropylated silica is a recently developed, efficient, and reusable catalyst that was also put into use for the synthesis of bis(pyrazolyl)methanes via a three-component tandem Knoevenagel–Michael reaction of 1-phenyl-3-methylpyrazolone with aldehydes (Figure 46) [104]. The formation of these products was very rapid (<45 min) and efficient (90–98% yields at room temperature).

3.3. Isocyanide-Type

The wide range of starting materials that can be used in isocyanide-based MCRs makes it one of the most researched families of reactions. There are various ways in which the isocyanide derivatives can react, but the most important one is the α-addition of electrophiles and nucleophiles at the isocyanide carbon [105]. The earliest studied reaction that falls into this category is the Passerini three-component reaction (3CR) between an aldehyde, a carboxylic acid, and an isocyanide to yield α-acyloxycarboxamides (Figure 47). A recently developed and successfully employed heterogeneous catalyst for the reaction is silica gel-immobilised 1,2-benzenedisulfonimide [106]. This catalyst gave the products in excellent yields (78–93%) in 20–30 h and, most importantly, at an enantiomeric excess, surpassing 92.7% in all trials. The carboxylic acids that were used were both aromatic and aliphatic, the aldehydes were mainly aromatic in nature, and the isocyanides were either t-butyl or naphthyl isocyanides. The catalyst could also be reused for five consecutive runs with a 3% reduction in yield.
Later on, it was postulated that if the aldehyde were to be replaced by an imine that can itself be formed in situ by the reaction of an amine and an aldehyde, the reaction mechanism should be followed in a similar manner to that of the Passerini reaction and result in similar products. In fact, the reaction worked to form peptide-like structures and became known as the Ugi reaction (first reported in 1959) [107]. A modern heterogeneous catalyst that could be used was fluorite which, in the presence of microwaves and ethanol solvent, yielded 2-chloroquinoline-3-carbaldehyde derivatives in excellent yields (Figure 48) [108].

4. Conclusions

As can be discerned, the diversity of multicomponent reactions is particularly broad, owing to the fact that if one reactant is replaced by one or more others, a completely new set of products will end up being produced, thus confirming the high versatility of this useful synthetic approach. All the catalysts that were employed in the selected multicomponent reactions could be reused at the end of the reaction, making them operationally heterogeneous. In mechanistic terms, some were homotopic and others were heterotopic, but few studies actually delved into such details. This is because the determination of the catalyst functionality needs to be coupled with detailed mechanistic studies which, in truth, fall outside of the scope of the studies for which the aim was simply to obtain products in an environmentally benign manner.

Author Contributions

The review has been cowritten by G.B. and R.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors thank the University of Malta for financial and scientific knowledge support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anastas, P.; Eghbali, N. Green chemistry: Principles and practice. Chem. Soc. Rev. 2010, 39, 301–312. [Google Scholar] [CrossRef] [PubMed]
  2. Bosica, G. Heterogeneous Catalysis. In Sustainable Organic Synthesis: Tools and Strategies; Protti, S., Palmieri, A., Eds.; Royal Society of Chemistry: London, UK, 2021; Chapter 3; pp. 45–67. [Google Scholar] [CrossRef]
  3. Gartner, D.; Sandl, S.; Wangelin, A. Homogeneous vs. heterogeneous: Mechanistic insights into iron group metal-catalyzed reductions from poisoning experiments. J. Cat. Sci. Tech. 2020, 11, 3502–3514. [Google Scholar] [CrossRef]
  4. Crabtree, R.H. Resolving heterogeneity problems and impurity artifacts in operationally homogeneous transition metal catalysts. Chem. Rev. 2011, 11, 1536–1554. [Google Scholar] [CrossRef] [PubMed]
  5. Singh, M.S.; Chowdhury, S. Recent developments in solvent-free multicomponent reactions: A perfect synergy for eco-compatible organic synthesis. RSC Adv. 2012, 2, 4547–4592. [Google Scholar] [CrossRef]
  6. Domling, A.; Wang, W.; Wang, K. Chemistry and biology of multicomponent reactions. Chem. Rev. 2012, 112, 3083–3135. [Google Scholar] [CrossRef] [Green Version]
  7. Sheldon, R.A. The E factor 25 years on: The rise of green chemistry and sustainability. Green Chem. 2017, 19, 18–43. [Google Scholar] [CrossRef]
  8. Bhagat, S.; Shah, P.; Garg, S.K.; Mishra, S.; Kaur, P.K.; Singh, S.; Chakraborti, A.K. α-Aminophosphonates as novel anti-leishmanial chemotypes: Synthesis, biological evaluation, and CoMFA studies. Med. Chem. Commun. 2014, 5, 665–670. [Google Scholar] [CrossRef]
  9. Borazjani, N.; Sepehri, S.; Behzadi, M.; Jarrahpour, A.M.; Rad, J.A.; Sasanipour, M.; Mohkam, M.; Ghasemi, J.; Akbarizadeh, A.R.; Digiorgio, C.; et al. Three-component synthesis of chromeno β-lactam hybrids for inflammation and cancer screening. Eur. J. Med. Chem. 2019, 179, 389–403. [Google Scholar] [CrossRef]
  10. Patil, R.; Ghosh, A.; Cao, P.S.; Sommer, R.D.; Grice, K.A.; Waris, G.; Patil, S. Novel 5-arylthio-5H-chromenopyridines as a new class of anti-fibrotic agents. Bioorg. Med. Chem. Lett. 2017, 27, 1129–1135. [Google Scholar] [CrossRef]
  11. Shi, F.; Zeng, X.; Zhang, G.; Ma, N.; Jiang, B.; Tu, S. Facile synthesis of new 4-aza-podophyllotoxin analogs via microwave-assisted multi-component reactions and evaluation of their cytotoxic activity. Bioorg. Med. Chem. Lett. 2011, 21, 7119–7123. [Google Scholar] [CrossRef]
  12. Ahmed, K.A.; El-Hennawi, H.M.; Shahin, A.A.; Ragheb, A.A. Preparation of nano disperse dye based on benzopyran in one pot reaction using microwave irradiation and its appliance in textile printing. SN Appl. Sci. 2019, 1, 1009. [Google Scholar] [CrossRef] [Green Version]
  13. John, S.E.; Gulati, S.; Shankaraiah, N. Recent advances in multi-component reactions and their mechanistic insights: A triennium review. Org. Chem. Front. 2021, 8, 4237–4287. [Google Scholar] [CrossRef]
  14. Li, B.; Zhang, M.; Hu, H.; Du, X.; Zhang, Z. Nano-CoFe2O4 supported molybdenum as an efficient and magnetically recoverable catalyst for a one-pot, four-component synthesis of functionalize pyrroles. New J. Chem. 2014, 38, 2435–2442. [Google Scholar] [CrossRef]
  15. Nandeesh, K.N.; Raghavendra, M.; Revanna, C.N.; Vijay, T.A.J.; Rangappa, K.S.; Mantelingu, K. Recyclable, graphite-catalyzed, four-component synthesis of functionalized pyrroles. Syn. Commun. 2014, 44, 1103–1110. [Google Scholar] [CrossRef]
  16. Raczko, J.; Jurczak, J. Furan in the synthesis of natural products. Stud. Nat. Prod. Chem. 1995, 16, 639–685. [Google Scholar]
  17. Ghomi, J.; Heidari-Baghbahadorani, E.; Shahbazi-Alavi, H. SnO nanoparticles: A robust and reusable heterogeneous catalyst for the synthesis of 3,4,5-substituted furan-2(5H)-ones. Chem. Mon. 2014, 146, 181–186. [Google Scholar] [CrossRef]
  18. Hantzsch, A.; Justus, L. On the synthesis of pyridine-like compounds from acetoacetic ether and aldehyde ammonia. Ann. Cheime. 1882, 215, 1–82. [Google Scholar] [CrossRef] [Green Version]
  19. Cosconati, S.; Marinelli, L.; Lavecchia, A.; Novellino, E. Characterizing the 1,4-dihydropyridines binding interactions in the L-Type Ca2+ channel: Model construction and docking calculations. J. Med. Chem. 2007, 50, 1504–1513. [Google Scholar] [CrossRef]
  20. Rahman, A.; Nehemia, P.N.; Nyambe, M.M. An efficient method for the synthesis of dihydropyridine by Hantzsch reaction with Fe/SiO2 nano heterogeneous catalysts. Bul. Chem. Rea. Eng. Cat. 2020, 15, 617–630. [Google Scholar] [CrossRef]
  21. Koduri, R.G.; Pagadala, R.; Varala, R.; Boodida, S. An effective process for the synthesis of dihydropyridines via SO42−/SnO2. J. Chi. Chem. Soc. 2020, 68, 333–337. [Google Scholar] [CrossRef]
  22. Bosica, G.; Demanuele, K.; Padron, J.M.; Puerta, A. One-pot multicomponent green Hantzsch synthesis of 1,2-dihydropyridine derivatives with antiproliferative activity. Beilstein J. Org. Chem. 2020, 16, 2862–2869. [Google Scholar] [CrossRef]
  23. Khalife, R.; Ghamari, M. A multicomponent synthesis of 2-amino-3-cyanopyridine derivatives catalyzed by heterogeneous and recyclable copper nanoparticles on charcoal. J. Braz. Soc. 2016, 27, 759–768. [Google Scholar] [CrossRef]
  24. Heravi, M.M.; Beheshtiha, S.Y.S.; Dehghani, M.; Hosseintash, N. Using magnetic nanoparticles Fe3O4 as a reusable catalyst for the synthesis of pyran and pyridine derivatives via one-pot multicomponent reaction. J. Iran. Chem. Soc. 2015, 12, 2075–2081. [Google Scholar] [CrossRef]
  25. Yahyazadeh, A.; Abbaspout-Gilandeh, E.; Aghaei-Hasjin, M. Four-component synthesis of 2-amino-3-cyanopyridine derivatives catalyzed by Cu@ImineZCMNPs as a novel, efficient and simple nanocatalyst under solvent-free conditions. Catal. Lett. 2018, 184, 1214–1262. [Google Scholar] [CrossRef]
  26. Afradi, M.; Pour, S.A.; Dolat, M.; Yazhdani-Elah-Abadi, A. Nanomagnetically modified vitamin B3 (Fe3O4@Niacin): An efficient and reusable green biocatalyst for microwave-assisted rapid synthesis of 2-amino-3-cyanopyridines in aqueous medium. Appl. Organometal. Chem. 2018, 32, e4103. [Google Scholar] [CrossRef]
  27. Challa, C.S.; Katari, N.K.; Nallanchakravarthula, V.; Nayakanti, D.; Kapavarapu, R.; Pal, M. Amberlyst-15 catalysed sonochemical synthesis of 2-amino-4,6-disubstituted nicotinonitrile derivatives and their biological evaluation. J. Mol. Struct. 2021, 1240, 130541–130546. [Google Scholar] [CrossRef]
  28. Achagar, R.; Elmakssoudi, A.; Thoume, A.; Mohamed, D.; Elamrani, A.; Zouheir, Y.; Zahouily, M.; Ait-Touchente, Z.; Jamaleddine, J.; Chehimi, M.M. Nanostructured Na2CaP2O7: A new and efficient catalyst for one-pot synthesis of 2-amino-3-cyanopyridine derivatives and evaluation of their antibacterial activity. Appl. Sci. 2022, 12, 5487. [Google Scholar] [CrossRef]
  29. Safaei, J.; Shahbazi-Alavi, H.; Baghbahadorani, E. SnO nanoparticles as efficient catalyst for the one-pot synthesis of chromenp[2,3-b]pyridines and 2-amino-3,5-dicyano-6-sulfanyl pyridines. RSC Adv. 2014, 4, 50668–50677. [Google Scholar] [CrossRef]
  30. Molla, A.; Hussain, S. Borax catalyzed domino reactions: Synthesis of highly functionalised pyridines, dienes, anilines and dihydropyrano[3,2-c]chromenes. RSC Adv. 2014, 4, 29750–29758. [Google Scholar] [CrossRef] [Green Version]
  31. Kumari, S.; Shekhar, A.; Pathak, D.D. Graphene oxide–TiO2 composite: An efficient heterogeneous catalyst for the green synthesis of pyrazoles and pyridines. New J. Chem. 2016, 40, 5053–5060. [Google Scholar] [CrossRef]
  32. Godugo, K.; Yadala, V.D.; Pinjari, M.K.M.; Gundala, T.R.; Sanapareddy, L.R.; Nallagondu, C.G.R. Natural dolomitic limestone-catalysed synthesis of benzimidazoles, Dihydropyrimidinones and highly substituted pyridines under ultrasound irradiation. Beilstein J. Org. Chem. 2020, 16, 1881–1900. [Google Scholar] [CrossRef] [PubMed]
  33. Ebrahimiasl, H.; Azarifar, D.; Rakthshah, J.; Keypour, H.; Mahmoudabadi, M. Application of novel and reusable Fe3O4@CoII(macrocyclic Schiff base ligand) for multicomponent reactions of highly substituted thiopyridine and 4H-chromene derivatives. Appl. Organomet. Chem. 2020, 34, 769–787. [Google Scholar] [CrossRef]
  34. Bektas, H.; Ceylan, S.; Demirbas, N.; Alpay-Karaoglu, S.; Sokmen, B.B. Antimicrobial and antiurease activities of newly synthesized morpholine derivatives containing an azole nucleus. Med. Chem. Res. 2013, 22, 3629–3639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hakimi, A.M.; Lashgari, N.; Mahernia, S.; Ziarani, G.M.; Amanlou, M. Facile one-pot four component synthesis of 3,4-dihydro-2-pyridone derivatives: Novel urease inhibitor scaffold. Res. Pharm. Sci. 2017, 12, 353–363. [Google Scholar]
  36. Dittmer, D.C.; Li, Q.; Avilov, D.V. Synthesis of coumarins, 4-hydroxycoumarins and 4-hydroxyquinolinones by Tellurium-Triggered Cyclizations. J. Org. Chem. 2005, 70, 4682–4686. [Google Scholar] [CrossRef]
  37. Khodabakshi, S.; Karami, B.; Eskandari, K.; Hoseini, S.J. Titanium dioxide nanowires as green and heterogeneous catalysts for the synthesis of novel pyranocoumarins. C. R. Chim. 2014, 17, 35–40. [Google Scholar] [CrossRef]
  38. Kerru, N.; Gummidi, L.; Gangu, K.K.; Maddila, S.; Jonnalagadda, B. Synthesis of novel furo[3,2-c]coumarin derivatives through multicomponent [4+1] cycloaddition reaction using ZnO/Fap as a sustainable catalyst. Chem. Sel. 2020, 5, 4104–4110. [Google Scholar] [CrossRef]
  39. Foye, W.O. Principi di Chimica Farmaceutica; Piccin: Padova, Italy, 1991; p. 416. [Google Scholar]
  40. Bahram, P.; Ahmad, M. Piperazine-functionalized nickel ferrite nanoparticles as efficient and reusable catalysts for the solvent-free synthesis of 2-amino-4H-chromenes. J. Chin. Chem. Soc. 2019, 66, 1356–1362. [Google Scholar]
  41. Zhu, H.; Xu, G.; Du, H.; Zhang, C.; Ma, N.; Zhang, W. Prolinamide functionalized polyacrylonitrile fiber with tunable linker length and surface microenvironment as efficient catalyst for Knoevenagel condensation and related multicomponent tandem reactions. J. Cat. 2019, 374, 217–229. [Google Scholar] [CrossRef]
  42. Kiasat, A.R.; Hamid, S.; Saghanezhad, S.J. Bipyridinium chloride supported rice husk silica: An efficient nanocomposite for the one-pot preparation of spirooxindole pyran and 2-amino-4H chromene derivatives. Rev. Roum. Chim. 2019, 64, 927–934. [Google Scholar] [CrossRef]
  43. Neela, A.; Clarina, T.; Rama, V. Nickel oxide-catalyzed synthesis of 4-amino-2H-chromenes: Its application in antimicrobial studies and towards protein docking. Chem. Asian J. 2019, 31, 1049–1056. [Google Scholar] [CrossRef]
  44. Khazee, A.; Jahanshahi, R.; Sobhani, S.; Skibsted, J.; Sansano, J.M. Immobilized piperazine on the surface of graphene oxide as a heterogeneous bifunctional acid-base catalyst for the multicomponent synthesis of 2-amino-3-cyano-4H-chromenes. Green Chem. 2020, 22, 4604–4616. [Google Scholar] [CrossRef]
  45. Chahkamali, F.O.; Sobhani, S.; Sansano, J.M. A novel base-metal multifunctional catalyst for the synthesis of 2-amino-3-cyano-4H-chromenes by a multicomponent tandem oxidation process. Sci. Rep. 2022, 12, 2867–2886. [Google Scholar] [CrossRef]
  46. Bansal, P.; Chaudhary, G.R.; Kaur, N.; Mehta, S.K. An efficient and green synthesis of xanthenes derivatives using CuS quantum dots as heterogeneous and reusable catalyst under solvent free condition. RSC Adv. 2015, 5, 8205–8209. [Google Scholar] [CrossRef]
  47. Kumar, A.; Rout, L.; Acahry, L.S.K.; Dhaka, R.S.; Dash, P. Greener route for synthesis of aryl and alkyl-14H-dibenzo[a.j]xanthenes using graphene-oxide copper ferrite nanocomposite as a recyclable heterogeneous catalyst. Sci. Rep. 2017, 7, 42975. [Google Scholar] [CrossRef]
  48. Chakraborty, H.; Grihanjali Devi, P.; Sarkar, M.; Dasgupta, D. Chapter 1—New functions of old drugs: Aureolic acid group of anti-cancer antibiotics and non-steroidal anti-inflammatory drugs. In Recent Advances in Medicinal Chemistry; Attaur-Rahman, Choudhary, M.I., Perry, G., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 3–55. [Google Scholar]
  49. Saleh, N.M.; El-Gazzar, M.; Aly, H.M.; Othman, R.A. Novel anticancer fused pyrazole derivatives as EGFR and VEGFR-2 dual TK inhibitors. Front. Chem. 2019, 7, 917–928. [Google Scholar] [CrossRef]
  50. Paul, S.; Pradhan, K.; Ghosh, S.; Das, A.R. Uncapped SnO2 qunatum dot catalyzed cascade assembling of four components: A rapid and green approach to the pyrano[2,3-c]pyrazole and spiro-2-oxindole derivatives. Tetrahedron 2014, 70, 6088–6099. [Google Scholar] [CrossRef]
  51. Fatahpour, M.; Noori Sadeh, F.; Hazeri, N.; Maghsoodlou, M.T.; Hadavi, M.S.; Mahnaei, S. Ag/TiO2 nano-thin films as robust heterogeneous catalyst for one-pot, multi-component synthesis of bis (pyrazol-5-ol) and dihydropyrano[2,3-c]pyrazole analogs. J. Saud. Chem. 2017, 21, 998–1006. [Google Scholar] [CrossRef]
  52. Shaabani, B.; Maleki, H.; Rakhtshah, J. Environmentally benign synthesis of pyranopyrazole derivatives by cobalt Schiff-base complexes immobilized on magnetic iron oxide nanoparticles. J. Org. Chem. 2019, 897, 139–147. [Google Scholar] [CrossRef]
  53. Pzharskii, A.F.; Soldatenkov, A.T.; Katritzky, A. Heterocycles in Life and Society: An Introduction to Heterocyclic Chemistry, Biochemistry and Applications; John Wiley & Sons, Ltd.: Chichester, United Kingdom, 2011. [Google Scholar]
  54. Varzi, Z.; Esmaeili, M.S.; Taheri-Ledari, R.; Maleki, A. Facile synthesis of imidazoles by an efficient and eco-friendly heterogeneous catalytic system constructed of Fe3O4 and Cu2O nanoparticles, and guarana as a natural basis. Inorg. Chem. Commun. 2021, 125, 108465–108475. [Google Scholar] [CrossRef]
  55. Shaterian, H.R.; Molaei, P. Fe3O4@vitamin B1 as a sustainable superparamagnetic heterogeneous nanocatalyst promoting green synthesis of trisubstituted 1,3-thiazole derivatives. App. Organomet. Chem. 2019, 33, 4964–4976. [Google Scholar] [CrossRef]
  56. Arabpoor, Z.; Shaterian, H.R. Synthesis of trisubstituted 1,3-thiazoles using Gly-Pro-Glu (tripeptide) supported on superparamagnetic silica-encapsulated γ-Fe2O3 nanoparticles through efficient multi-component reaction in water. Polycycl. Aromat. Compd. 2020, 1, 2247–2262. [Google Scholar] [CrossRef]
  57. Chakraborti, A.; Kumar, D.; Sonawane, M.; Pujala, B.W.; Jain, V.K. Supported protic acid-catalyzed synthesis of 2,3-disubstituted thiazolidin-4-ones: Enhancement of the catalytic potential of protic acid by adsorption on solid supports. Green Chem. 2013, 15, 2872–2883. [Google Scholar]
  58. Subhedar, D.D.; Yadavm, P.A.; Pawar, S.R.; Bhange, B.M. Environmentally benign synthesis of 4-thiazolidinone derivatives using a Co/Al hydrotalcite as heterogeneous catalyst. Cat. Lett. 2020, 151, 1776–1787. [Google Scholar] [CrossRef]
  59. Nimkar, A.; Ramana, M.M.V.; Betkar, R.; Ranade, P.; Mundhe, B. CsOH/γ-Al2O3: A heterogeneous reusable basic catalyst for one pot synthesis of 2-amino-4,6-diaryl pyrimidines. New J. Chem. 2016, 40, 2541–2546. [Google Scholar] [CrossRef]
  60. Wisniewska-Jarosinska, M.; Sliwinski, T.; Kasznicki, J. Cytotoxicity and genotoxicity of capecitabine in head and neck cancer and normal cells. Mol. Biol. Rep. 2011, 38, 3679–3688. [Google Scholar] [CrossRef] [Green Version]
  61. Crespo, A.; El Maatougui, A.; Biagini, P.; Azuaje, J.; Coelho, A.; Brea, J.; Loza, M.I.; Cadavid, M.I.; Garcia-Mera, X.; Gutierrez-de-Teran, H.; et al. Discovery of 3,4-dihydropyrimidin-2(1H)-ones as a novel class of potent and selective A2B adenosine receptor antagonists. ACS Med. Chem. Lett. 2013, 4, 1031–1036. [Google Scholar] [CrossRef] [Green Version]
  62. Do Nascimento, L.G.G.; Dias, I.M.; de Souza, G.B.M.; Dancini-Pontes, I.; Fernandes, N.R.C.; de Souza, P.S.; de Oliveira, G.R.R.; Alonso, C.G. Niobium oxides as heterogeneous catalysts for Biginelli multicomponent reaction. J. Org. Chem. 2020, 85, 11170–11180. [Google Scholar] [CrossRef]
  63. Patel, H.A.; Sawant, A.M.; Rao, V.J.; Patel, A.L.; Bedekar, A.V. Polyaniline-supported FeCl3: An effective heterogeneous catalyst for Biginelli reaction. Catal. Lett. 2017, 147, 2306–2312. [Google Scholar] [CrossRef]
  64. Al-Wahaibi, L.H.; Abdelhamid, A.A.; Saber, S.A.; Elkanzi, N.A.A.; Ali, A.M. Pumice as a novel natural heterogeneous catalyst for synthesis of 3,4-dihydropyrimidine-2-(1H)-ones/thiones via Biginelli reaction under solvent-free conditions. Preprints 2021, 2021050493. [Google Scholar] [CrossRef]
  65. Bosica, G.; Cachia, F.; De Nittis, R.; Mariotti, N. Efficient one-pot syntehsis of 3,4-dihydropyrimidin-2(1H)-ones via a three-component Biginelli reaction. Molecules 2021, 26, 3753. [Google Scholar] [CrossRef]
  66. Safari, J.; Gandomi-Ravandi, S. A novel protocol for solvent-free synthesis of 4,6-diaryl-3,4-dihydropyridimine-2(1H)-ones catalysed by metal oxide-MWCNTs nanocomposites. J. Mol. Struct. 2014, 1074, 71–78. [Google Scholar] [CrossRef]
  67. Sharma, V.; Kumar, V. Indolizine: A biologically active moiety. Med. Chem. Res. 2014, 23, 3593–3606. [Google Scholar] [CrossRef]
  68. Rajesh, U.C.; Purohit, G.; Rawat, D.S. One-pot synthesis of aminoindolizines and chalcones using CuI/CSP nanocomposites with anomalousselectiity under green conditions. ACS Sustain. Chem. Eng. 2015, 3, 2397–2404. [Google Scholar] [CrossRef]
  69. Solgi, S.; Ghorbani-Vaghei, R.; Alavinta, S. Application of copper iodide nanoparticles immobilized porous polysulfonamide gel as an effective nano catalyst for the synthesis of amindolizines. J. Por. Mater. 2020, 28, 289–298. [Google Scholar] [CrossRef]
  70. Baraldi, P.G.; Romagnoli, R.; Beria, I.; Cozzi, P.; Geroni, C.; Mongelli, N.; Bianchi, N.; Mischiati, C.; Gambari, R. Synthesis and antitumor activity of new benzoheterocyclic derivatives of distamycin A. J. Med. Chem. 2000, 43, 2675–2684. [Google Scholar] [CrossRef]
  71. Leung, A.Y.; Foster, S. Encyclopedia of Common Natural Ingredients Used in Food, Drugs, and Cosmetics; Wiley: New York, NY, USA, 1996. [Google Scholar]
  72. Safaei-Ghomi, J.; Ali Ghasemzadeh, M.; Kakavand-Qalenoei, A. CuI-nanoparticles-catalyzed one-pot synthesis of benzo[b]furans via three-component coupling of aldehydes, amines and alkyne. J. Saud. Chem. Soc. 2016, 20, 502–509. [Google Scholar] [CrossRef] [Green Version]
  73. Purohit, G.; Rajesh, U.C.; Rawat, D.S. Hierarchically porous sphere-like copper oxide (HS-CuO) nanocatalyzed synthesis of benzofuran isomers with anomalous selectivity and their ideal green chemistry metrics. ACS Sustain. Chem. Eng. 2017, 5, 6466–6477. [Google Scholar] [CrossRef]
  74. Dawood, D.H.; Abbas, E.M.H.; Farghaly, T.A.; Ali, M.M.; Ibrahim, M.F. ZnO nanoparticles catalyst in the synthesis of bioactive fused pyrimidines as anti-breast cancer agents targeting VEGFR-2. Med. Chem. 2019, 15, 277–286. [Google Scholar] [CrossRef]
  75. Dinh Dang, M.; Ho Thuy Nguyen, L.; Thi Thu Nguyen, T.; Xuan Dat Mai, N.; Hoang Tran, P.; Le Hoang Doan, T. Using sulfate-functionalized Hf-based metal–organic frameworks as a heterogeneous catalyst for solvent-free synthesis of pyrimido[1,2-a]benzimidazoles via one-pot three-component reaction. J. Ind. Eng. Chem. 2021, 103, 340–347. [Google Scholar] [CrossRef]
  76. Makhsous, M.; Shirini, F.; Seddighi, M.; Mazloumi, M. Efficient Synthesis of Pyrimido[1,2-a]benzimidazoles and ethyl pyrimido[1,2-a]benzimidazole-3-carboxylates using Brönsted acidic ionic liquid supported on nanoporous Na+-Montmorillonite. Polycycl. Arom. Comp. 2020, 40, 494–501. [Google Scholar] [CrossRef]
  77. Safaei-Ghomi, J.; Tavazo, M.; Vakili, M.R.; Shahbazi-Alavi, H. Chitosan functionalized by citric acid: An efficient catalyst for one-pot synthesis of 2,4-diamino-5H-benzopyrano[2,3-b]pyridine-3-carbonitriles 5-(arylthio) or 5-[(arylmethyl)thio] substituted. J. Sulphur Chem. 2017, 38, 236–248. [Google Scholar] [CrossRef]
  78. Kour, G.; Gupta, M. A nano silver-xerogel (Agnps@modified TEOS) as a newly developed nanocatalyst in the synthesis of benzopyranopyrimidines (with secondary and primary amines) and gem-bisamides. Dalton Trans. 2017, 46, 7039–7050. [Google Scholar] [CrossRef]
  79. Kabeer, S.A.; Reddy, G.R.; Sreelakshmi, P.; Manidhar, D.M.; Reddy, C.S. TiO2-SiO2 catalyzed eco-friendly synthesis and antioxidant activity of benzopyrano[2,3-d]pyrimidine derivatives. J. Heterocycl. Chem. 2017, 54, 2598–2604. [Google Scholar] [CrossRef]
  80. Sadjadi, S.; Heravi, M.M.; Malmir, M. Heteropolyacid@creatin-halloysite clay: An environmentally friendly, reusable and heterogeneous catalyst for the synthesis of benzopyranopyrimidines. Res. Chem. Intermed. 2017, 43, 6701–6717. [Google Scholar] [CrossRef]
  81. Mostafavi, M.M.; Movahedi, F. Synthesis, characterization, and heterogeneous catalytic activity of sulfamic acid functionalized magnetic IRMOF-3. Eur. J. Inorg. Chem. 2019, 2019, 787–793. [Google Scholar] [CrossRef]
  82. Liang, Y.; Hu, Y.; Zhou, X.; Wu, Q.; Lin, X. One-pot construction of spirooxindole backbone via biocatalytic domino reaction. Tetrahedron. Lett. 2017, 30, 2923–2926. [Google Scholar] [CrossRef]
  83. Arya, K.; Rawat, D.S.; Sasai, H. Zeolite supported Bronsted-acid ionic liquids: An eco-approach for synthesis of spiro[indole-pyrido[3,2-e]thiazine] in water under ultra-sonication. Green Chem. 2012, 14, 1956–1963. [Google Scholar] [CrossRef]
  84. Rohi, M.; Sadeghi, B.; Moslemin, M. Synthesis of spirooxindoles with three component reaction between β-ketoesters, malononitrile and isatin derivatives in the presence of Alum.SiO2 nanoparticles as a new catalyst. Ciencia eNatura. 2015, 37, 80–85. [Google Scholar] [CrossRef] [Green Version]
  85. Sadeghi, B.; Mousavi, S.A. Preparation and characterization of nano-coc-OSO3H as a novel nanocatalyst for the one-pot synthesis of spirooxindoles. Polycycl. Aromat. Compd. 2022, 42, 424–436. [Google Scholar] [CrossRef]
  86. Kleinman, E.F. The bimolecular aliphatic Mannich and related reactions. In Comprehensive Organic Synthesis; Trost, B.M., Fleming, I., Eds.; Elsevier Science Ltd.: Amsterdam, The Netherlands, 1991; Volume 2, pp. 893–951. [Google Scholar]
  87. Bosica, G.; Gabarretta, J. Unprecedented one-pot multicomponent synthesis of propargylamines using Amberlyst A-21 supported CuI under solvent-free conditions. RSC Adv. 2015, 57, 46074–46087. [Google Scholar] [CrossRef]
  88. Bosica, G.; Abdilla, R. The KA2 coupling reaction under green, solventless, heterogeneous catalysis. J. Mol. Cat. A. 2017, 426, 542–549. [Google Scholar] [CrossRef]
  89. Sun, W.; Xi, F.; Pan, W.; Gao, E. MIL-101(Cr)-SO3Ag: An efficient catalyst for solvent-free A3 coupling reactions. Mol. Cat. 2017, 430, 36–42. [Google Scholar] [CrossRef]
  90. Teimuri, R.; Gholamhosseini-Nazari, M.; Esmati, S.; Shahrisa, A. An efficient and green method for the synthesis of Betti base employing nano-SiO2-H3BO3 as a novel recyclable heterogenous catalyst. Res. Chem. Intermed. 2017, 43, 6845–6861. [Google Scholar] [CrossRef]
  91. Bosica, G.; Abdilla, R.; Demanuele, K. Revisiting the Betti synthesis: Using a cheap, readily available, recyclable clay catalyst under solventless conditions. Eur. J. Org. Chem. 2018, 44, 6127–6133. [Google Scholar] [CrossRef]
  92. Lei, Z.K.; Xiao, L.; Lu, X.Q.; Huang, H.; Liu, C.J. Graphite-supported perchloric acid (HClO4-C): An efficient and recyclable heterogeneous catalyst for the one-pot synthesis of amidoalkyl naphthols. Molecules 2013, 28, 1653–1659. [Google Scholar] [CrossRef] [Green Version]
  93. Kothandapani, J.; Ganesan, A.; Ganesan, S.S. Magnetically separable sulfonic acid catalysed one-pot synthesis of diverse indole derivatives. Tetrahedron. Lett. 2015, 56, 5568–5572. [Google Scholar] [CrossRef]
  94. Bosica, G.; Abdilla, R. Regioselective one-pot aza-Friedel-Crafts reaction for primary, secondary and tertiary anilines using heterogeneous catalyst. Green Chem. 2017, 19, 5683–5690. [Google Scholar] [CrossRef]
  95. Bosica, G.; Zammit, R. One-pot multicomponent nitro-Mannich reaction using a heterogeneous catalyst under solvent-free conditions. PeerJ 2018, 6, e5065. [Google Scholar] [CrossRef] [Green Version]
  96. Kulkarni, A.M.; Pandit, K.S.; Chavan, P.V.; Desai, U.V.; Wadgaonkar, P.P. Cobalt ferrite nanoparticles: A magnetically separable and reusable catalyst for Petasis-Borono-Mannich reaction. RSC Adv. 2015, 5, 70586–70594. [Google Scholar] [CrossRef]
  97. Sobhani, S.; Falatooni, Z.M.; Honarmand, M. Synthesis of phosphoric acid supported on magnetic core-shell nanoparticles: A novel recyclable heterogeneous catalyst for Kabachnik-Fields reaction in water. RSC Adv. 2014, 4, 15797–15806. [Google Scholar] [CrossRef]
  98. Teimuri, A.; Ghorbanian, L. One-pot three-component synthesis of β-amino carbonyl compounds using nanocrystalline solid acid catalyst. Int. J. Green Nanotechnol. 2013, 1, 1943089213507161. [Google Scholar] [CrossRef]
  99. Liu, D.; Zhang, J.; Zhang, C.; Kou, X. One-pot synthesis of β-amino carbonyl compounds catalysed silica supported phenylphosphinic acid. Green Process. Synth. 2015, 4, 11–15. [Google Scholar]
  100. Zhang, G.; Zheng, D.; Nie, J.; Wang, T.; Ma, J. Bronsted acid-catalyzed efficient Strecker reaction of ketones, amines and trimethylsilyl cyanide. Org. Biomol. Chem. 2010, 6, 1399–1405. [Google Scholar] [CrossRef]
  101. Gharib, A.; Pesyan, N.N.; Fard, L.V.; Roshani, M. Catalytic synthesis of α-aminonitriles using nano copper ferrite under green conditions. Org. Chem. Intern. 2014, 2014, 169803. [Google Scholar] [CrossRef] [Green Version]
  102. Knoevenagel, E. Condensation of malonic acid with aromatic aldehydes via ammonia and amines. Chem. Ber. 1898, 31, 2596–2619. [Google Scholar] [CrossRef] [Green Version]
  103. Horiuchi, Y.; Toyao, T.; Fujiwaki, M.; Dohshi, S.; Kim, T.; Matsuoka, M. Zeolitic imidazolate frameworks as heterogeneous catalysts for a one-pot P–C bond formation reaction via Knoevenagel condensation and phospha-Michael addition. RSC Adv. 2015, 5, 24687–24690. [Google Scholar] [CrossRef]
  104. Sobhani, S.; Hasaninejad, A.; Maleki, M.F.; Parizi, Z.P. Tandem Knoevenagel–Michael Reaction of 1-Phenyl-3-methyl-5-pyrazolone with Aldehydes Using 3-Aminopropylated Silica Gel as an Efficient and Reusable Heterogeneous Catalyst. Synth. Commun. 2012, 42, 2245–2255. [Google Scholar] [CrossRef]
  105. Tietze, L.F.; Brasche, G.; Gericke, K.M. Multicomponent reactions. In Domino Reactions in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2006; pp. 542–561. [Google Scholar]
  106. Antenucci, A.; Marra, F.; Dughera, S. Silica gel-immobilised chiral 1,2-benzenedisulfonimide: A Bronsted-acid heterogeneous catalyst for enantioselective multicomponent Passerini reaction. RSC Adv. 2021, 11, 26083–26092. [Google Scholar] [CrossRef]
  107. Ugi, I.; Meyr, R.; Fetzer, U. Versuche mit isonitrilen. Angew. Chem. 1959, 71, 386–397. [Google Scholar]
  108. Mohanram, I.; Meshram, J.; Shaikh, A. Kandpal, Microwave-assisted one-pot synthesis of bioactive UGI-4CR using fluorite as benign and heterogeneous catalyst. Synth. Commun. 2013, 43, 3322–3328. [Google Scholar] [CrossRef]
Figure 1. Homotopic catalysts have one active site, whilst heterotopic catalysts have several identical sites.
Figure 1. Homotopic catalysts have one active site, whilst heterotopic catalysts have several identical sites.
Catalysts 12 00725 g001
Figure 2. Convergent vs. divergent synthesis.
Figure 2. Convergent vs. divergent synthesis.
Catalysts 12 00725 g002
Figure 3. Pyrrole synthesis—four-component synthesis involving nitroalkanes, aldehydes, 1,3-dicarbonyl compounds, and primary amines.
Figure 3. Pyrrole synthesis—four-component synthesis involving nitroalkanes, aldehydes, 1,3-dicarbonyl compounds, and primary amines.
Catalysts 12 00725 g003
Figure 4. Furan synthesis using an aldehyde, an amine, and a dialkyl acetylene dicarboxylate.
Figure 4. Furan synthesis using an aldehyde, an amine, and a dialkyl acetylene dicarboxylate.
Catalysts 12 00725 g004
Figure 5. Hantzsch synthesis between aldehydes, ammonium acetate, and ethyl acetoacetate to yield 1,4-DHPs.
Figure 5. Hantzsch synthesis between aldehydes, ammonium acetate, and ethyl acetoacetate to yield 1,4-DHPs.
Catalysts 12 00725 g005
Figure 6. 1,4-DHPs obtained by the reaction of aldehydes, ammonium acetate, and acetoacetanilide using sulphated tin(IV) oxide (STO).
Figure 6. 1,4-DHPs obtained by the reaction of aldehydes, ammonium acetate, and acetoacetanilide using sulphated tin(IV) oxide (STO).
Catalysts 12 00725 g006
Figure 7. 1,2-DHP synthesis using phosphotungstic acid (PW) supported on alumina at room temperature in neat conditions.
Figure 7. 1,2-DHP synthesis using phosphotungstic acid (PW) supported on alumina at room temperature in neat conditions.
Catalysts 12 00725 g007
Figure 8. 2-Amino-3-cyanopyridine derivatives via the reaction of aldehydes, ketones, ammonium acetate, and malononitrile using CuI/C as catalyst.
Figure 8. 2-Amino-3-cyanopyridine derivatives via the reaction of aldehydes, ketones, ammonium acetate, and malononitrile using CuI/C as catalyst.
Catalysts 12 00725 g008
Figure 9. Pyridinone synthesis via the combination of Meldrum’s acid, aryl aldehydes, and methyl acetoacetate.
Figure 9. Pyridinone synthesis via the combination of Meldrum’s acid, aryl aldehydes, and methyl acetoacetate.
Catalysts 12 00725 g009
Figure 10. Pyrano [2,3-c] coumarin derivative synthesis using titania nanowires in ethanol/water solvent under reflux conditions.
Figure 10. Pyrano [2,3-c] coumarin derivative synthesis using titania nanowires in ethanol/water solvent under reflux conditions.
Catalysts 12 00725 g010
Figure 11. Furo [3,2-c]coumarin derivatives synthesized using zinc oxide supported on fluorapatite.
Figure 11. Furo [3,2-c]coumarin derivatives synthesized using zinc oxide supported on fluorapatite.
Catalysts 12 00725 g011
Figure 12. Chromene synthesis via the combination of 2-naphthol, an aldehyde, and malononitrile.
Figure 12. Chromene synthesis via the combination of 2-naphthol, an aldehyde, and malononitrile.
Catalysts 12 00725 g012
Figure 13. Xanthene synthesis using two equivalents of 2-naphthol and an aldehyde.
Figure 13. Xanthene synthesis using two equivalents of 2-naphthol and an aldehyde.
Catalysts 12 00725 g013
Figure 14. The structure of celecoxib, a pyrazole-containing NSAID drug.
Figure 14. The structure of celecoxib, a pyrazole-containing NSAID drug.
Catalysts 12 00725 g014
Figure 15. Pyranopyrazole synthesis using a β-dicarbonyl compound, an aldehyde, a malononitrile, and a hydrazine hydrate.
Figure 15. Pyranopyrazole synthesis using a β-dicarbonyl compound, an aldehyde, a malononitrile, and a hydrazine hydrate.
Catalysts 12 00725 g015
Figure 16. Pyranopyrazole synthesis via 3-methyl-1-phenyl-2-pyrazoline-5-one, aldehydes, and malononitrile.
Figure 16. Pyranopyrazole synthesis via 3-methyl-1-phenyl-2-pyrazoline-5-one, aldehydes, and malononitrile.
Catalysts 12 00725 g016
Figure 17. Histamine, an imidazole-containing hormone that regulates action in the gut and acts as a neurotransmitter for nerve impulses in the brain.
Figure 17. Histamine, an imidazole-containing hormone that regulates action in the gut and acts as a neurotransmitter for nerve impulses in the brain.
Catalysts 12 00725 g017
Figure 18. Copper(I) oxide–magnetite suspended on guarana nanocatalyst used to catalyse the formation of 2,4,5-triaryl-1H-imidazoles.
Figure 18. Copper(I) oxide–magnetite suspended on guarana nanocatalyst used to catalyse the formation of 2,4,5-triaryl-1H-imidazoles.
Catalysts 12 00725 g018
Figure 19. Vitamin B1, also known as thiamine.
Figure 19. Vitamin B1, also known as thiamine.
Catalysts 12 00725 g019
Figure 20. Synthesis of trisubstituted 1,3-thiazole derivatives.
Figure 20. Synthesis of trisubstituted 1,3-thiazole derivatives.
Catalysts 12 00725 g020
Figure 21. Synthesis of 4-thiazolidinones using perchloric acid–silica catalyst.
Figure 21. Synthesis of 4-thiazolidinones using perchloric acid–silica catalyst.
Catalysts 12 00725 g021
Figure 22. 4-Thiazolidinone derivative synthesis via benzyl alcohols.
Figure 22. 4-Thiazolidinone derivative synthesis via benzyl alcohols.
Catalysts 12 00725 g022
Figure 23. 5-fluorouracil—an anticancer drug usually used to treat colon tumours.
Figure 23. 5-fluorouracil—an anticancer drug usually used to treat colon tumours.
Catalysts 12 00725 g023
Figure 24. Synthesis of pyrimidines from aldehydes, acetophenone, and guanidine hydrochloride.
Figure 24. Synthesis of pyrimidines from aldehydes, acetophenone, and guanidine hydrochloride.
Catalysts 12 00725 g024
Figure 25. Biginelli reaction between an aldehyde, urea, and a β-dicarbonyl compound.
Figure 25. Biginelli reaction between an aldehyde, urea, and a β-dicarbonyl compound.
Catalysts 12 00725 g025
Figure 26. Synthesis of pyrimidinones via a Biginelli-like reaction of aldehydes, acetophenone, and urea.
Figure 26. Synthesis of pyrimidinones via a Biginelli-like reaction of aldehydes, acetophenone, and urea.
Catalysts 12 00725 g026
Figure 27. Synthesis of indolizine derivatives via A3 coupling–cyclization tandem reaction of an alkyne, an amine, and 2-pyridinecarboxaldehyde.
Figure 27. Synthesis of indolizine derivatives via A3 coupling–cyclization tandem reaction of an alkyne, an amine, and 2-pyridinecarboxaldehyde.
Catalysts 12 00725 g027
Figure 28. Benzo[b]furan synthesis via A3 coupling of aldehydes, amines, and alkynes.
Figure 28. Benzo[b]furan synthesis via A3 coupling of aldehydes, amines, and alkynes.
Catalysts 12 00725 g028
Figure 29. Selective synthesis of benzofurans and dihydrobenzofurans depending on the salicylaldehyde used.
Figure 29. Selective synthesis of benzofurans and dihydrobenzofurans depending on the salicylaldehyde used.
Catalysts 12 00725 g029
Figure 30. Synthesis of pyrimido [1,2-b]benzoimidazoles via the combination of aldehydes, 2-aminobenzoimidazole, and ethyl acetoacetate.
Figure 30. Synthesis of pyrimido [1,2-b]benzoimidazoles via the combination of aldehydes, 2-aminobenzoimidazole, and ethyl acetoacetate.
Catalysts 12 00725 g030
Figure 31. Synthesis of chromeno [2,3-b]pyridines from salicylaldehyde, thiophenol, and two equivalents of malononitrile.
Figure 31. Synthesis of chromeno [2,3-b]pyridines from salicylaldehyde, thiophenol, and two equivalents of malononitrile.
Catalysts 12 00725 g031
Figure 32. Synthesis of chromeno [2,3-d]pyrimidine from salicylaldehyde, thiophenol, and two equivalents of malononitrile.
Figure 32. Synthesis of chromeno [2,3-d]pyrimidine from salicylaldehyde, thiophenol, and two equivalents of malononitrile.
Catalysts 12 00725 g032
Figure 33. Formation of spirooxindole backbone via a papain-catalysed reaction.
Figure 33. Formation of spirooxindole backbone via a papain-catalysed reaction.
Catalysts 12 00725 g033
Figure 34. Synthesis of spiro[indole-pyrido [3,2-e]thiazine] under ultrasound radiation.
Figure 34. Synthesis of spiro[indole-pyrido [3,2-e]thiazine] under ultrasound radiation.
Catalysts 12 00725 g034
Figure 35. Spirooxindole synthesis via malononitrile, isatins, and β-ketoesters using alum–silica as a catalyst under reflux conditions.
Figure 35. Spirooxindole synthesis via malononitrile, isatins, and β-ketoesters using alum–silica as a catalyst under reflux conditions.
Catalysts 12 00725 g035
Figure 36. The A3coupling reaction catalysed by 10 mol% CuI/A21.
Figure 36. The A3coupling reaction catalysed by 10 mol% CuI/A21.
Catalysts 12 00725 g036
Figure 37. The Betti synthesis via the reaction of 2-naphthol, aldehydes, and amines.
Figure 37. The Betti synthesis via the reaction of 2-naphthol, aldehydes, and amines.
Catalysts 12 00725 g037
Figure 38. The Betti synthesis via the reaction of 2-naphthols, aldehydes, and amides.
Figure 38. The Betti synthesis via the reaction of 2-naphthols, aldehydes, and amides.
Catalysts 12 00725 g038
Figure 39. Aza-Friedel–Crafts reaction for secondary and tertiary aromatic amines gives two different products.
Figure 39. Aza-Friedel–Crafts reaction for secondary and tertiary aromatic amines gives two different products.
Catalysts 12 00725 g039
Figure 40. The nitro-Mannich reaction catalysed by copper(I) iodide supported on Amberlyst A21.
Figure 40. The nitro-Mannich reaction catalysed by copper(I) iodide supported on Amberlyst A21.
Catalysts 12 00725 g040
Figure 41. Borono–Mannich reaction using cobalt ferrite as a catalyst.
Figure 41. Borono–Mannich reaction using cobalt ferrite as a catalyst.
Catalysts 12 00725 g041
Figure 42. The Kabachnik–Fields reaction catalysed by phosphoric acid (PA) supported on magnetic core nanoparticles (γ-Fe3O4@SiO2-PA).
Figure 42. The Kabachnik–Fields reaction catalysed by phosphoric acid (PA) supported on magnetic core nanoparticles (γ-Fe3O4@SiO2-PA).
Catalysts 12 00725 g042
Figure 43. β-amino carbonyl compound synthesis via a Mannich-type reaction of an aldehyde, an amine, and acetophenones.
Figure 43. β-amino carbonyl compound synthesis via a Mannich-type reaction of an aldehyde, an amine, and acetophenones.
Catalysts 12 00725 g043
Figure 44. Synthesis of α-aminonitriles using trimethylsilyl cyanide as the CN source.
Figure 44. Synthesis of α-aminonitriles using trimethylsilyl cyanide as the CN source.
Catalysts 12 00725 g044
Figure 45. Tandem Knoevenagel condensation phospha-Michael reaction.
Figure 45. Tandem Knoevenagel condensation phospha-Michael reaction.
Catalysts 12 00725 g045
Figure 46. Synthesis of bis(pyrazolyl)methanes via a Knoevenagel–Michael reaction.
Figure 46. Synthesis of bis(pyrazolyl)methanes via a Knoevenagel–Michael reaction.
Catalysts 12 00725 g046
Figure 47. The Passerini three-component reaction results in the formation of α-acyloxycarboxamides via the combination of isonitriles, aldehydes, and carboxylic acids.
Figure 47. The Passerini three-component reaction results in the formation of α-acyloxycarboxamides via the combination of isonitriles, aldehydes, and carboxylic acids.
Catalysts 12 00725 g047
Figure 48. The Ugi reaction, a modification of the Passerini reaction.
Figure 48. The Ugi reaction, a modification of the Passerini reaction.
Catalysts 12 00725 g048
Table 1. Synthesis of 2-amino-3-cyanopyridines.
Table 1. Synthesis of 2-amino-3-cyanopyridines.
Author (Date) ReferenceCatalyst and ConditionsYields, Recycling Runs
Heravi et al. (2015) [24]Magnetite nanoparticles (1 mg)
80 °C
Neat conditions
2–3 h
Yields: 82–90%, 7 examples
Recycling runs: 4 (8% decrease in yield)
Yahyazadeh et al. (2018) [25]Aminopropylated-salicylaldehyde-copper(II)-zirconia-coated magnetite
80 °C
Neat conditions
Yields: 85–96%, 17 examples
Recycling runs: 5 (5% decrease in yield)
Afradi et al. (2017) [26]Niacin–magnetite
MW
7–10 min
Yields: 73–95%, 15 examples
Recycling runs: 6 (7% decrease in yield)
Challa et al. (2021) [27]Amberlyst 15
Ultrasound
RT
Acetonitrile solvent
10 min
Yields: 75–92%, 17 examples
Recycling runs: N/A
Achagar et al. (2022) [28]Na2CaP2O7 nanoparticles
Neat
30 min
Yields: 84–95%, 13 examples
Recycling runs: 3 (3% decrease in yield)
Table 2. Catalysts employed in the synthesis of 2-amino-3,5-dicarbonitrile-6-sulfanylpyridines.
Table 2. Catalysts employed in the synthesis of 2-amino-3,5-dicarbonitrile-6-sulfanylpyridines.
Catalysts 12 00725 i001
Synthesis of 2-amino-3,5-dicarbonitrile-6-sulfanylpyridines
Author (Date) ReferenceCatalyst and ConditionsReaction times, Yields, Recycling Runs
Safaei et al. (2014) [29]6 mol% tin(IV) oxide nanoparticles
Ethanol solvent
60 °C
54–142 min, 79–92%
Catalyst recycling: 5 runs (7% drop in yield)
Molla et al. (2014) [30]10 mol% Borax
Ethanol solvent
Reflux
30–120 min, 75–92%, 19 examples
Catalyst recycling: 5 runs (no specific data given)
Kumari et al. (2016) [31]5 mol% graphene oxide/titania
Water solvent
RT
60–120 min, 79–89%, 9 examples
Catalyst recycling: 7 runs (22% drop in yield)
Godugu et al. (2020) [32]Dolomitic limestone (5.0 wt.%)
50% Ethanol/water
45–50 °C
Ultrasound radiation
30–45 min, 90–98%, 18 examples
No recycling data specified
Ebrahimiasl et al. (2019) [33]Magnetite@Cobalt(II) Schiff base
100 °C
Neat conditions
11–20 min, 90–98%, 8 examples
Recycling runs: 5 (3% drop in yield)
Table 3. Recent heterogeneous catalysts employed in the synthesis of chromenes.
Table 3. Recent heterogeneous catalysts employed in the synthesis of chromenes.
Author (Date) ReferenceCatalyst and ConditionsProduct yields, Catalyst Recycling Runs
Banerjee et al. (2015) [42]10 mol% tetragonal zirconium(IV) oxide nanoparticles
Water solvent
80 °C
40–55 min
88–94%, 9 examples
Catalyst reused for 10 runs (92 → 80%)
Rama et al. (2019) [43]20 mol% nickel(II) oxide
Water solvent
50 °C
Ball-milled for 30–60 min
76–99%, 10 examples
Catalyst reused for 3 runs (98 → 93%)
Sobhani et al. (2020) [44]1 mol% piperazine–graphene oxide
50% aqueous ethanol solvent
50 °C
5–150 min
70–98%, 33 examples
Catalyst reused for 6 runs (90 → 85%)
Chahkamali et al. (2022) [45]5 mol% (Γ-Fe2O3-Im-Py)2WO4
90 °C
Neat
5.5–19 h
45–91% (17 examples)
Catalyst reused for 5 runs (91 → 85%)
Table 4. Recent catalysts employed in the Biginelli reaction.
Table 4. Recent catalysts employed in the Biginelli reaction.
Author (Date) ReferenceCatalyst and ConditionsReaction times, Yields, Recycling Runs
Patel et al. (2017) [62]13 mol% PANI-FeCl3
Acetonitrile solvent
Reflux
24 h, 61–90%, 19 examples
No recycling runs specified
Alonso et al. (2020) [63]5 mol% Nb2O5
Neat
130 °C
38–70 min, 60–94%, 9 examples
Recycling runs: 4 (3% yield drop)
Ali et al. (2021) [64]Pumice stone (0.4 g)
Neat
Heat (Temperature not specified)
2–3 min, 91–97%, 18 examples
No recycling runs specified
Bosica et al. (2021) [65]Silicotungstic acid on Amberlyst 15 (40% w/w) (0.05 g per mmol)
Neat
92 °C
4.5–23 h, 37–86%, 17 examples
Recycling runs: 5 (12% yield drop)
Table 5. Recent catalysts employed in the synthesis of chromeno [2,3-d]pyrimidines.
Table 5. Recent catalysts employed in the synthesis of chromeno [2,3-d]pyrimidines.
Author (Date) ReferenceCatalyst and ConditionsReaction Times, Yields, Recycling Runs
Kour et al. (2017) [78]0.1 g of nano-silver-xerogel
60 °C
Ethanol solvent (2 mL per mmol)
45 min
79–93%, 7 examples (secondary amines); 63%—using 1 primary amine
Kabeer et al. (2017) [79]Titania–silica (5 mol%)
80 °C
Neat
10–20 min
89–94%, 12 examples
Catalyst reused 3 times
(92 → 85%)
Sadjadi, S. et al. (2017) [80]Heteropolyacid@creatin-halloysite clay (0.03 g per mmol)
Ultrasound radiation
RT
Water solvent (10 mL per mmol)
4 min
78-92%, 8 examples
Catalyst reused 5 times
(95 → 88%)
Mostafavi et al. (2019) [81]Sulfamic acid-functionalized magnetic IRMOF-3 nanocomposite (10 mg per mmol)
RT
4 mL EtOH (per mmol)
78–97%, 21 examples
3 h
Catalyst reused 5 times (95 → 93%)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bosica, G.; Abdilla, R. Recent Advances in Multicomponent Reactions Catalysed under Operationally Heterogeneous Conditions. Catalysts 2022, 12, 725. https://doi.org/10.3390/catal12070725

AMA Style

Bosica G, Abdilla R. Recent Advances in Multicomponent Reactions Catalysed under Operationally Heterogeneous Conditions. Catalysts. 2022; 12(7):725. https://doi.org/10.3390/catal12070725

Chicago/Turabian Style

Bosica, Giovanna, and Roderick Abdilla. 2022. "Recent Advances in Multicomponent Reactions Catalysed under Operationally Heterogeneous Conditions" Catalysts 12, no. 7: 725. https://doi.org/10.3390/catal12070725

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop