Highly Enantioselective Oxidation of Spirocyclic Hydrocarbons by

Feb 7, 2018 - Similarly, the ee value of the product 2a was low in the reaction of the catalyst 4b (R,R-PDMB-Mn) (Table 1, entry 12). .... On the basi...
0 downloads 6 Views 928KB Size
Subscriber access provided by READING UNIV

Article

Highly Enantioselective Oxidation of Spirocyclic Hydrocarbons by Bioinspired Manganese Catalysts and Hydrogen Peroxide Bin Qiu, daqian xu, Qiangsheng Sun, Chengxia Miao, Yong-Min Lee, Xiao-Xi Li, Wonwoo Nam, and Wei Sun ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b03601 • Publication Date (Web): 07 Feb 2018 Downloaded from http://pubs.acs.org on February 7, 2018

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 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 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 10 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

Bin Qiu,†,‡ Daqian Xu,† Qiangsheng Sun,† Chengxia Miao,† Yong-Min Lee,§ Xiao-Xi Li,§ Wonwoo Nam,*,†,§ and Wei Sun*,† †

State Key Laboratory for Oxo Synthesis and Selective Oxidation, Center for Excellence in Molecular Synthesis, Suzhou Research Institute of LICP, Lanzhou Institute of Chemical Physics (LICP), Chinese Academy of Sciences, Lanzhou 730000, China ‡

University of Chinese Academy of Sciences, Beijing 100049, China

§

Department of Chemistry and Nano Science, Ewha Womans University, Seoul 03760, Korea

ABSTRACT: Bioinspired manganese complexes have emerged as attractive catalysts for a number of selective oxidation reactions over the past several decades. In the present study, we report the enantioselective oxidation of spirocyclic compounds with manganese complexes bearing tetradentate N4 ligands as catalysts and aqueous H2O2 as a terminal oxidant under mild conditions; spirocyclic tetralone (1a) and its derivatives bearing electron-donating and -withdrawing substituents are converted to their corresponding chiral spirocyclic β,β’-diketones with high yields and enantioselectivities. Spirocyclic indanones are also converted to the β,β’-spirobiindanones with high enantioselectivities. Indeed, the reaction expands the diversity of chiral spirocyclic diketones via a late-stage oxidative process. In addition, it is of importance to note that the catalytic reaction can be easily scaled up and the chiral spirocyclic β,β’-diketones can be transformed into diol products. In mechanistic studies, we have shown that (1) ketones were yielded as products via the initial formation of alcohols, followed by the further oxidation of the alcohols to ketones, (2) hydrogen atom (H-atom) abstraction from the methylene C-H bonds of 1a by a putative Mn(V)-oxo intermediate was proposed to be the rate-determining step, and (3) the C-H bond hydroxylation of 1a by the Mn(V)-oxo species was proposed to occur via oxygen rebound mechanism. Based on these results, we have proposed a plausible mechanism for the selective C-H bond oxidation of hydrocarbons by bioinspired manganese catalysts and hydrogen peroxide. KEYWORDS: C-H Oxidation, Manganese Catalyst, Asymmetric Oxidation, Hydrogen Peroxide, Mechanism, Manganese-Oxo

INTRODUCTION The selective oxidation of hydrocarbon C-H bonds is a highly challenging reaction in synthetic organic chemistry, but common in enzymatic reactions.1 For example, metalloenzymes (e.g., Cytochromes P450 and methane monooxygenases) generate highly reactive intermediates (e.g., high-valent metal-oxo species) for the stereo-, regio-, and enantioselective oxidation of organic substrates.2 Therefore, development of efficient and selective oxidation of hydrocarbons, especially for the enantioselective oxidation of C– H bonds, using bioinspired metal catalysts and environmentally benign oxidants has been a long-standing goal in the communities of synthetic organic, oxidation, and biomimetic/bioinorganic chemistry.3 In biomimetic studies, tremendous efforts have been devoted to developing artificial catalysts that mimic the reactivities of metalloenzymes in catalytic oxidation reactions.3-5 In addition, bioinspired metal complexes have been used in the synthesis and characterization of metal-oxygen intermediates (e.g., high-valent metal-oxo species), and reactivities and mechanisms of the synthetic metal-oxygen intermediates have been intensively investigated in various oxidation reactions, with the purpose of elucidating enzymatic reactions in biology.6 One notable example is the synthetic metalloporphyrins used as catalysts in the (enantio)selective oxidation of organic substrates as well as chemical models of Cytochromes P450.3b,7 For example, Groves and co-workers observed a

moderate level of enantioselectivity in the asymmetric hydroxylation of ethylbenzene by a chiral iron porphyrin complex.8 While significant progress has been made in the aliphatic C−H bond oxidation catalysis over the past several decades,9-12 asymmetric oxidation of aliphatic C-H bonds by chiral metal catalysts has been reported only in few cases.13-16 To develop efficient asymmetric C-H bond oxidation catalysis, two major challenges should be overcome, such as (1) the C–H oxidation of hydrocarbons by catalysts yields chiral alcohol products with a poor to moderate enantioselectivity and (2) the chiral alcohol products are further oxidized rapidly to ketones, thereby losing the stereogenic center (Scheme 1A). Fortunately, the problem of the overoxidation of alcohols to the corresponding ketones has the potential application in enantioselective desymmetrization reactions.17 For example, Bach and coworkers reported an elegant study for the oxidative desymmetrization of spirocyclic oxindoles by a chiral ruthenium porphyrin catalyst and 2,6-dichloropyridine N-oxide, in which the methylene C-H bonds were oxidized to yield ketone products with high enantioselectivities (Scheme 1B).18 In the study, enantioselectivities were achieved by building hydrogen bonding between substrates and the porphyrin ligand of the ruthenium complex (Scheme 1B).18 If such an oxidative strategy can be expanded to all-carbon spirocyclic precursors, a catalytic enantioselective oxidation of spirocyclic compounds to their corresponding spiro β,β’-diketone derivatives can be successfully established (Scheme 1C). Although these opti-

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

A. Hydroxylation of benzylic C-H bonds to alcohol and ketone products

B. Enantiotopos-selective C-H bond oxidation by Ru porphyrin complex

Page 2 of 10

(e.g., carboxylic acids or H2SO4);5,23,24 the asymmetric epoxidation reactions are highly efficient in terms of product yields and enantioselectivities. In contrast, there are only a couple of reports for the enantioselective oxidation of aliphatic C-H bonds by nonheme manganese catalysts and H2O2.14,15 Since we have reported an efficient manganese catalytic system using H2O2 as a terminal oxidant for the oxidation of benzylic and aliphatic methylene C-H bonds to ketones under mild reaction conditions,13 we attempted to examine the previously reported catalytic systems (i.e., using a bioinspired manganese catalyst and an environmentally benign oxidant) in the enantioselective synthesis of spirocyclic compounds. Herein, we report an enantioselective oxidation of benzylic methylene C-H bonds in spirocyclic precursors by a manganese catalyst bearing a tetradentate N4 ligand and aqueous H2O2, affording a vast array of chiral spirocyclic β,β’-diketones in high yields with excellent enantioselectivities (Scheme 1C). Mechanistic aspects for the manganese-catalyzed enantioselective C-H bond activation reaction have also been discussed in the present study. RESULTS AND DISCUSSION

C. This work: Site- and enantioselective oxidation of methylene C-H bonds of spirocyclic compounds by Mn catalyst and H2O2

Scheme 1. Strategy for the Enantioselective Oxidation of C-H Bonds in This Study cally pure spirocyclic compounds are important structural motifs that are found not only in many important chiral ligands of asymmetric catalysis but also in biologically active natural products (Figure 1),19-21 catalytic enantioselective methods for the synthesis and functionalization of these chiral spirocyclic compounds remain significantly underdeveloped.22 Recently, a great advance has been achieved in the enantioselective epoxidation of olefins by bioinspired manganese complexes bearing nonheme ligands and H2O2 in the presence of additives

Figure 1. Examples of chiral ligands and natural product containing the spirocyclic structural motif.

To examine the feasibility of our proposed strategy, we conducted the oxidation of methylene C-H bonds of spirocyclic compounds by nonheme manganese complexes and aqueous H2O2 in the presence of carboxylic acids (see Table 1 for the structures of manganese complexes and carboxylic acids and the reaction scheme). First, the oxidation of spirocyclic tetralone (1a) by [(S-PEB)Mn] (3a)13 (2.0 mol%) and H2O2 (7.0 equiv to 1a) at 0 oC afforded spirocyclic diketone (2a) in 80% yield with 87% enantiomeric excess (ee) in the presence of 14 equiv of 2,2-dimethylbutanoic acid (DMBA) at 0 oC (Table 1, entry 1). To optimize the reaction conditions, the oxidation of 1a was carried out at different reaction temperatures with different amounts of catalyst loading and carboxylic acid additive, and the catalytic activity and enantioselectivity of several manganese catalysts were also examined (Table 1). When the reaction was carried out at –30 oC, the ee value of 2a increased to 94% (Table 1, entry 3). Reducing the catalyst loading (0.20 mol% catalyst) also afforded a high enantioselectivity albeit with a moderate yield (Table 1, entry 5). In the presence of 0.50 mol% catalyst 3a, the amount of oxidant could be reduced to 5.0 equiv while maintaining both reactivity and enantioselectivity (Table 1, entry 6). Changing the N-substituent of benzimidazole motif in catalyst 3 to methyl (3b) and isopropyl (3c) slightly decreased the ee value of 2a (Table 1, entries 9 and 10). Compared to the catalyst 3b, the catalyst 4a (R,R-MCMB-Mn)23d bearing a C2symmetric cyclohexane-1,2-diamine backbone afforded a low enantioselectivity (Table 1, entry 11 (75% ee) versus entry 9 (92% ee)). Similarly, the ee value of the product 2a was low in the reaction of the catalyst 4b (R,R-PDMB-Mn) (Table 1, entry 12).23c Since it has been well documented that the reactivity and enantioselectivity of nonheme metal complexes are affected significantly by carboxylic acid additives in the catalytic asymmetric epoxidation of olefins by H2O2,24-27 other carboxylic acids, such as 2-ethylhexanoic acid (EHA) and acetic acid (AcOH), were examined. In these reactions, we found that the ee value of 2a was decreased slightly (e.g., ~90% ee) (Table 1, entries 13 and 14). Finally, it is notable that only trace amounts of products were formed from the methylene C-H bonds of six-membered ring in 1a, demonstrating that the oxidation of

ACS Paragon Plus Environment

Page 3 of 10 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

Table 1. Optimization of Site- and Enantiotopos-selective Oxidation of Methylene C-H Bonds of spirocyclic tetralonea

entry

catalyst (mol%)

H2O2 (equiv)

temp. (oC)

conv.b (%)

yieldc (%)

(%)

eed

1

3a (2.0)

7.0

0

94

80

87

2

3a (2.0)

7.0

–20

88

77

92

3

3a (2.0)

7.0

–30

86

76

94

4

3a (0.50)

7.0

–30

88

74

94

5

3a (0.20)

7.0

–30

60

54

93

6

3a (0.50)

5.0

–30

90

80

94

7

3a (0.50)

3.5

–30

86

74

93

3a (0.50)

5.0

–30

78

56

92

9

3b (0.50)

5.0

–30

86

74

92

10

3c (0.50)

5.0

–30

80

67

86

11

4a (0.50)

5.0

–30

92

75

75h

12

4b (0.50)

5.0

–30

92

78

80h

13f

3a (0.50)

5.0

–30

80

64

89

14g

3a (0.50)

5.0

–30

92

83

90

8

e

Table 2. Products Formed in the Oxidation of Spirocyclic Tetralone Derivatives (1a – 1i)a,b

a

Reaction conditions: Substrate 1a (0.20 mmol), manganese catalyst, and DMBA (14 equiv) were dissolved in CH2Cl2 (1.0 mL) at –30 oC under an Ar atmosphere. Then, H2O2 (30% aqueous solution diluted in 1.0 mL of MeCN) was added to the solution dropwise over 2 h using a syringe pump, and the reaction solution was stirred for additional 2 h. b Conversion was based on the amount of the recovered 1a. c Isolated yield. d The ee values were determined by HPLC using chiral stationary phases. e DMBA (7 equiv) was used. f EHA (14 equiv) was used instead of DMBA. gAcOH (14 equiv) was used instead of DMBA. h The configuration of 2a was opposite due to the use of (R,R) manganese catalyst.

spirocyclic compounds in this manganese catalytic system is highly site-selective (Scheme 1C). With the optimized reaction conditions (Table 1, entry 6), we investigated the oxidation of spirocyclic tetralone derivatives (1a – 1i; see Supporting Information (SI) for the syntheses of the spiro-

a

Reaction conditions: Substrate 1 (0.20 mmol), manganese catalyst 3a (0.50 mol%), and DMBA (14 equiv) were dissolved in CH2Cl2 (1.0 mL) at –30 oC under an Ar atmosphere. Then, 5.0 equiv of H2O2 (30% aqueous solution diluted in 1.0 mL of MeCN) was added dropwise over 2 h using a syringe pump, and the reaction solution was stirred for additional 2 h. The ee values were determined by HPLC using chiral stationary phases. b See SI for the product analyses for 2a – 2i (see also SI, CIF files for the X-ray structures of 2g and 2i). c 4b was used as a catalyst, and the configuration of 2a was opposite due to the use of (R,R) manganese catalyst.

cyclic monoketone substrates), showing that spirocyclic diketones were formed with moderate to good yields in all of the reactions (Table 2). More importantly, enantioselectivities were excellent in these reactions (up to 94% ee; see 2a and 2e in Table 2). Only the substrate 1d bearing a methoxy group at the C5 position afforded a low product yield, it was probably due to the steric effect of methoxy group at the C5 position (see the data of 2d in Table 2, such as 34% yield and 77% ee). In addition, a low ee value (67%) was observed only in the case of 2g. Interestingly, catalyst 4b with a bipyrrolidine backbone (see the Mn structure in Table 1) afforded a high enantioselectivity for this substrate (89% ee). Furthermore, spirocyclic chromanone 1i was readily converted to the diketone product (2i) with a high enantioselectivity (90% ee). The struc-

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 4 of 10

Table 3. Spirocyclic Indanone Derivatives (5) as Substratea

Scheme 2. Gram-Scale Synthesis of Spirocyclic Diketone (2a) and Its Derivatization

a

Reaction conditions: Substrate 5 (0.20 mmol), manganese catalyst 3a (0.50 mol%), and DMBA (14 equiv) were dissolved in CH2Cl2 (1.0 mL) at –30 oC under an Ar atmosphere. Then, H2O2 (5.0 equiv; 30% aqueous solution diluted in 1.0 mL of MeCN) was added to the solution dropwise over 2 h using a syringe pump, and the reaction solution was stirred for additional 2 h. The ee values were measured by HPLC analysis using chiral stationary phases.

tures of the diketone products, 2g and 2i, were determined by single crystal X-ray diffraction analysis (Table 2; also see SI). To examine the substrate scope of the catalytic system, oxidation of spirocyclic indanone derivatives by H2O2 in the presence of catalyst 3a and DMBA as additive was investigated (see the reaction scheme in Table 3). To our delight, the oxidation of spirocyclic indanones (5a – 5k; see SI for the syntheses of the spirocyclic monoketone substrates) yielded the corresponding spirocyclic β,β’spirobiindanones with good to excellent enantioselectivities (Table 3, 68% – 98% ee). Interestingly, spirocyclic compounds containing two methoxy groups in indanone side (5f) and in indane ring (5k) were converted to the same product (6f and 6k) but with different ee values (92% and 74% ees for 5f and 5k, respectively). The absolute configurations of these spirobiindanone products (6a – 6k) were confirmed to be R-configuration by comparing with the reported compounds (see SI for the product analyses).21e To demonstrate the synthetic utility of the present catalytic system, a gram-scale reaction was performed as shown in Scheme 2. The oxidative desymmetrization of 1a (1.24 g) yielded the desired

spirocyclic diketone S-2a in 70% isolated yield (0.91 g) with a high enantioselectivity (92% ee) (Scheme 2), and the enantiopure S-2a (>99% ee) could be obtained after recrystallization in ethanol. Further, since the reduction of spirobiindanone 6a was reported previously using DIBALH and t-BuLi in THF at –78 oC and the resulting chiral diol could be transformed into bisphosphinite ligand,21e,28 we reduced 2a with the literature procedure and obtained the spirocyclic diol product (7a) without the loss of the optical purity (Scheme 2). The structure of the diol product, 7a, was determined by single crystal X-ray diffraction analysis (Scheme 2; also see SI, CIF file). To gain mechanistic insights into the oxidation of spirocyclic hydrocarbons catalyzed by manganese catalyst and H2O2, we first conducted the oxidation of 1a by reducing the amounts of the Mn catalyst (3a) and the oxidant (H2O2) (Scheme 3A). In this reaction, spirocyclic alcohol (8a) was isolated with high diastereoselectivity and enantioselectivity (13:1 diastereomer ratio (dr) and 97% ee; see SI, Figure S1) (Scheme 3A, eq 1), and further oxidation of 8a produced the corresponding spirocyclic diketone (2a) with 92% yield and 95% ee (Scheme 3A, eq 2). These results demonstrate unambiguously that alcohols were produced initially in the oxidation of spirocyclic compounds by 3a and H2O2, followed by the further oxidation of the alcohols to the spirocyclic diketones.14,29 Secondly, we determined a kinetic isotope effect (KIE) value of 3.0(2) by carrying out an intermolecular competitive oxidation of spirocyclic tetralone (1a) and its deuterated spirocyclic tetralone (1a-d4; SI, Figure S2 for the synthesis) (Scheme 3B), proposing that hydrogen atom (H-atom) abstraction from the methylene CH bonds of 1a by a putative Mn(V)-oxo intermediate is the ratelimiting step (vide infra); a KIE value of 3.9 was reported in the oxidation of cyclohexane and deuterated cyclohexane by 3b and H2O213 and KIE values of ~3 – 4 were obtained in the hydroxylation of cumene, ethylbenzene, and cyclohexane by nonheme manganese catalysts and H2O2.11b,15 Further, when the catalytic oxidation of 1a by 3a and H2O2 was carried out using H216O2 under an 18O2 atmosphere, most (>99%) of oxygen atoms in 2a were found to derive from H216O2 (Scheme 3C; SI, Figure S3 for GC-MS spectrum). Furthermore, when the catalytic oxidation of 1a by 3a and H2O2 was carried out in the presence of CCl3Br, the ketone product (2a)

ACS Paragon Plus Environment

Page 5 of 10 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

A. Stepwise Oxidation for the Formation of Spirocyclic Diketone (2a)

B. Determination of Kinetic Isotope Effect (KIE) Value

Scheme 4. Proposed Mechanism for the Methylene Oxidation by Mn/H2O2 catalytic system OH and carbon radical species was obtained in the 18O2 and CCl3Br experiments (Scheme 4, pathway e). Finally, the alcohol product is further oxidized by another molecule of Mn(V)-oxo species to yield the ketone product (Scheme 4, pathway f).

C. Isotopically 18O-Labeled Experiments

CONCLUSION

Scheme 3. Summary of Mechanistic Studies in the Oxidation of Spirocyclic Tetralone (1a) was formed predominantly with a small amount of brominated product (99% ee). Product analyses for (S)-3',4'-dihydro-1'H-spiro[indene-2,2'20 naphthalene]-1,1'(3H)-dione (S-2a). [α] D = –125.5 (c = 0.20 in CHCl3), 1H NMR (400 MHz, CDCl3) in ppm: δ = 8.07 (d, J = 8.4 Hz, 1H), δ = 7.80 (d, J = 7.6 Hz, 1H), 7.67 (t, J = 7.2 Hz, 1H), 7.537.58 (m, 2H), 7.44 (t, J = 7.6 Hz, 1H), 7.31-7.39 (m, 2H), 3.88 (d, J = 17.2 Hz, 1H), 3.50-3.57 (m, 1H), 3.09 (d, J = 17.2 Hz, 1H), 3.013.09 (m, 1H), 2.57-2.62 (m, 1H), 2.32-2.39 (m, 1H); 13C NMR (100 MHz, CDCl3) in ppm: δ = 204.1, 196.4, 153.0, 144.3, 135.2, 133.9, 131.5, 128.8, 128.2, 127.9, 126.8, 126.5, 124.7, 61.1, 38.0,

Page 6 of 10

32.2, 25.5; HRMS [M+H]+: calculated for C18H15O2: 263.1071, found: 263.1066; HPLC-separation conditions: Chiralcel AD-H, 20 oC, 210 nm, 90/10 hexane/iPrOH, 1.0 mL/min; tR = 11.62 min and 15.24 min, 94% ee. Reduction of chiral spirocyclic diketone 2a. Reduction of chiral spirocyclic diketone 2a was performed according to the literature procedures.21e,28 Under an Ar atmosphere, tert-butyllithium (1.6 M in pentane, 1.9 mL, 3.0 mmol, 6.0 equiv) was slowly added to a solution of DIBAL-H (1.0 M in hexane, 3.3 mL, 3.3 mmol, 6.6 equiv) at –78 °C. The colorless solution was stirred for 5 min, and then warmed to room temperature. The solution was cooled again to –78 °C. To this solution, a suspension of the recrystallized 2a (131 mg, 0.50 mmol, 1.0 equiv) in THF (3.0 mL) was added dropwise over 10 min (additional 1.0 mL of THF was used to rinse the syringe) and the reaction mixture was stirred overnight at – 78 °C. Afterwards, saturated aqueous ammonium chloride was added to quench the reaction at –78 °C, and then the mixture was warmed to room temperature. The mixture was transferred to a beaker containing chloroform (10 mL) and saturated aqueous ammonium chloride (5.0 mL) and stirred for 30 min. The precipitated aluminium salts were removed and the biphasic mixtures were extracted by addition of chloroform (10 mL × 3 times). The combined organic extracts were dried over anhydrous Na2SO4 and filtered to remove Na2SO4 added. The solution phase was then purified by silica gel chromatography (petroleum ether/ethyl acetate = 10:1 – 5:1) to give the desired spirocyclic compound 7a as a white solid (114 mg, 86%, > 20:1 dr). Stepwise oxidation of 1a. Substrate 1a (0.40 mmol), catalyst 3a (0.25 mol %), and DMBA (7.0 equiv) were dissolved in CH2Cl2 (2.0 mL) under an Ar atmosphere at –30 oC. H2O2 (1.0 mmol, 2.5 equiv; 30% aqueous solution diluted in 0.50 mL of MeCN) was added dropwise to this mixture over 30 min using a syringe pump. The reaction mixture was stirred for additional 10 min and then quenched with sodium sulfite. The final reaction mixture was purified by silica gel chromatography (petroleum ether/ethyl acetate = 20:1 – 10:1) to give the desired spirocyclic alcohol 8a and diketone 2a. Based on the 1H NMR, dr of 8a was 13 (SI, Figure S1). Then, the catalytic oxidation of 8a was performed separately. Spirocyclic alcohol 8a (0.060 mmol), manganese catalyst 3a (0.50 mol%) and DMBA (14 equiv) were dissolved in CH2Cl2 (0.50 mL) under an Ar atmosphere at –30 oC. H2O2 (3.0 equiv; 30% aqueous solution diluted in 0.50 mL of MeCN) was added dropwise to this mixture over 30 min using a syringe pump. The reaction mixture was stirred for additional 10 min and then quenched with sodium sulfite. The final reaction mixture was purified by silica gel chromatography (petroleum ether/ethyl acetate = 20:1 – 10:1) to give the desired spirocyclic diketone 2a with the isolated yield of 92% and 95% ee. Deuterium kinetic isotope effect in intermolecular competitive reaction. The deuterated spirocyclic tetralone (1a-d4) was synthesized according to the same synthetic method of 1a with 1,2bis(bromomethyl-d4)benzene (SI, Figure S2). Substrate 1a-h4 (0.10 mmol), 1a-d4 (0.10 mmol), catalyst 3a (0.50 mol%) and DMBA (14 equiv) were dissolved in CH2Cl2 (1.0 mL) under an Ar atmosphere at –30 oC. H2O2 (0.10 mmol, 0.50 equiv; 30% aqueous solution diluted in 0.50 mL of MeCN) was added dropwise to this mixture over 30 min using a syringe pump. The reaction mixture was stirred for additional 10 min and then quenched with sodium sulfite. The final reaction solution was purified by silica gel chroma-

ACS Paragon Plus Environment

Page 7 of 10 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

tography (petroleum ether/ethyl acetate = 20:1 – 10:1) to recover the unreacted substrate and ketone and alcohol. The KIE value (kH/kD = 3.0(2)) was determined based on the 1H NMR of the isolated alcohol and ketone products (SI, Figures S4 – S6).

Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: Syntheses of the spirocyclic monoketone substrates (1a – 1i and 5a – 5k), General procedure for the oxidation of 1a – 1i and 5a – 5k catalyzed by manganese complex 3a and analyses of the products (2a – 2i and 6a – 6k) Product analyses for 7a and 8a, and Figures S1 – S5. (PDF) Crystallographic data for 2g (CIF) Crystallographic data for 2i (CIF) Crystallographic data for 7a (CIF)

*[email protected] *[email protected]

The authors declare no competing financial interest.

We acknowledge financial support of this work from the National Natural Science Foundation of China (21473226 and 21773273 to W.S.) and the NRF of Korea through CRI (NRF-2012R1A3A2048842 to W.N.) and GRL (NRF-2010-00353 to W.N.).

(1) Newhouse, T.; Baran, P. S., If C–H Bonds Could Talk: Selective C–H Bond Oxidation. Angew. Chem., Int. Ed. 2011, 50, 3362-3374. (2) (a) Abu-Omar, M. M.; Loaiza, A.; Hontzeas, N. Reaction Mechanisms of Mononuclear Non-Heme Iron Oxygenases. Chem. Rev. 2005, 105, 2227-2252. (b) Krebs, C.; Fujimori, D. G.; Walsh, C. T.; Bollinger, J. M., Jr., Non-Heme Fe(IV)–Oxo Intermediates. Acc. Chem. Res. 2007, 40, 484492. (c) Ortiz de Montellano, P. R., Hydrocarbon Hydroxylation by Cytochrome P450 Enzymes. Chem. Rev. 2010, 110, 932-948. (d) Solomon, E. I.; Goudarzi, S.; Sutherlin, K. D., O2 Activation by Non-Heme Iron Enzymes. Biochemistry 2016, 55, 6363-6374. (e) Wang, V. C.-C.; Maji, S.; Chen, P. P.-Y.; Lee, H. K.; Yu, S. S.-F.; Chan, S. I., Alkane Oxidation: Methane Monooxygenases, Related Enzymes, and Their Biomimetics. Chem. Rev. 2017, 117, 8574-8621. (3) (a) Que, L., Jr.; Tolman, W. B., Biologically Inspired Oxidation Catalysis. Nature 2008, 455, 333-340. (b) Che, C.-M.; Lo, V. K.-Y.; Zhou, C.-Y.; Huang, J.-S., Selective Functionalisation of Saturated C–H Bonds with Metalloporphyrin Catalysts Chem. Soc. Rev. 2011, 40, 1950-1975. (c) Lindhorst, A. C.; Haslinger, S.; Kühn, F. E., Molecular Iron Complexes as Catalysts for Selective C–H Bond Oxygenation Reactions. Chem. Commun. 2015, 51, 17193-17212. (d) Oloo, W. N.; Que, L., Jr., Bioinspired Nonheme Iron Catalysts for C–H and C=C Bond Oxidation: Insights into the Nature of the Metal-Based Oxidants. Acc. Chem. Res. 2015, 48, 2612-2621. (e) Upp, D. M.; Lewis, J. C., Selective C–H Bond Functionalization Using Repurposed or Artificial Metalloenzymes. Curr. Opin. Chem. Biol. 2017, 37, 48-55.

(4) (a) Cussό, O.; Ribas, X.; Costas, M., Biologically Inspired Non-heme Iron-Catalysts for Asymmetric Epoxidation; Design Principles and Perspectives. Chem. Commun. 2015, 51, 14285-14298. (b) Olivo, G.; Cussό, O.; Costas, M., Biologically Inspired C–H and C=C Oxidations with H2O2 Catalyzed by Iron Coordination Complexes. Chem. - Asian J. 2016, 11, 3148-3158. (c) Canta, M.; Rodrίguez, M.; Costas, M., Recent Advances in the Selective Oxidation of Alkyl C–H Bonds Catalyzed by Iron Coordination Complexes. Top Curr. Chem. 2016, 372, 27-54. (d) Olivo, G.; Cussό, O.; Borrell, M. Costas, M., Oxidation of Alkane and Alkene Moieties with Biologically Inspired Nonheme Iron Catalysts and Hydrogen Peroxide: from Free Radicals to Stereoselective Transformations. J. Biol. Inorg. Chem. 2017, 22, 425-452. (e) Gamba, I.; Codolà, Z.; Lloret-Fillol, J.; Costas, M., Making and Breaking of the O–O Bond at Iron Complexes. Coord. Chem. Rev. 2017, 334, 2-24. (5) (a) Talsi, E. P.; Bryliakov, K. P., Chemo- and Stereoselective C–H Oxidations and Epoxidations/cis-Dihydroxylations with H2O2, Catalyzed by Non-Heme Iron and Manganese Complexes. Coord. Chem. Rev. 2012, 256, 1418-1434. (b) Bryliakov, K. P.; Talsi, E. P., Active Sites and Mechanisms of Bioinspired Oxidation with H2O2, Catalyzed by Non-heme Fe and Related Mn Complexes. Coord. Chem. Rev. 2014, 276, 73-96. (c) Ottenbacher, R. V.; Talsi, E. P.; Bryliakov, K. P., Direct Selective Oxidative Functionalization of C–H Bonds with H2O2: Mn-Aminopyridine Complexes Challenge the Dominance of Non-Heme Fe Catalysts. Molecules 2016, 21, 1454. (d) Bryliakov, K. P., Catalytic Asymmetric Oxygenations with the Environmentally Benign Oxidants H2O2 and O2. Chem. Rev. 2017, 117, 11406-11459. (6) (a) Nam, W., High-Valent Iron(IV)–Oxo Complexes of Heme and Non-Heme Ligands in Oxygenation Reactions. Acc. Chem. Res. 2007, 40, 522-531. (b) Ray, K.; Pfaff, F. F.; Wang, B.; Nam, W., Status of Reactive Non-Heme Metal-Oxygen Intermediates in Chemical and Enzymatic Reactions. J. Am. Chem. Soc. 2014, 136, 13942-13958. (c) Chen, Z.; Yin, G., The Reactivity of the Active Metal Oxo and Hydroxo Intermediates and Their Implications in Oxidations. Chem. Soc. Rev. 2015, 44, 1083-1100. (d) Nam, W.; Lee, Y.-M.; Fukuzumi, S., Tuning Reactivity and Mechanism in Oxidation Reactions by Mononuclear Nonheme Iron(IV)-Oxo Complexes. Acc. Chem. Res. 2014, 47, 1146-1154. (e) Cook, S. A.; Borovik, A. S., Molecular Designs for Controlling the Local Environments Around Metal Ions. Acc. Chem. Res. 2015, 48, 2407-2414. (f) Puri, M.; Que, L., Jr., Toward the Synthesis of More Reactive S = 2 Non-Heme Oxoiron(IV) Complexes. Acc. Chem. Res. 2015, 48, 2443-2452. (g) Neu, H. M.; Baglia, R. A.; Goldberg, D. P., A Balancing Act: Stability versus Reactivity of Mn(O) Complexes. Acc. Chem. Res. 2015, 48, 2754-2764. (h) Engelmann, X.; Monte-Pérez, I.; Ray, K., Oxidation Reactions with Bioinspired Mononuclear Non-Heme Metal–Oxo Complexes. Angew. Chem., Int. Ed. 2016, 55, 7632-7649. (7) (a) Costas, M., Selective C–H Oxidation Catalyzed by Metalloporphyrins. Coord. Chem. Rev. 2011, 255, 2912-2932. (b) Lu, H.; Zhang, X. P., Catalytic C–H Functionalization by Metalloporphyrins: Recent Developments and Future Directions. Chem. Soc. Rev. 2011, 40, 1899-1909. (8) (a) Groves, J. T.; Viski, P., Asymmetric Hydroxylation by a Chiral Iron Porphyrin. J. Am. Chem. Soc. 1989, 111, 8537-8538. (b) Groves, J. T.; Viski, P., Asymmetric Hydroxylation, Epoxidation, and Sulfoxidation Catalyzed by Vaulted Binaphthyl Metalloporphyrins. J. Org. Chem. 1990, 55, 3628-3634. (9) (a) Chen, M. S.; White, M. C., A Predictably Selective Aliphatic C-H Oxidation Reaction for Complex Molecule Synthesis. Science 2007, 318, 783-787. (b) Chen, M. S.; White, M. C., Combined Effects on Selectivity in Fe-Catalyzed Methylene Oxidation. Science 2010, 327, 566-571. (c) White, M. C., Adding Aliphatic C-H Bond Oxidations to Synthesis. Science 2012, 335, 807-809. (d) Bigi, M. A.; Reed, S. A.; White, M. C., Directed Metal (Oxo) Aliphatic C–H Hydroxylations: Overriding Substrate Bias. J. Am. Chem. Soc. 2012, 134, 9721-9726. (e) Howell, J. M.; Feng, K.; Clark, J. R.; Trzepkowski, L. J.; White, M. C., Remote Oxidation of Aliphatic C–H

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

Bonds in Nitrogen-Containing Molecules. J. Am. Chem. Soc. 2015, 137, 14590-14593. (10) (a) Company, A.; Gómez, L.; Güell, M.; Ribas, X.; Luis, J. M.; Que, L., Jr.; Costas, M., Alkane Hydroxylation by a Nonheme Iron Catalyst that Challenges the Heme Paradigm for Oxygenase Action. J. Am. Chem. Soc. 2007, 129, 15766-15767. (b) Gómez, L.; Garcia-Bosch, I.; Company, A.; Benet-Buchholz, J.; Polo, A.; Sala, X.; Ribas, X.; Costas, M., Stereospecific C-H Oxidation with H2O2 Catalyzed by a Chemically Robust Site-Isolated Iron Catalyst. Angew. Chem., Int. Ed. 2009, 48, 5720-5723. (c) Prat, I.; Company, A.; Postils, V.; Ribas, X.; Que, L., Jr.; Luis, J. M.; Costas, M., The Mechanism of Stereospecific C–H Oxidation by Fe(Pytacn) Complexes: Bioinspired Non-Heme Iron Catalysts Containing cis-Labile Exchangeable Sites. Chem. - Eur. J. 2013, 19, 6724-6738. (d) Mitra, M.; Lloret-Fillol, J.; Haukka, M.; Costas, M.; Nordlander, E., Evidence that Steric Factors Modulate Reactivity of Tautomeric Iron-Oxo Species in Stereospecific Alkane C–H Hydroxylation. Chem. Commun. 2014, 50, 1408-1410. (e) Mitra, M.; Nimir, H.; Hrovat, D. A.; Shteinman, A. A.; Richmond, M. G.; Costas, M.; Nordlander, E., Catalytic C–H Oxidations by Nonheme Mononuclear Fe(II) Complexes of Two Pentadentate Ligands: Evidence for an Fe(IV) Oxo Intermediate. J. Mol. Catal. A: Chem. 2017, 426, 350356. (f) Chen, J.; Lutz, M.; Milan, M.; Costas, M.; Otte, M.; Gebbink, R. J. M. K., Non-Heme Iron Catalysts with a Rigid Bis-Isoindoline Backbone and Their Use in Selective Aliphatic C-H Oxidation. Adv. Synth. Catal. 2017, 359, 2590-2595. (11) (a) Ottenbacher, R. V.; Samsonenko, D. G.; Talsi, E. P.; Bryliakov, K. P., Highly Efficient, Regioselective, and Stereospecific Oxidation of Aliphatic C–H Groups with H2O2, Catalyzed by Aminopyridine Manganese Complexes. Org. Lett. 2012, 14, 4310-4313. (b) 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. (c) Talsi, E. P.; Ottenbacher, R. V.; Bryliakov, K. P., Bioinspired Oxidations of Aliphatic C–H Groups with H2O2 in the Presence of Manganese Complexes. J. Organomet. Chem. 2015, 793, 102107. (12) (a) Costas, M.; Chen, K.; Que, L., Jr., Biomimetic Nonheme Iron Catalysts for Alkane Hydroxylation. Coord. Chem. Rev. 2000, 200-202, 517-544. (b) Chen, K.; Que, L., Jr., Stereospecific Alkane Hydroxylation by Non-Heme Iron Catalysts: Mechanistic Evidence for an FeV=O Active Species. J. Am. Chem. Soc. 2001, 123, 6327-6337. (c) Chen, K.; Costas, M.; Que, L., Jr., Spin State Tuning of Non-Heme Iron-Catalyzed Hydrocarbon Oxidations: Participation of FeIII–OOH and Fev=O Intermediates. J. Chem. Soc., Dalton Trans. 2002, 672-679. (13) Shen, D.; Miao, C.; Wang, S.; Xia, C.; Sun, W., Efficient Benzylic and Aliphatic C–H Oxidation with Selectivity for Methylenic Sites Catalyzed by a Bioinspired Manganese Complex. Org. Lett. 2014, 16, 1108-1111. (14) Milan, M.; Bietti, M.; Costas, M., Highly Enantioselective Oxidation of Nonactivated Aliphatic C–H Bonds with Hydrogen Peroxide Catalyzed by Manganese Complexes. ACS Cent. Sci. 2017, 3, 196-204. (15) Talsi, E. P.; Samsonenko, D. G.; Ottenbacher, R. V.; Bryliakov, K. P., Highly Enantioselective C−H Oxidation of Arylalkanes with H2O2 in the Presence of Chiral Mn-Aminopyridine Complexes. ChemCatChem 2017, 9, 4580-4586. (16) (a) Hamachi, K.; Irie, R.; Katsuki, T., Asymmetric Benzylic Oxidation Using a Mn-Salen Complex as Catalyst. Tetrahedron Lett. 1996, 37, 49794982. (b) Hamada, T.; Irie, R.; Mihara, J.; Hamachi, K.; Katsuki, T., Highly Enantioselective Benzylic Hydroxylation with Concave Type of (Salen)manganese(III) Complex. Tetrahedron 1998, 54, 10017-10028. (c) Zhang, R.; Yu, W.-Y.; Lai, T.-S.; Che, C.-M., Enantioselective Hydroxylation of Benzylic C−H Bonds by D4-Symmetric Chiral Oxoruthenium Porphyrins. Chem. Commun. 1999, 1791-1792. (d) Zhang, R.; Yu, W.-Y.; Che, C.-M., Catalytic Enantioselective Oxidation of Aromatic Hydrocarbons with D4-Symmetric Chiral Ruthenium Porphyrin Catalysts. Tetrahedron: Asymmetry 2005, 16, 3520-3526. (e) Srour, H.; Le Maux, P.; Simonneaux, G., Enantioselective Manganese-Porphyrin-

Page 8 of 10

Catalyzed Epoxidation and C–H Hydroxylation with Hydrogen Peroxide in Water/Methanol Solutions. Inorg. Chem. 2012, 51, 5850-5856. (17) (a) Zeng, X.-P.; Cao, Z.-Y.; Wang, Y.-H.; Zhou, F.; Zhou, J., Catalytic Enantioselective Desymmetrization Reactions to All-Carbon Quaternary Stereocenters. Chem. Rev. 2016, 116, 7330-7396. (b) Komiya, N.; Noji, S.; Murahashi, S.-I., Manganese Catalyzed Asymmetric Oxidation of Alkanes to Optically Active Ketones Bearing Asymmetric Center at the α-Position. Tetrahedron Lett. 1998, 39, 7921-7924. (18) Frost, J. R.; Huber, S. M.; Breitenlechner, S.; Bannwarth, C.; Bach, T., Enantiotopos-Selective C–H Oxygenation Catalyzed by a Supramolecular Ruthenium Complex. Angew. Chem., Int. Ed. 2015, 54, 691-695. (19) (a) Xie, J.-H.; Zhou, Q.-L., Chiral Diphosphine and Monodentate Phosphorus Ligands on a Spiro Scaffold for Transition-Metal-Catalyzed Asymmetric Reactions. Acc. Chem. Res. 2008, 41, 581-593. (b) Chan, A. S. C.; Hu, W.; Pai, C.-C.; Lau, C.-P.; Jiang, Y.; Mi, A.; Yan, M.; Sun, J.; Lou, R.; Deng, J., Novel Spiro Phosphinite Ligands and Their Application in Homogeneous Catalytic Hydrogenation Reactions. J. Am. Chem. Soc. 1997, 119, 9570-9571. (c) Xie, J.-H.; Wang, L.-X.; Fu, Y.; Zhu, S.-F.; Fan, B.-M.; Duan, H.-F.; Zhou, Q.-L., Synthesis of Spiro Diphosphines and Their Application in Asymmetric Hydrogenation of Ketones. J. Am. Chem. Soc. 2003, 125, 4404-4405. (d) Zhu, S.-F.; Cai, Y.; Mao, H.-X.; Xie, J.-H.; Zhou, Q.-L., Enantioselective Iron-Catalyzed O–H Βond Insertions. Nat. Chem. 2010, 2, 546-551. (e) Han, Z.; Wang, Z.; Zhang, X.; Ding, K., Spiro[4,4]-1,6-nonadiene-Based Phosphine–Oxazoline Ligands for Iridium-Catalyzed Enantioselective Hydrogenation of Ketimines. Angew. Chem., Int. Ed. 2009, 48, 5345-5349. (f) Wang, X.; Guo, P.; Han, Z.; Wang, X.; Wang, Z.; Ding, K., Spiroketal-Based Diphosphine Ligands in PdCatalyzed Asymmetric Allylic Amination of Morita–Baylis–Hillman Adducts: Exceptionally High Efficiency and New Mechanism. J. Am. Chem. Soc. 2014, 136, 405-411. (20) D’yakonov, V. A.; Trapeznikova, O. A.; de Meijere, A.; Dzhemilev, U. M., Metal Complex Catalysis in the Synthesis of Spirocarbocycles. Chem. Rev. 2014, 114, 5775-5814. (21) (a) Zheng, J.; Wang, S.-B.; Zheng, C.; You, S.-L., Asymmetric Dearomatization of Naphthols via a Rh-Catalyzed C(sp2)–H Functionalization/Annulation Reaction. J. Am. Chem. Soc. 2015, 137, 4880-4883. (b) Yang, B.-M.; Cai, P.-J.; Tu, Y.-Q.; Yu, Z.-X.; Chen, Z.-M.; Wang S.-H.; Wang, S.-H.; Zhang, F.-M., Organocatalytic Asymmetric Tandem Nazarov Cyclization/Semipinacol Rearrangement: Rapid Construction of Chiral Spiro[4.4]nonane-1,6-diones. J. Am. Chem. Soc. 2015, 137, 8344-8347. (c) Yoshida, K.; Itatsu, Y.; Fujino, Y.; Inoue, H.; Takao, K., Enantioselective Organocatalytic Construction of Spiroindane Derivatives by Intramolecular Friedel–Crafts-Type 1,4-Addition. Angew. Chem., Int. Ed. 2016, 55, 6734-6738. (d) Chai, Z.; Rainey, T. J., Pd(II)/Brønsted Acid Catalyzed Enantioselective Allylic C–H Activation for the Synthesis of Spirocyclic Rings. J. Am. Chem. Soc. 2012, 134, 36153618. (e) Rahemtulla, B. F.; Clark, H. F.; Smith, M. D., Catalytic Enantioselective Synthesis of C1- and C2-Symmetric Spirobiindanones through Counterion-Directed Enolate C-Acylation. Angew. Chem., Int. Ed. 2016, 55, 13180-13183. (22) Franz, A. K.; Hanhan, N. V.; Ball-Jones, N. R., Asymmetric Catalysis for the Synthesis of Spirocyclic Compounds. ACS Catal. 2013, 3, 540-553. (23) (a) Miao, C.; Wang, B.; Wang, Y.; Xia, C.; Lee, Y.-M.; Nam, W.; Sun, W., Proton-Promoted and Anion-Enhanced Epoxidation of Olefins by Hydrogen Peroxide in the Presence of Nonheme Manganese Catalysts. J. Am. Chem. Soc. 2016, 138, 936-943. (b) Shen, D.; Qiu, B.; Xu, D.; Miao, C.; Xia, C.; Sun, W., Enantioselective Epoxidation of Olefins with H2O2 Catalyzed by Bioinspired Aminopyridine Manganese Complexes. Org. Lett. 2016, 18, 372-375. (c) Shen, D.; Miao, C.; Wang, S.; Xia, C.; Sun, W., A Mononuclear Manganese Complex of a Tetradentate Nitrogen Ligand – Synthesis, Characterizations, and Application in the Asymmetric Epoxidation of Olefins. Eur. J. Inorg. Chem. 2014, 5777-5782. (d) Wang, X.; Miao, C.; Wang, S.; Xia, C.; Sun, W., Bioinspired Manganese and Iron Complexes with Tetradentate N Ligands for the Asymmetric Epoxidation of Olefins. ChemCatChem 2013, 5, 2489-2494. (e) Wang, B.; Miao, C.;

ACS Paragon Plus Environment

Page 9 of 10 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

Wang, S.; Xia, C.; Sun, W., Manganese Catalysts with C1-Symmetric N4 Ligand for Enantioselective Epoxidation of Olefins. Chem. - Eur. J. 2012, 18, 6750-6753. (f) Wu, M.; Wang, B.; Wang, S.; Xia, C.; Sun, W., Asymmetric Epoxidation of Olefins with Chiral Bioinspired Manganese Complexes. Org. Lett. 2009, 11, 3622-3625. (24) (a) Ottenbacher, R. V.; Talsi, E. P.; Bryliakov, K. P., Bioinspired MnAminopyridine Catalyzed Epoxidations of Olefins with Various Oxidants: Enantioselectivity and Mechanism. Catal. Today 2016, 278, 30-39. (b) Ottenbacher, R. V.; Samsonenko, D. G.; Talsi, E. P.; Bryliakov, K. P., Enantioselective Epoxidations of Olefins with Various Oxidants on Bioinspired Mn Complexes: Evidence for Different Mechanisms and Chiral Additive Amplification. ACS Catal. 2016, 6, 979-988. (c) Ottenbacher, R. V.; Samsonenko, D. G.; Talsi, E. P.; Bryliakov, K. P., Highly Enantioselective Bioinspired Epoxidation of Electron-Deficient Olefins with H2O2 on Aminopyridine Mn Catalysts. ACS Catal. 2014, 4, 1599-1606. (d) Lyakin, O. Y.; Ottenbacher, R. V.; Bryliakov, K. P.; Talsi, E. P., Asymmetric Epoxidations with H2O2 on Fe and Mn Aminopyridine Catalysts: Probing the Nature of Active Species by Combined Electron Paramagnetic Resonance and Enantioselectivity Study. ACS Catal. 2012, 2, 1196-1202. (25) (a) Zima, A. M.; Lyakin, O. Y.; Ottenbacher, R. V.; Bryliakov, K. P.; Talsi E. P., Dramatic Effect of Carboxylic Acid on the Electronic Structure of the Active Species in Fe(PDP)-Catalyzed Asymmetric Epoxidation. ACS Catal. 2016, 6, 5399-5404. (b) Lyakin, O. Y.; Zima, A. M.; Samsonenko, D. G.; Bryliakov, K. P.; Talsi, E. P., EPR Spectroscopic Detection of the Elusive FeV=O Intermediates in Selective Catalytic Oxofunctionalizations of Hydrocarbons Mediated by Biomimetic Ferric Complexes. ACS Catal. 2015, 5, 2702-2707. (26) (a) Mas-Ballesté, R.; Que, L., Jr., Iron-Catalyzed Olefin Epoxidation in the Presence of Acetic Acid: Insights into the Nature of the Metal-Based Oxidant. J. Am. Chem. Soc. 2007, 129, 15964-15972. (b) Wang, Y.; Janardanan, D.; Usharani, D.; Han, K.; Que, L., Jr.; Shaik, S., Nonheme Iron Oxidant Formed in the Presence of H2O2 and Acetic Acid Is the Cyclic Ferric Peracetate Complex, Not a Perferryloxo Complex. ACS Catal. 2013, 3, 1334-1341. (27) (a) Cussó, O.; Garcia-Bosch, I.; Ribas, X.; Lloret-Fillol, J.; Costas, M., Asymmetric Epoxidation with H2O2 by Manipulating the Electronic Properties of Non-heme Iron Catalysts. J. Am. Chem. Soc. 2013, 135, 14871-14878. (b) Cussó, O.; Ribas, X.; Lloret-Fillol, J.; Costas, M., Synergistic Interplay of a Non-Heme Iron Catalyst and Amino Acid Coligands in H2O2 Activation for Asymmetric Epoxidation of α-AlkylSubstituted Styrenes. Angew. Chem., Int. Ed. 2015, 54, 2729-2733. (28) (a) Nieman, J. A.; Keay, B. A., An Improved Synthesis and Resolution of (+)-cis,cis-2,2΄-Spirobiindane-1,1΄-diol. Tetrahedron: Asymmetry 1995, 6, 1575-1583. (b) Guo, Z.; Guan, X.; Chen, Z., Synthesis of a Novel Spiro Bisphosphinite Ligand and Its Application in Rh-catalyzed Asymmetric Hydrogenation. Tetrahedron: Asymmetry 2006, 17, 468-473. (29) (a) Shen, D.; Miao, C.; Xu, D.; Xia, C.; Sun, W., Highly Efficient Oxidation of Secondary Alcohols to Ketones Catalyzed by Manganese Complexes of N4 Ligands with H2O2. Org. Lett. 2015, 17, 54-57. (b) Miao, C.; Li, X.-X.; Lee, Y.-M.; Xia, C.; Wang, Y.; Nam, W.; Sun, W., Manganese Complex-Catalyzed Oxidation and Oxidative Kinetic Resolution of Secondary Alcohols by Hydrogen Peroxide. Chem. Sci. 2017, 8, 7476-7482. (c) Talsi, E. P.; Samsonenko, D. G.; Bryliakov, K. P., Asymmetric Autoamplification in the Oxidative Kinetic Resolution of Secondary Benzylic Alcohols Catalyzed by Manganese Complexes. ChemCatChem 2017, 9, 2599-2607. (30) (a) Huyser, E. S.; Schimke, H.; Burham, R. L., Competition Reactions of Cycloalkanes with Trichloromethanesulfonyl Chloride and Bromotrichloromethane. J. Org. Chem. 1963, 28, 2141-2143. (b) Wang, C.; Shalyaev, K. V.; Bonchio, M.; Carofiglio, T.; Groves, J. T., Fast Catalytic Hydroxylation of Hydrocarbons with Ruthenium Porphyrins. Inorg. Chem. 2006, 45, 4769-4782. (31) (a) Huang, X.; Groves, J. T., Beyond Ferryl-Mediated Hydroxylation: 40 Years of the Rebound Mechanism and C–H Activation. J. Biol. Inorg.

Chem. 2017, 22, 185-207. (b) Liu, W.; Cheng, M.-J.; Nielsen, R. J.; Goddard, W. A., III; Groves, J. T., Probing the C–O Bond-Formation Step in Metalloporphyrin-Catalyzed C–H Oxygenation Reactions. ACS Catal. 2017, 7, 4182-4188. (c) Groves, J. T., Key Elements of the Chemistry of Cytochrome P-450: The Oxygen Rebound Mechanism. J. Chem. Educ. 1985, 62, 928-931. (32) (a) Cho, K.-B.; Hirao, H.; Shaik, S.; Nam, W., To Rebound or Dissociate? This Is the Mechanistic Question in C–H Hydroxylation by Heme and Nonheme Metal–oxo Complexes. Chem. Soc. Rev. 2016, 45, 1197-1210. (b) Bae, S. H.; Seo, M. S.; Lee, Y.-M.; Cho, K.- B.; Kim, W.-S.; Nam, W., Mononuclear Nonheme High-Spin (S = 2) versus IntermediateSpin (S = 1) Iron(IV)–Oxo Complexes in Oxidation Reactions. Angew. Chem., Int. Ed. 2016, 55, 8027-8031. (c) Cho, K.-B.; Nam, W., A Theoretical Study into a trans-Dioxo MnV Porphyrin Complex That Does Not Follow the Oxygen Rebound Mechanism in C–H Bond Activation Reactions. Chem. Commun. 2016, 52, 904-907. (d) Dhuri, S. N.; Cho, K.B.; Lee, Y.-M.; Shin, S. Y.; Kim, J. H.; Mandal, D.; Shaik, S.; Nam, W., Interplay of Experiment and Theory in Elucidating Mechanisms of Oxidation Reactions by a Nonheme RuIVO Complex. J. Am. Chem. Soc. 2015, 137, 8623-8632. (e) Kwon, E.; Cho, K.-B.; Hong, S.; Nam, W., Mechanistic Insight into the Hydroxylation of Alkanes by a Nonheme Iron(V)–Oxo Complex. Chem. Commun. 2014, 50, 5572-5575. (f) Cho, K.-B.; Wu, X.; Lee, Y.-M.; Kwon, Y. H.; Shaik, S.; Nam, W., Evidence for an Alternative to the Oxygen Rebound Mechanism in C–H Bond Activation by Non-Heme FeIVO Complexes. J. Am. Chem. Soc. 2012, 134, 2022220225. (g) Wu, X.; Seo, M. S.; Davis, K. M.; Lee, Y.-M.; Chen, J.; Cho, K.B.; Pushkar, Y. N.; Nam, W., A Highly Reactive Mononuclear Non-heme Manganese(IV)–Oxo Complex that Can Activate the Strong C–H Bonds of Alkanes. J. Am. Chem. Soc. 2011, 133, 20088-20091. (33) Armarego, W. L. F.; Chai, C. L. L. in Purification of Laboratory Chemicals, 6th ed., Pergamon Press, Oxford, 2009.

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

Table of Contents

ACS Paragon Plus Environment

Page 10 of 10