Next Article in Journal
Impact of Maturity of Malay Cherry (Lepisanthes alata) Leaves on the Inhibitory Activity of Starch Hydrolases
Previous Article in Journal
Encapsulation of the Antioxidant R-(+)-α-Lipoic Acid in Permethylated α- and β-Cyclodextrins: Thermal and X-ray Structural Characterization of the 1:1 Inclusion Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in C–C and C–N Bond Forming Reactions Catalysed by Polystyrene-Supported Copper Complexes

Institute of Organic Chemistry and Technology, Faculty of Chemical Technology, University of Pardubice, Studentská 573, 532 10 Pardubice, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(6), 865; https://doi.org/10.3390/molecules22060865
Submission received: 13 April 2017 / Revised: 11 May 2017 / Accepted: 19 May 2017 / Published: 24 May 2017
(This article belongs to the Section Organic Chemistry)

Abstract

:
This present mini-review covers recently published results on Cu(I) and Cu(II) complexes immobilized on polystyrene carriers, which are used as heterogeneous, eco-friendly reusable catalysts applied for carbon–carbon and carbon–nitrogen forming reactions. Recent advances and trends in this area are demonstrated in the examples of oxidative homocoupling of terminal alkynes, the synthesis of propargylamines, nitroaldolization reactions, azide alkyne cycloaddition, N-arylation of nitrogen containing compounds, aza-Michael additions, asymmetric Friedel–Crafts reactions, asymmetric Mukaiyama aldol reactions, and asymmetric 1,3-dipolar cycloaddition of azomethine ylides. The type of polystyrene matrix used for the immobilization of complexes is discussed in this paper, and particularly, the efficiency of the catalysts from the point of view of the overall reaction yield, and possible enantioselectivity and potential reusing, is reviewed.

Graphical Abstract

1. Introduction

Preservation of the current standard of living for the human population is not further possible without the development of ecologically sustainable chemical processes and technologies [1,2]. One of the methods for contributing to responsible handling with raw materials and energy resources is the utilization of recyclable catalysts [1,2,3,4,5]. The strategy of how to obtain recyclable catalysts belongs the immobilization of known homogeneous catalysts to the solid carriers [3,4,5]. Organic polymers, especially commercially available cross-linked pearl-like copolymers of styrene e.g., Merrifield resin™ [6] and JandaJel resin™ [7] belong among the most relevant carriers. The remarkable advantage of polymer-supported catalysts lies in their simple separation and reusability, their low price, and also in the possibility of their usage in selective continuous-flow procedures [5,7]. The immobilization of the catalyst can be performed by the reaction of ligands or catalysts with suitably chemically activated polystyrene (post-modification strategy) [7]. Another possibility is the grafting of the ligand or catalyst containing a double bond from the reaction, with monomers generating the scaffold of the polymeric carrier (copolymerization strategy) [7]. The convenience of one or other strategies cannot be generally stated—it depends on many factors for each concrete case. Nowadays, there are many examples in chemical databases of recyclable catalysts based on complex compounds immobilized on different polymers. This review is focused on the area of highly efficient recyclable catalysts based on Cu(I) and Cu(II) complexes immobilized on polystyrene carriers. The aim of this review was to summarize, discuss and evaluate recently published results of catalysts, which were designed particularly for reactions, where new carbon–carbon respective carbon–nitrogen bonds are formed.

2. Carbon–Carbon Forming Reactions

2.1. Oxidative Homocoupling of Terminal Alkynes

Carbon–carbon bond forming reactions are essential tools for building the carbon skeleton of organic compounds, and therefore they represent an elemental content of organic synthesis. Copper-catalysed oxidative homocoupling of terminal alkynes known also as Glaser coupling [8] introduce a suitable method for the preparation of 1,3-diynes, which are needful basic building blocks in the synthesis of many natural and pharmaceutical products, as well as some advanced functional materials [9]. A successful application of the heterogeneous catalyst intended for this reaction can be documented on the example of recyclable catalyst 1, which is based on the complex consisting of N,N,N′,N′-tetraethyldiethylenetriamine and CuSO4, anchored on polystyrene cross-linked with 1% divinylbenzene (DVB) (Scheme 1) [10].
The efficiency of catalyst 1 was confirmed from the syntheses of 20 substituted symmetrical 1,3-diynes, where high yields of isolated products were achieved (58–99%). During the reuse of catalyst 1, release of small amount of Cu(II) ions was observed, however this did not affect significantly the yield of 1,4-diphenylbutane-1,3-diyne, even after nine cycles [10]. Another recently published successful heterogeneous catalyst [11] for Glaser coupling was based on the Cu(II) complex of copolymer N-(1,10-phenanthroline-5-yl)acrylamide and DVB. This catalyst was also used for Huisgen 1,3-dipolar cycloaddition [11]. For oxidative C–C homocoupling of acetylenes, a heterogeneous catalyst based on copper nanoparticles (3–9 nm), stabilized onto nitro functionalized polystyrene resin, was prepared. This catalyst was tested in the synthesis of seven 1,4-disubstituted-1,3-diynes (92–95%) [12].

2.2. Sonogashira Cross-Coupling Reaction

Sonogashira cross-coupling of ten substituted iodobenzenes with phenylacetylene catalysed by polystyrene-supported Cu(II) complex 2 was described by Mandapati et al. [13] (Scheme 2). Heterogeneous catalyst 2 was prepared by the reaction of commercial chloromethylated polystyrene (5.5% DVB) with N,N-dimethylethylenediamine and subsequent complexing with CuBr2. Cross-coupling reactions proceeded in aqueous media in the presence of K2CO3 and cetyltrimethylammonium bromide (CTAB) at 110 °C for 10 h. The yields of substituted acetylenes ranged between 66–85%. Reusability of the catalyst was tested for the reaction of iodobenzene with phenylacetylene (Cycle 1: 85%; Cycle 2: 83%; Cycle 3: 82%; Cycle 4: 77%) (Scheme 2).
The main advantage of this system lies in easy preparation of the catalyst from inexpensive and commercially available starting materials. Moreover, a significant benefit is the possibility of reusing of the catalyst, and the eco-friendly performance of the reaction in aqueous media [13].

2.3. Carbene Transfer Reaction

Reactions where carbenes were transferred onto unsaturated substrates (alkenes, alkynes, arenes, imines, aldehydes etc.), have provided a number of useful products [14,15,16,17,18,19]. The mentioned carbenes can be generated in situ from diazo compounds by the application of transition metal-based catalysis [14]. The most widespread example of such reaction implemented at large scale is copper-catalysed cyclopropanation of substituted alkenes with alkyl diazoacetates. Therefore, many homogeneous as well as heterogeneous catalytic systems have been developed for this reaction [14,15,16,17,18,19,20]. In the case of Cu(I) complexes, bisoxazoline ligands [15,16,17,18] and azabisoxazoline ligands [19] were used, anchored on pearl-like polystyrenes. For the immobilization of ligands either a copolymerization strategy (catalysts 3 and 4) or post-modification strategy (catalysts 57) was utilized. The catalysts were generated in situ by the coordination of copper(II) triflate and consequent reduction with phenylhydrazine (Scheme 3) [15,16,17,18,19,20].
Catalyst 3 was prepared by copolymerization of corresponding bisoxazoline-bearing styrene (4%) with styrene (42%) and DVB (54%), which gave a highly cross-linked polymer having a compact structure (<5 m2 g−1) [15,16]. Cyclopropanation catalysed by 3 was tested on the reaction of ethyl diazoacetate with styrene (60%, trans/cis: 67/33, 93% enantioselective excess (ee)) (Scheme 3); with 1,1-diphenylethene (97%, 98% ee) and with 2-methylpropene (92%, 97% ee) [16]. Catalyst 3 was fivefold reused in the reaction of styrene with ethyl diazoacetate without any significant decrease in yield and enantioselectivity. Modification in the preparation of catalyst 3 by performing of copolymerization in different ratios of bisoxazoline-bearing styrene (6%) with styrene (92%) and DVB (2%) did not affect either the isolated yield (67%), or the diastereoselectivity (67/33) and enantioselectivity of cyclopropanation (92% ee) [16]. Likewise, a change of strategy of preparation for catalyst 3 (post-modification of Merrifield resin™) did not lead to any significant impact on the catalyst′s efficiency [16]. Poor enantioselectivity found for catalysts 4 (36%, trans/cis: 37/63, 78% ee) [17] and 5 (18%, trans/cis: 66/34, 26% ee) [17], and 6 (60%, trans/cis: 64/36, 65% ee) [18], was explained by lower binding affinity towards copper, caused by the change of geometry of the ligand as a consequence of the replacement of the gem-dimethyl substitution box-bridge with sterically large groups [19]. A notable improvement in binding affinity towards copper was achieved by using of azabisoxazoline ligand; moreover, it also led to an increase of the catalyst 7’s efficiency (94%, trans/cis: 74/26, 99% ee) [19]. An improvement in enantioselectivity (79% ee, cis)—explained by the effect of the “chiral-cavity” of the rigid polymer—was later achieved for the highly-crosslinked catalyst 4, which was prepared according to the copolymerization strategy [20]. Even in this case, the influence of post-modification strategy or copolymerization strategy was not significant. On the other hand, testing confirmed that the Merrifield resin™ used in catalyst 7 was a much better support than TentaGel™ (35%, trans/cis: 64/34, 47% ee) [19].
The next reusable catalyst 8 was prepared by coordination of [Cu(MeCN)4][PF6] on tris(triazolyl)methane anchored on commercial Merrifield resin™ (Scheme 4) [20]. Reactions of catalyst 8 with ethyl diazoacetate led to formation of carbene (:CHCO2Et), which was used for six reactions with following substrates: cyclohexane (C–H insertion), tetrahydrofuran (THF) (C–H insertion), ethanol (O–H insertion), aniline (N–H insertion), phenylacetylene (cyclopropenation), and benzene (cyclopropanation/ring expansion). In the case of the usage of benzene as substrate, this was the first representative of the Büchner reaction performed under the conditions of heterogeneous catalysis (Scheme 4) [21].
The reactions were performed both in batch and, for the first time, also in continuous flow. In the first variant, the reaction proceeded in a Schlenk flask under an inert atmosphere. Minor by-products were detected—only diethyl fumarate and diethyl maleate. Products of the Cu-catalysed carbene transfer addition or insert reaction 8af were obtained with high yields (8a: 98%; 8b: 82%; 8c: 99%; 8d: 97%; 8e: 98%; 8f: 88%). Moreover, the yields were still satisfactory during experiments with catalyst recycling and its reuse; e.g., after the fifth cycle the results were as follows—8b: 89%; 8c: 99%; 8d: 86%; 8e: 94%; 8f: 82%. A significant decrease of yield was observed with product 8a (67%), which was explained by the lower reactivity of cyclohexane during insertion into the C–H bond [21]. In the continuous flow arrangement, a vertically mounted and fritted low-pressure Omnifit column loaded with catalyst 8 was used. Very stable flow was afforded after two hours with the following results—8a: 0.94 mmol; 41%; 4 mmolproduct·mmolCu−1 h−1; 8b: 2.53 mmol; 56%; 10.8 mmolproduct·mmolCu−1 h−1; 8d: 3.91 mmol; 89%; 16.7 mmolproduct·mmolCu−1 h−1; 8e: 0.54 mmol, 70%; 2.3 mmolproduct·mmolCu−1 h−1; 8c: (after 48 h) 12.5 g; 17.1 mmolproduct·mmolCu−1 h−1; turnover number (TON) = 820. From the given results, catalyst 8 represents a highly efficient catalytic system for carbene transfer reactions to various types of substrates both in batch and in continuous flow [21].

2.4. Addition of Alkynes on Double Bond Carbon–Nitrogen

Propargylamines—which represent important intermediates for the synthesis of significant pharmaceutical polyfunctional substances—can be synthesized by the addition of alkynes on the imine bond (Scheme 5 and Scheme 6) [22,23,24,25,26]. Propargylamines had been traditionally prepared by nucleophilic addition of magnesium or lithium acetylides on imines or their derivatives [22]. The disadvantage of such synthesis consists in the necessity of use of stoichiometric amounts of acetylides under very strict reaction conditions and absolute moisture exclusion. Simplification of this reaction brought the utilization of transition metal complexes—especially Cu(I) complexes [22]. Unfortunately, the isolation of these homogeneous complexes from the reaction mixture as well as their recycling is unfeasible. Further facilitation started with the application of Cu(II) complexes immobilized on pearl-like polystyrene and with the reaction performed as a one-pot multi-component reaction [23,24,25,26]. Catalyst 9 was prepared by the reaction of l-proline with chloroacetylated pearl-like polystyrene and subsequent complexation with CuI (Scheme 5) [23]. The reaction of all three components (alkyne/aldehyde/amine: 1/2/1.25; 5 mol % Cu) catalysed by 9 proceeded under laboratory conditions [Dimethyl sulfoxide (DMSO), 2 h or 12 h at room temperature] with high yields of isolated propargylamines (48–99%). Formaldehyde and aliphatic aldehydes provided higher yields (78–99%) than aromatic aldehydes (e.g., Ph-C≡CH/4-MeO-C6H4CH=O/piperidine: 48%). For the reaction of phenylacetylene, formaldehyde and piperidine catalysed by catalyst 9, several solvents were tested: water (0%), methanol (25%), toluene (30%), THF (42%), MeCN (87%), DMSO (99%). For the same reaction in DMSO, the recovery and reuse of catalyst 9 was tested, when only slight decrease in yields for each cycle was observed: 99%, 99%, 98%, 95%, 90%, 90% (Scheme 5) [23]. Corresponding polymer complexes with CuBr, CuCl and CuSO4 were also studied; however, they were less efficient [23].
Catalyst 10 was prepared by the reaction of β-alanine with polystyrene containing the –CH=O group and by subsequent complexation with CuI (Scheme 5) [24]. The reaction catalysed by catalyst 10 was performed under reflux in water (6–16 h) with the following ratio of educts: alkyne/aldehyde/amine: 1/1/1.2 (2 mol % Cu), while the obtained yields of isolated propargylamines ranged between 48–99%. For the reaction of phenylacetylene, benzaldehyde and piperidine catalysed by 10 also different solvents were tested: dichloromethane (DCM) (15%), toluene (58%), MeCN (62%), THF (68%), water (95%). Furthermore, the recovery and reuse of catalyst 10 was studied for the same reaction in water, where a slight decrease in yields was observed: Cycle 1: 95%; Cycle 5: 93%; Cycle 10: 88%. Besides, the authors monitored the copper content in catalyst 10, which remained unchanged in all of the cycles (Scheme 5) [24].
Catalyst 2 (Scheme 2) was also used by the same authors [13] for the synthesis of substituted propargylamines (Scheme 6) [25]. The reaction of all three components (alkyne/aldehyde/amine: 1.1/1/1.1; 1 mol % Cu) catalysed by catalyst 2 proceeded in toluene (6 h, 110 °C) with yields of isolated propargylamines in the range of 64–95% (Scheme 6). Also in this case, different solvents were tested for the reaction of phenylacetylene, benzaldehyde and piperidine catalysed by catalyst 2: water (trace), methanol (trace), ethanol (trace), THF (22%), MeCN (36%), dimethylformamide (DMF) (48%), dichloroethane (DCE) (46%), DMSO (50%) and toluene (85%). The recovery and reusing of catalyst 2 was studied for the same reaction, again with only slight decrease in yields: Cycle 1: 85%; Cycle 5: 83% (Scheme 6) [25]. Catalyst 11 was prepared by the reaction of commercial chloromethylated polystyrene with N-phenyl piperazine and subsequent complexation with CuBr2 (Scheme 6) [26]. The reaction of all three components (alkyne/ketone/amine: 1/1/1; 0.7 mol % Cu) catalysed by 11 was performed under solvent free conditions (6 h, 110 °C) with the yields of isolated propargylamines in the range of 88–95%. The recovery and reusing of catalyst 11 was tested for the reaction of phenylacetylene, cyclohexanone and piperidine, while only negligible decrease in yields was also observed: Cycle 1: 94%; Cycle 5: 92% (Scheme 6) [26].
From the presented results, the prepared heterogeneous catalysts 2, 911 represent highly efficient catalytic systems. Particularly, catalyst 11 showed itself to be an eco-friendly variant that allowed the synthesis of propargylamines in aqueous media. From the chemical structure of catalyst 9, it contains a defined stereogenic centre. However, the eventual optical purity of chiral propargylamines prepared with the use of this catalyst was not discussed in the cited paper [23]. Elsewhere, e.g., in [27,28,29,30], several enantioselective homogeneous catalysts based on Cu(II) complexes designed for the synthesis of optically pure propargylamines were described. For the heterogeneous variant of enantioselective synthesis of propargylamines, only a version of a Cu(I) pybox complex anchored on magnetic nanoparticles (Fe3O4@SiO2) have been described so far [31].

2.5. Nitroaldolization Reaction

The nitroaldolization (Henry) reaction catalysed by chiral optically pure catalysts is important among reactions; it involves the formation of a new carbon–carbon bond [32,33,34,35,36,37,38,39]. This reaction occurs in the synthesis of substituted 2-nitroalcohols used e.g., for the preparation of biologically active compounds and medicinal drugs [34]. A recently published comprehensive paper [35] covers works concerning catalysts based on supported transition metal complexes aimed for an asymmetric variant of the Henry reaction. Three recent works were not covered in the above mentioned review [36,38,39]. The first work [39] included a study of a library of substituted polystyrylsulfonyl-imidazoline-aminophenol copper complexes as solid supported catalysts for the Henry reaction. The selected complex of copper(II) acetate, based on (S,S)-diphenylethylenediamine-derived imidazoline, (S)-phenylethylamine, and dibromophenol, was highly efficient and provided the products of the reaction (nitromethane with substituted aldehydes) in high yields (76–99%) and high enantioselectivity (75–95% ee). The same ligand as a complex with copper(II) triflate was examined in the Friedel–Crafts alkylation of indole by substituted nitroalkenes, to give the adducts in high yields (86–99%) and high enantioselectivity—up to 83% ee [39]. The second variant of immobilization represented recyclable the heterogeneous catalyst 12 (Scheme 7), containing imidazolidine-4-one covalently bound on swellable pearl-like polystyrene having an –SH group (200–800 µm) [36]. Covalent anchoring of the ligand was performed by radical thiol-alkene click reaction, initiated thermally (azobisisobutyronitrile (AIBN), toluene, reflux 24 h), as well as photochemically (2,2-dimethoxy-2-phenylacetophenone, λmax = 365 nm, 40 W, DCM, 25 °C, 24 h). For the subsequent reaction of the copolymer with copper(II) acetate, heterogeneous catalyst 12 was prepared, and was characterized by elemental microanalysis and Raman spectroscopy [36]. Henry reactions proceeded in the polymeric matrix of swellable catalyst 12 with a rate comparable with the use of catalyst 13 in homogeneous media [37]. The corresponding substituted (R)-2-nitroethanols were produced in high yields (65–99%) and with high enantioselectivity (up to 92% ee) comparable with the homogeneous catalyst 13 (up to 96% ee). The reuse of catalyst 12 was studied for the reaction of pivalaldehyde with nitromethane. After five-fold recovery and reuse of heterogeneous catalyst 12, a decrease in yield of about 24% occurred; on the other hand, the decrease in enantioselectivity was negligible (3% ee) (Scheme 7) [36].
The third variant represented heterogeneous catalysts 14a,b based on Cu(II) complexes of substituted imidazolidine-4-thiones immobilized by means of sulfur atoms on different types of polystyrene carriers (polystyrene-co-4-vinylbenzyl chloride-co-tetra(ethylene glycol)-bis(4-vinylbenzyl)ether), Merrifield resin™ and JandaJel resin™ (Scheme 8) [38]. It was found, that catalysts 14a,b catalyse the Henry reaction with high yields and high enantioselectivity, regardless of the polymeric carrier used (up to 98% ee). The obtained yields and enantioselectivities were comparable with the individual cases of reactions catalysed by analogical homogeneous catalysts 15a,b (Scheme 8). Beside monitoring the yields and enantioselectivity, the possibility of recovery and reusing of catalysts 14a,b was also studied. For instance, in the reaction of 2-methoxybenzaldehyde with nitromethane, the catalyst was reused more than 10 times without any decrease in enantioselectivity (81% ee) (Scheme 8) [38].

2.6. Asymmetric Friedel–Crafts Reaction

Beletskaya et al. [40] studied asymmetric Friedel–Crafts alkylation of indole and its derivatives in the presence of complex of Cu(OTf)2 with chiral isopropyl bis(oxazoline) ligand immobilized on Merrifield resin™ according to the “click” procedure 16. The target products were obtained with a yield of up to 99% and up to 97% ee. Catalyst 16 ensured high yield and enantioselectivity even after five-fold reuse (Scheme 9) [40].
The heterogeneous catalyst 16 was also tested in the alkylation of N-methylindole by methyl (E)-2-oxo-4-phenylbut-3-enoate with yields of 22–94% and lower enantioselectivity (49% ee) [40].

2.7. Asymmetric Mukaiyama Aldol Reaction

The enantioselective Mukaiyama aldol reaction of methyl pyruvate with silylated ketene thioacetal was tested with a heterogeneous polystyrene copper complex catalyst based on chiral bis(oxazoline) [41] (Scheme 10).
The highly-crosslinked polymer 3 was prepared by a copolymerization strategy in the presence of toluene as a porogen agent [41]. The heterogeneous catalyst 3/Cu(OTf)2 was less active than the analogous homogeneous catalysts, and aldol products were formed after a longer time; however, high overall yield (90%) and with an enantiomeric excess comparable to those obtained with the soluble ligands (90% ee), was obtained. Moreover, the catalyst based on polymer complex 3 was easily recovered by simple filtration and was reused several times without any loss of activity or stereoselectivity (Scheme 10) [41]. In the second work, the azabis(oxazoline) complexes 17 bonded to Merrifield resin™ were confirmed for the aldol reaction of different α-ketoesters with silylated ketone enolates [42]. The catalysis results showed that these heterogeneous complexes 17 accomplished enantioselectivities only slightly lower than those obtained in solution. Catalyst recovery depends on the ligand, so whereas polymers (R3: Ph) and (R3: i-Pr) were recovered with the same unsatisfactory results (51%, 38–39% ee), the recovered catalyst coming from (R3: t-Bu) led to even slightly lower yields (25%) but with the higher enantioselectivity (85% ee) (Scheme 10) [42].

3. Carbon–Nitrogen Forming Reactions

3.1. Copper(I)-Catalysed Azide Alkyne Cycloaddition

Beside the reactions leading to carbon–carbon bond formation, the reactions providing new carbon–nitrogen bond also play a relevant role in organic synthesis. Particularly, this includes the Huisgen 1,3-dipolar cycloaddition of alkynes with azides providing 1,4- and 1,5- isomers of 1,2,3-triazole [43]. Copper(I)-catalysed azide alkyne cycloaddition (CuAAC) represents a current variant, which converse to the original thermic version, proceeds quickly, regiospecifically (1,4-), and with high yields [44,45]. These properties rank this reaction among the most significant “near-perfect” bond-forming reactions known as “click” reactions [46]. The importance of this reaction consists not only in the efficient preparation of substituted 1,2,3-triazoles, but especially in the possibility for connecting different molecules and macromolecules, or to functionalize surfaces in a controlled way. The first polystyrene-supported copper(I) catalyst for CuAAC was based on the Amberlyst A21–CuI complex and was published in 2006 [47]. This and many further papers are discussed in a recently published review concerning immobilized copper complexes anchored on polystyrene and other heterogeneous carriers [48]. Also, two other papers deal with CuAAC [49,50].
The authors [49] prepared the catalyst 18 based on copper(I) iodide complex with cryptand-22 immobilized on commercial chloromethylated polystyrene (1% DVB) (Scheme 11). This heterogeneous heat- and air-stable catalyst 18 was used for a two-component CuAAC reaction of substituted organic azides with substituted alkynes, and this provided excellent yields of 1,4-disubstituted 1,2,3-triazoles (78–99%). The second three-component variant of this reaction involving organic halides, sodium azide and alkynes, also resulted in the production of 1,4-disubstituted 1,2,3-triazoles with good yields (45–99%). Both variants of the reaction were performed under eco-friendly conditions (0.6 mol % Cu, aerobic, water, room temperature). The reusing of catalyst 18 was successfully tested for the reaction of benzyl azide with phenylacetylene (Cycle 1: 99%; Cycle 4: 87%) and for the variant of the reaction of benzyl bromide, sodium azide and phenylacetylene (Cycle 1: 99%; Cycle 4: 70%) (Scheme 11) [49]. The catalyst 18 was also successfully used for the formation of aryl–sulphur bonds in aqueous media [50]. In the next work [51], the heterogeneous catalyst 19 based on a copper(II) complex of 1,2-bis(4-pyridylthio)ethane anchored on chloromethylated polystyrene was prepared (Scheme 12). Catalyst 19 was tested for three-component reaction of organic halides, sodium azide and alkynes, which was performed in the presence of sodium ascorbate as a reduction agent in poly(ethylene glycol) (PEG-400), (0.2 mol % Cu), at 65 °C, 10–25 min. The yields of isolated 1,4-disubstituted 1,2,3-triazoles were excellent, ranging from 87–99%. For the reaction of 4-bromobenzylbromide, sodium azide and phenylacetylene catalysed by 19 were further tested different solvents (0.1–0.3 mol % Cu, 15–30 min): water (78%), methanol (70%), toluene (40%), DMF (65%), MeCN (28%), PEG–water 1:1 (89%), ethanol–water 1:1 (75%).
The recovery and reuse of catalyst 19 was successfully confirmed with the same three-component reaction (Cycle 1: 99%; Cycle 4: 95%; Cycle 7: 93%) (Scheme 12). This catalytic system also showed excellent activity for the synthesis of bis-1,4-disubstituted 1,2,3-triazoles (66–88%) [51].

3.2. Other Ring-Formation Reactions

Carretero et al. [52] studied the Cu(I)-catalysed asymmetric 1,3-dipolar cycloaddition of azomethine ylides generated from substituted arylimines derived from glycine esters with N-phenylmaleimide or dimethyl fumarate (Scheme 13). Except for the homogeneous variant, cycloaddition was studied in the heterogeneous system with the use of in situ prepared complexes Cu(MeCN)4ClO4 with Fesulfos-based chiral ligands immobilized on Merrifield′s peptide resin (1% cross-linked, 200–400 mesh, 2 mmol Cl/g) 20 or on trichloroacetimidate Wang resin (copoly(styrene-1% DVB), 200–400 mesh, 0.6–1.00 mmol trichloroacetimidate/g resin) 21.
Interestingly, 20 showed much better catalytic performance than 21 (in all cases from 91 to >99% ee). Furthermore, the polystyrene-supported Fesulphos ligand 20 were recovered and reused in successive catalytic 1,3-dipolar cycloadditions without further addition of Cu(I), maintaining excellent enantioselectivity in up to three runs (Scheme 13) [52].
Halligudi et al. studied biomimetic oxidative coupling of 2-aminophenol to phenoxazine-2-one catalysed by bis(2′-[2-hydroxyethyl]benzimidazolato)copper(II) complex anchored onto chloromethylated polystyrene [53]. The kinetics and the effects of the catalyst concentration, substrate concentration, air pressure and temperature on the reaction course were studied. 2-Aminophenol in the presence of air (70 °C, 800 psig, DMF, 8 h) gave 62% conversion and the immobilized catalyst was more stable and more active than its homogeneous analogue. The catalyst recycling study demonstrated that the anchored catalyst did not leach out the metal complex during the reaction, and thus its catalytic activity was reproducible [53].

3.3. N-Arylation of Amines, Amides and Nitrogen Containing Heterocycles

Aryl–nitrogen bonds formed via cross-coupling reactions represent a powerful tool for the production of a number of compounds with practical uses [54]. The original variant of the Ullmann reaction is still utilized for this purpose [55]. However, this method of performing of the reaction is constrained by several disadvantages, e.g., it needs high temperature (150–200 °C) and the use of stoichiometric amounts of copper [54,55]. A remarkable simplification was achieved by the use of substituted phenylboronic acids for N-arylation of imidazole catalysed by copper(II) salts [56]. The next facilitation was the utilization of heterogeneous catalyst 22 (Scheme 14) [57]. Catalyst 22 was prepared from commercial chloromethylated polystyrene (5.5% DVB), on which β-alanine was attached through the oxygen atom, and copper(II) acetate was coordinated. Less efficient variants of this catalyst were prepared by the coordination with CuCl2 and CuI. Catalyst 22 was successfully tested for the reaction of substituted phenylboronic acids with nitrogen containing heterocycles (imidazole, benzimidazole; 65–98%) and with amides (phthalimide, benzamide, benzenesulfonamide; 37–98%) under very mild conditions (methanol, 40 °C, 10–20 h). The reaction with amines was performed in DMSO in the presence of K2CO3 at 140 °C (aniline, benzylamine; 58–91%) (Scheme 14) [57].
The recovery and reuse of catalyst 22 was studied for the reaction of phenylboronic acid with imidazole (Cycle 1: 98%; Cycle 6: 90%) and aniline (Cycle 1: 86%; Cycle 4: 82%). In both cases, only a slight decrease in yields was observed (Scheme 14) [57]. Under comparable conditions as in [57] (catalyst 22), the same authors achieved analogical N-arylations of N(H)-heterocycles with aryl halogenides and arylboronic acids with good or excellent results with the use of heterogeneous catalysts based on complexes of polystyrene–Schiff bases (PS-4-C6H4N=CH-2-pyridyl Cu(OAc)2) [58], or (PS-4-C6H4N=CH-2-furyl Cu(OAc)2) [59].
Another variant of this reaction is the N-arylation of imidazole, pyrazole, 1,2,4-triazole, benzylamine and substituted anilines with substituted iodobenzenes catalysed by catalyst 23 (Scheme 15) [60].
Catalyst 23 was prepared from commercial chloromethylated polystyrene (2% DVB), on which 5-phenyl-1H-tetrazole was attached, and CuCl2 was coordinated (Scheme 15). The reactions were performed in DMSO at 120 °C for 12–16 h in the presence of K2CO3 (N-heterocycles; 67–95%), or in the presence of KOH (amines; 67–85%). In different solvents (THF, MeCN, DME, N-methylpyrrolidone (NMP), DMF) and under different conditions (Cs2CO3, NaHCO3, K3PO4, TEA; 80 °C, 130 °C) N-arylation of 1H-imidazole with iodobenzene occurred with lower yields than the reaction in DMSO. The recovery and reusing of catalyst 23 was tested with success for the reaction of imidazole with iodobenzene (Cycle 1: 94%; Cycle 6: 89%). The N-arylation of amides and lactams with substituted iodobenzenes, bromobenzene, 2-bromopyridine or 2-iodopyridine catalysed by 24 was also successful (Scheme 16) [61].
Catalyst 24 was prepared from commercial chloromethylated polystyrene, where the chloromethyl group was transformed to an aldehydic group. Subsequent reaction with β-alanine led to the production of an imine, which was then coordinated with Cu(II) acetate. N-arylations were performed without a solvent and in the presence of K2CO3 at 120 °C, for 15–20 h, with yields in the range of 62–92%. The recovery and reusing of catalyst 24 was successfully tested for the reaction of benzamide with iodobenzene (Cycle 1: 92%; Cycle 8: 89%) (Scheme 16) [61]. Further examples of reactions proceeding in aqueous media were N-arylation of substituted bromobenzenes with substituted anilines and (cyclo)alkylamines, catalysed by catalyst 25 in the presence of tetrabutylammonium bromide (TBAB) (Scheme 17) [62]. Catalyst 25 was prepared by the sequence of reactions from atactic polystyrene. Firstly, polystyrene was nitrated, reduced to amine, diazotated, and reduced to hydrazine, which was then acylated by 1H-pyrrole-2-carboxylic acid. The catalyst 25 itself was generated directly in the reaction mixture from the functionalized polystyrene and Cu(I) iodide.
The reaction of substituted bromobenzenes with benzylamine and substituted anilines (70 °C, 16 h) proceeded with very good yields of up to 82%. The lowest yield was attained in the reaction of 4-bromonitrobenzene with benzylamine (44%), while the reaction with sterically demanding 2,4,6-trimethylaniline did not proceed. Analogical reactions with (cyclo)alkylamines (90 °C, 16 h) provided yields of up to 93%, while the lowest yield was achieved in the reaction of 4-bromonitrobenzene with butylamine (51%). Substituted bromopyridines provided products with yields of 33–85%. The reactions of substituted iodobenzenes with imidazole (120 °C, 16 h) led to products with yields ranging from 65% to 87%. The lowest yield was obtained for iodobenzene itself (65%), while the reaction with 2-methoxyiodobenzene did not proceed. Catalyst 25 was also applied in the intramolecular N-arylation of N-(2-iodophenyl)-1H-imidazole-2-carboxamide giving imidazo[1,2-a]quinoxaline. The recovery and reusing of catalyst 25 was successfully tested for the reaction of 4-methoxyiodobenzene with imidazole (Cycle 1: 79%; Cycle 5: 82%) (Scheme 17) [62].

3.4. Aza-Michael Addition

Nucleophilic addition of N-nucleophiles on activated multiple bonds (aza-Michael addition) is a significant reaction in the synthesis of β-amino carbonyl compounds as crucial intermediates in the synthesis of β-amino acids, β-lactam antibiotics, and chiral auxiliaries [63]. Simple, green and efficient preparation of N-alkylated imidazoles was tested, with a reaction of substituted imidazoles with different α,β-unsaturated compounds catalysed by polystyrene-supported CuI–imidazole complex catalyst 26 (Scheme 18) [64]. Catalyst 26 was prepared via surfactant free emulsion polymerization with 4-VBC (4-vinylbenzylchloride) and DVB (2%), whereby uniform globular particles (400–500 nm) with high Cl content (21.5%) were obtained initially. Afterwards, these particles were highly functionalized by imidazole and coordinated by CuI.
The efficiency of catalyst 26 was firstly confirmed for the addition of imidazole on acrylonitrile (8.5 mol % Cu; 60 °C; 4–10 h); while the influence of the solvent was studied: methanol (58%); MeCN (63%); DMSO (52%); DMF (71%); acetone (47%); THF (58%). The reactions were then performed in DMF and the temperature optimized to 75 °C. The addition of substituted imidazoles on α,β-unsaturated substrates proceeded with very good yields up to 74–84%. The lowest yield was achieved for the addition of 4-nitro-1H-imidazole on acrylonitrile. Catalyst 26 was separated by filtration and was further recycled for the reaction of imidazole with acrylonitrile (Cycle 1: 92%; Cycle 2: 91%; Cycle 5: 89%). The given results show that catalyst 26 represents an excellent reusable catalyst; which could be potentially applied at an industrial scale [64].

4. Conclusions

In this mini-review, we focused on Cu(I) and Cu(II) complexes anchored on pearl-like polystyrenes. The prepared heterogeneous catalysts were beneficially used for oxidative homocoupling of terminal alkynes, synthesis of propargylamines, nitroaldolization reaction, azide alkyne cycloaddition, N-arylation of nitrogen containing compounds, aza-Michael addition asymmetric Friedel–Crafts reactions, asymmetric Mukaiyama aldol reactions, and asymmetric 1,3-dipolar cycloaddition of azomethine ylides. In several given examples, any difference between the efficiency of individual catalysts was not found depending on the method of their preparation (copolymerization strategy or post-modification strategy) [15,16,17,18,19]. In a prevalent number of examples, any significant decrease in chemical yields was not observed in the comparison of heterogeneous catalysts with corresponding homogeneous catalysts. The advantages of discussed heterogeneous catalysts consisted in their simple preparation from inexpensive and easily available starting materials, their simple isolation, and the possibility of reuse. Easy performance of the reaction in aqueous media brings great benefits in the case of cross-coupling reactions [13], Huisgen cycloaddition [49], N-arylation [62], and solvent-free N-arylation [61]. Most of the mentioned reactions possess high potential for industrial usage, hence the catalytic systems presented in this review significantly contribute to further development of ecologically sustainable chemical processes and technologies.

Acknowledgments

The authors thank the Czech Science Foundation (project number 17-08499S) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van Santen, R.; Khoe, D.; Vermeer, B. Sustainable materials. In 2030 Technology that will Change the World; Oxford University Press: New York, NY, USA, 2010. [Google Scholar]
  2. Anastas, P.T.; Warner, J.C. Tools of Green Chemistry. In Green Chemistry: Theory and Practice; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  3. Cozzi, F. Immobilization of organic catalysts: When, why, and how. Adv. Synth. Catal. 2006, 348, 1367–1390. [Google Scholar] [CrossRef]
  4. Benaglia, M. Insoluble Resin-supported Catalysts. In Recoverable and Recyclable Catalysts; John Wiley & Sons: Wiltshire, UK, 2009. [Google Scholar]
  5. Jimeno, C.; Sayalero, S.; Pericàs, M.A. Covalent Heterogenization of Asymmetric Catalysts on Polymers and Nanoparticles. In Heterogenized Homogenous Catalysts for Fine Chemicals Production: Materials and Processes; Barbaro, P., Liguori, F., Eds.; Springer Science and Business Media: Dordrecht, the Netherlands, 2010. [Google Scholar]
  6. Merrifield, R.B.J. Solid phase peptide synthesis. I. The synthesis of a tetrapeptide. J. Am. Chem. Soc. 1963, 85, 2149–2154. [Google Scholar] [CrossRef]
  7. Kristensen, T.E.; Hansen, T. Polymer-supported chiral organocatalysts: Synthetic strategies for the road towards affordable polymeric immobilization. Eur. J. Org. Chem. 2010, 3179–3204. [Google Scholar] [CrossRef]
  8. Glaser, C. Beiträge zur Kenntniss des Acetenylbenzols. Ber. Dtsch. Chem. Ges. 1869, 2, 422–424. [Google Scholar] [CrossRef]
  9. Sindhu, K.S.; Anilkumar, G. Recent advances and applications of Glaser coupling employing greener protocols. RSC Adv. 2014, 4, 27867–27887. [Google Scholar] [CrossRef]
  10. Yan, S.; Pan, S.; Osako, T.; Uozumi, Y. Recyclable polystyrene-supported copper catalysts for the aerobic oxidative homocoupling of terminal alkynes. Synlett 2016, 27, 1232–1236. [Google Scholar] [CrossRef]
  11. Sun, Q.; Lv, Z.; Du, Y.; Wu, Q.; Wang, L.; Zhu, L.; Meng, X.; Chen, W.; Xiao, F.-S. Recyclable porous polymer-supported copper catalysts for Glaser and Huisgen 1,3-dipolar cycloaddition reactions. Chem. Asian J. 2013, 8, 2822–2827. [Google Scholar] [CrossRef] [PubMed]
  12. Barot, N.; Patel, S.B.; Kaur, H. Nitro resin supported copper nanoparticles: An effective heterogeneous catalyst for C–N cross coupling and oxidative C–C homocoupling. J. Mol. Catal. A Chem. 2016, 423, 77–84. [Google Scholar] [CrossRef]
  13. Kodicherla, B.; Perumgani, P.C.; Mandapati, M.R. A reusable polystyrene-supported copper(II) catalytic system for N-arylation of indoles and Sonogashira coupling reactions in water. Appl. Catal. A Gen. 2014, 483, 110–115. [Google Scholar] [CrossRef]
  14. Doyle, M.P.; McKervey, M.A.; Ye, T. Catalysts for Metal Carbene Transformations. In Modern Catalytic Methods for Organic Synthesis with Diazo Compounds: From Cyclopropanes to Ylides; Wiley-VCH: Weinheim, Germany, 1998. [Google Scholar]
  15. Mandoli, A.; Orlandi, S.; Pini, D.; Salvadori, P. A reusable, insoluble polymer-bound bis(oxazoline) (IPB-box) for highly enantioselective heterogeneous cyclopropanation reactions. Chem. Commun. 2003, 2466–2467. [Google Scholar] [CrossRef]
  16. Mandoli, A.; Garzelli, R.; Orlandi, S.; Pini, D.; Lessi, M.; Salvadori, P. Some factors affecting the catalytic efficiency in the enantioselective cyclopropanation of olefins by the use of insoluble polystyrene-bound bisoxazoline-copper(I) complex. Catal. Today 2009, 140, 51–57. [Google Scholar] [CrossRef]
  17. Burguete, M.I.; Fraile, J.M.; García, J.I.; García-Verdugo, E.; Herrerías, C.I.; Luis, S.V.; Mayoral, J.A. Bis(oxazoline)copper complexes covalently bonded to insoluble support as catalysts in cyclopropanation reactions. J. Org. Chem. 2001, 66, 8893–8901. [Google Scholar] [CrossRef] [PubMed]
  18. Knight, J.; Belcher, P.E. 2-Aryl-5,5-bisoxazolin-2-yl[1,3]dioxanes as solution phase and immobilised ligands for highly enantioselective cyclopropanations. Tetrahedron Asymmetry 2005, 16, 1415–1418. [Google Scholar] [CrossRef]
  19. Werner, H.; Herrerías, C.I.; Glos, M.; Gissibl, A.; Fraile, J.M.; Péres, I.; Mayoral, J.A.; Reiser, O. Synthesis of polymer bound azabis(oxazoline) ligands and their application in asymmetric cyclopropanations. Adv. Synth. Catal. 2006, 348, 125–132. [Google Scholar] [CrossRef]
  20. Burguete, M.I.; Fraile, J.M.; García-Verdugo, E.; Luis, S.V.; Martínez-Merino, V.; Mayoral, J.A. Polymer-supported bis(oxazolines) and related systems: Toward new heterogeneous enantioselective catalysts. Ind. Eng. Chem. Res. 2006, 44, 8580–8587. [Google Scholar] [CrossRef]
  21. Maestre, L.; Ozkal, E.; Ayats, C.; Beltrán, A.; Díaz-Requejo, M.M.; Pérez, P.J.; Pericàs, M.A. A fully recyclable heterogenized Cu catalyst for the general carbene transfer reaction in batch and flow. Chem. Sci. 2015, 6, 1510–1515. [Google Scholar] [CrossRef]
  22. Peshkov, V.A.; Pereshivko, O.P.; Van der Eycken, E.V. A walk around the A3-coupling. Chem. Soc. Rev. 2012, 41, 3790–3807. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, L.; Cai, C. Reusable polymer-anchored amino acid copper complex for the synthesis of propargylamines. J. Chem. Res. 2008, 32, 538–541. [Google Scholar] [CrossRef]
  24. Islam, M.; Roy, A.S.; Islam, S.M. Functionalized polystyrene supported copper(I) complex as an effective and reusable catalyst for propargylamines synthesis in aqueous medium. Catal. Lett. 2016, 146, 1128–1137. [Google Scholar] [CrossRef]
  25. Kodicherla, B.; Perumgani, P.C.; Mandapati, M.R. Polymer-anchored copper(II) complex: An efficient reusable catalyst for the synthesis of propargylamines. Appl. Organometal. Chem. 2014, 28, 756–759. [Google Scholar] [CrossRef]
  26. Perumgani, P.C.; Keesara, S.; Parvathaneni, S.; Mandapati, M.R. Polystyrene supported N-phenylpiperazine–Cu(II) complex: An efficient and reusable catalyst for KA2-coupling reactions under solvent-free conditions. New J. Chem. 2016, 40, 5113–5120. [Google Scholar] [CrossRef]
  27. Gommermann, N.; Knochel, P. Practical highly enantioselective synthesis of propargylamines through a copper-catalyzed one-pot three-component condensation reaction. Chem. Eur. J. 2006, 12, 4380–4392. [Google Scholar] [CrossRef] [PubMed]
  28. Nakamura, S.; Ohara, M.; Nakamura, Y.; Shibata, N.; Toru, T. Copper-catalyzed enantioselective three-component synthesis of optically active propargylamines from aldehydes, amines, and aliphatic alkynes. Chem. Eur. J. 2010, 16, 2360–2362. [Google Scholar] [CrossRef] [PubMed]
  29. Naeimi, H.; Moradian, M. Thioether-based copper(I) Schiff base complex as a catalyst for a direct and asymmetric A3-coupling reaction. Tetrahedron Asymmetry 2014, 25, 427–434. [Google Scholar] [CrossRef]
  30. Drabina, P.; Funk, P.; Růžička, A.; Hanusek, J.; Sedlák, M. Synthesis, copper(II) complexes and catalytic activity of substituted 6-(1,3-oxazolin-2-yl)pyridine-2-carboxylates. Transit. Met. Chem. 2010, 35, 363–371. [Google Scholar] [CrossRef]
  31. Zeng, T.; Yang, L.; Hudson, R.; Song, G.; Moores, A.R.; Li, C.-J. Fe3O4 Nanoparticle-supported copper(I) pybox catalyst: Magnetically recoverable catalyst for enantioselective direct-addition of terminal alkynes to imines. Org. Lett. 2011, 13, 442–445. [Google Scholar] [CrossRef] [PubMed]
  32. Sasai, H.; Suzuki, T.; Arai, S.; Arai, T.; Shibasaki, M. Basic character of rare earth metal alkoxides. Utilization in catalytic C–C bond-forming reactions and catalytic asymmetric nitroaldol reactions. J. Am. Chem. Soc. 1992, 114, 4418–4420. [Google Scholar] [CrossRef]
  33. Ananthi, N.; Velmathi, S. Asymmetric Henry reaction catalyzed by transition metal complexes: A short review. Ind. J. Chem. 2013, 52, 87–108. [Google Scholar]
  34. Corey, E.J.; Zhang, F. re- and si-Face-selective nitroaldol reactions catalyzed by a rigid chiral quaternary ammonium salt: A highly stereoselective synthesis of the HIV protease inhibitor Amprenavir (Vertex 478). Angew. Chem. Int. Ed. 1999, 38, 1931–1934. [Google Scholar] [CrossRef]
  35. Drabina, P.; Harmand, L.; Sedlák, M. Enantioselective Henry reaction catalyzed by supported transition metal complexes. Curr. Org. Synth. 2014, 11, 879–888. [Google Scholar] [CrossRef]
  36. Harmand, L.; Drabina, P.; Pejchal, V.; Husáková, L.; Sedlák, M. Recyclable catalyst for the asymmetric Henry reaction based on functionalized imidazolidine-4-one-copper(II) complexes supported by a polystyrene copolymer. Tetrahedron Lett. 2015, 56, 6240–6243. [Google Scholar] [CrossRef]
  37. Panov, I.; Drabina, P.; Padělková, Z.; Šimůnek, P.; Sedlák, M. Highly enantioselective nitroaldol reactions catalyzed by copper(II) complexes derived from substituted 2-(pyridin-2-yl)imidazolidin-4-one ligands. J. Org. Chem. 2011, 76, 4787–4793. [Google Scholar] [CrossRef] [PubMed]
  38. Nováková, G.; Drabina, P.; Frumarová, B.; Sedlák, M. Recyclable enantioselective catalysts based on copper(II) complexes of 2-(pyridine-2-yl)imidazolidine-4-thione: Their application in asymmetric Henry reactions. Adv. Synth. Catal. 2016, 358, 2541–2552. [Google Scholar] [CrossRef]
  39. Arai, T.; Yokoyama, N.; Yanagisawa, A. A library of chiral imidazoline–aminophenol ligands: Discovery of an efficient reaction sphere. Chem. Eur. J. 2008, 14, 2052–2059. [Google Scholar] [CrossRef] [PubMed]
  40. Desyatkin, V.G.; Anokhin, M.V.; Rodionov, V.O.; Beletskaya, I.P. Polystyrene-supported Cu(II)-R-Box as recyclable catalyst in asymmetric Friedel–Crafts reaction. Russ. J. Org. Chem. 2016, 52, 1717–1727. [Google Scholar] [CrossRef]
  41. Orlandi, S.; Mandoli, A.; Pini, D.; Salvadori, P. An insoluble polymer-bound bis-oxazoline copper(II) complex: A highly efficient heterogeneous catalyst for the enantioselective Mukaiyama aldol reaction. Angew. Chem. Int. Ed. 2001, 40, 2519–2521. [Google Scholar] [CrossRef]
  42. Fraile, J.M.; Pérez, I.; Mayoral, J.A. Comparison of immobilized Box and azaBox–Cu(II) complexes as catalysts for enantioselective Mukaiyama aldol reactions. J. Catal. 2007, 252, 303–311. [Google Scholar] [CrossRef]
  43. Huisgen, R. 1,3-Dipolar Cycloadditions. Proc. Chem. Soc. 1961, 357–396. [Google Scholar]
  44. Tornøe, C.W.; Christensen, C.; Medal, M. Peptidotriazoles on solid phase: [1,2,3]-Triazoles by regiospecific copper(I)-catalyzed 1,3-dipolar cycloadditions of terminal alkynes to azides. J. Org. Chem. 2002, 67, 3057–3064. [Google Scholar]
  45. Rostovtset, V.V.; Green, L.G.; Fokin, V.V.; Sharpless, K.B.A. Stepwise Huisgen cycloaddition process: Copper(I)-catalyzed regioselective ligation of azides and terminal alkynes. Angew. Chem. Int. Ed. 2002, 41, 2596–2599. [Google Scholar] [CrossRef]
  46. Kolb, H.C.; Finn, M.G.; Sharpless, K.B.A. Click chemistry: Diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  47. Girard, C.; Onen, E.; Aufort, M.; Beauviere, S.; Samson, E.; Herscovici, J. Reusable polymer-supported catalyst for the [3+2] Huisgen cycloaddition in automation protocols. Org. Lett. 2006, 8, 1689–1692. [Google Scholar] [CrossRef] [PubMed]
  48. Chassaing, S.; Bénéteau, V.; Pale, P. When CuAAC Click Chemistry′ goes heterogenous. Catal. Sci. Technol. 2016, 6, 923–957. [Google Scholar] [CrossRef]
  49. Movassagh, B.; Rezaei, N. Polystyrene resin-supported CuI-cryptand 22 complex: A highly efficient and reusable catalyst for three-component synthesis of 1,4-disubstituted 1,2,3-triazoles under aerobic conditions in water. Tetrahedron 2014, 70, 8885–8892. [Google Scholar] [CrossRef]
  50. Rezaei, N.; Movassagh, B. Polystyrene resin-supported CuI-cryptand 22 complex: A highly efficient and reusable catalyst for the formation of aryl-sulfur bonds in aqueous media. Tetrahedron Lett. 2016, 57, 1625–1628. [Google Scholar] [CrossRef]
  51. Tavassoli, M.; Landarani-Isfahani, A.; Moghadam, M.; Tangestaninejad, S.; Valliolah Mirkhani, V.; Mohammadpoor-Baltork, I. Polystyrene-supported ionic liquid copper complex: A reusable catalyst for one-pot three-component click reaction. Appl. Catal. A 2015, 503, 186–195. [Google Scholar] [CrossRef]
  52. Martín-Matute, B.; Pereira, S.I.; Peña-Cabrera, E.; Adrio, J.; Silva, A.M.S.; Carretero, J.C. Synthesis of polymer-supported Fesulphos ligands and their application in asymmetric catalysis. Adv. Synth. Catal. 2007, 349, 1714–1724. [Google Scholar] [CrossRef]
  53. Maurya, M.R.; Sikarwar, S.; Joseph, T.; Halligudi, S.B. Bis(2-[α-hydroxyethyl]benzimidazolato)copper(II) anchored onto chloromethylated polystyrene for the biomimetic oxidative coupling of 2-aminophenol to 2-aminophenoxazine-3-one. J. Mol. Catal. A Chem. 2005, 236, 132–138. [Google Scholar] [CrossRef]
  54. Hassan, J.; Sévignon, M.; Gozzi, C.; Schulz, E.; Lemaire, M. Aryl-aryl bond formation one century after the discovery of the Ullmann reaction. Chem. Rev. 2002, 102, 1359–1470. [Google Scholar] [CrossRef] [PubMed]
  55. Ullmann, F.; Jean Bielecki, J. Ueber Synthesen in der Biphenylreihe. Chem. Ber. 1901, 34, 2174–2185. [Google Scholar] [CrossRef]
  56. Lan, J.-B.; Chen, L.; Yu, X.-Q.; You, J.-S.; Xie, R.-G. A simple copper salt catalyzed the coupling of imidazole with arylboronic acids in protic solvent. Chem. Commun. 2004, 188–189. [Google Scholar] [CrossRef] [PubMed]
  57. Islam, S.M.; Mondal, S.; Mondal, P.; Roy, A.S.; Tuhina, K; Salam, N.; Mobarak, M. A reusable polymer supported copper catalyst for the C–N and C–O bond cross-coupling reaction of aryl halides as well as arylboronic acids. J. Organomet. Chem. 2012, 696, 4264–4274. [Google Scholar] [CrossRef]
  58. Islam, S.M.; Mondal, S.; Mondal, P.; Roy, A.S.; Tuhina, K.; Salam, N.; Sumantra, P.; Hossain, D.; Mobarok, M. An efficient polymer-supported copper(II) catalyst for the N-arylation reaction of N(H)-heterocycles with aryl halides as well as arylboronic acids. Transit. Met. Chem. 2011, 36, 447–458. [Google Scholar] [CrossRef]
  59. Islam, M.; Salam, N.; Mondal, P.; Roy, A.S.; Ghosh, K.; Tuhina, K. A highly active reusable polymer anchored copper catalyst for C–O, C–N and C–S cross coupling reactions. J. Mol. Catal. A Chem. 2014, 387, 7–19. [Google Scholar] [CrossRef]
  60. Nasrollahzadeh, M.; Zahraei, A.; Pourbasheer, E. Catalytic activity and antibacterial properties of nanopolymer-supported copper complex for C–N coupling reactions of amines and nitrogen-containing heterocycles with aryl halides. Monatsh. Chem. 2015, 146, 1329–1334. [Google Scholar] [CrossRef]
  61. Islam, M.M.; Halder, H.; Roy, A.S.; Chatterjee, S.; Bhaumik, A.; Islam, S.M. Copper(II) incorporated functionalized polystyrene catalyzed N-arylation of amides under solvent free condition with broad substrate scope. RSC Adv. 2016, 6, 109692–109701. [Google Scholar] [CrossRef]
  62. Huang, L.; Yu, R.; Zhu, X.; Wan, Y. A recyclable Cu-catalyzed C–N coupling reaction in water and its application to synthesis of imidazo[1,2-a]quinoxaline. Tetrahedron 2013, 69, 8974–8977. [Google Scholar] [CrossRef]
  63. Cardillo, G.; Tomasini, C. Asymmetric synthesis of β-amino acids and α-substituted β-amino acids. Chem. Soc. Rev. 1996, 25, 117–128. [Google Scholar] [CrossRef]
  64. Li, L.; Liu, Z.; Ling, Q.; Xing, X. Polystyrene-supported CuI–imidazole complex catalyst for aza-Michael reaction of imidazoles with α,β-unsaturated compounds. J. Mol. Catal. A Chem. 2012, 353, 178–184. [Google Scholar] [CrossRef]
Scheme 1. Oxidative homocoupling of terminal alkynes catalysed by polymeric complex 1.
Scheme 1. Oxidative homocoupling of terminal alkynes catalysed by polymeric complex 1.
Molecules 22 00865 sch001
Scheme 2. Sonogashira cross-coupling reaction catalysed by catalyst 2.
Scheme 2. Sonogashira cross-coupling reaction catalysed by catalyst 2.
Molecules 22 00865 sch002
Scheme 3. Cyclopropanation catalysed by different polystyrene-supported complexes 37.
Scheme 3. Cyclopropanation catalysed by different polystyrene-supported complexes 37.
Molecules 22 00865 sch003
Scheme 4. Carbene transfer reactions catalysed by polymeric complex 8.
Scheme 4. Carbene transfer reactions catalysed by polymeric complex 8.
Molecules 22 00865 sch004
Scheme 5. Addition of terminal alkynes to imines catalysed by catalysts 9 and 10.
Scheme 5. Addition of terminal alkynes to imines catalysed by catalysts 9 and 10.
Molecules 22 00865 sch005
Scheme 6. Addition of terminal alkynes to imines catalysed by catalysts 2 and 11.
Scheme 6. Addition of terminal alkynes to imines catalysed by catalysts 2 and 11.
Molecules 22 00865 sch006
Scheme 7. Henry reaction catalysed by copper(II) complex 13 and its immobilized form 12.
Scheme 7. Henry reaction catalysed by copper(II) complex 13 and its immobilized form 12.
Molecules 22 00865 sch007
Scheme 8. Henry reaction catalysed by copper(II) complexes 15a,b and their immobilized forms 14a,b.
Scheme 8. Henry reaction catalysed by copper(II) complexes 15a,b and their immobilized forms 14a,b.
Molecules 22 00865 sch008
Scheme 9. Asymmetric Friedel–Crafts alkylation of indole catalysed by polymeric complex 16.
Scheme 9. Asymmetric Friedel–Crafts alkylation of indole catalysed by polymeric complex 16.
Molecules 22 00865 sch009
Scheme 10. Asymmetric Mukaiyama aldol reactions catalysed by complexes of Cu(OTf)2/3 and/or 17.
Scheme 10. Asymmetric Mukaiyama aldol reactions catalysed by complexes of Cu(OTf)2/3 and/or 17.
Molecules 22 00865 sch010
Scheme 11. Huisgen 1,3-dipolar cycloaddition catalysed by polystyrene supported complex 18.
Scheme 11. Huisgen 1,3-dipolar cycloaddition catalysed by polystyrene supported complex 18.
Molecules 22 00865 sch011
Scheme 12. Huisgen 1,3-dipolar cycloaddition catalysed by polystyrene supported complex 19.
Scheme 12. Huisgen 1,3-dipolar cycloaddition catalysed by polystyrene supported complex 19.
Molecules 22 00865 sch012
Scheme 13. Asymmetric 1,3-dipolar cycloadditions catalysed by polymeric complexes of Cu(MeCN)4ClO4 with 20 and or 21.
Scheme 13. Asymmetric 1,3-dipolar cycloadditions catalysed by polymeric complexes of Cu(MeCN)4ClO4 with 20 and or 21.
Molecules 22 00865 sch013
Scheme 14. N-Arylation of N-nucleophiles by substituted phenylboronic acids catalysed by polystyrene supported complex 22.
Scheme 14. N-Arylation of N-nucleophiles by substituted phenylboronic acids catalysed by polystyrene supported complex 22.
Molecules 22 00865 sch014
Scheme 15. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 23.
Scheme 15. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 23.
Molecules 22 00865 sch015
Scheme 16. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 24.
Scheme 16. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 24.
Molecules 22 00865 sch016
Scheme 17. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 25.
Scheme 17. N-Arylation of N-nucleophiles by substituted aryl halides catalysed by polystyrene supported complex 25.
Molecules 22 00865 sch017
Scheme 18. Aza-Michael addition catalysed by polystyrene supported complex 26.
Scheme 18. Aza-Michael addition catalysed by polystyrene supported complex 26.
Molecules 22 00865 sch018

Share and Cite

MDPI and ACS Style

Drabina, P.; Svoboda, J.; Sedlák, M. Recent Advances in C–C and C–N Bond Forming Reactions Catalysed by Polystyrene-Supported Copper Complexes. Molecules 2017, 22, 865. https://doi.org/10.3390/molecules22060865

AMA Style

Drabina P, Svoboda J, Sedlák M. Recent Advances in C–C and C–N Bond Forming Reactions Catalysed by Polystyrene-Supported Copper Complexes. Molecules. 2017; 22(6):865. https://doi.org/10.3390/molecules22060865

Chicago/Turabian Style

Drabina, Pavel, Jan Svoboda, and Miloš Sedlák. 2017. "Recent Advances in C–C and C–N Bond Forming Reactions Catalysed by Polystyrene-Supported Copper Complexes" Molecules 22, no. 6: 865. https://doi.org/10.3390/molecules22060865

Article Metrics

Back to TopTop