Next Article in Journal
Highly Selective Photocatalytic Reduction of o-Dinitrobenzene to o-Phenylenediamine over Non-Metal-Doped TiO2 under Simulated Solar Light Irradiation
Previous Article in Journal
Catalytic Oxidative/Extractive Desulfurization of Model Oil using Transition Metal Substituted Phosphomolybdates-Based Ionic Liquids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Homogeneous Metal-Catalyzed Aerobic C–H Oxidation of Benzylic Compounds

Department of Organic Chemistry II, Faculty of Science and Technology, University of the Basque Country, 48940 Leioa, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(12), 640; https://doi.org/10.3390/catal8120640
Submission received: 9 November 2018 / Revised: 4 December 2018 / Accepted: 5 December 2018 / Published: 8 December 2018

Abstract

:
Csp3–H oxidation of benzylic methylene compounds is an established strategy for the synthesis of aromatic ketones, esters, and amides. The need for more sustainable oxidizers has encouraged researchers to explore the use of molecular oxygen. In particular, homogeneous metal-catalyzed aerobic oxidation of benzylic methylenes has attracted much attention. This account summarizes the development of this oxidative strategy in the last two decades, examining key factors such as reaction yields, substrate:catalyst ratio, substrate scope, selectivity over other oxidation byproducts, and reaction conditions including solvents and temperature. Finally, several mechanistic proposals to explain the observed results will be discussed.

Graphical Abstract

1. Introduction

Aryl ketones, benzoate esters, and benzamides are ubiquitous structures found in natural and biologically active compounds (e.g., hypericin [1,2,3], acetosyringone [4,5], oxcarbazepine [6,7], daidzein [8,9], sitaxentan [10,11], radicicol [12,13], arnamial [14,15,16], epicatechingallate [17], Rucaparib [18,19,20,21], and Leucine-rich repeat kinase-2 (LRRK2) [22,23], Figure 1). In addition, such carbonyl compounds are also extensively used in advanced materials [24,25,26,27,28] and as key intermediates for the synthesis of fine chemicals, including structurally complex compounds. [29,30,31]. Among the many synthetic procedures devised for the preparation of aromatic ketones, esters and amides [32,33,34,35,36,37], including Friedel–Crafts acylation, esterification and amidation of benzoic acid derivatives [38,39,40,41,42,43,44,45,46], the oxidation of benzylic methylene compounds stands out as a simple approach that has attracted much attention in the last decades. Many reagents (KMnO4, iodine(V) species, tBuOOH, m-CPBA, H2O2, NHPI, inter alia) can oxidize the activated benzylic position of aromatic alkanes, benzyl ethers, and amines to provide the corresponding arylcarbonyl moiety (Scheme 1) by a process opposite to Clemmensen, Wolf–Kisnerr and related reduction reactions [47,48,49,50,51,52,53,54,55,56,57,58,59,60,61]. Although good yields and selectivities have been obtained in most cases, stoichiometric amounts of relatively harmful or unsafe oxidizing agents are required. For safety and environmental reasons, molecular oxygen is the oxidant of choice in most oxidative transformations, and an increasing number of reports concerning oxidation of arylmethylene compounds under aerobic conditions can be found in the literature [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78]. Interestingly, although a theoretically easier catalyst–product separation has boosted research on heterogeneous catalysts (e.g., polymer- or montmorillonite clay-supported catalysts, Pd, Au, or Co nanoparticles) [79,80,81,82,83,84,85,86,87,88,89,90,91,92,93], leaching problems and efficiency or selectivity issues [79,85,86,90,93] have encouraged investigations on the development of new homogeneous catalyst systems. This review will cover recent literature from 2000 on the use of homogenous metal catalysts for the aerobic benzylic methylene C–H oxidation leading to aromatic ketones, esters, and amides, with a focus on relevant parameters such as reaction yields, substrate:catalyst ratio, selectivity towards target carbonyl compounds, reaction media, and other experimental conditions. Considering that metal complexes and combination of metal salts and ligands have been described as catalysts or precatalysts in this field, this review is organized according to the metal core involved in the reaction. Although mechanism for this transformation is far from being completely elucidated, several proposals will be discussed.

2. Cobalt Catalysts

2.1. Co(II)-NHPI

Following the pioneering works by Ishii [63,64,65,66], cobalt(II) salts combined with N-hydroxyphtalimide (NHPI) derivatives have been amply used for the aerobic C–H oxidation of methylene compounds at atmospheric pressure. An in situ-generated phthalimide-N-oxyl radical (PINO) is proposed to play a key catalytic role as hydrogen atom abstracting species in the reaction, and other N-hydroxylated ligand/additives prone to form related radical species have also been evaluated and compared with NHPI. Among other transition metals, cobalt salts have been defined as promoters for the formation of PINO. Nolte and coworkers published the use of 0.5 mol% of Co(OAc)2·4H2O combined with NHPI (10 mol%) in acetic acid to perform benzylic oxidation of 2-(hetero)arylacetic esters to the corresponding α-ketoesters at relatively low temperatures (40–80 °C). Lower yields (<30%) were obtained from some o-alkoxylated and brominated substrates, as well as from all the heteroaromatic esters. Finally, with regard to the oxidation of ethylbenzene to acetophenone, several N-hydroxy-phthalimides were evaluated as co-catalysts, achieving good conversion rates (except for nitro-substituted NHPIs) and relatively good acetophenone:1-phenylethanol ratios (59–88:2–5) [94].
The solvent-free oxidation of several alkylbenzenes to the corresponding ketones using di-n-decyl-di-methyl ammonium bromide (DDAB) and other phase transfer agents was explored by Patil et al. A 0.5 mol% of the latter quaternary ammonium salt was required along with NHPI (3 mol%) and CoCl2·6H2O (1 mol%) to perform the oxidation under oxygen gas flow rate of 110 mL/min at 65 °C, although other metal sources (MnCl2(H2O)4 and CrCl3(H2O)6) and ammonium phase transfer catalysts (TBAB and TOAB; CTAB and Aliquat 336) were also tested and provided poorer results in most cases (CrCl3(H2O)6 performed better with tetralin). Excellent selectivity was observed for ethylbenzene and indane, but other oxidation products (17–20%) were detected from tetralin (Scheme 2). Catalyst–product separation by vacuum distillation is described along with upscaling of the reaction leading to acetophenone [95].
In addition to a new electrochemical NHPI-catalyzed oxygenation of methylene compounds, Stahl and coworkers reported the Co/NHPI co-catalyzed benzylic C–H oxidation of a number of alkyl(hetero)arenes, including several di(hetero)arylmethanes precursors of pharmacologically active compounds. After observing selectivity problems (side-products from overoxidation or oxidative cleavage of the alkyl chain, among others), a careful optimization of the reaction conditions led to a procedure involving a greener solvent, n-butyl acetate (Scheme 3). In some cases pyridine was added as co-solvent (7:3 proportion) in order to minimize the above side-reactions. They also showed that the presence of certain heterocyclic moieties could inhibit the radical oxygenation reaction, probably by chelation with the metal catalyst. In those cases the aforementioned electrochemical procedure can be used as a valid alternative. (Benzoimidazol-2-yl)(4-fluorophenyl)methanone, precursor to a clinical candidate for schizophrenia, was prepared in 2 g scale by a variation of the procedure that involved the use of 500 psi of a mixture (O2:N2 9:91) in order to maintain the concentration of molecular oxygen below the limiting oxygen concentration of EtOAc (Scheme 3) [96].
Wang et al. prepared an analog to NHPI (Py-NHPI, Figure 2) and used it for the oxidation of N-benzyl-acetamides, -propionamides, and -benzamides to the corresponding imides as well as for the synthesis of several aromatic ketones and esters by oxidation of isochomane and alkylarenes. Under 1 atm of molecular oxygen, Co(PF6)2 (1 mol%) and Py-NHPI (5 mol%) catalyzed the above oxidations at 65 °C using for the first time a ionic liquid as solvent ([bmim][PF6]) [97]. A Dields–Alder cycloaddition was the key for the preparation of several chiral anthrone-derived NHPI derivatives performed by Shen and Tan. Ethylbenzene, 1,2-dihydroacenaphthylene, and 2-methoxy-2-phenyl-2,3-dihydroindene were reacted with molecular oxygen (1 atm) in the presence of Co(OAc)2 (5 mol%) and one of the above chiral anthrone-derived N-hydroxyphthalimides (10 mol%) in acetonitrile at 60 °C. Mixtures of monohydroxylated product along with the corresponding ketone were obtained for the first two substrates, and 2-methoxy-2-phenyl-2,3-dihydroinden-1-one was isolated in a 48% yield (Scheme 4). Enantiomeric excesses for chiral products ranged from 4 to 13% [98].
As an alternative to NHPI-based catalysts, Sheldon and coworkers reported the use of NHS (N-hydroxysaccharin) (10 mol%) for the oxidation of ethylbenzene. A 0.5 mol% of Co(OAc)2 or Co(acac)2 were tested as co-catalysts in four different solvents (AcOH, CH3CN, PhCF3, and BuOAc) at 100 °C and 85 °C, and the results compared with those obtained with 10 mol% of NHPI. Higher conversion rates and selectivities were observed in the NHPI-co-catalyzed reactions, and only when benzoic acid (5 mol%) was added to the NHS-catalyzed reaction performed in BuOAc, the conversion rate and selectivity towards acetophenone improved to 71% and 90%, respectively [99]. Among other oxidations of the alkyl group of alkylacetophenones, the group of Ishii reported the molecular oxygen-mediated oxidation of 4-ethylacetophenone by using Co(OAc)2 (0.5 mol%) and a 5 mol% of N,N′,N″-trihydroxyisocyanuric acid (THICA) in acetic acid. Although 1,1′-(1,4-phenylene)bis(ethan-1-one) was obtained by this method, a higher yield was observed when N-hydroxy imide THICA was replaced with NHPI (10 mol%) as catalyst for the generation of carbon radicals from the methylene sp3 C–H bonds (Scheme 5) [100].

2.2. Co(II) without Hydroxyimides

Shaabani et al. carried out the oxidation of several benzylic methylene compounds to the corresponding ketones employing cobalt(II) phthalocyanine (5 mol%) in a ionic liquid ([bmim]Br) at 100 °C under oxygen at one atmospheric pressure. Good yields were obtained, and recycling of the catalyst–solvent system was also reported, as after several cycles no decrease in the catalytic activity was observed. The authors propose a radical pathway which starts with hydrogen abstraction by superoxocobalt(III) species [101]. Activation of molecular oxygen by metalloporphyrins toward hydrocarbon oxidation was proposed by Wang et al. as the key for the Co(II) tetraphenylporphyrin (CoTPP)-catalyzed solventless oxidation of alkylaromatics under the special temperature control method (reaction initiation at high temperatures and progress at lower temperatures). Depending on the substrate, the reaction mixture containing 10−3 mol% of CoTPP and 1 atm of O2 was initially heated to 136–200 °C for 1 min, and then heating was continued at 130 °C for 12 h. Moderate yields (<50%) and selectivities for target ketones were observed in most cases [102].
A Co(II)/Mn(II) co-catalyzed aerobic oxidation of ethylbenzene was performed in a gas–liquid continuous flow reactor. Different reaction parameters (flow rates, temperature, reaction time, etc.) were optimized (CoBr2 (2.5 mol%), Mn(OAc)2 (2.5 mol%) in AcOH, 0.6 mL/min for the liquid stream and 100 mL/min for the synthetic air flow, 12 atm, 100 °C, ~20 min) in order to get an 84% of acetophenone mixed with small proportions of benzoic acid (5%), phenylethanol (2%), benzaldehyde (1%), 1-phenylethyl acetate (2%), and 2-bromoacetophenone (5%) by a comparatively fast oxidative transformation. Higher temperatures (150 °C) provoked oxidative cleavage of ethylbenzene to benzoic acid [103]. It should be pointed out that the reaction conditions were close to those for the industrial oxidation of p-xylene to terephthalic acid.
Medium-pressure oxidation of ethyl benzene and other alkylbenzenes such as n-propylbenzene, tetraline, 1,2-dihydroacenaphthylene, and fluorene is described in the report by Wang and He. The authors propose that Co(OAc)2 acts not only as an initiator for generating polyethylene glycol (PEG) radicals but also as a promoter of the decomposition of intermediate hydroperoxides into the product, in a similar way to the radical reactions performed in the presence of the couple NHPI/Co(OAc)2. High-pressure CO2 is introduced in the system in order to “expand” PEG so that lowered melting points, raised diffusion rates and lowered solvent viscosity are achieved. Although oxidation of ethylbenzene to acetophenone performed at 120 °C (2 mol% of Co(OAc)2, 36 mol% of PEG-1000, 24.7 atm O2, 69 atm CO2, 12 h) provided target ketone with a 65% yield, a lower temperature (100 °C) was chosen in order to explore the generality of the procedure. Moderate to good yields ranging from 13 to 83% were obtained [104]. Another polyol (ethylene glycol) is the solvent of the procedure described by Ji and coworkers. A number of 4-benzyl- and 4-ethylphenols were oxidized to the corresponding 4-hydroxybenzophenones and 4-hydroxyacetophenones by the basic conditions displayed in Scheme 6. On the basis of the comparison of the yields obtained from 4-benzyl- and 4-ethyl derivatives and of the isolation and/or reactivity of presumed intermediates, the authors propose a mechanistic path initiated by a phenoxy radical formed by single electron transfer between the phenoxide and dioxygen-induced Co(III) species. Disproportionation or hydrogen abstraction leads to a benzoquinone methide which undergoes conjugate addition by ethylene glycol, and after formation of the corresponding phenoxy radical and subsequent benzoquinone methide, hydrolysis provides target ketone (Scheme 6) [105].

3. Copper Catalysts

3.1. Cu-NHPI

In comparison with the Co(II)-NHPI system, a notable smaller number of reports on molecular oxygen-mediated benzylic C–H oxidation in the presence of copper-imide co-catalysts can be found in the literature. Following previous research by Orlinska and Romanowska on aerobic oxidation of cumene derivatives [106], Chen et al. published the medium pressure (11.8 atm of O2) oxidation of tetralin using a 10 mol% of NHPI and 2.5 mol% of CuCl2 in CH3CN. 1-Tetralone was produced with a selectivity of 46% over other undetermined oxidation products. An in situ generation of PINO radical from NHPI through the redox reaction of copper salts is suggested by the authors as the key step of the reaction mechanism [107]. Several chiral and achiral N-hydroxyimides were evaluated as co-catalysts for the aerobic oxidation of indane to 1-indanone published by Einhorn and coworkers. N-hydroxy-3,4,5,6-tetraphenylphthalimide (NHTPPI) resulted very convenient for the latter transformation performed with O2 at atmospheric pressure at 35 °C in MeCN. The reaction scope was then extended to other alkylbenzenes bearing benzylic methylene groups, providing the results displayed in Scheme 7. In addition to the mild conditions employed, the relatively high conversion rates (69–98%) and the high selectivities observed (1–3% of the corresponding partially alcohol derivatives were detected), the procedure was also efficient, as only a 1 mol% of NHTPPI and 5 mol% of CuCl were required. Such enhanced catalytic activity may be due to the higher stability of the NHTPPI-derived radical [108].

3.2. Additive-Free Procedures

A simple catalyst system comprising CuCl2·2H2O (10 mol%) was enough to catalyze the oxidation at atmospheric pressure of the benzylic methylene unit of 4-ethylpyridine, 4-benzylpyridine and 2-benzylpyridine to the corresponding ketones. DMSO was the best solvent for the presented procedure, as model reaction in other reaction media provided negligible results. In contrast with 4-ethylpyridine, good conversion rates and selectivity towards ketones was observed for the latter benzyl derivatives [109]. The additive- and ligand-free procedure described by Ji and coworkers provides m,m′-disubstituted p-hydroxyaryl ketones from m,m′-disubstituted p-alkylphenols by treatment with 1 atm of air in the presence of 1 mol% of Cu(OAc)2 in ethylene glycol at 50 °C. In order to explain this process, a mechanistic path akin to the one proposed for the cobalt-catalyzed oxidation under similar conditions [105] was postulated by the authors [110]. A rather specific benzylic oxidation was observed by Chiba and coworkers during the addition of Grignard reagents to o-cyanodiarylmethanes. The authors described the aerobic oxidation of the benzylic methylene unit to the corresponding ketone by using Cu(OAc)2 (10 mol%) in DMF, and suggested that complexation of the copper catalyst to the so-formed o-carboximino group enables an anchimeric assistance leading to o-acylbenzophenone derivatives [111]. A copper-catalyzed and carbonate base-accelerated oxidation of deoxybenzoins and oxindoles to benzils and isatins was reported by Yung et al. Dimethylformamide was found to be the best solvent for the reaction, which involves the use of the conditions displayed in Scheme 8. On the basis of a radical trapping experiment with 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) the authors propose the participation of a benzyl radical, that after trapping with oxygen and hydrogen abstraction undergoes dehydration to provide target diketones/ketoamides [112].

3.3. With Ligand and/or Additives

Following their initial research on the oxidation of the methylene group of aryl(di)azinylmethanes [113], an extensive mechanistic study designed to understand the parameters governing the copper-catalyzed aerobic oxidation of a number of benzylpyridines and other benzylheterocyles (2-benzylbenzooxazoles, -benzothiazoles, and -imidazoles, inter alia) was performed by Maes and coworkers. As a result, a direct correlation between calculated imine–enamine tautomerization equilibrium constants of the substrates (DFT calculations) and the yields of the obtained ketones was found, so that Keq values serve as tool to predict if a substrate can be oxidized by the optimized procedure (O2 (1 atm), CuI (10 mol%) AcOH (Equation (1)), DMSO, 100 °C). In addition, it was also demonstrated that benzyl alcohol is not an intermediate but a side-product, and that mononuclear and less active dinuclear catalytic species are involved in a relatively complex catalytic cycle (Scheme 9) [114].
The benefits of introducing other activating groups for the oxidation of aliphatic C–H bonds of N-heterocyclic compounds with molecular oxygen to produce N-heterocyclic ketones were explored by Liu et al. They found that ethyl chloroacetate (Equation (1)) combined with O2 (1 atm), CuCl2·2H2O (10 mol%) in DMF at 130 °C could effectively promote the selective oxidation of a number of N-heterocyclic methylene compounds, including pyridines, quinolines, benzoquinolines, quinoxalines, benzimidazoles, and benzothiazoles. A number of experiments (isotopic labeling, kinetic, byproduct isolation, electron paramagnetic resonance, etc.) led them to propose that N-alkylation of heterocyclic nitrogen with ethyl chloroacetate activates the benzylic methylene unit for a catalytic hydrogen abstraction, and the resulting benzyl radical interacts with dioxygen to generate a hydroperoxide, that upon dehydration and copper induced reduction, provides target ketone along with ethyl acetate [115].
An efficient oxidation of ethylbenzene, tetraline, and indane was reported under medium pressure by Murahashi and coworkers. Acetaldehyde (2.5 mol%) was added to the reaction mixture containing the benzylic methylene compound and Cu(OAc)2 (6 × 10−4 mol%) in a mixture of acetonitrile and dichloromethane under O2 (1 atm)/N2 (8 atm). The in situ formation of peracetic acid was proposed by the authors to explain the oxidation results. However, selectivity towards the corresponding ketones was poor, as mixtures with the partially oxidized benzylic alcohol were obtained in all cases [116]. The increase in oxidizing ability of quinones upon irradiation with visible light was exploited by Finney et al. in order to perform the benzylic C–H oxidation of a number of isochromanes, phthalans and other benzyl ethers, benzyltosylamines and diarylmethanes. 1,4-Hydroquinone, readily available from the naturally occurring β-d-glucopyranoside Arbutin was employed (5 mol%) as precursor of benzoquinone and thus as a pre-oxidant, along with CuCl2·2H2O (2 mol%) as an electron transfer mediator and molecular oxygen as the thermal oxidant in dimethyl carbonate at 25–30 °C. Although minor selectivity issues were encountered (e.g., oxidation of 5-methoxyphthalan or of 1-methoxy-4-(methoxymethyl)benzene), it should be pointed out the access to benzamides by this more sustainable method (Scheme 10) [117].
Goggiamani and coworkers found that the previously reported DABCO/air oxidation of deoxybenzoins failed when applied to some of their diarylethanones. Accordingly, they devised a more reliable procedure for the copper-catalyzed aerobic oxidation of deoxybenzoins that provided the corresponding benzils. Indeed, moderate to good yields (45–95%) were obtained for a number of benzils bearing methyl, methoxy, cyano, acetyl, methoxycarbonyl, chloro, bromo, and iodo functionalities using Cu(OAc)2 (15 mol%) and PPh3 (30 mol%) in 1,2,4-trimethylbenzene at 100 °C under air (1 atm) [118]. Structurally related α,β-diketotriazoles were selectively prepared by aerobic oxidation of α-ketotriazoles, with the 1-alkyl-1,2,3-triazol-4-yl moiety acting as an intramolecular assisting group. The 1-Alkyl group (benzyl group in most cases) remained untouched under the reaction conditions (Scheme 11). The authors propose active participation of both the substrate carbonyl oxygen and triazolyl N3-nitrogen in the reaction mechanism as copper(II) chelating positions that favor interaction with dioxygen [119]. Navarro’s group reported the evaluation of two discrete Cu(II) imidazolyl complexes as catalysts for the oxidation of tetralin. Under the optimized reaction conditions (air (35 atm), Cu complex (0.14 mol%), 100 °C), α-tetralone was obtained with a moderate selectivity over α-tetralol (39–43:13) with a moderate conversion yield (52–57%) [120].

4. Iron Catalysts

4.1. Fe-NHPI

Up to 15 benzylic methylene compounds were oxidized by the system Fe(NO3)3·9H2O (8 mol%)/NHPI (10 mol%) in acetonitrile under 1 atm of O2 at 25 °C. Sun and coworkers performed a series of experiments in order to get information on the reaction mechanism. As a result, they concluded that substrates bearing electron-donating groups have a higher oxidation rate, and proposed a catalytic cycle involving hydrogen abstraction by PINO to generate a benzylic radical followed by interaction with dioxygen and nitrate anions [121].

4.2. Light-Induced Oxidations

Several photochemical approaches to the iron-catalyzed benzylic C–H oxidation can be found in the recent literature. Thus, Pasau and coworkers described the flow synthesis of benzylic ketones using a combination of riboflavin tetraacetate (10 mol%) as the photocatalyst, Fe(ClO4)2 (5 mol%) as an additive to avoid photocatalyst bleaching, a flow of molecular oxygen of 1.0 mL/min and 106 W UV lamp for the mesofluidic C–H oxidation process of benzylic methylene compounds. Moderate to good conversions were obtained for substrates bearing amide, cyano, acetyl, carboxyl and halogen groups. When a 24 W 240-nm blue LED light was employed conversion dropped to 32%. The reaction mechanism is supposed to the initiated by a benzylic hydrogen abstraction with the photoactivated riboflavin tetraacetate catalyst aided the iron salt. The so-formed benzyl radical upon incorporation of singlet oxygen provides target ketones [122]. FeCl3·6H2O (10 mol%) is the additive chosen by Tang et al. to perform the photocatalytic activation of C–H bonds using riboflavin tetraacetate (10 mol%) and N-hydroxysuccinimide (20 mol%) organophotocatalysts and a 90 W 450-nm LED light at room temperature. Tetralin, n-butylbenzene, n-propylbenzene, and three ethylbenzene derivatives were oxidized in moderate to good yields by this method [123].
A photochemical aerobic oxidation of a number of alkylbenzene derivatives was performed in the presence of FeCl3 (5.0 mol%), LiBr (5.0 mol%) in MeCN at 25 °C. Irradiation with 3 W blue LED light under 1 atm of air induced selective methylene C–H oxidation to the corresponding ketones (Scheme 12). Good yields were obtained in most cases, although the presence of nitro or ciano electron-withdrawing groups in the arene moiety resulted in a decrease a yield (34–55%). The authors postulated that an excited iron complex related to Li(FeCl3Br) driven by visible light oxidize benzylic methylene compounds through a single-electron transfer process [124].

4.3. Other Approaches

After an initial report on the oxidation of aryl(di)azinylmethanes using the FeCl2·4H2O/AcOH/DMSO system [113], Maes’ group reported a further investigation on extension of the substrate scope (including benzoannulated azines and diazines, papaverine, or the total synthesis of antimalarial agent Mefloquine), chemoselectivity, alternative solvents, and trace metal analysis. In addition, an extensive comparison with alternative copper catalysts discovered by the same group [114] is made, revealing that oxidation for 2-benzylpyridine derivatives under the optimized conditions (O2 (1 atm), FeCl2·4H2O (10 mol%), AcOH (Equation (1)), DMSO, 100–130 °C) provided comparable and in some cases superior results [125]. A high-temperature/pressure gas–liquid continuous flow approach to 2-aroylpyridines was described by Pieber and Kappe. As shown in Scheme 13, several 2-benzylpyridines were quickly (13 min) oxidized in propylene carbonate, a high-boiling point solvent alternative to traditional aprotic polar solvents such as dimethyl sulfoxide or N-methylpyrrolidone. In addition, this procedure avoids the need of stoichiometric amounts of additives such as AcOH [126].
Several alkylbenzenes were oxidized to the corresponding ketones at room temperature by using substoichiometric amounts of an additional oxidant along with molecular oxygen. Li and Wang reported a system comprising O2 (1 atm), chloramine-T (1 mmol) and Fe(TPP)Cl (5 mol%) in acetonitrile (TPP meso-tetrakisphenyl porphyrin). Good to excellent yields were obtained in all cases with excellent selectivity [127]. Feringa’s group published the results from the oxidation of ethylbenzene under 1 atm of molecular oxygen using a 0.1 mol% of different non-heme iron complexes. Initial temperature of 30 °C was increased to 80 °C in order to enhance the catalytic activity. However, the reactions showed a poor selectivity towards acetophenone over 1-phenylethanol [128].

5. Palladium Catalysts

In addition to their investigation with copper catalysts (see Section 3.3), Navarro and coworkers evaluated the catalytic activity of a palladium bis-imidazolyl complex (0.14 mol%) for the medium-pressure (35 atm) aerobic oxidation of tetralin at 100 °C. A low (30%) conversion rate was observed for a mixture (22:8) of α-tetralone and α-tetralol respectively [119]. Two palladium NCN and CNC tricoordinated complexes were tested as catalysts for the benzylic C–H oxidation of several alkylbenzenes under 1 atm of molecular oxygen. Good to excellent yields were obtained in all cases with complete selectivity towards the corresponding ketones, and the procedure allowed the use of unusually low catalyst loadings (10−2 mol%) [129]. Later the same group published similar results with a more simple palladium system comprising a combination of Pd(OAc)2 and a bis-triazolyl ligand (Scheme 14). A further improvement in efficiency had been achieved, as catalyst amounts as low as 10−5 mol% were enough to catalyze the reaction in a environmentally friendly solvent, polyethylene glycol [130].
A base-mediated (Cs2CO3, Equation (1)) oxidation of 2-benzylbenzothiazole and 2-benzylbenzimidazole derivatives under air atmosphere was presented by Dos Santos et al. and the results were compared with those from the oxidation using palladium catalysis (Pd(OAc)2 (10 mol%) or PdCl2(CH3CN)2 (10 mol%), CsOPiv (Equation (1)), DMF, 90 °C)). Both methods turned out to be complementary, as in some cases, depending on the substrate, only one of them provided the target 2-aroylbenzothiazole or 2-aroylbenzoimidazole with acceptable yields [131].

6. Other Metal Catalysts

6.1. Mg/V-NHPI

Xu’s group explored the use of alkali-earth metal salts as promoters for the formation of PINO radical, key intermediate for NHPI-catalyzed oxidations. After evaluating the results from different alkali-earth metal chlorides, they found that MgCl2 gave better conversion rates and selectivities in the oxidation of ethylbenzene. Other Mg(II) salts were also tested, providing inferior results. In addition to ethylbenzene, fluorene, tetralin, and diphenylmethane were submitted to the optimized reaction conditions (O2 (3 atm), MgCl2 (2.5 mol%), NHPI (10 mol%), CH3CN, 90 °C)) to afford the corresponding ketones with moderate to good conversion rates (47–93%) and selectivities (75–99%) over other oxidation products such as benzyl alcohol and hydroperoxide derivatives [132]. A lower pressure of molecular oxygen (1 atm) was employed by Figiel and Sobczak to perform the NHPI/VO(acac)2-co-catalyzed oxidation of ethylbenzene. Depending on the chloride additive (LiCl or Bu4NCl) two procedures were devised and applied to several alkanes, including ethylbenzene. As shown in Scheme 15, a better conversion rate and selectivity was observed for the latter Bu4NCl-based procedure [133].

6.2. Other Approaches

A manganese(II) catalyst (MnCO3) proved to be active for the medium-pressure (PO2: 10 atm) oxidation of ethylbenzene. When the latter alkylbenzene was submitted to the optimized conditions (MnCO3 (0.5 mol%), 190 °C) a mixture of acetophenone (26%) and 1-phenylethanol (7%) was detected and quantified by gas chromatography-mass spectroscopy (GC-MS). Gao et al. propose that high-valence Mn species, formed from MnCO3 in the presence of oxygen, can activate aromatic hydrocarbons to the corresponding radical intermediates [134]. A more efficient solvent-free oxidation of ethylbenzene, tetralin, and diphenylmethane was carried out at lower temperatures (120 °C) in the presence of O2 (1 atm), [Cp*IrCl2]2 (1.6 × 10−3 mol%, Cp* = pentamethylcyclopentadienyl), picolinic acid (3 × 10−3 mol%) as nitrogen ligand and iodobenzene (1.9 mol%) as radical initiator. Although good to excelent TON values are reported, moderate selectivity was achieved from ethylbenzene (73:27 acetophenone/1-phenylethanol) and tetralin (61:39 α-tetralone/α-tetralol). A radical oxidation mechanism initiated by decomposition of iodobenzene under the reaction conditions to generate phenyl radical is postulated by the authors [135].
Following their research on palladium-catalyzed aerobic oxidations [129,130], Urgoitia et al. found that a combination of NiBr2 (10−5 mol%), triazole ligands (10−5 mol%), and sodium acetate (10−1 mol%) could promote the selective oxidation of seven alkylarenes (ethylbenzene, diphenylmethane, 2-phenylacetonitrile, 4-benzylpyridine, fluorene, 9,10-dihydroanthracene, and xanthene) to the corresponding ketones at 120 °C under 1 atm of molecular oxygen in polyethylene glycol. Yields ranged from 50 to 98% with reaction times of 48–72 h [136].
A ruthenium catalyst system was devised by Kojima, Oisaki and Kanai for the selective photooxidation of fluorene derivatives at room temperature in acetonitrile. As shown in Scheme 16, Cu(MeCN)4OTf was employed as a metal additive, and this copper additive along with tris(bipyridyl)ruthenium(II) dichloride and irradiation with visible white LED light were essential to get the target transformation. A high compatibility with a number of functional groups was observed, and the authors discarded any autooxidation process after examining the results from several experiments performed under limited irradiation time, then in the dark [137].

7. Conclusions

As in many other fields of synthetic chemistry, the molecular oxygen-mediated C–H oxidation of benzylic methylene compounds has witnessed a clear evolution towards more sustainable procedures. Starting from the need of lower pressures of the terminal oxidant or reagent itself (O2, air in some cases), more active and selective homogeneous metal catalysts have been discovered, thus rivaling other strategies based on heterogeneous or metal-free approaches. Indeed, although chemoselectivity under oxidation conditions and product selectivity towards the carbonyl compound (ketone, ester or amide, depending on the starting material) are still issues to be addressed, a substantial decrease in the catalyst amount can be easily noticed by comparison with earlier literature data. In addition to procedures with a wider substrate scope, some more specific methods applicable to certain compounds (e.g., benzylazine derivatives) have been devised. Moreover, new protocols involving visible light induction and alternative, greener solvents have been introduced. Some of the above methods allow reaction performance under mild conditions, or by continuous flow technology. Considering also that a number of synthetic intermediates for pharmacologically active compounds have been prepared by this aerobic strategy, and although research is needed to improve selectivity and efficiency in some reactions, it can be concluded that the use of homogeneous metal catalyst for this benzylic C–H oxidation has reached a mature state.

Funding

(IT-774-13 (Basque Government) and CTQ2017-86630-P (Spanish Ministry of Economy and Competitiveness) projects.

Acknowledgments

This research was supported by the Basque Government (IT-774-13) and the Spanish Ministry of Economy and Competitiveness (CTQ2017-86630-P). Finally, technical and human support provided by SGIker is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. van de Putte, M.; Roskams, T.; Vandenheede, J.R.; Agostinis, P.; de Witte, P.A.M. Elucidation of the tumoritropic principle of hypericin. Br. J. Cancer 2005, 92, 1406–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Prince, A.M.; Pascual, D.; Meruelo, D.; Liebes, L.; Mazur, Y.; Dubovi, E.; Mandel, M.; Lavie, G. Strategies for Evaluation of Enveloped Virus Inactivation in Red Cell Concentrates Using Hypericin. Photochem. Photobiol. 2000, 71, 188–195. [Google Scholar] [CrossRef]
  3. Rahimipour, S.; Palivan, C.; Barbosa, F.; Bilkis, I.; Koch, Y.; Weiner, L.; Fridkin, M.; Mazur, Y.; Gescheidt, G. Chemical and Photochemical Electron Transfer of New Helianthrone Derivatives: Aspects of Their Photodynamic Activity. J. Am. Chem. Soc. 2003, 125, 1376–1384. [Google Scholar] [CrossRef] [PubMed]
  4. Hargrave, K.D.; Hess, F.K.; Oliver, J.T. N-(4-Substituted-thiazolyl)oxamic acid derivatives, new series of potent, orally active antiallergy agents. J. Med. Chem. 1983, 26, 1158–1163. [Google Scholar] [CrossRef]
  5. Morton, J.G.M.; Draghici, C.; Kwon, L.D.; Njadarson, J.T. Rapid Assembly of Vinigrol’s Unique Carbocyclic Skeleton. Org. Lett. 2009, 11, 4492–4495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Maiti, R.; Mishra, B.R.; Jowhar, J.; Mohapatra, D.; Parida, S.; Bisoi, D. Effect of Oxcarbazepine on Serum Brain Derived Neurotrophic Factor in Bipolar Mania: An Exploratory Study. Clin. Psychopharmacol. Neurosci. 2017, 15, 170–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bang, L.M.; Goa, K.L. Spotlight on oxcarbazepine in epilepsy. CNS Drugs 2004, 18, 57–61. [Google Scholar] [CrossRef]
  8. Liu, M.; Yanagihara, N.; Toyohira, Y.; Tsutsui, M.; Ueno, S.; Shinohara, Y. Dual Effects of Daidzein, a Soy Isoflavone, on Catecholamine Synthesis and Secretion in Cultured Bovine Adrenal Medullary Cells. Endocrinology 2007, 148, 5348–5354. [Google Scholar] [CrossRef] [Green Version]
  9. Choi, E.J.; Kim, G.-H. The antioxidant activity of daidzein metabolites, O-desmethylangolensin and equol, in HepG2 cells. Mol. Med. Rep. 2018, 18, 4163–4174. [Google Scholar] [CrossRef]
  10. Scott, L.J. Sitaxentan: In Pulmonary Arterial Hypertension. Drugs 2007, 67, 761–770. [Google Scholar] [CrossRef]
  11. Erve, J.C.L.; Gauby, S.; Maynard, J.W., Jr.; Svensson, M.A.; Tonn†, G.; Quinn, K.P. Bioactivation of Sitaxentan in Liver Microsomes, Hepatocytes, and Expressed Human P450s with Characterization of the Glutathione Conjugate by Liquid Chromatography Tandem Mass Spectrometry. Chem. Res. Toxicol. 2013, 26, 926–936. [Google Scholar] [CrossRef] [PubMed]
  12. Wicklow, D.T.; Joshi, B.K.; Gamble, W.R.; Gloer, J.B.; Dowd, P.F. Antifungal Metabolites (Monorden, Monocillin IV, and Cerebrosides) from Humicola fuscoatra Traaen NRRL 22980, a Mycoparasite of Aspergillus flavus Sclerotia. Appl. Environ. Microbiol. 1998, 64, 4482–4484. [Google Scholar] [PubMed]
  13. Moulin, E.; Zoete, V.; Barluenga, S.; Karplus, M.; Winssinger, N. Design, Synthesis, and Biological Evaluation of HSP90 Inhibitors Based on Conformational Analysis of Radicicol and Its Analogues. J. Am. Chem. Soc. 2005, 127, 6999–7004. [Google Scholar] [CrossRef] [PubMed]
  14. Misiek, M.; Williams, J.; Schmich, K.; Hüttel, W.; Merfort, I.; Salomon, C.E.; Aldrich, C.C.; Hoffmeister, D. Structure and Cytotoxicity of Arnamial and Related Fungal Sesquiterpene Aryl Esters. J. Nat. Prod. 2009, 72, 1888–1891. [Google Scholar] [CrossRef] [PubMed]
  15. Misiek, M.; Hoffmeister, D. Sesquiterpene aryl ester natural products in North American Armillaria species. Mycol. Prog. 2012, 11, 7–15. [Google Scholar] [CrossRef]
  16. Hajime Kobori, H.; Sekiya, A.; Suzuki, T.; Choi, J.-H.; Hirai, H.; Kawagishi, H. Bioactive Sesquiterpene Aryl Esters from the Culture Broth of Armillaria sp. J. Nat. Prod. 2015, 78, 163–167. [Google Scholar] [CrossRef]
  17. Shiota, S.; Shimizu, M.; Mizushima, T.; Ito, H.; Hatano, T.; Yoshida, T.; Tsuchiya, T. Marked reduction in the minimum inhibitory concentration (MIC) of beta-lactams in methicillin-resistant Staphylococcus aureus produced by epicatechin gallate, an ingredient of green tea (Camellia sinensis). Biol. Pharm. Bull. 1999, 22, 1388–1390. [Google Scholar] [CrossRef]
  18. Hughes, D.L. Patent Review of Manufacturing Routes to Recently Approved PARP Inhibitors: Olaparib, Rucaparib, and Niraparib. Org. Process Res. Dev. 2017, 21, 1227–1244. [Google Scholar] [CrossRef]
  19. Merin, B.; Josini, C.J. Rucaparib: A PARP inhibitor for the treatment of advanced ovarian cancer. Int. J. Basic Clin. Pharmacol. 2017, 6, 1018–1019. [Google Scholar] [CrossRef]
  20. Coleman, R.L.; Oza, A.M.; Lorusso, D.; Aghajanian, C.; Oaknin, A.; Dean, A.; Colombo, N.; Weberpals, J.I.; Clamp, A.; Scambia, G.; et al. Rucaparib maintenance treatment for recurrent ovarian carcinoma after response to platinum therapy (ARIEL3): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2017, 390, 1949–1961. [Google Scholar] [CrossRef]
  21. White, A.W.; Almassy, R.; Calvert, A.H.; Curtin, N.J.; Griffin, R.J.; Hostomsky, Z.; Maegley, K.; Newell, D.R.; Srinivasan, S.; Golding, B.T. Resistance-Modifying Agents. 9.1 Synthesis and Biological Properties of Benzimidazole Inhibitors of the DNA Repair Enzyme Poly(ADP-ribose) Polymerase. J. Med. Chem. 2000, 43, 4084–4097. [Google Scholar] [CrossRef] [PubMed]
  22. Deng, X.; Dzamko, N.; Prescott, A.; Davies, P.; Liu, Q.; Yang, Q.; Lee, J.-D.; Patricelli, M.P.; Nomanbhoy, T.K.; Alessi, D.R.; et al. Characterization of a selective inhibitor of the Parkinson’s disease kinase LRRK2. Nat. Chem. Biol. 2011, 7, 203–205. [Google Scholar] [CrossRef] [PubMed]
  23. Fan, Y.; Howden, A.J.M.; Sarhan, A.R.; Lis, P.; Ito, G.; Martinez, T.N.; Brockmann, K.; Gasser, T.; Alessi, D.R.; Sammler, E.M. Interrogating Parkinson’s disease LRRK2 kinase pathway activity by assessing Rab10 phosphorylation in human neutrophils. Biochem. J. 2018, 475, 23–44. [Google Scholar] [CrossRef] [PubMed]
  24. Choudhary, U.; Mir, A.A.; Emrick, T. Soluble, Allyl-Functionalized Deoxybenzoin Polymers. Macromolecules 2017, 50, 3772–3778. [Google Scholar] [CrossRef]
  25. Chena, C.-H.; Lina, C.-H.; Hona, J.-M.; Wanga, M.-W.; Juang, T.-Y. First halogen and phosphorus-free, flame-retardant benzoxazine thermosets derived from main-chain type bishydroxydeoxybenzoin-based benzoxazine polymers. Polymer 2018, 154, 35–41. [Google Scholar] [CrossRef]
  26. Hu, X.; Wang, Y.; Yu, J.; Zhu, J.; Hu, Z. Nonhalogen flame retarded poly(butylene terephthalate) composite using aluminum phosphinate and phosphorus-containing deoxybenzoin polymer. J. Appl. Polym. Sci. 2017, 134, 45537–45546. [Google Scholar] [CrossRef]
  27. Moon, S.; Ku, B.-C.; Emrick, T.; Coughlin, B.E.; Farris, R.J. Flame Resistant Electrospun Polymer Nanofibers from Deoxybenzoin-based Polymers. J. Appl. Polym. Sci. 2009, 111, 301–307. [Google Scholar] [CrossRef]
  28. Zhang, L.; Wu, W.; Zhong, Y.; Zhu, S.; Wang, Z.; Zou, Z. Synergistic Effects of BHDB-IPC with AlPi/MCA on Flame Retarding TPEE. RSC Adv. 2015, 5, 87609–87615. [Google Scholar] [CrossRef]
  29. Reis, D.C.; Pinto, M.C.X.; Souza-Fagundes, E.M.; Wardell, S.M.S.V.; Wardell, J.L.; Beraldo, H. Antimony(III) complexes with 2-benzoylpyridine-derived thiosemicarbazones: Cytotoxicity against human leukemia cell lines. Eur. J. Med. Chem. 2010, 45, 3904–3910. [Google Scholar] [CrossRef]
  30. Hochgürtel, M.; Biesinger, R.; Kroth, H.; Piecha, D.; Hofmann, M.W.; Krause, S.; Schaaf, O.; Nicolau, C.; Eliseev, A.V. Ketones as Building Blocks for Dynamic Combinatorial Libraries:  Highly Active Neuraminidase Inhibitors Generated via Selection Pressure of the Biological Target. J. Med. Chem. 2003, 46, 356–358. [Google Scholar] [CrossRef]
  31. Rong, F.; Chow, S.; Yan, S.; Larson, G.; Hong, Z.; Wu, J. Structure–activity relationship (SAR) studies of quinoxalines as novel HCV NS5B RNA-dependent RNA polymerase inhibitors. Bioorg. Med. Chem. Lett. 2007, 17, 1663–1666. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, X.-J.; Zhang, L.; Sun, X.; Xu, Y.; Krishnamurthy, D.; Senanayake, C.H. Addition of Grignard Reagents to Aryl Acid Chlorides:  An Efficient Synthesis of Aryl Ketones. Org. Lett. 2005, 7, 5593–5595. [Google Scholar] [CrossRef] [PubMed]
  33. Urgoitia, G.; SanMartin, R.; Herrero, M.T.; Domínguez, E. An Aerobic Alternative to Oxidative Ozonolysis of Styrenes. Adv. Synth. Catal. 2016, 358, 1150–1156. [Google Scholar] [CrossRef]
  34. Cheng, L.-J.; Mankad, N.P. Cu-Catalyzed Hydrocarbonylative C−C Coupling of Terminal Alkynes with Alkyl Iodides. J. Am. Chem. Soc. 2017, 139, 10200–10203. [Google Scholar] [CrossRef] [PubMed]
  35. Iranpoor, N.; Firouzabadi, H.; Khalili, D.; Motevalli, S. Easily Prepared Azopyridines As Potent and Recyclable Reagents for Facile Esterification Reactions. An Efficient Modified Mitsunobu Reaction. J. Org. Chem. 2008, 73, 4882–4887. [Google Scholar] [CrossRef] [PubMed]
  36. But, T.Y.S.; Toy, P.H. Organocatalytic Mitsunobu Reactions. J. Am. Chem. Soc. 2006, 128, 9636–9637. [Google Scholar] [CrossRef] [PubMed]
  37. Ratchadapiban, K.; Praserthdam, P.; Tungasmita, D.N.; Tangku, C.; Anutrasakda, W. Effect of Surface Modifications of SBA-15 with Aminosilanes and 12-Tungstophosphoric Acid on Catalytic Properties in Environmentally Friendly Esterification of Glycerol with Oleic Acid to Produce Monoolein. Catalysts 2018, 8, 360. [Google Scholar] [CrossRef]
  38. Makihara, M.; Aoki, H.; Komura, K. Reaction Profiles of High Silica MOR Zeolite Catalyzed Friedel–Crafts Acylation of Anisole Using Acetic Anhydride in Acetic Acid. Catal. Lett. 2018, 148, 2974–2979. [Google Scholar] [CrossRef]
  39. Li, L.-H.; Niua, Z.-J.; Liang, Y.-M. New Friedel–Crafts strategy for preparing 3-acylindoles. Org. Biomol. Chem. 2018, 16, 7792–7796. [Google Scholar] [CrossRef]
  40. Suchand, B.; Satyanarayana, G. Palladium-Catalyzed Direct Acylation: One-Pot Relay Synthesis of Anthraquinones. Synthesis 2018, 50. [Google Scholar] [CrossRef]
  41. Chakraborti, A.K.; Singh, B.; Chankeshwara, S.V.; Patel, A.R. Protic Acid Immobilized on Solid Support as an Extremely Efficient Recyclable Catalyst System for a Direct and Atom Economical Esterification of Carboxylic Acids with Alcohols. J. Org. Chem. 2009, 74, 5967–5974. [Google Scholar] [CrossRef] [PubMed]
  42. Moumne, R.; Lavielle, S.; Karoyan, P. Efficient Synthesis of β2-Amino Acid by Homologation of α-Amino Acids Involving the Reformatsky Reaction and Mannich-Type Imminium Electrophile. J. Org. Chem. 2006, 71, 3332–3334. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, Z.; Wen, Y.; Fu, Y.; Chen, H.; Ye, M.; Luo, G. Graphene Oxide: An Efficient Acid Catalyst for the Construction of Esters from Acids and Alcohols. Synlett 2017, 28, 981–985. [Google Scholar] [CrossRef]
  44. Pathak, G.; Das, D.; Rokhum, L. A microwave-assisted highly practical chemoselective esterification and amidation of carboxylic acids. RSC Adv. 2016, 6, 93729–93740. [Google Scholar] [CrossRef]
  45. Jiang, L.; Yu, J.; Niu, F.; Zhang, D.; Sun, X. A high-efficient method for the amidation of carboxylic acids promoted by triphenylphosphine oxide and oxalyl chloride. Heteroat. Chem. 2017, 28, e21364. [Google Scholar] [CrossRef]
  46. Moon, H.K.; Sung, G.S.; Kim, B.R.; Park, J.K.; Yoon, Y.-J.; Yoon, H.J. One for Many: A Universal Reagent for Acylation Processes. Adv. Synth. Catal. 2016, 358, 1725–1730. [Google Scholar] [CrossRef]
  47. Partenheimer, W. Methodology and scope of metal/bromide autoxidation of hydrocarbons. Catal. Today 1995, 23, 69–158. [Google Scholar] [CrossRef]
  48. Bäckvall, J.-E. Modern Oxidation Methods, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2004; ISBN 978-3-527-63203-9. [Google Scholar]
  49. Bahador Karami, B.; Montazerozohori, M. Tungstate Sulfuric Acid (TSA)/KMnO4 as a Novel Heterogeneous System for Rapid Deoximation. Molecules 2006, 11, 720–725. [Google Scholar] [CrossRef]
  50. Lounisa, Z.; Riahi, A.; Djafri, F.; Muzart, J. Chromium-exchanged zeolite (CrE-ZSM-5) as catalyst for alcohol oxidation and benzylic oxidation with t-BuOOH. Appl. Catal. A Gen. 2006, 309, 270–272. [Google Scholar] [CrossRef]
  51. Moghadam, M.; Mirkhani, V.; Tangestaninejad, S.; Mohammdpoor-Baltorka, I.; Kargar, H. Silica supported Mn(Br8TPP)Cl and Mn(TPP)Cl as efficient and reusable catalysts for selective hydrocarbon oxidation under various reaction conditions: The effect of substituted bromines on the catalytic activity and reusability. J. Mol. Catal. A Chem. 2008, 288, 116–124. [Google Scholar] [CrossRef]
  52. Singh, S.J.; Jayaram, R.V. Oxidation of alkylaromatics to benzylic ketones using TBHP as an oxidant over LaMO3 (M = Cr, Co, Fe, Mn, Ni) perovskites. Catal. Commun. 2009, 10, 2004–2007. [Google Scholar] [CrossRef]
  53. Mayilmurugan, R.; Stoeckli-Evans, H.; Suresh, E.; Palaniandavar, M. Chemoselective and biomimetic hydroxylation of hydrocarbons by non-heme μ-oxo-bridged diiron(III) catalysts using m-CPBA as oxidant. Dalton Trans. 2009, 26, 5101–5114. [Google Scholar] [CrossRef] [PubMed]
  54. Yusubov, M.S.; Nemykin, V.N.; Zhdankin, V.Z. Transition metal-mediated oxidations utilizing monomeric iodosyl- and iodylarene species. Tetrahedron 2010, 66, 5745–5752. [Google Scholar] [CrossRef]
  55. Zhou, M.; Schley, N.D.; Crabtree, R.H. Cp* Iridium Complexes Give Catalytic Alkane Hydroxylation with Retention of Stereochemistry. J. Am. Chem. Soc. 2010, 132, 12550–12551. [Google Scholar] [CrossRef] [PubMed]
  56. Srour, H.; Maux, P.L.; Simonneaux, G. Enantioselective Manganese-Porphyrin-Catalyzed Epoxidation and C–H Hydroxylation with Hydrogen Peroxide in Water/Methanol Solutions. Inorg. Chem. 2012, 51, 5850–5856. [Google Scholar] [CrossRef] [PubMed]
  57. Tu, D.-H.; Yang Li, Y.; Jiangwei Li, J.; Gu, Y.-J.; Wang, B.; Liu, Z.-T.; Liu, Z.-W.; Lu, J. Selective oxidation of arylalkanes with N-Graphitic-Modified cobalt nanoparticles in water. Catal. Commun. 2017, 97, 130–133. [Google Scholar] [CrossRef]
  58. Zhang, Q.; Gorden, J.D.; Goldsmith, C.R. C–H Oxidation by H2O2 and O2 Catalyzed by a Non-Heme Iron Complex with a Sterically Encumbered Tetradentate N-Donor Ligand. Inorg. Chem. 2013, 52, 13546–13554. [Google Scholar] [CrossRef]
  59. Stepovik, L.P.; Potkina, A.Y. Oxidation of Alkylarene C–H Bonds by tert-Butyl Hydroperoxide in the Presence of Cobalt, Chromium, and Vanadium Acetylacetonates. Russ. J. Gen. Chem. 2013, 83, 1047–1059. [Google Scholar] [CrossRef]
  60. Yang, Z.; Liu, Z.; Zhang, H.; Yu, B.; Zhao, Y.; Wang, H.; Ji, G.; Chen, Y.; Liu, X.; Liu, Z. N-Doped porous carbon nanotubes: synthesis and application in catalysis. Chem. Commun. 2017, 53, 929–932. [Google Scholar] [CrossRef]
  61. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Mechanism of Selective C–H Hydroxylation Mediated by Manganese Aminopyridine Enzyme Models. ACS Catal. 2015, 5, 39–44. [Google Scholar] [CrossRef]
  62. Miyamoto, M.; Minami, Y.; Ukaji, Y.; Kinoshita, H.; Inomata, K. A New Aerobic Oxidation System Using Pd-Cu Catalysts in the Presence of CO. Chem. Lett. 1994, 1149–1152. [Google Scholar] [CrossRef]
  63. Ishii, Y.; Nakayama, K.; Takeno, M.; Sakaguchi, S.; Iwahama, T.; Nishiyama, Y. Yasutaka Ishii, Kouichi Nakayama, Mitsuhiro Takeno, Satoshi Sakaguchi, Takahiro Iwahama, and Yutaka Nishiyama. J. Org. Chem. 1995, 60, 3934–3935. [Google Scholar] [CrossRef]
  64. Ishii, Y.; Iwahama, T.; Sakaguchi, S.; Nakayama, K.; Nishiyama, Y. Alkane Oxidation with Molecular Oxygen Using a New Efficient Catalytic System:  N-Hydroxyphthalimide (NHPI) Combined with Co(acac)n (n = 2 or 3). J. Org. Chem. 1996, 61, 4520–4526. [Google Scholar] [CrossRef] [PubMed]
  65. Yoshino, Y.; Hayashi, Y.; Iwahama, T.; Sakaguchi, S.; Ishii, Y. Catalytic Oxidation of Alkylbenzenes with Molecular Oxygen under Normal Pressure and Temperature by N-Hydroxyphthalimide Combined with Co(OAc)2. J. Org. Chem. 1997, 62, 6810–6813. [Google Scholar] [CrossRef]
  66. Ishii, Y.; Sakaguchi, S. A new strategy for alkane oxidation with O2 Using N-hydroxyphthalimide (NHPI) as a radical catalyst. Catal. Surv. Jpn. 1999, 3, 27–35. [Google Scholar] [CrossRef]
  67. Ishii, Y.; Sakaguchi, S.; Iwahama, T. Innovation of Hydrocarbon Oxidation with Molecular Oxygen and Related Reactions. Adv. Synth. Catal. 2001, 343, 393–427. [Google Scholar] [CrossRef]
  68. Sheldon, R.A.; Arends, I.W.C.E. Organocatalytic Oxidations Mediated by Nitroxyl Radicals. Adv. Synth. Catal. 2004, 346, 1051–1071. [Google Scholar] [CrossRef]
  69. Yang, G.; Zhang, Q.; Miao, H.; Tong, X.; Xu, J. Selective Organocatalytic Oxygenation of Hydrocarbons by Dioxygen Using Anthraquinones and N-Hydroxyphthalimide. Org. Lett. 2005, 7, 263–266. [Google Scholar] [CrossRef]
  70. Matsushita, Y.; Sugamoto, K.; Iwakiri, Y.; Yoshida, S.; Chaen, T.; Matsui, T. Aerobic oxidation of 8,11,13-abietatrienes catalyzed by N-hydroxyphthalimide combined with 2,20-azobis(4-methoxy-2,4-dimethylvaleronitrile) and its application to synthesis of naturally occurring diterpenes. Tetrahedron Lett. 2010, 51, 3931–3934. [Google Scholar] [CrossRef]
  71. Hollmann, F.; Arends, I.W.C.E.; Buehler, K.; Schallmey, A.; Bühler, B. Enzyme-mediated oxidations for the chemist. Green Chem. 2011, 13, 226–265. [Google Scholar] [CrossRef]
  72. Rickert, A.; Krombach, V.; Hamers, O.; Zorn, H.; Maison, W. Enzymatic allylic oxidations with a lyophilisate of the edible fungus Pleurotus sapidus. Green Chem. 2012, 14, 639–644. [Google Scholar] [CrossRef]
  73. Weidmann, V.; Schaffrath, M.; Zorn, H.; Rehbein, J.; Maison, W. Elucidation of the regio- and chemoselectivity of enzymatic allylic oxidations with Pleurotus sapidus–conversion of selected spirocyclic terpenoids and computational analysis. Beilstein J. Org. Chem. 2013, 9, 2233–2241. [Google Scholar] [CrossRef] [PubMed]
  74. Kexian Chen, K.; Zhang, P.; Wang, Y.; Li, H. Metal-free allylic/benzylic oxidation strategies with molecular oxygen: Recent advances and future prospects. Green Chem. 2014, 16, 2344–2374. [Google Scholar] [CrossRef]
  75. Ren, L.; Wang, L.; Lv, Y.; Li, G.; Gao, S. Synergistic H4NI−AcOH Catalyzed Oxidation of the Csp3−H Bonds of Benzylpyridines with Molecular Oxygen. Org. Lett. 2015, 17, 2078–2081. [Google Scholar] [CrossRef]
  76. Wang, H.; Wang, Z.; Huang, H.; Tan, J.; Xu, K. KOtBu-Promoted Oxidation of (Hetero)benzylic Csp3−H to Ketones with Molecular Oxygen. Org. Lett. 2016, 18, 5680–5683. [Google Scholar] [CrossRef]
  77. Jin, W.; Zheng, P.; Wing-Tak Wong, W.-T.; Law, G.-L. Efficient Selenium-Catalyzed Selective C(sp3)-H Oxidation of Benzylpyridines with Molecular Oxygen. Adv. Synth. Catal. 2017, 359, 1588–1593. [Google Scholar] [CrossRef]
  78. Zhang, H.-H.; Wang, Y.-Q.; Huang, L.-T.; Zhu, L.-Q.; Feng, Y.-Y.; Lu, Y.-M.; Zhao, Q.; Wang, X.-Q.; Wang, Z. NaI-mediated divergent synthesis of isatins and isoindigoes: A new protocol enabled by an oxidation relay strategy. Chem. Commun. 2018, 54, 8265–8268. [Google Scholar] [CrossRef]
  79. Gao, B.; Kong, D.; Zhang, Y. Preparation and catalytic activity of P4VP–Cu(II)complex supported on silica gel. J. Mol. Catal. A Chem. 2008, 286, 143–148. [Google Scholar] [CrossRef]
  80. Shaabani, A.; Rahmati, A. Aerobic oxidation of alkyl arenes using a combination of N-hydroxy phthalimide and recyclable cobalt(II) tetrasulfophthalocyanine supported on silica. Catal. Commun. 2008, 9, 1692–1697. [Google Scholar] [CrossRef]
  81. Murwani, I.K.; Scheurell, K.; Kemnitz, E. Liquid phase oxidation of ethylbenzene on pure and metal doped HS-AlF3. Catal. Commun. 2008, 10, 227–231. [Google Scholar] [CrossRef]
  82. Shaabani, A.; Hezarkhani, Z.; Badali, E. Wool supported manganese dioxide nano-scale dispersion: A biopolymer based catalyst for the aerobic oxidation of organic compounds. RSC Adv. 2015, 5, 61759–61767. [Google Scholar] [CrossRef]
  83. Chauhan, P.; Yan, N. Nanocrystalline cellulose grafted phthalocyanine: A heterogeneous catalyst for selective aerobic oxidation of alcohols and alkyl arenes at room temperature in a green solvent. RSC Adv. 2015, 5, 37517–37520. [Google Scholar] [CrossRef]
  84. Rezaeifard, A.; Khoshyan, A.; Jafarpour, M.; Pourtahmasb, M. Selective aerobic benzylic C–H oxidation co-catalyzed by N-hydroxyphthalimide and Keplerate {Mo72V30} nanocluster. RSC Adv. 2017, 7, 15754–15761. [Google Scholar] [CrossRef]
  85. Lin, X.; Nie, Z.; Zhang, L.; Mei, S.; Chen, Y.; Zhang, B.; Zhu, R.; Liu, Z. Nitrogen-doped carbon nanotubes encapsulate cobalt nanoparticles as efficient catalysts for aerobic and solvent-free selective oxidation of hydrocarbons. Green Chem. 2017, 19, 2164–2173. [Google Scholar] [CrossRef]
  86. Kuwahara, Y.; Yoshimura, Y.; Yamashita, H. Liquid-phase oxidation of alkylaromatics to aromatic ketones with molecular oxygen over a Mn-based metal–organic framework. Dalton Trans. 2017, 46, 8415–8421. [Google Scholar] [CrossRef] [PubMed]
  87. Mouchaham, G.; Abeykoon, B.; Giménez-Marqués, M.; Navalon, S.; Santiago-Portillo, A.; Affram, M.; Guillou, N.; Martineau, C.; Garcia, H.; Fateeva, A.; et al. Adaptability of the metal(III, IV) 1,2,3-trioxobenzene rod secondary building unit for the production of chemically stable and catalytically active MOFs. Chem. Commun. 2017, 53, 7661–7664. [Google Scholar] [CrossRef] [PubMed]
  88. Guo, S.; Zhang, Q.; Li, H.; Guo, H.; He, W. Ag/C nanoparticles catalysed aerobic oxidation of diaryl and aryl(hetero) methylenes into ketones. Nano Res. 2017, 10, 3261–3267. [Google Scholar] [CrossRef]
  89. Wang, X.; Liu, M.; Wang, Y.; Fan, H.; Wu, J.; Huang, C.; Hou, H. Cu(I) Coordination Polymers as the Green Heterogeneous Catalysts for Direct C–H Bonds Activation of Arylalkanes to Ketones in Water with Spatial Confinement Effect. Inorg. Chem. 2017, 56, 13329–13336. [Google Scholar] [CrossRef]
  90. Zhang, G.; Wang, D.; Feng, P.; Shi, S.; Wang, C.; Zheng, A.; Lüa, G.; Tiana, Z. Synthesis of zeolite Beta containing ultra-small CoO particles for ethylbenzene oxidation. Chin. J. Catal. 2017, 38, 1207–1215. [Google Scholar] [CrossRef]
  91. Sharma, A.S.; Kaur, H. Sustainable Protocol for Benzylic -CH2 Oxidation with Dioxygen to Phenones Using AuNPs@ Resin Beads. Chem. Sel. 2017, 31, 10112–10117. [Google Scholar] [CrossRef]
  92. Dai, X.; Chen, D.; Zhou, W.; Yang, S.; Sun, F.; Qian, J.; He, M.; Chen, Q. Aromatization of hydrocarbons by oxidative dehydrogenation catalyzed by nickel porphyrin with molecular oxygen. Catal. Commun. 2018, 117, 85–89. [Google Scholar] [CrossRef]
  93. Jiang, J.; Wang, J.-X.; Zhou, X.-T.; Chen, H.-Y.; Ji, H.-B. Mechanistic Understanding towards the Role of Cyclohexene in Enhancing the Efficiency of Manganese Porphyrin-Catalyzed Aerobic Oxidation of Diphenylmethane. Eur. J. Inorg. Chem. 2018, 2018, 2666–2674. [Google Scholar] [CrossRef]
  94. Wentzel, B.B.; Donners, M.P.J.; Alsters, P.L.; Feiters, M.C.; Nolte, R.J.M. N-Hydroxyphthalimide/Cobalt(II) Catalyzed Low Temperature Benzylic Oxidation Using Molecular Oxygen. Tetrahedron 2000, 56, 7797–7803. [Google Scholar] [CrossRef]
  95. Patil, R.D.; Fuchs, B.; Taha, N.; Sasson, Y. Solvent–free and Selective Autooxidation of Alkylbenzenes Catalyzed by Co/NHPI under Phase Transfer Conditions. Chem. Sel. 2016, 1, 3791–3796. [Google Scholar] [CrossRef]
  96. Hruszkewycz, D.P.; Miles, K.S.; Thiel, O.R.; Stahl, S.S. Co/NHPI-mediated aerobic oxygenation of benzylic C–H bonds in pharmaceutically relevant molecules. Chem. Sci. 2017, 8, 1282–1287. [Google Scholar] [CrossRef] [Green Version]
  97. Wang, J.-R.; Liu, L.; Wang, Y.-F.; Zhang, Y.; Deng, W.; Guo, Q.-X. Aerobic oxidation with N-hydroxyphthalimide catalysts in ionic liquid. Tetrahedron Lett. 2005, 46, 4647–4651. [Google Scholar] [CrossRef]
  98. Shen, J.; Tan, C.-H. Anthrone-derived NHPI analogues as catalysts in reactions using oxygen as an oxidant. Org. Biomol. Chem. 2008, 6, 4096–4098. [Google Scholar] [CrossRef]
  99. Baucherel, X.; Gonsalvi, L.; Arends, I.W.C.E.; Ellwood, S.; Sheldon, R.A. Aerobic Oxidation of Cycloalkanes, Alcohols and Ethylbenzene Catalyzed by the Novel Carbon Radical Chain Promoter NHS (N-Hydroxysaccharin). Adv. Synth. Catal. 2004, 346, 286–296. [Google Scholar] [CrossRef]
  100. Nakamura, R.; Obora, Y.; Ishii, Y. Selective Oxidation of Acetophenones Bearing Various Functional Groups to Benzoic Acid Derivatives with Molecular Oxygen. Adv. Synth. Catal. 2009, 351, 1677–1684. [Google Scholar] [CrossRef]
  101. Shaabani, A.; Farhangi, E.; Rahmati, A. Aerobic oxidation of alkyl arenes and alcohols using cobalt(II) phthalocyanine as a catalyst in 1-butyl-3-methyl-imidazolium bromide. Appl. Catal. A Gen. 2008, 338, 14–19. [Google Scholar] [CrossRef]
  102. Wang, P.; She, Y.; Fu, H.; Zhao, W.; Wang, M. Oxidation of alkylaromatics to aromatic ketones catalyzed by metalloporphyrins under the special temperature control method. Can. J. Chem. 2014, 92, 1059–1065. [Google Scholar] [CrossRef]
  103. Gutmann, B.; Elsner, P.; Roberge, D.; Kappe, C.O. Homogeneous Liquid-Phase Oxidation of Ethylbenzene to Acetophenone in Continuous Flow Mode. ACS Catal. 2013, 3, 2669–2676. [Google Scholar] [CrossRef]
  104. Wang, J.-Q.; He, L.-N. Polyethylene glycol radical-initiated benzylic C–H bond oxygenation in compressed carbon dioxide. New J. Chem. 2009, 33, 1637–1640. [Google Scholar] [CrossRef]
  105. Huang, J.-G.; Guo, Y.; Jiang, J.-A.; Liu, H.-W.; Ji, Y.-F. An eco-friendly Co(OAc)2-catalyzed aerobic oxidation of 4-benzylphenols into 4-hydroxybenzophenones. Res. Chem. Intermed. 2015, 41, 7115–7124. [Google Scholar] [CrossRef]
  106. Orlińska, B.; Romanowska, I. N-Hydroxyphthalimide and transition metal salts as catalysts of the liquid-phase oxidation of 1-methoxy-4-(1-methylethyl)benzene with oxygen. Cent. Eur. J. Chem. 2011, 9, 670–676. [Google Scholar] [CrossRef]
  107. Chen, L.; Tang, R.; Li, Z.; Liang, S. An Efficient and Mild Oxidation of α-Isophorone to Ketoisophorone Catalyzed by N-Hydroxyphthalimide and Copper Chloride. Bull. Korean Chem. Soc. 2012, 33, 459–463. [Google Scholar] [CrossRef] [Green Version]
  108. Nechab, M.; Einhorn, C.; Einhorn, J. New aerobic oxidation of benzylic compounds: Efficient catalysis by N-hydroxy-3,4,5,6-tetraphenylphthalimide (NHTPPI)/CuCl under mild conditions and low catalyst loading. Chem. Commun. 2004, 1500–1501. [Google Scholar] [CrossRef]
  109. Abe, T.; Tanaka, S.; Ogawa, A.; Tamura, M.; Sato, K.; Itoh, S. Copper-Catalyzed Selective Oxygenation of Methyl and Benzyl Substituents in Pyridine with O2. Chem. Lett. 2017, 46, 348–350. [Google Scholar] [CrossRef]
  110. Jiang, J.-A.; Chen, C.; Huang, J.-G.; Liu, H.-W.; Cao, S.; Ji, Y.-F. Cu(OAc)2-catalyzed remote benzylic C(sp3)–H oxyfunctionalization for C=O formation directed by the hindered para-hydroxyl group with ambient air as the terminal oxidant under ligand- and additive-free conditions. Green Chem. 2014, 16, 1248–1254. [Google Scholar] [CrossRef]
  111. Zhang, L.; Ang, G.Y.; Chiba, S. Copper-Catalyzed Benzylic C–H Oxygenation under an Oxygen Atmosphere via N-H Imines as an Intramolecular Directing Group. Org. Lett. 2011, 13, 1622–1625. [Google Scholar] [CrossRef]
  112. Yu, J.-W.; Mao, S.; Wang, Y.-Q. Copper-catalyzed base-accelerated direct oxidation of C–H bond to synthesize benzils, isatins, and quinoxalines with molecular oxygen as terminal oxidant. Tetrahedron Lett. 2015, 56, 1575–1580. [Google Scholar] [CrossRef]
  113. De Houwer, J.; Tehrani, K.A.; Maes, B.U.W. Synthesis of Aryl(di)azinyl Ketones through Copper- and Ironcatalyzed Oxidation of the Methylene Group of Aryl(di)azinylmethanes. Angew. Int. Chem. Ed. 2012, 51, 2745–2748. [Google Scholar] [CrossRef] [PubMed]
  114. Sterckx, H.; De Houwer, J.; Mensch, C.; Caretti, I.; Tehrani, K.A.; Herrebout, W.A.; Van Doorslaerb, S.; Maes, B.U.W. Mechanism of the CuII-catalyzed benzylic oxygenation of (aryl)(heteroaryl)methanes with oxygen. Chem. Sci. 2016, 7, 346–357. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, J.; Zhang, X.; Yi, H.; Liu, C.; Liu, R.; Zhang, H.; Zhuo, K.; Lei, A. Chloroacetate-Promoted Selective Oxidation of Heterobenzylic Methylenes under Copper Catalysis. Angew. Chem. Int. Ed. 2015, 54, 1261–1265. [Google Scholar] [CrossRef] [PubMed]
  116. Hayashi, Y.; Komiya, N.; Suzuki, K.; Murahashi, S.-I. Copper-catalyzed aerobic oxidative functionalization of C–Hbonds of alkanes in the presence of acetaldehyde under mild conditions. Tetrahedron Lett. 2013, 54, 2706–2709. [Google Scholar] [CrossRef]
  117. Finney, L.C.; Mitchell, L.J.; Moody, C.J. Visible light mediated oxidation of benzylic sp3 C–H bonds using catalytic 1,4-hydroquinone, or its biorenewable glucoside, arbutin, as a pre-oxidant. Green Chem. 2018, 20, 2242–2249. [Google Scholar] [CrossRef]
  118. Cacchi, S.; Fabrizi, G.; Goggiamani, A.; Iazzetti, A.; Verdiglione, R. Copper-Catalyzed Oxidation of Deoxybenzoins to Benzils under Aerobic Conditions. Synthesis 2013, 45, 1701–1707. [Google Scholar]
  119. Menendez, C.; Gau, S.; Ladeira, S.; Lherbet, C.; Baltas, M. Synthesis of α,β-Diketotriazoles by Aerobic Copper-Catalyzed Oxygenation with Triazole as an Intramolecular Assisting Group. Eur. J. Org. Chem. 2012, 2012, 409–416. [Google Scholar] [CrossRef]
  120. Navarro, M.; Escobar, A.; Landaeta, V.R.; Visbal, G.; Lopez-Linares, F.; Luis, M.L.; Fuentes, A. Catalytic oxidation of tetralin by biologically active copper and palladium complexes. Appl. Catal. A Gen. 2009, 363, 27–31. [Google Scholar] [CrossRef]
  121. Miao, C.; Zhao, H.; Zhao, Q.; Xiaa, C.; Sun, W. NHPI and ferric nitrate: A mild and selective system for aerobic oxidation of benzylic methylenes. Catal. Sci. Technol. 2016, 6, 1378–1383. [Google Scholar] [CrossRef]
  122. Lesieur, M.; Genicot, C.; Pasau, P. Development of a Flow Photochemical Aerobic Oxidation of Benzylic C−H Bonds. Org. Lett. 2018, 20, 1987–1990. [Google Scholar] [CrossRef] [PubMed]
  123. Tang, G.; Gong, Z.; Han, W.; Sun, X. Visible light mediated aerobic photocatalytic activation of CAH bond by riboflavin tetraacetate and N-hydroxysuccinimide. Tetrahedron Lett. 2018, 59, 658–662. [Google Scholar] [CrossRef]
  124. Li, S.; Zhu, B.; Lee, R.; Qiao, B.; Jiang, Z. Visible light-induced selective aerobic oxidative transposition of vinyl halides using a tetrahalogenoferrate(III) complex catalyst. Org. Chem. Front. 2018, 5, 380–385. [Google Scholar] [CrossRef]
  125. Sterckx, H.; De Houwer, J.; Mensch, C.; Herrebout, W.; Tehrani, K.A.; Maes, B.U.W. Base metal catalyzed benzylic oxidation of (aryl)(heteroaryl)methanes with molecular oxygen. Beilstein J. Org. Chem. 2016, 12, 144–153. [Google Scholar] [CrossRef] [PubMed]
  126. Pieber, B.; Kappe, C.O. Direct aerobic oxidation of 2-benzylpyridines in a gas–liquid continuous-flow regime using propylene carbonate as a solvent. Green Chem. 2013, 15, 320–324. [Google Scholar] [CrossRef]
  127. Li, S.-J.; Wang, Y.-G. A novel and selective catalytic oxidation of hydrocarbons to ketones using chloramine-T/O2/Fe(TPP)Cl system. Tetrahedron Lett. 2005, 46, 8013–8015. [Google Scholar] [CrossRef]
  128. Klopstra, M.; Hage, R.; Kelloggc, R.M.; Feringa, B.L. Non-heme iron catalysts for the benzylic oxidation: A parallel ligand screening approach. Tetrahedron Lett. 2003, 44, 4581–4584. [Google Scholar] [CrossRef]
  129. Urgoitia, G.; SanMartin, R.; Herrero, M.T.; Domínguez, E. Palladium NCN and CNC pincer complexes as exceptionally active catalysts for aerobic oxidation in sustainable media. Green Chem. 2011, 13, 2161–2166. [Google Scholar] [CrossRef]
  130. Urgoitia, G.; Maiztegi, A.; SanMartin, R.; Herrero, M.T.; Domízguez, E. Aerobic oxidation at benzylic positions catalyzed by a simple Pd(OAc)2/bis-triazole system. RSC Adv. 2015, 5, 103210–103217. [Google Scholar] [CrossRef]
  131. Dos Santos, A.; Kaïm, L.E.; Grimaud, L. Metal-free aerobic oxidation of benzazole derivatives. Org. Biomol. Chem. 2013, 11, 3282–3287. [Google Scholar] [CrossRef]
  132. Yang, X.; Zhou, L.; Chen, Y.; Chen, C.; Su, Y.; Miao, H.; Xu, J. A promotion effect of alkaline-earth chloride on N-hydroxyphthalimide-catalyzed aerobic oxidation of hydrocarbons. Catal. Commun. 2009, 11, 171–174. [Google Scholar] [CrossRef]
  133. Figiel, P.J.; Sobczak, J.M. Aerobic oxidation of alcohols and alkylaromatics with dioxygen catalysed by N-hydroxyphthalimide with vanadium co-catalysts. New J. Chem. 2007, 31, 1668–1673. [Google Scholar] [CrossRef]
  134. Gao, J.; Tong, X.; Li, X.; Miao, H.; Xu, J. The efficient liquid-phase oxidation of aromatic hydrocarbons by molecular oxygen in the presence of MnCO3. J. Chem. Technol. Biotechnol. 2007, 82, 620–625. [Google Scholar] [CrossRef]
  135. Yan, Y.; Chen, Y.; Yan, M.; Li, X.; Zeng, W. Iridium-catalyzed aerobic oxidation of alkylarenes with excellentturnover numbers. Catal. Commun. 2013, 35, 64–67. [Google Scholar] [CrossRef]
  136. Urgoitia, G.; SanMartin, R.; Herrero, M.T.; Domínguez, E. An outstanding catalyst for the oxygen-mediated oxidation of arylcarbinols, arylmethylene and arylacetylene compounds. Chem. Commun. 2015, 51, 4799–4802. [Google Scholar] [CrossRef] [PubMed]
  137. Kojima, M.; Oisaki, K.; Kanai, M. Chemoselective aerobic photo-oxidation of 9H-fluorenes for the synthesis of 9-fluorenones. Tetrahedron Lett. 2014, 55, 4736–4738. [Google Scholar] [CrossRef]
Figure 1. A selection of biologically relevant aromatic ketones and esters.
Figure 1. A selection of biologically relevant aromatic ketones and esters.
Catalysts 08 00640 g001
Scheme 1. Benzylic C–H oxidation with oxidants other than molecular oxygen.
Scheme 1. Benzylic C–H oxidation with oxidants other than molecular oxygen.
Catalysts 08 00640 sch001
Scheme 2. Solvent-free oxidation of alkylbenzenes by Patil et al.
Scheme 2. Solvent-free oxidation of alkylbenzenes by Patil et al.
Catalysts 08 00640 sch002
Scheme 3. Synthesis of a clinical candidate for the treatment of schizophrenia by Stahl and col.
Scheme 3. Synthesis of a clinical candidate for the treatment of schizophrenia by Stahl and col.
Catalysts 08 00640 sch003
Figure 2. Structure of Py-NHPI prepared by Wang et al.
Figure 2. Structure of Py-NHPI prepared by Wang et al.
Catalysts 08 00640 g002
Scheme 4. Aerobic oxidation of several alkylbenzenes in the presence of a chiral NHPI derivative.
Scheme 4. Aerobic oxidation of several alkylbenzenes in the presence of a chiral NHPI derivative.
Catalysts 08 00640 sch004
Scheme 5. THICA- or NHPI-co-catalyzed benzylic oxidation of 4-ethylacetophenone.
Scheme 5. THICA- or NHPI-co-catalyzed benzylic oxidation of 4-ethylacetophenone.
Catalysts 08 00640 sch005
Scheme 6. Oxidation to 4-hydroxybenzophenones and 4-hydroxyacetophenones.
Scheme 6. Oxidation to 4-hydroxybenzophenones and 4-hydroxyacetophenones.
Catalysts 08 00640 sch006
Scheme 7. NHTPPI-CuCl co-catalyzed oxidation of alkylbenzenes at 35 °C.
Scheme 7. NHTPPI-CuCl co-catalyzed oxidation of alkylbenzenes at 35 °C.
Catalysts 08 00640 sch007
Scheme 8. Oxidation of deoxybenzoins and oxindoles to benzils and isatins.
Scheme 8. Oxidation of deoxybenzoins and oxindoles to benzils and isatins.
Catalysts 08 00640 sch008
Scheme 9. Postulated mechanism for the oxidation of 2-benzylpyridine using O2/CuI/AcOH/DMSO.
Scheme 9. Postulated mechanism for the oxidation of 2-benzylpyridine using O2/CuI/AcOH/DMSO.
Catalysts 08 00640 sch009
Scheme 10. Visible light-mediated oxidation of N-benzyltosylamides.
Scheme 10. Visible light-mediated oxidation of N-benzyltosylamides.
Catalysts 08 00640 sch010
Scheme 11. Aerobic oxidation to α,β-diketotriazoles.
Scheme 11. Aerobic oxidation to α,β-diketotriazoles.
Catalysts 08 00640 sch011
Scheme 12. Visible light-induced oxidation of methylene compounds by Quio and coworkers.
Scheme 12. Visible light-induced oxidation of methylene compounds by Quio and coworkers.
Catalysts 08 00640 sch012
Scheme 13. Oxidation of 2-benzylpyridines under high-temperature/pressure gas–liquid continuous conditions.
Scheme 13. Oxidation of 2-benzylpyridines under high-temperature/pressure gas–liquid continuous conditions.
Catalysts 08 00640 sch013
Scheme 14. Benzylic C–H oxidation of methylene compounds by SanMartin and coworkers.
Scheme 14. Benzylic C–H oxidation of methylene compounds by SanMartin and coworkers.
Catalysts 08 00640 sch014
Scheme 15. NHPI/VO(acac)2-co-catalyzed oxidation of ethylbenzene.
Scheme 15. NHPI/VO(acac)2-co-catalyzed oxidation of ethylbenzene.
Catalysts 08 00640 sch015
Scheme 16. Synthesis of 9-fluorenones by aerobic photooxidation of fluorenes.
Scheme 16. Synthesis of 9-fluorenones by aerobic photooxidation of fluorenes.
Catalysts 08 00640 sch016

Share and Cite

MDPI and ACS Style

Urgoitia, G.; SanMartin, R.; Herrero, M.T.; Domínguez, E. Recent Advances in Homogeneous Metal-Catalyzed Aerobic C–H Oxidation of Benzylic Compounds. Catalysts 2018, 8, 640. https://doi.org/10.3390/catal8120640

AMA Style

Urgoitia G, SanMartin R, Herrero MT, Domínguez E. Recent Advances in Homogeneous Metal-Catalyzed Aerobic C–H Oxidation of Benzylic Compounds. Catalysts. 2018; 8(12):640. https://doi.org/10.3390/catal8120640

Chicago/Turabian Style

Urgoitia, Garazi, Raul SanMartin, María Teresa Herrero, and Esther Domínguez. 2018. "Recent Advances in Homogeneous Metal-Catalyzed Aerobic C–H Oxidation of Benzylic Compounds" Catalysts 8, no. 12: 640. https://doi.org/10.3390/catal8120640

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