Aerobic Cleavage of Alkenes and Alkynes into Carbonyl and Carboxyl

Mar 29, 2017 - Layla Filiciotto , Alina M. Balu , Antonio A. Romero , Enrique Rodríguez-Castellón , Jan C. van der Waal , Rafael Luque. Green Chem. ...
0 downloads 0 Views 2MB Size
Review pubs.acs.org/acscatalysis

Aerobic Cleavage of Alkenes and Alkynes into Carbonyl and Carboxyl Compounds Garazi Urgoitia, Raul SanMartin,* María Teresa Herrero, and Esther Domínguez* Department of Organic Chemistry II, Faculty of Science and Technology, University of the Basque Country (UPV-EHU), 48940 Leioa, Spain ABSTRACT: This review provides an overview of the recent developments on the oxygen-mediated cleavage of olefins and acetylene derivatives to provide carbonyl and carboxylic compounds. A number of photo, organo, and metal catalysts have been reviewed with an emphasis on those exhibiting superior performance. For this purpose, crucial aspects such as pressure, substrate scope, product selectivity for the target aldehydes, ketones, or carboxylic acids, catalyst amount, and recyclability are discussed. Although many improvements are required to match the results provided by conventional oxidants, aerobic cleavage of such unsaturated hydrocarbons is a very promising field in progress on account of the significant achievements extracted from the recent literature. Several postulated mechanisms are also provided to gain insight into the role of the catalysts and the incorporation of oxygen atoms into the products. KEYWORDS: oxidative cleavage, molecular oxygen, alkenes, alkynes, carbonyl compounds, carboxylic acids



INTRODUCTION The abundance and diversity of alkenes and alkynes from natural sources and coal/petroleum products make them a suitable source for the preparation of structurally more complex compounds.1 Among the many ways to functionalize such unsaturated hydrocarbons, oxidative cleavage to generate carbonyl or carboxy compounds, a transformation easily exemplified by ozonolysis, has been exploited for decades and still constitutes a preliminary step in many synthetic approaches.2 Moreover, oxidative cleavages can be also found at advanced stages of routes toward products of interest (Grandilodine C alkaloid, Penifulvin B and C sesquiterpenoids, and neuroactive benz[e]indenes, inter alia).3 The toxicity and potential risk of ozone and other oxidants and the stoichiometric amounts of waste generated from the use of other oxidizing agents (KMnO4, NaIO4, TBHP, OsO4, oxone, I(III) reagents)4,5 have encouraged the search for more sustainable reagents. In this context, despite the interest in hydrogen peroxide mediated oxidations described so far,6 oxygen is the oxidant of choice in terms of abundance, cheapness, safety, and lack of harmful waste, and a number of reports on the oxygenmediated cleavage of alkenes and alkynes have appeared in the literature. A review on this subject can be found here, with an emphasis on metal-catalyzed aerobic cleavages, especially those aided by tailored ligands. Only the direct preparation of carbonyl and carboxylic compounds from olefins/acetylenes will be covered (Scheme 1), thus leaving aside the topic of aerobic stepwise © XXXX American Chemical Society

Scheme 1. Schematic Generalization of the Molecular Oxygen Mediated Cleavage of Unsaturated Hydrocarbons

cleavages via formation of epoxides or 1,2-diols. Homogeneous and heterogeneous metal catalysts as well as other metal-free systems have been successfully applied to the subject reaction, and this review has been organized according to the catalyst type and to literature reports on the comparison of different metal sources associated with the same catalyst system. As in many other oxygenation reactions,7 the mechanism of such aerobic transformations is far from being completely understood, but several proposals have been presented and will be briefly discussed, showing that in some cases the pathways for the reaction seem to be catalyst dependent but several meaningful similarities can be found. The formation of different products along with the target carbonyl and carboxyl compounds due to competing side reactions (alkene isomerization, epoxidation, Wacker-type Received: December 23, 2016 Revised: March 20, 2017

3050

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis

reaction media, and the experience gained in their previous reports led the authors to propose a mechanism where an initial iodomethoxylation in the presence of trifluoroacetic acid followed by a visible-light-induced homolytic cleavage of the C−I bond generates benzyl radical species A. Interaction with molecular oxygen provides peroxy radical B, and the corresponding hydroperoxide C undergoes C−C bond cleavage under acidic conditions (Scheme 3).11

oxidation, etc.) is a common feature of many of the reported procedures. Therefore, special attention will be paid to product selectivity and to substrate limitations as well. In accordance with the sustainability principles derived from the use of molecular oxygen, aspects such as recyclability for heterogeneous systems and/or catalyst loadings for homogeneous catalysts will also be discussed.



METAL-FREE CATALYSTS In order to avoid the use of precious-metal catalysts, metal-free systems have been developed with relative success in the molecular oxygen mediated cleavage of alkenes. Light irradiation and the use of substoichiometric amounts of several organic oxidizers or free radical promoters are the most common strategies reported to date. As such, Zhao and coworkers reported the preparation of benzaldehydes from styrene derivatives in the absence of any catalyst under UV light irradiation with a 100 W medium-pressure Hg lamp (λ 330−400 nm). Water was the optimal solvent for a better selectivity toward carbonyl compounds (benzyl alcohols were also detected), but yields were moderate and the procedure could be applied to neither stilbenes nor aliphatic alkenes.8 The presence of O2•− and •OOH reactive species was detected by electron paramagnetic resonance, and the participation of singlet oxygen or hydroxy radicals was discarded on the basis of several experiments (negative results from reactions using Rose Bengal or a photo-Fenton reaction, partial cleavage of 1,1diphenylethylene, etc.). The Itoh group has published several photooxidative cleavages, mainly from aromatic alkenes and alkynes (Scheme 2). In 2009, they described the combined use of CBr4 and UV

Scheme 3. Reaction Pathway for the Iodine-Catalyzed Cleavage of Stilbene under Visible Light Irradiation

Visible light along with an organophotoredox catalyst or photosensitizer has been also exploited by You12 and Yadav.13 5,10,15-Triphenyl-20-(4-hydroxyphenyl)-21H,23H-porphyrin (TPP-OH) provided carbonyl compounds with variable results highly dependent on the structure of the olefin. According to the authors, a cycloaddition between photosensitizer-induced singlet oxygen and the alkene generates a 1,2-dioxetane which upon reaction conditions collapses to provide the corresponding aldehydes (Scheme 4).12 A series of benzaldehydes were obtained in good yields from styrenes under mild conditions (room temperature in DMSO) by irradiation with green lightemitting diodes in the presence of catalytic amounts (1 mol %) of Eosin Y.13 A very recent report by Wang and co-workers shows an advantageous cleavage of styrene derivatives under white-LED

Scheme 2. Photooxidative Cleavages Reported by Itoh and Co-workers

Scheme 4. Examples and Proposed Mechanism of the Formation of Aldehydes from Alkenes in the Presence of 1 equiv of TPP-OH Photosensitizer

irradiation (4000W Hg lamp) as a procedure to generate carboxylic acids from terminal olefins. An aliphatic alkene (1dodecene) could be also cleaved with a moderate 32% yield even in the presence of higher amounts (60 mol %) of CBr4.9 Water was employed as additive in a similar procedure applied to terminal and internal alkynes.10 A recent contribution by the group is based on the use of visible light and molecular iodine along with trifluoroacetic acid for the synthesis of aromatic aldehydes from trans-stilbene derivatives. Isolation of several intermediates, the inhibition observed when 3,5-di-tert-butyl-4hydroxytoluene (BHT) radical inhibitor was added to the 3051

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis visible-light irradiation using a catalytic amount (5 mol %) of bis(4-methoxyphenyl) disulfide as a photocatalyst. Mono-, 1,1di-, and 1,2-disubstituted styrenes are cleaved to the corresponding aldehydes and ketones under 1 atm of O2 at 25 °C. A number of experiments (detection or trapping of proposed intermediates, control and crossover experiments, NMR) and DFT calculations were carried out to prove that the olefin acts as a sensitizer for the homolysis of the disulfide by initial formation of a disulfide-olefin charge-transfer complex. Interaction between the so-generated thiyl radical and the alkene and trapping by molecular oxygen would provide the key dioxetane intermediate (Scheme 5).14 Certain activated olefins,

Scheme 6. Examples and Proposed Mechanism of the Formation of Aldehydes from Alkenes in the Presence of 1 equiv of TPP-OH Photosensitizer

Scheme 5. Disulfide-Catalyzed Oxidative Cleavage of Alkenes Mediated by Visible Light: Reaction Scope and Proposed Mechanism for α-Methylstyrene

thylpiperidine 1-oxyl (TEMPO) catalyzed one-pot oxidative cleavage/nitrogenation in the presence of TMSN3 (1.5 equiv), leading to oxonitriles from internal alkenes and aromatic carbonyls from styrene derivatives (Scheme 7).18 Scheme 7. TEMPO-Catalyzed Aerobic Oxygenation and Nitrogenation of Olefins

such as azulen-1-yl N-sulfonyl enamides, undergo aerobic cleavage to azulen-1-yl ketones on treatment with Cs2CO3 in air and natural sunlight and in the absence of a photosensitizer.15 N-Hydroxyphthalimide (NHPI) is a common oxidant employed in substoichiometric amounts by Tong et al. for the oxidation of cyclohexene and styrene under a 0.3 MPa pressure of O2 and 1 mol % of DADCAQ (1,4-diamino-2,3dichloroanthraquinone). Low selectivities were observed, with mixtures of epoxides, hydroperoxides, and allylic oxidation products being obtained.16 Jiao’s group described the use of NHPI (10 mol %) in dimethylacetamide (DMA) at 1 atm of O2 and 80 °C, a procedure that provided better yields for α,αdisubstituted aromatic alkenes. From a series of experiments (labeling with 18O2, inhibition by radical scavengers, etc.) the authors proposed a mechanism where formation of phthalimide N-oxyl (PINO) radical under aerobic conditions and electrophilic addition to the alkene generate a carbon radical which would be trapped by molecular oxygen. 1,2-Dioxetane formation would release PINO radical, and thermal cleavage of the former would provide target ketones and aldehydes (Scheme 6).17 The same group reported a 2,2,6,6-tetrame-

Although it occurred at the cost of very high pressures of O2 and CO2, Miao et al. reported the cleavage of aromatic alkenes in the presence of the radical initiator tert-butyl nitrite. The use of scCO2 at a suitable pressure ensured a proper selectivity toward carbonyls over mixtures with carboxylic acids (Scheme 8). 19 Another radical initiator, AIBN (2,2- azobis(isobutyronitrile)), was employed in substoichiometric amounts (25 mol %) for the cleavage of gem-disubstituted Scheme 8. Cleavage of Styrenes in the Presence of tert-Butyl Nitrite

3052

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis

co-workers reported the use of commercially available Co(acac)3 for the cleavage of vinyl aromatics to a mixture of aldehyde and arenecarboxylic acids. THF was the solvent of choice to get a higher proportion (∼1:4) of the latter, whereas EtOAc provided the former as the main product.25 The use of ligandless CoCl2 in tBuOH also provided mixtures of oxidation products along with polymers, and only moderate yields and low selectivities were obtained for substrates other than αsubstituted styrenes. Formation of the benzyl radical intermediate E by interaction of the alkene with cobalt superoxo complex D and subsequent dioxetane F formation and thermal cleavage was considered by the authors as the most likely mechanism to explain this reaction (Scheme 11).26

aromatic alkenes to the corresponding ketones, and some examples of the preparation of benzaldehydes from styrene and stilbene derivatives were also described by this method.20 Biocatalysis has been also explored in this field. A cell-free extract or lyophilized cells of the wood-degrading fungus Trametes hirsuta FCC047 can oxidatively cleave several aromatic alkenes in a similar way to reductive ozonolysis. Good to excellent chemoselectivities were obtained at room temperature, and the use of labeled 18O2 with a cyclic alkene such as indene proved the uptake of both oxygen atoms from molecular oxygen (Scheme 9).21 It has been observed that Scheme 9. Biocatalyzed and Catalyst- and Solvent-Free Cleavages of Aromatic Alkenes

Scheme 11. CoCl2-Catalyzed Cleavage of Aromatic Alkenes

some α-substituted styrene derivatives undergo oxidative cleavage even in the absence of any catalyst or solvent at 1 atm of O2. The authors discovered that the reactivity of the alkenes corresponded well with their HOMO energy (obtained by semiempirical molecular orbital calculations at AM1 level) and that radical species are involved in the mechanism. Tri- and tetrasubstituted alkenes provided reaction mixtures with epoxides.22 Garcia and co-workers reported the use of sulfurdoped graphene as a carbocatalyst for the solventless aerobic oxidation of styrene, p-chlorostyrene, and α-methylstyrene. Good selectivities toward benzaldehyde and acetophenone cleavage products were observed at the expense of low conversion rates (90% yield. Nitroethane was found to be an excellent solvent for the reaction, and calculations at the G4 level were performed to check if a homolytic cleavage of RNO2 to provide R• and NO2 took place. On account of the results obtained, the authors conclude that the nitroalkane underwent a C−H bond activation of the acidic hydrogen atom by the polyoxometalate, and reaction with O2 resulted in a cascade of autoxidation processes that released NO2.48 Aromatic gemdisubstituted alkenes can be cleaved to the corresponding ketones (acetophenone and benzophenone derivatives) by means of [Cu(μ-Cl)Cl(phen)]2 catalyst. Slightly better yields were obtained under 4 atm of O2 in a THF−H2O mixture in comparison with other Cu(II) catalyst systems (CuCl2·2H2O, CuCl2·2H2O and phenanthroline). However, monosubstituted styrenes, stilbenes, provided the corresponding aldehydes with moderate to poor yields, and no reaction was observed from aliphatic olefins. From a series of experiments carried out to clarify the reaction pathway (observation of the products from oxidation of styrene and 1,1-diphenylethylene oxides, detection of small amounts of the latter compounds as well as γbutyrolactone in the reaction mixture, and the need for both cosolvents THF/H2O for an effective cleavage, inter alia), the oxidation of THF to the corresponding hydroperoxide is proposed as the key step in the reaction mechanism (Scheme 20).49



OTHER METAL CATALYSTS Hydrolysis of tetraethyl orthosilicate (TEOS) and Cr(NO3)3· 9H2O in ethylene glycol followed by calcination provided a CrSiO2 catalyst which, upon visible photoirradiation, promoted the oxidation of several styrene and stilbene derivatives. Epoxides, alkene isomerization products, Wacker oxidation, and cleavage products (carbonyls) were observed. These results were compared in terms of yield and selectivity with those obtained from other known photosensitizers, such as TiO2 and TPT (2,4,6-triphenylpyrylium tetrafluoroborate), under similar conditions.53 In the last few years two reports have been published on the use of a common oxidant for other oxidation processes, cerium ammonium nitrate (CAN), for aerobic oxidation of aromatic alkenes. K10-montmorillonite clay supported CAN turned out to be a much better catalyst for the cleavage of aromatic alkenes, including chalcones, than overstoichiometric amounts of CAN in methanol. Moreover, this heterogeneous K10montmorillonite clay supported CAN (60 wt % substrate/ catalyst ratio) could be reused four times with a minimal loss of activity. The authors suggest that the constrained environment inside the layers of montmorillonite clay facilitates the formation of an adduct between molecular oxygen and the olefin and its close interaction with CAN, which oxidizes the adduct to generate the dioxetane radical cation K and Ce(III) species. Reduction of K with concomitant oxidation of Ce(III) to Ce(IV) provides an unstable dioxetane derivative similar to the aforementioned F or I, which undergoes rearrangement to yield carbonyl products (Scheme 21).54 Following their research on the oxidation of alkylenecyclobutanes, Yu et al. described the formation of benzaldehydes (along with moderate amounts of the corresponding benzoic acids) from mono- and 1,2-disubstituted aromatic alkenes by using 5 mol % of CAN in dioxane.55 Xie recently reported a Rh−Cu cocatalyzed ortho acylation of N-phenoxyacetamide with internal alkynes as acyl sources. This good-yielding aerobic reaction leading to o-acylphenols implied an oxidative cleavage of the C−C triple bond, although no carboxy or carbonyl derivatives were detected.56 Following our research on the oxidation of alcohols and methylene compounds,57 we envisaged that such oxidation could be also promoted by a combination of a suitably designed triazole ligand and commercially available nickel(II) salt. After optimization of reaction conditions, a number of carbinols,

Scheme 20. Synthesis of Benzophenone in the Presence of [Cu(μ-Cl)Cl(phen)]2

Gold nanoparticles (small 55-atom clusters) supported on boron nitride and SiO2 (0.65 mol %) have been reported to catalyze the oxidative cleavage of styrene to benzaldehyde under 1.5 atm of O2 with moderate conversion rates (up to 21%) and selectivities (up to 82%) over styrene oxide and acetophenone.50 Following their research on the epoxidation of cyclooctene, Li et al. discovered that gold nanoparticles supported on multiwalled carbon nanotubes (Au/CNTs) could also render benzaldehyde (conversion rate 61%, selectivity 91%) from styrene on submission to a pressure of 1 MPa of O2 in the presence of 5 mol % of tert3057

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis Scheme 21. Suggested Mechanism for the Cleavage of Styrene Catalyzed by K10-Montmorillonite Clay Supported Cerium Ammonium Nitrate

Scheme 23. NiBr2/L3-Catalyzed Oxidative Cleavage of Aromatic Alkenes

including 3-phenylprop-2-ynol, were efficiently oxidized. However, the oxidation product from the latter propargyl alcohol (benzoic acid) implied the cleavage of a C−C triple bond. This fact led us to explore this transformation, and to our delight, the optimized procedure (O2 (1 atm), 10−5 mol % of NiBr2, 10−5 mol % of L3, 10 mol % of NaOAc, PEG-400, 120 °C) was a most general and efficient protocol for the oxidative cleavage of internal and terminal arylacetylenes (Scheme 22).

the transformation of α-methylstyrene into acetophenone was carried out on a 1.5 g scale with no significant decrease in the yield. The homogenenous nature of the catalytic system was proved by several kinetic and poisoning experiments, and a possible mechanism was postulated in which hydride transfer from the decomposition of polyethylene glycol under the reaction conditions generates hydride complex L. Formation of hydroperoxide and η-peroxide species M and N by oxygen insertion and subsequent interaction with alkene would provide the five-membered peroxo-nickelacycle O, which upon reductive elimination would render a dioxetane intermediate such as the aforementioned F or I (Scheme 24).59

Scheme 22. Molecular Oxygen Promoted Cleavage of Arylacetylenes in the Presence of Very Low Amounts of NiBr2/L3



SUMMARY AND OUTLOOK Considerable effort has been devoted to the molecular oxygen mediated cleavage of alkenes and alkynes due to the abundance and diversity of such unsaturated raw materials and the Scheme 24. Proposed Pathway for the Aerobic Cleavage of α-Methylstyrene in the Presence of NiBr2 and L3 Ligand

Such infinitesimal amounts of nickel salt afforded oxidation products with nickel levels (1.2 ppm) below the concentration limit for oral and parenteral drugs with a daily dose of 10 g/day. In addition, reactions were conducted in PEG-400, an environmentally friendly solvent that enabled a simple catalyst product separation by simple extraction.58 The excellent results collected from the oxidation of alkynes encouraged us to attempt a similar approach for the aerobic cleavage of aromatic alkenes. Again, a whole array of ligands was assayed along with nickel sources in order to find the optimized conditions for such an alternative synthesis of carboxylic acids and ketones from olefins. The L3 bis-triazolyl ligand along with NiBr2 was once more the elected system to perform the reaction, which turned out to be applicable not only to monosubstituted styrenes but also to 1,1- and 1,2disubstituted and tri- and tetrasubstituted alkenes. TON and TOF values ranged from 6000000 to 10000000 and from 125000 to 208333 h−1, respectively (Scheme 23). Moreover, 3058

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis

2006, 1, 453−458. (i) Xing, D.; Guan, B.; Cai, G.; Fang, Z.; Yang, L.; Shi, Z. Org. Lett. 2006, 8, 693−696. (j) Samanta, S.; Adak, L.; Jana, R.; Mostafa, G.; Tuononen, H. M.; Ranu, B. C.; Goswami, S. Inorg. Chem. 2008, 47, 11062−11070. (k) Hart, R.; Whitehead, D. C.; Travis, B. R.; Borhan, B. Org. Biomol. Chem. 2011, 9, 4741−4744. (l) Shaikh, T. M.; Hong, F.-E. Adv. Synth. Catal. 2011, 353, 1491−1496. (m) Ghosh, K.; Kumar, P.; Goyal, I. Inorg. Chem. Commun. 2012, 24, 81−86. (n) Maurya, M. R.; Kumar, N. J. Mol. Catal. A: Chem. 2014, 383− 384, 172−181. (o) Romanowski, G.; Kira, J.; Wera, M. J. Mol. Catal. A: Chem. 2014, 381, 148−160. (5) (a) Yang, D.; Chen, F.; Dong, Z.-M.; Zhang, D.-W. J. Org. Chem. 2004, 69, 2221−2223. (b) Bailey, S. Chem. Rev. 1958, 58, 925−1010. (c) Bailey, P. S.; Chang, V. S.; Kwie, W. W. L. J. Org. Chem. 1962, 27, 1198−1201. (d) Ando, W.; Miyazaki, H.; Ito, K.; Auchi, D. Tetrahedron Lett. 1982, 23, 555−556. (e) Shing, T. K. M. In Comprehensive Organic Synthesis; Trost, B. M., Flemming, I., Eds.; Pergamon Press: Oxford, U.K., 1991; Vol. 7, p 703. (f) Moriarty, R. M.; Penmasta, R.; Awasthi, A. K.; Prakash, I. J. Org. Chem. 1988, 53, 6124−6125. (g) Banerjee, A.; Hazra, B.; Bhattacharya, A.; Banerjee, S.; Banerjee, G. C.; Sengupta, S. Synthesis 1989, 1989, 765−766. (h) Ishii, Y.; Sakata, Y. J. Org. Chem. 1990, 55, 5545−5547. (i) Ranu, B. C.; Bhadra, S.; Adak, L. Tetrahedron Lett. 2008, 49, 2588−2591. (j) Lee, H.; Kim, Y.-H.; Han, S. B.; Kang, H.; Park, S.; Seo, W. S.; Park, J. T.; Kim, B.; Chang, S. J. Am. Chem. Soc. 2003, 125, 6844−845. (k) Kumar A., V.; Reddy, V. P.; Sridhar, R.; Srinivas, S.; Rao, K. R. Synlett 2009, 2009, 739−742. (l) Whitehead, D. C.; Travis, B. R.; Borhan, B. Tetrahedron Lett. 2006, 47, 3797−3800. (6) For a recent review on hydrogen peroxide oxidation of unsaturated hydrocarbons mediated by H2O2, see: (a) Halina Wójtowicz-Młochowska, H. Arkivoc 2017, 12−58. See also: (b) Goti, A.; Cardona, F. In Green Chemical Reactions; Tundo, P., Ed.; Springer: Dordrecht, The Netherlands, 2006; pp 191−212. (c) Rajabi, F.; Karimi, N.; Saidi, M. R.; Primo, A.; Varma, R. S.; Luque, R. Adv. Synth. Catal. 2012, 354, 1707−1711. (d) Döbler, C.; Mehltretter, G.; Beller, M. Angew. Chem., Int. Ed. 1999, 38, 3026− 3028. (7) See for example: (a) Khan, M. M. T.; Rao, A. P.; Bhatt, S. D. J. Mol. Catal. 1992, 75, 41−51. (b) Nehru, K.; Kim, S. J.; Kim, I. Y.; Seo, M. S.; Kim, Y.; Kim, S.-J.; Kim, J.; Nam, W. Chem. Commun. 2007, 4623−4625. (c) Yoon, J.; Wilson, S. A.; Jang, Y. K.; Seo, M. S.; Nehru, K.; Hedman, B.; Hodgson, K. O.; Bill, E.; Solomon, E. I.; Nam, W. Angew. Chem., Int. Ed. 2009, 48, 1257−60. (d) Lindhorst, A. C.; Haslinger, S.; Kühn, F. E. Chem. Commun. 2015, 51, 17193−17212. (8) Ren, Y.; Che, Y.; Ma, W.; Zhang, X.; Shen, T.; Zhao, J. New J. Chem. 2004, 28, 1464−1469. (9) Hirashima, S.; Kudo, Y.; Nobuta, T.; Tada, N.; Itoh, A. Tetrahedron Lett. 2009, 50, 4328−4330. (10) Yamaguchi, T.; Kudo, Y.; Hirashima, S.; Tada, N.; Miura, T.; Itoh, A. Synlett 2013, 24, 607−610. (11) Fujiya, A.; Kariya, A.; Nobuta, T.; Tada, N.; Miura, T.; Itoh, A. Synlett 2014, 25, 884−888. (12) Murthy, R. S.; Bio, M.; You, Y. Tetrahedron Lett. 2009, 50, 1041−1044. (13) (a) Singh, A. K.; Chawla, R.; Yadav, L. D. S. Tetrahedron Lett. 2015, 56, 653−656. (b) Cleavage of α- and β-methylstyrene and 1,1diphenylethene to the corresponding acetophenone, benzaldehyde, and benzophenone was observed as a side reaction in the photoredoxcatalyzed dimerization of arylalkenes using 9-mesityl-10-methylacridinium perchlorate as organic photosensitizer. See: Wei, D.; Li, Y.; Liang, F. Adv. Synth. Catal. 2016, 358, 3887−3896. (14) Deng, Y.; Wei, X. J.; Wang, H.; Sun, Y.; Noël, T.; Wang, X. Angew. Chem., Int. Ed. 2017, 56, 832−836. (15) Park, S.; Jeon, W. H.; Yong, W. S.; Lee, P. H. Org. Lett. 2015, 17, 5060−5063. (16) Tong, X.; Xu, J.; Miao, H.; Yang, G.; Ma, H.; Zhang, Q. Tetrahedron 2007, 63, 7634−7639. (17) (a) Lin, R.; Chen, F.; Jiao, N. Org. Lett. 2012, 14, 4158−4161. (b) For a recent report on the use of a polymer-anchored NHPI for a similar transformation, see: Łątka, P.; Kasperczyk, K.; Orlińska, B.;

convenience associated with the use of such a sustainable oxidizing agent. A number of different catalysts have been investigated in this field, and in many cases the use of a suitably constructed ligand bound to a metal center has provided better results in terms of efficacy and efficiency. Epoxides, 1,2dicarbonyl compounds, and 1,2-dioxetanes have been proposed as plausible intermediates in the reaction mechanism. However, the substrate scope, which is often restricted to simple styrene with very scarce examples of alkyne oxidation, the need for pressures higher than atmospheric, and relatively low substrate/ catalyst ratios are still present in several reports on this issue. Therefore, it would be highly desirable to develop more general procedures with a broader scope, including most of the aliphatic olefins and of course acetylene derivatives as well. Probably the reduction of catalyst amounts and required pressures will owe a great deal to careful catalyst design.



AUTHOR INFORMATION

Corresponding Authors

*E-mail for R.S.: [email protected]. *E-mail for E.D.: [email protected]. ORCID

Raul SanMartin: 0000-0002-8105-1947 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported by the Basque Government (IT774-13), the Spanish Ministry of Economy and Competitiveness (CTQ2013-46970-P), and the University of the Basque Country (UFI QOSYC 11/12). G.U. thanks the University of the Basque Country (UPV/EHU) for a postdoctoral scholarship. Finally, technical and human support provided by SGIker is gratefully acknowledged.



REFERENCES

(1) (a) Arora, A. Hydrocarbons (Alkanes, Alkenes And Alkynes); Discovery Publishing House: New Delhi, India, 2006. (b) Modern Alkyne Chemistry: Catalytic and Atom-Economic Transformations; Trost, B. M., Li, C.-J., Eds.; Wiley-VCH: Weinheim, Germany, 2014. (c) Yun, H.; Danishefsky, S. J. J. Org. Chem. 2003, 68, 4519−4522. (d) Yun, H.; Chou, T. C.; Dong, H.; Tian, Y.; Li, Y. M.; Danishefsky, S. J. J. Org. Chem. 2005, 70, 10375−10380. (2) (a) Kurata, K.; Inoue, K.; Nishimura, K.; Hoshiya, N.; Kawai, N.; Uenishi, J. Synthesis 2015, 47, 1238−1244. (b) Collins, J.; Drouin, M.; Sun, X.; Rinner, U.; Hudlicky, T. Org. Lett. 2008, 10, 361−364. (c) Labadie, G. R.; Luna, L. E.; Gonzalez-Sierra, M.; Cravero, R. M. Eur. J. Org. Chem. 2003, 2003, 3429−3434. (d) Hilf, J. A.; Holzwarth, M. S.; Rychnovsky, S. D. J. Org. Chem. 2016, 81, 10376−10382. (3) (a) Nakajima, M.; Arai, S.; Nishida, A. Angew. Chem., Int. Ed. 2016, 55, 3473−3476. (b) Gaich, T.; Mulzer, J. Org. Lett. 2010, 12, 272−275. (c) Becheanu, A.; Baro, A.; Laschat, S.; Frey, W. Eur. J. Org. Chem. 2006, 2006, 2215−2225. (d) Zeng, C.; Han, M.; Covey, D. F. J. Org. Chem. 2000, 65, 2264−2266. (e) Ketcham, J. M.; Volchkov, I.; Chen, T. Y.; Blumberg, P. M.; Kedei, N.; Lewin, N. E.; Krische, M. J. J. Am. Chem. Soc. 2016, 138, 13415−13423. (4) (a) Sam, D. J.; Simmons, H. E. J. Am. Chem. Soc. 1972, 94, 4024− 4025. (b) Lee, D. G.; Chang, V. S. J. Org. Chem. 1978, 43, 1532−1536. (c) Sato, K.; Aoki, M.; Noyori, R. Science 1998, 281, 1646−1647. (d) Antonelli, E.; D’Aloisio, R.; Gambaro, M.; Fiorani, T.; Venturello, C. J. Org. Chem. 1998, 63, 7190−7026. (e) Oakley, M. A.; Woodward, S.; Coupland, K.; Parker, D.; Temple-Heald, C. J. Mol. Catal. A: Chem. 1999, 150, 105−111. (f) Yang, D.; Zhang, C. J. Org. Chem. 2001, 66, 4814−4818. (g) Noyori, R.; Aoki, M.; Sato, K. Chem. Commun. 2003, 1977−1986. (h) Che, C.-M.; Yip, W.-P.; Yu, W.-Y. Chem. - Asian J. 3059

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060

Review

ACS Catalysis Drozdek, M.; Skorupska, B.; Witek, E. Catal. Lett. 2016, 146, 1991− 2000. (18) Wang, T.; Jiao, N. J. Am. Chem. Soc. 2013, 135, 11692−11695. (19) Miao, C.-X.; Yu, B.; He, L.-N. Green Chem. 2011, 13, 541−544. (20) Wang, G.-Z.; Li, X.-L.; Dai, J.-J.; Xu, H.-J. J. Org. Chem. 2014, 79, 7220−7225. (21) (a) Mang, H.; Gross, J.; Lara, M.; Goessler, C.; Schoemaker, H. E.; Guebitz, G. M.; Kroutil, W. Angew. Chem., Int. Ed. 2006, 45, 5201− 5203. (b) Lara, M.; Mutti, F. G.; Glueck, S. M.; Kroutil, W. Eur. J. Org. Chem. 2008, 2008, 3668−3672. (c) For a related procedure using tertbutyl hydroperoxide as co-oxidant, see: Hajnal, I.; Faber, K.; Schwab, H.; Hall, M.; Steiner, K. Adv. Synth. Catal. 2015, 357, 3309−3316. (22) Hayashi, Y.; Takeda, M.; Miyamoto, Y.; Shoji, M. Chem. Lett. 2002, 31, 414−415. (23) Dhakshinamoorthy, A.; Latorre-Sanchez, M.; Asiri, A. M.; Primo, A.; Garcia, H. Catal. Commun. 2015, 65, 10−13. (24) Ganeshpure, P. A.; Satish, S. Tetrahedron Lett. 1988, 29, 6629− 6632. (25) (a) Reetz, M. T.; Töllner, K. Tetrahedron Lett. 1995, 36, 9461− 9464. (b) For a recent report on the nonselective formation of carbonyls from isolated alkene derivatives using Co(II)-enolate complexes, see: Alves, T. M. F.; Costa, M. O.; Bispo, B. A. D.; Pedrosa, F. L.; Ferreira, M. A. B. Tetrahedron Lett. 2016, 57, 3334− 3338. (26) Lin, Y. H.; Williams, I. D.; Li, P. Appl. Catal., A 1997, 150, 221− 229. (27) (a) Pathan, S.; Patel, A. Catal. Sci. Technol. 2014, 4, 648−656. (b) Pathan, S.; Patel, A. Chem. Eng. J. 2014, 243, 183−191. (c) Singh, S.; Narkhede, N.; Patel, A. RSC Adv. 2015, 5, 36270−36278. (28) Wang, R.-M.; Hao, C.-J.; He, Y.-F.; Xia, C.-G.; Wang, J.-R.; Wang, Y.-P. J. Appl. Polym. Sci. 2000, 75, 1138−1143. (29) Haber, J.; Kłosowski, M.; Połtowicz, J. J. Mol. Catal. A: Chem. 2003, 201, 167−178. (30) Zhou, X.; Ji, H. Chin. J. Chem. 2012, 30, 2103−2108. (31) (a) Chen, H.; Ji, H.; Zhou, X.; Xu, J.; Wang, L. Catal. Commun. 2009, 10, 828−832. (b) See also: Farokhi, A.; Hosseini-Monfared, H. New J. Chem. 2016, 40, 5032−5043. (32) Varkey, S. P.; Ratnasamy, C.; Ratnasamy, P. J. Mol. Catal. A: Chem. 1998, 135, 295−306. (33) (a) Zsigmond, Á .; Horváth, A.; Notheisz, F. J. Mol. Catal. A: Chem. 2001, 171, 95−102. See also: (b) Wang, R.-M.; Hao, C.-J.; He, Y.-F.; Xia, C.-G.; Wang, J.-R.; Wang, Y.-P. J. Appl. Polym. Sci. 2000, 75, 1138−1143. (34) Baucherel, X.; Uziel, J.; Jugé, S. J. Org. Chem. 2001, 66, 4504− 4510. (35) Zeng, W.; Li, J.; Qin, S. Inorg. Chem. Commun. 2006, 9, 10−12. (36) Kaneda, K.; Haruna, S.; Imanaka, T.; Kawamoto, K. J. Chem. Soc., Chem. Commun. 1990, 1467−1468. (37) Pillai, U. R.; Sahle-Demessie, E.; Namboodiri, V. V.; Varma, R. S. Green Chem. 2002, 4, 495−497. (38) (a) Hong, H.; Hu, L.; Li, M.; Zheng, J.; Sun, X.; Lu, X.; Cao, X.; Lu, J.; Gu, H. Chem. - Eur. J. 2011, 17, 8726−8730. (b) See also: Hadian-Dehkordi, L.; Hosseini-Monfared, H. Green Chem. 2016, 18, 497−507. (39) Rak, M. J.; Lerro, M.; Moores, A. Chem. Commun. 2014, 50, 12482−12485. (40) Wong, W.-K.; Chen, X.-P.; Pan, W.-X.; Guo, J.-P.; Wong, W.-Y. Eur. J. Inorg. Chem. 2002, 2002, 231−237. (41) Gonzalez-de-Castro, A.; Xiao, J. J. Am. Chem. Soc. 2015, 137, 8206−8218. (42) Zhou, Y.; Rao, C.; Mai, S.; Song, Q. J. Org. Chem. 2016, 81, 2027−2034. (43) Wang, J.-Q.; Cai, F.; Wang, E.; He, L.-N. Green Chem. 2007, 9, 882−887. (44) Wang, A.; Jiang, H. J. Org. Chem. 2010, 75, 2321−2326. (45) Gu, L.; Liu, J.; Zhang, H. Chin. J. Chem. 2014, 32, 1267−1270. (46) Feng, B.; Hou, Z.; Wang, X.; Hu, Y.; Li, H.; Qiao, Y. Green Chem. 2009, 11, 1446−1452. (47) Wang, A.; Jiang, H. J. Am. Chem. Soc. 2008, 130, 5030−5031.

(48) Rubinstein, A.; Jiménez-Lozanao, P.; Carbó, J. J.; Poblet, J. M.; Neumann, R. J. Am. Chem. Soc. 2014, 136, 10941−10948. (49) (a) Hossain, Md. M.; Shyu, S.-G. Tetrahedron 2014, 70, 251− 255. (b) See also: Agasti, S.; Dey, A.; Maiti, D. Chem. Commun. 2016, 52, 12191−12194. (50) Turner, M.; Golovko, V. B.; Vaughan, O. P. H.; Abdulkin, P.; Berenguer-Murcia, A.; Tikhov, M. S.; Johnson, B. F. G.; Lambert, R. M. Nature 2008, 454, 981−983. (51) Li, B.; He, P.; Yi, G.; Lin, H.; Yuan, Y. Catal. Lett. 2009, 133, 33−40. (52) Hu, L.; Cao, X.; Yang, J.; Li, M.; Hong, H.; Xu, Q.; Ge, J.; Wang, L.; Lu, J.; Chen, L.; Gu, H. Chem. Commun. 2011, 47, 1303−1305. (53) Shiraishi, Y.; Teshima, Y.; Hirai, T. J. Phys. Chem. B 2006, 110, 6257−6263. (54) Dhakshinamoorthy, A.; Pitchumani, K. Catal. Commun. 2009, 10, 872−878. (55) (a) Yu, L.; Huang, Y.; Bai, Z.; Zhu, B.; Ding, K.; Chen, T.; Ding, Y.; Wang, Y. J. Chin. Chem. Soc. 2015, 62, 479−482. See also: (b) Yu, L.; Ren, L.-F.; Yi, R.; Guo, R. Synth. Commun. 2011, 41, 2530−2538. (56) Xie, Y. Chem. Commun. 2016, 52, 12372−12375. (57) (a) Urgoitia, G.; SanMartin, R.; Herrero, M. T.; Domínguez, E. Green Chem. 2011, 13, 2161−2166. (b) Urgoitia, G.; Maiztegi, A.; SanMartin, R.; Herrero, M. T.; Domínguez, E. RSC Adv. 2015, 5, 103210−103217. (58) Urgoitia, G.; SanMartin, R.; Herrero, M. T.; Domínguez, E. Chem. Commun. 2015, 51, 4799−4802. (59) Urgoitia, G.; SanMartin, R.; Herrero, M. T.; Domínguez, E. Adv. Synth. Catal. 2016, 358, 1150−1156.

3060

DOI: 10.1021/acscatal.6b03654 ACS Catal. 2017, 7, 3050−3060