Norbornene

the Microscale, University of Science and Technology of China, 96 Jinzhai Road, Hefei, Anhui 230026, P. R. China. ABSTRACT: Three-component ...
31 downloads 3 Views 472KB Size
Subscriber access provided by Kaohsiung Medical University

Letter

Enantioselective Synthesis of Biaryl Atropisomers via Pd/ Norbornene-Catalyzed Three-Component Cross-Couplings Linlin Ding, Xianwei Sui, and Zhenhua Gu ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b01037 • Publication Date (Web): 18 May 2018 Downloaded from http://pubs.acs.org on May 18, 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.

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 7 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

Enantioselective Synthesis of Biaryl Atropisomers Pd/Norbornene-Catalyzed Three-Component Cross-Couplings

via

Linlin Ding, Xianwei Sui and Zhenhua Gu* Department of Chemistry, Center for Excellence in Molecular Synthesis, Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and Technology of China, 96 Jinzhai Road, Hefei, Anhui 230026, P. R. China ABSTRACT: Three-component cross-coupling cocatalyzed by palladium and norbornene is reported for the synthesis of biaryl atropisomers. This domino reaction gave optimal yield and enantioselectivity with a P,C-type ligand bearing axial chirality and P chiral center. The process showed advantages over traditional cross-coupling because of its step economy and its compatibility with readily available ortho-substituted aryl halides, which could, therefore, be used instead of continuously trisubstituted aryl halides.

INTRODUCTION Biaryl atropisomers show chirality due to restricted rotation about a single aryl-aryl bond, and biaryl atropisomeric substructures exist in many biologically active natural products (Figure 1, A and B).1-4 These substructures provide chiral skeletons for ligands and catalysts in asymmetric catalysis (Figure 1, C and D).5-8

Figure 1. Atropisomerism in Natural Products and Ligands The most straightforward way to prepare biaryl atropisomers is asymmetric aryl-aryl cross-coupling or oxidative coupling of phenols.9-20 Nevertheless other methods have been developed to construct axially chiral biaryls with good substituent-loading capability.21 For example, the (dynamic) kinetic resolution, asymmetric [2+2+2]cycloaddition were efficient protocols for the construction of chiral atropisomers.22-37 Chiral atropisomers have also been created via organo-catalyzed asymmetric [3,3]sigmatropic rearrangement of N,N’-binaphthyl hydrazines and by asymmetric Michael addition of 2-naphthols to quinones.3840 Till now, several excellent ligands and catalytic systems have been developed for preparing chiral atropisomers. Some of these ligands and catalysts show extremely high activity and enantioselectivity even in the case of bulky aryl-aryl asymmetric cross-couplings (Scheme 1A). However, these methods of preparing biaryl atropisomers still have their own limitations. For example, continuously trisubstituted aryl halides must be prepared through a tedious sequential procedure (Scheme 1B), and some multi-substituted aryl halides with certain substituents at certain positions simply cannot be prepared through the known methods. Some reactions may be adaptable to the synthesis of biaryl atropisomers, thereby allowing the preparation of a greater

diversity of these compounds. The Pd/norbornene-catalyzed domino reaction known as the Catellani reaction, developed in 1997, enables both ipso and ortho functionalization of aryl halides.41-49 The groups of Catellani, Lautens and others have synthesized numerous complex aromatic compounds through ortho alkylation or arylation of aryl halides50-59 or indoles,60 or through C-H functionalization of arenes.61-62 In 2013, the Dong group reported the first palladium/norbornene-catalyzed orthoamination.63 In 2015, our group and the groups of Liang and Dong independently achieved ortho-acylation.64-67 Subsequently, Dong and co-workers selectively cleaved the C(O)-O bond of mixed carbonate anhydrides and introduced an α-alkoxylcarbonyl group.68 We hypothesized that we could use a three-component coupling strategy to prepare biaryl atropisomers. Based on the Catellani group's pioneering work on Suzuki-type termination,69 we reasoned that palladium and norbornene could cocatalyze addition of the R2 group adjacent to the C-X bond, followed by asymmetric cross-coupling to construct the axially chiral bond (Scheme 1C). Ortho-substituted aryl halides in the presence of appropriate electrophiles can take the place of continuously trisubstituted aryl halides, which are more difficult to prepare. This three-component strategy should be step-economical and highly divergent through variations of the three components. On the other hand, the strategy is more complex than classic cross-couplings, and the stoichiometric norbornene may coordinate with the palladium catalyst, affecting stereoinduction and reactivity.

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 2 of 7

atropisomeric chirality of ligands L6a, L6b, L7a and L7b was highly stable; HPLC analysis on the chiral stations revealed no evidence of racemization. Ligand L7a was isolated from the mixture of L7a and L7b and used as the optimized ligand in cross-coupling reactions. The configuration at the P chiral center was unambiguously established through single-crystal X-ray diffraction analysis of PdCl2-L7a complex 9 (Scheme 2).73 Cσ-Pd bonding was observed in the crystal structure. L7a acts as a P,C-ligand, whose analogues demonstrated a substantial acceleration of Suzuki-Miyaura coupling in the previous reports.74 With these ligands in hand, we continued to optimize the reaction conditions. The diastereomers L6a and L6b differed slightly in stereoinduction (entries 6-7). L7a showed the highest enantioselectivity of 86:14 er, while L7b afforded product 4a without any enantioselectivity (entries 8-9). This enantioselectivity rose to 89:11 after the reaction temperature was lowered to 60 oC (entry 10). Screening of palladium sources identified the combination of Pd(TFA)2 and bicyclo[2.2.1]hept-5-ene-2-carbaldehyde as optimal in terms of yield and enantioselectivity (entries 11-14). Table 1. Optimization of Reaction Conditionsa

Scheme 1. (A) Representative Ligands or Catalysts for Asymmetric Aryl-Aryl Coupling. (B) Traditional CrossCoupling for Atropisomer Synthesis. (C) Three-Component Strategy. RESULTS AND DISCUSSION In our initial trials, we used chloromethyl benzoate 2a as the electrophile; this compound can serve as a functionalized methylene group. We reasoned that the resulting (benzoyloxy)methyl group in the products would readily undergo postmodification, providing access to structurally divergent atropisomers. In our experiments, reactions with bidentate phosphines such as DIOP or BINAP afforded very low yield or even no product. Screening of monodentate phosphine ligands L1-L3 (Scheme 2) led to enantioselectivities up to 76:33 er (Table 1, entries 1-3). Replacing the methoxyl group in the phosphine ligand L3 with a dimethylamino group [L4 (MAP) and L5 (KenPhos)] improved both yield and enantioselectivity (entries 4 and 5). Next we focused on the modification of an L4-type ligand. Known procedures were used to synthesize a class of ligands bearing a chiral phosphine atom (L6a, L6b, L7a and L7b; Scheme 2). Amino iodide 6 was prepared on the gram scale from (R)-[1,1'-binaphthalene]-2,2'-diamine 5 within four steps.70-71 Palladium-catalyzed phosphorylation afforded phosphine oxides 7 and 8 in excellent yields with a pair of diastereomers bearing a P chiral center.72 These diastereomers were separated by column chromatography on silica gel, but they isomerized during reduction with phosphine oxide at high temperature. HSiCl3 directly reduced stereoisomers 7 and 8 to give the corresponding phosphines, whose diastereomers were successfully separated by column chromatography on silica gel. Heating L7b in toluene created an equilibrium between L7a and L7b. The ratio of L7a : L7b gradually increased upon heating to 110 oC in toluene, reaching a maximal value of 0.94 : 1.0. This ratio did not increase further upon prolonged heating (see Supporting Information). The

entry

ligand

Pd

T/oC

yield/%b

er

1

L1

Pd(OAc)2

90

33

54:46

2

L2

Pd(OAc)2

90

27

65:35

3

L3

Pd(OAc)2

90

41

67:33

4

L4

Pd(OAc)2

90

53

71:29

5

L5

Pd(OAc)2

90

47

68:32

6

L6a

Pd(OAc)2

90

46

79:21

7

L6b

Pd(OAc)2

90

42

72:28

8

L7a

Pd(OAc)2

90

61

86:14

9

L7b

Pd(OAc)2

90

52

50:50

10

L7a

Pd(OAc)2

60

52

89:11

11

L7a

Pd2(dba)3

60

44

90:10

12

L7a

PdCl2

60

46

90:10

13

L7a

Pd(acac)2

60

46

89:11

14c

L7a

Pd(TFA)2

60

60

90:10

a

The reactions were carried out with 1a (0.157 mmol), 2a (0.392 mmol, 2.5 equiv), 3a (0.314 mmol, 2.0 equiv), palladium (10 mol %), norbornene (0.314 mmol, 2.0 equiv), phosphine ligand (11 mol %) and K2CO3 (0.471 mmol, 3.0 equiv) in CH3CN (2.0 ml). Reactions were incubated at the indicated temperatures for 18-22 h. b Isolated yield of 4a. c Bicyclo[2.2.1]hept-5-ene-2carbaldehyde was used instead of norbornene.

ACS Paragon Plus Environment

Page 3 of 7 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 or nitrile groups failed to give the desired products under the optimal conditions (see Supporting Information). Table 2. Substrate Scopea

Scheme 2. Chiral Ligands After optimizing the conditions of this three-component asymmetric cross-coupling, we examined its substrate scope (Table 2). The following chloromethyl coupling components were found to give similar enantioselectivity: substituted benzoates (4b-4i, 4l-4v), pivalate (4j) and 2-naphthoate (4k). In the end, chloromethyl 3-nitrobenzoate was preferable because the resulting products were easier to purify from byproducts. The ortho aldehyde functionality in aryl boronic acids was necessary for good enantioselectivity, whereas introducing a fluoro or chloro atom adjacent to the aldehyde group increased er to 94:6 and 95:5, respectively (4c and 4d). Addition of fluoro, chloro or trifluoromethyl groups meta to the aldehyde did not substantially affect yield or enantioselectivity (4e-4g). Adding an electron-donating methoxyl group adjacent to the aldehyde significantly reduced product yield (4h), while adding a methyl group para to the aldehyde slightly decreased enantioselectivity (4i). Adding electron-withdrawing or donating substituents (nitro, methyl, fluoro) to the 4-position of 1-iodonaphthalene (4l-4n) did not substantially affect the reaction. Enantioselectivity reached 96:4-98:2 er when we used 4sulfonamide-substituted 1-iodonaphthalenes (4o-4s). A rotamer pair was detected by NMR at room temperature, which may arise because the substituted benzoate interferes sterically with the sulfonamide group, restricting the latter's rotation. However, the single isomer was observed and clear NMR spectroscopies could be obtained by taking NMR at an elevated temperature. Performing the reaction at 1.0 mmol scale gave the desired product 4t without loss of enantioselectivity. Introducing a 6methoxyl group into 1-iodonaphthalene only marginally affected yield and selectivity (4u). The absolute configuration of the axial chirality of 4v was unambiguously determined as R based on single-crystal x-ray diffraction.75 Phenyl boronic acids bearing ortho nitro, methoxylcarbonyl, methylcarbonyl

a

The reactions were carried out with 1 (0.157 mmol), 2 (0.392 mmol, 2.5 equiv), 3 (0.314 mmol, 2.0 equiv), palladium (10 mol %), NBE* (0.314 mmol, 2.0 equiv), phosphine ligand (11 mol %) and K2CO3 (0.471 mmol, 3.0 equiv) in CH3CN (2.0 ml). Reactions were incubated at 60 °C for 22 h. The yields are isolated yields of 4.NBE* = Bicyclo[2.2.1]hept-5-ene-2-carbaldehyde.

To rationalize this domino reaction, we propose a plausible catalytic cycle based on previous studies by Catellani's group and others (Scheme 3). Oxidative addition of Pd(0) with 1iodonaphthalene 1a gives arylpalladium intermediate I. Insertion of norbornene into the C-Pd bond, followed by carbopalladation, gives II. A second oxidative addition of II with chloromethyl benzoate 2a affords Pd(IV) intermediate III, which undergoes reductive elimination to yield IV. Norbornene extrusion gives α-alkylated arylpalladium V, which undergoes classic Suzuki coupling to afford the final product 4a.

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

single crystal structure of compounds 9 and (R)-4v (CIF). This material is available free of charge via the internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author * [email protected]

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We are grateful for financial supported from NSFC (21622206, 21472179), the '973' project from the MOST of China (2015CB856600), and Strategic Priority Research Program of the Chinese Academy of Sciences (XDB20000000).

REFERENCES Scheme 3. A Plausible Catalytic Cycle Finally, we demonstrated the synthetic usefulness of our novel three-component coupling by preparing optically active atropisomers (Scheme 4). Vinylation of 4t delivered styrene analogue 10t in excellent yield, and the (benzoyloxy)methyl group was then converted into an aldehyde functionality without loss of enantiopurity. Reduction of 4t with LiAlH4 gave diol 12t in 74% yield, and this diol was smoothly transformed into the corresponding dibromo compound 13t. This dibromide can serve as a versatile precursor in nucleophilic substitutions.

Scheme 4. Synthetic Applications

CONCLUSION In summary, we report the first asymmetric catalytic synthesis of biaryl atropisomers via palladium/norbornene co-catalysis. The best enantioselectivity of up to 98:2 er was obtained using the P,C-type ligand L7a bearing axial and P center chirality. This strategy shows high step economy and relies on readily available ortho-substituted aryl iodides instead of continuously trisubstituted aryl halides, which are more tedious to synthesize.

ASSOCIATED CONTENT Supporting Information. Experimental procedure, spectroscopic data, and the 1H, 13C NMR spectra of the products (PDF), the

(1) Smyth, J. E.; Butler, N. M.; Keller, P. A. A Twist of Nature – the Significance of Atropisomers in Biological Systems. Nat. Prod. Rep. 2015, 32, 1562-1583. (2) Bringmann, G.; Mortimer, A. J.; Keller, P. A.; Gresser, M. J.; Garner, J.; Breuning, M. Atroposelective Synthesis of Axially Chiral Biaryl Compounds. Angew. Chem. Int. Ed. 2005, 44, 5384-5427. (3) Zask, A.; Murphy, J.; Ellestad, G. A. Biological Stereoselectivity of Atropisomeric Natural Products and Drugs. Chirality 2013, 25, 265-274. (4) Clayden, J.; Moran, W. J.; Edwards, P. J.; LaPlante, S. R. The Challenge of Atropisomerism in Drug Discovery. Angew. Chem. Int. Ed. 2009, 48, 6398-6401. (5) Chen, Y.; Yekta, S.; Yudin, A. Modified BINOL Ligands in Asymmetric Catalysis. Chem. Rev. 2003, 103, 3155-3212. (6) Berthod, M.; Mignani, G.; Woodward, G.; Lemaire, M. Modified BINAP:  The How and the Why. Chem. Rev. 2005, 105, 18011836. (7) Akiyama, T.; Mori, K. Stronger Brønsted Acids: Recent Progress. Chem. Rev. 2015, 115, 9277-9306. (8) Rokade, B. V.; Guiry, P. J. Axially Chiral P,N-Ligands: Some Recent Twists and Turns. ACS Catal. 2018, 8, 624-643. (9) Zhang, D.; Wang, Q. Palladium Catalyzed Asymmetric SuzukiMiyaura Coupling Reactions to Axially Chiral Biaryl Compounds: Chiral Ligands and Recent Advances, Coord. Chem. Rev. 2015, 286, 1-16. (10) Hayashi, T.; Hayashizaki, K.; Kiyoi, T.; Ito, Y. Asymmetric Synthesis Catalyzed by Chiral Ferrocenylphosphine-Transition-Metal Complexes. 6 Practical Asymmetric Synthesis of 1, 1'-Binaphthyls via Asymmetric Cross-Coupling with a Chiral [(alkoxyalkyl)ferrocenyl]monophosphine/Nickel Catalyst. J. Am. Chem. Soc. 1988, 110, 8153. (11) Genov, M.; Almorín, A.; Espinet, P. Efficient Synthesis of Chiral 1,1′-Binaphthalenes by the Asymmetric Suzuki–Miyaura Reaction: Dramatic Synthetic Improvement by Simple Purification of Naphthylboronic Acids. Chem. Eur. J. 2006, 12, 9346-9352. (12) Wang, S.; Li, J.; Miao, T.; Wu, W.; Li, Q.; Zhuang, Y.; Zhou, Z.; Qiu, L. Highly Efficient Synthesis of a Class of Novel ChiralBridged Atropisomeric Monophosphine Ligands via Simple Desymmetrization and Their Applications in Asymmetric Suzuki–Miyaura Coupling Reaction, Org. Lett. 2012, 14, 1966-1969. (13) Uozumi, Y.; Matsuura, Y.; Arakawa, T.; Yamada, Y. M. A. Asymmetric Suzuki–Miyaura Coupling in Water with a Chiral Palladium Catalyst Supported on an Amphiphilic Resin. Angew. Chem. Int. Ed. 2009, 48, 2708-2710. (14) Bermejo, A.; Ros, A.; Fernández, R.; Lassaletta, J. M. C2Symmetric Bis-Hydrazones as Ligands in the Asymmetric Suzuki−Miyaura Cross-Coupling. J. Am. Chem. Soc. 2008, 130, 15798-15799.

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 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 (15) Shen, X.; Jones, G. O.; Watson, D. A.; Bhayana, B.; Buchwald, S. L. Enantioselective Synthesis of Axially Chiral Biaryls by the Pd-Catalyzed Suzuki−Miyaura Reaction: Substrate Scope and Quantum Mechanical Investigations. J. Am. Chem. Soc. 2012, 132, 11278-11287. (16) Xu, G.; Fu, W.; Liu, G.; Senanayake, C. H.; Tang, W. Efficient Syntheses of Korupensamines A, B and Michellamine B by Asymmetric Suzuki-Miyaura Coupling Reactions. J. Am. Chem. Soc. 2014, 136, 570-573. (17) Benhamou, L.; Besnard, C.; Kundig, E. P. Chiral PEPPSI Complexes: Synthesis, Characterization, and Application in Asymmetric Suzuki–Miyaura Coupling Reactions, Organometallics, 2014, 33, 260-266. (18) Luo, Z.; Liu, Q.; Gong, L.; Cui, X.; Mi, A.; Jiang, Y. Novel Achiral Biphenol-Derived Diastereomeric Oxovanadium(IV) Complexes for Highly Enantioselective Oxidative Coupling of 2Naphthols. Angew. Chem. Int. Ed. 2002, 41, 4714-4717. (19) Guo, Q.; Wu, Z.; Luo, Z.; Liu, Q.; Ye, J.; Luo, S.; Cun, L.; Gong, L. Highly Enantioselective Oxidative Couplings of 2Naphthols Catalyzed by Chiral Bimetallic Oxovanadium Complexes with Either Oxygen or Air as Oxidant. J. Am. Chem. Soc. 2007, 129, 13927-13938. (20) Egami, H.; Katsuki, T. Iron-Catalyzed Asymmetric Aerobic Oxidation: Oxidative Coupling of 2-Naphthols. J. Am. Chem. Soc. 2009, 131, 6082-6083. (21) Wencel-Delord, J.; Panossian, A.; Leroux, F. R.; Colobert, F. Recent advances and new concepts for the synthesis of axially stereoenriched biaryls. Chem. Soc. Rev. 2015, 44, 3418-3430. (22) Hazra, C. K.; Dherbassy, Q.; Wencel-Delord, J.; Colobert, F. Synthesis of Axially Chiral Biaryls through Sulfoxide-Directed Asymmetric Mild C-H Activation and Dynamic Kinetic Resolution. Angew. Chem. Int. Ed. 2014, 53, 13871-13875. (23) Bringmann, G.; Menche, D. Stereoselective Total Synthesis of Axially Chiral Natural Products via Biaryl Lactones. Acc. Chem. Res. 2001, 34, 615-624. (24) Mori, K.; Itakura, T.; Akiyama, T. Enantiodivergent Atroposelective Synthesis of Chiral Biaryls by Asymmetric Transfer Hydrogenation: Chiral Phosphoric Acid Catalyzed Dynamic Kinetic Resolution. Angew. Chem. Int. Ed. 2016, 55, 11642-11646. (25) Zhang, J.; Wang, J. Atropoenantioselective Redox-Neutral Amination of Biaryl Compounds through Borrowing Hydrogen and Dynamic Kinetic Resolution. Angew. Chem. Int. Ed. 2018, 57, 465469. (26) Gustafson, J. L.; Lim, D.; Miller, S. J. Dynamic Kinetic Resolution of Biaryl Atropisomers via Peptide-Catalyzed Asymmetric Bromination. Science, 2010, 328, 1251-1255. (27) Diener, M. E.; Metrano, A. J.; Kusano, S.; Miller, S. J. Enantioselective Synthesis of 3-Arylquinazolin-4(3H)-ones via Peptide-Catalyzed Atroposelective Bromination. J. Am. Chem. Soc. 2015, 137, 12369-12377. (28) Crawford, J. M.; Stone, E. A.; Metrano, A. J.; Miller, S. J.; Sigman, M. S. Parameterization and Analysis of Peptide-Based Catalysts for the Atroposelective Bromination of 3-Arylquinazolin4(3H)-ones. J. Am. Chem. Soc. 2018, 140, 868-871. (29) Zheng, J.; You, S.-L. Construction of Axial Chirality by Rhodium-Catalyzed Asymmetric Dehydrogenative Heck Coupling of Biaryl Compounds with Alkenes. Angew. Chem. Int. Ed. 2014, 53, 13244-13463. (30) Zheng, J.; Cui, W.-J.; Zheng, C.; You, S.-L. Synthesis and Application of Chiral Spiro Cp Ligands in Rhodium-Catalyzed Asymmetric Oxidative Coupling of Biaryl Compounds with Alkenes. J. Am. Chem. Soc. 2016, 138, 5242-5245. (31) Bhat, V.; Wang, S.; Stoltz, B. M.; Virgil, S. C. Asymmetric Synthesis of QUINAP via Dynamic Kinetic Resolution. J. Am. Chem. Soc. 2013, 135, 16829-16832. (32) Ma, Y.-N.; Li, S.-X.; Yang, S.-D. New Approaches for BiarylBased Phosphine Ligand Synthesis via P=O Directed C – H Functionalizations. Acc. Chem. Res. 2017, 50, 1480-1492. (33) Ros, A.; Estepa, B.; Ramírez-López, Álvarez, E.; Fernández, R.; Lassaletta, J. M. Dynamic Kinetic Cross-Coupling Strategy for the

Asymmetric Synthesis of Axially Chiral Heterobiaryls. J. Am. Chem. Soc., 2013, 135, 15730-15733. (34) Yao, Q.-J.; Zhang, S.; Zhan, B.-B.; Shi, B.-F. Atroposelective Synthesis of Axially Chiral Biaryls by Palladium-Catalyzed Asymmetric C−H Olefination Enabled by a Transient Chiral Auxiliary. Angew. Chem. Int. Ed. 2017, 56, 6617-6621. (35) Feng, J.; Li, B.; Jiang, J.; Zhang, M.; Ouyang, W.; Li, C.; Fu, Y.; Gu, Z. Visible Light Accelerated Vinyl C–H Arylation in PdCatalysis: Application in the Synthesis of ortho Tetra-substituted Vinylarene Atropisomers. Chin. J. Chem. 2018, 36, 11-14. (36) Zhao, K.; Duan, L.; Xu, S.; Jiang, J.; Fu, Y.; Gu, Z. Enhanced Reactivity by Torsional Strain of Cyclic Diaryliodonium in CuCatalyzed Enantioselective Ring-Opening Reaction. Chem, 2018, 4, 599-612. (37) Transition-Metal-Mediated Aromatic Ring Construction, Tanaka, K. Eds, John Wiley & Sons, Inc., Hoboken, New Jersey, 2013. (38) Li, Q.-Q.; Gao, H.; Keene, C.; Devonas, M.; Ess, D. H.; Kurti, L. Organocatalytic Aryl–Aryl Bond Formation: An Atroposelective [3,3]-Rearrangement Approach to BINAM Derivatives. J. Am. Chem. Soc. 2013, 135, 7414-7417. (39) Chen, Y.-H.; Cheng, D.-J.; Zhang, J.; Wang, Y.; Liu, X.-Y.; Tan, B. Atroposelective Synthesis of Axially Chiral Biaryldiols via Organocatalytic Arylation of 2-Naphthols. J. Am. Chem. Soc. 2015, 137, 15062-15065. (40) Moliterno, M.; Cari, R.; Puglisi, A.; Antenucci, A.; Sperandio, C.; Moretti, E.; Di Sabato, A.; Salvio, R.; Bella, M. QuinineCatalyzed Asymmetric Synthesis of 2,2′-Binaphthol-Type Biaryls under Mild Reaction Conditions. Angew. Chem. Int. Ed. 2016, 55, 6525-6529. (41) Catellani, M.; Frignani, F.; Rangoni, A. A Complex Catalytic Cycle Leading to a Regioselective Synthesis of o,o’-Disubstituted Vinylarenes, Angew. Chem. Int. Ed. Engl. 1997, 36, 119-112. (42) Catellani, M. Catalytic Multistep Reactions via Palladacycles. Synlett 2003, 298-313. (43) Lautens, M.; Alberico, D.; Bressy, C.; Fang, Y.-Q.; Mariampillai, B.; Wilhelm, T. Palladium-catalyzed Ring-forming Reactions: Methods and Applications. Pure Appl. Chem. 2006, 78, 351361. (44) Catellani, M.; Motti, E.; Della Ca’, N.; Ferraccioli, R. Recent Developments in Catalytic Aryl Coupling Reactions. Eur. J. Org. Chem. 2007, 4153-4165. (45) Catellani, M.; Motti, E.; Della Ca’, N. Catalytic Sequential Reactions Involving Palladacycle-Directed Aryl Coupling Steps. Acc. Chem. Res. 2008, 41, 1512-1522. (46) Martins, A.; Mariampillai, B.; Lautens, Synthesis in the Key of Catellani: Norbornene-Mediated ortho C–H Functionalization. Top. Curr. Chem. 2010, 292, 1-33. (47) Ferraccioli, R. Palladium-Catalyzed Synthesis of Carbo- and Heterocycles through Norbornene-Mediated ortho C–H Functionalization. Synthesis, 2013, 45, 581-591. (48) Sui, X.; Gu, Z. Palladium-Catalyzed Chemoselective Catellani ortho-Arylation Reactions and Their Applications in Natural Products Synthesis, Synlett, 2013, 24, 2023-2031. (49) Ye, J.-T.; Lautens, M. Palladium-catalysed Norbornenemediated C–H Functionalization of Arenes. Nat. Chem. 2015, 7, 863870. (50) Lautens, M.; Piguel, S. A New Route to Fused Aromatic Compounds by Using a Palladium-Catalyzed Alkylation –  Alkenylation Sequence. Angew. Chem. Int. Ed. 2000, 39, 1045-1046. (51) Chai, D. I.; Thansandote, P.; Lautens, M. Mechanistic Studies of Pd-Catalyzed Regioselective Aryl C-H Bond Functionalization with Strained Alkenes: Origin of Regioselectivity. Chem. Eur. J. 2011, 17, 8175-8188. (52) Larraufie, M.-H.; Maestri, G.; Beaume, A.; Derat, E.; Ollivier, C.; Fensterbank, L.; Courillon, C.; Lacote, E.; Catellani, M.; Malacria, M. Exception to the ortho Effect in Palladium/Norbornene Catalysis. Angew. Chem. Int. Ed. 2011, 50, 12461-12464. (53) Liu, H.; El-Salfiti, M.; Lautens, M. Expeditious Synthesis of Tetrasubstituted Helical Alkenes by a Cascade of Palladium-

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

Catalyzed C-H Activations. Angew. Chem., Int. Ed. 2012, 51, 98469850. (54) Zhang, H.; Chen, P.; Liu, G. Palladium-Catalyzed Cascade CH Trifluoroethylation of Aryl Iodides and Heck Reaction: Efficient Synthesis of ortho-Trifluoroethylstyrenes. Angew. Chem., Int. Ed. 2014, 53, 10174-10178. (55) Lei, C.; Jin, X.; Zhou, J. Palladium-Catalyzed Heteroarylation and Concomitant ortho-Alkylation of Aryl Iodides. Angew. Chem. Int. Ed. 2015, 54, 13397-13400. (56) Shi, H.; Babinski, D. J.; Ritter, T. Modular C–H Functionalization Cascade of Aryl Iodides. J. Am. Chem. Soc. 2015, 137, 37753778. (57) Fan, L.; Liu, J.; Bai, L.; Wang, Y.; Luan, X. Rapid Assembly of Diversely Functionalized Spiroindenes by a Three-Component Palladium-Catalyzed C−H Amination/Phenol Dearomatization Domino Reaction. Angew. Chem. Int. Ed. 2017, 56, 14257-14261. (58) Cheng, H.; Wu, C.; Chen, H.; Chen, R.; Qian, G.; Geng, Z.; Wei, Q.; Xia, Y.; Zhang, J.; Zhang, Y.; Zhou, Q. Epoxides as Alkylating Reagents for the Catellani Reaction. Angew. Chem. Int. Ed. 2018, 57, 3444-3448. (59) Chang, T.-W.; Ho, P.-Y.; Map, K.-C.; Hong, F.-E. Developing five-membered heterocycle substituted phosphinous acids as ligands for palladium-catalyzed Suzuki–Miyaura and Catellani reactions, Dalton Trans., 2015, 44, 17129-17142. (60) Jiao, L.; Bach, T. Palladium-Catalyzed Direct 2-Alkylation of Indoles by Norbornene-Mediated Regioselective Cascade C–H Activation. J. Am. Chem. Soc. 2011, 133, 12990-12993. (61) Wang, X.-C.; Gong, W.; Fang, L.-Z; Zhu, R.-Y.; Li, S.; Engle, K. M.; Yu, J.-Q. Ligand-enabled meta-C–H activation using a transient mediator. Nature, 2015, 519, 334-338. (62) Dong, Z.; Wang, J.; Dong, G. Simple Amine-Directed MetaSelective C–H Arylation via Pd/Norbornene Catalysis. J. Am. Chem. Soc. 2015, 137, 5887-5890. (63) Dong, Z.; Dong, G. Ortho vs Ipso: Site-Selective Pd and Norbornene-Catalyzed Arene C–H Amination Using Aryl Halides, J. Am. Chem. Soc. 2013, 135, 18350-18353. (64) Zhou, P.-X.; Ye, Y.-Y.; Liu, C.; Zhao, L.-B.; Hou, J.-Y.; Chen, D.-Q.; Tang, Q.; Wang, A.-Q.; Zhang, J.-Y.; Huang, Q.-X.; Xu, P.-F.; Liang, Y.-M. Palladium-Catalyzed Acylation/Alkenylation of Aryl Iodide: A Domino Approach Based on the Catellani–Lautens Reaction. ACS Catal. 2015, 5, 4927-4931.

Page 6 of 7

(65) Dong, Z.; Wang, J.; Ren, Z.; Dong, G. Ortho C-H Acylation of Aryl Iodides by Palladium/Norbornene Catalysis. Angew. Chem. Int. Ed. 2015, 54, 12664-1266. (66) Huang, Y.; Zhu, R.; Zhao, K.; Gu, Z. Palladium-Catalyzed Catellani ortho-Acylation Reaction: An Efficient and Regiospecific Synthesis of Diaryl Ketones. Angew. Chem. Int. Ed. 2015, 54, 1266912672. (67) Sun, F.; Li, M.; He, C.; Wang, B.; Li, B.; Sui, X.; Gu, Z. Cleavage of the C(O)–S Bond of Thioesters by Palladium/Norbornene/Copper Cooperative Catalysis: An Efficient Synthesis of 2-(Arylthio)aryl Ketones. J. Am. Chem. Soc. 2016, 138, 7456-7459. (68) Wang, J.; Zhang, L.; Dong, Z.; Dong, G. Reagent-Enabled ortho-Alkoxycarbonylation of Aryl Iodides via Palladium/Norbornene Catalysis, Chem 2016, 1, 581-591. (69) Catellani, M.; Motti, E.; Minari, M. Symmetrical and unsymmetrical 2,6-dialkyl-1,1’-biaryls by combined catalysis of aromatic alkylation via palladacycles and Suzuki-type coupling, Chem. Commun. 2000, 157-158. (70) Bekkaye, M.; Masson, G. Synthesis of New Axially Chiral Iodoarenes. Synthesis, 2016, 48, 302-312. (71) Lindqvist, M.; Borre, K.; Axenov, K.; Kótai, B.; Nieger, M.; Leskelä, M.; Pápai, I.; Repo, T. Chiral Molecular Tweezers: Synthesis and Reactivity in Asymmetric Hydrogenation. J. Am. Chem. Soc. 2015, 137, 4038-4041. (72) Wang, Y.; Ji, B.-M.; Ding, K.-L. Synthesis of Aminophosphine Ligands with Binaphthyl Backbones for Silver(I)-catalyzed Enantioselective Allylation of Benzaldehyde. Chin. J. Chem. 2002, 20, 1300-1312. (73) CCDC 1825403 contains the supplementary crystallographic data for compound 9. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre. (74) Kočovský, P.; Vyskočil, Š.; Císařová, I.; Sejbal, J.; Tišlerová, I.; Smrčina, M.; Lloyd-Jones, G. C.; Stephen, S. C.; Butts, C. P.; Murray, M.; Langer, V. Palladium(II) Complexes of 2-Dimethylamino-2’diphenylphosphino-1,1’-binaphthyl (MAP) with Unique P,CσCoordination and Their Catalytic Activity in Allylic Substitution, Hartwig-Buchwald Amination, and Suzuki Coupling. J. Am. Chem. Soc. 1999, 121, 7714-7715. (75) CCDC 1825397 contains the supplementary crystallographic data for compound 4v. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre.

ACS Paragon Plus Environment

Page 7 of 7 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

ACS Paragon Plus Environment

7