Next Article in Journal
Thermal Inactivation Mechanism and Structural Features Providing Enhanced Thermal Stability of Hyperthermophilic Thermococcus sibiricus L-Asparaginase in Comparison with Mesophilic and Thermophilic L-Asparaginases
Next Article in Special Issue
Synthesis and Applications of Copper-Based Catalysts
Previous Article in Journal
Investigation on the Edge Doping Process of Nitrogen-Doped Carbon Materials by In Situ Pyrolysis Mass Spectrometry and Laser-Induced Acoustic Desorption Mass Spectrometry
Previous Article in Special Issue
Progress on the Cu-Catalyzed 1,4-Conjugate Addition to Thiochromones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Achievements in the Copper-Catalyzed Arylation of Adamantane-Containing Amines, Di- and Polyamines

by
Alexei D. Averin
1,*,
Svetlana P. Panchenko
1,
Arina V. Murashkina
1,
Varvara I. Fomenko
1,
Daria S. Kuliukhina
1,
Anna S. Malysheva
1,
Alexei A. Yakushev
1,
Anton S. Abel
1 and
Irina P. Beletskaya
1,2
1
Department of Chemistry, Lomonosov Moscow State University, Leninskie Gory, 1-3, Moscow 119991, Russia
2
A.N. Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Leninsky Pr. 31-4, Moscow 119071, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(5), 831; https://doi.org/10.3390/catal13050831
Submission received: 31 March 2023 / Revised: 23 April 2023 / Accepted: 26 April 2023 / Published: 1 May 2023
(This article belongs to the Special Issue Synthesis and Applications of Copper-Based Catalysts)

Abstract

:
Rapid development of the copper-catalyzed amination of aryl halides in the beginning of the 21st century, known as the Renaissance of the Ullmann chemistry, laid foundations for the use of this method as a powerful tool for the construction of the C(sp2)-N bond and became a rival of the Buchwald–Hartwig amination reaction. Various applications of this approach are well-documented in a number of comprehensive and more specialized reviews, and this overview in the form of a personal account of the Cu-catalyzed arylation and heteroarylation of the adamantane-containing amines, and di- and polyamines, covers a more specific area, showing the possibilities of the method and outlining general regularities, considering reagents structure, copper source and ligands, scope, and limitations. The material of the last decade is mainly considered, and recent data on the application of the unsupported copper nanoparticles and possibilities of the Chan-Lam reaction as an alternative to the use of aryl halides are also discussed.

Graphical Abstract

1. Introduction

The beginning of the 20th century was marked by the birth of the copper-mediated reactions of carbon–carbon and carbon–heteroatom bond formation. In the pioneer research carried out by Ullmann and Goldberg, copper powder was shown to provide access to biphenyls [1], aryl amines [2], diaryl ethers [3], and N-aryl amides [4]. These reactions employed stoichiometric amounts of copper and strong bases, and demanded very harsh conditions; thus, the substrates’ scope was strictly limited. However, the simplicity of the approach allowed for the synthesis of a series of valuable compounds via C–C, C–M, C–O, and C–S bond formation, and some “classical” reactions of this type are still used in industry [5,6,7,8].
The need for compounds bearing the N-(hetero)aryl moiety cannot be overestimated. They are widely used in the synthesis of pharmaceuticals [9,10] and agrochemicals [11,12], and for the fabrication of modern organic materials [13,14]. It is obvious that convenient, reliable, and inexpensive methods for the synthesis of compounds with the C(sp2)–N bond are always highly demanded. The so-called “Renaissance of the Ullmann chemistry” dates back to the beginning of the 21st century and is associated with the application of the ligands which allowed for the use of copper in catalytic amounts and the application of rather mild reaction conditions. To note, the effect of ketones or esters on the rate of the Ullmann reactions was noted even in 1964 [15]; however, at that time, it was thought to be a result of a better solubility of the copper compounds. In 1997, the reaction of phenols with aryl halides containing triazene fragments at the ortho position was reported [16], and this fragment was described as an appended ligand for the copper catalyst. A real breakthrough was a series of works by Buchwald [17], Ma [18], and Taillefer [19] which used the catalytic systems on the basis of in situ-formed Cu(II) complexes with various ligands. This approach proved to be exceedingly fruitful and further development of the Cu-catalyzed amination and amidation reactions has been decisively linked with the design and investigation of the ligands [20,21]. These ligands can be divided into three main groups: N,N-ligands [22,23,24,25,26], N,O-ligands [27,28,29,30,31], and O,O-ligands [32,33,34,35,36], of which the most widely used and efficient representatives are shown in Figure 1. The role of the anionic ligands, which favor the oxidative addition and thus ensure milder conditions and a wider scope of the aryl halides, was emphasized in [37,38,39], the oxalamide ligands being essentially efficient for this purpose [40,41,42].
The development of the copper-catalyzed amination has been already described in several comprehensive reviews. Two of them were published before 2010 by Ma [43] and Taillefer [19]. Two other reviews are dedicated to the comparison of the copper- and palladium-catalyzed amination reactions [44,45]. Further, a more thorough review has been published [46], and some accounts emphasize the application of the method to the synthesis of natural compounds [47,48]. During the last decade, the achievements in the copper-catalyzed amination became a focus of three comprehensive studies [49,50,51], and special cases of this approach have also been discussed recently [52,53].
One of the important trends in the study of the Cu-catalyzed amination and amidation reactions is the use of heterogeneous and heterogenized copper catalysts immobilized on various supports. The most frequently used copper salt for this purpose is CuI. The nature of supports is versatile, and different organic linkers may serve as the ligands for the copper cation. Cu catalysts are often immobilized on polymer supports [54,55,56,57,58,59,60]. For the needs of the development of reusable catalysts, those containing magnetite particles were found to be especially attractive [61,62,63,64].
The studies on the mechanism of Cu-catalyzed amination put forward several hypotheses. Taking into consideration the fact that copper in different oxidation states can be successfully employed in catalysis, it was supposed that Cu(I) is a genuine catalytic particle which emerges in the course of the reduction process of Cu(II) compounds [15,65,66,67]. In some protocols, additional reagents such as ascorbic acid are added to ensure this reduction [68,69], but many others do not envisage any special reductant [70,71]. In the case of Cu(0), it is thought that, on contrary, its oxidation into Cu2O gives rise to a catalytically active Cu(I) [65,72]. The main types of proposed mechanisms are the following: (1) oxidative addition of ArHal leading to Cu(III) complex followed by the reductive elimination; (2) formation of the π-complex of aryl halide with Cu(I) followed by the nucleophilic substitution; (3) single electron transfer (SET) or hydrogen atom transfer (HAT); (4) σ-bond metathesis. The majority of researchers support the idea of a key role of Cu(III) intermediate [46,73,74,75,76,77]; for this purpose, some researchers synthesized several stable complexes of Cu(III) in order to study the coordination sphere of the metal cation and to carry out their reactions with nucleophiles [78,79,80]. Ma investigated a series of reactions without additional ligands and postulated that, in this case, aryl halides, acting as ligands, form the π-complexes with the copper cation [27]. SET and HAT mechanisms involving radicals were thought to be verified [66] by the use of the radical traps which hindered the reactions. At last, metathesis with a four-membered transition state was proposed by Bacon [81]; this mechanism suggests the coordination of the copper cation with the halogen atom of ArHal. It is quite plausible that, with different reagents and in the presence of various ligands, different mechanisms can be actualized.
Our research interests in the field of Cu-catalyzed amination cover the reactions with adamantane-containing amines, diamines, oxadiamines, and polyamines. These substrates are interesting due to their often complicated and unpredictable reactivity in spite of the fact that the reactions proceed at primary amino groups. The outcome of the reactions is strongly dependent on the spatial environment of the amino group in the case of adamantaneamines, and the number of methylene groups, or oxygen or nitrogen atoms in the chain for di- and polyamines. Previously, we thoroughly studied the (hetero)arylation of these compounds in the presence of palladium catalysts [21] and found the scope and limitations of this approach. In the following investigations, we compared the possibilities of the application of a much cheaper copper and found the regularities of this process. Our ongoing research deals with the employment of the copper nanoparticles instead of the homogeneous CuI-catalyzed reactions as well as the search for new applications of the Chan-Lam reaction; the present review will describe all of these versions of the C(sp2)-N bond formation. The main objective of this review is to show the scope and limitations of the copper catalysis with regard to the arylation and heteroarylation of these amines, and the following aspects will be addressed: the dependence of the most appropriate catalytic system on the nature of reagents, possibilities of N,N′-di(hetero)arylation of the diamines, oxadiamines, and polyamines, the use of unsupported copper nanoparticles, and the character of the Chan–Lam amination employing adamantaneamines, diamines and oxadiamines.

2. CuI-Catalyzed Arylation of Adamantane-Containing Amines

Adamantane-containing amines and their derivatives are well-known for their versatile biological activity; it is sufficient to mention that amantadine (adamantane-1-amine hydrochloride) has been used as an antiviral and antiparkonsonian drug for decades [82], memantine for the treatment of Alzheimer’s disease [83], and bromantane (N-(4-bromophenyl)adamantane-2-amine) [84] and its fluorine-containing analogue [85] as neurostimulation agents. Thus, it is quite natural that, previously, we paid much attention to the elaboration of efficient methods of N-(hetero)arylation of the adamantane-containing amines employing Pd(0)-catalyzed amination reactions.
In the trend of replacing expensive palladium catalysis with much cheaper copper-catalyzed reactions, we carried out an investigation of the arylation of a series of adamantane-containing amines, 15, differing by the steric hindrances at the amino group, with a model iodobenzene in the presence of various copper catalysts (Scheme 1) [86]. Several widely employed ligands, L1–L7, were tested in the reaction and two of them, L1 (2-isobityrylcyclohexanone) and L2 (rac-BINOL (BINOL = 2,2′-bi(1-naphthol)), both O,O-type, were found to be the most efficient in the reactions run in DMF at 140 °C and catalyzed by a standard CuI in the presence of Cs2CO3. Other solvents which were tested such as propionitrile (EtCN) gave poorer yields, as well as did other copper sources (CuBr, CuOAc, CuOTf, Cu2O, and CuO). Depending on the structure of the amine, they provided the yields of the N-arylation products in the range of 63–90% and were further used for the reactions with adamantane-containing amines. The highest yield (compound 6, 90%) was observed in the reaction with the amine 1, in which the amino group is located furthest from the adamantane core. The attempt to use bromobenzene instead of iodobenzene resulted in a dramatic diminishing of the product yield.
The same amines, 15, were introduced in the reactions with para-substituted iodobenzenes containing typical electron donor and electron withdrawing substituents (Scheme 2). It was interesting to follow the dependence of the yields of the N-arylated products; however, in the majority of cases, they ranged from 50 to 75%, and the advantage of the electron deficient 4-iodotrifluoromethylbenzene over electron donor 4-iodoanisole was quite moderate. Thus, in the reactions of 14, CF3-containing products 12, 15, 19, and 22 were obtained in 55–79% yields, while MeO-containing products 13, 16, 20, and 23 were synthesized in 44–65% yields. Only more sterically hindered amine 5 gave generally lower yields with all of the iodobenzenes which were tested (compounds 2426).
The extension of the series of the adamantane-containing amines (compounds 2740) was undertaken to reveal the dependence of the amines’ structure on the outcome of the copper-catalyzed N-arylation (Scheme 3) [87]. The reactions were run under optimized conditions with iodobenzene and its several p-substituted analogues. The amines 2729, 31, 32, 36, and 3840 provided similar results as described above, obviously, due to a lack of notable steric hindrances. The yields of the corresponding derivatives with the methoxy group were somewhat lower and ranged from 36 to 58%. To note, the presence of the hydroxyl group in the amines 39 and 40 did not alter its good reactivity in the N-arylation reactions. In the case of the amines 30 and 33, steric hindrances were more pronounced, which led to diminished yields of the corresponding products 5356 and 6265. Amines 34 and 35 were less reactive and the use of three equiv. of iodobenzene was inevitable, allowing for a 54% yield of the product 66, but only a 15% yield of 67. The reactivity of the amine 37 was also substantially lower than that of its analogue, 38, probably due to the presence of the ethylenediamine moiety in the first, which could better bind copper than the NCH2CH2CH2N fragment in 38, thus removing it from the catalytic cycle.
The possibility to conduct an N,N′-diarylation reaction of the adamantane-containing diamines 85 and 86 was shown using several iodobenzenes (Scheme 4) [88]. Depending on the substituent in iodobenzene, ligand L1 or L2 was found to be optimal. Better yields were obtained with the diamine 85, where amino groups are located further form the adamantane core, and the yields of the corresponding derivatives 8792 ranged from 50 to 79%. It is interesting that the best yield was observed in the reaction with the electron-enriched p-iodoanisole (compound 92), probably due to fewer side reactions with this less active iodide. In the reactions with the diamine 86, generally, the yields of the diarylated products were notably lower, and several by-products were isolated, evidencing a substantial contribution of the formation of the amine-imine followed by its transformation into the corresponding aldehyde and subsequent processes.
A special investigation was dedicated to the possibilities of the reactions run in DMSO instead of DMF [89]. The reactions with various amines, 1, 35, 28, 29, 34, 35, featuring a different spatial environment of the amino group were carried out with iodobenzene at 110 °C (Scheme 5). The results were quite comparable with those obtained in DMF at 140 °C; moreover, the arylation of the most hindered amines, 34 and 35, allowed for increasing the yields of the corresponding derivatives, 66 and 67, when three equiv. of PhI were employed. The reactions of the amine 1 with various para- and meta-substituted iodobenzenes provided good to high yields of the products 99114 (63–83%). The influence of the electron properties of the substituents was poorly manifested though the yields in the reactions with the electron donors p-iodotoluene and p-iodoanisole, and these were among the smallest (63–64%). DMSO also helped to carry out the reactions with much more problematic ortho-substituted iodobenzenes (R = Me, F, Cl, CN). Further, 2-iodotoluene, taken in a three-fold excess, gave a 66% yield of the arylation product 115, while other o-substituted iodobenzenes provided 47–48% yields of the corresponding compounds 116118.

3. CuI-Catalyzed Heteroarylation of Adamantane-Containing Amines

Molecules which combine adamantane and pyridine moieties in their structure are of great importance due to their versatile biological activities. A 2-aminoadamantyl-containing derivative of pyrrolopyridine known as Peficitinib is used for rheumatoid arthritis treatment [90], and Cu(II) complexes of certain adamantane-substituted pyridines possess anticancer activity [91]. Derivatives of 2-aminopyridine with adamantane fragments were shown to inhibit 11β-hydrosteroid dehydrogenase 11β-HSD1 [92,93,94], and adamantane derivatives of three-substituted pyridine act as antifungal [95] and antinicotinic [96] agents. A great variety of methods exists to introduce heteroaromatic moieties in the adamantane-containing compounds. Interesting radical reactions for the direct transformations of adamantane have been described in a recent review [97]. Thus, quinolinyl-substituted adamantane was obtained via a radical process [98], and modern photocatalytic [99] and photoredox [100] processes were applied for the introduction of quinoline and trifluoromethylpyridine structural fragments into adamantane moiety.
We investigated the possibilities of the Cu(I)-catalyzed amination of 2- and 3-iodopyridines (Scheme 6). Ligands L1-L4 were tested and the best efficiency of L1 was undoubtedly shown. The reactivity of 2-iodopyridine was compared with that of 2-bromopyridine, and while the first afforded a 72% yield in the reaction with the amine 1, the latter provided 57% of the N-pyrid-2-yl containing product 119. A series of the adamantane-containing amines was tested in the reaction with 2-iodopyridine, and those with the unhindered amino group provided good to high yields of the target products (55–90%) [101]. In the case of the more sterically demanding amines 4 and 5, the use of three equiv. of PhI was necessary to increase the yields of 123 and 124; the more hindered amines 35 and 36 provided low yields of the corresponding products 127 and 128 (15 and 36%, respectively).
Similar results were obtained with 3-iodopyridine (Scheme 6). Amines 13, 28, and 31 afforded 67–85% yields of the N-heteroarylated products and, in the case of the more sterically hindered amines 4 and 5, the application of three equiv. of PhI was required to increase the yields of compounds 132, 133. To note, the secondary cyclic amino group in the piperazinyl derivative of adamantane 32 turned out to be less reactive than the primary amino groups, and it also demanded the use of PhI excess.
As a special case, we studied the amination of 2-fluoro-5-iodopyridine with a series of adamantane-containing amines (Scheme 7) [102]. The main peculiarity of this process is the competition of the Cu(I)-catalyzed substitution of the iodine atom and the non-catalytic substitution of the fluorine atom. In the case of the application of 10 mol% catalyst and 1:1 ratio of the reagents, generally, only the mixtures of the products were obtained, but the use of 20 mol% catalyst and two-fold amount of the amine led to a substantial increase in the yields of the target 5-amino-2-fluoropyridines. In some cases (compounds 137, 140, 146, and 147), they reached 92–98%, and it is noteworthy that enough reluctant amines such as 34 and 37 successfully participated in this reaction.
The amines 15 and 27 were introduced in the reactions with other dihalopyridines to explore the selectivity of the iodine substitution (Scheme 8). The reactions with 2-bromo-5-iodopyridine turned out to be totally non-selective, and the amination of the isomeric 5-bromo-2-iodopyridine led to the corresponding products of the iodine substitution 148152 in moderate yields (31–50%). The amination of 2-chloro-5-iodopyridine with the same amines produced compounds 153157 in somewhat higher yields (44–63%). Nevertheless, the presence of much-less reactive additional halogen atoms (Cl, Br) diminished the selectivity of the amination process, giving rise to unidentified by-products.
Though 3-bromopyridine was found to be unreactive in Cu(I)-catalyzed amination, 6-bromoquinoline was successfully aminated by a series of adamantane-containing amines, and the yields of the products 158163 varied from 46 to 74%. In all of the abovementioned processes, the CuI/L1 catalytic system was the most efficient.
It is well known that the adamantane derivatives bearing fluorine substituents are of substantial interest for a drug design. Indeed, compounds combining adamantane and fluoroaryl moieties display antituberculosis [103] and antiinflammatory [104] activities, as they inhibit tropoisomerase II [105] and purino receptor P2RX7 [106]. Adamantane derivatives with trifluoromethyl substituent inhibit 11β-HSD1 [93,107], melanin concentrating hormone MHC1 [108], and the ligands of cannabinoid receptor CB2 [109].
In this regard, we studied the heteroarylation of several adamantane-containing amines, 1, 35, 27, and 28, with fluoro- and triflurormethyl-substituted 2-bromopyridines [110,111]. The reactions were run with 2-bromo-3-fluoro-, 5-fluoropyridines, and 3-, 4-, 5-, and 6-trifluoromethylsubstituted 2-fluoropyridines using the optimized CuI/L1 catalytic system (Scheme 9). It was found that 2-bromo-3-fluoropyridine and especially 2-bromo-3-trifluoromethylpyrdine were not reactive enough under copper catalysis conditions due to steric hindrances at the bromine atom; thus, they were aminated using alternative Pd(0)-catalyzed reactions. The other 2-bromopyridines allowed for good to high yields of the corresponding products. The best results were obtained with the amine 3, as it produced N-pyridyl derivatives 165, 171, 177, and 183 in 79–89% yields. Among the CF3 derivatives, 2-bromo-6-trifluoromethylpyridine allowed for the formation of the amination products in the highest yields (up to 97% in the case of 186). Probably, the reactivity of the bromine atom in this compound was optimal: on one hand, enough for the amination reaction to proceed at a sufficient rate, on the other hand, allowing for avoidance of the formation of many by-products in the course of other catalytic processes.

4. C-N Bond Formation Using Unsupported Copper Nanoparticles

A special case is the use of copper nanoparticles (Cu NPs) in catalytic C–N bond formation. Often, copper and copper oxide nanoparticles immobilized on various supports are used, such as nanocatalysts immobilized on silica gel [112] and maghemite-silica magnetic support [113]; the following supports are also well-known: TiO2 [114], graphene [115], and multiwall carbon nanotubes (MWCNTs) [116]. The use of unsupported copper nanoparticles in amination reactions is quite rare, and mainly CuO NPs were described for this purpose [117,118,119,120,121].
We began an investigation of the possibilities of the unsupported copper and copper oxide nanoparticles to promote the arylation of various amines, including adamantane-containing compounds. At first, the optimal conditions for the Cu-catalyzed amination of PhI with a model n-octylamine were established (Scheme 10) using copper nanoparticles of various average sizes: 10/80 nm (mixture of two fractions), 25 nm, 72 nm, 86 nm, and CuO 65 nm [122,123]. The reactions were run in DMSO or DMF at 110 °C or 140 °C, and a variety of ligands were tested. For the arylation with iodobenzene, L2 was found to be the best choice, as it allowed for the synthesis of N-octylaniline 188 in 86–95% yields. The best result (82%) for the heteroarylation of n-octylamine with 2-iodopyridine was obtained using the Cu NPs 25 nm/L3 catalytic system, and other combinations of nanoparticles and ligands gave poorer yields. The possibilities of the catalyst recycling were revealed to be strongly dependent on the copper source, and the best result was obtained with the Cu NPs 25 nm/L1 system (up to eight cycles without loss of the catalyst’s activity).
Optimized conditions were applied to the arylation of a variety of adamantane-containing amines, 1, 35, 28, and 29, with iodobenzene (Scheme 11) [89].
The results were encouraging, as the yields of the target products (85–92%) were substantially higher than those obtained with CuI catalyst (65–76%, Scheme 5). The application of various substituted iodobenzenes for the arylation of amine 3 in the presence of another efficient system Cu NPs 10/80 nm/L2 (Scheme 12) was equally successful, and the yields of many products reached 90% and more (18, 20, 190192, 194, and 197). Some of these reactions gave better results in DMSO at 110 °C, and others successfully proceeded in DMF at 140 °C. It is clearly seen that the high yields could be obtained both with electron donor and electron withdrawing substituents in iodobenzenes. The experiments on catalyst recycling demonstrated the possibility of the use of Cu NPs 25 nm/L1 for the arylation of amine 3 with iodobenzene at least in seven cycles without a decrease in the product’s yield.
The N-heteroarylation of the adamantane-containing amines turned out to be more problematic (Scheme 13) [124]. The reactions of the amines 1, 35, and 28 with 2-iodopyridine gave 62–76% yields and it was noted that the effect of the ligand (L1 or L3) was dependent on the structure of the amine. In the reactions with the fluorine-containing 2-bromopyridines, the Cu NPs 25 nm/L3 catalytic system was found to be the best; however, the yields of the products ranged from humble, 24% (166), to high, 75–78% (183, 184). Generally, as in the reactions catalyzed by CuI/L1, the best results were obtained with 2-bromo-6-trifluoromethylpyridine.
While copper nanoparticles were more efficient than CuI in the N-arylation of the adamantane-containing amines with iodoarenes, they provided lower yields with bromoarenes under the same conditions, e.g., the attempts to introduce 6-bomoquinoline in the amination with several amines in the presence of Cu NPs/L1 resulted in 11–25% yields of the corresponding products 158, 160, and 161, while these reactions catalyzed by the CuI/L1 system provided 47–74% yields (Scheme 8). However, the application of Cu NPs/L2 for these reactions provided stable 70–75% yields of the products.

5. N,N′-Diarylation of the Di- and Polyamines Using Cu(I) Catalysis

Naturally occurring diamines such as putrescine 201, cadaverine 202, and some polyamines, are biologically active compounds which play a key role in cells’ proliferation and apoptosis [125], and more and more attention is attracted to their N-aryl derivatives [126,127,128], E.g., such derivatives of putrescine and cadaverine showed cytotoxic and antiproliferative activity [129], N-(p-tolyl) substituted cadaverine, and hexane-1,6-diamine demonstrated an affinity to NMDA receptors and antileishmaniasis activity [130,131].
We were mainly interested in the synthesis of N,N′-diarylsubstituted diamines and, for this purpose, conducted a series of reactions of the diamines 200203 with iodobenzene and some p-substituted derivatives (Scheme 14) [132]. The catalytic system CuI/L1 was shown to be optimal and the yields of the reaction products 204207 reached 70–80% in the best cases. It was noted that the electron-enriched p-iodoanisole was less reactive, giving 20–50% yields of the target compounds 204e207e; on the contrary, the reactions with electron-deficient 4-iodobenzotrifluoride provided enough high yields of the compounds 204d207d.
The investigation of the N,N′-diarylation of the tatraamine 208 with iodobenzene and a number of aryl iodides revealed that the best ligand in this case was L3 with MeCN as a solvent (Scheme 15) [133]. Bromobenzene was shown to be equally reactive with this tetraamine, which could be due to the coordination of copper to the tetraamine, resulting in a higher reactivity of the catalytic species. On the other hand, it caused the side reaction of the N-arylation of the secondary amino groups in the tetraamine, and moderate to good yields of the target compounds 209ag (42–67%) were obtained independently of the electron properties of the substituents in iodobenzenes.
Further, several iodobenzenes were introduced in the reactions with the naturally occurring tri- and tetra-amines 210213 (Scheme 16). The yields of the target N,N′-diaryl derivatives 214217 varied substantially depending on the nature of the polyamines and substituents in the aryl iodides. To note, in the case of 210 and 212, containing only trimethylenediamine moieties, the CuI/L3/EtCN catalytic system was more efficient, while in the reactions of 211 and 213 with tetramethylenediamine fragments, CuI/L1/DMF was found to act better. Again, p-iodoanisole provided the lowest yields of the target compounds, and the N,N′-diarylation of 210 and 213 generally proceeded more smoothly and resulted in higher yields of the products compared to the two other polyamines.
Oxadiamines, due to their hydrophilicity and chelating ability, are employed as flexible linkers for the synthesis of biologically active compounds. The structure and the length of the linker can modify the activity [134], and linkers can be incorporated in the organic molecules [135,136,137,138] or serve as a bidentate ligand for binding two Pd or Pt complexes [139].
We investigated the N,N′-diarylation of three different oxadiamines, 218220, varying their length, and number of oxygen atoms and methylene groups between N and O atoms (Scheme 17) [140]. CuI/L1 in DMF was shown to be the most active catalytic system, and trioxadiamine 218 was reacted with a large series of aryl iodides. In the majority of cases, the yields were good to high, attaining 91% in the case of 221m (with 4-CN substituent in the phenyl ring). A lower, 43%, yield was recorded for the product 221s, obtained from the electron-enriched p-iodoanisole, and, suddenly, the reactions with p-fluoroiodobenzene and 4-iodobenzotrifluorides gave very small yields of the target compounds 221h,k. Similar observations were true for the two other oxadiamines, 219 and 220, which also encountered difficulties upon reacting with these aryl iodides, especially with the latter one. On the contrary, much higher yields were observed in the reactions with o-fluoroiodonbenzene (77% yield of 221j and 58% yield of 222j).
Special investigations were also carried out to verify the absence of the side N,N-diarylation process, which is common for palladium-catalyzed amination reactions. For this purpose, diamines and oxadiamines were reacted with 10 equiv. of PhI and only the traces of the polyarylated derivatives were detected in the reaction mixtures in several cases. On the other hand, it helped to increase the yields of some N,N′-diphenyl derivatives to 80–90%.

6. N,N′-Diheteroarylation of Di- and Polyamines Using Cu(I) Catalysis

The next part of our research was dedicated to the N,N′-heteroarylation of the di- and poly-amines. The diamines 200203 reacted with 2- and 3-iodopyridines in the presence of the CuI/L1 catalytic system, and the yields of the target dipyridinyl derivatives were dependent on the nature of the diamines and were the highest for the shortest diamine (compounds 224 and 228) (Scheme 18) [141]. The use of 20 mol% catalyst was crucial for the N,N′-diheteroarylation; however, in many cases, monopyridinyl derivatives were also isolated.
More complicated were the reactions of tri- and tetraamine 210, 212, 232, and 233 with 2- and 3-iodopyridines due to many side reactions such as the heteroarylation of the secondary amino group and the N,N-diheteroarylation of the primary amino group (Scheme 19) [142].
In several cases, the formation of great amounts of monopyridinyl derivatives was observed. Almost in every case, the adjustment of the catalytic system was needed, including the ligands (L3 or L4), the catalyst loading, and a proper choice between 2-iodopyridine and less-active 2-bromopyridine. As a result, it was possible to obtain the N,N′-di(pyridine-2-yl) derivative of the triamine 234 and the N,N′-di(pyridine-3-yl) derivative of the tetraamine 241 in good yields (76 and 68%, respectively), while other products, 235240, were isolated in moderate yields: 30–52%. Spermidine 211 and spermine 213 reacted better with 2-bromopyridine and 3-iodopyridine, providing 50–64% yields of 242245 (Scheme 20), but they demanded the application of the another catalytic system, i.e., CuI/L1/DMF, due to their lower reactivity, probably due to the presence of the tetramethylenediamine moieties [141].
The N,N′-diheteroarylation of the trioxadiamine 218 with 2- and 3-iodopyridines as well as fluoro- and trifluoromethylsubstituted 2-bromopyridine was successful in the presence of CuI/L1 (Scheme 21), giving good to high yields of the desired compounds. Notable is the 79% yield of the compound 248, the product of the reaction with 2-bromo-3-fluropyridine; in this case, the substitution at the ortho-position was unexpectedly very efficient.
The possibility of introducing simple fluorophore groups such as naphthalene and quinoline in the polyamines was shown using several exemplary reactions (Scheme 22) [143]. Triamine 210 reacted with 2-iodonaphthalene to give only an 18% yield of the target diaryl derivative 253, while tetraamine 233 was able to produce the corresponding compound 254 in a 28% yield using less-active 2-bromonaphthalene. Note that the reaction of 210 with 2-bromonaphthalene afforded only monoaryl derivative 257 in 90% yields and the reaction of the tetraamine 233 with a more active 2-iodonaphthalene was totally non-selective. Similar results were obtained when reacting tri- and tetra-amines with 6-bromoquinoline (formation of monoquinolinyl derivative of the triamine 259 and N,N′-diquinolinyl derivative of the tetraamine 256). These facts demonstrate a strong dependence of the reactivity of polyamines on the number of ethylenediamine and trimethylenediamine fragments in their structure, as well as on the activity of (hetero)aryl halides.
Trioxadiamine 218 was less active and helped to produce its N,N′-dinaphthyl derivative 255 in a higher yield (46%) using 2-iodonaphthalene.
The reactions of the triamine 210 and several tetraamines 212, 232, and 233 with 3-iodothiophene were run in the presence of CuI/L3 in boiling EtCN, and the yields of the desired N,N′-diheteroaryl derivatives 260263 were also strongly dependent on the nature of the starting polyamines (Scheme 23) [144]. The highest yield (60%) was obtained with the tetraamine 212, possessing only trimethylenediamine moieties.

7. Copper-Catalyzed Amination in the Modifications of the Aza- and Diazacrown Ethers

Our investigations of the synthesis of the macrocycles definitely showed that copper catalysis is not applicable to this process, as all macrocyclization reactions require diluted enough conditions and Cu-catalyzed amination normally proceeds at 0.25–0.5 M concentrations of the reagents. However, it can be used for some modifications of the macrocycles possessing iodobenzyl substituents. It was shown to be possible to synthesize diamino derivatives of the azacrown ethers using this approach (Scheme 24). The reactions of N-(3-iodobenzyl)-substituted 1-aza-18-crown-6 264 were successfully aminated with the excess of the diamines 200 and 201, and the oxadiamines 218 and 219 to give the corresponding derivatives 265268 in 50–80% yields [145].
Note that the isomeric N-(4-iodobenzyl) derivative was much less reactive and provided 20% or less yields of the monoamination products. This method can be employed in the synthesis of bismacrocyclic compounds with the trioxadamine linkers 270 and 271; in this case the excess of azacrown derivatives should be used (Scheme 25). Also, the diamination of the N,N′-di(3-iodobenzyl) derivative of diazacrown 272 was carried out using an excess of propane-1,3-diamine 200 and trioxadiamine 218, with the yields of the corresponding products 273 and 274 being 36 and 76%, respectively.
The attempts to apply the analogous reactions to the iodobenzyl derivatives of porphyrins were not successful and the introduction of the diamine and oxadiamine moieties in these compounds could be achieved only using Pd-catalyzed amination reactions [146].

8. Chan-Lam Reactions for the Arylation of Adamantane-Containing Amines, Diamines and Oxadiamines

At the present time, the amines’ arylation with arylboronic acids and their derivatives known as the Chan–Lam reaction has been developed steadily [147], and we also explored this alternative to arylation with aryl halides. At first, we optimized the reaction conditions using a model p-tolylboronic acid in the reactions with various adamantane-containing amines (Scheme 26). Standard copper (II) acetate in MeCN in the presence of DBU (2 equiv.) and the use of p-tolylboronic acid excess were found to provide the highest yields. They ranged from 58 to 74% for the N-tolyl derivatives 17, 275278.
The possibility of introducing naphthalene and quinoline moieties in the adamantane-containing amines using this method was also shown (Scheme 27). The reactions with 2-naphthalenelboronic acid gave 34–66% yields of the corresponding derivatives 279282, while the reactions with 6-quinolinyl picolinyl borate produced the compounds 158, 160, 161, and 283 in 54–79% yields.
At the next step, the conditions for the N,N′-diarylation of the diamines 200203 and the oxadiamines 218220, 284 were elaborated (four equiv. of p-tolylboronic acid and three equiv. of DBU), and the corresponding target compounds 285288 and 289292 were synthesized (Scheme 28) [148]. Except for the propane-1,3-diamine 200 with the shortest chain, which obviously formed an inert complex with the Cu(II) cation, the other diamines and all of the oxadiamines provided good to high yields of the N,N′-darylated products (46–80%). Thus, the Chan–Lam reaction can be judged as a valuable prospective addition to the previously developed methods for the introduction of the aryl and heteroaryl substituents in the adamantane-containing amines, diamines, and oxadiamines.

9. Conclusions and Prospects

The results presented in this review clearly demonstrate wide possibilities of copper-catalyzed amination reactions for the synthesis of N-(hetero)aryl derivatives of the valuable adamantane-containing amines, diamines, oxadiamines, and polyamines. The reactions catalyzed by the most-commonly used CuI were shown to proceed in DMF at 140 °C or in DMSO at 110 °C in the presence of O,O-ligands (2-isobutirylcyclohexanone and rac-BINOL), in the case of the adamantane-containing amines, diamines, and oxadiamines or in the presence of N,O ligands (proline, N,N-dimethylglycine) in the case of tri- and tetra-amines. Under these conditions, iodoarenes and iodopyridines were found to be substantially more reactive than their bromosubstituted analogues, although with certain exclusions. Thus, bromobenzene and 2-bromopyridine were suitable for reactions with some polyamines, and more reactive F- and CF3-substituted 2-bromopyridines were successfully employed in reactions with the adamantane-containing amines. The majority of the adamantane-containing amines were found to normally participate in the copper-catalyzed N-(hetero)arylation reactions; nevertheless, several bulky amines with enough hindered amino groups were reluctant in this process and could provide target products in normal yields only in the Pd-catalyzed amination reactions. Copper-catalyzed amination has an important advantage over its Pd-catalyzed alternative, as it does not afford N,N-diarylation products even in the presence of a great excess of the aryl halide. On the other hand, some drawbacks should be mentioned such as insufficient yields of the N,N′-diaryl compounds in certain cases due to the formation of substantial amounts of mono-derivatives. Another drawback is a possibility of the side reaction of the secondary dialkylamino group arylation in linear polyamines. At last, the reactions need concentrated solutions making macrocyclization reactions impossible.
The possibility of using commercially available Cu and CuO nanoparticles for reactions with the adamantane-containing amines has been investigated recently, and the possibility of obtaining high yields of the N-aryl derivatives was demonstrated. An important advantage of this approach is the possibility of reuse of the catalyst for up to seven cycles without a loss of reactivity. Further development of this method is required to engage di- and poly-amines in (hetero)arylation using copper nanoparticles. Equally, it is important to continue investigations revealing the action of nano catalysts, metal leaching during the reaction, and catalyst reusability to study the dependence of the catalyst activity on the ligands and solvents employed. This first successful syntheses of the arylated adamantane-containing amines, diamines, and oxadiamines under the conditions of the Chan–Lam reaction indicate a need to further search for its scope and limitations with our compounds and the application of the copper nanocatalysts in this process.

Author Contributions

Conceptualization, A.D.A. and I.P.B.; investigation, S.P.P., A.V.M., V.I.F., D.S.K., A.S.M. and A.A.Y.; writing—original draft preparation, A.D.A. and A.S.A.; writing—review and editing, A.S.A., A.D.A. and I.P.B.; funding acquisition, A.D.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (grant no. 22-23-00518).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ullmann, F.; Bielecki, J. Ueber Synthesen in der Biphenylreihe. Ber. Dtsch. Chem. Ges. 1901, 34, 2174–2185. [Google Scholar] [CrossRef]
  2. Ullmann, F. Ueber eine neue Bildungsweise von Diphenylaminderivaten. Ber. Dtsch. Chem. Ges. 1903, 36, 2382–2384. [Google Scholar] [CrossRef]
  3. Ullmann, F.; Sponagel, P. Ueber Phenylirung von Phenolen. Justus Liebigs Ann. Chem. 1906, 350, 83–107. [Google Scholar] [CrossRef]
  4. Goldberg, I. Ueber Phenylirungen bei Gegenwart von Kupfer als Katalysator. Ber. Dtsch. Chem. Ges. 1906, 39, 1691–1692. [Google Scholar] [CrossRef]
  5. Lindley, J. Tetrahedron report number 163: Copper assisted nucleophilic substitution of aryl halogen. Tetrahedron 1984, 40, 1433–1456. [Google Scholar] [CrossRef]
  6. Kunz, K.; Scholz, U.; Ganzer, D. Renaissance of Ullmann and Goldberg Reactions—Progress in Copper Catalyzed C-N-, C-O- and C-S-Coupling. Synlett 2003, 2003, 2428–2439. [Google Scholar] [CrossRef]
  7. Fanta, P.E. The Ullmann Synthesis of Biaryls. Synthesis 1974, 1974, 9–21. [Google Scholar] [CrossRef]
  8. Tuong, T.D.; Hida, M. Mechanism of the Ullmann Condensation. I. Kinetic and Thermodynamic Studies. Bull. Chem. Soc. Jpn. 1970, 43, 1763–1768. [Google Scholar] [CrossRef]
  9. Affouard, C.; Crockett, R.D.; Diker, K.; Farrell, R.P.; Gorins, G.; Huckins, J.R.; Caille, S. Multi-Kilo Delivery of AMG 925 Featuring a Buchwald–Hartwig Amination and Processing with Insoluble Synthetic Intermediates. Org. Proc. Res. Dev. 2015, 19, 476–485. [Google Scholar] [CrossRef]
  10. Ku, Y.-Y.; Chan, V.S.; Christesen, A.; Grieme, T.; Mulhern, M.; Pu, Y.-M.; Wendt, M.D. Development of a Convergent Large-Scale Synthesis for Venetoclax, a First-in-Class BCL-2 Selective Inhibitor. J. Org. Chem. 2019, 84, 4814–4829. [Google Scholar] [CrossRef]
  11. Yang, Q.; Zhao, Y.; Ma, D. Cu-Mediated Ullmann-Type Cross-Coupling and Industrial Applications in Route Design, Process Development, and Scale-up of Pharmaceutical and Agrochemical Processes. Org. Proc. Res. Dev. 2022, 26, 1690–1750. [Google Scholar] [CrossRef]
  12. Evendar, P.; Qu, R.-Y.; Kang, W.-M.; He, B.; Yang, G.-F. Palladium-Catalyzed Cross-Coupling Reactions: A Powerful Tool for the Synthesis of Agrochemicals. J. Agric. Food. Chem. 2018, 66, 8914–8934. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, J.; Yan, W.; Townsend, E.J.; Feng, J.; Pan, L.; Del Angel Hernandez, V.; Faul, C.F.J. Tunable Surface Area, Porosity, and Function in Conjugated Microporous Polymers. Angew. Chem. Int. Ed. 2019, 58, 11715–11719. [Google Scholar] [CrossRef]
  14. Astridge, D.D.; Hoffman, J.B.; Zhang, F.; Park, S.Y.; Zhu, K.; Sellinger, A. Polymer Hole Transport Materials for Perovskite Solar Cells via Buchwald–Hartwig Amination. ACS Appl. Polym. Mater. 2021, 3, 5578–5587. [Google Scholar] [CrossRef]
  15. Weingarten, H. Mechanism of the Ullmann Condensation. J. Org. Chem. 1964, 29, 3624–3626. [Google Scholar] [CrossRef]
  16. Nicolaou, K.C.; Boddy, C.N.C.; Natarajan, S.; Yue, T.Y.; Li, H.; Bräse, S.; Ramanjulu, J.M. New Synthetic Technology for the Synthesis of Aryl Ethers: Construction of C-O-D and D-O-E Ring Model Systems of Vancomycin. J. Am. Chem. Soc. 1997, 119, 3421–3422. [Google Scholar] [CrossRef]
  17. Marcoux, J.-F.; Doye, S.; Buchwald, S.L. A General Copper-Catalyzed Synthesis of Diaryl Ethers. J. Am. Chem. Soc. 1997, 119, 10539–10540. [Google Scholar] [CrossRef]
  18. Ma, D.; Zhang, Y.; Yao, J.; Wu, S.; Tao, F. Accelerating Effect Induced by the Structure of α-Amino Acid in the Copper-Catalyzed Coupling Reaction of Aryl Halides with α-Amino Acids. Synthesis of Benzolactam-V8. J. Am. Chem. Soc. 1998, 120, 12459–12467. [Google Scholar] [CrossRef]
  19. Monnier, F.; Taillefer, M. Catalytic C–C, C–N, and C–O Ullmann-Type Coupling Reactions. Angew. Chem. Int. Ed. 2009, 48, 6954–6971. [Google Scholar] [CrossRef]
  20. Beletskaya, I.P.; Cheprakov, A.V. Copper in cross-coupling reactions: The post-Ullmann chemistry. Coord. Chem. Rev. 2004, 248, 2337–2364. [Google Scholar] [CrossRef]
  21. Beletskaya, I.P.; Averin, A.D. Metal-catalyzed reactions for the C(sp2)–N bond formation: Achievements of recent years. Russ. Chem. Rev. 2021, 90, 1359. [Google Scholar] [CrossRef]
  22. Goodbrand, H.B.; Hu, N.-X. Ligand-Accelerated Catalysis of the Ullmann Condensation: Application to Hole Conducting Triarylamines. J. Org. Chem. 1999, 64, 670–674. [Google Scholar] [CrossRef]
  23. Gujadhur, R.K.; Bates, C.G.; Venkataraman, D. Formation of Aryl–Nitrogen, Aryl–Oxygen, and Aryl–Carbon Bonds Using Well-Defined Copper(I)-Based Catalysts. Org. Lett. 2001, 3, 4315–4317. [Google Scholar] [CrossRef] [PubMed]
  24. Surry, D.S.; Buchwald, S.L. Diamine ligands in copper-catalyzed reactions. Chem. Sci. 2010, 1, 13–31. [Google Scholar] [CrossRef] [PubMed]
  25. Klapars, A.; Huang, X.; Buchwald, S.L. A General and Efficient Copper Catalyst for the Amidation of Aryl Halides. J. Am. Chem. Soc. 2002, 124, 7421–7428. [Google Scholar] [CrossRef] [PubMed]
  26. Antilla, J.C.; Baskin, J.M.; Barder, T.E.; Buchwald, S.L. Copper–Diamine-Catalyzed N-Arylation of Pyrroles, Pyrazoles, Indazoles, Imidazoles, and Triazoles. J. Org. Chem. 2004, 69, 5578–5587. [Google Scholar] [CrossRef]
  27. Ma, D.; Cai, Q.; Zhang, H. Mild Method for Ullmann Coupling Reaction of Amines and Aryl Halides. Org. Lett. 2003, 5, 2453–2455. [Google Scholar] [CrossRef]
  28. Zhang, H.; Cai, Q.; Ma, D. Amino Acid Promoted CuI-Catalyzed C–N Bond Formation between Aryl Halides and Amines or N-Containing Heterocycles. J. Org. Chem. 2005, 70, 5164–5173. [Google Scholar] [CrossRef]
  29. Jiang, Q.; Jiang, D.; Jiang, Y.; Fu, H.; Zhao, Y. A Mild and Efficient Method for Copper-Catalyzed Ullmann-Type N-Arylation of Aliphatic Amines and Amino Acids. Synlett 2007, 2007, 1836–1842. [Google Scholar] [CrossRef]
  30. Fan, M.; Zhou, W.; Jiang, Y.; Ma, D. Assembly of Primary (Hetero)Arylamines via CuI/Oxalic Diamide-Catalyzed Coupling of Aryl Chlorides and Ammonia. Org. Lett. 2015, 17, 5934–5937. [Google Scholar] [CrossRef]
  31. Gao, J.; Bhunia, S.; Wang, K.; Gan, L.; Xia, S.; Ma, D. Discovery of N-(Naphthalen-1-yl)-N′-alkyl Oxalamide Ligands Enables Cu-Catalyzed Aryl Amination with High Turnovers. Org. Lett. 2017, 19, 2809–2812. [Google Scholar] [CrossRef] [PubMed]
  32. Kwong, F.Y.; Klapars, A.; Buchwald, S.L. Copper-Catalyzed Coupling of Alkylamines and Aryl Iodides: An Efficient System Even in an Air Atmosphere. Org. Lett. 2002, 4, 581–584. [Google Scholar] [CrossRef] [PubMed]
  33. Jiang, D.; Fu, H.; Jiang, Y.; Zhao, Y. CuBr/rac-BINOL-Catalyzed N-Arylations of Aliphatic Amines at Room Temperature. J. Org. Chem. 2007, 72, 672–674. [Google Scholar] [CrossRef] [PubMed]
  34. Shafir, A.; Buchwald, S.L. Highly Selective Room-Temperature Copper-Catalyzed C–N Coupling Reactions. J. Am. Chem. Soc. 2006, 128, 8742–8743. [Google Scholar] [CrossRef]
  35. Shafir, A.; Lichtor, P.A.; Buchwald, S.L. N- versus O-Arylation of Aminoalcohols: Orthogonal Selectivity in Copper-Based Catalysts. J. Am. Chem. Soc. 2007, 129, 3490–3491. [Google Scholar] [CrossRef]
  36. Kwong, F.Y.; Buchwald, S.L. Mild and Efficient Copper-Catalyzed Amination of Aryl Bromides with Primary Alkylamines. Org. Lett. 2003, 5, 793–796. [Google Scholar] [CrossRef]
  37. Bernhardson, D.J.; Widlicka, D.W.; Singer, R.A. Cu-Catalyzed Couplings of Heteroaryl Primary Amines and (Hetero)aryl Bromides with 6-Hydroxypicolinamide Ligands. Org. Proc. Res. Dev. 2019, 23, 1538–1551. [Google Scholar] [CrossRef]
  38. Modak, A.; Nett, A.J.; Swift, E.C.; Haibach, M.C.; Chan, V.S.; Franczyk, T.S.; Shekhar, S.; Cook, S.P. Cu-Catalyzed C–N Coupling with Sterically Hindered Partners. ACS Catal. 2020, 10, 10495–10499. [Google Scholar] [CrossRef]
  39. Kim, S.-T.; Strauss, M.J.; Cabré, A.; Buchwald, S.L. Room-Temperature Cu-Catalyzed Amination of Aryl Bromides Enabled by DFT-Guided Ligand Design. J. Am. Chem. Soc. 2023, 145, 6966–6975. [Google Scholar] [CrossRef]
  40. Zhou, W.; Fan, M.; Yin, J.; Jiang, Y.; Ma, D. CuI/Oxalic Diamide Catalyzed Coupling Reaction of (Hetero)Aryl Chlorides and Amines. J. Am. Chem. Soc. 2015, 137, 11942–11945. [Google Scholar] [CrossRef]
  41. Chen, Z.; Ma, D. Cu/N,N′-Dibenzyloxalamide-Catalyzed N-Arylation of Heteroanilines. Org. Lett. 2019, 21, 6874–6878. [Google Scholar] [CrossRef] [PubMed]
  42. Li, S.; Huang, X.; Gao, Y.; Jin, J. Oxalamide/Amide Ligands: Enhanced and Copper-Catalyzed C–N Cross-Coupling for Triarylamine Synthesis. Org. Lett. 2022, 24, 5817–5824. [Google Scholar] [CrossRef] [PubMed]
  43. Ma, D.; Cai, Q. Copper/Amino Acid Catalyzed Cross-Couplings of Aryl and Vinyl Halides with Nucleophiles. Acc. Chem. Res. 2008, 41, 1450–1460. [Google Scholar] [CrossRef] [PubMed]
  44. Beletskaya, I.P.; Cheprakov, A.V. The Complementary Competitors: Palladium and Copper in C–N Cross-Coupling Reactions. Organometallics 2012, 31, 7753–7808. [Google Scholar] [CrossRef]
  45. Senra, J.D.; Aguiar, L.C.S.; Simas, A.B.C. Recent Progress in Transition-Metal-Catalyzed C-N Cross-Couplings: Emerging Approaches Towards Sustainability. Curr. Org. Synth. 2011, 8, 53–78. [Google Scholar] [CrossRef]
  46. Sambiagio, C.; Marsden, S.P.; Blacker, A.J.; McGowan, P.C. Copper catalysed Ullmann type chemistry: From mechanistic aspects to modern development. Chem. Soc. Rev. 2014, 43, 3525–3550. [Google Scholar] [CrossRef] [PubMed]
  47. Evano, G.; Blanchard, N.; Toumi, M. Copper-Mediated Coupling Reactions and Their Applications in Natural Products and Designed Biomolecules Synthesis. Chem. Rev. 2008, 108, 3054–3131. [Google Scholar] [CrossRef]
  48. Okano, K.; Tokuyama, H.; Fukuyama, T. Copper-mediated aromatic amination reaction and its application to the total synthesis of natural products. Chem. Commun. 2014, 50, 13650–13663. [Google Scholar] [CrossRef]
  49. Lee, J.; Panek, J.S. Application of Copper-Mediated C–N Bond Formation in Complex Molecules Synthesis. In Copper-Mediated Cross-Coupling Reactions; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2013; pp. 589–641. [Google Scholar] [CrossRef]
  50. Junge, K.; Wienhöfer, G.; Beller, M.; Tlili, A.; Evano, G.; Taillefer, M.; Kempe, R.; Malbertz, C.; Klankermayer, J. New Trends in Organometallic Catalysts. In Applied Homogeneous Catalysis with Organometallic Compounds: A Comprehensive Handbook in Three Volumes, 3rd ed.; Cornils, B., Herrmann, W.A., Beller, M., Paciello, R., Eds.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2017. [Google Scholar]
  51. Shaughnessy, K.H.; Ciganek, E.; De Vasher, R.B.; Denmark, S.E. Copper-Catalyzed Amination of Aryl and Alkenyl Electrophiles; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2017. [Google Scholar]
  52. Neetha, M.; Saranya, S.; Ann Harry, N.; Anilkumar, G. Recent Advances and Perspectives in the Copper-Catalysed Amination of Aryl and Heteroaryl Halides. ChemistrySelect 2020, 5, 736–753. [Google Scholar] [CrossRef]
  53. Weidlich, T.; Špryncová, M.; Čegan, A. Copper-Catalyzed Reactions of Aryl Halides with N-Nucleophiles and Their Possible Application for Degradation of Halogenated Aromatic Contaminants. Catalysts 2022, 12, 911. [Google Scholar] [CrossRef]
  54. Hosseinzadeh, R.; Aghili, N.; Tajbakhsh, M. SBA-15 Immobilized Phenanthroline–Copper(I) Complex as a Recyclable Efficient Catalyst for N-Arylation of Amides and N–H Heterocycles with Aryl Halides. Catal. Lett. 2016, 146, 193–203. [Google Scholar] [CrossRef]
  55. Niakan, M.; Asadi, Z.; Zare, S. Preparation, Characterization and Application of Copper Schiff base Complex Supported on MCM-41 as a Recyclable Catalyst for the Ullmann-type N-arylation Reaction. ChemistrySelect 2020, 5, 40–48. [Google Scholar] [CrossRef]
  56. Veisi, H.; Hamelian, M.; Hemmati, S.; Dalvand, A. CuI catalyst heterogenized on melamine-pyridines immobilized SBA-15: Heterogeneous and recyclable nanocatalyst for Ullmann-type CN coupling reactions. Tetrahedron Lett. 2017, 58, 4440–4446. [Google Scholar] [CrossRef]
  57. Hemmati, S.; Ahany Kamangar, S.; Yousefi, M.; Hashemi Salehi, M.; Hekmati, M. Cu(I)-anchored polyvinyl alcohol coated-magnetic nanoparticles as heterogeneous nanocatalyst in Ullmann-type C–N coupling reactions. Appl. Organomet. Chem. 2020, 34, e5611. [Google Scholar] [CrossRef]
  58. Islam, M.; Mondal, S.; Mondal, P.; Roy, A.S.; Tuhina, K.; Mobarok, M.; Paul, S.; Salam, N.; Hossain, D. An Efficient Recyclable Polymer Supported Copper(II) Catalyst for C–N Bond Formation by N-Arylation. Catal. Lett. 2011, 141, 1171–1181. [Google Scholar] [CrossRef]
  59. Arundhathi, R.; Kumar, D.C.; Sreedhar, B. C–N Bond Formation Catalysed by CuI Bonded to Polyaniline Nanofiber. Eur. J. Org. Chem. 2010, 2010, 3621–3630. [Google Scholar] [CrossRef]
  60. Islam, S.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]
  61. Esmaeilpour, M.; Sardarian, A.R.; Firouzabadi, H. Dendrimer-encapsulated Cu(Π) nanoparticles immobilized on superparamagnetic Fe3O4@SiO2 nanoparticles as a novel recyclable catalyst for N-arylation of nitrogen heterocycles and green synthesis of 5-substituted 1H-tetrazoles. Appl. Organomet. Chem. 2018, 32, e4300. [Google Scholar] [CrossRef]
  62. Chouhan, G.; Wang, D.; Alper, H. Magnetic nanoparticle-supported proline as a recyclable and recoverable ligand for the CuI catalyzed arylation of nitrogen nucleophiles. Chem. Commun. 2007, 45, 4809–4811. [Google Scholar] [CrossRef]
  63. Zahmatkesh, S.; Esmaeilpour, M.; Javidi, J. 1,4-Dihydroxyanthraquinone–copper(ii) supported on superparamagnetic Fe3O4@SiO2: An efficient catalyst for N-arylation of nitrogen heterocycles and alkylamines with aryl halides and click synthesis of 1-aryl-1,2,3-triazole derivatives. RSC Adv. 2016, 6, 90154–90164. [Google Scholar] [CrossRef]
  64. Sardarian, A.R.; Eslahi, H.; Esmaeilpour, M. Copper(II) Complex Supported on Fe3O4@SiO2 Coated by Polyvinyl Alcohol as Reusable Nanocatalyst in N-Arylation of Amines and N(H)- Heterocycles and Green Synthesis of 1H-Tetrazoles. ChemistrySelect 2018, 3, 1499–1511. [Google Scholar] [CrossRef]
  65. Paine, A.J. Mechanisms and models for copper mediated nucleophilic aromatic substitution. 2. Single catalytic species from three different oxidation states of copper in an Ullmann synthesis of triarylamines. J. Am. Chem. Soc. 1987, 109, 1496–1502. [Google Scholar] [CrossRef]
  66. Aalten, H.L.; van Koten, G.; Grove, D.M.; Kuilman, T.; Piekstra, O.G.; Hulshof, L.A.; Sheldon, R.A. The copper catalysed reaction of sodium methoxide with aryl bromides. A mechanistic study leading to a facile synthesis of anisole derivatives. Tetrahedron 1989, 45, 5565–5578. [Google Scholar] [CrossRef]
  67. Komori, T.; Satoh, N.; Yokoshima, S.; Fukuyama, T. Copper-Mediated Aryl Amination: In Situ Generation of an Active Copper(I) Species. Synlett 2011, 2011, 1859–1862. [Google Scholar] [CrossRef]
  68. Meng, F.; Zhu, X.; Li, Y.; Xie, J.; Wang, B.; Yao, J.; Wan, Y. Efficient Copper-Catalyzed Direct Amination of Aryl Halides Using Aqueous Ammonia in Water. Eur. J. Org. Chem. 2010, 2010, 6149–6152. [Google Scholar] [CrossRef]
  69. Quan, Z.; Xia, H.; Zhang, Z.; Da, Y.; Wang, X. Copper-Catalyzed Amination of Aryl Halides with Aqueous Ammonia under Mild Conditions. Chin. J. Chem. 2013, 31, 501–506. [Google Scholar] [CrossRef]
  70. Fantasia, S.; Windisch, J.; Scalone, M. Ligandless Copper-Catalyzed Coupling of Heteroaryl Bromides with Gaseous Ammonia. Adv. Synth. Catal. 2013, 355, 627–631. [Google Scholar] [CrossRef]
  71. Shang, Z.; Yang, L.; Chang, G. Synthesis of high-performance polymers via copper-catalyzed amination of dibromoarenes with primary aromatic ether diamines. Macromol. Res. 2015, 23, 937–943. [Google Scholar] [CrossRef]
  72. Jiao, J.; Zhang, X.-R.; Chang, N.-H.; Wang, J.; Wei, J.-F.; Shi, X.-Y.; Chen, Z.-G. A Facile and Practical Copper Powder-Catalyzed, Organic Solvent- and Ligand-Free Ullmann Amination of Aryl Halides. J. Org. Chem. 2011, 76, 1180–1183. [Google Scholar] [CrossRef]
  73. Sperotto, E.; van Klink, G.P.M.; van Koten, G.; de Vries, J.G. The mechanism of the modified Ullmann reaction. Dalton Trans. 2010, 39, 10338–10351. [Google Scholar] [CrossRef]
  74. Mansour, M.; Giacovazzi, R.; Ouali, A.; Taillefer, M.; Jutand, A. Activation of aryl halides by Cu0/1,10-phenanthroline: Cu0 as precursor of CuI catalyst in cross-coupling reactions. Chem. Commun. 2008, 45, 6051–6053. [Google Scholar] [CrossRef] [PubMed]
  75. Guo, Z.; Guo, J.; Song, Y.; Wang, L.; Zou, G. Hemilabile-coordinated copper promoted amination of aryl halides with ammonia in aqueous ethylene glycol under atmosphere pressure. Appl. Organomet. Chem. 2009, 23, 150–153. [Google Scholar] [CrossRef]
  76. Lefèvre, G.; Franc, G.; Adamo, C.; Jutand, A.; Ciofini, I. Influence of the Formation of the Halogen Bond ArX---N on the Mechanism of Diketonate Ligated Copper-Catalyzed Amination of Aromatic Halides. Organometallics 2012, 31, 914–920. [Google Scholar] [CrossRef]
  77. Xiang, S.-K.; Zhang, D.-X.; Hu, H.; Shi, J.-L.; Liao, L.-G.; Feng, C.; Wang, B.-Q.; Zhao, K.-Q.; Hu, P.; Yang, H.; et al. Synthesis of N-Arylamides by Copper-Catalyzed Amination of Aryl Halides with Nitriles. Adv. Synth. Catal. 2013, 355, 1495–1499. [Google Scholar] [CrossRef]
  78. Huffman, L.M.; Stahl, S.S. Carbon–Nitrogen Bond Formation Involving Well-Defined Aryl–Copper(III) Complexes. J. Am. Chem. Soc. 2008, 130, 9196–9197. [Google Scholar] [CrossRef] [PubMed]
  79. Casitas, A.; King, A.E.; Parella, T.; Costas, M.; Stahl, S.S.; Ribas, X. Direct observation of CuI/CuIII redox steps relevant to Ullmann-type coupling reactions. Chem. Sci. 2010, 1, 326–330. [Google Scholar] [CrossRef]
  80. Casitas, A.; Ribas, X. The role of organometallic copper(III) complexes in homogeneous catalysis. Chem. Sci. 2013, 4, 2301–2318. [Google Scholar] [CrossRef]
  81. Bacon, R.G.R.; Karim, A. Metal ions and complexes in organic reactions. Part XV. Copper-catalysed substitutions of aryl halides by phthalimide ion. J. Chem. Soc. Perkin Trans. 1973, 272–278. [Google Scholar] [CrossRef]
  82. Wanka, L.; Iqbal, K.; Schreiner, P.R. The Lipophilic Bullet Hits the Targets: Medicinal Chemistry of Adamantane Derivatives. Chem. Rev. 2013, 113, 3516–3604. [Google Scholar] [CrossRef]
  83. Sonkusare, S.K.; Kaul, C.L.; Ramarao, P. Dementia of Alzheimer’s disease and other neurodegenerative disorders—Memantine, a new hope. Pharmacol. Res. 2005, 51, 1–17. [Google Scholar] [CrossRef]
  84. Grekhova, T.V.; Gainetdinov, R.R.; Sotnikova, T.D.; Krasnykh, L.M.; Kudrin, V.S.; Sergeeva, S.A.; Morozov, I.S. Effect of bromantane, a new immunostimulating agent with psychostimulating activity, on the release and metabolism of dopamine in the striatum of freely moving rats. A microdialysis study. Bull. Exp. Biol. Med. 1995, 119, 294–296. [Google Scholar] [CrossRef]
  85. Morozov, I.S.; Klimova, N.V.; Lavrova, L.N.; Avdyunina, N.I.; Pyatin, B.M.; Troitskaya, V.S.; Bykov, N.P. N-adamantyl derivatives of aromatic amines. Part I. Synthesis and neurotropic activity of N-(adamant-2-yl)anilines. Pharm. Chem. J. 1998, 32, 1–4. [Google Scholar] [CrossRef]
  86. Panchenko, S.P.; Abel, A.S.; Averin, A.D.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Arylation of adamantanamines: VIII. Optimization of the catalytic system for copper-catalyzed arylation of adamantane-containing amines. Russ. J. Org. Chem. 2017, 53, 1497–1504. [Google Scholar] [CrossRef]
  87. Averin, A.D.; Panchenko, S.P.; Abel, A.S.; Maloshitskaya, O.A.; Butov, G.M.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Arylation of adamantanamines: IX. Copper(I)-catalyzed arylation of adamantane-containing amines. Russ. J. Org. Chem. 2017, 53, 1788–1798. [Google Scholar] [CrossRef]
  88. Panchenko, S.P.; Abel, A.C.; Averin, A.D.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. CuI-catalyzed N,N′-diarylation of diamines of adamantane series. Russ. Chem. Bull. 2016, 65, 1550–1555. [Google Scholar] [CrossRef]
  89. Murashkina, A.V.; Averin, A.D.; Panchenko, S.P.; Abel, A.S.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Correia, C.R.D.; Beletskaya, I.P. Comparison of the Catalytic Activities of Copper(I) Iodide and Copper Nanoparticles in the N-Arylation of Adamantane-Containing Amines. Russ. J. Org. Chem. 2022, 58, 15–24. [Google Scholar] [CrossRef]
  90. Kivitz, A.J.; Gutierrez-Ureña, S.R.; Poiley, J.; Genovese, M.C.; Kristy, R.; Shay, K.; Wang, X.; Garg, J.P.; Zubrzycka-Sienkiewicz, A. Peficitinib, a JAK Inhibitor, in the Treatment of Moderate-to-Severe Rheumatoid Arthritis in Patients With an Inadequate Response to Methotrexate. Arthritis Rheumatol. 2017, 69, 709–719. [Google Scholar] [CrossRef]
  91. Leovac, V.M.; Rodić, M.V.; Jovanović, L.S.; Joksović, M.D.; Stanojković, T.; Vujčić, M.; Sladić, D.; Marković, V.; Vojinović-Ješić, L.S. Transition Metal Complexes with 1-Adamantoyl Hydrazones—Cytotoxic Copper(II) Complexes of Tri- and Tetradentate Pyridine Chelators Containing an Adamantane Ring System. Eur. J. Inorg. Chem. 2015, 2015, 882–895. [Google Scholar] [CrossRef]
  92. Ryu, J.H.; Kim, S.; Han, H.Y.; Son, H.J.; Lee, H.J.; Shin, Y.A.; Kim, J.-S.; Park, H.-g. Synthesis and biological evaluation of picolinamides as potent inhibitors of 11β-hydroxysteroid dehydrogenase type 1 (11β-HSD1). Bioorg. Med. Chem. Lett. 2015, 25, 695–700. [Google Scholar] [CrossRef]
  93. Rohde, J.J.; Pliushchev, M.A.; Sorensen, B.K.; Wodka, D.; Shuai, Q.; Wang, J.; Fung, S.; Monzon, K.M.; Chiou, W.J.; Pan, L.; et al. Discovery and Metabolic Stabilization of Potent and Selective 2-Amino-N-(adamant-2-yl) Acetamide 11β-Hydroxysteroid Dehydrogenase Type 1 Inhibitors. J. Med. Chem. 2007, 50, 149–164. [Google Scholar] [CrossRef]
  94. Sorensen, B.; Rohde, J.; Wang, J.; Fung, S.; Monzon, K.; Chiou, W.; Pan, L.; Deng, X.; Stolarik, D.; Frevert, E.U.; et al. Adamantane 11-β-HSD-1 inhibitors: Application of an isocyanide multicomponent reaction. Bioorg. Med. Chem. Lett. 2006, 16, 5958–5962. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, S.; Qian, P.; Wan, F.-X.; Shi, Y.-H.; Jiang, L. Design, synthesis, and biological activity of novel 2-(pyridin-3-yl)ethan-1-one oxime ethers bearing adamantane moiety. J. Chin. Chem. Soc. 2019, 66, 330–334. [Google Scholar] [CrossRef]
  96. Collins, K.C.; Janda, K.D. Investigating Hapten Clustering as a Strategy to Enhance Vaccines against Drugs of Abuse. Bioconjug. Chem. 2014, 25, 593–600. [Google Scholar] [CrossRef] [PubMed]
  97. Weigel, W.K., III; Dang, H.T.; Feceub, A.; Martin, D.B.C. Direct radical functionalization methods to access substituted adamantanes and diamondoids. Org. Biomol. Chem. 2022, 20, 10–36. [Google Scholar] [CrossRef]
  98. Zhou, L.; Togo, H. Introduction of Heteroaromatic Bases onto Cycloalkanes with BPO. Eur. J. Org. Chem. 2019, 2019, 1627–1634. [Google Scholar] [CrossRef]
  99. Zhao, H.; Li, Z.; Jin, J. Green oxidant H2O2 as a hydrogen atom transfer reagent for visible light-mediated Minisci reaction. New J. Chem. 2019, 43, 12533–12537. [Google Scholar] [CrossRef]
  100. Perry, I.B.; Brewer, T.F.; Sarver, P.J.; Schultz, D.M.; DiRocco, D.A.; MacMillan, D.W.C. Direct arylation of strong aliphatic C–H bonds. Nature 2018, 560, 70–75. [Google Scholar] [CrossRef]
  101. Abel, A.S.; Averin, A.D.; Anokhin, M.V.; Maloshitskaya, O.A.; Butov, G.M.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Arylation of adamantanamines: VII. Copper(I)-catalyzed N-heteroarylation of adamantane-containing amines with halopyridines. Russ. J. Org. Chem. 2015, 51, 301–308. [Google Scholar] [CrossRef]
  102. Abel, A.S.; Kotovshchikov, Y.N.; Averin, A.D.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Problem of Regioselectivity in the Amination of 2-Fluoro-5-iodopyridine with Adamantylalkyl Amines. Heterocycles 2019, 99, 1342–1354. [Google Scholar] [CrossRef]
  103. Scherman, M.S.; North, E.J.; Jones, V.; Hess, T.N.; Grzegorzewicz, A.E.; Kasagami, T.; Kim, I.-H.; Merzlikin, O.; Lenaerts, A.J.; Lee, R.E.; et al. Screening a library of 1600 adamantyl ureas for anti-Mycobacterium tuberculosis activity in vitro and for better physical chemical properties for bioavailability. Bioorg. Med. Chem. 2012, 20, 3255–3262. [Google Scholar] [CrossRef]
  104. Al-Omar, M.A.; Al-Abdullah, E.S.; Shehata, I.A.; Habib, E.E.; Ibrahim, T.M.; El-Emam, A.A. Synthesis, Antimicrobial, and Anti-inflammatory Activities of Novel 5-(1-Adamantyl)-4-arylideneamino-3-mercapto-1,2,4-triazoles and Related Derivatives. Molecules 2010, 15, 2526–2550. [Google Scholar] [CrossRef] [PubMed]
  105. Yu, X.; Zhang, M.; Annamalai, T.; Bansod, P.; Narula, G.; Tse-Dinh, Y.-C.; Sun, D. Synthesis, evaluation, and CoMFA study of fluoroquinophenoxazine derivatives as bacterial topoisomerase IA inhibitors. Eur. J. Med. Chem. 2017, 125, 515–527. [Google Scholar] [CrossRef] [PubMed]
  106. O’Brien-Brown, J.; Jackson, A.; Reekie, T.A.; Barron, M.L.; Werry, E.L.; Schiavini, P.; McDonnell, M.; Munoz, L.; Wilkinson, S.; Noll, B.; et al. Discovery and pharmacological evaluation of a novel series of adamantyl cyanoguanidines as P2X7 receptor antagonists. Eur. J. Med. Chem. 2017, 130, 433–439. [Google Scholar] [CrossRef] [PubMed]
  107. Udagawa, S.; Sakami, S.; Takemura, T.; Sato, M.; Arai, T.; Nitta, A.; Aoki, T.; Kawai, K.; Iwamura, T.; Okazaki, S.; et al. Discovery of novel 7-membered cyclic amide derivatives that inhibit 11beta-hydroxysteroid dehydrogenase type 1. Bioorg. Med. Chem. Lett. 2013, 23, 1617–1621. [Google Scholar] [CrossRef] [PubMed]
  108. Berglund, S.; Egner, B.J.; Gradén, H.; Gradén, J.; Morgan, D.G.A.; Inghardt, T.; Giordanetto, F. Optimization of piperidin-4-yl-urea-containing melanin-concentrating hormone receptor 1 (MCH-R1) antagonists: Reducing hERG-associated liabilities. Bioorg. Med. Chem. Lett. 2009, 19, 4274–4279. [Google Scholar] [CrossRef]
  109. Brogi, S.; Corelli, F.; Di Marzo, V.; Ligresti, A.; Mugnaini, C.; Pasquini, S.; Tafi, A. Three-dimensional quantitative structure–selectivity relationships analysis guided rational design of a highly selective ligand for the cannabinoid receptor 2. Eur. J. Med. Chem. 2011, 46, 547–555. [Google Scholar] [CrossRef]
  110. Lyakhovich, M.S.; Murashkina, A.V.; Averin, A.D.; Abel, A.S.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Beletskaya, I.P. Arylation of Adamantanamines: X. Palladium- and Copper-Catalyzed Heteroarylation of Adamantane-Containing Amines with Bromopyridines. Russ. J. Org. Chem. 2019, 55, 737–747. [Google Scholar] [CrossRef]
  111. Lyakhovich, M.S.; Murashkina, A.V.; Panchenko, S.P.; Averin, A.D.; Abel, A.S.; Maloshitskaya, O.A.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Arylation of Adamantanamines: XI. Comparison of the Catalytic Efficiency of Palladium and Copper Complexes in Reactions of Adamantanamines with Fluorinated 2-Bromopyridines. Russ. J. Org. Chem. 2021, 57, 768–783. [Google Scholar] [CrossRef]
  112. Hajipour, A.R.; Dordahan, F.; Rafiee, F.; Mahdavi, M. C–N cross-coupling reaction catalysed by efficient and reusable CuO/SiO2 nanoparticles under ligand-free conditions. Appl. Organomet. Chem. 2014, 28, 809–813. [Google Scholar] [CrossRef]
  113. Nador, F.; Volpe, M.A.; Alonso, F.; Radivoy, G. Synthesis of N-aryl imidazoles catalyzed by copper nanoparticles on nanosized silica-coated maghemite. Tetrahedron 2014, 70, 6082–6087. [Google Scholar] [CrossRef]
  114. Mitrofanov, A.Y.; Murashkina, A.V.; Martín-García, I.; Alonso, F.; Beletskaya, I.P. Formation of C–C, C–S and C–N bonds catalysed by supported copper nanoparticles. Catal. Sci. Technol. 2017, 7, 4401–4412. [Google Scholar] [CrossRef]
  115. Mondal, P.; Sinha, A.; Salam, N.; Roy, A.S.; Jana, N.R.; Islam, S.M. Enhanced catalytic performance by copper nanoparticle–graphene based composite. RSC Adv. 2013, 3, 5615–5623. [Google Scholar] [CrossRef]
  116. Gopiraman, M.; Ganesh Babu, S.; Khatri, Z.; Kai, W.; Kim, Y.A.; Endo, M.; Karvembu, R.; Kim, I.S. An efficient, reusable copper-oxide/carbon-nanotube catalyst for N-arylation of imidazole. Carbon 2013, 62, 135–148. [Google Scholar] [CrossRef]
  117. Khalil, A.; Jouiad, M.; Khraisheh, M.; Hashaikeh, R. Facile Synthesis of Copper Oxide Nanoparticles via Electrospinning. J. Nanomater. 2014, 2014, 438407. [Google Scholar] [CrossRef]
  118. Jammi, S.; Sakthivel, S.; Rout, L.; Mukherjee, T.; Mandal, S.; Mitra, R.; Saha, P.; Punniyamurthy, T. CuO Nanoparticles Catalyzed C–N, C–O, and C–S Cross-Coupling Reactions: Scope and Mechanism. J. Org. Chem. 2009, 74, 1971–1976. [Google Scholar] [CrossRef]
  119. Suramwar, N.V.; Thakare, S.R.; Karade, N.N.; Khaty, N.T. Green synthesis of predominant (111) facet CuO nanoparticles: Heterogeneous and recyclable catalyst for N-arylation of indoles. J. Mol. Catal. A Chem. 2012, 359, 28–34. [Google Scholar] [CrossRef]
  120. Rout, L.; Jammi, S.; Punniyamurthy, T. Novel CuO Nanoparticle Catalyzed C–N Cross Coupling of Amines with Iodobenzene. Org. Lett. 2007, 9, 3397–3399. [Google Scholar] [CrossRef]
  121. Reddy, K.H.V.; Satish, G.; Ramesh, K.; Karnakar, K.; Nageswar, Y.V.D. An efficient synthesis of N-substituted indoles from indoline/indoline carboxylic acid via aromatization followed by C–N cross-coupling reaction by using nano copper oxide as a recyclable catalyst. Tetrahedron Lett. 2012, 53, 3061–3065. [Google Scholar] [CrossRef]
  122. Murashkina, A.V.; Kuliukhina, D.S.; Averin, A.D.; Abel, A.S.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Correia, C.R.D.; Beletskaya, I.P. A comparison of homogeneous and heterogeneous copper catalyzed arylation of amines. Mendeleev Commun. 2022, 32, 91–93. [Google Scholar] [CrossRef]
  123. Fomenko, V.I.; Murashkina, A.V.; Averin, A.D.; Shesterkina, A.A.; Beletskaya, I.P. Unsupported Copper Nanoparticles in the Arylation of Amines. Catalysts 2023, 13, 331. [Google Scholar] [CrossRef]
  124. Kuliukhina, D.S.; Averin, A.D.; Panchenko, S.P.; Abel, A.S.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Correia, C.R.D.; Beletskaya, I.P. CuI and Copper Nanoparticles in the Catalytic Amination of 2-Halopyridines. Russ. J. Org. Chem. 2022, 58, 167–174. [Google Scholar] [CrossRef]
  125. Pegg, A.E.; Casero, R.A. Current Status of the Polyamine Research Field. In Polyamines: Methods and Protocols; Pegg, A., Casero, R., Jr., Eds.; Humana Press: Totowa, NJ, USA, 2011; Volume 720, pp. 3–35. [Google Scholar] [CrossRef]
  126. Díaz, J.E.; Bisceglia, J.Á.; Mollo, M.C.; Orelli, L.R. 1,n-Diamines. Part 2: Synthesis of acyclic and heterocyclic N-arylputrescine derivatives. Tetrahedron Lett. 2011, 52, 1895–1897. [Google Scholar] [CrossRef]
  127. Bisceglia, J.Á.; García, M.B.; Massa, R.; Magri, M.L.; Zani, M.; Gutkind, G.O.; Orelli, L.R. Synthesis, characterization and biological activity of bis(3-Aryl-1-hexahydropyrimidinyl)methanes. Novel heterocyclic polyamine derivatives. J. Heterocycl. Chem. 2004, 41, 85–90. [Google Scholar] [CrossRef]
  128. Haffner, C.D.; Thomson, S.A.; Guo, Y.; Petrov, K.; Larkin, A.; Banker, P.; Schaaf, G.; Dickerson, S.; Gobel, J.; Gillie, D.; et al. Substituted N-{3-[(1,1-dioxido-1,2-benzothiazol-3-yl)(phenyl)amino]propyl}benzamide analogs as potent Kv1.3 ion channel blockers. Part 2. Bioorg. Med. Chem. Lett. 2010, 20, 6989–6992. [Google Scholar] [CrossRef] [PubMed]
  129. Burns, M.R.; LaTurner, S.; Ziemer, J.; McVean, M.; Devens, B.; Carlson, C.L.; Graminski, G.F.; Vanderwerf, S.M.; Weeks, R.S.; Carreon, J. Induction of apoptosis by aryl-substituted diamines: Role of aromatic group substituents and distance between nitrogens. Bioorg. Med. Chem. Lett. 2002, 12, 1263–1267. [Google Scholar] [CrossRef] [PubMed]
  130. Bergeron, R.J.; Weimar, W.R.; Wu, Q.; Feng, Y.; McManis, J.S. Polyamine Analogue Regulation of NMDA MK-801 Binding: A Structure–Activity Study. J. Med. Chem. 1996, 39, 5257–5266. [Google Scholar] [CrossRef] [PubMed]
  131. da Costa, C.F.; Coimbra, E.S.; Braga, F.G.; dos Reis, R.C.N.; da Silva, A.D.; de Almeida, M.V. Preparation and antileishmanial activity of lipophilic N-alkyl diamines. Biomed. Pharmacother. 2009, 63, 40–42. [Google Scholar] [CrossRef] [PubMed]
  132. Panchenko, S.P.; Averin, A.D.; Anokhin, M.V.; Maloshitskaya, O.A.; Beletskaya, I.P. Cu(I)-catalyzed N,N′-diarylation of natural diamines and polyamines with aryl iodides. Beilstein J. Org. Chem. 2015, 11, 2297–2305. [Google Scholar] [CrossRef]
  133. Anokhin, M.V.; Averin, A.D.; Beletskaya, I.P. Copper-Catalyzed Arylation of Oxadiamines and Polyamines. Eur. J. Org. Chem. 2011, 2011, 6240–6253. [Google Scholar] [CrossRef]
  134. Albert, J.; Bosque, R.; Cadena, M.; D’Andrea, L.; Granell, J.; González, A.; Quirante, J.; Calvis, C.; Messeguer, R.; Badía, J.; et al. A New Family of Doubly Cyclopalladated Diimines. A Remarkable Effect of the Linker between the Metalated Units on Their Cytotoxicity. Organometallics 2014, 33, 2862–2873. [Google Scholar] [CrossRef]
  135. Li, S.A.; Cadelis, M.M.; Sue, K.; Blanchet, M.; Vidal, N.; Brunel, J.M.; Bourguet-Kondracki, M.-L.; Copp, B.R. 6-Bromoindolglyoxylamido derivatives as antimicrobial agents and antibiotic enhancers. Bioorg. Med. Chem. 2019, 27, 2090–2099. [Google Scholar] [CrossRef]
  136. Liew, L.P.P.; Pearce, A.N.; Kaiser, M.; Copp, B.R. Synthesis and in vitro and in vivo evaluation of antimalarial polyamines. Eur. J. Med. Chem. 2013, 69, 22–31. [Google Scholar] [CrossRef] [PubMed]
  137. Devi, J.; Devi, S.; Kumar, A. Synthesis, antibacterial evaluation and QSAR analysis of Schiff base complexes derived from [2,2′-(ethylenedioxy)bis(ethylamine)] and aromatic aldehydes. MedChemComm 2016, 7, 932–947. [Google Scholar] [CrossRef]
  138. Vennerstrom, J.L.; Ager, A.L.; Dorn, A.; Andersen, S.L.; Gerena, L.; Ridley, R.G.; Milhous, W.K. Bisquinolines. 2. Antimalarial N,N-Bis(7-chloroquinolin-4-yl)heteroalkanediamines. J. Med. Chem. 1998, 41, 4360–4364. [Google Scholar] [CrossRef] [PubMed]
  139. Chtchigrovsky, M.; Eloy, L.; Jullien, H.; Saker, L.; Ségal-Bendirdjian, E.; Poupon, J.; Bombard, S.; Cresteil, T.; Retailleau, P.; Marinetti, A. Antitumor trans-N-Heterocyclic Carbene–Amine–Pt(II) Complexes: Synthesis of Dinuclear Species and Exploratory Investigations of DNA Binding and Cytotoxicity Mechanisms. J. Med. Chem. 2013, 56, 2074–2086. [Google Scholar] [CrossRef]
  140. Lyakhovich, M.S.; Averin, A.D.; Grigorova, O.K.; Roznyatovsky, V.A.; Maloshitskaya, O.A.; Beletskaya, I.P. Cu(I)- and Pd(0)-Catalyzed Arylation of Oxadiamines with Fluorinated Halogenobenzenes: Comparison of Efficiency. Molecules 2020, 25, 1084. [Google Scholar] [CrossRef]
  141. Panchenko, S.P.; Averin, A.D.; Lyakhovich, M.S.; Abel, A.S.; Maloshitskaya, O.A.; Beletskaya, I.P. CuI-catalyzed hetarylation of natural di- and polyamines with halopyridines. Russ. Chem. Bull. 2017, 66, 1611–1617. [Google Scholar] [CrossRef]
  142. Anokhin, M.V.; Averin, A.D.; Panchenko, S.P.; Maloshitskaya, O.A.; Buryak, A.K.; Beletskaya, I.P. Copper(I)-Catalyzed Amination of Halogenopyridines with Polyamines. Helv. Chim. Acta 2015, 98, 47–59. [Google Scholar] [CrossRef]
  143. Anokhin, M.V.; Averin, A.D.; Panchenko, S.P.; Maloshitskaya, O.A.; Beletskaya, I.P. CuI-mediated modification of polyamines with fluorophore groups. Mendeleev Commun. 2015, 25, 245–247. [Google Scholar] [CrossRef]
  144. Anokhin, M.V.; Averin, A.D.; Panchenko, S.P.; Maloshitskaya, O.A.; Beletskaya, I.P. Copper(I)-catalyzed amination of halothiophenes with polyamines. Russ. J. Org. Chem. 2014, 50, 923–927. [Google Scholar] [CrossRef]
  145. Yakushev, A.A.; Averin, A.D.; Anokhin, M.V.; Maloshitskaya, O.A.; Lamaty, F.; Beletskaya, I.P. Copper-catalyzed amination in the synthesis of polyoxadiamine derivatives of aza- and diazacrown ethers. Macroheterocycles 2014, 7, 358–364. [Google Scholar] [CrossRef]
  146. Yakushev, A.A.; Averin, A.D.; Maloshitskaya, O.A.; Syrbu, S.A.; Koifman, O.I.; Beletskaya, I.P. Palladium- and Copper-Catalyzed Amination of Halogenophenyl Substituted Porphyrins for the Synthesis of Porphyrin-Azacrown Ethers Conjugates and Evaluation of Their Sensing Properties. Macroheterocycles 2016, 9, 65–72. [Google Scholar] [CrossRef]
  147. West, M.J.; Fyfe, J.W.B.; Vantourout, J.C.; Watson, A.J.B. Mechanistic Development and Recent Applications of the Chan–Lam Amination. Chem. Rev. 2019, 119, 12491–12523. [Google Scholar] [CrossRef] [PubMed]
  148. Kuliukhina, D.S.; Yakushev, A.A.; Malysheva, A.S.; Averin, A.D.; Beletskaya, I.P. Synthesis of N,N′-Diaryl Diamines and Oxadiamines via Chan–Lam Amination. Russ. J. Org. Chem. 2022, 58, 1752–1758. [Google Scholar] [CrossRef]
Figure 1. Most important ligands for Cu-catalyzed C–N bond formation.
Figure 1. Most important ligands for Cu-catalyzed C–N bond formation.
Catalysts 13 00831 g001
Scheme 1. CuI-catalyzed N-arylation of the adamantane-containing amines with iodobenzene.
Scheme 1. CuI-catalyzed N-arylation of the adamantane-containing amines with iodobenzene.
Catalysts 13 00831 sch001
Scheme 2. CuI-catalyzed N-arylation of the adamantane-containing amines with para-substituted iodobenzenes.
Scheme 2. CuI-catalyzed N-arylation of the adamantane-containing amines with para-substituted iodobenzenes.
Catalysts 13 00831 sch002
Scheme 3. CuI-catalyzed N-arylation of a variety of adamantane-containing amines with iodobenzenes.
Scheme 3. CuI-catalyzed N-arylation of a variety of adamantane-containing amines with iodobenzenes.
Catalysts 13 00831 sch003
Scheme 4. CuI-catalyzed N,N′-diarylation of the adamantane-containing diamines with iodobenzenes.
Scheme 4. CuI-catalyzed N,N′-diarylation of the adamantane-containing diamines with iodobenzenes.
Catalysts 13 00831 sch004
Scheme 5. CuI-catalyzed N-arylation of the adamantane-containing amines with iodobenzenes–effect of DMSO.
Scheme 5. CuI-catalyzed N-arylation of the adamantane-containing amines with iodobenzenes–effect of DMSO.
Catalysts 13 00831 sch005
Scheme 6. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 2- and 3-iodopyridines.
Scheme 6. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 2- and 3-iodopyridines.
Catalysts 13 00831 sch006
Scheme 7. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 2-fluoro-5-iodopyridne.
Scheme 7. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 2-fluoro-5-iodopyridne.
Catalysts 13 00831 sch007
Scheme 8. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 5-bromo-2-iodopyridine, 2-chloro-5-iodopyridine, and 6-bromoquinoline.
Scheme 8. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with 5-bromo-2-iodopyridine, 2-chloro-5-iodopyridine, and 6-bromoquinoline.
Catalysts 13 00831 sch008
Scheme 9. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with fluoro- and trifluoromethyl-substituted 2-bromopyridines.
Scheme 9. CuI-catalyzed N-heteroarylation of the adamantane-containing amines with fluoro- and trifluoromethyl-substituted 2-bromopyridines.
Catalysts 13 00831 sch009
Scheme 10. Initial studies of the (hetero)arylation of a model n-octylamine with iodobenzene and 2-iodopyridine using unsupported Cu and CuO nanoparticles.
Scheme 10. Initial studies of the (hetero)arylation of a model n-octylamine with iodobenzene and 2-iodopyridine using unsupported Cu and CuO nanoparticles.
Catalysts 13 00831 sch010
Scheme 11. N-arylation of the adamantane-containing amines with iodobenzene using unsupported Cu NPs.
Scheme 11. N-arylation of the adamantane-containing amines with iodobenzene using unsupported Cu NPs.
Catalysts 13 00831 sch011
Scheme 12. N-arylation of the adamantane-containing amine 3 with substituted iodobenzenes using unsupported Cu NPs.
Scheme 12. N-arylation of the adamantane-containing amine 3 with substituted iodobenzenes using unsupported Cu NPs.
Catalysts 13 00831 sch012
Scheme 13. N-heteroarylation of the adamantane-containing amines with substituted iodopyridines using unsupported Cu NPs.
Scheme 13. N-heteroarylation of the adamantane-containing amines with substituted iodopyridines using unsupported Cu NPs.
Catalysts 13 00831 sch013
Scheme 14. CuI-catalyzed N,N′-diarylation of the diamines 200203 with substituted iodobenzenes.
Scheme 14. CuI-catalyzed N,N′-diarylation of the diamines 200203 with substituted iodobenzenes.
Catalysts 13 00831 sch014
Scheme 15. CuI-catalyzed N,N′-diarylation of the tetraamine 208 with substituted halobenzenes.
Scheme 15. CuI-catalyzed N,N′-diarylation of the tetraamine 208 with substituted halobenzenes.
Catalysts 13 00831 sch015
Scheme 16. CuI-catalyzed N,N′-diarylation of the tri and tetra-amines 210213 with substituted iodobenzenes.
Scheme 16. CuI-catalyzed N,N′-diarylation of the tri and tetra-amines 210213 with substituted iodobenzenes.
Catalysts 13 00831 sch016
Scheme 17. CuI-catalyzed N,N′-diarylation of the oxadiamines 218220 with substituted iodobenzenes.
Scheme 17. CuI-catalyzed N,N′-diarylation of the oxadiamines 218220 with substituted iodobenzenes.
Catalysts 13 00831 sch017
Scheme 18. CuI-catalyzed N,N′-diheteroarylation of the diamines 200203 with 2- and 3-iodopyridines.
Scheme 18. CuI-catalyzed N,N′-diheteroarylation of the diamines 200203 with 2- and 3-iodopyridines.
Catalysts 13 00831 sch018
Scheme 19. CuI-catalyzed N,N′-diheteroarylation of the tetraamines 210, 212, 232, and 233 with 2- and 3-iodopyridines.
Scheme 19. CuI-catalyzed N,N′-diheteroarylation of the tetraamines 210, 212, 232, and 233 with 2- and 3-iodopyridines.
Catalysts 13 00831 sch019
Scheme 20. CuI-catalyzed N,N′-diheteroarylation of the naturally occurring triamine 211 and tetraamine 213 with 2- and 3-iodopyridines.
Scheme 20. CuI-catalyzed N,N′-diheteroarylation of the naturally occurring triamine 211 and tetraamine 213 with 2- and 3-iodopyridines.
Catalysts 13 00831 sch020
Scheme 21. CuI-catalyzed N,N′-diheteroarylation of the trioxadiamine 218 with 2- and 3-iodopyridines and substituted 2-bromopyridine.
Scheme 21. CuI-catalyzed N,N′-diheteroarylation of the trioxadiamine 218 with 2- and 3-iodopyridines and substituted 2-bromopyridine.
Catalysts 13 00831 sch021
Scheme 22. CuI-catalyzed N,N′-di(hetero)arylation of the polyamines 210, 218, and 233 with 2-bromo- and 2-iodonaphthalenes and 6-bromoquinoline.
Scheme 22. CuI-catalyzed N,N′-di(hetero)arylation of the polyamines 210, 218, and 233 with 2-bromo- and 2-iodonaphthalenes and 6-bromoquinoline.
Catalysts 13 00831 sch022
Scheme 23. CuI-catalyzed N,N′-di(hetero)arylation of the polyamines with 3-iodothiophene.
Scheme 23. CuI-catalyzed N,N′-di(hetero)arylation of the polyamines with 3-iodothiophene.
Catalysts 13 00831 sch023
Scheme 24. CuI-catalyzed amination of 3-iodobenzyl-substituted 1-aza-18-crown-6 ether 264.
Scheme 24. CuI-catalyzed amination of 3-iodobenzyl-substituted 1-aza-18-crown-6 ether 264.
Catalysts 13 00831 sch024
Scheme 25. CuI-catalyzed N,N′-diarylation and diamination of the derivatives of the aza- and diazacrown ethers.
Scheme 25. CuI-catalyzed N,N′-diarylation and diamination of the derivatives of the aza- and diazacrown ethers.
Catalysts 13 00831 sch025
Scheme 26. Chan–Lam N-arylation of the adamantane-containing amines with the model p-tolylboronic acid.
Scheme 26. Chan–Lam N-arylation of the adamantane-containing amines with the model p-tolylboronic acid.
Catalysts 13 00831 sch026
Scheme 27. Introduction of the naphthyl and quinolinyl substituents to adamantane-containing amines via Chan–Lam reaction.
Scheme 27. Introduction of the naphthyl and quinolinyl substituents to adamantane-containing amines via Chan–Lam reaction.
Catalysts 13 00831 sch027
Scheme 28. N,N′-diarylation of the diamines and oxadiamines with the model p-tolylboronic acid.
Scheme 28. N,N′-diarylation of the diamines and oxadiamines with the model p-tolylboronic acid.
Catalysts 13 00831 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

Averin, A.D.; Panchenko, S.P.; Murashkina, A.V.; Fomenko, V.I.; Kuliukhina, D.S.; Malysheva, A.S.; Yakushev, A.A.; Abel, A.S.; Beletskaya, I.P. Recent Achievements in the Copper-Catalyzed Arylation of Adamantane-Containing Amines, Di- and Polyamines. Catalysts 2023, 13, 831. https://doi.org/10.3390/catal13050831

AMA Style

Averin AD, Panchenko SP, Murashkina AV, Fomenko VI, Kuliukhina DS, Malysheva AS, Yakushev AA, Abel AS, Beletskaya IP. Recent Achievements in the Copper-Catalyzed Arylation of Adamantane-Containing Amines, Di- and Polyamines. Catalysts. 2023; 13(5):831. https://doi.org/10.3390/catal13050831

Chicago/Turabian Style

Averin, Alexei D., Svetlana P. Panchenko, Arina V. Murashkina, Varvara I. Fomenko, Daria S. Kuliukhina, Anna S. Malysheva, Alexei A. Yakushev, Anton S. Abel, and Irina P. Beletskaya. 2023. "Recent Achievements in the Copper-Catalyzed Arylation of Adamantane-Containing Amines, Di- and Polyamines" Catalysts 13, no. 5: 831. https://doi.org/10.3390/catal13050831

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