Next Article in Journal
Enhanced Photocatalytic Degradation of Herbicide 2,4-Dichlorophenoxyacetic Acid Using Sulfated CeO2
Previous Article in Journal
The Role of Catalysts in Life Cycle Assessment Applied to Biogas Reforming
Previous Article in Special Issue
Palladium-Catalyzed Regioselective [3+2] Cycloadditions of α,β-Unsaturated Imines with Vinylethylene Carbonates: Access to Oxazolidines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Green Catalysts and/or Green Solvents for Sustainable Multi-Component Reactions

Institute of Organic and Analytical Chemistry (ICOA), University of Orleans, UMR-CNRS 7311, BP 6759, Rue de Chartres, 45067 Orleans CEDEX2, France
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(9), 593; https://doi.org/10.3390/catal14090593
Submission received: 31 July 2024 / Revised: 24 August 2024 / Accepted: 31 August 2024 / Published: 4 September 2024

Abstract

:
Here, we describe some well-known multicomponent reactions and the progress made over the past decade to make these processes even more environmentally friendly. We focus on the Mannich, Hantzsch, Biginelli, Ugi, Passerini, Petasis, and Groebke–Blackburn–Bienaymé reactions. After describing the origin of the reactions and their mechanisms, we summarize some advances in terms of the eco-compatibility of these different MCRs. These are followed by examples of some reactions, considered as variants, which are less well documented but which are promising in terms of structures generated or synthetic routes.

1. Introduction

Multi-component reactions, sometimes confused with tandem or domino reactions, are very important in green chemistry, given the diversity of products they can generate.
The multicomponent reaction itself is sustainable, since most of atoms present in the starting substrates are found in the final product [1], but it can also be improved by using green catalysts and/or green solvents. Indeed, this kind of reaction is quite old, dating back to the previous century, but the emergence of combinatorial chemistry in industry during the 1990’s brought it back onto the stage.
The aim of combinatorial chemistry and its various methods is to minimize the time needed to synthesize products of interest in a given core structure, in order to maximize patent exploitation time for pharmaceutical companies for economic purposes. Industries have all turned to these new processes that use mixtures in solution or supported synthesis. This impetus was based on Bruce Merrifield’s work in 1963 about synthesis of supported polypeptides [2]. This type of synthesis is widely used in this field at the industrial level. These reactions have thus emerged as the methodology of choice for making large numbers of compounds with great diversity in a minimum of time, especially as these reactions involve amino, carbonyl (aldehyde, ketone, carboxylic acid) and other commercially available substrates with a wide variety of structures.
The limiting factor at the outset was the small number of isonitriles often involved in these reactions. The availability of isonitriles is no longer an issue, and work on these reactions has moved towards the use of green reagents or solvents to improve durability.
The positive effects of catalysis in terms of improving processes to meet certain principles of green chemistry, as well as action on the solvents used, can speed up the transition to eco-compatible chemistry. Indeed, solvents are the most important constituents in terms of quantity in a chemical reaction, and they account for a large proportion of environmental performance. This is why, in addition to reducing catalyst quantities, the use of less toxic or even recyclable and/or biomass-derived alternatives as reaction solvents can have a significant impact. For this reason, here, we have chosen examples using mainly solvents possibly derived from biomass and not from petroleum.
In this review, we discuss the different criteria that can characterize a multi-component reaction, and we focus on improvements made in recent years in terms of eco-compatibility and increased diversity. A domino reaction involves chemical transformations with the formation of bonds and new functionalities that enable subsequent reactions. The differences between so-called cascade or tandem reactions are not obvious, and it is common to see approximations in naming, although this depends on how the reaction is initiated or how the reactions propagate. However, definitions often vary. Here, the processed reactions are domino-type reactions involving at least three substrates and often more. Moreover, most of the substrates’ atoms are found in the final product [3,4,5].

2. Multi-Component Reactions (MCRs)

Multi-component reactions are processes that approximate to what nature is capable of creating from various compounds in the same pot. They are efficient and selective reactions that can generate complex structures from standard but multifunctional compounds. The mechanisms can be reversible or irreversible and involve both intermolecular and intramolecular reactions. It is noteworthy that intramolecular reactions often generate a degree of stereoselectivity. The most well-known MCRs are the three or four-component Mannich [6,7], Hantzsch [8,9], Biginelli [10,11], Ugi [12,13], Passerini [14,15], Petasis [16], Groebke–Blackburn–Bienaymé [17] reactions. Transition metal catalysts are sometimes used in multicomponent processes to increase reactivity and selectivity, which may be limited in some reactions. In fact, the problems with multi-component reactions are the potential by-products generated if the reaction is not complete. Increasing reactivity therefore guarantees fewer purification problems. Although the use of catalysts does not always guarantee a long-lasting reaction, it is necessary to weigh up the benefit-risk balance, as with drugs. Is it not better to use a few percent of metal catalysts, even those recognized as non-green, than to carry out solvent-intensive purification? Similarly, is it not better to limit the impact of the reaction solvent rather than the catalyst? The main improvement associated with these MCR reactions concerns the atomic savings they provide, both in terms of the reaction itself to obtain elaborated products and the enormous atomic gain due to the absence of intermediate purifications required in conventional multi-step processes previously used to synthesize these target compounds.
In this study, we summarize the currently available literature about the multicomponent reactions we consider the most sustainable. In addition to the best-known MCR reactions, a large number of anonymous MCRs have been developed and reported over the past two or three decades and can be considered variants of the classical MCRs listed above. It is also interesting to note that enantioselective syntheses under environmentally friendly conditions have been developed. Where possible, we mention one such example at the end of each section.

2.1. Mannich 3-MCR

The Mannich reaction, which involves the reaction of an aldehyde derivative, an amine, and a compound with at least one acidic hydrogen, has been used to generate a variety of pharmaceutical products. This three-component reaction takes place under acid catalysis and can produce a wide variety of compounds [7]. The mechanism begins with the condensation of an amine with an enolizable carbonyl compound to form an iminium ion, which represents the electrophilic compound. The iminium reacts with a nucleophile represented by an enolizable ion, leading to the so-called Mannich base (Scheme 1). Lewis acids generally catalyze this reaction, and the search for sustainable processes involves the use of green catalysts. Green catalysts are non-toxic and/or recyclable and/or abundant catalysts that do not exploit or overexploit planetary resources. Similarly, solvents may include water or ionic liquids, with the use of microwave irradiation or ultrasonic irradiation, as well as any solvent that is biodegradable and/or reusable, or the process may even be solvent- and catalyst-free.
An example of this catalyst- and solvent-free methodology is the synthesis of 7-azagramine analogues by the Mannich reaction involving 7-azaindole derivatives in the presence of aromatic aldehydes and heteroaromatic amines [18]. The researchers tested different solvent and catalysts conditions using Lewis or Brönsted acids (p-TSA, FeCl3, ZnCl2, HCl and others) for the synthesis of these indole derivatives. Under the different reaction conditions tested, the nature of the catalyst was found to have no significant influence on the expected product yield. Even in the absence of catalyst and solvent, this M3CR afforded 7-azagramine analogues in excellent yields (71–87%) after heating within 3–4 h at 80–85 °C. This represents an interesting process given the therapeutic appeal of indole and azaindole derivatives (Scheme 2). Other Mannich reactions were carried out without a catalyst, implying that the ketone in the reaction mechanism was sometimes sufficiently reactive to undergo nucleophilic attack by the amine without a catalyst being required to increase its electrophilicity [19,20,21,22].
Lewis acid catalysts used to increase the electrophilicity of the carbonyl and to promote enolization of the ketone to generate a good nucleophilic partner include Zinc [23], with solvent-free (Scheme 3a) or ZnO nanoparticles enabling the catalyst to be reused, thus making it greener (Scheme 3b) [24]. Both aqueous and alcoholic media involve the use of InCl3 (20 mol%). This reaction can take place in water, but is more easily performed with methanol, and possibly ethanol, so the reaction can be carried out overnight with excellent results and can even be diastereoselective, starting from a chiral amine (Scheme 3c) [25,26].
Similarly, other simple catalysts such as iron derivatives (Scheme 4) [27,28,29,30] have been used to increase electrophilicity or promote nucleophilicity of compounds. Here, we focus on the abundant metal iron; in this example, Fe3+ acts as a catalyst for imine formation. It also increases carbon’s electrophilicity by complexing the aldehyde and then, the imine. Finally, it catalyzes the ketone enolization step. The results are very interesting, enabling solvent-free products to be obtained with very good yields and optimized reaction times ranging from 10 min to 2 h depending on the reagents used. Other iron (III) salts have also been used.
Organic synthesis using microwave irradiation is now well documented. In addition, ultrasound-assisted organic synthesis to decrease reaction time is beginning to gain a strong following among chemists involved in green processes [31]. Regarding the Mannich MCR reaction, Pelit and Turgut [32] reported that, in solvent-free conditions and in presence of camphorsulfonic acid, naphthoxazines were obtained in good yields (Scheme 5). The various reagents played the role of reagents and solvents for the liquids and allow mobility and interactions of the molecules, whether under microwave irradiation or ultrasound. These reactions can also take place in green solvents.
The first study of the catalytic and enantioselective Mannich reaction performed in aqueous medium, an ideal solvent due to its environmentally friendly nature, was carried out and described by Lu et al. in 2007 [33,34]. The use of a threonine-derived organocatalyst proved effective (Scheme 6). This method was applied to aliphatic as well as aromatic aldehydes and enabled them to obtain β-amino-a-hydroxy-ketones in good yields. However, enantioselectivities varied from moderate to high (68–97% ees). In 2008, Wu et al. described an asymmetric Mannich MCR, again in aqueous media, with a hydrophobic organocatalyst O-silyl-serine. Results obtained with various aldehydes and aromatic amines in the presence of various carbonyl donors led to the expected Mannich products with good enantioselectivities (up to 92% ee) [35].

2.2. Hantzsch MCRs

The Hantzsch reaction [8,9] to generate substituted pyridines takes place between an aldehyde with two β-ketoester equivalents and a reagent bearing the future pyridine nitrogen. This compound can be either ammonia or ammonium acetate for the most common reactions. This reaction yields dihydropyridine, which is oxidizable into pyridine with various oxidants. This MCR allows the use of green solvents such as water or ethanol. It therefore features three substrates, but only one reacts twice during the process; so, it is comparable to a four-component reaction. The mechanism (Scheme 7) begins with an aldolization reaction between the aldehyde and the ketoester, followed by a conjugated Michael addition to the enone. The next step involves addition of the amino derivative to the ketone function, followed by cyclization of this imine or enamine, as appropriate, onto the other ketone. Yields are generally good and the environmental impact is reduced, but here, we can choose another oxidant to be even greener. FeCl3 is often used to oxidize dihydropyridine to pyridine.

Variants of the Hantzsch MCR

Variations of this reaction using malonitrile instead of the ketoester provide oxidant-free access to pyridines via Knoevenagel condensation, followed by dehydration. An example of this reaction was carried out in a green solvent from biomass and was produced by our team in 2021 [36]. To achieve the synthesis of highly functionalized pyridines, the reaction was performed in eucalyptol [37] with the stoichiometry already reported in the literature [38]. In this case, it turns out that the addition of a catalyst did not improve the yield. Finally, some derivatives were synthesized using benzaldehyde (1 equiv.), pyrrolidine or another amine (2 equiv.), and malonitrile (2 equiv.). The yields were comparable to those described in conventional solvents (Scheme 8) [36].
Another variant was developed in 2017 to synthesize 2,6-diamino-pyran-3,5-dicarbonitriles, via heterogeneous catalysis using benzaldehyde, various malonitriles, and cyanoacetamide (Scheme 9) [39]. In this example, chitosan, an abundant natural polysaccharide, was used in combination with calcium hydroxyapatite to enhance the reactivity [40]. Here, the expected products were obtained in good yields after 30 min of stirring at room temperature. The same authors carried out other studies using less green mixed Cu/ZrO2, Sm/ZrO2, or Mn/ZrO2 catalysts, but chitosan was found to be more efficient, making it an excellent option for easy separation and recyclability.
Due to their low vapor pressure and great resistance to oxidation and reduction, ionic liquids are considered as green solvents. Srivastava et al. [41] used this type of medium for the synthesis of 3,4-dihydropyrano[c]chromene and tetrahydrobenzo[b]pyrans, heterocyclic scaffolds found in many natural products with high therapeutic potential. Here again, multicomponent methods have been reported, including the use of an aldehyde, a malonitrile, and a 1,3-dicarbonyl compound to synthesize pyran derivatives (Scheme 10).
Despite the efforts already made in this field, the search for new methods to synthesize these derivatives from non-toxic products remains necessary. The authors report the use of an ionic liquid as a catalyst due to its low toxicity. In fact, the saccharin-based ionic liquid [Bmim]sac is involved in 3,4-dihydropyrano[c]chromene and tetrahydrobenzo[b]pyran synthesis. After 10 min of stirring at 80 °C, the desired product was obtained with yields ranging from 70 to 95%. In addition, the reusability of the ionic liquid [Bmim]sac was also investigated, and it could be reused 5 times. Indeed, the purity was preserved even if a slight drop in yield was observed. Consequently, this ionic liquid remains a promising tool for large-scale applications.
An example of highly diastereoselective synthesis of optically active 1,4-dihydropyridines DHPs via the multicomponent Hantzsch reaction using chiral C-glycosyl aldehyde was reported by Dondoni et al. [42] (Scheme 11). That study was carried out in view of the presence of a chiral hydroxylated tetrahydrofuran moiety in biologically active natural compounds [43], with methanol as solvent, which could be advantageously replaced by ethanol or even an ethanol/water mixture. The reason why we detail it here is that we believe it to be of great interest. Various furanosyl derivatives were brought into the presence of acetylacetone and enamine ester in this diastereoselective Hantzsch reaction. When it was carried out with (S)- or (R)-proline (10 mol%) as catalyst, C-glyco-DHPs were obtained with excellent diastereoisomeric excess (>95%), although yields were low (47–50%). The results obtained with either proline suggested that the chiral furanosyl reagent and not only the proline catalyst determined the stereochemical selectivity.

2.3. Biginelli MCR

Pietro Biginelli developed the synthesis of 3,4-dihydropyrimidin-2(1H)-ones by condensation reaction between an aromatic aldehyde, urea, and ethyl acetoacetate. The reaction is similar to all long-lasting MCRs and generally takes place in ethanol [44,45]. This provides access to various pyridines and pyrimidines. Therefore, this reaction can give access to numerous potential therapeutic applications. Here again, the initial synthesis conditions were modified in order to improve yields. As a matter of fact, catalyst recyclability and milder conditions have been studied in the literature [46,47]; consequently, Biginelli synthesis has been carried out without solvent [48] and under microwave or ultrasound activation [49,50]. Brønsted or Lewis catalysts have also been studied, notably rare earths. It now seems established that despite the various possible mechanisms considered, the Biginelli reaction is an MCR catalyzed by urea via the formation of iminium (Scheme 12).
Chitra et al. efficiently synthesized antibacterial compounds via a Biginelli MCR in the presence of isopropyl acetoacetate aldehydes and urea or thiourea in ethanol, catalyzed by strontium chloride hexahydrate. Unfortunately, strontium is not very abundant (Scheme 13) [51].
Many examples of Biginelli reactions were mentioned by M. Marinescu [52], including reactions carried out in an alcoholic solvent, using sustainable heating (microwaves or ultrasound) and non-toxic catalysts. M. Marinescu was thus able to demonstrate the varieties of compounds that could be synthesized in an environmentally friendly way.
Rai et al. have developed a diastereoselective Biginelli reaction using a novel methylene substrate and unprotected aldoses. Synthesis can be carried out in a few minutes under microwave irradiation and does not require additional protection and deprotection steps (Scheme 14) [53].
The reaction takes place, solvent-free, in the presence of urea or thiourea and Montmorillonite K-10 nanoclay. The use of deprotected aldoses as aldehydes, in addition to the structure of the scaffold generated, enabled them to increase the water solubility of the synthesized compound. Rai et al. obtained dihydropyrimidines in good yields (76–89%), with a diastereomeric excess of over 92% [53].

2.4. Ugi MCR

The Ugi-4CR reaction, which dates back to 1959, is one of the best-known MCRs. This reaction has been used in the pharmaceutical industry to easily generate a diversified library of compounds for high-throughput bioassays and detection of positive results [1,12,13,54,55,56,57]. In this reaction, an amine, a carbonyl derivative (aldehyde or ketone), a carboxylic acid, and finally, an isocyanide react together to synthesize an α-acetoamidocarboxamide, according to the mechanism described in Scheme 15.
Recently, W. Zhang et al. [58] have documented different variants leading to a very large diversity of molecules. Indeed, different types of acids, carbonyl compounds, and amino derivatives (ranging from amines, hydrazines, and azides to isocyanides) have been used. For this reason, we will not dwell on the most detailed reactions described in the literature. Ugi MCRs are generally carried out in conventional solvents, using conventional catalysts. To our knowledge, no biomass-derived solvents have yet been used. However, many reactions can be carried out in aqueous or alcoholic media with the help of microwave or ultrasonic activation.
While the enantioselective UGI reaction requires the use of chiral catalysts and generally conventional solvent conditions, this is not the case for diastereoselective Ugi reactions, which can be carried out in alcoholic solvents. For diastereoselective U-4CR, the most promising results have been obtained using chiral amines [59,60,61]. Dyker et al. described the synthesis of chiral isoindoles and dihydroisoquinolines [62]. The 5-center, 4-component Ugi reaction (U-5C-4CR) using a chiral amino acid, an alkyne, and an isocyanate afforded the desired Ugi compounds in moderate yields (68–71%) and diastereoselectivities (66–82% des) (Scheme 16). The reaction was carried out in methanol, a solvent that could be advantageously replaced by ethanol or even a more environmentally friendly solvent. The derivative obtained by subsequent reaction with the alkyne moiety led to dihydroisoquinolines by 6-endo-dig cyclization or to isoindoles by 5-exo-dig cyclization under suitable conditions.

2.5. Passerini MCR

The Passerini mechanism is comparable to the one involved in Ugi’s reaction, even though there are only three components, with the same catalysis conditions (Scheme 17). However, variants involving derivatives with carboxylic acid characters are possible; the isocyanide and the carbonyl derivative are identical to those reported by W. Zhang et al. [58]. This MCR reaction, which like the UGI reaction is one of the best-known MCRs, has not, to our knowledge, been carried out in biomass-based solvents. So, we do not describe it further and focus instead on less common MCRs. However, we believe it would be interesting to work in this direction in future research.
Some enantioselective catalytic Passerini-3CR reactions have been carried out, but these require the use of asymmetric catalysts and generally take place in solvents such as dichloromethane, toluene, etc., which does not allow the desired sustainability [63,64,65].

2.6. Petasis MCR

The Petasis MCR, also known as the Petasis Borono–Mannich reaction, involves a carbonyl derivative, an amine, and a boronic acid to lead to substituted amines with good yields, whether catalyzed or not. The reaction is stereoselective and proceeds easily under mild conditions. Moreover, no protection or deprotection step is required. Finally, Lewis acids or metals may also be added to catalyze this reaction [66,67,68]. It is noteworthy that this MCR was more recently developed than those previously described. Some variations have been described in the literature, involving different substrates [69]. The generally accepted mechanism is described in Scheme 18, as follows.
Reddy et al. [70] described the reaction catalyzed by rare earth triflates, notably lanthanum (La(OTf)3) under microwave irradiation, leading to the desired products in good yields (80–98%). However, the reaction is always carried out in dioxane. Other catalysts, such as indium bromide, have also been used effectively, but always in a conventional solvent [71]. Mojzych et al. [72] have recently listed the operating conditions that can be used in order to synthetize various products via this MCR; although various catalysts considered green are listed, little effort has been made to use accessible solvents, with the exception of Yadav et al. [73], who have worked on the use of [Bmim]BF4 ionic liquids. Their use as a solvent increases the reactivity of boronic acids and therefore give much better results in terms of yield. In addition, the ease of separation and recycling of the ionic liquid makes it a green solvent and an environmentally friendly catalyst (Scheme 19).
Azizi et al. suggested an alternative to conventional solvents by using deep eutectic solvents, due to their ability to be recycled, giving them the characteristic of green solvents [74]. Similarly, several Petasis reaction procedures have been carried out solvent-free and under microwave irradiation conditions, resulting in reasonable yields of the expected products [75,76].
With regard to Petasis MCR, diastereoselective reactions have been described but are often carried out in dichloromethane or toluene as solvent. However, Schreiber et al. used a combination of L-phenylalanine with an enantiopure aldehyde in lactol form [77]. They observed an opposite enantiopreference (S) or (R) depending on the lactol used and they hypothesized that the secondary hydroxyl group, adjacent to the imine intermediate, was the factor directing the stereochemistry of this reaction. The expected stereochemical diversity of products could thus be achieved specifically by choosing the right combination of lactols and amino acids (Scheme 20).

2.7. Groebke-Blackburn-Bienaymé: GBB-3-CR

The Groebke–Blackburn–Bienaymé is a three-component reaction generally referred as GBB-3CR [4,17]. This reaction involves an aldehyde, an isocyanide, and a variable amino derivative, which can lead to a variety of therapeutically important amino heterocycles. Among these is the imidazo[1,2-a]pyridine family, widely used in central nervous system drugs. In general, this reaction is catalyzed in acidic media and implemented in conventional solvents such as ethyl acetate (EtOAc), dichloromethane (CH2Cl2), and alcoholic solvents (usually ethanol or methanol) (Scheme 21).
Most of the time, this reaction is catalyzed by Lewis acids. Longo Jr et al. [78] used microwave irradiation at 150 °C in methanol or ethanol in order to evaluate a number of rare earth triflates. Sc(OTf)3 was used for the synthesis of imidazo[1,2-a]pyridines and study was carried out on other available metal triflates—M(OTf)3, with M = Sc, Y, La, Eu, Gd, and Yb. In(OTf)3 and Bi(OTf)3 were also added to the study as Lewis acids in the presence of 2-aminopyridine, benzaldehyde, and tert-butyl isocyanide; yields ranged from 60 to 95%. It should be noted that Sc(OTf)3 proved more efficient than Y(OTf)3. Other, less expensive rare earth triflates such as La(OTf)3 and Gd(OTf)3 had the same catalytic activity as Sc(OTF)3. This is not always the case in other processes, where scandium generally performs better (Scheme 22). Rare earth derivatives, which are not all that rare, have been widely mined since the 1950s, and demand continues to grow. However, the extraction of these metals requires rather polluting methods. Here again, the benefit–risk ratio has yet to be taken into consideration to determine the ecological interest of this type of catalyst.
We are currently working on optimizing the reaction conditions for this MCR-GBB in bio-based green solvents.

2.8. Others Examples of Undefined MCR

The review by Jonnalagadda et al. in 2021 showed the extent of available recyclable catalysts applied to the organic synthesis of pyranopyrazoles via the use of multicomponent reactions that respect many green chemistry principles [79]. That review showed the value of easily recyclable heterogeneous metal-based catalysts, enabling multiple reaction cycles to be carried out without loss of performance. The reactions reported were often carried out in green solvents, or even better, solvent-free. Two years later, in 2023, Punder et al. [80] referred to the same examples in a review of pyrazole synthesis methods, focusing on reactions in aqueous solvents. The reactions mentioned by these two groups concern variants of multicomponent 3-component or 4-component MCR reactions, often involving hydrazine or arylhydrazines. Here, we describe just a few of the most significant examples.
Bansal et al. [81] reported in 2021 the synthesis of tetra-substituted pyrazoles via a multicomponent reaction in an aqueous medium, an environmentally friendly protocol, as shown in Scheme 23. Arylaldehydes, ethyl acetoacetate, and PhNHNH2 or NH2NH2 in the presence of cetyltrimethylammonium bromide (CTAB) were used for this green protocol, leading to interesting compounds for therapeutic applications.
Using L-proline as a catalyst, Jun et al. [82] demonstrated the efficient (84–93%) synthesis of spiro[indoline-3,4-pyrano[2,3-c]pyrazole] derivatives via reaction of four components. These spiro derivatives were obtained by reacting α,β-dicarbonyl derivative, malonitrile, symmetrical acetylenic diester, and phenyl hydrazine variously substituted in the para position. In addition, to ensure the viability of the four-component reaction, it was carried out in a mixture of EtOH/H2O solvents (1:1) and under ultrasound to obtain the expected compounds in 30–60 min (Scheme 24).
Dihydropyrano[2,3-c]pyrazoles can be obtained using a variety of processes. Banerjee et al. [83] used ZrO2 nanoparticles at room temperature, with yields ranging from 90 to 98%. In addition, these ZrO2 nanoparticles were recyclable and reusable. In 2020, Zahoor et al., synthesized some pyrano[2,3-c]pyrazoles hydrazine with various arylaldehydes in the presence of L-cysteine (0.5 mole) as a natural catalyst by heating the aqueous medium containing 10% ethanol (9:1) to 90 °C (Scheme 25) [84].
The synthesis of numerous pyrazolo[4′,3′:5,6]pyrido[2,3-d]pyrimidinediones derivatives was achieved by Daraie and Heravi via the same multicomponent sustainable approach, from ethyl acetoacetate, arylaldehyde, hydrazine, and 6-amino-1,3-dimethyluracil. This reaction was performed in water with simple organocatalysts such as triethylamine or L-proline (Scheme 26). The authors were able to obtain the desired heterocycles in very good yields with both triethylamine (82–92%) and L-proline (75–90%) [85].
These MCRs can be likened to a Hantzsch variant in which the amine is replaced by a hydrazine or arylhydrazine.
The last examples of this type of reaction to be mentioned here, using arylcarbaldehyde, malonitrile, ethylacetoacetate, and hydrazine to lead to pyranopyrazoles, involve the use of unusual natural, bio-based catalysts, including citric acid from lemon juice [86], caffeine [87], and sodium ascorbate [88]. Reactions took place in water or in water with ethanol as co-solvent. Overall, this is a particularly sustainable route for the production of various pyranopyrazoles (Scheme 27).
To conclude this series of examples on this note of strong sustainability, we consider the use of another natural source: mango, to achieve the synthesis of 1H-Pyrazolo[1,2-b]phthalazine-5,10-dione derivatives. These heterocycles have shown activity as potential antipyretic, antifungal, and antiproliferative agents. Here again, the use of waste products from the fruit juice and compote industry, for example, can be useful for the development of biologically interesting structures. Indeed, Kamana et al. used dried mango peels to develop a catalyst for the multicomponent synthesis of various 1H-pyrazolo[1,2-b]phthalazine-5,10-diones by condensation of an aldehyde, malononitrile, and phthalhydrazide [89]. This aqueous extract of mango peel ash, known as WEMPA, was prepared as follows: the peel was dried in the sun and then, the fruit peel was burnt to produce ash, which was collected at a ratio of 10 g to 100 mL of water. After agitation and filtration, the filtrate recovered was called aqueous extract of mango peel ash (WEMPA). WEMPA has been shown to contain various metal oxides, which are likely to be responsible for its ability to catalyze this multi-component reaction. Mixing the reagents in molecular quantities solubilized in ethanol in the presence of this WEMPA solution under microwave irradiation enabled the team to obtain the desired products (12 examples) in just a few minutes (6–8 min) with very satisfactory yields (83–89%), making this method a particularly durable alternative (Scheme 28).

3. Conclusions

In this study, we set out to identify sustainable methods for building biologically interesting structures via multi-component reactions that are already very atom-efficient. We are also interested in coupling these MCRs to so-called green catalysts and solvents, to limit further our environmental impact. Of course, the list of possibilities is far from exhaustive, but we have attempted to group together the possibilities for synthesizing O,N,S heterocycles, whether stereo-controlled or not, with commercially available reagents, enabling us to easily generate libraries of compounds. We have avoided description of results with complex catalysts requiring previous synthesis. The use of catalysts derived from food, or simply from what nature provides us, also seem to us to be very interesting routes to be highlighted and explored in future research.

Author Contributions

Conceptualization, S.B.-R.; methodology, G.M., E.C. and S.B.-R.; validation, G.M., E.C. and S.B.-R.; investigation, G.M. and S.B.-R.; resources, S.B.-R.; data curation, G.M., E.C. and S.B.-R.; writing—original draft preparation, S.B.-R.; writing—review and editing, G.M., E.C. and S.B.-R.; supervision, S.B.-R.; project administration, S.B.-R.; funding acquisition, S.B.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dömling, A.; Ugi, I. Multicomponent Reactions with Isocyanides. Angew. Chem. Int. Ed. 2000, 39, 3168–3210. [Google Scholar] [CrossRef]
  2. Merrifield, R.B. Solid-state peptide synthesis. I. Synthesis of a tetrapeptide. J. Am. Chem. Soc. 1963, 85, 2149–2154. [Google Scholar] [CrossRef]
  3. Pellissier, H. Stereocontrolled Domino Reactions. Chem. Rev. 2013, 113, 442–524. [Google Scholar] [CrossRef]
  4. Zhu, J.; Bienayme, H. Multicomponent Reactions; Wiley-VCH: Weinheim, Germany, 2005. [Google Scholar]
  5. Orru, R.V.A.; Ruijter, E. Synthesis of Heterocycles via Multicomponent Reactions. In Topics in Heterocyclic Chemistry; Springer: Berlin/Heidelberg, Germany, 2010; Volumes I and II. [Google Scholar]
  6. Mannich, C.; Krösche, W. Ueber ein Kondensationsprodukt aus Formaldehyd, Ammoniak und Antipyrin. Arch. Pharm. 1912, 250, 647–667. [Google Scholar] [CrossRef]
  7. Filho, J.F.A.; Lemos, B.C.; de Souza, A.S.; Pinheiro, S.; Greco, S.J. Multicomponent Mannich reactions: General aspects, methodologies and applications. Tetrahedron 2017, 73, 6977–7004. [Google Scholar] [CrossRef]
  8. Hantzsch, A. Ueber die Synthese pyridinartiger Verbindungen aus Acetessigaẗher und Aldehydammoniak. Justus Liebigs. Ann. Chem. 1882, 215, 1–82. [Google Scholar] [CrossRef]
  9. Debache, A.; Boulcina, R.; Belfaitah, A.; Rhouati, S.; Carboni, B. One-Pot Synthesis of 1,4-Dihydropyridines via a Phenylboronic Acid Catalyzed Hantzsch Three-Component Reaction. Synlett 2008, 2008, 509–512. [Google Scholar] [CrossRef]
  10. Biginelli, P. Ueber Aldehyduramide des Acetessigaẗhers. Eur. J. Inorg. Chem. 2006, 24, 1317–1319. [Google Scholar] [CrossRef]
  11. Alvim, H.G.O.; Pinheiro, D.L.J.; Carvalho-Silva, V.H.; Fioramonte, M.; Gozzo, F.C.; da Silva, W.A.; Amarante, G.W.; Neto, B.A.D. Combined Role of the Asymmetric Counteranion-Directed Catalysis (ACDC) and Ionic Liquid Effect for the Enantioselective Biginelli Multicomponent Reaction. J. Org. Chem. 2018, 83, 12143–12153. [Google Scholar] [CrossRef]
  12. Ugi, I.; Meyr, R.; Fetzer, U.; Steinbrückner, C. Versammlungs- berichte. Angew. Chem. 1959, 71, 373–388. [Google Scholar] [CrossRef]
  13. Van der Heijden, G.; Jong, J.A.W.S.; Ruijter, E.; Orru, R.V.A. 2-Bromo-6-isocyanopyridine as a Universal Convertible Isocyanide for Multicomponent Chemistry. Org. Lett. 2016, 18, 984–987. [Google Scholar] [CrossRef] [PubMed]
  14. Passerini, M. Sopra gli isonitrili (I). Composto del p-isonitril- azobenzolo con acetone ed acido acetico. Gazz. Chim. Ital. 1921, 51, 126–129. [Google Scholar]
  15. Yang, K.; Zhang, F.; Fang, T.; Li, C.; Li, W.; Song, Q. Passerini- type Reaction of Boronic Acids Enables α-hydroxyketones Synthesis. Nat. Commun. 2021, 12, 441. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, P.; Givskov, M.; Nielsen, T.E. Reactivity and Synthetic Applications of Multicomponent Petasis Reactions. Chem. Rev. 2019, 119, 11245–11290. [Google Scholar] [CrossRef]
  17. Boltjes, A.; Dömling, A. The Groebke-Blackburn-Bienaymé Reaction. Eur. J. Org. Chem. 2019, 2019, 7007–7049. [Google Scholar] [CrossRef]
  18. Dongare, S.B.; Chavan, H.V.; Bhale, P.S.; Mule, Y.B.; Kotmale, A.S.; Bandgar, B.P. A catalyst- and solvent-free multicomponent synthesis of 7-azagramine analogues via a Mannich type reaction. Chin. Chem. Lett. 2016, 27, 99–103. [Google Scholar] [CrossRef]
  19. Abonia, R.; Castillo, J.C.; Garay, A.; Insuasty, B.; Quiroga, J.; Nogueras, M.; Cobo, J.; D’Vries, R. A facile synthesis of stable β-amino-N-/O-hemiacetals through a catalyst-free three-component Mannich-type reaction. Tetrahedron Lett. 2017, 58, 1490–1494. [Google Scholar] [CrossRef]
  20. Yang, Z.-K.; Chen, Q.-H.; Wang, F.-P. Construction of AE ring system for the C19-diterpenoid alkaloids with a 5β-hydroxyl group. Tetrahedron 2011, 67, 4192–4195. [Google Scholar] [CrossRef]
  21. Ma, K.; Zhang, C.; Liu, M.; Chu, Y.; Zhou, L.; Hu, C.; Ye, D. First total synthesis of (+)-Carainterol A. Tetrahedron Lett. 2010, 51, 1870–1872. [Google Scholar] [CrossRef]
  22. Prakash, G.K.S.; Gurung, L.; Jog, P.V.; Tanaka, S.; Thomas, T.E.; Ganesh, N.; Haiges, R.; Mathew, T.; Olah, G.A. Direct Synthesis of Diverse β-Fluoroethylamines by a Multicomponent Protocol. Chem. Eur. J. 2013, 19, 3579–3583. [Google Scholar] [CrossRef]
  23. Salter, M.M.; Kobayashi, J.; Shimizu, Y.; Kobayashi, S. Direct-Type Catalytic Three-Component Mannich Reactions Leading to an Efficient Synthesis of α,β-Diamino Acid Derivatives. Org. Lett. 2006, 8, 3533–3536. [Google Scholar] [CrossRef]
  24. Garib, A.; Pesyan, N.N.; Vojdanifard, I.; Jahangir, M.; Roshani, M.; Moghadasi, S. Synthesis of β-amino carbonyl compounds using ZnO nanoparticles as a green, effective and reusable catalyst. Bulg. Chem. Commun. 2014, 46, 486–496. [Google Scholar]
  25. Loh, T.P.; Chen, S.L. InCl3-Catalyzed Three-Component Asymmetric Mannich-Type Reaction in Methanol. Org. Lett. 2002, 4, 3647–3650. [Google Scholar] [CrossRef]
  26. Loh, T.P.; Liung, S.B.K.W.; Tan, K.L.; Wei, L.L. Three Component Synthesis of β-Amino Carbonyl Compounds Using Indium Trichloride-Catalyzed One-pot Mannich-type Reaction in Water. Tetrahedron 2000, 56, 3227–3237. [Google Scholar] [CrossRef]
  27. Behbahani, F.K.; Ziarani, L.M. One pot three-component Mannich reaction catalyzed by FePO4. Eur. Chem. Bull. 2013, 2, 782–784. [Google Scholar] [CrossRef]
  28. Sathesh, V.; Sathishkumar, M.; Ramachandran, G.; Rathore, R.S.; Sathiyanarayanan, K.I. A green approach for the one-pot multi-component synthesis of N-substituted γ, δ and ε-lactams involving C–N bond formation catalyzed by FeCl3. RSC Adv. 2013, 3, 23035–23045. [Google Scholar] [CrossRef]
  29. Behbahani, F.K.; Ziaei, P. A Green Route for the One-Pot Synthesis of 1,2-Disubstituted Benzimidazoles Using Iron(III) Phosphate under Solventless Conditions. Chin. J. Chem. 2012, 30, 65–70. [Google Scholar] [CrossRef]
  30. Halli, J.; Manolikakes, G. Iron-Catalyzed Three-Component Synthesis of α-Amino Acid Derivatives. Eur. J. Org. Chem. 2013, 33, 7471–7475. [Google Scholar] [CrossRef]
  31. Xu, H.; Liao, W.M.; Li, H.H. A mild and efficient ultrasound-assisted synthesis of diaryl ethers without any catalyst. Ultrason. Sonochem. 2007, 14, 779–782. [Google Scholar] [CrossRef]
  32. Pelit, E.; Turgut, Z. (+)-CSA Catalyzed Multicomponent Synthesis of 1-[(1,3-Thiazol-2-ylamino)methyl]-2-naphthols and Their Ring-Closure Reaction under Ultrasonic Irradiation. J. Chem. 2016, 2016, 1–9. [Google Scholar] [CrossRef]
  33. Cheng, L.; Wu, X.; Lu, Y. Direct asymmetric three-component organocatalytic anti-selective Mannich reactions in a purely aqueous system. Org. Biomol. Chem. 2007, 5, 1018–1020. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, L.-W.; Luo, J.; Lu, Y. Asymmetric catalysis with chiral primary amine-based organocatalysts. Chem. Commun. 2009, 14, 1807–1821. [Google Scholar] [CrossRef]
  35. Dziedzic, P.; Ibrahem, I.; Cordova, A. Direct catalytic asymmetric three-component Mannich reactions with dihydroxyacetone: Enantioselective synthesis of amino sugar derivatives. Tetrahedron Lett. 2007, 49, 803–807. [Google Scholar] [CrossRef]
  36. Campos, J.F.; Ferreira, V.; Berteina-Raboin, S. Eucalyptol: A bio-based solvent for the synthesis of O,S,N-Heterocycles. Application to Hiyama Coupling, Cyanation, and Multicomponent Reactions Catalysts 2021, 11, 222. [Google Scholar] [CrossRef]
  37. Campos, J.F.; Scherrmann, M.-C.; Berteina-Raboin, S. Eucalyptol: A new solvent for the synthesis of heterocycles containing oxygen, sulfur and nitrogen. Green. Chem. 2019, 21, 1531–1539. [Google Scholar] [CrossRef]
  38. Sarkar, S.; Das, D.K.; Khan, A.T. Synthesis of fully substituted pyridines and dihydropyridines in a highly chemoselective manner utilizing a multicomponent reaction (MCR) strategy. RSC Adv. 2014, 4, 53752–53760. [Google Scholar] [CrossRef]
  39. Maddila, S.; Gangu, K.K.; Maddila, S.N.; Jonnalagadda, S.B. A facile, efficient, and sustainable chitosan/CaHAp catalyst and one-pot synthesis of novel 2,6-diamino-pyran-3,5-dicarbonitriles. Mol. Divers. 2017, 21, 247–255. [Google Scholar] [CrossRef]
  40. Zeng, M.; Yuan, X.; Yang, Z.; Qi, C. Novel macroporous palladium cation crosslinked chitosan membrane for heterogeneous catalysis application. Inter. J. Biolog. Macromol. 2014, 68, 189–197. [Google Scholar] [CrossRef]
  41. Sharma, H.; Srivastava, S. Anion–cation co-operative catalysis by artificial sweetener saccharine-based ionic liquid forsustainable synthesis of 3,4-dihydropyrano[c]chromenes, 4,5-dihydropyrano [4,3-b]pyran and tetrahydrobenzo[b]pyrans in aqueous medium. RSC Adv. 2018, 8, 38974–38979. [Google Scholar] [CrossRef]
  42. Ducatti, D.R.B.; Massi, A.; Noseda, M.D.; Duarte, M.E.R.; Dondoni, A. DihydropyridineC-glycoconjugates by organocatalytic Hantzsch cyclocondensation. Stereoselective synthesis of α-threofuranoseC-nucleoside enantiomers. Org. Biomol. Chem. 2009, 7, 1980–1986. [Google Scholar] [CrossRef]
  43. Wang, Z.; Huynh, H.K.; Han, B.; Krishnamurthy, R.; Eschenmoser, A. 2,6-Diamino-5,8-diaza-7,9-dicarba-purine. Org. Lett. 2003, 5, 2067–2070. [Google Scholar] [CrossRef] [PubMed]
  44. Biginelli, P. Aldehyde-urea derivatives of aceto-and oxaloacetic acids. Gazz. Chim. Ital. 1893, 23, 360–413. [Google Scholar]
  45. Kappe, C.O. 100 years of the Biginelli dihydropyrimidine synthesis. Tetrahedron 1993, 49, 6937–6963. [Google Scholar] [CrossRef]
  46. Panda, S.S.; Khanna, P.; Khanna, L. Biginelli Reaction: A Green Perspective. Curr. Org. Chem. 2012, 16, 507–520. [Google Scholar] [CrossRef]
  47. Chopda, L.V.; Dave, P.N. Recent Advances in Homogeneous and Heterogeneous Catalyst in Biginelli Reaction from 2015-19: A Concise Review. Chem. Sel. 2020, 5, 5552–5572. [Google Scholar] [CrossRef]
  48. Dondoni, A.; Massi, A. Parallel synthesis of dihydropyrimidinones using Yb(III)-resin and polymer-supported scavengers under solvent-free conditions. A green chemistry approach to the Biginelli reaction. Tetrahedron Lett. 2001, 42, 7975–7978. [Google Scholar] [CrossRef]
  49. Zhang, X.; Li, Y.; Liu, C.; Wang, J. An efficient synthesis of 4-substituted pyrazolyl-3,4-dihydropyrimidin-2(1H)-(thio)ones catalyzed by Mg(ClO4)2 under ultrasound irradiation. J. Mol. Catal. A Chem. 2006, 253, 207–211. [Google Scholar] [CrossRef]
  50. Ahn, B.J.; Gang, M.S.; Chae, K.; Oh, Y.; Shin, J.; Chang, W. A microwave-assisted synthesis of 3,4-dihydropyrimidin-2-(1H)-ones catalyzed by FeCl3-supported Nanopore Silica under solvent-free conditions. J. Ind. Eng. Chem. 2008, 14, 401–405. [Google Scholar] [CrossRef]
  51. Chitra, S.; Devanathan, D.; Pandiarajan, K. Synthesis and in vitro microbiological evaluation of novel 4-aryl-5-isopropoxycarbonyl-6-methyl-3,4-dihydropyrimidinones. Eur. J. Med. Chem. 2010, 45, 367–371. [Google Scholar] [CrossRef]
  52. Marinescu, M. Biginelli Reaction Mediated Synthesis of Antimicrobial Pyrimidine Derivatives and Their Therapeutic Properties. Molecules 2021, 26, 6022. [Google Scholar] [CrossRef]
  53. Yadav, L.D.S.; Awasthi, C.; Rai, V.K.; Rai, A. Biorenewable and mercaptoacetylating building blocks in the Biginelli reaction: Synthesis of thiosugar-annulated dihydropyrimidines. Tetrahedron Lett. 2007, 48, 4899–4902. [Google Scholar] [CrossRef]
  54. Ugi, I. Versuche Mit Isonitrilen. Angew. Chem., Int. Ed. 1959, 71, 386. [Google Scholar]
  55. Alvim, H.G.O.; da Silva Junior, E.N.; Neto, B.A.D. What do we know about multicomponent reactions? Mechanisms and trends for the Biginelli, Hantzsch, Mannich, Passerini and Ugi MCRs. RSC Adv. 2014, 4, 54282–54299. [Google Scholar] [CrossRef]
  56. Madej, A.; Paprocki, D.; Koszelewski, D.; Zadlo-Dobrowolska, A.; Brzozowska, A.; Walde, P.; Ostaszewski, R. Efficient Ugi reactions in an aqueous vesicle system. RSC Adv. 2017, 7, 33344–33354. [Google Scholar] [CrossRef]
  57. Rocha, R.O.; Rodrigues, M.-O.; Neto, B.A.D. Review on the Ugi Multicomponent Reaction Mechanism and the Use of Fluorescent Derivatives as Functional Chromophores. ACS Omega 2020, 5, 972–979. [Google Scholar] [CrossRef]
  58. Hooshmand, S.E.; Zhang, W. Ugi Four-Component Reactions Using Alternative Reactants. Molecules 2023, 28, 1642. [Google Scholar] [CrossRef]
  59. Linderman, R.J.; Binet, S.; Petrich, S.R. Enhanced Diastereoselectivity in the Asymmetric Ugi Reaction Using a New “Convertible” Isonitrile. J. Org. Chem. 1999, 64, 336–337. [Google Scholar] [CrossRef]
  60. Ugi, I.; Demharter, A.; Hörl, W.; Smid, T. Ugi reactions with trifunctional α-amino acids, aldehydes, isocyanides and alcohols. Tetrahedron 1996, 52, 11657–11664. [Google Scholar] [CrossRef]
  61. Ugi, I.; Hörl, W.; Hanusch-Kompa, C.; Smid, T.; Herdtweck, E. MCR 6: Chiral 2,6-Piperazinedione via UGI reactions with α-amino acids, carbonyl compounds, isocyanides and alcohols. Heterocycles 1998, 47, 965–975. [Google Scholar] [CrossRef]
  62. Kadzimirsz, D.; Hildebrandt, D.; Merz, K.; Dyker, G. Isoindoles and dihydroisoquinolines by gold-catalyzed intramolecular hydroamination of alkynes. Chem. Commun. 2006, 6, 661–662. [Google Scholar] [CrossRef]
  63. Andreana, P.R.; Liu, C.C.; Schreiber, S.L. Stereochemical Control of the Passerini Reaction. Org. Lett. 2004, 6, 4231–4233. [Google Scholar] [CrossRef] [PubMed]
  64. Yue, T.; Wang, M.-X.; Wang, D.-X.; Zhu, J. Asymmetric synthesis of 5-(1-hydroxyalkyl)tetrazoles by catalytic enantioselective Passerini-type reactions. Angew. Chem., Int. Ed. 2008, 47, 9454–9457. [Google Scholar] [CrossRef] [PubMed]
  65. Denmark, S.E.; Fan, Y. Catalytic, Enantioselective α-Additions of Isocyanides:  Lewis Base Catalyzed Passerini-Type Reactions. J. Org. Chem. 2005, 70, 9667–9676. [Google Scholar] [CrossRef] [PubMed]
  66. Fodor, A.; Hell, Z.; Pirault-Roy, L. Catalytic activity of metal-doped porous materials in the salicylaldehyde Petasis-BoronoMannich reaction. Monatsh. Chem. 2016, 147, 749–753. [Google Scholar] [CrossRef]
  67. Akhtar, R.; Zahoor, A.F.; Rasool, N.; Ahmad, M.; Ali, K.G. Recent trends in the chemistry of Sandmeyer reaction: A review. Mol. Divers. 2022, 26, 1837–1873. [Google Scholar] [CrossRef]
  68. Bering, L.; Antonchick, A.P. Regioselective metal-free cross-coupling of quinoline N-oxides with boronic acids. Org. Lett. 2015, 17, 3134–3137. [Google Scholar] [CrossRef]
  69. Abonia, R.; Garay, A.; Castillo, J.C.; Insuasty, B.; Quiroga, J.; Nogueras, M.; Cobo, J.; Butassi, E.; Zacchino, S. Design of two alternative routes for the synthesis of naftifine and analogues as potential antifungal agents. Molecules 2018, 23, 520. [Google Scholar] [CrossRef]
  70. Reddy, B.N.; Rani, C.R.; Reddy, S.M.; Pathak, M. An efficient and green La(OTf)3 catalyzed Petasis Borono–Mannich reaction for the synthesis of tertiary amines. Res. Chem. Intermed. 2016, 42, 7533–7549. [Google Scholar] [CrossRef]
  71. Li, Y.; Xu, M.H. Lewis acid promoted highly diastereoselective Petasis borono-Mannich reaction: Efficient synthesis of optically active β,γ-unsaturated α-amino acids. Org. Lett. 2012, 14, 2062–2065. [Google Scholar] [CrossRef]
  72. Saeed, S.; Munawar, S.; Ahmad, S.; Mansha, A.; Zahoor, A.F.; Irfan, A.; Irfan, A.; Kotwica-Mojzych, K.; Soroka, M.; Głowacka, M.; et al. Recent Trends in the Petasis Reaction: A Review of Novel Catalytic Synthetic Approaches with Applications of the Petasis Reaction. Molecules 2023, 28, 8032. [Google Scholar] [CrossRef]
  73. Yadav, J.S.; Reddy, B.S.; Lakshmi, P.N. Ionic liquid accelerated Petasis reaction: A green protocol for the synthesis of alky-laminophenols. J. Mol. Catal. A Chem. 2007, 274, 101–104. [Google Scholar] [CrossRef]
  74. Azizi, N.; Farhadi, E. Straightforward and rapid Petasis multicomponent reactions in deep eutectic solvent. Curr. Res. Green Sustain. Chem. 2020, 4, 100220. [Google Scholar] [CrossRef]
  75. Nun, P.; Martinez, J.; Lamaty, F. Microwave-assisted neat procedure for the Petasis reaction. Synthesis 2010, 12, 2063–2068. [Google Scholar] [CrossRef]
  76. Ayaz, M.; Dietrich, J.; Hulme, C. A novel route to synthesize libraries of quinoxalines via Petasis methodology in two synthetic operations. Tetrahedron Lett. 2011, 52, 4821–4823. [Google Scholar] [CrossRef]
  77. Kumagai, N.; Muncipinto, G.; Schreiber, S.L. Short Synthesis of Skeletally and Stereochemically Diverse Small Molecules by Coupling Petasis Condensation Reactions to Cyclization Reactions. Angew. Chem. Int. Ed. 2006, 45, 3635–3638. [Google Scholar] [CrossRef] [PubMed]
  78. Santos, G.F.D.; Anjos, N.S.; Gibeli, M.M.; Silva, G.A.; Fernandes, P.C.S.; Fiorentino, E.S.C.; Longo, L.S., Jr. A Comparative Study on the Groebke-Blackburn-Bienaymé Three-Component Reaction Catalyzed by Rare Earth Triflates under Microwave Heating. J. Braz. Chem. Soc. 2020, 31, 1434–1444. [Google Scholar] [CrossRef]
  79. Ganta, R.K.; Kerru, N.; Maddila, S.; Jonnalagadda, S.B. Advances in pyranopyrazole scaffolds’s syntheses using sustainable catalysts—A review. Molecules 2021, 26, 3270. [Google Scholar] [CrossRef]
  80. Singh, S.; Yadav, S.; Minakshi; Pundeer, R. Green Synthesis of Pyrazoles: Recent Developments in Aqueous Methods. Synopen 2023, 7, 297–312. [Google Scholar] [CrossRef]
  81. Bansal, R.; Soni, P.K.; Gupta, N.; Bhagyawant, S.S.; Halve, A.K. Synthesis and Antibacterial Screening of Some Pyrazole Derivatives Catalyzed by Cetyltrimethylammoniumbromide (CTAB). Curr. Org. Synth. 2021, 18, 225–231. [Google Scholar] [CrossRef]
  82. Liju, W.; Ablajan, K.; Jun, F. Rapid and efficient one-pot synthesis of spiro[indoline-3,4′-pyrano [2, 3-c]pyrazole] derivatives catalyzed by l-proline under ultrasound irradiation. Ultrason. Sonochem. 2015, 22, 113–118. [Google Scholar] [CrossRef]
  83. Saha, A.; Payra, S.; Banerjee, S. One-pot multicomponent synthesis of highly functionalized bio-active pyrano [2,3-c]pyrazole and benzylpyrazolyl coumarin derivatives using ZrO2 nanoparticles as a reusable catalyst. Green Chem. 2015, 17, 2859–2866. [Google Scholar] [CrossRef]
  84. Sikandar, S.; Zahoor, A.F.; Ahmad, S.; Anjum, M.N.; Ahmad, M.N.; Shah, M.S.U. L-Cysteine Catalyzed Environmentally BenignOne-Pot Multicomponent Approach Towards the Synthesis of Dihydropyrano [2,3-c]Pyrazole Derivatives. Curr. Org. Synth. 2020, 17, 457–463. [Google Scholar] [CrossRef]
  85. Daraie, M.; Heravi, M.M. Molecular diversity of four-component synthesis of pyrazole-basedpyrido [2,3-d]pyrimidine-diones in water: A green synthesis. ARKIVOC 2016, (iv), 328–338. [Google Scholar] [CrossRef]
  86. Govindaraju, S.; Tabassum, S.; Pasha, M.A. Citric-Acid-Catalyzed Green and Sustainable Synthesis of Novel Functionalized Pyrano [2, 3-e]pyrimidin- and Pyrano [2, 3-d]pyrazol-amines in Water via One-Pot Multicomponent Approaches. ChemistrySelect 2018, 3, 3832–3838. [Google Scholar] [CrossRef]
  87. Mohamadpour, F. Caffeine as a Naturally Green and Biodegradable Catalyst for Preparation of Dihydropyrano [2,3-c]pyrazoles. Org. Prep. Proced. Int. 2020, 52, 453–457. [Google Scholar] [CrossRef]
  88. Kiyani, H.; Bamdad, M. Sodium ascorbate as an expedient catalyst for green synthesis of polysubstituted 5-aminopyrazole-4-carbonitriles and 6-amino-1,4-dihydropyrano [2,3-c]pyrazole-5-carbonitriles. Res. Chem. Intermed. 2018, 44, 2761–2778. [Google Scholar] [CrossRef]
  89. Hiremath, P.B.; Kamanna, K. Microwave-Accelerated Facile Synthesis of 1H-Pyrazolo [1,2-b]Phthalazine-5,10-Dione Derivatives Catalyzed by WEMPA. Polycycl. Aromat. Compd. 2022, 42, 2162–2178. [Google Scholar] [CrossRef]
Scheme 1. Mannich reaction and mechanism.
Scheme 1. Mannich reaction and mechanism.
Catalysts 14 00593 sch001
Scheme 2. Synthesis of 7-azagramine derivatives.
Scheme 2. Synthesis of 7-azagramine derivatives.
Catalysts 14 00593 sch002
Scheme 3. Lewis acid-catalyzed Mannich reaction (a) ZnI2, (b) ZnO-nanoparticles or (c) InCl3.
Scheme 3. Lewis acid-catalyzed Mannich reaction (a) ZnI2, (b) ZnO-nanoparticles or (c) InCl3.
Catalysts 14 00593 sch003
Scheme 4. Mannich reaction in solvent-free conditions.
Scheme 4. Mannich reaction in solvent-free conditions.
Catalysts 14 00593 sch004
Scheme 5. Napththoxazines synthesis with Mannich 3-MCR under ultrasound activation.
Scheme 5. Napththoxazines synthesis with Mannich 3-MCR under ultrasound activation.
Catalysts 14 00593 sch005
Scheme 6. Catalytic enantioselective Mannich 3-MCR; (PMP = p-methoxyphenyl and TBDMS = tert-butyldiphenylsilyl protective group).
Scheme 6. Catalytic enantioselective Mannich 3-MCR; (PMP = p-methoxyphenyl and TBDMS = tert-butyldiphenylsilyl protective group).
Catalysts 14 00593 sch006
Scheme 7. Hantzsch reaction and mechanism for dihydropyridines synthesis.
Scheme 7. Hantzsch reaction and mechanism for dihydropyridines synthesis.
Catalysts 14 00593 sch007
Scheme 8. Catalyst-free Hantzsch MCR in eucalyptol.
Scheme 8. Catalyst-free Hantzsch MCR in eucalyptol.
Catalysts 14 00593 sch008
Scheme 9. Heterogeneous catalysis for symmetrical 4H-pyrans preparation.
Scheme 9. Heterogeneous catalysis for symmetrical 4H-pyrans preparation.
Catalysts 14 00593 sch009
Scheme 10. Functionalized pyran synthesis using ionic liquid.
Scheme 10. Functionalized pyran synthesis using ionic liquid.
Catalysts 14 00593 sch010
Scheme 11. Diastereoselective Hantzsch MCR using chiral aldehyde. * represents the created stereocenter.
Scheme 11. Diastereoselective Hantzsch MCR using chiral aldehyde. * represents the created stereocenter.
Catalysts 14 00593 sch011
Scheme 12. Biginelli reaction and mechanism.
Scheme 12. Biginelli reaction and mechanism.
Catalysts 14 00593 sch012
Scheme 13. Synthesis of biological compounds by Biginelli MCR.
Scheme 13. Synthesis of biological compounds by Biginelli MCR.
Catalysts 14 00593 sch013
Scheme 14. Diastereoselective Biginelli reaction, * represents a stereocenter [53].
Scheme 14. Diastereoselective Biginelli reaction, * represents a stereocenter [53].
Catalysts 14 00593 sch014
Scheme 15. Ugi reaction and mechanism.
Scheme 15. Ugi reaction and mechanism.
Catalysts 14 00593 sch015
Scheme 16. Diastereoselective Ugi reaction using chiral amine derivative.
Scheme 16. Diastereoselective Ugi reaction using chiral amine derivative.
Catalysts 14 00593 sch016
Scheme 17. Passerini reaction and mechanism.
Scheme 17. Passerini reaction and mechanism.
Catalysts 14 00593 sch017
Scheme 18. Petasis reaction and mechanism.
Scheme 18. Petasis reaction and mechanism.
Catalysts 14 00593 sch018
Scheme 19. Alkyl aminophenoll preparation with Petasis under green conditions using ionic liquid.
Scheme 19. Alkyl aminophenoll preparation with Petasis under green conditions using ionic liquid.
Catalysts 14 00593 sch019
Scheme 20. Diastereoselective Petasis MCR with both chiral amine and aldehyde reagents.
Scheme 20. Diastereoselective Petasis MCR with both chiral amine and aldehyde reagents.
Catalysts 14 00593 sch020
Scheme 21. The multicomponent GBB reaction and mechanism.
Scheme 21. The multicomponent GBB reaction and mechanism.
Catalysts 14 00593 sch021
Scheme 22. Rare earth-catalyzed GBB 3-CR conditions in ethanol.
Scheme 22. Rare earth-catalyzed GBB 3-CR conditions in ethanol.
Catalysts 14 00593 sch022
Scheme 23. Cetyltrimethylammonium bromide (CTAB)-catalyzed synthesis of tetrasubstituted pyrazoles.
Scheme 23. Cetyltrimethylammonium bromide (CTAB)-catalyzed synthesis of tetrasubstituted pyrazoles.
Catalysts 14 00593 sch023
Scheme 24. Spiro[indoline-3,4-pyrano[2,3-c]pyrazole] synthesis.
Scheme 24. Spiro[indoline-3,4-pyrano[2,3-c]pyrazole] synthesis.
Catalysts 14 00593 sch024
Scheme 25. Dihydropyrano[2,3-c]pyrazoles synthesis with nano- or L-Cysteine catalysts [83,84].
Scheme 25. Dihydropyrano[2,3-c]pyrazoles synthesis with nano- or L-Cysteine catalysts [83,84].
Catalysts 14 00593 sch025
Scheme 26. Pyrazolo[4′,3′:5,6]pyrido[2,3-d]pyrimidinediones syntheses using TEA or L-proline as catalyst.
Scheme 26. Pyrazolo[4′,3′:5,6]pyrido[2,3-d]pyrimidinediones syntheses using TEA or L-proline as catalyst.
Catalysts 14 00593 sch026
Scheme 27. Pyranopyrazoles synthesis using natural organocatalysts.
Scheme 27. Pyranopyrazoles synthesis using natural organocatalysts.
Catalysts 14 00593 sch027
Scheme 28. 1H-pyrazolo[1,2-b]phthalazine-5,10-diones synthesis catalyzed by WEMPA and proposed mechanism.
Scheme 28. 1H-pyrazolo[1,2-b]phthalazine-5,10-diones synthesis catalyzed by WEMPA and proposed mechanism.
Catalysts 14 00593 sch028
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Messire, G.; Caillet, E.; Berteina-Raboin, S. Green Catalysts and/or Green Solvents for Sustainable Multi-Component Reactions. Catalysts 2024, 14, 593. https://doi.org/10.3390/catal14090593

AMA Style

Messire G, Caillet E, Berteina-Raboin S. Green Catalysts and/or Green Solvents for Sustainable Multi-Component Reactions. Catalysts. 2024; 14(9):593. https://doi.org/10.3390/catal14090593

Chicago/Turabian Style

Messire, Gatien, Emma Caillet, and Sabine Berteina-Raboin. 2024. "Green Catalysts and/or Green Solvents for Sustainable Multi-Component Reactions" Catalysts 14, no. 9: 593. https://doi.org/10.3390/catal14090593

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