Direct Reductive Quinolyl β-C–H Alkylation by ... - ACS Publications

Jun 12, 2017 - Shudi Zhang , Zhenda Tan , Biao Xiong , Huan Feng Jiang , Min Zhang. Organic & Biomolecular Chemistry 2018 16 (4), 531-535 ...
0 downloads 0 Views 731KB Size
Subscriber access provided by University of Florida | Smathers Libraries

Letter

Direct Reductive Quinolyl #-C-H Alkylation by Multi-Spherical Cavity Carbon-Supported Cobalt Oxide Nanocatalysts Feng Xie, Rong Xie, Jiaxi Zhang, Huanfeng Jiang, Li Du, and Min Zhang ACS Catal., Just Accepted Manuscript • Publication Date (Web): 12 Jun 2017 Downloaded from http://pubs.acs.org on June 12, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Catalysis is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Direct Reductive Quinolyl β-C-H Alkylation by Multi-Spherical Cavity Carbon-Supported Cobalt Oxide Nanocatalysts Feng Xie, Rong Xie, Jia-Xi Zhang, Huan-Feng Jiang, Li Du*, and Min Zhang* Key Lab of Functional Molecular Engineering of Guangdong Province, School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510641, P. R. China. ABSTRACT: Until now, the selective (hetero)aryl C-H alkylation without directing groups’ assistance or pre-installation of functionalities still remains a highly challenging goal. Herein, by developing acid-resistant multi-spherical cavity carbon-supported cobalt oxide nanocatalysts (CoOx/MSCC) and a hydrogen transfer-mediated activation mode for non-activated N-heteroaromatics, we present a direct reductive quinolyl and isoquinolyl β-C-H alkylation with various aldehydes as the alkylating agents. The catalytic transformation features broad substrate scope, good functional tolerance, use of earth-abundant and reusable cobalt catalysts, and no need for pre-functionalizations, demonstrating that the developed nanocatalysts enable to directly functionalize inert N-heteroaryl systems that are difficult to realize by organometallic complexes. KEYWORDS: heterogeneous catalysis, cobalt, multi-spherical cavity carbon, hydrogen transfer, quinolyl β-C-H alkylation

Due to the widespread occurrence of N-heteroaromatics in bioactive molecules, functional materials, dyes, agrochemicals, pharmaceuticals, and natural products, the search for alkylation approaches to access such compounds is of significant importance. Pioneered by the Friedel–Crafts alkylation,1 much attention has been directed towards the development of more efficient alternatives, which mainly involve the hydroarylation of olefins,2 directing-group the assisted N-heteroaryl C-H alkylation,3 4 cross-dehydrogenative alkylation, the addition of strained rings,5 carbene-involved insertion,6 radical coupling7a-d including Minisci-type reactions,7e-g and Catellani-Lauten-type reactions.8 Despite these significant achievements, the success is mainly based on the use of noble metal complex catalysts that are difficult to recover, the need for pre-functionalizations to generate reactive functionalities (such as halogens, -OTf, organometallic centers) or directing groups. However, the direct and site-selective alkylation of N-heteroaromatics without directing groups’ assistance or pre-installation steps still remains elusive and highly challenging.9 In consideration of the interesting functions of 3-alkyl quinolines, we recently reported an aerobic ruthenium-catalyzed dehydrogenative β-C(sp3)-H benzylation of tetrahydroquinolines with aryl aldehydes (Scheme 1, eq.1).10 However, the transformation suffers from a pre-preparation step, via catalytic hydrogenation of the quinoline derivatives or the cycloaddition between aryl nitriles and alkenes,11 to afford the tetrahydroquinolyl starting materials. From the viewpoint of step- and atom-economic concerns, the direct alkylation of quinolyl β-C(sp2)-H bonds with easily available alkylating agents would be highly desirable, but it remains an unresolved goal to date. Inspired by our efforts in creation of N-heterocycles by transfer hydrogenation (TH) strategy12 and the potential of TH in reduction of inert chemical bonds,13 we envisioned a hydrogen transfer-mediated activation mode for non-activated quinolines, and wished to realize the direct reductive quinolyl β-C-H alkylation with aldehydes. As shown in Scheme 2, in the presence of a suitable hydrogen donor (HD) and catalyst

(Cat), the first TH of quinoline 1 affords an allylic amine A and its tautomers (enamine B and imine C). Then, the β-site of B undergoes nucleophilic addition to the electrophilic α-carbon of aldehyde 2 under the assistance of acid additive, thereby releasing the 3-alkenyl intermediate E by elimination of H2O and H+ from the coupling adduct D. Finally, the tautomerization of E would give the β-alkylated quinoline 3.

Scheme 1. Previous Work and the New Attempts

Scheme 2. Envisioned New Synthetic Protocol However, it is worth mentioning that the presence of transition metal complexes with HD can easily form in-situ the highly reactive metal hydride species,12-13 which would unavoidably lead to the occurrence of the second transfer hydrogenation (from C to 1’) to produce the over-hydrogenated tetrahydroquinoline 1’ and its coupling

ACS Paragon Plus Environment

ACS Catalysis

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

by-products such as 3’. Hence, it is not surprising that the utilization of organometallic complexes including the Ru and Ir ones for the β-alkylation reaction of quinoline 1a and benzaldehyde 2a with formic acid as the HD mainly gave the N-substituted tetrahydroquinolines (Scheme 1, 3aa1 and 3aa2). Hence, to achieve selective generation of product 3 (Scheme 2), at least two challenging issues have to be solved: (1) There should be an compatible catalyst system to ensure that the trapping rate of enamine B by aldehyde 2 is much faster than the second TH process, thus suppressing the formation of undesired intermediate 1’. (2) Aldehyde 2 should not be reduced under the TH conditions. In sharp contrast with organometallic catalysis, the heterogeneous metal oxide nanocatalysts show at least three aspects of advantages for the TH reactions: (1) Unlike the metal hydride species, such catalysts preferentially attach the hydrogen onto the surface of catalytic sites, which exhibit relatively low TH efficiency and would be favorable for the capture of enamine (B) by aldehyde 2. (2) The catalysts are unable to undergo oxidative addition to the carbon-halogen bonds, thus prohibiting the hydrodehalogenation of halo-substrates in protonic conditions, and offering the potential for further synthetic diversity. (3) High stability and easy recyclability. In recent years, by adopting specific carbon supporting materials, heterogeneous non-noble metal catalysts (such as Co and Fe) have been elegantly developed, some of them exhibited even better performance than the conventional Pt/C catalysts for electrochemical oxygen reduction reaction (ORR).14 Moreover, the applications on organic transformations, such as the hydrogenation of nitro compounds,15 aerobic dehydrogenation and selective reduction of N-heteroaromatics,16 oxidation of alcohols to esters,17 have also been nicely demonstrated. Nevertheless, the micropores are ubiquitous in the carbon materials, to some extent, which result in diffusion resistance to hinder mass transfer. To overcome such an inherent obstacle, we believe that the exploitation of novel hollow mesoporous spheres carbon materials would offer a promising way, because such materials feature high specific surface area and adjustable pore diameter,18a-c which have been applied as the electrode materials.18d-g Herein, we report the preparation, characterization of novel multi-spherical cavity carbon-supported cobalt oxide nanocatalysts, and describe for the first time their application in direct reductive quinolyl β-C-H alkylation using aldehydes as the alkylating agents.

Page 2 of 6

precursor and aniline as the sources of carbon material. As depicted in Figure 1, spherical SiO2 templates were initially synthesized using a regrowth procedure.19 Next, the Co precursor in ethanol was introduced onto the PVP (polyvinylpyrrolidone) coated SiO2 nanospheres by in-situ impregnating method. The presence of aniline followed by a polymerization step forms in-situ the PVP-Co-PA (PA: polyaniline) co-coated SiO2 nanospheres due to the hydrogen bonding and metal coordinating effects. Finally, the resulting composites were pyrolyzed under argon flow at 800 °C for 2 hour, and the carbonized product was treated with NaOH solution to etch the silica templates, affording the multi-spherical cavity carbon-embedded cobalt oxide catalysts (CoOx/MSCC, see details in the Supporting Information (SI)). The XRD pattern of the CoOx/MSCC (SI, Figure S1) did not present any evident peaks ascribed to cobalt metal except for peaks of carbon, which indicates that the cobalt species in the catalyst are highly dispersed or amorphous. A typical IV isotherm with hysteresis loop arose in the N2 adsorption-desorption isotherms of CoOx/MSCC, which suggests a mesoporous structure (SI, Figure S2). The BET specific surface area shows to be 150 m2g-1, and the corresponding distribution of pore diameters proves the existence of micropores and mesoporous structure (Figure S2). Notably, there is a concentrated distribution of pore size around 20 nm, which represents the multi-spherical cavity structure that makes for substrates transfer into the active sites. a

b

100 nm

c

d

Figure 2. (a) TEM images of CoOx/MSCC. (b) HAADF-STEM image of CoOx/MSCC. EDS elemental maps for combined image (c), Co (d).

Figure 1. Schematic illustration for the preparation of catalyst.

Typically, the new CoOx/MSCC catalyst was prepared by an in-situ hard template method, using Co(OAc)2 as the cobalt

The morphology and structure of CoOx/MSCC was investigated by TEM techniques. The TEM image (Figure 2: a) demonstrates a multi-spherical cavity structure of CoOx/MSCC with cavity size of 21 nm and it evenly distributes in the carbon materials. It is worth noting that no individual cobalt-encompassing nanoparticles were detected in the STEM image (Figure 2: b), suggesting that the cobalt species are highly dispersed, which is in good agreement with the XRD results. In addition, as shown in the elemental mapping results (Figure 2: c-d, and Figure S4 (SI): d-f), the signals of Co, O, N and C are completely overlapped and 2

ACS Paragon Plus Environment

Page 3 of 6

interweaved for each other with nanoscale size, implying that the Co is bonded to heteroatoms prior to aniline polymerization, which facilitates the uniform adsorption of cobalt precursor on the SiO2 nanospheres and prevents it from agglomeration during the carbonization process. To identify the surface chemistry of the developed new catalysts, the X-ray photoelectron spectroscopy (XPS) was then performed. The surface of CoOx/MSCC is dominated by carbon, nitrogen, oxygen and cobalt, the corresponding content is 74.18, 11.83, 9.89 and 4.1 wt %, respectively. As shown in N 1s spectrum of the catalysts (see SI, Figure S5: c), it can be deconvoluted into three peaks with pyridinic N (398.6 eV), graphitic N (400.8 eV).20 The pyridinic N results from coordination with cobalt, while the graphitic N originates from the graphitization of uncoordinated aniline. The spectrum of Co 2p (Figure 3) can be fitted to four components that ascribe to CoO, Co3O4. The first maximum peak is found with typical binding energy of 780.7 eV, which is characteristic of CoO. In addition, two peaks at 786.9 eV and 802.8 eV are the satellite peaks of oxidic Co.20 In the cobalt region, the peak at 796.5 eV assigns to Co3O4 with Co 2p 1/2 electrons.17d In agreement with elemental mapping results, the XPS results suggest that the nanoparticles are cobalt oxides.

electron-withdrawing substituents could enhance the electrophilicity of the aldehyde carbonyl groups, thus favoring the coupling process. In comparison with para- and meta-substituted aldehydes (3ag, 3ah), the ortho-substituted one (2i) afforded the product (3ai) in a lower yield, which is assigned to the effect of steric hindrance. Furthermore, heteroaryl aldehydes were also proven to be effective coupling partners, affording the products (3al-3an) in reasonable yields. Noteworthy, terephthalaldehyde (2e) reacted with 1a in a selective manner, giving product 3ae with retention of one aldehyde group, which is useful for further modifications. CoOx/MSCC (5 mol%) HCOOH (5 eq), p-xylene

+ R 1a

N

N

2

CHO

N

O S

N 3ad, 62%

N R 3, isolated yield%

CF3CO2H (80 mol%) 130 oC, 16 h

3aa, 70%

N

O

3ab, 73%

3ae, 41%

Cl

N 3ac, 63%

CHO

N 3af, 67%

N

N 3ah, 61%

O 3ag, 64%

N

7500

N 3am, 53%

N

O

CN

CH3 O

O

N 3aj, 57%

8000

Counts / s

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

N

3ai, 52%

N N 3al, 42%

3ak, 54%

S N 3an, 55%

7000

Scheme 3. Quinolyl β-Alkylation with Various Aldehydes 6500

6000

CoO Satelite Satelite Co 3O 4

5500 820

810

800

790

780

770

Binding energy/ eV Figure 3. Co 2p XPS spectra of CoOx/MSCC.

To test the catalytic performance of the the prepared CoOx/MSCC material, we chose the coupling of quinoline 1a and benzaldehyde 2a as a model system to evaluate different reaction parameters, including the effects of the catalysts, acidic additives, solvents, hydrogen donors and temperatures (see supporting information (SI), Table S1). An optimal GC yield (82%) of product 3aa with high chemo-selectivity was obtained at 130 oC by using 5 mol% of catalyst, CF3CO2H (0.8 equiv), formic acid (5 equiv) and p-xylene as the additive, hydrogen donor and solvent, respecitively. With the availability of the optimal conditions, we then examined the generality of the synthetic protocol. First, the reactions of 1a in combination with a variety of aldehydes (see SI, Scheme S1) were tested. As shown in Scheme 3, all the reactions proceeded smoothly and furnished the β-alkylated products in moderate to excellent yields upon isolation (see 3aa-3an). Among all the examples presented, it was found that the electronic property of substituents on the aryl ring of benzaldehydes slightly affected the product yields. Specially, electron-deficient groups afforded the products (3ab-3ad) in relatively higher yields than those of electron-donating ones (3af-3aj). This phenomenon can be rationalized as the

Subsequently, we focused on the variation of quinoline derivatives with different types of aldehydes (see SI, Scheme S1). Gratifyingly, all the substrates underwent efficient reductive quinolyl β-C-H alkylation and furnished the desired products in moderate to excellent isolated yields (Scheme 4). Except for the substituent influence of aldehydes as described in Scheme 3, the electronic property of substituents on the quinolines significantly affected the product formation. In general, the electron-donating group containing quinolines afforded the desired product in higher yields (see 3bb, 3bf, 3da-3dd) than the electron-poor one (3eb), which is ascribed to the electron-donating groups could enhance the nucleophilicity of the enamines arising from the first transfer hydrogenation of the quinolines 1, thus favoring the cross-coupling process with aldehydes (Scheme 2). Noteworthy, alkyl aldehydes such as heptanal (2p) and cyclohexanecarboxaldehyde (2q) were also amenable to the transformation, albeit the yields were somewhat low (3dp, 3dq) due to the occurrence of decarbonylation. Interestingly, the polycyclic quinoline (1f) and isoquinolines (1g and 1h) effectively coupled with different aldehydes, affording the alkylated products in reasonable yields (3fb, 3fj, 3ga, 3gb, 3gf and 3hf). However, the alkylation of the challenging pyridine (1i) was not applicable with our catalyst system. It is worth mentioning that a wide range of functional groups (such as -Cl, -CN, -SO2Me, -CHO, -Me, -OMe, -N(Me)2, -CF3, ester) are well tolerated in the developed new reaction (Scheme 3 and 4), thus offering the potential for molecular complexity via further transformations. Moreover, all the products possess a benzylic site, which can be employed for fabrication of ketones via direct C-H oxidation.21 3

ACS Paragon Plus Environment

ACS Catalysis

4% yields, respectively (Scheme 5, eq.1). Then, treating quinoline (1a) with benzylic alcohol (2a’) under standard conditions gave only 4% of product 3aa (eq.2). Further investigations showed that only addition of CF3COOH was sufficient to furnish the coupling of the prepared mixture of enamine (B1a) and imine (C1a) with benzaldehyde 2a, giving product 3aa in 81% yield (eq.3). These results clearly show that the reaction involving tetrahydroquinoline 1a’ and alcohol 2a’ as the main intermediates is less likely, and the formation of 3aa is contributed by the coupling of B1a and C1a with aldehyde 2a, which is in agreement with the pathway proposed in Scheme 2. Moreover, by replacing HCO2H with DCO2D in the standard conditions, the reaction of 8-methylquinoline (1d) with 4-chlorobenzaldehyde (2b) gave product 3db-dn in 57% yield with different deuterium ratios at position -2, -4

and the benzylic site, supporting that the reaction undergoes a transfer hydrogenative coupling pathway. (Scheme 5, eq.4).

Scheme 4. Both Variations of Quinolines and Aldehydes Further, we examined the stability and reusability of our developed new catalysts CoOx/MSCC pyrolyzed at 800 oC. As shown in Figure 4, five more recycles were conducted with the model reaction, it indicates that the catalytic activity maintains very well in these five consecutive runs, albeit with slight decline of yield, which is accounted for the catalyst loss during the processes of recycling and mechanical abrasion (SI, Table S2). Noteworthy, the obstacle of using formic acid and trifluoroacetic acid is that they can react with base metals such as cobalt and its oxide composites. However, the cobalt oxide nanoparticles embedded in the multi-spherical cavity carbon material exhibit to be resistant to acids and keep its inherent activity without agglomeration of nanoparticles. As shown in Figure S3: c (see SI), the TEM study clearly presents that the catalyst remains its morphology even after five reuses. 90

3aa

80 70 60

GC Yield (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 6

50 40 30 20 10 0 0

1

2

3

4

5

Recycle number

Figure 4. Reuse of the CoOx/MSCC catalysts.

To gain insight into the reaction information, several verification experiments were performed. First, under standard conditions, the reaction of tetrahydroquinoline 1a’ with benzaldehyde 2a gave 3aa, N-formylated (3aa1) and the Bruneau-type 1,3-dialkylated product (3aa3)22 in 1%, 90% and

Scheme 5. Verification Experiments Finally, a time-concentration profile of the model reaction was depicted in Figure S6 (SI). 1a was fully consumed within 20 h to afford a maximum yield of 3aa (82%) along with small portion of N-formylated product 3aa1. Enamine (B1a) and imine (C1a) were not detected, showing that the capture of both intermediates by aldehyde 2a is much faster than the second TH process that leads to generation of by-product 3aa1. In addition, benzylic alcohol 2a’ was also not detected in the whole process, presumably because quinoline 1a, rather than aldehyde 2a, preferentially interacts with the nanocatalysts and leads to selective hydrogen transfer to quinoline 1a. In summary, by developing novel and acid-resistant multi-spherical cavity carbon-supported cobalt oxide nanocatalysts (CoOx/MSCC) and a hydrogen transfer-mediated activation mode for non-activated N-heteroaromatics, we have demonstrated a direct reductive β-C-H alkylation of non-activated quinolines and isoquinolines using aldehydes as the alkylating agents for the first time. The catalytic transformation features broad substrate scope, excellent functional tolerance, use of earth-abundant, reusable and acid-resistant cobalt metal catalysts, and no need for pre-functionalization steps. Further development of MSCC-supported nanocatalysts and exploitation of their 4

ACS Paragon Plus Environment

Page 5 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

applications including effective functionalization of other N-heteroaryl systems are curretly underway.

(8)

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: The detail of catalyst preparation, more characterization results, complete experimental procedures and spectral data (PDF)

(9) (10) (11)

AUTHOR INFORMATION (12)

Corresponding Author * [email protected] * [email protected]

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We thank the Science and Technology Program of GuangZhou (201607010306), National Natural Science Foundation of China (21472052), Science Foundation for Distinguished Young Scholars of Guangdong Province (2014A030306018) and ‘‘1000 Youth Talents Plan’’ for financial support.

(13)

(14)

REFERENCES (1) Selected examples, see: (a) Ge, C.; Liu, R.-R.; Gao, J.-R.; Jia, Y.-X. Org. Lett. 2016, 18, 3122-3125. (b) García-García, C.; Ortiz-Rojano, L.; Álvarez, S.; Álvarez, R.; Ribagorda, M.; Carreño, M. C. Org. Lett. 2016, 18, 2224-2227. (2) Selected examples, see: (a) Dong, Z.; Ren, Z.; Thompson, S. J.; Xu, Y.; Dong, G. Chem. Rev. 2017, DOI: 10.1021/acs.chemrev.6b00574, in press. (b) Rit, R. K.; Ghosh, K.; Mandal, R.; Sahoo, A. K. J. Org. Chem. 2016, 81, 8552-8560. (c) Lee, S. Y.; Villani-Gale, A.; Eichman, C. C. Org. Lett. 2016, 18, 5034-5037. (d) Friis, S. D.; Pirnot, M. T.; Buchwald, S. L. J. Am. Chem. Soc. 2016, 138, 8372-8375. (3) Selected examples, see: (a) Soni, V.; Jagtap, R. A.; Gonnade, R. G.; Punji, B. ACS Catal. 2016, 6, 5666-5672. (b) Sarkar, D.; Gevorgyan, V. Chem. Eur. J. 2016, 22, 11201-11204. (c) Wang, H.; Yu, S.; Qi, Z.; Li, X. Org. Lett. 2015, 17, 2812-2815. (4) Selected examples, see: (a) Salman, M.; Zhu, Z. Q.; Huang, Z. Z. Org. Lett. 2016, 18, 1526-1529. (b) Wan, M.; Lou, H.; Liu, L. Chem. Commun. 2015, 51, 13953-13956. (c) Wang, F.-F.; Luo, C.-P.; Deng, G.; Yang, L. Green Chem. 2014, 16, 2428-2431. (d) Liu, X.; Zhang, J.; Ma, S.; Ma, Y.; Wang, R. Chem. Commun. 2014, 50, 15714-15717. (5) Selected examples, see: (a) Zhou, X.; Yu, S.; Qi, Z.; Kong, L.; Li, X. J. Org. Chem. 2016, 81, 4869-4875. (b) Cheng, G. L.; Li, T. J.; Yu, J. Q. J. Am. Chem. Soc. 2015, 137, 10950-10953. (6) Selected examples, see: (a) Mishra, N. K.; Choi, M.; Jo, H.; Oh, Y.; Sharma, S.; Han, S. H.; Jeong, T.; Han, S.; Lee, S.-Y.; Kim, I. S. Chem. Commun. 2015, 51, 17229-17232. (b) Cui, X.; Xu, X.; Jin, L.-M.; Wojtas, L.; Zhang, X. P. Chem. Sci. 2015, 6, 1219-1224. (7) Selected examples, see: (a) Cheng, W.-M.; Shang, R.; Fu, M.-C.; Fu, Y. Chem. Eur. J. 2017, 23, 2537-2541. (b) Babu, K. R.; Zhu, N.; Bao, H. Org. Lett. 2017, 19, 46-49. (c) Kandukuri, S. R.; Bahamonde, A.; Chatterjee, I.; Jurberg, I. D.; Escudero-Adán, E. C.; Melchiorre, P. Angew. Chem. In. Ed. 2015, 54, 1485-1489. (d) Wu, X.; See, J. W. T.; Xu, K.; Hirao, H.; Roger, J.; Hierso, J.-C.; Zhou, J. Angew. Chem. In. Ed. 2014, 53, 13573-13577. (e) Hara, F. O.; Blackmond, D. G.; Baran, P. S. J. Am. Chem. Soc. 2013, 135, 12122-12134. (f) Fujiwara, Y.; Dixon, J. A.; O’Hara, F.; Funder, E. D.; Dixon, D. D.; Rodriguez, R. A.; Baxter, R. D.; Herle, B.; Sach, N.; Collins, M. R.; Ishihara, Y.; Baran, P. S.

(15)

(16)

(17)

(18)

(19) (20)

Nature 2012, 492, 95-99. (g) Bowman, W. R.; Storey, J. M. D. Chem. Soc. Rev. 2007, 36, 1803-1822. Selected examples, see: (a) Shen, P.-X.; Wang, X.-C.; Wang, P.; Zhu, R.-Y.; Yu, J.-Q. J. Am. Chem. Soc. 2015, 137, 11574-11577. (b) Lei, C.; Jin, X.; Zhou, J. Angew. Chem. Int. Ed. 2015, 54, 13397-13400. Ye, J.; Shi, Z.; Sperger, T.; Yasukawa, Y.; Kingston, C.; Schoenebeck, F.; Lautens, M. Nat. Chem. 2017, 9, 361-368. Tan, Z. D.; Jiang, H. F.; Zhang, M. Org. Lett. 2016, 18, 3174-3177. Cui, X.; Li, Y.; Bachmann, S.; Scalone, M.; Surkus, A. E.; Junge, K.; Topf, C.; Beller, M. J. Am. Chem. Soc. 2015, 137, 10652-10658. (a) Xiong, B.; Jiang, J. X.; Zhang, S. D.; Jiang, H. F.; Ke, Z. F.; Zhang, M. Org. Lett. 2017, 19, 2730-2733. (b) Lv, W.; Xiong, B.; Jiang, H.; Zhang, M. Adv. Synth. Catal. 2017, 359, 1202-1207. (c) Tan, Z.; Jiang, H.; Zhang, M. Chem. Commun. 2016, 52, 9359-9362. (d) Xiong, B.; Zhang, S. D.; Chen, L.; Li, B.; Jiang, H. F.; Zhang, M. Chem. Commun. 2016, 52, 10636– 10639. (e) Xiong, B.; Li, Y.; Lv, W.; Tan, Z. D.; Jiang, H. F.; Zhang, M. Org. Lett. 2015, 17, 4054-4057. (f) Xie, F.; Zhang, M.; Jiang, H. F.; Chen, M. M.; Lv, W.; Zheng, A.; Jian, X. Green Chem. 2015, 17, 279–284. (a) Chen, Z.-W.; Zeng, H.-Y.; Gong, H.; Wang, H.-N.; Li, C.-J. Chem. Sci. 2015, 6, 4174–4178. (b) Chen, Z.-W.; Zeng, H.-Y.; Girard, S. A.; Wang, F.; Chen, N. Li, C.-J, Angew. Chem. Int. Ed. 2015, 54, 14487–14491. Selected examples, see: (a) Deng, D.; Yu, L.; Chen, X.; Wang, G.; Jin, L.; Pan, X.; Deng, J.; Sun, G.; Bao, X. Angew. Chem. Int. Ed. 2013, 52, 371-375. (b) Liang Y.; Li, Y.; Wang, H.; Zhou, J.; Wang, J.; Regier, T.; Dai, H. Nat. Mater. 2011, 10, 780-786. Selected examples, see: (a) Zhou, P.; Jiang, L.; Wang, F.; Deng, K.; Lv, K.; Zhang, Z. Science Adv. 2017, 3, DOI:10.1126/sciadv.1601945, in press. (b) Schwob, T.; Kempe, R. Angew. Chem. Int. Ed. 2016, 55, 15175-15179. (c) Liu, W.; Zhang, L.; Yan, W.; Liu, X.; Yang, X.; Miao, S.; Wang, W.; Wang, A.; Zhang, T. Chem. Sci. 2016, 7, 5758-5764. (d) Jagadeesh, R. V.; Surkus, A.-E.; Junge, H.; Pohl, M.-M.; Radnik, J.; Rabeah, J.; Huan, H.; Schuenemann, V.; Brueckner, A.; Beller, M. Science 2013, 342, 1073-1076. Selected examples, see: (a) Wei, Z.; Chen, Y.; Wang, J.; Su, D.; Tang, M.; Mao, S.; Wang, Y. ACS Catal. 2016, 5816-5822. (b) Chen, F.; Surkus, A. E.; He, L.; Pohl, M. M.; Radnik, J.; Topf, C.; Junge, K.; Beller, M. J. Am. Chem. Soc. 2015, 137, 11718-11724. (a) Su, H.; Zhang, K.-X.; Zhang, B.; Wang, H.-H.; Yu, Q.-Y.; Li, X.-H.; Antonietti, M.; Chen, J.-S. J. Am. Chem. Soc. 2017, 139, 811-818. (b) Zhou, Y. X.; Chen, Y. Z.; Cao, L.; Lu, J.; Jiang, H. L. Chem. Commun. 2015, 51, 8292-8295. (c) Zhong, W.; Liu, H.; Bai, C.; Liao, S.; Li, Y. ACS Catal. 2015, 5, 1850-1856. (d) Jagadeesh, R. V.; Junge, H.; Pohl, M. M.; Radnik, J.; Bruckner, A.; Beller, M. J. Am. Chem. Soc. 2013, 135, 10776-10782. Selected examples, see: (a) Yang, T.; Li, K.; Pu, L.; Liu, Z.; Ge, B.; Pan, Y.; Liu, Y. Biosens. Bioelectron. 2016, 86, 129-134. (b) Lin, X.; Fu, L.; Chen, Y.; Zhu, R.; Wang, S.; Liu, Z. ACS Appl. Mat. Interfaces 2016, 8, 26809-26816. (c) Wang, H.; Bo, X.; Wang, A.; Guo, L. Electrochem. Commun. 2013, 36, 75-79. (d) Tang, J.; Yamauchi, Y. Nat. Chem. 2016, 8, 638-639. (e) Li, Y.; Bastakoti, B. P.; Yamauchi, Y. APL Mater. 2016, 4, 040703-040708. (f) Teng, Z.; Su, X.; Zheng, Y.; Zhang, J.; Liu, Y.; Wang, S.; Wu, J.; Chen, G.; Wang, J.; Zhao, D.; Lu, G. J. Am. Chem. Soc. 2015, 137, 7935-7944. Hartlen, K. D.; Athanasopoulos, A. P. T.; Kitaev, V. Langmuir 2008, 24, 1714-1720. (a) Yang, C.; Fu, L.; Zhu, R.; Liu, Z. Phys. Chem. Chem. Phys. 2016, 18, 4635-4642. (b) Casanovas, J.; Ricart, J. M.; Rubio, J.; Illas, F.; Jimenez-Mateos, J. M. J. Am. Chem. Soc. 1996, 118, 8071-8076.

5

ACS Paragon Plus Environment

ACS Catalysis

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 6

(21) De Houwer, J.; Abbaspour Tehrani, K.; Maes, B. U. W. Angew. Chem., Int. Ed. 2012, 51, 2745-2748. (22) (a) Yuan, K.; Jiang, F.; Sahli, Z.; Achard, M.; Roisnel, T.; Bruneau, C. Angew. Chem. Int. Ed. 2012, 51, 8876-8880. (b) Sundararaju, B.; Achard, M.; Sharma, G. V. M.; Bruneau, C. J. Am. Chem. Soc. 2011, 133, 10340-10343.

A direct reductive β-C-H alkylation of non-activated quinolines and isoquinolines with aldehydes as the alkylating agents, enabled by novel and acid-resistant multi-spherical cavity carbon-supported cobalt oxide nanocatalysts (CoOx/MSCC), has been demonstrated.

6

ACS Paragon Plus Environment