Skip to content
BY 4.0 license Open Access Published by De Gruyter March 8, 2020

The oxidative coupling between benzaldehyde derivatives and phenylacetylene catalyzed by rhodium complexes via C-H bond activation

  • Xinyi Zhao , Hongge Jia EMAIL logo , Qingji Wang , Heming Song , Yanan Tang , Liqun Ma , Yongqiang Shi , Guoxing Yang , Yazhen Wang , Yu Zang and Shuangping Xu

Abstract

This paper reports the use of rhodium (Rh) catalysts for the oxidative coupling reaction between phenylacetylene and benzaldehyde derivatives via C-H bond activation. These reactions were catalyzed by Rh(l-amino acid)(cod) (the l-amino acid is l-phenylalanine, l-valine or l-proline; cod is 1,5-cyclooctadiene) to obtain chromones in 12.7–88.3% yield. These new Rh catalysts have excellent activity for the coupling reaction between phenylacetylene and different benzaldehyde derivatives. It was found that the electronic effects of the benzaldehyde derivative substituent affected the reaction yield, which is in accordance with the proposed mechanism.

Introduction

In previous studies, rhodium catalysts have been used in many different domains, such as oxidative coupling reactions [1,2], activated chemical bond reactions [3,4], polymerizations [5,6], and asymmetric hydrogenation reactions [7,8]. Rh complexes have shown high catalytic activity, especially in the case of polymerizations of various substituted acetylenes [9,10,11]. We previously reported two kinds of novel Rh complexes containing l-amino acids and 1,5-cyclooctadiene (cod) as ligands to catalyze the helical selective polymerization of m-dihydroxyphenylacetylene, 3,4,5-trisubstituted phenylacetylene, and p‐dodecyloxy‐m,m‐dihydroxyphenylacetylene [11].

Oxidative coupling reactions have been carried out by using Au catalysts [12,13], Fe catalysts [14], Ir catalysts [15,16], Cu catalysts [17,18], Pd catalysts [19,20], and Ru catalysts [21]. Even if Rh catalysts are expensive, they have unique advantages because of their functional group tolerance, wide range of substrate scope, high activity, and selectivity for synthetic transformations [22, 23, 24, 25, 26]. Rh catalysts are used by many researchers in C-H activation reactions [27, 28, 29]. One kind of Rh catalyst, which contains Cp* (Cp* = 1,2,3,4,5-Pentamethylcyclopentadiene (C5Me5) ligands, has been used to catalyze oxidative coupling reactions. For example, Lu et al. reported that [Cp*RhCl2]2 catalyzed the C-H activation of pyrazolones with symmetric 1,6-enynes to obtain functional products in 34%–72% yield [27]. Krieger’s team used [Cp*RhCl2]2 to obtain macrocyclic pyridones in excellent yields, and proposed that the catalyst could become more efficient in redox-neutral processes when the O-pivaloyl hydroxamate is the directing group [28]. The alkyne annulation with peresters, as the oxidizing directing group, was also catalyzed by [Cp*RhCl2]2 via C−H activation [29].

We synthesized Rh(l-amino acid)(cod)s, which are chiral, highly selective and highly active Rh complexes [11]. In this paper, these Rh complexes were used in oxidative coupling reactions between phenylacetylene and benzaldehyde derivatives. However, we did not determine the actual mechanism of the Rh-diene complexes, i.e., Rh(l-amino acid)(cod), in oxidative coupling reactions, but we have discussed a possible mechanism in our previous articles [30]. In this study, the oxidative coupling reaction between phenylacetylene and benzaldehyde derivatives was catalyzed by Rh with an l-amino acid ligand, plus the catalytic efficiencies and substituent electronic effects are discussed.

Results and discussion

The oxidative coupling reactions between phenylacetylene and the benzaldehyde derivatives (1a-1g) were carried out by using Rh catalysts at 120°C in an o-xylene solution (Scheme 2). The Rh catalysts can activate the C-H bond in compounds 1a-1g to react with the triple bond of phenylacetylene, which produces final compounds (3a-3g). Products were purified by extraction and silica gel column chromatography, before analysis of the product chemical structures by 1H-NMR, 13C-NMR and IR (See SI, Figures S13-S30), it was concluded that the novel compounds (3a-3g) were synthesized successfully.

Scheme 1 Synthetic route to Rh(l-amino acids)(cod).
Scheme 1

Synthetic route to Rh(l-amino acids)(cod).

Scheme 2 Synthetic route to chromones.
Scheme 2

Synthetic route to chromones.

We hypothesize that the Rh-N bond between the l-amino acid ligand and Rh metal centre is active [26]. Rhodium easily reacts with an atom that has more electrons in its outermost shell to form a new active Rh species [31]. In the oxidative coupling reaction process, the hydroxyl oxygen atom in the benzaldehyde derivative replaces the nitrogen atom of l-amino acid in Rh(l-amino acid)(cod) to coordinate with Rh resulting in intermediate A (Scheme 3). The triple bond of phenylacetylene attacks the Rh centre displacing the L-amino acid ligand, leaving an empty catalytic site (see B in Scheme 3). Then, the triple bond of phenylacetylene inserts into this unoccupied site at Rh to afford ring structure C [30]. Rh(l-amino acid)(cod) is regenerated using a Cu oxidant [32], and the desired product (3a) is obtained. Rh catalyst with a cyclopentadiene ligand and Cu oxidant effectively catalyzes the cyclization reaction of salicylaldehydes and alkynes to produce chromone derivatives [32].

Scheme 3 Plausible reaction mechanism.
Scheme 3

Plausible reaction mechanism.

We discussed the influence of l-amino acid ligands for these oxidative coupling reactions [30]. When Rh(l-Phe) (cod) was used as catalyst to obtain the highest yield of 3a (see Table 1), as Rh(l-Phe)(cod) has the best catalytic effect among the obtained Rh complexes. We considered that as l-Phe has a large phenyl group, which can easily dissociate from Rh, providing an empty coordination site at Rh for phenylacetylene insertion, yielding a RhIX species [30].

Table 1

Oxidative coupling reactions between dihydroxybenzaldehyde [1a] and phenylacetylene catalyzed by rhodium with amino acid ligands [a]

No.ReactantRh CatalystYield (%)
1Rh(cod)(l-Phe)82.9
2Rh(cod)(l-Val)63.7
3Rh(cod)(l-Pro)54.2

To investigate the effect of the benzaldehyde derivatives on the oxidative coupling reaction, seven chemical reactions between 1a-1g and 2 were catalyzed by Rh(l-Phe) (cod), as shown in Table 2. All of these benzaldehyde derivatives (1a-1g) reacted with phenylacetylene to obtain chromones 3a-3g in 12.7%–88.3% yield. These results illustrated that Rh catalysts with l-amino acid ligands have a wide range of applications for different substrates.

The reason for such a wide range of yields in Table 2 can be explained by the role of the substituent effects. Aldehyde groups and hydroxyl groups on the phenyl ring can react with the triple bond of phenylacetylene, therefore the nature and position of ring substituents affects the reactivity of the aromatic compound. The benzaldehyde that has two hydroxyl groups in the ortho- and para-positions (Table 2, 1a) reacted with phenylacetylene to obtain the product in 82.9% yield, which is lower than the yield obtained by unsubstituted benzaldehyde as the substrate (Table 2, 1b, 88.0%) but higher than the yield afforded by the substrate containing three hydroxyl groups (Table 2, 1c and 1d). In particular, 2,4,6-trihydroxybenzaldehyde had the lowest yield of 22.6% (Table 2, 1d). The reason for these results could be that hydroxyl groups in the ortho-or para-position can increase the electron density of the benzene ring. The electronegativity of the carbon atom closest to the aldehyde group is enhanced by the hydroxyl groups (Table 2, 1a, 1c, 1d). As the number of hydroxyl groups increases, the electronegativity of the benzaldehyde also increases. Compound 1d with two ortho and one para hydroxyl group has a higher electronegativity than 1c, in which the ortho-position, meta-position and para-position each contain one hydroxyl group. Therefore, the electronegativity order of the carbon in position 1 is 1d > 1c >1a> 1b (see chart 1), which is contrary to their yield order (1d < 1c <1a < 1b in Table 2). The reason for this phenomenon can be explained as the electronegative triple bond of phenylacetylene can enter into the unoccupied active site on Rh. This sequence results in ring structure formation with the positive carbon atom of the aldehyde group in 1a-1d. The electronegativity of the carbon atom at position 1 is weakened by the positive electricity of carbon in the aldehyde group. This is not beneficial for the formation of ring structures, between benzaldehyde derivative and phenylacetylene in oxidative coupling reactions. Consequently, the yield order is opposite to the number of hydroxyl groups because of the effect of the electron density in the aromatic ring.

Chart 1 Sequence of carbon in 1a-1g.
Chart 1

Sequence of carbon in 1a-1g.

There are also similar influences for the oxidative coupling reactions between halogen substituted benzaldehydes (1e-1g) and phenylacetylene (2). Chlorine and bromine are electron withdrawing groups, as an electronic effect, but are also electron-donating groups via conjugation effects. The conjugation effect of 1g with two bromine groups is stronger than 1e and 1f, and so the electronegativity of carbon in position 1 of 1g is the strongest among 1e-1g. Increasing the electronegativity of the carbon in position 1 is not conducive to form a ring structure between benzaldehyde derivative and phenylacetylene.

Table 2

Oxidative coupling reactions between aldehyde derivatives (1a-1g) and phenylacetylene using Rh(l-Phe)(cod) as the catalyst[a]

So, the yield of 1g is low (12.7%) (Table 2, 7). The carbon atom in position 1 of 1e has a lower electronegativity than those of 1f and 1g, so the yield of 3e was higher than those of 3f and 3g (Table 2, 5-7). However, all of these reaction yields were lower than the yield provided by benzaldehyde and phenylacetylene (Table 2, 1). The reason for this result is that halogen and hydroxyl can weaken the ability of benzaldehyde derivatives to form a ring structure with phenylacetylene.

Conclusions

In summary, the C-H bond activation reaction of benzaldehyde derivatives was demonstrated in the presence of novel Rh catalysts with l-amino acid ligand. Seven kinds of chromones were synthesized. Rh(l-Phe) (cod) had the best catalytic effect among the obtained Rh complexes. l-amino acids containing a large aromatic group can dissociate easily to afford an empty Rh active site, which can improve the catalytic activity of the Rh complexes. The kind of substituent and its position on the aromatic ring of benzaldehyde affects the oxidative coupling reaction by electronic effects.

Experimental

Synthesis of 3-phenyl-7-hydroxy-1H-chromen-4-one (3a): Rh(l-phe)(cod) (1.50mg, 3.98 × 10-6mol), 1,2,3,4-tetraphenyl-1,3-cyclopentadiene (7.04mg, 1.90 × 10-5mol) and Cu(OAc)2·H2O (200mg, 1.00 × 10-3mol) were dissolved in o-xylene (2.76mL, 2.29 × 10-5mol), 2,4-dihydroxybenzaldehyde (6.04 × 102mg, 4.37 × 10-3mol) and phenylacetylene (4.91 × 10-1mL, 4.47 × 10-3mol) were added. The mixture was stirred under N2 at 120°C for 6 h. The product was extracted by ether and water. The final product was isolated in n-hexane and ethyl acetate by silica gel column chromatography. Other chromones were obtained by a similar process as described for 3a.

3-phenyl-7-hydroxy-1H-chromen-4-one (3a): The compound was obtained in 82.9% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.57, (d, 2H, H-2, H-5, J=1.2Hz), 7.56 (d, 2H, H-6, H-8, J=1.2Hz), 7.40 (t, 1H, H=4', J=3Hz), 7.39 (t, 1H, H-5', J=6Hz), 7.38 (t, 1H, H-3', J=3Hz), 7.37 (s, 1H, H-1'), 7.36 (s, 1H, H-2'), 7.35 (t, 1H, H-4', J=3Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 72.65, 80.54, 120.73, 127.65, 128.19, 131.72 ppm. IR (KBr): 3063, 2964, 2924, 2851, 2218, 1951, 1485, 1023 cm-1. (See Figures S13-S15 of Supporting Information)

3-phenyl-1H-chromen-4-one (3b): The compound was obtained in 88.3% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.63 (s, 2H, H-2, H-5), 7.62 (d, 3H, H-6, H-7, H-8, J=1.2Hz), 7.52 (t, 1H, H-4', J=3Hz), 7.51 (t, 1H, H-5', J=6Hz), 7.50 (t, 1H, H-3', J=3Hz), 7.47 (s, 1H, H-1'), 7.46 (s, 1H, H-2'), 7.45 (t, 1H, H-4', J=3Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 77.24, 81.57, 121.83, 128.46, 129.23, 132.53 ppm. IR (KBr): 3050, 2953, 2926, 2855, 2191, 1741, 1485, 1252 cm-1. (See Figures S16-S18 of Supporting Information)

3-phenyl-7,8-dihydroxy-1H-chromen-4-one (3c): The compound was obtained in 69.7% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.54 (d, 2H, H-2, H-5, J=2.4Hz), 7.53 (d, 1H, H-6, J=3.6Hz), 7.39 (t, 1H, H=4', J=3Hz), 7.38 (t, 1H, H-5', J=6Hz), 7.37 (t, 1H, H-3', J=3Hz), 7.35 (s, 1H, H-1'), 7.34 (d, 1H, H-2', J=1.2Hz), 7.33 (t, 1H, H-4', J=3.6Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 73.91, 81.56, 121.82, 128.45, 129.22, 132.52 ppm. IR (KBr): 3064, 2973, 2918, 2198, 1961, 1487, 1460, 1013 cm-1. (See Figures S19-S21 of Supporting Information)

3-phenyl-5,7-dihydroxy-1H-chromen-4-one (3d): The compound was obtained in 22.6% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.54, (d, 2H, H-2, J=2.4Hz), 7.53 (d, 2H, H-6, H-8, J=3.6Hz), 7.39 (t, 1H, H=4', J=3Hz), 7.38 (t, 1H, H-5', J=6Hz), 7.37 (t, 1H, H-3', J=3Hz), 7.36 (s, 1H, H-1'), 7.34 (d, 1H, H-2', J=1.2Hz), 7.33 (t, 1H, H-4', J=3Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 73.93, 81.57, 121.81, 128.46, 129.23, 132.53 ppm. IR (KBr): 2978, 2931, 2912, 2891, 1714, 1443, 1055, 882 cm-1. (See Figures S22-S24 of Supporting Information)

3-phenyl-6-chloro-1H-chromen-4-one (3e): The compound was obtained in 49.5% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.55, (s, 2H, H-2, H-5), 7.54 (d, 2H, H-7, H-8, J=1.2Hz), 7.39 (s, 1H, H=4'), 7.38 (s, 1H, H-5'), 7.37 (s, 1H, H-3'), 7.36 (s, 1H, H-1'), 7.35 (s, 1H, H-2'), 7.33 (d, 1H, H-4'). 13C-NMR (150 MHz, CDCl3, TMS, δ): 72.99, 80.54, 118.45, 120.82, 128.26, 127.49, 131.39 ppm. IR (KBr): 3047, 2949, 2921, 2848, 2142, 1951, 1660, 1468 cm-1. (See Figures S25-S27 of Supporting Information)

3-phenyl-6-bromo-1H-chromen-4-one (3f): The compound was obtained in 29.3% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.54, (s, 2H, H-2, H-5), 7.53 (d, 2H, H-7, H-8, J=1.2Hz), 7.39 (s, 1H, H=4'), 7.37 (t, 1H, H-5', J=6Hz), 7.36 (s, 1H, H-3'), 7.35 (s, 1H, H-1'), 7.34 (s, 1H, H-2'), 7.33 (t, 1H, H-4', J=3Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 72.91, 80.47, 118.78, 120.85, 127.42, 128.18, 131.47 ppm. IR (KBr): 3057, 2954, 2912, 2849, 2214, 1951, 1598, 1266 cm-1. (See Figures S28-S30 of Supporting Information)

3-phenyl-6,8-dibromo-1H-chromen-4-one (3g): The compound was obtained in 12.7% yield as a brown oil. 1H-NMR (600 MHz, CDCl3, TMS, δ): 7.54, (s, 2H, H-2, H-5), 7.53 (d, 2H, H-7, J=1.2Hz), 7.39 (s, 1H, H=4'), 7.38 (t, 1H, H-5', J=6Hz), 7.37 (s, 1H, H-3'), 7.36 (s, 1H, H-1'), 7.34 (s, 1H, H-2'), 7.33 (d, 1H, H-4', J=1.2Hz). 13C-NMR (150 MHz, CDCl3, TMS, δ): 72.89, 80.53, 120.86, 127.42, 128.23, 131.44 ppm. IR (KBr): 3053, 2966, 2917, 2857, 2144, 1880, 1484, 1260 cm-1. (See Figures S31-S33 of Supporting Information)

Instruments: The molecular structures of the products were determined by NMR [Bruker, Germany; 1H (600 MHz) and 13C (150 MHz)] and IR [Perkin-Elmer, America; Spectrum One B IR spectrophotometer]. Yield = Actual product weight ÷Theoretical product weight. Theoretical product weight = Theoretical mole number of product × molecular weight of product.


Tel.: +86-0452-2738752

Acknowledgments

Financial and facility support for this research came from the Fundamental Research Funds in Heilongjiang provincial universities (YSTSXK201862 135309110, 135309503), and the National Natural Science Foundation of Heilongjiang province, China (LH2019B032).

References

[1] Matsumoto K, Tachikawa S, Hashimoto N, Nakano R, Yoshida M, Shindo M. Aerobic C-H oxidation of arenes using a recyclable, heterogeneous rhodium catalyst. J Org Chem. 2017;82:4305–16.10.1021/acs.joc.7b00300Search in Google Scholar PubMed

[2] Elouarzaki K, Goff AL, Holzinger M, Thery J, Cosnier S. Electrocatalytic oxidation of glucose by rhodium porphyrinfunctionalized MWCNT electrodes: application to a fully molecular catalyst-based glucose/O2 fuel cell. J Am Chem Soc. 2012;134:14078–85.10.1021/ja304589mSearch in Google Scholar PubMed

[3] Ghorai D, Choudhury J. Rhodium(III)-N-Heterocyclic carbene-driven cascade C-H activation catalysis. ACS Catal. 2015;5:2692–6.10.1021/acscatal.5b00243Search in Google Scholar

[4] Luo X, Bai R, Liu S, Shan C, Chen C, Lan Y. Mechanism of rhodium-catalyzed formyl activation: a computational study. J Org Chem. 2016;81:2320–6.10.1021/acs.joc.5b02828Search in Google Scholar PubMed

[5] Abe S, Hirata K, Ueno T, Morino K, Shimizu N, Yamamoto M, et al. Polymerization of phenylacetylene by rhodium complexes within a discrete space of apo-ferritin. J. ACS. 2009;131:6958–60.10.1021/ja901234jSearch in Google Scholar PubMed

[6] Nikishkin N, Huskens J, Verboom W. Highly active and robust rhodium(I) catalyst for the polymerization of arylacetylenes in polar and aqueous medium under air atmosphere. Polymer (Guildf). 2013;54:3175–81.10.1016/j.polymer.2013.04.028Search in Google Scholar

[7] Yang H, Huo N, Yang P, Pei H, Lv H, Zhang X. Rhodium catalyzed asymmetric hydrogenation of 2-pyridine ketones. Org Lett. 2015;17:4144–7.10.1021/acs.orglett.5b01878Search in Google Scholar PubMed

[8] Li P, Zhou M, Zhao Q, Wu W, Hu X, Dong X, et al. Synthesis of chiral β‑amino nitroalkanes via rhodium-catalyzed asymmetric hydrogenation. Org Lett. 2016;18:40–3.10.1021/acs.orglett.5b03158Search in Google Scholar PubMed

[9] Jia H, Teraguchi M, Aoki T, Abe Y, Kaneko T, Hadano S, et al. Two modes of asymmetric polymerization of phenylacetylene having a l-valinol residue and two hydroxy groups. Macromolecules. 2009;42:17–9.10.1021/ma802313zSearch in Google Scholar

[10] Liu L, Zang Y, Hadano S, Aoki T, Teraguchi M, Kaneko T, et al. New achiral phenylacetylene monomers having an oligosiloxanyl group most suitable for helix-sense-selective polymerization and for obtaining good optical resolution membrane materials. Macromolecules. 2010;43:9268–76.10.1021/ma101999kSearch in Google Scholar

[11] Jia H, Shi Y, Ma L, Gao X, Wang Y, Zang Y, et al. Novel isolated, l-amino acid ligand rhodium catalysts that induce highly helix-sense-selective polymerization of an achiral 3,4,5-trisubstituted phenylacetylene. Polym Chem. 2016;54:2346–51.10.1002/pola.28106Search in Google Scholar

[12] Brenzovich WE Jr, Brazeau JF, Toste FD. Gold-catalyzed oxidative coupling reactions with aryltrimethylsilanes. Org Lett. 2010;12:4728–31.10.1021/ol102194cSearch in Google Scholar PubMed PubMed Central

[13] Xu B, Madix RJ, Friend CM. Predicting gold-mediated catalytic from oxidative coupling reactions from single crystal studies. Acc Chem Res. 2014;47:761–72.10.1021/ar4002476Search in Google Scholar PubMed

[14] Du W, Tian L, Lai J, Huo X, Xie X, She X, et al. Iron-catalyzed radical oxidative coupling reaction of aryl olefins with 1,3-dithiane. Org Lett. 2014;16:2470–3.10.1021/ol500850dSearch in Google Scholar PubMed

[15] Zhou T, Li L, Li B, Song H, Wang B. Ir(III)-catalyzed oxidative coupling of NH isoquinolones with benzoquinone. Org Lett. 2015;17:4204–7.10.1021/acs.orglett.5b01974Search in Google Scholar PubMed

[16] Wang Z, Yang M, Yang Y. Ir(III)-catalyzed oxidative annulation of phenylglyoxylic acids with benzobthiophenes. Org Lett. 2018;20:3001–5.10.1021/acs.orglett.8b01033Search in Google Scholar PubMed

[17] Boess E, Schmitz C, Klussmann M. A comparative mechanistic study of cu-catalyzed oxidative coupling reaction with N-phenyltetrahydroisoquinoline. J Am Chem Soc. 2012;134:5317–25.10.1021/ja211697sSearch in Google Scholar PubMed

[18] Zhang Q, Liu Q, Wang T, Zhang X, Long C, Wu Y, et al. Mechanistic study on cu(II)-catalyzed oxidative cross-coupling reaction between arenes and boronic acids under aerobic conditions. J Am Chem Soc. 2018;140:5579–87.10.1021/jacs.8b01896Search in Google Scholar PubMed

[19] Wang D, Stahl S. Pd-catalyzed aerobic oxidative biaryl coupling: non-redox cocatalysis by Cu(OTf)2 and discovery of Fe(OTf)3 as a highly effective cocatalyst. J Am Chem Soc. 2017;139:5704–7.10.1021/jacs.7b01970Search in Google Scholar PubMed PubMed Central

[20] Liu W, Li Y, Wang Y, Kuang C. Pd-catalyzed oxidative CH/CH direct coupling of heterocyclic N‑oxides. Org Lett. 2013;15:4682–5.10.1021/ol4019776Search in Google Scholar PubMed

[21] Baruah S, Kaishap P, Gogoi S. Ru(II)-catalyzed C–H activation and annulation of salicylaldehydes with monosubstituted and disubstituted alkynes. Chem Commun (Camb). 2016;52:13004–7.10.1039/C6CC07204FSearch in Google Scholar

[22] Park YJ, Park JW, Jun CH. Metal−organic cooperative catalysis in C−H and C−C bond activation and its concurrent recovery. Acc Chem Res. 2008;41:222–34.10.1021/ar700133ySearch in Google Scholar PubMed

[23] Satoh T, Miura M. Oxidative coupling of aromatic substrates with alkynes and alkenes under rhodium catalysis. Chemistry. 2010;16:11212–22.10.1002/chem.201001363Search in Google Scholar PubMed

[24] Colby DA, Bergman RG, Ellman J. Rhodium-catalyzed C−C bond formation via heteroatom-directed C−H bond activation. A. Chem. Rev. 2010;110:624–55.10.1021/cr900005nSearch in Google Scholar PubMed PubMed Central

[25] Colby DA, Tsai AS, Bergman RG, Ellman J. Rhodium catalyzed chelation-assisted C–H bond functionalization reactions. A. Acc. Chem. Res. 2012;45:814–25.10.1021/ar200190gSearch in Google Scholar PubMed PubMed Central

[26] Azpíroz R, Rubio-Perez L, Giuseppe AD, Passarelli V, Lahoz FJ, Castarlenas R, et al. Rhodium(I)-N-Heterocyclic carbene catalyst for selective coupling of N-vinylpyrazoles with alkynes via C-H activation. ACS Catal. 2014;4:4244–53.10.1021/cs501366qSearch in Google Scholar

[27] Lu H, Fan Z, Xiong C, Zhang A. Highly stereo selective assembly of polycyclic molecules from 1,6-enynes triggered by rhodium(III)-catalyzed C-H activation. Org Lett. 2018;20:3065–9.10.1021/acs.orglett.8b01099Search in Google Scholar PubMed

[28] Krieger J, Lesuisse D, Ricci G, Perrin M, Meyer C, Cossy J. Rhodium(III)-catalyzed C-H activation /heterocyclization as a macrocyclization strategy synthesis of macrocyclic pyridones. Org Lett. 2017;19:2706–9.10.1021/acs.orglett.7b01051Search in Google Scholar PubMed

[29] Mo J, Wang L, Cui X. Rhodium(III)-catalyzed C−H activation/ alkyne annulation by weak coordination of peresters with O−O bond as an internal oxidant. Org Lett. 2015;17:4960–3.10.1021/acs.orglett.5b02291Search in Google Scholar PubMed

[30] Jia H, Tang Y, Shi Y, Ma L, He Z, Lai W, et al. Rhodium complexes catalyze oxidative coupling between salicylaldehyde and phenylacetylene via C–H bond activation. Chem Pap. 2017;71:1791–5.10.1007/s11696-017-0153-4Search in Google Scholar

[31] Luo C, Jayakumar J, Gandeepan P, Wu Y, Cheng C. Rhodium(III)-catalyzed vinylic C−H activation: a direct route toward pyridinium salts. Org Lett. 2015;17:924–7.10.1021/acs.orglett.5b00028Search in Google Scholar PubMed

[32] Shimizu M, Tsurugi H, Satoh T, Miura M. Rhodium-catalyzed oxidative coupling between salicylaldehydes and internal alkynes with C-H bond cleavage to produce 2,3-disubstituted chromones. Chem Asian J. 2008;3:881–6.10.1002/asia.200800037Search in Google Scholar PubMed

Received: 2019-09-25
Accepted: 2020-01-30
Published Online: 2020-03-08

© 2020 Zhao et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 5.5.2024 from https://www.degruyter.com/document/doi/10.1515/hc-2020-0004/html
Scroll to top button