RhCl3.3H2O-Catalyzed Ligand-Enabled Highly Regioselective

7 hours ago - It is noteworthy that [Cp*RhCl2]2 was not as efficient as RhCl3 3H2O in this strategy. Furthermore, the product could be transferred to ...
0 downloads 0 Views 387KB Size
Subscriber access provided by Nottingham Trent University

Article

RhCl3.3H2O-Catalyzed Ligand-Enabled Highly Regioselective Thiolation of Acrylic Acids Can Liu, Yi Fang, Shun-Yi Wang, and Shun-Jun Ji ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.9b02982 • Publication Date (Web): 22 Aug 2019 Downloaded from pubs.acs.org on August 22, 2019

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 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

RhCl33H2O-Catalyzed Ligand-Enabled Highly Regioselective Thiolation of Acrylic Acids Can Liu,†,‡ Yi Fang,†,‡ Shun-Yi Wang,*,†,‡ and Shun-Jun Ji*,†,‡ † Key

Laboratory of Organic Synthesis of Jiangsu Province, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou, 215123, China. ‡Collaborative

Innovation Center of Suzhou Nano Science and Technology, Soochow University, Suzhou, 215123, China.

Supporting Information Placeholder ABSTRACT: A simple inorganic rhodium salt RhCl33H2O was used as the

catalyst to achieve the thiolation of acrylic acids, with a series of (Z)-alkenyl sulfides obtained exclusively. It is noteworthy that [Cp*RhCl2]2 was not as efficient as RhCl33H2O in this strategy. Furthermore, the product could be transferred to biological and pharmacological thioflavone conveniently.

R1

COOH

R2

H

INTRODUCTION

RhCl3 3H2O

R1

COOH

R2

SR3

* Higher catalytic efficiency than [Cp*RhCl2]2 * Easier derivative carboxyl group as DG

KEYWORDS: RhCl33H2O, carboxylate-directed, alkenyl sulfide, C-H activation, regioselectivity



R3SSR3

(Scheme 1).3 Nevertheless, limited research has been conducted on non-aromatic C(sp2)-H activation, and aromatic reaction still predominates in current research. Therefore, developing carboxylate-directed non-aromatic C(sp2)-H activation has important practical significance.

C-H activation has attracted much attention in recent years due to its high atom economy and diversity of reaction, and has become one of the important strategies of organic synthesis. However, in order to ensure regioselectivity, directing group (DG) is usually indispensable. Strong directing groups such as pyridine, pyrimidine and 8-aminoquinoline were first used in C-H activation reactions.1 Soon after, weak guiding groups such as oxime ether, ketone and amide have also been reported.2 Since 2007, carboxyl group has been initially used as directing group of C-H activation reaction because of its non-toxicity, easy removal, widespread existence in nature and abundant derivative reactions. During the past decade, carboxyl group show extraordinary talents in this field. Palladium-, rhodium-, ruthenium- and iridiumcatalyzed systems have been developed successfully to achieve arylation, alkylation, alkenylation, thiolation and selenation

[Cp*RhCl2]2 (Cp* = 1,2,3,4,5-pentamethylcyclopentadienyl) is the most used catalyst in existing rhodium-catalyzed C-H activation reactions. The Cp* ligand is believed to be helpful in stabilizing rhodium intermediates, as well as promoting eliminating processes due to the steric hindrance effect. Indeed, a large number of reported facts prove that the catalyst has high efficiency, regioselectivity and excellent functional group tolerance.4 [Cp*RhCl2]2 is generally prepared from the coordination reaction of RhCl33H2O with excess 1,2,3,4,5pentamethylcyclopentadiene in methanol under an inert atmosphere. However, as the precursor of [Cp*RhCl2]2, COOH

R Ar 2015, Gooßen, ref 3m COOH

COOH

R

COOH

R R

XR' X = S or Se

HOOC

[Ru(p-cymene)Cl2]2

COOH R

Ar

Ar 2007, Daugulis, ref 3a

2017, Baidya, ref 3i

2018, Baidya, ref 3h

R

R

[IrCp*Cl2]2

2011, Larrosa, ref 3b

Pd(OAc)2 O COOH

COOH

R

R

O O

N R'

n

O

[Rh(nbd)Cl2]2

[RhCp*Cl2]2

2017, Ackermann & Baidya ref 3j & ref 3k

2018, Gooßen, ref 3l

CH2OBn

Bu

2013, Yu, ref 3c

R'

O

R

R

2015, Yu, ref 3d

R Het 2015, Su, ref 3f O R

COOH R R HOOC 2015, Li, ref 3e

HOOC 2019, Gooßen, ref 3g

Scheme 1. Carboxylate-directed C-H activation of benzoic acids under Pd, Rh, Ru and Ir catalysis

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

RhCl33H2O was used to be the catalyst of C-H activation rarely.5 In 1990, the RhCl(PPh3)3/Cu2+ redox couple was used in carboxylation of aromatic rings. Inspired by this work, Witulski and cooperators used RhCl33H2O and Cu(II) salts to achieve intramolecular C-H/C-H Coupling.6 Since then, there are only a few articles which have reported RhCl33H2O-catalyzed C-H activation and relevant research.7 Alkenyl sulfides are a class of important sulfur-containing compounds, which are widely used in organic synthesis, material science and biopharmaceuticals.8 With the development of related research, its synthesis is becoming more simple and efficient. Initial synthetic strategies required unfriendly sulfur sources, low atomic economy, poor selectivity and harsh reaction conditions (Scheme 2, A & B),9 but now these shortcomings have been overcome. In 2018, we reported an N-tosylamide-assisted synthesis of (Z)--alkenyl sulfides through [Cp*RhCl2]2-catalyzed thiolation of alkenes with high efficiency and good substrate tolerance (Scheme 2, C).10 However, we failed to remove the directing group, which hinders the subsequent application. Therefore, we hope to develop a system with directing group which is easier to remove or derive. Meanwhile, we also desire to apply RhCl33H2O to this system with the help of ligand. Herein, we demonstrated a RhCl3.3H2O-catalyzed ligand-enabled highly regioselective thiolation of acrylic acids to afford (Z)--alkenyl sulfides(Scheme 2, D). Scheme 2. Synthesis of alkenyl sulfides through different processes. A. Nucleophilic addition of thiols to alkynes R1SH

R2

R 1S

R2

1

R S

R2

smelly and toxic thiol poor regioselectivity

B. Cross-coupling of thiols and vinyl halides X R1SH

R2 or

X

R 1S

R2

[M]

R2

harsh conditions low atom economy

or R 1S

R2

C. Our previous work: regioselective synthesis of (Z)--alkenyl sulfides O R1 R

N H

2

Ts

H

R3XXR3

[Cp*RhCl2]2

X = S, Se

O R1 R

2

N H XR3

Ts

difficult-removed DG complicated catalyst

Page 2 of 6

the yield up to 75% further (Table 1, entry 16). By the following investigation, reaction time and temperature were determined to 6 h and 120 oC, respectively. Under this condition, the ligand was replaced by cheaper tris(2,6-dimethoxyphenyl)phosphane (L37), and the yield could still maintain at 79% (Table 1, entry 20). In addition, when [Cp*RhCl2]2 was employed in this reaction, the yield decreased to 69% (Table 1, entry 21). Table 1. Optimization of the Rh-catalyzed thiolation of 1a with 2aa H PhSSPh

Cu(OAc)2

ligand

T (oC)

Time(h)

solvent, T, 6 h

2a

4a

1

/

/

120

12

0

2

10

/

120

12

17

3

20

/

120

12

27

4

50

/

120

12

40

5

100

/

120

12

49

6

150

/

120

12

38

7

100

L2

120

12

25

8

100

L8

120

12

25

9

100

L19

120

12

52

10

100

L33

120

12

57

11

100

L34

120

12

59

12

100

L35

120

12

37

13

100

L36

120

12

71

14

100

L36

80

12

Trace

15

100

L36

100

12

73

16c

100

L36

120

12

75

17 c

100

L36

140

12

72

18 c

100

L36

120

6

80

19 c

100

L36

120

18

75

20 c

100

L37

120

6

79(77 d)

21 c,e

100

L37

120

6

69

PPh2 PPh2

NHAc

R2

COOH

N

H

COOH

N

R3XXR3 X = S, Se

RhCl3 3H2O

R1 R2

COOH

L2

XR3

RESULTS AND DISCUSSION

O

Initially, atropic acid 1a and diphenyl disulfide 2a were chose to investigate the reaction conditions (Table 1, see Supporting Information for more details). Rh catalyst could not be replaced by other transition metals such as Pd, Co, Ni, Ru. The reaction could not undergo in common solvents such as ether, benzene and alcohols other than amide solvents and MeCN. The amount and type of Cu salts are key factors of the reaction. The yield peaked when 1 equiv Cu(OAc)2 was added (Table 1, entry 5). Subsequently, different ligands were applied to the reaction. Phosphine ligands were found to be beneficial to the reaction. The desired product could be obtained in 68% yield when 30 mol % tris(2,4,6-trimethoxyphenyl)phosphane (L36) was added. And the yield would increase to 71% when the amount of ligand reduced to 15 mol % (Table 1, entry 13). The increment of 2a also brings

O

O P O

simple DG simple catalyst

O



LC-Yield(%)b

(mol %)

D. This work R1

COOH

COOH 1a

Entry

SPh

RhCl3 3H2O (5 mol %) Cu(OAc)2 Ligand

P

O

L8

L19

O

O

O

L34

O O

P

O O

O O

O O P

P O

L35

L33

OO O

OO

L36

O

L37

aReaction

conditions: 1a (0.1 mmol, 1.0 equiv), 2a (0.1 mmol, 1.0 equiv), RhCl33H2O (5 mol %), ligand (15 mol %), Cu(OAc)2(1 equiv), DMF (1 mL), 120 oC, 12 h. bLC yield using biphenyl as the internal standard. cPhSSPh 2 equiv. dIsolated yield. e[Cp*RhCl ] (2.5 mol %). DMF = N,N-dimethylformamide. 2 2 With the optimal conditions in hand, we explored the substrate scope of acrylic acids (Table 2). Ortho- and meta-substituted substrates could undergo the reaction but the yield went down (3a and 3b). Electron effect did not show obvious effect on this reaction, corresponding products were obtained in 72-78% yield,

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 respectively (3c-3e). Benzyl acrylic acid could also react smoothly with 76% yield (3f). Unactivated alkyl acrylic acids could also adapt to this system to give desired products in moderate yield (3g and 3h). It is worth mentioning that this strategy could also be applied to the thiolation of nonterminal alkenes, although the yields were not so ideal (3i-3k). However, cinnamic acid could not afford products successfully (3l). We assume that it is due to the steric hindrance effect. Table 2. Substrate scope of acrylic

Table 3. Substrate scope of disulfidesa

COOH

RhCl3 3H2O (5 mol %) Cu(OAc)2 (1 equiv) L37 (15 mol %)

RSSR

1a (1 equiv)

4

2 (2 equiv)

OMe

acidsa

S R2

COOH

1 (1 equiv)

Ph

RhCl3 3H2O (5 mol %) Cu(OAc)2 (1 equiv) L37 (15 mol %)

H S

COOH

DMF, 120 oC

S

R1

SR

R

DMF, 120 oC

1

S

R2

S

COOH

Ph

Ph

4b: 77%

4a: 77%

COOH 3

2b (2 equiv)

COOH

S

Cl

Br

S

S

COOH 4c: 73%

F S

S

S

S

S

COOH

COOH

COOH

COOH

MeO 3a: 59%

3b: 56%

COOH

Ph

4d: 75%

3c: 76%

3d: 72%

S

S

S

S

COOH

COOH

COOH

COOH

F 3f: 76%

Ph

COOH

COOH 4f: 62%

4e: 73%

COOH

Ph

4g: 22%b

Cl

3h: 48%

3g: 52%

COOH

4h: 52%b

S

S

S Ph

3e: 78%

Ph

S

Ph

Ph

COOH

4i: 36%

COOH

Se Ph

COOH

4k: 22%b

4j: 31%c

b

aReaction conditions: 1 (0.3 mmol), 2b (0.6 mmol), RhCl 3H O 3 2 (5 mol %), Cu(OAc)2 (1 equiv), L37 (15 mol %), DMF (3 mL), 120 oC, 6 h. b140 oC, 12 h. cRhCl33H2O (10 mol %), L37 (30 mol %), 140 oC, 12 h.

Scheme 3. Scale-up synthesis. S COOH 3i: 39%

S COOH 3j: 35%b

S COOH 3k: 26%

Ph

S COOH

COOH

3l: trace

aReaction conditions: 1 (0.3 mmol), 2b (0.6 mmol), RhCl 3H O 3 2 (5 mol %), Cu(OAc)2 (1 equiv), L37 (15 mol %), DMF (3 mL), 120 oC, 6 h. b140 oC, 12 h.

Next, we investigate the tolerance of disulfides (Table 3). For electron-rich substrates such as methyl- (4b) and methoxy- (4c), weak electron-deficient substrates such as F-, Cl-, Br-substituted disulfides (4d-4f), the reaction underwent smoothly, and the corresponding target products could be obtained in 62-77% yields, respectively. But for strong electron-deficient nitro-substituted disulfides, the reaction could not occur. For alkyl disulfides, the reaction temperature should be increased to 140 oC, and the yield decreased in varying degrees (4g-4j). We also tried to apply diselenides to this system, but the result was not ideal. The target product (4k) was merely obtained in 22% yield. Meanwhile, we attempted to amplify the template reaction to gram level. With reaction time prolonged to 24 h, 4a could be obtained in 60% isolated yield (scheme 3), which implies the possibility of large-scale industrial production.

1a (6 mmol)

PhSSPh

RhCl3 3H2O (5 mol %) Cu(OAc)2 (1 equiv) L37 (15 mol %)

SPh COOH

DMF, 120 oC, 24 h

2a (12 mmol)

4a, 0.9260g isolated yield: 60%

Thioflavone is the core moiety of various natural products with biopharmaceutical activities. In addition, thioflavone and its oxide can be further derived to protect many light-unstable functional groups11. Based on the structure of aforementioned products, Friedel-Crafts acylation was considered to achieve the construction of thioflavone. When 4a was stirred in 0.5 mL trifluoromethanesulfonic acid for 3 hours under 120 oC heating, the derivative thioflavone could be obtained in 88% yield. Meanwhile, 4a was successfully reduced to corresponding alcohol in 50% yield (Scheme 4). Scheme 4. Derivative reaction. S

TfOH (0.5 ml)

LiAlH4 (3 equiv) S

120 oC, 3 h

Ph

S

THF, reflux, 1 h

O Ph der1, 88%

ACS Paragon Plus Environment

COOH 4a

Ph

CH2OH

der2, 50%

ACS Catalysis Scheme 6. XPS & AES spectra of yellow precipitate. Cu2p Scan

LC1 Survey

3.00E+05

1.00E+06 9.00E+05 8.00E+05 7.00E+05

2.00E+05 Counts / s

6.00E+05 5.00E+05 4.00E+05

1.00E+05

3.00E+05 2.00E+05 1.00E+05

0.00E+00 0

0.00E+00 1200

1000

800

600

400

200

960

950

940

930

Binding Energy (eV)

Binding Energy (eV)

a. XPS survey spectrum of yellow precipitate

b. XPS sprectra of yellow precipitate

CuLM2 Scan 6.00E+04

5.00E+04

Counts / s

To explore the plausible mechanism, a series controlled experiments were conducted. Under standard condition, atropic acid 1a was replaced by ethyl 2-phenylacrylate to undergo the reaction (Scheme 5, A). However, corresponding product was not obtained, which reveals the direction of carboxyl group is indispensable. At the same time, the path of Michael additionelimination was ruled out. Next, we investigated Kinetic Isotope Effect (KIE) of the system (Scheme 5, B). Through a competitive experiment between 1a and 1a-d2, the KIE value was calculated as 3.0, which suggested a concerted metalation deprotonation (CMD) process. In addition, it was found that a large amount of precipitate was produced after the reaction. The yellow precipitate is poorly soluble in organic solvents, but it can dissolve in concentrated acids to generate diphenyl disulfide. Therefore, we speculate that the solid may be a low valent copper species bound to disulfide. By X-ray photoelectron spectroscopy (XPS) characterization, from the survey spectrum (Scheme 6, a), we could confirm the existence of copper. From the high resolution XPS spectrum (Scheme 6, b), characteristic satellite peaks of Cu(II) (standard spectra see Scheme 6, c) did not appear, which indicated the precipitate did not contain Cu(II). By Auger Electron Spectroscopy (AES) characterization (Scheme 6, d), two groups of peaks were found, which reveals the existence of Cu(0) and Cu(I). That is to say, the precipitate is the mixture of low valence copper species.

Counts / s

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

Page 4 of 6

4.00E+04

3.00E+04

2.00E+04

1.00E+04 582 580 578 576 574 572 570 568 566 564 562 560 558 Binding Energy (eV)

Scheme 5. Controlled experiments. A. Replaced by ethyl 2-phenylacrylate

c. Standard XPS spectra of different kinds of Cu

RhCl3 3H2O (5 mol %) Cu(OAc)2 (1 equiv) L37 (15 mol %) Ph

COOEt

PhSSPh DMF, 120 oC, 6 h

SPh

Scheme 7. Plausible mechanism. Ph

COOEt

RhCl3 3H2O Ligand

B. Competitive experiment between 1a and 1a-d2

RhCl3 3H2O (5 mol %) Cu(OAc)2 (1 equiv) L37 (15 mol %)

COOH 1a S

D

D

S

DMF, 120 oC, 1.5 h kH/kD = 3.0

Ph

COOH 1a-d2

D/H Ph

A

S

Ph

HOAc

4c and 4c-d

[LRhIIISPh]OAc

2AcOH

C L SPh Ph

RhIII O

COOH 4a



In summary, a RhCl33H2O-Catalyzed thiolation of acrylic acids was reported. With higher efficiency than [Cp*RhCl2]2, this strategy provides a novel way to construct C-S bond through C-H activation. The reaction can be successfully enlarged to gram scale, and the product can be efficiently converted into thioflavone with potential biopharmaceutical activity, which shows practical application in industrial production.

COOH 1a

CuII

COOH

OMe

CONCLUSION

H

LRhIII(OAc)2

PhSCu0, I

Based on above experimental results and related literatures, we proposed a plausible mechanism (Scheme 7). Ligand and Cu(OAc)2 enabled RhCl33H2O to form the active RhIII species A. A reacts with substrate 1a to generate five-membered rhodacycle B through a CMD process. Then, product 4a and intermediate C are obtained by the nucleophilic substitution of B and disulfide. At last, the active species A is regenerated through oxidation of inactive intermediate C by Cu(OAc)2, and the low valence Cu species is formed simultaneously.



Cu(OAc)2

OMe

OMe Ph

d. AES sprectra of yellow precipitate

Ph PhSSPh HOAc

O B

ASSOCIATED CONTENT

Supporting Information

Detailed experimental procedures. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

[email protected]; [email protected]

Notes The authors declare no competing financial interests.

ACS Paragon Plus Environment

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

ACS Catalysis



Rh(III)-Catalyzed Acceptorless Dehydrogenative Coupling of (Hetero)arenes with 2-Carboxyl Allylic Alcohols. Org. Lett. 2018, 20, 740-743.

ACKNOWLEDGMENT

We gratefully acknowledge the National Natural Science Foundation of China (21971174, 21772137, 21672157), the Major Basic Research Project of the Natural Science Foundation of the Jiangsu Higher Education Institutions (No. 16KJA150002), the Ph.D. Programs Foundation of PAPD, the project of scientific and technologic infrastructure of Suzhou (SZS201708), and Soochow University for financial support. We thank Rong Zhang in this group for reproducing the result of 3f, 3i, 4b, 4f.



(5)

She, Z.-J.; Wang, Y.; Wang, D.-P.; Zhao, Y.-S.; Wang, T.-B.; Zheng, X.S.; Yu, Z.-X.; Gao, G.; You, J.-S. Two-Fold C−H/C−H Cross-Coupling Using RhCl3·3H2O as the Catalyst: Direct Fusion of N(Hetero)arylimidazolium Salts and (Hetero)arenes. J. Am. Chem. Soc. 2018, 140, 12566−12573.

(6)

Bernhard, W.; Torsten, S.; Synthesis of Indolo[2,3-a]pyrrolo[3,4c]carbazoles by Oxidative Cyclization of Bisindolylmaleimides with a Rhodium(III) Copper(II) Catalytic System. Synthesis 2005, 1959–1966.

(7)

(a) Vladimir, V. G.; William, J. M.; David L. T. A New, Highly Selective Rh(III) Catalytic System for Carboxylation of Arenes via C±H Activation under Mild Conditions. Adv. Synth. Catal. 2001, 343, 161-165. (b) Wang, P.; Rao, H.-H.; Hua, R.-M.; Li, C.-J. Rhodium-Catalyzed Xanthone Formation from 2-Aryloxybenzaldehydes via Cross-Dehydrogenative Coupling (CDC). Org. Lett. 2012, 14, 902-905. (c) Tiffany, P.; Fedor, R.M.; Maria, R.-M.; Kelvin, E. J.; Natthawat, S.; Trevor, D. T.; Brian, S. N.; Christopher, D. R.; Robert, S. P.; Tomislav, R. Correlating Reactivity and Selectivity to Cyclopentadienyl Ligand Properties in Rh(III)-Catalyzed C−H Activation Reactions: An Experimental and Computational Study. J. Am. Chem. Soc. 2017, 139, 1296−1310. (d) Ran, Y.; Yang, Y.-D.; You, H.S.; You, J.-S. RhCl3‑Catalyzed Oxidative C−H/C−H Cross-Coupling of (Hetero)aromatic Sulfonamides with (Hetero)arenes. ACS Catal. 2018, 8, 1796−1801. (e) Shi, Y.; Zhang, L.-Q.; Lan, J.-B.; Zhang, M.; Zhou, F.-L.; Wei, W.-L.; You, J.-S. Oxidative C-H/C-H Cross-Coupling Reactions between N-Acylanilines and Benzamides Enabled by a Cp*-Free RhCl3/TFA Catalytic System. Angew. Chem. Int. Ed. 2018, 57, 9108 – 9112. (f) Zhao, K.; Du, R.-R.; Wang, B.-Y.; Liu, J.-H. Xia, C.-G. Yang, L. RhCl3·3H2O‑Catalyzed Regioselective C(sp2)−H Alkoxycarbonylation: Efficient Synthesis of Indole- and Pyrrole-2-carboxylic Acid Esters. ACS Catal. 2019, 9, 5545-5551.

(8)

(a) Dvorak, C. A.; Schmitz, W. D.; Poon, D. J.; Pryde, D. C.; Lawson, J. P.; Amos, R. A.; Meyers, A. I. The Synthesis of Streptogramin Antibiotics: (−)-Griseoviridin and Its C-8 Epimer. Angew. Chem., Int. Ed. 2000, 39, 1664−1666. (b) Sader, H. S.; Johnson, D. M.; Jones, R. N. In Vitro Activities of the Novel Cephalosporin LB 11058 against MultidrugResistant Staphylococci and Streptococci. Antimicrob. Agents Chemother. 2004, 48, 53−62. (c) Johannesson, P.; Lindeberg, G.; Johansson, A.; Nikiforovich, G. V.; Gogoll, A.; Synnergren, B.; Le Grèves, M.; Nyberg, F.; Karlén, A.; Hallberg, A. Vinyl Sulfide Cyclized Analogues of Angiotensin II with High Affinity and Full Agonist Activity at the AT1 Receptor. J. Med. Chem. 2002, 45, 1767−1777.

(9)

(a) Lee, C.-F.; Liu, Y.-C.; Badsara, S. S. Transition-Metal Catalyzed C−S Bond Coupling Reaction. Chem. - Asian J. 2014, 9, 706−722. (b) Beletskaya, I. P.; Ananikov, V. P. Transition-Metal Catalyzed C−S, C−Se, and C−Te Bond Formation via Cross- Coupling and Atom-Economic Addition Reactions. Chem. Rev. 2011, 111, 1596−1636. (c) Dondoni, A.; Marra, A. Metal-Catalyzed and Metal-Free Alkyne Hydrothiolation: Synthetic Aspects and Application Trends. Eur. J. Org. Chem. 2014, 2014, 3955−3969. (d) Castarlenas, R.; Di Giuseppe, A.; Pérez-Torrente, J. J.; Oro, L. A. The Emergence of Transition-Metal-Mediated Hydrothiolation of Unsaturated Carbon−Carbon Bonds: A Mechanistic Outlook. Angew. Chem., Int. Ed. 2013, 52, 211−222. (e) Strydom, I.; Guisado-Barrios, G.; Fernández, I.; Liles, D. C.; Peris, E.; Bezuidenhout, D. I. A Hemilabile and Cooperative N-Donor Functionalized 1,2,3-Triazol-5- ylidene Ligand for Alkyne Hydrothiolation Reactions. Chem. - Eur. J. 2017, 23, 1393−1401. (f) Degtyareva, E. S.; Burykina, J. V.; Fakhrutdinov, A. N.; Gordeev, E. G.; Khrustalev, V. N.; Ananikov, V. P. Pd−NHC Catalytic System for the Efficient Atom-Economic Synthesis of Vinyl Sulfides from Tertiary, Secondary, or Primary Thiols. ACS Catal. 2015, 5, 7208−7213. (g) Di Giuseppe, A.; Castarlenas, R.; Pérez-Torrente, J. J.; Crucianelli, M.; Polo, V.; Sancho, R.; Lahoz, F. J.; Oro, L. A. LigandControlled Regioselectivity in the Hydrothiolation of Alkynes by Rhodium N-Heterocyclic Carbene Catalysts. J. Am. Chem. Soc. 2012, 134, 8171−8183. (h) Weiss, C. J.; Marks, T. J. Organozirconium Complexes as Catalysts for Markovnikov-Selective Intermolecular Hydrothiolation of Terminal Alkynes: Scope and Mechanism. J. Am. Chem. Soc. 2010, 132, 10533−10546. (i) Cao, F.; Fraser, L. R.; Love, J. A. Rhodium Catalyzed Alkyne Hydrothiolation with Aromatic and Aliphatic Thiols. J. Am. Chem. Soc. 2005, 127, 17614−17615. (j) Pan, X.-Q.; Zou, J.-P.; Yi, W.-B.; Zhang, W. Recent Advances in Sulfur- and Phosphorous-Centered Radical Reactions for the Formation of S−C and P−C Bonds. Tetrahedron 2015, 71, 7481− 7529. (k) Kao, H.-L.; Lee, C.-F. Efficient CopperCatalyzed SVinylation of Thiols with Vinyl Halides. Org. Lett. 2011, 13, 5204−5207. (l) Kabir, M. S.; Lorenz, M.; Van Linn, M. L.; Namjoshi, O. A.; Ara, S.; Cook, J. M. A Very Active Cu-Catalytic System for the Synthesis of Aryl, Heteroaryl, and Vinyl Sulfides. J. Org. Chem. 2010, 75, 3626−3643. (m) Velasco, N.; Virumbrales, C.; Sanz, R.; Suárez Pantiga, S.; Fernández-Rodríguez, M. A. General Synthesis of Alkenyl Sulfides by Palladium-Catalyzed Thioetherification of Alkenyl Halides and Tosylates. Org. Lett. 2018, 20, 2848−2852. (n) Rao, W.-H.; Shi, B.-F. Copper(II)Catalyzed Direct Sulfonylation of C(sp2)−H Bonds with Sodium Sulfinates. Org. Lett. 2015, 17, 2784−2787. (o) Chen, F.-J.; Liao, G.; Li,

REFERENCES (1)

(a) Liu, T.; Zhou, W.; Wu, J. Palladium-Catalyzed Direct C–H Functionalization of Indoles with the Insertion of Sulfur Dioxide: Synthesis of 2-Sulfonated Indoles. Org. Lett., 2017, 19, 6638−6641. (b) Parthasarathy, G.; Julian, K.; Lutz, A. Expedient C–H Chalcogenation of Indolines and Indoles by Positional-Selective Copper Catalysis. ACS Catal., 2017, 7, 1030–1034. (c) Lin, C.; Yu, W.-L.; Yao, J.-Z.; Wang, B.J.; Liu, Z.-X.; Zhang, Y.-H. Nickel-Catalyzed Direct Thioetherification of β-C(sp3)–H Bonds of Aliphatic Amides. Org. Lett. 2015, 17, 1340−1343. (d) Yan, S.-Y.; Liu, Y.-J.; Liu, B.; Liu, Y.-H.; Shi, B.-F. Nickel-catalyzed thiolation of unactivated aryl C–H bonds: efficient access to diverse aryl sulfides. Chem. Commun., 2015, 51, 4069-4072.

(2)

(a) Korkit, K.; Nikolaos, K.; Torben, R.; Svenja, W.; A. Claudia, S.; Lutz, A. Sequential meta-/ortho-C–H Functionalizations by One-Pot Ruthenium(II/III) Catalysis. ACS Catal. 2018, 8, 886−892. (b) Yuan, C.C.; Zhu, L.; Zeng, R.-S.; Lan, Yu. Zhao, Y.-S. Ruthenium(II)‐Catalyzed C−H Difluoromethylation of Ketoximes: Tuning the Regioselectivity from the meta to the para Position. Angew. Chem. Int. Ed. 2018, 57, 1277 –1281. (c) Hojoon, P.; Nikita, C.; Shen, P.-X.; Yu, J.-Q. Ligand-Enabled, Palladium-Catalyzed β-C(sp3)–H Arylation of Weinreb Amides. ACS Catal. 2018, 8, 9292−9297. (d) Desta, D. B.; Cody, R. P.; Fereshte, G.; Stefan, A. H.; Joseph N. C.; Maxime, A. S.; Thomas, L. Ketones as directing groups in photocatalytic sp3 C–H fluorination. Chem. Sci. 2017, 8, 6918-6923.

(3)

(4)

(a) Hendrich, A. C.; Quynh-Nhu, P.; Olafs, D. Two Methods for Direct ortho-Arylation of Benzoic Acids. J. Am. Chem. Soc. 2007, 129, 98799884. (b) Josep, C.; Marika, R.; Igor, L. Carboxylic Acids as Traceless Directing Groups for Formal meta ‐ Selective Direct Arylation. Angew. Chem. Int. Ed. 2011, 50, 9429-9432. (c) Peter, S. T.-B.; Giorgio, V.; Devin, D.; Paul, R.; Su, S.; Yu, J.-Q. Ligand-Accelerated ortho-C–H Alkylation of Arylcarboxylic Acids using Alkyl Boron Reagents. J. Am. Chem. Soc. 2013, 135, 17508-17513. (d) Cheng, G.-L.; Li, T.-J.; Yu, J.-Q. Practical Pd(II)-Catalyzed C–H Alkylation with Epoxides: One-Step Syntheses of 3,4-Dihydroisocoumarins. J. Am. Chem. Soc. 2015, 137, 10950-10953. (e) Gong, H.; Zeng, H.-Y.; Zhou, F.; Li, C.‐J. Rhodium(I) ‐Catalyzed Regiospecific Dimerization of Aromatic Acids: Two Direct C-H Bond Activations in Water. Angew. Chem. Int. Ed. 2015, 54, 5718 5721. (f) Zhang, Y.-F.; Zhao, H.-Q.; Zhang, M.; Su, W.-P. Carboxylic Acids as Traceless Directing Groups for the Rhodium(III) ‐ Catalyzed Decarboxylative C-H Arylation of Thiophenes. Angew. Chem. Int. Ed. 2015, 54, 3817 –3821. (g) Zhang, G.-D.; Hu, Z.-Y.; Florian, B.; Ou, Y.; Nico, P.; Lukas. J. G. Rhodium‐Catalyzed Annelation of Benzoic Acids with α,β‐Unsaturated Ketones with Cleavage of C−H, CO−OH, and C−C Bonds. Angew. Chem. Int. Ed. 2019, 58, 6435 –6439. (h) Suman, D.; Deepan, C.; Anup, M.; Francis, A. S. C.; Mahiuddin, B. Ruthenium(II) Catalysis/Noncovalent Interaction Synergy for Cross-Dehydrogenative Coupling of Arene Carboxylic Acids. ACS Catal. 2018, 8, 10173−10179. (i) Anup, M.; Suman, D.; Harekrishna, S.; Gowri, S. G.; Mahiuddin, B. Ruthenium(II)-Catalyzed ortho-C–H Chalcogenation of Benzoic Acids via Weak O-Coordination: Synthesis of Chalcogenoxanthones. Org. Lett. 2017, 19, 2430−2433. (j) N. Y. Phani, K.; Torben, R.; Santhivardhana, R. Y.; Alexander, B.; Eric, C.; Lutz, A. Mild Decarboxylative C−H Alkylation: Computational Insights for Solvent‐Robust Ruthenium(II) Domino Manifold. Chem. Eur. J. 2017, 23, 17449-17453. (k) Anup MandalHarekrishna SahooSuman DanaMahiuddin Baidya. Ruthenium(II)-Catalyzed Hydroarylation of Maleimides Using Carboxylic Acids as a Traceless Directing Group. Org. Lett. 2017, 19, 4138−4141. (l) Guodong Zhang Fan Jia Prof. Dr. Lukas J. Gooßen. Regioselective C−H Alkylation via Carboxylate‐Directed Hydroarylation in Water. Chem. Eur. J. 2018, 24, 4537 – 4541. (m) Huang, L.-B.; Dagmar, H.; Lukas, J. G. Iridium‐Catalyzed ortho‐Arylation of Benzoic Acids with Arenediazonium Salts. Angew. Chem. Int. Ed. 2015, 54, 12607 –12611. (a) Tian, M.-M., Bai, D.-C., Zheng, G.-F., Chang, J.-B., Li, X.-W. Rh(III)-Catalyzed Asymmetric Synthesis of Axially Chiral Biindolyls by Merging C–H Activation and Nucleophilic Cyclization. J. Am. Chem. Soc. 2019, 141, 9527-9532. (b) Tian, M.-M., Liu, B.-X., Sun, J.-Q., Li, X.-W. Rh(III)-Catalyzed C-C Coupling of Diverse Arenes and 4-Acyl-1sulfonyltriazoles via C-H Activation. Org. Lett. 2018, 20, 4946-4949. (c) Xia, J.-T., Huang, Z.-P., Zhou, X.-K., Yang, X.-F., Wang F., Li, X.-W.

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

X.; Wu, J.; Shi, B.- F. Cu(II)-Mediated C−S/N−S Bond Formation via C−H Activation: Access to Benzoisothiazolones Using Elemental Sulfur. Org. Lett. 2014, 16, 5644−5647. (p) Liu, S.-L.; Li, X.-H.; Shi, T.-H.; Yang, G.- C.; Wang, H.-L.; Gong, J.-F.; Song, M.-P. Copper-Promoted Thiolation of C(sp2)−H Bonds Using a 2-Amino Alkylbenzimidazole Directing Group. Eur. J. Org. Chem. 2017, 2017, 2280−2289. (q) Lin, C.; Li, D.-Y.; Wang, B.-J.; Yao, J.-Z.; Zhang, Y.-H. Direct ortho- Thiolation of Arenes and Alkenes by Nickel Catalysis. Org. Lett. 2015, 17, 1328−1331. (r) Yang, L.; Wen, Q.; Xiao, F.-H.; Deng, G.-J. SilverMediated Oxidative Vinylic C−H Bond Sulfenylation of Enamides with Disulfides. Org. Biomol. Chem. 2014, 12, 9519−9523. (10) Liu, C.; Fang, Y.; Wang, S.-Y.; Ji, S.-J. Highly Regioselective RhIIICatalyzed Thiolation of N‑Tosyl Acrylamides: General Access to (Z)‑β- Alkenyl Sulfide. Org. Lett. 2018, 20, 6112-6116. (11) Subramani, S.; Govindasamy, S.; Copper-Catalyzed One-Pot Synthesis of 2‑Arylthiochromenones: An in Situ Recycle of Waste Byproduct as Useful Reagent. Org. Lett. 2019, 21, 75−79.

ACS Paragon Plus Environment

Page 6 of 6