Clean Oxidation of (Hetero)benzylic Csp3–H Bonds with Molecular

May 28, 2019 - Road. No. 960,. 410114,. China. E-mail: [email protected]. Total number of pages: 65. Total number of Figures: 51 (pages S15 – S6...
0 downloads 0 Views 1MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

pubs.acs.org/journal/ascecg

Clean Oxidation of (Hetero)benzylic Csp3−H Bonds with Molecular Oxygen Kai-Jian Liu,† Zhen-Hong Duan,† Xiu-Ling Zeng,† Meng Sun,‡ Zilong Tang,‡ Si Jiang,§ Zhong Cao,§ and Wei-Min He*,† †

Department of Chemistry, Hunan University of Science and Engineering, Yangzi Tang Road No. 130, Yongzhou 425100, China School of Chemistry and Chemical Engineering, Hunan University of Science and Technology, Taoyuan Road No. 9, Xiangtan 411201, China § Hunan Provincial Key Laboratory of Materials Protection for Electric Power and Transportation, Changsha University of Science and Technology, Wanjiali Road No. 960, Changsha 410114, China

Downloaded via BUFFALO STATE on July 30, 2019 at 02:20:46 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: By employing atmospheric dioxygen as the sole oxidant, an eco-friendly protocol for the clean preparation of various aromatic (cyclic) ketones through phosphinatepromoted oxidation of benzylic Csp3−H bonds under base-, metal-, additive-, organic-solvent-free conditions has been developed.

KEYWORDS: Base free, Metal free, Dioxygen, Dioxygenation, Aromatic ketone



Scheme 1. Oxidation of Benzylic Csp3−H Bonds

INTRODUCTION (Hetero)aromatic ketones are widely present in natural products, biologically active compounds, and artificial drugs.1−4 They also act as useful building blocks and versatile synthetic precursors in organic synthesis and material sciences.5−8 As a consequence, much effort and energy have been dedicated to developing various methodologies for the preparation of such motifs. Among the known protocols, the benzylic Csp3−H bond oxidation reaction is one of the most efficient methods, given the inexpensiveness, abundance, and accessibility of alkyl aromatics as industrial materials.9,10 Traditionally, (super)stoichiometric amounts of hazardous oxidants (e.g., high-valent metal salts, halides, and peroxides) were employed for these chemical transformations, generally generating large amounts of chemical waste and unwanted byproducts (Scheme 1a).11−17 Therefore, developing an ecofriendly and general protocol for the oxidation of alkyl (hetero)aromatics using safe and environment friendly oxidants is highly appealing. Compared to conventional inorganic or organic oxidants, molecular oxygen is a perfect oxidant because of its environmental benignancy, low cost, and natural abundance.18−39 Over the past decades, significant achievements have been obtained in the oxidation of alkyl aromatics with dioxygen catalyzed by transition metal−ligand complex, uncommercialized metal-supported catalyst, or activated carbon catalyst (Scheme 1b).40−51 Nevertheless, those protocols suffer from limitations of requirement of catalyst, tedious procedures, and meagre substrate generality, and few © 2019 American Chemical Society

methods are available for the direct oxidation of heterobenzylic Csp3−H bonds to construct heteroaryl ketones and dioxygenation products. Due to the low threshold of transition-metal Received: January 1, 2019 Revised: May 24, 2019 Published: May 28, 2019 10293

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

Research Article

ACS Sustainable Chemistry & Engineering

providing the best result (entry 5). Increasing the promoter loading had little influence on the oxidation (entry 11), whereas a lower yield was observed when the promoter loading was reduced (entry 12). Replacing dioxygen with air diminished the yield of 2a (entry 13). Elevating the reaction temperature did not improve the oxidation outcome (entry 14), whereas a decrease in the yield of 2a was detected when the temperature was dropped to 120 °C (entry 15). As a blank experiment, no oxidation reaction occurred when the reaction was conducted in the absence of phosphinate promoter (entry 16). With the optimal reaction conditions in hand (Table 1, entry 5), the substrate scope and limitations of the aerobic oxidation were investigated, where various alkyl aromatics were subjected to the standard conditions, and these results are summarized in Table 2. First, ethylbenzenes modified with either electron-rich

residual tolerance for pharmaceuticals, it is highly desirable to synthesize them under transition-metal-free conditions. Recently, Xu, Tan, and Wang reported the KOtBu-promoted oxidation of benzylic methylenes to diarylketones with dioxygen under metal-free conditions. Unfortunately, the reaction substrate scope was relatively limited,t and it only worked with diarylmethanes (Scheme 1c).52 Therefore, developing an eco-friendly and general protocol for the oxidation of alkyl (hetero)aromatics and diarylmethanes using O2 as the sole oxidant would be of great synthetic value in organic chemistry and the pharmaceutical industry. In 2012, Xiang’s group reported the Mn(II)/Co(II)/ PhP(O)HOnBu-cocatalyzed oxidative dimerization of styrenes with dioxygen as the oxidant.53 To the best of our knowledge, there have been no reports of alkyl phenylphosphinatepromoted oxidation of alkyl arenes using O2 as the sole oxidant. In line with our ongoing interest in green chemistry,54−60 herein, we report an eco-friendly protocol for the clean preparation of diverse aromatic ketones through the aerobic oxidation of a series of (hetero)benzylic methylenes with atmospheric dioxygen as the oxidant under metal-, base-, additive-, and organic-solvent-free conditions (Scheme 1d). This developed catalytic system is general and efficient with remarkable functional-group compatibility even toward 1,4diacetylbenzene and anthracenedione.

Table 2. Reaction Scopea



RESULTS AND DISCUSSION Initially, oxidation of ethylbenzene 1a with atmospheric dioxygen was performed in the presence of 3 equiv of methyl phenylphosphinate [PhP(O)H(OMe)] at 130 °C. To our delight, the expected acetophenone 2a was formed in 32% GCMS yield (Table 1, entry 1). As anticipated, the oxidation efficiency was found to be strongly dependent on the nature of the promoter (entries 2−10). Among all of the promoters analyzed, we found that phenylphosphinate motifs promoted the desired transformation, with n-octyl phenylphosphinate Table 1. Optimization of Reaction Conditionsa

entry

promoter

temp.

yieldb(%)

1 2 3 4 5 6 7 8 9 10 11 12 13c 14 15 16

PhP(O)H(OMe) (3 equiv) PhP(O)H(OEt) (3 equiv) PhP(O)H(On-Pr) (3 equiv) PhP(O)H(On-penty) (3 equiv) PhP(O)H(On-octyl) (3 equiv) PhP(O)H(On-decyl) (3 equiv) PhP(O)(OEt)2 (3 equiv) P(OEt)3 (3 equiv) P(O)Et3 (3 equiv) PEt3 (3 equiv) PhP(O)H(On-octyl) (4 equiv) PhP(O)H(On-octyl) (2 equiv) PhP(O)H(On-octyl) (3 equiv) PhP(O)H(On-octyl) (3 equiv) PhP(O)H(On-octyl) (3 equiv)

130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 130 °C 140 °C 120 °C 130 °C

32 41 47 86 90 88 N.R. N.R. N.R. N.R. 90 81 41 89 80 N.D.

All reactions were carried out in a 5 mL round-bottom flask in the presence of 1 (0.6 mmol), PhP(O)H(On-octyl) (1.8 mmol), O2 balloon, 130 °C. The reaction time is usually 15 h; isolated yields are reported. a

Conditions: 1a (0.3 mmol), promoter, 130 °C, O2 balloon, 15 h. b Yield was determined by GC-MS. cAtmospheric air was used instead of O2. a

10294

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

Research Article

ACS Sustainable Chemistry & Engineering or electron-poor substituents at the para position of the phenyl ring underwent the oxidation smoothly to afford the expected oxygenation products in good to excellent yields (2a−2p). Higher steric hindrance of the substituent of ethylbenzene substrate indicated reduced reactivity; hence, a relatively lower yield of oxidation product was obtained (2b vs 2q−2r). We were pleased to find that the oxidation conditions could tolerate sensitive functional groups, such as base-labile silyl ester group (2f) and oxidizable sensitive aldehyde (2m) groups. In previous reported oxidation reactions it was hard to construct dioxygenated products because of the electronwithdrawing effect of the carbonyl groups on the monoketones which are less reactive than their dialkyl arene precursors. Pleasingly, the expected double- oxidation product (2s) was obtained smoothly in 90% yield under the standard reaction conditions. Both 1-ethylnaphthalene and 2-ethylnaphthalene smoothly underwent the oxidation process to form the products 2t and 2u in 92% and 84% yield, respectively. It should be noted that O-, S-, and N-containing heteroaromatic motifs were first applied under the base- and metal-free conditions to yield the desired products (2v−2x) with good yields. Delightfully, a series of β-substituted alkyl benzenes could be oxidized to generate the desired ketones in good yield. The alkyl chain lengths and isomeric structures did not significantly affect the reaction outcome (2y−2aa). Furthermore, (haloalkyl)benzenes could be converted exclusively into the desired haloalkylketones (2ab−2ad). Notably, substituted diarylmethanes were also found to be suitable reaction partners, and the corresponding oxidation products were isolated in good yields (2ae−2aj). We were pleased to find that various (hetero)aromatic hydrocarbons bearing a methylene moiety in the (hetero)cyclic framework underwent oxidation to furnish the aromatic cyclic ketones in good yields (2ak− 2ap), thus further enhancing the reaction scope of the developed transformation. Moreover, the anthracene-9,10dione (2aq) could be selectively prepared from the readily available 9,10-dihydroanthracene raw material. The limitation of the present reaction was also realized. Prolonging the reaction time to 25 h, no double-oxidation product was detected in the oxidation of 2,3-dihydro-1H-indene and 1,4dihydronaphthalene. Treatment of methylbenzene under the optimized reaction conditions could not deliver any oxidation product. When 4-ethylbenzonitrile, 1-ethyl-4-nitrobenzene, 3ethyl-1H-indole, or 2-ethyl-1H-pyrrole was employed as the substrate, no oxidation reaction occurred. Ethylbenzenes bearing easily oxidized groups, such as methyl sulfide (1ar) and alkene (1as), are not suitable as substrates in the current oxidation reaction. Both the easily oxidized and the ethyl groups were oxidized and gave the double-oxidation products (2ar and 2s′) in moderate to good yields. To prove the practicality of the developed aerobic oxidation reaction, a scale-up reaction of p-diethylbenzene (1s,8 mmol) was carried out under the standard reaction conditions (Scheme 2a). Pleasingly, the metal-free oxidation provided the desired p-diacetylbenzene 2s in 89% yield, showing a promising application potential for pharmaceutical industry scale-up. The late-stage structural modification of drugs is an important strategy for pharmaceutical research. Thus, three drug derivatives were submitted to the standard conditions (Scheme 2b). All of them could furnish the corresponding ketones (2at−2av). Considering that this type of aerobic oxidation of ethylbenzene may involve the α-phenylethanol (3a) intermediate,

Scheme 2. Gram-Scale Synthesis and Late-Stage Structural Modification

we attempted to oxidize 3a under standard conditions, and an 88% yield of 2a was detected (Scheme 3a). This result proved that α-phenylethanol is the likely reaction intermediate for the developed oxidation. The addition of radical scavenger (0.1 or 2 equiv of TEMPO or BHT) to the aerobic oxidation led to no conversion of ethylbenzene, indicating that a radical pathway might be involved in the present oxidation reaction (Scheme 3b). When anhydrous TBHP or H2O2 was applied as the sole Scheme 3. Control Experiments

10295

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

Research Article

ACS Sustainable Chemistry & Engineering Scheme 4. Plausible Reaction Mechanism

conditions and easy construction of dioxygenation products; (e) the present oxidation reaction could be easily scaled up; (f) the general and metal-free oxidation reaction allows for production of pharmaceutical and bioactive polycyclic aromatic ketones, xanthenone, and dihydroquinolinone from easily available starting materials. The cleanness and simplicity of the present reaction will make it attractive in the pharmaceutical industry.

oxidant under nitrogen atmosphere, only a trace of acetophenone was observed (Scheme 3c). These results suggested that molecular oxygen played a critical role in the oxidation reaction. By using 18O2 as the sole oxidant, oxidation of 1a afforded the 18O-labeled 2a, which confirmed that the oxygen atom of carbonyl originated from atmospheric dioxygen (Scheme 3d). When the reaction was performed in the presence of 2 equiv of H218O, no 18O-2a was detected (Scheme 3e). An intermolecular kinetic isotope effect (KIE) of 1.2 was obtained in a 1:1 mixture of ethylbenzene 1a and ethylbenzene-D10 (1a-D10) (Scheme 3f), suggesting that cleavage of the methylene sp3 C−H bond might not be involved in the rate-limiting step in the present oxidation reaction. Furthermore, decomposition of n-octyl phenylphosphinate affording phenylphosphinic acid and di-n-octyl phenylphosphonate was observed by using 31P NMR experiments (Figures S1 and S2, Supporting Information). On the basis of the above-mentioned observations and previous reports,42,52 a possible mechanistic pathway for the developed oxidation reaction was proposed in Scheme 4. First, the phosphorus-center radical B was generated in situ from the oxidation of alkyl phenylphosphinate A with molecular oxygen. Second, the radical B abstracted a hydrogen atom from the benzylic C−H bond of ethylbenzene 1 to form the radical intermediate C, which was trapped by molecular oxygen to deliver a peroxide radical intermediate D. Then the intermediate D abstracted a proton from alkyl phenylphosphinate A to yield a molecule of E, which underwent βH cleavage and elimination of a molecule of water to generate the desired product 2. Compound E may also be converted to the benzyl alcohol F (detected by GC-MS), which was followed by dioxygen oxidation, affording the target ketone 2.



EXPERIMENTAL SECTION



ASSOCIATED CONTENT

General Procedure for the Synthesis of Ketones 2. A mixture of ethylbenzene 1 (0.6 mmol) and n-octyl phenylphosphinate (457 mg, 1.8 mmol) was added to a 5 mL round-bottom flask with an O2 balloon at room temperature; then the contents were stirred at 130 °C. The reaction typically took 15 h. The progress of the reaction was monitored by TLC or GC-MS. Upon completion, the reaction was cooled down to room temperature. The resultant residue was purified by silica gel column chromatography to afford the desired product 2. S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.9b00002. 1 H and 13C NMR spectra of compounds 2 (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Wei-Min He: 0000-0002-9481-6697



Notes

The authors declare no competing financial interest.



CONCLUSIONS In conclusion, a facile and eco-friendly protocol for the preparation of various (hetero)aromatic ketones has been developed. In the presence of 3 equiv of PhP(O)H(On-octyl), 35 examples of (hetero)arylalkyl ketones, 6 examples of diaryl ketones, and 7 examples of aromatic cyclic ketones were efficiently prepared with dioxygen as the sole oxidant at 130 °C under metal-, base-, additive-, and organic-solvent-free conditions. The principal advantages of this process are outlined as follows: (a) the abundance and easy accessibility of raw materials; (b) water as the sole side product; (c) phosphinate has a dual role as promoter and reaction medium; (d) the oxidation reaction proceeded with excellent functionalgroup tolerance, as proved by the fact that base- and oxidantsensitive groups were intact under the optimal reaction

ACKNOWLEDGMENTS We are grateful for financial support from the Hunan Provincial Natural Science Foundation of China (Nos. 2019JJ40090 and 2019JJ20008).



REFERENCES

(1) Erve, J. C. L.; Gauby, S.; Maynard, J. W.; 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. (2) Choi, E. J.; Kim, G.-H. The antioxidant activity of daidzein metabolites, O-desmethylangolensin and equol, in HepG2 cells. Mol. Med. Rep. 2014, 9, 328−332.

10296

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

Research Article

ACS Sustainable Chemistry & Engineering (3) Xie, W.; Wu, Y.; Zhang, J.; Mei, Q.; Zhang, Y.; Zhu, N.; Liu, R.; Zhang, H. Design, synthesis and biological evaluations of novel pyridone-thiazole hybrid molecules as antitumor agents. Eur. J. Med. Chem. 2018, 145, 35−40. (4) Liang, Q.; Zhang, Y.; Zeng, M.; Guan, L.; Xiao, Y.; Xiao, F. The role of IP3R-SOCCs in Cr(vi)-induced cytosolic Ca2+ overload and apoptosis in L-02 hepatocytes. Toxicol. Res. 2018, 7, 521−528. (5) Moon, S.; Ku, B.-C.; Emrick, T.; Coughlin, B. E.; Farris, R. J. Flame resistant electrospun polymer nanofibers from deoxybenzoinbased polymers. J. Appl. Polym. Sci. 2009, 111, 301−307. (6) 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 2benzoylpyridine-derived thiosemicarbazones: Cytotoxicity against human leukemia cell lines. Eur. J. Med. Chem. 2010, 45, 3904−3910. (7) Xie, W.; Xie, S.; Zhou, Y.; Tang, X.; Liu, J.; Yang, W.; Qiu, M. Design and synthesis of novel 5,6-disubstituted pyridine-2,3-dione-3thiosemicarbazone derivatives as potential anticancer agents. Eur. J. Med. Chem. 2014, 81, 22−27. (8) Choudhary, U.; Mir, A. A.; Emrick, T. Soluble, AllylFunctionalized Deoxybenzoin Polymers. Macromolecules 2017, 50, 3772−3778. (9) Shi, Z.; Zhang, C.; Tang, C.; Jiao, N. Recent advances in transition-metal catalyzed reactions using molecular oxygen as the oxidant. Chem. Soc. Rev. 2012, 41, 3381−3430. (10) Urgoitia, G.; SanMartin, R.; Herrero, M.; Domínguez, E. Recent Advances in Homogeneous Metal-Catalyzed Aerobic C&#-H Oxidation of Benzylic Compounds. Catalysts 2018, 8, 640−661. (11) Dohi, T.; Takenaga, N.; Goto, A.; Fujioka, H.; Kita, Y. Clean and Efficient Benzylic C−H Oxidation in Water Using a Hypervalent Iodine Reagent: Activation of Polymeric Iodosobenzene with KBr in the Presence of Montmorillonite-K10. J. Org. Chem. 2008, 73, 7365− 7368. (12) Yin, L.; Wu, J.; Xiao, J.; Cao, S. Oxidation of benzylic methylenes to ketones with Oxone-KBr in aqueous acetonitrile under transition metal free conditions. Tetrahedron Lett. 2012, 53, 4418− 4421. (13) Hussain, H.; Al-Harrasi, A.; Green, I. R.; Ahmed, I.; Abbas, G.; Rehman, N. U. meta-Chloroperbenzoic acid (mCPBA): a versatile reagent in organic synthesis. RSC Adv. 2014, 4, 12882−12917. (14) Xiang, M.; Xin, Z.-K.; Chen, B.; Tung, C.-H.; Wu, L.-Z. Exploring the Reducing Ability of Organic Dye (Acr+-Mes) for Fluorination and Oxidation of Benzylic C(sp3)-H Bonds under Visible Light Irradiation. Org. Lett. 2017, 19, 3009−3012. (15) Hu, Y.; Zhou, L.; Lu, W. Transition-Metal- and Halogen-Free Oxidation of Benzylic sp3 C-H Bonds to Carbonyl Groups Using Potassium Persulfate. Synthesis 2017, 49, 4007−4016. (16) Jiang, H.; Tang, X.; Xu, Z.; Wang, H.; Han, K.; Yang, X.; Zhou, Y.; Feng, Y.-L.; Yu, X.-Y.; Gui, Q. TBAI-catalyzed selective synthesis of sulfonamides and β-aryl sulfonyl enamines: coupling of arenesulfonyl chlorides and sodium sulfinates with tert-amines. Org. Biomol. Chem. 2019, 17, 2715−2720. (17) Wang, Y.-H.; Ouyang, B.; Qiu, G.; Zhou, H.; Liu, J.-B. Oxidative oxy-cyclization of 2-alkynylbenzamide enabled by TBAB/ Oxone: switchable synthesis of isocoumarin-1-imines and isobenzofuran-1-imine. Org. Biomol. Chem. 2019, 17, 4335−4341. (18) Chen, X.; Li, X.; Chen, X.-L.; Qu, L.-B.; Chen, J.-Y.; Sun, K.; Liu, Z.-D.; Bi, W.-Z.; Xia, Y.-Y.; Wu, H.-T.; Zhao, Y.-F. A one-pot strategy to synthesize β-ketophosphonates: silver/copper catalyzed direct oxyphosphorylation of alkynes with H-phosphonates and oxygen in the air. Chem. Commun. 2015, 51, 3846−3849. (19) Zhu, X.; Li, P.; Shi, Q.; Wang, L. Thiyl radical catalyzed oxidation of diarylalkynes to α-diketones by molecular oxygen under visible-light irradiation. Green Chem. 2016, 18, 6373−6379. (20) Sun, K.; Wang, X.; Zhang, C.; Zhang, S.; Chen, Y.; Jiao, H.; Du, W. Amidoselenation and Amidotelluration of Alkenes using Oxygen as Terminal Oxidant. Chem. - Asian J. 2017, 12, 713−717. (21) Liu, Y.; Xue, D.; Li, C.; Xiao, J.; Wang, C. Reactions catalyzed by a binuclear copper complex: selective oxidation of alkenes to carbonyls with O2. Catal. Sci. Technol. 2017, 7, 5510−5514.

(22) Cui, H.; Wei, W.; Yang, D.; Zhang, Y.; Zhao, H.; Wang, L.; Wang, H. Visible-light-induced selective synthesis of sulfoxides from alkenes and thiols using air as the oxidant. Green Chem. 2017, 19, 3520−3524. (23) Song, S.; Zhang, Y.; Yeerlan, A.; Zhu, B.; Liu, J.; Jiao, N. Cs2CO3-Catalyzed Aerobic Oxidative Cross-Dehydrogenative Coupling of Thiols with Phosphonates and Arenes. Angew. Chem., Int. Ed. 2017, 56, 2487−2491. (24) Liu, B.; Jin, F.; Wang, T.; Yuan, X.; Han, W. Wacker-Type Oxidation Using an Iron Catalyst and Ambient Air: Application to Late-Stage Oxidation of Complex Molecules. Angew. Chem., Int. Ed. 2017, 56, 12712−12717. (25) Fu, Q.; Yi, D.; Zhang, Z.; Liang, W.; Chen, S.; Yang, L.; Zhang, Q.; Ji, J.; Wei, W. Copper-catalyzed aerobic oxidative coupling of ketones with P(O)-H compounds leading to β-ketophosphine oxides. Org. Chem. Front. 2017, 4, 1385−1389. (26) Yi, D.; Fu, Q.; Chen, S.-Y.; Gao, M.; Yang, L.; Zhang, Z.-J.; Liang, W.; Zhang, Q.; Ji, J.-X.; Wei, W. Copper-catalyzed direct hydroxyphosphorylation of electron-deficient alkenes with Hphosphine oxides and dioxygen. Tetrahedron Lett. 2017, 58, 2058− 2061. (27) Xiang, L.; Yang, K.; Song, Q.-L. Pd-catalyzed ortho-olefination of aromatic acetyl esters. Chin. Chem. Lett. 2017, 28, 517−520. (28) Liang, W.; Zhang, Z.; Yi, D.; Fu, Q.; Chen, S.; Yang, L.; Du, F.; Ji, J.; Wei, W. Copper-Catalyzed Direct Oxyphosphorylation of Enamides with P(O)-H Compounds and Dioxygen. Chin. J. Chem. 2017, 35, 1378−1382. (29) Niu, H.-Y.; Su, L.-Y.; Bai, S.-X.; Li, J.-P.; Feng, X.-L.; Guo, H.M. Synthesis of C8-alkyl-substituted purine analogues by direct alkylation of 8-H purines with tetrahydrofuran catalyzed by CoCl2· 6H2O. Chin. Chem. Lett. 2017, 28, 105−108. (30) Guo, T.; Wei, X.-N.; Wang, H.-Y.; Zhu, Y.-L.; Zhao, Y.-H.; Ma, Y.-C. Copper-catalyzed chalcogenation of imidazoheterocycles with sulfur/selenium powder and coumarinyl triflates. Org. Biomol. Chem. 2017, 15, 9455−9464. (31) Jin, W.; Zheng, P.; Wong, W.-T.; Law, G.-L. Efficient SeleniumCatalyzed Selective C(sp3)−H Oxidation of Benzylpyridines with Molecular Oxygen. Adv. Synth. Catal. 2017, 359, 1588−1593. (32) Wei, W.; Cui, H.; Yue, H.; Yang, D. Visible-light-enabled oxyazidation of alkenes leading to α-azidoketones in air. Green Chem. 2018, 20, 3197−3202. (33) Li, R.; Chen, X.; Wei, S.; Sun, K.; Fan, L.; Liu, Y.; Qu, L.; Zhao, Y.; Yu, B. A Visible-Light-Promoted Metal-Free Strategy towards Arylphosphonates: Organic-Dye-Catalyzed Phosphorylation of Arylhydrazines with Trialkylphosphites. Adv. Synth. Catal. 2018, 360, 4807−4813. (34) Guo, T.; Liang, J.-J.; Yang, S.; Chen, H.; Fu, Y.-N.; Han, S.-L.; Zhao, Y.-H. Palladium-catalyzed oxidative C-H/C-H cross-coupling of imidazopyridines with azoles. Org. Biomol. Chem. 2018, 16, 6039− 6046. (35) Liu, Q.; Wang, L.; Yue, H.; Li, J.-S.; Luo, Z.; Wei, W. Catalystfree visible-light-initiated oxidative coupling of aryldiazo sulfones with thiols leading to unsymmetrical sulfoxides in air. Green Chem. 2019, 21, 1609−1603. (36) Li, G.-H.; Dong, D.-Q.; Yang, Y.; Yu, X.-Y.; Wang, Z.-L. Direct Carbamoylation of Quinoline N-oxides with Hydrazinecarboxamides via C−H Bond Activation Catalyzed by Copper Catalyst. Adv. Synth. Catal. 2018, 361, 832−835. (37) Zeng, F.-L.; Chen, X.-L.; He, S.-Q.; Sun, K.; Liu, Y.; Fu, R.; Qu, L.-B.; Zhao, Y.-F.; Yu, B. Copper-catalyzed one-pot three-component thioamination of 1,4-naphthoquinone. Org. Chem. Front. 2019, 6, 1476−1480. (38) Cheng, Z.; Jin, W.; Liu, C. B2pin2-catalyzed oxidative cleavage of a CC double bond with molecular oxygen. Org. Chem. Front. 2019, 6, 841−845. (39) Zhang, J.; Xie, W.; Ye, S.; Wu, J. Synthesis of β-hydroxysulfones through a copper(II)-catalyzed multicomponent reaction with the insertion of sulfur dioxide. Org. Chem. Front. 2019, DOI: 10.1039/ C9QO00520J. 10297

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298

Research Article

ACS Sustainable Chemistry & Engineering (40) Yi, H.; Bian, C.; Hu, X.; Niu, L.; Lei, A. Visible light mediated efficient oxidative benzylic sp3 C-H to ketone derivatives obtained under mild conditions using O2. Chem. Commun. 2015, 51, 14046− 14049. (41) Patil, R. D.; Fuchs, B.; Taha, N.; Sasson, Y. Solvent-free and Selective Autooxidation of Alkylbenzenes Catalyzed by Co/NHPI under Phase Transfer Conditions. Chemistry Select 2016, 1, 3791− 3796. (42) Miao, C.; Zhao, H.; Zhao, Q.; Xia, 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. (43) Mühldorf, B.; Wolf, R. C-H Photooxygenation of Alkyl Benzenes Catalyzed by Riboflavin Tetraacetate and a Non-Heme Iron Catalyst. Angew. Chem., Int. Ed. 2016, 55, 427−430. (44) Hruszkewycz, D. P.; Miles, K. C.; 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. (45) 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. (46) Fan, S.; Dong, W.; Huang, X.; Gao, H.; Wang, J.; Jin, Z.; Tang, J.; Wang, G. In Situ-Induced Synthesis of Magnetic Cu-CuFe2O4@ HKUST-1 Heterostructures with Enhanced Catalytic Performance for Selective Aerobic Benzylic C-H Oxidation. ACS Catal. 2017, 7, 243− 249. (47) Liu, J.; Hu, K.-F.; Qu, J.-P.; Kang, Y.-B. Organopromoted Selectivity-Switchable Synthesis of Polyketones. Org. Lett. 2017, 19, 5593−5596. (48) Wu, J.; Liu, Y.; Ma, X.; Liu, P.; Gu, C.; Dai, B. Cu(II)Catalyzed Ligand-Free Oxidation of Diarylmethanes and Second Alcohols in Water. Chin. J. Chem. 2017, 35, 1391−1395. (49) Sun, W.; Gao, L.; Feng, X.; Sun, X.; Zheng, G. Highly Efficient Catalytic Oxidation of Inert C(sp3)-H Bonds by the Synergistic Effect of Copper Nanoparticles/N-Doped Graphene. Eur. J. Org. Chem. 2018, 2018, 1121−1129. (50) 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. (51) Lesieur, M.; Genicot, C.; Pasau, P. Development of a Flow Photochemical Aerobic Oxidation of Benzylic C-H Bonds. Org. Lett. 2018, 20, 1987−1990. (52) Wang, H.; Wang, Z.; Huang, H.; Tan, J.; Xu, K. KOtBuPromoted Oxidation of (Hetero)benzylic Csp3-H to Ketones with Molecular Oxygen. Org. Lett. 2016, 18, 5680−5683. (53) Liu, K.; Wang, Y.; Li, P.; Cai, W.; Xiang, J. Dimerization of styrenes catalyzed by PhP(O)HOR/Mn(II)/Co(II)/O2. Catal. Commun. 2012, 20, 107−110. (54) Xie, L.-Y.; Duan, Y.; Lu, L.-H.; Li, Y.-J.; Peng, S.; Wu, C.; Liu, K.-J.; Wang, Z.; He, W.-M. Fast, Base-Free and Aqueous Synthesis of Quinolin-2(1H)-ones under Ambient Conditions. ACS Sustainable Chem. Eng. 2017, 5, 10407−10412. (55) Sheng, Y.; You, Y.; Cao, Z.; Liu, L.; Wu, H.-C. Rapid and selective DNA-based detection of melamine using α-hemolysin nanopores. Analyst 2018, 143, 2411−2415. (56) Bao, W.-H.; He, M.; Wang, J.-T.; Peng, X.; Sung, M.; Tang, Z.; Jiang, S.; Cao, Z.; He, W.-M. Iodine-Catalyzed Odorless Synthesis of S-Thiocarbamates with Sulfonyl Chlorides as a Sulfur Source. J. Org. Chem. 2019, 84, 6065−6071. (57) Lu, L.-H.; Wang, Z.; Xia, W.; Cheng, P.; Zhang, B.; Cao, Z.; He, W.-M. Sustainable routes for quantitative green selenocyanation of activated alkynes. Chin. Chem. Lett. 2019, 30, 1237−1240. (58) You, Y.; Zhou, K.; Guo, B.; Liu, Q.; Cao, Z.; Liu, L.; Wu, H.-C. Measuring Binding Constants of Cucurbituril-Based Host-Guest Interactions at the Single-Molecule Level with Nanopores. ACS Sens. 2019, 4, 774−779.

(59) Wu, C.; Hong, L.; Shu, H.; Zhou, Q.-H.; Wang, Y.; Su, N.; Jiang, S.; Cao, Z.; He, W.-M. Practical Approach for Clean Preparation of Z-β-Thiocyanate Alkenyl Esters. ACS Sustainable Chem. Eng. 2019, 7, 8798−8803. (60) Xie, L.-Y.; Peng, S.; Fan, T.-G.; Liu, Y.-F.; Sun, M.; Jiang, L.-L.; Wang, X.-X.; Cao, Z.; He, W.-M. Metal-free C3-Alkoxycarbonylation of Quinoxalin-2(1H)-ones with Carbazates as Ecofriendly Ester Sources. Sci. China: Chem. 2019, 62, 460−464.

10298

DOI: 10.1021/acssuschemeng.9b00002 ACS Sustainable Chem. Eng. 2019, 7, 10293−10298