Experimental and Computational Studies on ... - ACS Publications

Oct 25, 2017 - Arghya Deb, Sukriti Singh, Kapileswar Seth, Sandeep Pimparkar, Bangaru Bhaskararao, Srimanta Guin,. Raghavan B. Sunoj,* and Debabrata ...
11 downloads 0 Views 414KB Size
Subscriber access provided by READING UNIV

Article

Experimental and Computational Studies on Remote #-C(sp3)H Silylation and Germanylation of Aliphatic Carboxamides Arghya Deb, Sukriti Singh, Kapileswar Seth, Sandeep Pimparkar, Bangaru Bhaskararao, Srimanta Guin, Raghavan B. Sunoj, and Debabrata Maiti ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b03056 • Publication Date (Web): 25 Oct 2017 Downloaded from http://pubs.acs.org on October 25, 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

Experimental and Computational Studies on Remote γ-C(sp3)−H Silylation and Germanylation of Aliphatic Carboxamides Arghya Deb, Sukriti Singh, Kapileswar Seth, Sandeep Pimparkar, Bangaru Bhaskararao, Srimanta Guin, Raghavan B. Sunoj,* and Debabrata Maiti* Department of Chemistry, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India ABSTRACT: A Pd(II) catalyzed protocol for highly regioselective distal γ-C–H silylation and germanylation of aliphatic carboxylic acids has been disclosed. Bidentate 8aminoquinoline as the directing group was found to stabilize the six membered palladacycle. A variety of aliphatic carboxylic acids and amino acids were silylated and germanylated with good yield and high diasteroselectivity. Detailed mechanistic studies involving isolation of a Pd(II) intermediate, reaction rate and order determination, control experiments, isotopic labeling and DFT studies was found to be crucial for elucidating the elementary steps involved in this distal aliphatic functionalization. KEYWORDS: γ-C–H activation, silylation, germanylation, organopalladium crystal, mechanistic studies, DFT

silylation.13 In this context, earlier reports on C−H silylation of aliphatic alcohol were based on intramolecular approach.9d,9i,9j Intermolecular aliphatic silylation remained restricted to benzylic (sp3)C−H bond9a,9e or to substrates containing (sp3)C−H bonds located α to a heteroatom.9c,9h Directing group assisted approach was found to be essential for intermolecular silylation of unbiased aliphatic C−H bond.14 Kanai group first developed a palladium catalyzed intermolecular silylation protocol using aliphatic β-(sp3)C−H bond.14a An impressive procedure for β-(sp3)C−H silylation of phthalimido protected amino acids was recently reported by Zhang14b and Shi14c group. Herein we report a palladium silylation and catalyzed intermolecular (sp3)C−H germanylation15 reaction reaching at the farthest γ-position of aliphatic carboxylic acids (Scheme 1). a)

Transition

metal catalyzed selective aliphatic C−H functionalization has opened new avenues for synthesizing complex molecular scaffolds.1 Despite the challenges in sp3 C−H bond activation, a number of metal catalyzed protocols have provided avenues to controlling selectivity.2 In this regard, palladium catalyst played a pivotal role toward predictably selective C−H functionalizations.3 In the last decade, efforts for directed C(sp3)−H activation of carboxylic acid derivatives,4 imines5 and aliphatic amines6 was explored with five membered palladacycle. However, directing group assisted approaches to functionalize distal γ-C(sp3)−H bond of aliphatic carboxylic acid has remained challenging. Despite significant advances, distal C−H activation in aliphatic carboxylic acids has remained elusive.7,8 Although a variety of palladium catalyzed aliphatic functionalizations has already been reported, the scarcity of these functionalizations at distal positions may be attributed to the involvement of less favored palladacycle intermediates. The potential of C(sp3)−H activation reactions can be exploited more by introducing new functional groups at distal carbon of aliphatic acids. Driven by such motivation, we began exploring various carbon-heteroatom bond forming reactions to selectively introduce different functional groups at the distal position of aliphatic acids. Out of different carbon-heteroatom bond forming reactions, aliphatic C−H silylation9 is one of the most popular fields of research owing to the synthetic utility in addition to unique chemical and physical properties of organosilicon compounds.10 The core structural units of many important therapeutical agents contain organosilicon moieties.11 Incorporation of silyl group into α- amino acids leads to notable change in its chemical properties.12 Expectedly, transition metal catalyzed strategies have emerged as one of the most powerful tools for widening the scope of C−H silylation.9, 13 Due to the higher steric hindrance of bulky trimethylsilyl group, aliphatic (sp3)C−H silylation is even more challenging as compared to (sp2)C−H

DG

O

Me3X XMe3

PdII

Me R1

O

DG removal

R2

− −

Remote γ C H Silylation and Germanylation Detailed mechanistc and DFT studies

CCDC 1549713

H2N

COOH L valine Me

Me

H2N

COOH

COOH L isoleucine Me

H2N

XMe3

Me

H2N Me

OH γ β XMe3 R1 R2 X = Si, Ge

α

H2N L tert leucine Me

XMe3

Me

Amino acid diversification b)

Me Me c)

Me

Me Me

XMe3 Me

O Ar PdII DG Me3Si SiMe3 Ar SiMe3 − − Sequential γ C H functionalization

Ar

DG Ar

Me

O

Me Ar

PdII Me3Si SiMe3

DG

PdII D4 AcOH

− − γ C H deuteration

Me

DG DG Ar SiMe3 Me - ua ernar cen re mme r ruc uns ons o c on t ti βq t t ti f y y C O

DG

Me

COOH

O PdII Ar I

Me

Me Me

PdII Ar I

DG

O

d) O

O

DG

O

Me Me

H2N

COOH

Me

COOH

CD3 D3C

CD3

O PdII Me3Si SiMe3

DG D2 C

D3C

SiMe3 CD3

Scheme 1. Directed γ-C−H Functionalization In continuation with our efforts to promote remote (sp3)C−H functionalization, we chose 8-aminoquinolamide16 of 3,3-dimethyl-butanoic acid as the model substrate for promoting palladium catalyzed γ-(sp3)C−H silylation and germanylation of simple aliphatic carboxylic acids. To our delight, a trace amount of γ-silylated product was obtained by using Pd(OAc)2 and hexamethyldisilane (HMDS) in dioxane. Detailed optimization was resulted highly regioselectiveγ-silylated product by using Pd(OPiv)2 as the catalyst in the presence of Ag2CO3 and NaHCO3 in tBuOH or dioxane.17 Since exogenous ligands are known to facilitate C−H and accelerate the metal catalyzed sp3 functionalizations, various ligands were systematically introduced to study their effect (Scheme 2).3c,3d,17 Inclusion of

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

2-chloroquinoline was found to provide the best yield under the present reaction condition (Scheme 2). 10 mol% Pd(OPiv)2 O 20 mol% Ligand R 3 equiv Ag2CO3 NHQ R1 2 equiv NaHCO3 SiMe3 R1 2 mL tBuOH, 130 °C, 24 h 2

O R R1 R1

NHQ + Me e 3Si SiM 3 e M (5 equiv) 1

a)

Page 2 of 6

+ Me3Si-SiMe3 Me (5 equiv) Me

Ar

N

3 O Me

QHN

O Me

QHN

Q

O Me

QHN

O Me

QHN

SiMe3 Ar

4

O Me

QHN SiMe3

SiMe3

SiMe3

SiMe3

O Me

QHN 10 mol% Pd(OPiv)2 20 mol% L4 3 equiv Ag2CO3 2 equiv NaHCO3 2 mL tBuOH, 130 °C, 24 h

NHQ

O

SiMe3

Br without ligand N

N Cl L2 45%

F

L1 45%

53%

N

F N L3 7%

L4 71%

O Me

QHN

F3C N

N

L6 5%

L5 19%

N

OMe

N

L7 15%

L8 25%

SiMe3

SiMe3

L9 14%

4f, 66%

b)

N

Me

4g, 72%

OMe

4h, 69%

CF3

Ar

O

Me

O

MeO2C

NHQ + Me e 3Si SiM 3 Me (5 equiv) 1

O Me

QHN

NHQ

Me 2a, 71 %

Me

NHQ , Me 2d 63 %

Me Me O

O

Me Me3Si

QHN

6a, 54%

6b, 57%

MeO

NHQ Me3Si CO2Me 6c, 65%

EtO

F3C

O

O NHQ

Me3Si

6d, 52%

O NHQ

OMe

Me3Si

O

COMe

NHQ

OEt F3C Me3Si

6e, 61%

CF3

6f, 61% CF3

SiMe3

PhthN

O

O

MeOC

Me3Si

Me3Si

10 mol% Pd(OPiv)2 O 20 mol% L4 R v u e 3 q i Ag2CO3 NHQ R1 N 2 equiv NaHCO3 SiMe3 , , ° m u R1 2 L tB OH 130 C 24 h Q 2 NPhth= N Phthalimide

DG SiMe3

Ar

NHQ

O

Ar

6 O

Table 1. Scope of Aliphatic γ-C−H Silylation17

QHN

CO2Et

Condition: same as above

+ Me3Si-SiMe3 (5 equiv)

DG

L11 14%

Scheme 2. Ligand Screening for γ-C−H Silylation17

R1

4i, 72%

OEt

O Ar

5

R

SiMe3

OH N O L10 56%

R1

O Me

QHN

Br

N

Me tBu

NO2 COMe 4d, 75% 4e, 71%

O Me

QHN

O Me

QHN SiMe3

NH2

N

OH

tBu 4c, 61%

OMe 4b, 65%

CHO

4a, 67%

Cl

PhthN Me

SiMe3

SiMe3 Me 2c, 65 %

2b, 63 %

NHQ 2e, 69 % d r 18:1 SiMe3

O PhthN 2f, 70 % d r 37:1

Me

NHQ SiMe3

Subsequently, we have explored the scope of this selective silylation protocol with various aliphatic acetamides (Table 1). Expectedly, 8-amino quinolamide of 3,5,5trimethylhexanoic acetamide (2b) and isovaleric acid (2c) gave selective mono γ-silylated product. Owing to the importance of silicon-containing amino acids in biological research,12 a number of naturally occurring α-amino acids like L-valine (2f), L-isoleucine (2g) and tert-butyl leucine (2d) have been γ-silylated by following our method. Table 2. Scope of γ-C−H Silylation Quinolamides17

with γ-Arylated

After exploring various commercially available aliphatic carboxylic acids, we further thought to explore the scope of this γ-silylation reaction by using the γ-arylated product from our earlier report.8 A number of electronically different γarylated product of 3,3-dimethyl-butanamide successfully participated in γ-silylation (Table 2a). A modified quaternary carbon centre was generated by the successful γ-C−H silylation of the γ-di-arylated carboxylic acid. To the best of our knowledge, generation of such quaternary substituted carbon centres as well as desymmetrization of aliphatic system by iterative γ-C−H activation is unprecedented in literature. Accessing these quaternary γ-silylated molecules by this protocol is, therefore, expected to enhance the significance of the current method. Table 3. Scope of Aliphatic γ-C−H Germanylation17 NHQ

O

Me

R

+ Me3Ge GeMe3 (5 equiv)

R1 R2 7 QHN

10 mol% Pd(OPiv)2 R 20 mol% L4 R v u 1 3 eq i Ag2CO3 R1 2 equiv NaHCO3 , , ° 2 mL tBuOH 130 C 24 h

QHN

O Me GeMe3 Me 8a, 61 %

Me

Me Me3Ge

Me Me

O PhthN

O

O NHQ GeMe3 8 O

QHN GeMe3

GeMe3 Me , c 8 51 %

8b, 57 %

O

O NHQ

Me

8d, 63 %

PhthN Me

NHQ 8e, 69 % d r 19:1

GeMe3

PhthN 8f, 59 % d r 20:1

Me

NHQ GeMe3

Owing to the closely similar reactivites of silicon and germanium15, a further extension of the present γ-C−H functionalization protocol was carried out with

ACS Paragon Plus Environment

Page 3 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

hexamethyldigermane. Aliphatic carboxylic acids and amino acids resulted in the γ-germanylated product with high diastereoselectivity. This exemplary set of γ-germanylated products represents new chemical entities as well as reflects the generality of the present protocol (Table 3). Removal of 8-aminoquinoline directing group by acid calalyzed hydrodrolysis resulted 90% yield of silylated aliphatic ester.17

considered as the starting point of the catalytic cycle (Scheme 4). 2-chloro quinolone acts as an effective ligand in the active catalyst and in the formation of catalyst-substrate complex. N

O N Pd II

O

Me

Me

CCDC 1549713

1

PivO

O NPhth

QHN

SiMe3 Me isolated yield 65%

Oxidation

O

2

Pd(0)

Ag

O D 20% D D3C

NH CD3 CD3

10 mol% Pd(OPiv)2 D N 20 mol% L4 100% NH O Me3Si SiMe3 v u e 3 q i Ag2CO3 D (5 equiv) CD3 2 equiv NaHCO3 20% D 2 mL tBuOH, 130 °C, time D2C CD3

D 100%

NH O 20% D CD3 D 91% D3C CD3 D1

. kH/kD = 1 07

− C H Activation

HN O N

8

S

SiMe3

Scheme 3. Mechanistic studies for γ-C−H silylation17 A stoichiometric reaction of valine quinolamide and Pd(OAc)2 resulted a 6-membered palladacycle intermediate via γ-C−H activation which was confirmed by X-ray crystallography (Scheme 3). The organopalladium species (9) was found to be chemically and kinetically competent with Kinetic studies revealed first order rate Pd(OPiv)2. dependency of the substrate for this γ-silylation reaction. Yield of the product was monitored by gas chromatography and was plotted against time. Due to the excess use (5 equiv) of HMDS, steady state approximation was considered for the same. For intermolecular KIE experiment, gram scale synthesis of D1-quinolamide (D1-1) was carried out by treating quinolamide with Pd(OAc)2 and D4-AcOH at 100 oC. A 91% deuterium incorporation in each γ-methyl group hypothesized the reversibility of this γ-C−H activation process. A small KIE value of 1.07 further supported our hypothesis (Scheme 3). To develop a better understanding on the mechanism of this catalytic transformation, we have performed density functional theory computations at the SMD/M06/6-31G**, SDD (Pd, Ag) level of theory.18a The key intermediates and transition states of the catalytic cycle were fully optimized in condensed phase using tert-butanol as the solvent continuum. A Pd(II) species with two 2-chloro quinoline ligands and two pivalate groups trans to each other Pd(OPiv)2L2 (1) is noted as energetically most favorable active species (4.2 kcal/mol).18b Out of different possibilities, the substrate binds as an amidate to the pre-catalyst 1 to form a catalyst-substrate chelate complex (2)18c which is

- ‡ e [3 4] M

N Pd

S O

Ag

4

Me Me Me

S Me Me

L

O Me Me

N Pd

7

N

O

O

Ag

O

Me

PivOH

O

PivOH + Ag CO 2 3

N

N

Pd L

Me Me

Me Me

O N Pd N

6 SiMe3

75% D

Me Me Me

H O

Me3Si

75% D 61% D

N

O

3

S

- ‡ [6 7]

91%

O

Ag O

Reductive Elimination

D1

Pd

S

O

Pd

Me3Si

Me3Si

D 100%

N

H

O

Me3Si

N

O

L

N

Me Me

Me3Si

D 75%

d) KIE experiment with labeled D 1 61% D 75% D 75% D N

L

N

S

10 mol% Pd(OAc)2 . 2 equiv Na2CO3 2 mL D4 AcOH, 120oC, 72 h

Me

Ag2CO3

Pd

3

61% D

NH

N H

Piv

N

N

PivO

+ PivOSiMe

b) Chemical and kinetic competence studies O 10 mol% Intermediate 9 NPhth 20 mol% L4 QHN + Me3Si-SiMe3 3 equiv Ag2CO3 v u e Me Me (5 q i ) 2 equiv NaHCO3 2 mL tBuOH 130 °C, 24 h c) Preparation of D 1 D 75%

O

(2b) NaOPiv + L

O

Me3Si

O

N

O Me Me

Me O Intermediate 9 [X Ray]

N

1

+ OPiv

N

N Me

Cl

O Na

L Pd

L

S = t BuOH (solvent)

a) Crystal structure of 6-membered palladacycle

N

PivO

L=

N

Oxidative e Addition M 3Si

- ‡ [5 6]

O

N L

4a

Me3Si SiMe3

Pd N Si Me3

5

Scheme 4. Mechanistic Cycle for Palladium Catalyzed γSilylation and Germanylation A concerted metalation deprotonation (CMD) in 3 takes place to form a six-membered palladacycle intermediate 4. We note that the carbonate ligand is involved in the ligandassisted C−H activation (free energy 19.2 kcal/mol for [3-4]‡). A direct pivalate-assisted CMD is found to be 4.3 kcal/mol higher in energy.18d In 4, ligand (L) can replace the labile pivalic acid and silver carbonate (formed by combining silver pivalate and silver bicarbonate), to give another intermediate 4a. Interestingly, we could characterize a palladacycle intermediate (9, Scheme 3a) similar to 4a by using X-ray crystallography. Independent geometry optimization of 9 revealed very good agreement with the experimental structure (A root mean square deviation (RMSD) of only 1.274 Å is noted between the computed and X-ray structures for 9; See Figure S5 in the Supporting Information). The second substrate, (SiMe3)2 (or (GeMe3)2) can now displace ligand (L) from 4a and bind to the Pd(II) center through a weak interaction with the two silyl atoms in disilane (or Ge) (intermediate 5). It is important to note that the structure and energetics of transition states for the insertion of disilane derivatives to a palladium center is seldom reported.19 Two possible modes of silylation were considered, namely, (i) a lower energy oxidative addition, and (ii) a higher energy transmetallation.18e In the crucial step, an oxidative addition of Pd(II) to the disilane bond in HMDS takes place via [5-6]‡ to form 6 (23.0 kcal/mol).18e, 20 The intermediate 6 can now undergo reductive elimination via [6-7]‡ to give the silylated intermediate 7 (20.5 kcal/mol). The PivOH formed during the C−H activation step may interact with intermediate 7, initially through a hydrogen-bonding and subsequently transfer its proton to the amidate nitrogen to form productcatalyst complex 8. Thus, the computed energetics reveal that the silylation/germanylation mechanism involves a Pd(II)/Pd(IV) redox cycle.18

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

Certain vital insights emerged through our computational study and its relevance to the experimental observations are worth considering at this juncture. It can be noted (Figure 1) that the highest energy point on the potential energy profiles for silylation (blue) and germanylation (green) are respectively the oxidative addition and reductive elimination TSs.18f This prediction is in concert with our low KIE values for silylation (1.07) and germanylation (1.27). A closer look at the energy profile reveals that the oxidative addition is the rate-determining step for silylation while it is the reductive elimination in germanylation reaction.

Figure 1. Gibbs free energy profile (kcal/mol) of the γC(sp3)−H silylation and germanylation obtained at the SMD/M06/6-31G**, SDD(Pd) level of theory. In summary, we have developed a novel method for distal γ-C(sp3)−H silylation and germanylation of aliphatic acids. Preliminary mechanistic studies with organopalladium intermediate suggested a facile γ-C−H activation process with subsequent silylation. Detailed studies and further development towards the stereoselective silylation by rational design of directing groups is currently underway in our laboratory. ACKNOWLEDGEMENT This research was supported by SERB(EMR/2015/000164). Fellowships from CSIR-India (AD), UGC-India (BB), IITB (SS), SERB (KS and SG) and SpaceTime supercomputing facility are gratefully acknowledged. AUTHOR INFORMATION Corresponding Authors: [email protected] (experimental studies), [email protected] (computational studies) Supporting Information Available: Experimental details, computational methods and characterization data. This material is available free of charge via the Internet at http://pubs.acs.org

REFERENCES 1.

For selected examples, see: (a) Feng, Y.; Chen, G. Angew. Chem., Int. Ed. 2010, 49, 958-961. (b) Miao, J.; Ge, H. Eur. J. Org. Chem. 2015, 2015, 7859-7868. (c) Topczewski, J. J.; Cabrera, P. J.; Saper, N. I.; Sanford, M. S. Nature 2016, 531, 220-224. (d) Chapman, L. M.; Beck, J. C.; Wu, L.; Reisman, S. E. J. Am. Chem. Soc. 2016, 138, 9803-9806. (e) Gemoets, H. P. L.; Kalvet, I.; Nyuchev, A. V.; Erdmann, N.; Hessel, V.; Schoenebeck, F.; Noël, T. Chem. Sci. 2017, 8, 1046-1055. (f) Gemoets, H. P. L.; Laudadio, G.; Verstraete,

Page 4 of 6

K.; Hessel, V.; Noël, T. Angew. Chem., Int. Ed., 2017, 56, 71617165. 2. For selected examples, see: (a) Jazzar, R.; Hitce, J.; Renaudat, A.; Sofack-Kreutzer, J.; Baudoin, O. Chem. -Eur. J. 2010, 16, 26542672. (b) Baudoin, O. Chem. Soc. Rev. 2011, 40, 4902-4911. 3. For selected examples, see: (a) Dastbaravardeh, N.; Christakakou, M.; Haider, M.; Schnürch, M. Synthesis 2014, 46, 1421-1439. (b) Pedroni, J.; Cramer, N. Chem. Commun. 2015, 51, 17647-17657. (c) He, J.; Wasa, M.; Chan, S. L. K.; Shao, Q.; Yu, J.-Q. Chem. Rev. 2017, 117, 8754-8786. (d) Kapdi, A.; Maiti, D. Strategies for Palladium-Catalyzed Non-directed and Directed C−H Bond Functionalization 2017, ISBN: 9780128052549. 4. For selected examples see: (a) Shabashov, D.; Daugulis, O. J. Am. Chem. Soc. 2010, 132, 3965-3972. (b) Ano, Y.; Tobisu, M.; Chatani, N. J. Am. Chem. Soc. 2011, 133, 12984-12986. (c) Hoshiya, N.; Kobayashi, T.; Arisawa, M.; Shuto, S. Org. Lett. 2013, 15, 62026205. (d) Zhang, S.-Y.; Li, Q.; He, G.; Nack, W. A.; Chen, G. J. Am. Chem. Soc. 2013, 135, 12135-12141.(e)Yan, S.-B.; Zhang, S.; Duan, W.-L. Org. Lett. 2015, 17, 2458-2461. (f) Shan, G.; Huang, G.; Rao, Y. Org. Biomol. Chem. 2015, 13, 697-701. (g) Gopalakrishnan, B.; Mohan, S.; Parella, R.; Babu, S. A. J. Org. Chem. 2016, 81, 89889005. 5. For selected examples, see: (a) Desai, L. V.; Hull, K. L.; Sanford, M. S. J. Am. Chem. Soc. 2004, 126, 9542-9543. (b) Thu, H.-Y.; Yu, W.Y.; Che, C.-M. J. Am. Chem. Soc. 2006, 128, 9048-9049. (c) Ren, Z.; Mo, F.; Dong, G. J. Am. Chem. Soc. 2012, 134, 16991-16994. (d) Sharpe, R. J.; Johnson, J. S. J. Am. Chem. Soc. 2015, 137, 49684971. 6. For selected examples, see: (a) Zaitsev, V. G.; Shabashov, D.; Daugulis, O. J. Am. Chem. Soc. 2005, 127, 13154-13155. (b) Zhang, S.-Y.; He, G.; Nack, W. A.; Zhao, Y.; Li, Q.; Chen, G. J. Am. Chem. Soc. 2013, 135, 2124-2127. 7. For selected examples, see: (a) Reddy, B. V. S.; Reddy, L. R.; Corey, E. J. Org. Lett. 2006, 8, 3391-3394. (b) Rodriguez, N.; Romero-Revilla, J. A.; Fernandez-Ibanez, M. A.; Carretero, J. C. Chem. Sci. 2013, 4, 175-179. (c) He, G.; Zhang, S.-Y.; Nack, W. A.; Li, Q.; Chen, G. Angew. Chem., Int. Ed. 2013, 52, 11124-11128. (d) He, G.; Zhang, S.-Y.; Nack, W. A.; Pearson, R.; Rabb-Lynch, J.; Chen, G. Org. Lett. 2014, 16, 6488-6491. (e) Li, S.; Chen, G.; Feng, C.-G.; Gong, W.; Yu, J.-Q. J. Am. Chem. Soc. 2014, 136, 52675270. (f) Li, S.; Zhu, R.-Y.; Xiao, K.-J.; Yu, J.-Q. Angew. Chem., Int. Ed. 2016, 55, 4317-4321. (g) Xu, J.-W.; Zhang, Z.-Z.; Rao, W.H.; Shi, B.-F. J. Am. Chem. Soc. 2016, 138, 10750-10753. (h) Liu, B.; Shi, B.-F. Synlett, 2016, 27, 2396-2400. (i) Ling, P.-X.; Fang, S.L.; Yin, X.-S.; Zhang, Q.; Chen, K.; Shi, B.-F. Chem. Commun. 2017, 53, 6351-6454. (j) We are involved in developing distal C-H functionalization of aliphatic substrates. For example, remote γ-sp3 C−H thioarylation and selenoarylation were successfully developed in our laboratory. 8. Dey, A.; Pimparkar, S.; Deb, A.; Guin, S.; Maiti, D. Adv. Synth. Catal. 2017, 359, 1301-1307. 9. (a) Kakiuchi, F.; Tsuchiya, K.; Matsumoto, M.; Mizushima, E.; Chatani, N. J. Am. Chem. Soc. 2004, 126, 12792-12793. (b) Larsson, J. M.; Zhao, T. S. N.; Szabó, K. J. Org. Lett. 2011, 13, 1888-1891. (c) Ihara, H.; Ueda, A.; Suginome, M. Chem. Lett. 2011, 40, 916918. (d) Simmons, E. M.; Hartwig, J. F. Nature, 2012, 483, 70-73. (e) Mita, T.; Michigami, K.; Sato, Y. Org. Lett. 2012, 14, 34623465. (f) Frihed, T. G.; Heuckendorff, M.; Pedersen, C. M.; Bols, M. Angew. Chem., Int. Ed. 2012, 51, 12285-12288. (g) Kuninobu, Y.; Nakahara, T.; Takeshima, H.; Takai, K. Org. Lett. 2013, 15, 426428. (h) Mita, T.; Michigami, K.; Sato, Y. Chem.−Asian J. 2013, 8, 2970-2973. (i) Ghavtadze, N.; Melkonyan, F. S.; Gulevich, A. V.; Huang, C.; Gevorgyan, V. Nat. Chem. 2014, 6, 122-125. (j) Li, B.; Driess, M.; Hartwig, J. F. J. Am. Chem. Soc. 2014, 136, 6586-6589. (k) Li, Q.; Driess, M.; Hartwig, J. F. Angew. Chem., Int. Ed. 2014, 53, 8471-8474. (l) Li, W.; Huang, X.; You, J. Org. Lett. 2016, 18, 666-668. (m) Parija, A.; Sunoj, R. B. Org. Lett. 2013, 15, 40664069. 10. (a) Rappoport, Z., Apeloig, Y., Eds. Chemistry of Organosilicon Compounds; Wiley-VCH: New York, 2001; Vol. 3. 1-1172 (b) Liu, X.-M.; He, C.; Xu, J. Chem. Mater. 2005, 17, 434-441. (c) Sanchez J. C.; Trogler, W. C. Macromol. Chem. Phys. 2008, 209, 1527-1540. (d) Iida, A.; Nagura, K.; Yamaguchi, S. Chem.−Asian J. 2008, 3, 1456-1464. (e) Bai, D.; Han, S.; Lu, Z.-H.; Wang, S. Can. J. Chem.

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

11.

12.

13.

14.

15.

16.

17. 18.

19. 20.

ACS Catalysis 2008, 86, 230-237. (f) Nakao, Y.; Hiyama, T. Chem. Soc. Rev. 2011, 40, 4893-4901. (a) Tacke, R.; Heinrich, T.; Bertermann, R.; Burschka, C.; Hamacher, A.; Kassack, M. U. Organometallics 2004, 23, 44684477. (b) Barnes, M. J.; Conroy, R.; Miller, D. J.; Mills, J. S.; Montana, J. G.; Pooni, P. K.; Showell, G. A.; Walsh, L. M.; Warneck, J. B. H. Bioorg. Med. Chem. Lett. 2007, 17, 354-357. (a) Mortensen, M.; Husmann, R.; Veri, E.; Bolm, C. Chem. Soc. Rev. 2009, 38, 1002-1010. (b) Franz, A. K.; Wilson, S. O. J. Med. Chem. 2013, 56, 388-405. (a) Hartwig, J. F. Acc. Chem. Res. 2012, 45, 864-873. (b) Cheng, C.; Hartwig, J. F. Chem. Rev. 2015, 115, 8946-8975 and the references cited therein. (a) Kanyiva, K. S.; Kuninobu, Y.; Kanai, M. Org. Lett. 2014, 16, 1968-1971. (b) Pan, J.-L.; Li, Q.-Z.; Zhang, T.-Y.; Hou, S.-H.; Kang, J.-C.; Zhang, S.-Y. Chem. Commun. 2016, 52, 13151-13154. (c) Liu, Y.-J.; Liu, Y.-H. Zhang, Z.-Z.; Yan, S.-Y. Chen, K.; Shi, B.-F. Angew. Chem., Int. Ed. 2016, 55, 13859-13862. For selected examples of synthetic application of organogermanes, see: (a) Kinoshita, H.; Nakamura, T.; Kakiya, H.; Shinokubo, H.; Matsubara, S.; Oshima, K. Org. Lett. 2001, 3, 2521-3167. (b) Nakamura, T.; Yorimitsu, H.; Shinokubo, H.; Oshima, K. Tetrahedron 2001, 57, 9827-9836. (c) Kinoshita, H.; Shinokubo, H.; Oshima, K. J. Am. Chem. Soc. 2002, 124, 4220-4221. (d) Nakamura, T.; Kinoshita, H.; Shinokubo, H.; Oshima, K. Org. Lett. 2002, 4, 3165-3167. (e) Chen, C.; Guan, M.; Zhang, J.; Wen, Z.; Zhao, Y. Org. Lett. 2015, 17, 3646-3649. (a) Rouquet, G.; Chatani, N. Angew. Chem., Int. Ed. 2013, 52, 11726-11743. (b) Corbet, M.; Campo, D. F. Angew. Chem., Int. Ed. 2013, 52, 9896-9898. (c) Daugulis, O.; Roane, J.; Tran, L. D. Acc. Chem. Res. 2015, 48, 1053-1064. See supporting information for experimental details. (a) Gaussian 09 quantum chemical suite has been used for the computations. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross,J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman,J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision D.01. Gaussian, Inc., Wallingford, CT, 2013. See Supporting Information for (b) additional possibilities of the active catalyst (Figure S1), (c) more details of substrate binding (Scheme S1), (d) other higher energy pathways for C−H activation (Figure S2), (e) higher energy transmetallation (Figure S3), and (f) details on germanylation (Figure S4) Bottoni, A.; Higueruelo, A. P.; Miscione, G. P. J. Am. Chem. Soc. 2002, 124, 5506-5513. (a) Tanabe, M.; Mawatari, A.; Osakada, K. Organometallics 2007, 26, 2937-2940. (b) Pan, Y.; Mague, J. T.; Fink, M. J. Organometallics 1992, 11, 3495-3497.

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

TOC Graphic DG

O

Me3X XMe3

PdII

Me R1

O

OH γ XMe3

R2

R1

6 Membered Palladacycle

R2

X = Si, Ge 27 examples Excellent yield and selectivity Mechanistic studies DFT calculations

ACS Paragon Plus Environment

Page 6 of 6