Fluorination of Malonic Acid Derivatives - American Chemical Society

Recently, the development of novel transformations with carboxylic acids as ... decarboxylative reactions and exploiting malonic acid as a .... 2/3 = ...
0 downloads 0 Views 492KB Size
Subscriber access provided by Drexel University Libraries

Communication

Ag-Catalyzed Chemoselective Decarboxylative Monoand gem-Di-Fluorination of Malonic Acid Derivatives Zhen Wang, Cong-Ying Guo, Cheng Yang, and Jian-Ping Chen J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 25 Mar 2019 Downloaded from http://pubs.acs.org on March 25, 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

Journal of the American Chemical Society

Ag-Catalyzed Chemoselective Decarboxylative Mono- and gem-DiFluorination of Malonic Acid Derivatives Zhen Wang, Cong-Ying Guo, Cheng Yang, Jian-Ping Chen* Institute of Advanced Synthesis, School of Chemistry and Molecular Engineering, Nanjing Tech University, Nanjing 211816, China Supporting Information Placeholder ABSTRACT: Malonic acid derivatives have been successfully applied in a Ag-catalyzed decarboxylative fluorination reaction, providing an unprecedented route to either gem-difluoroalkanes or α-fluorocarboxylic acids by the judicious selection of base and solvent. This reaction features the use of readily available starting materials, tunable chemoselectivity and good functional group compatibility as well as gram-scale synthetic capability. The advantage of using malonic acid derivatives in this radical decarboxylative functionalization is further highlighted by the facile transformations of the α-fluorocarboxylic acid to valuable fluorine-containing compounds. Preliminary mechanistic studies suggest that an α-carboxylic acid radical is involved in this reaction.

dedicated toward the incorporation of gem-difluoromethylene group into small molecules. However, achievements have mainly a) Radical decarboxylative functionalization (RDF) of monocarboxylic acid - CO2

R1 COOH monocarboxylic acids

R2

2

R = F, Cl, Br, N3, Bpin, OH, CF3, alkynyl, allyl, Ar, alkyl etc.

alkyl radical

b) Double Hunsdiecker and oxidative decarboxylation reactions of malonic acid derivatives Br

Br

HgO (stoichiometric) HOOC

R1 R2

COOH Pb(OAc)4 (stoichiometric)

R1 R2

Br2

O R1

or Electrolysis

R2

c) Photoredox hydrodecarboxylation of malonic acid derivatives Mes-Acr-Ph (cat.) PhS-SPh (cat.)

COOH

HOOC

R1 R2

Carboxylic acids are appealing starting materials because of their stability, broad availability and low cost. Recently, the development of novel transformations with carboxylic acids as starting materials has attracted much attention.1 Among them, radical decarboxylative functionalization (RDF) has emerged as a powerful strategy to efficiently construct carbon-carbon and carbon-heteroatom bonds (Scheme 1a).2 The strength of this methodology lies in its ability to generate alkyl radicals through a radical decarboxylation and capture the carbon radicals under mild conditions. Nevertheless, the carboxylic acids used in RDFs have been predominantly restricted to monocarboxylic acids. Malonic acid derivatives are classical building blocks, which can be readily accessible via alkylations of malonic ester and subsequent hydrolysis;3 however, the corresponding RDFs of malonic acids have rarely been reported. Early studies demonstrated that disubstituted malonic acid derivatives could undergo double Hunsdiecker or oxidative decarboxylation reactions in the presence of excess toxic metals or under electrochemical conditions (Scheme 1b).4,5 A notable example was reported by the Nicewicz group, who applied gem-dicarboxylic acids in a photoredox hydrodecarboxylation reaction, but the increased oxidation potential caused by the second carboxylic acid moiety through hydrogen bonding makes them challenging substrates, and a strong base was necessary for this transformation (Scheme 1c).6 Being aware that the utilization of malonic acid derivatives in RDF remains largely elusive, especially in a catalytic manner, we are particularly interested in exploring their novel radical decarboxylative reactions and exploiting malonic acid as a promising synthon in organic synthesis. gem-Difluoromethylene group is a valuable fluorinated motif that is widely present in pharmaceuticals and biologically active compounds (Scheme 1e).7,8 As a result, great efforts have been

R1

R1

H

H

R1 R2

KOt-Bu 450 nm LEDs

d) This work: F HOOC

COOH

AgNO3 (cat.)

HOOC

F

R1 R2

R1 R2

R1 R2

Selectfluor -carboxylic acid radical base, solvent key intermediate malonic acid derivatives

F

COOH

R1 R2

e) Representative drugs containing a gem-difluoromethylene group F F F

F

NH2

O

N HO

HN

O

Me N

HO

O O

N N

Maraviroc HIV-1 therapeutic agent

N

OH

O

Tafluprost ocular hypertension drug

N

O

F

OH F Gemcitabine chemotherapy drug

Scheme 1. Radical Decarboxylative Functionalization and Representative Drugs Containing a gem-Difluoromethylene Group been made in the formation of aryl-CF2R bonds,8 but methods for the synthesis of gem-difluoroalkanes (Csp3-CF2-Csp3) are less developed. Conventional approaches for preparing difluoroalkylated alkanes include deoxyfluorination of ketones and hydrofluorination of alkynes.9,10 Alternatively, gemdifluoroalkanes can be obtained via a light-induced difluoroalkylation or a nickel-catalyzed tandem difluoroalkylationarylation of activated alkenes.11,12 Despite recent advances, the limitations of the aforementioned methods such as harsh reaction conditions, poor functional group tolerance, poor availability of starting materials and limited scope have impeded their applications in organic synthesis to a certain extent. Given their

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

significance in medicinal chemistry, the exploration of a general and practical approach for the efficient synthesis of gemdifluoroalkanes is highly desirable. Although the application of monocarboxylic acids in radical decarboxylative fluorinations has been well-documented,13 to the best of our knowledge, the generation of α-carboxylic acid radicals through radical decarboxylation of malonic acid derivatives and subsequent trapping of the resulting radicals with a fluorine reagent is unprecedented. Herein, we describe the successful application of malonic acid derivatives in a silver-catalyzed decarboxylative fluorination. The chemoselectivity of this reaction was tunable, providing valuable access to either gem-difluoroalkanes or αfluorocarboxylic acids (Scheme 1d). Inspired by Li’s elegant work on the Ag-catalyzed decarboxylative fluorination of monocarboxylic acids,13a we commenced our study with the reaction of 2-(4-chlorophenethyl)2-ethylmalonic acid (1a) and Selectfluor in the presence of AgNO3 (20 mol%) in aqueous acetone or acetonitrile at 55 °C. Delightfully, gem-difluorinated product 2a and monofluorinated product 3a were produced after 12 h, albeit in low yields and poor ratios of 2a/3a (Table 1, entries 1 and 2). Next, we examined the effect of base on this reaction. Generally, the addition of base could promote the formation of 2a (entries 4-9), except for KOH (entry 3). When PhCO2K or PhCO2Na were employed, 2a was produced as the major product in a 31% yield with the ratio of 2a/3a being 84/16 (entries 6 and 7). The investigation of the solvents with PhCO2Na as the base revealed that the solvent played an important role in controlling the chemoselectivity (entry 7 vs entries 10-13). Interestingly, when biphasic systems (CH2Cl2/H2O or CPME/H2O) were employed, the monofluorinated product 3a was generated exclusively (entries 10 and 13).

CH3CN/H2O/n-Hex (1/1/3) in the presence of AgNO3 (30 mol%) by using PhCO2Na (3.0 equiv) as the base (entry 15). When the base was changed to K2HPO4, the reaction in CPME/H2O (1/1) at room temperature exclusively provided α-fluorocarboxylic acid 3a in an isolated yield of 54% (entry 14). This tunable chemoselectivity realized by simply varying the reaction conditions offers an expedient way of selectively introducing fluorine atoms into small molecules. Control experiments showed that the solvent had a substantial impact on the chemoselectivity of this reaction (entries 15 and 16) and the silver catalyst was essential for these transformations (entries 17 and 18).

Table 2. Scope of Decarboxylative gem-Difluorinationa-c HOOC COOH

Cl

AgNO3 (20 mol%) Selectfluor (4.0 equiv.)

F

COOH

Base, Solvent 55 °C

F

1a

Cl

1

COOH

F F

Cl

Cl

Ph

2a yield (%)b

3a yield (%)b

1

-

Acetone/H2O (1/1)

48/52

10

11

-

CH3CN/H2O (1/1)

21/79

4

15

Ph 2ff , 72%; 2/3 > 95/5

HOOC

F

32/68

6

13

Acetone/H2O (1/1)

76/24

16

5

1u

5

K2HPO4 (2.0)

Acetone/H2O (1/1)

71/29

17

6

6

PhCO2K (2.0)

Acetone/H2O (1/1)

84/16

31

6

7

PhCO2Na (2.0)

Acetone/H2O (1/1)

84/16

31

6

8

K2CO3 (2.0)

Acetone/H2O (1/1)

59/41

10

7

9

Et3N (2.0)

Acetone/H2O (1/1)

63/37

17

10

10

PhCO2Na (2.0)

DCM/H2O (1/1)

0/100

0

18

11

PhCO2Na (2.0)

CH3CN/H2O (1/1)

21/79

4

15

12

PhCO2Na (2.0)

n-Hex/H2O (1/1)

27/73

6

16

4/96

1

25

2/98

1

56 (54)

15d

PhCO2Na (3.0)

CH3CN/H2O/n-Hex (1/1/3)

100/0

69 (60)

0

16d

PhCO2Na (3.0)

CPME/H2O (1/1)

7/93

1

14

17 e

K2HPO4 (4.0)

CPME/H2O (1/1)

0

0

0

18e

PhCO2Na (3.0)

CH3CN/H2O/n-Hex (1/1/3)

0

0

0

aReaction

conditions: 1a (0.1 mmol), AgNO3 (0.02 mmol), Selectfluor (0.4 mmol), base, solvent (2 mL), 55 °C, 12 h under N2. b19F NMR yield with 4-fluorobenzoic acid as the internal standard, isolated yield in parentheses. cSelectfluor (0.2 mmol), rt. d1a (0.1 mmol), AgNO3 (0.03 mmol), Selectfluor (0.6 mmol), base, solvent (1.5 mL). eWithout AgNO3. CPME = cyclopentyl methyl ether; n-Hex = n-hexane.

Based on the above findings, an extensive study of the reaction parameters was conducted to further improve the yields. Ultimately, the gem-difluorinated compound 2a could be obtained predominantly in an isolated yield of 60% by the reaction of 1a (0.1 mmol) with Selectfluor (0.6 mmol) in a mixed solvent of

F BzN Cl

2od , 79%; 2/3 > 95/5

Br

2pg,i , 81%; 2/3 = 90/10 O

F F Ph

N

F

O 2ti,j , 21%; 2/3 = 54/46

H

F

OH

2rf , 74%; 2/3 > 95/5

HOOC COOH

Acetone/H2O (1/1)

CPME/H2O (1/1)

F

2sf , 35%; 2/3 > 95/5

unsuccessful substrates

KOH (2.0)

CPME/H2O (1/1)

F

O S O

O

2qf , 58%; 2/3 = 95/5

2lf , 69%; 2/3 > 95/5 F

F

F F

K3PO4 (2.0)

K2HPO4 (4.0)

2kf , 55%; 2/3 > 95/5

2nf , 85%; 2/3 > 95/5

2m, 62%; 2/3 = 90/10

O

O

2jd , 77%; 2/3 > 95/5

F F

F F

F O

F

O 2ih, 66%; 2/3 > 95/5

2h, 80%; 2/3 > 95/5 F

F

N

4

PhCO2Na (2.0)

2gf , 55%; 2/3 = 90/10

O

F F

3

13

F F F

Ph

2e, 60%; 2/3 > 95/5

5

14c

, 77%; 2/3 > 95/5

F

F F

F F Ph

2dd,e

2c, 51%; 2/3 > 95/5

, 69%; 2/3 > 95/5

O

ratio (2/ 3)

R2 3

F F

2bd

2a, 60%; 2/3 > 95/5

3a

solvent (V/ V)

2

Ph Cl

Cl

base (equiv)

R1

F F

F

2a

F COOH

+

R2

R1

F F

F F

+

F F

AgNO3 (30 mol%) PhCO2Na, Selectfluor CH3CN/H2O/n-hexane

CN

entry

2

R2

R1

Table 1. Optimization of Reaction Conditionsa COOH

Page 2 of 6

COOH COOH

Ph Et 1v

AcHN COOH Ph COOH 1w

Br

EtO

COOH COOH

1x

aReaction

condition: 1 (0.2 mmol), AgNO3 (0.06 mmol), Selectfluor (1.2 mmol), PhCO2Na (0.5 mmol), CH3CN (0.6 mL), H2O (0.6 mL), n-hexane (1.8 mL), 55 °C, under N2. bIsolated yield. cThe 2/3 ratio was determined by 19F NMR. dK2HPO4 (0.4 mmol) instead of PhCO2Na. e19F NMR yield with 4-fluorobenzoic acid as the internal standard. fPhCO2Na (0.7 mmol), CH3CN (1.0 mL), H2O (1.0 mL) and n-hexane (3.0 mL) as solvent, rt. gAgNO3 (0.04 mmol), K2HPO4 (0.6 mmol). hK2CO3 (0.6 mmol), DCE (2.0 mL) and H2O (2.0 mL). iCH3CN (2.0 mL) and H2O (2.0 mL) were used as the solvent. jSelectfluor (1.6 mmol), K2HPO4 (0.8 mmol).

With the optimal conditions established, the substrate scope of decarboxylative gem-difluorination was examined (Table 2). Generally, for disubstituted malonic acids, the reaction proceeded smoothly, affording gem-difluorinated compounds 2 in moderate to good yields with excellent chemoselectivities. The reactions of simple dialkyl-substituted gem-dicarboxylic acids exclusively produced gem-difluoroalkanes 2a-2g in 51-77% yields. Increasing the steric hindrance at the α-position of the malonic acids had no obvious deleterious effects on the reactivity or selectivity (2a-2c). Aryl- and alkyl-substituted malonic acids were tolerated in this transformation, and two fluoro substituents could be effectively installed at the benzylic position (2h and 2i). In addition, cyclic malonic acids were also suitable substrates, furnishing 2o and 2p in good yields. Notably, malonic acids bearing functional groups, including phthalimide (2j), ester (2k), ether (2q), sulfone (2n), benzoic acid (2i), cyanide (2m), amide (2p) and ketone (2l and 2r) were well tolerated in this system, providing gem-difluoro compounds 2 in 55-85% yields with excellent chemoselectivities.

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

Journal of the American Chemical Society Secondary alcohol partially survived to afford product 2s in 35% yield, along with 33% yield of corresponding ketone product.14 Alkynyl and alkenyl groups (1u and 1v) were not compatible with such oxidative conditions. Only oxidative decarboxylation products were observed for the α-heteroatom substituted malonic acids (1w and 1x).5 Furthermore, the decarboxylative gemdifluorination of monosubstituted malonic acid was also examined; unfortunately, product 2t was obtained in a low yield with poor chemoselectivity.15 Subsequently, we turned our attention to extending the substrate scope of decarboxylative monofluorination. As illustrated in Table 3, the substrates that can participate in the decarboxylative gemdifluorination are also suitable for this monofluorination reaction, predominately producing α-fluorocarboxylic acids in 30-92% yields;16 the exception is aryl-substituted malonic acids, which only produced gem-difluorides (2h and 2i). We attribute the different chemoselectivity of aryl-substituted malonic acids in this system to the reaction rate of the second decarboxylative fluorination being accelerated due to the stabilization of the corresponding benzylic radical. In contrast to the decarboxylative gem-difluorination, introduction of sterically encumbered groups at the α-position of the malonic acids compromised the reaction efficiency (3c vs 3b and 3e vs 3d), but 3c and 3e could still be obtained in synthetically useful yields. Generally, methods for accessing tertiary αfluorocarboxylic acid derivatives rely on electrophilic fluorination of the corresponding ester enolates and subsequent hydrolysis;17 however, the harsh reaction conditions of these protocols showed limited functional group compatibility. In comparison, our mild reaction conditions for α-fluorocarboxylic acids accommodate a variety of functional groups, such as aryl halide (3a-3c, 3o and 3r), phthalimide (3j), ester (3k), ether (3q), sulfone (3n), cyanide (3m), amide (3p), ketone (3l and 3r), alcohol (3s) and alkyne (3u). Substrates 1v-1x are not viable for this transformation. Pleasingly, the reaction of monosubstituted malonic acid delivered product 3t in 41% yield with a 2/3 ratio of 21/79.

To demonstrate the practical utility of this unprecedented decarboxylative fluorination, gram-scale reactions were carried out (Scheme 2a). From the same starting material 1p, gemdifluoroalkane 2p and α-fluorocarboxylic acid 3p were obtained in good yields and excellent selectivities. Notably, α-fluorocarboxylic acids 3 are of particular versatility, as the remaining carboxylic acid group provides spacious room for further elaboration. The utility of the α-fluorocarboxylic acid has been showcased by the facile conversions of compound 3p to a series of valuable fluorinecontaining building blocks, such as gem-chloro fluoroalkane 4,2a,18 tertiary propargylic fluoride 52b,19 and γ-fluoride Michael adduct 62g via radical decarboxylative functionalization (Scheme 2b). These results clearly demonstrate the potential applications of this protocol for the late-stage diversification of fluorinated bioactive compounds. a) Gram-scale synthesis of 2p and 3p F F

HOOC COOH Table 2 reaction condition for 2p 4h 78% yield; 2p/3p > 95/5

N Bz 2p

5h 95% yield; 2p/3p < 5/95

N Bz 1p 1.0 gram

N Bz 3p

b) Transformations of compound 3p F Cl Ag(Phen)2OTf (10 mol%) t-BuOCl, CH3CN, 50 °C 83%

N Bz 4 TIPS

F COOH F

AgNO3 (30 mol%) K2S2O8, TIPS-EBX

N Bz 3p

CH3CN/H2O, 55 °C 58%

N Bz 5

Table 3. Scope of Decarboxylative Monofluorinationa-c HOOC COOH R2

R1

F F

AgNO3 (20 mol%) K2HPO4, Selectfluor CPME/H2O

1 F COOH

+

R2

R1 2

F COOH

F COOH F COOH Ph

Cl

Cl

F COOH Ph

Ph

3cd,e, 30%; 2/3 = 22/78 O

F COOH

N

Ph

3ed , 47%; 2/3 = 18/82 O

HOOC

HOOC F

F COOH

F

O

3p, 87%; 2/3 < 5/95

F COOH

N

F COOH H

O 3te,f , 41%; 2/3 = 21/79 aReaction

5 3u, 58%; 2/3 < 5/95

Selectfluor

3k, 63%; 2/3 < 5/95 COOH

Cl

2 BF4 8

HOOC

HOOC

path b

COOAg(I)

F COOH Ph OH 3s, 78%; 2/3 = 9/91; dr = 1/1

F

11

-CO2

N

R1 R2

HOOC

COOAg(II)

R1 R2 10

path a

Selectfluor 8

Cl

N 9

F

BF4

base

F

HOOC

R1 R2 2

Ag(I)

R1 R2 1

9 -CO2

Selectfluor COOH

base

1

R R2

F

F

8

COOH

8

COOH

R1 R2 3

8

3o, 58%; 2/3 = 17/83; dr = 3.5/1 Br

6

Cl

R1 R2 7

F Ph

N

O O

O 3r, 78%; 2/3 < 5/95

3q, 80%; 2/3 < 5/95

O

COOH

O S O

F COOH

COOH BzN

N

COOH

3n, 80%; 2/3 < 5/95

3m, 74%; 2/3 < 5/95

F

F

F

CN 3l, 78%; 2/3 < 5/95

3d, 76%; 2/3 < 5/95

O 3j, 92%; 2/3 < 5/95

3f, 60%; 2/3 < 5/95

F COOH

F

N Bz

Scheme 2. Synthetic Applications

Cl 3b, 61%; 2/3 < 5/95

3a, 54%; 2/3 < 5/95

CO2Et

Diethyl ethylidenemalonate K2HPO4, blue LED DMF, rt 78%

R2 3

CO2Et

F

Ir[df(CF3)ppy]2(dtbbpy)PF6 (3 mol%)

F COOH R1

F COOH

Table 3 standard condition

F

path b

COOAg(II)

R1 R2 13

COOAg(I)

R1 R2 12

path a

Selectfluor

8

14

Et 3v, 14%; 2/3 < 5/95

Scheme 3. Proposed Mechanism

condition: 1 (0.2 mmol), AgNO3 (0.04 mmol), Selectfluor (0.4 mmol), K2HPO4 (0.8 mmol), CPME (2.0 mL), H2O (2.0 mL), rt under N2. bIsolated yield for the corresponding methyl ester. cThe ratio of 2/3 was determined by 19F NMR. dAgNO3 (0.06 mmol), Selectfluor (0.8 mmol), PhCO2Na (0.7 mmol), CPME (9.0 mL), CH3CN (1.0 mL), n-hexane (3.0 mL), H2O (1.0 mL), rt. e55 °C. fAgNO3 (0.06 mmol), Selectfluor (1.2 mmol), CH3CN (0.8 mL), H2O (3.2 mL).

Based upon our experimental results and literature precedents,13a,20 a tentative reaction mechanism was proposed (Scheme 3). Initially, facilitated by a weak base, malonic acid 1

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

reacts with a catalytic amount of Ag(I) to form malonic acid monosilver(I) salt 7,21 which is readily oxidized to Ag(II) 10 by Selectfluor with concomitant generation of 8; alternatively, the newly formed Selectfluor radical cation 8 is also capable of oxidizing Ag(I) to Ag(II).20 Then, the carboxylate anion loses an electron to Ag(II), and subsequent CO2 extrusion delivers an αcarboxylic acid radical 11 and regenerates Ag(I). The resulting radical 11 abstracts a fluorine atom from Selectfluor to give product 3. In the decarboxylative monofluorination, the reaction ceases at this stage and α-fluorocarboxylic acid 3 is the final product. In the decarboxylative gem-difluorination, compound 3 would undergo a second decarboxylative fluorination (3 → 12 → 13 → 14 → 2) to eventually produce gem-difluorinated product 2. Although the mechanism of the silver-catalyzed decarboxylative fluorination of monocarboxylic acids has been well studied,20 the radical decarboxylation of malonic acid derivatives to generate αcarboxylic acid radicals in a Ag(I)/Selectfluor system is unknown. To support the above mechanistic hypothesis, radical clock 15 was subjected to our standard decarboxylative monofluorination conditions, and the ring-opening product 16 was obtained in 54% yield (Scheme 4, eq 1), which suggested the involvement of αcarboxylic acid radical 11 in this transformation. Moreover, the successful conversion of 3p to gem-difluoroalkane 2p (eq 2) and the observation of monofluorinated product 3 as a side product in the decarboxylative gem-difluorination (Table 2) implied that αfluorinated carboxylic acid 3 was the intermediate in the formation of gem-difluoroalkanes.22 Additionally, 3p could not be converted to 2p by decarboxylative monofluorination (eq 3). Although the origin of this intriguing chemoselectivity is not yet clear,23 a detailed mechanistic study of this reaction is underway in our laboratory. Ph HOOC 15

Table 3 F standard condition Me Ph COOH

F COOH

16 54% yield Table 2 standard condition

N Bz 3p

Me COOH F 17 not observed

(1)

F F (2) N Bz 2p 67% yield

N Bz 3p

F COOH

COOH + Ph

Table 3 standard condition

F F

N Bz 2p 19 < 5% FNMR yield

F COOH (3)

+ N Bz 3p recovered in 90% yield

Scheme 4. Mechanistic Experiments In summary, we have successfully developed a silver-catalyzed decarboxylative fluorination reaction of malonic acid derivatives. Tunable chemoselectivity was realized by the judicious selection of base and solvent, which offers an expedient way to access gemdifluoroalkanes and α-fluorocarboxylic acids. The advantage of using malonic acid derivatives in RDF was further highlighted by the facile transformations of the α-fluorocarboxylic acid to some valuable fluorine-containing compounds. More importantly, this work constitutes the first example of the generation of α-carboxylic acid radicals through the decarboxylation of malonic acid derivatives in a Ag(I)/Selectfluor system. Given the operational simplicity and good functional group tolerance as well as the use of readily available starting materials, we believe that the present protocol will provide a powerful tool for the preparation of valuable fluorinated compounds.

ASSOCIATED CONTENT Supporting Information Detailed experimental procedures, spectral data, and analytical data are available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author Jian-Ping Chen: [email protected]

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This paper is dedicated to Professor Li-Xin Dai on the occasion of his 95th birthday. Financial support from the National Natural Science Foundation of China (21702104), the Natural Science Foundation of Jiangsu Province, China (BK20170983), and a Startup Grant from Nanjing Tech University (39837120) are gratefully acknowledged. We also appreciate the SICAM Fellowship by Jiangsu National Synergetic Innovation Center for Advanced Materials.

REFERENCES (1) For recent reviews on decarboxylation reactions, see: (a) Weaver, J. D.; Recio, A.; Grenning, A. J.; Tunge, J. A. Transition metal-catalyzed decarboxylative allylation and benzylation reactions. Chem. Rev. 2011, 111, 1846. (b) Rodríguez, N.; Goossen, L. J. Decarboxylative coupling reactions: a modern strategy for C–C-bond formation. Chem. Soc. Rev. 2011, 40, 5030. (c) Cornella, J.; Larrosa, I. Decarboxylative carbon-carbon bondforming transformations of (hetero)aromatic carboxylic acids. Synthesis 2012, 44, 653. (d) Wei, Y.; Hu, P.; Zhang, M.; Su, W. Metal-catalyzed decarboxylative C–H functionalization. Chem. Rev. 2017, 117, 8864. (e) Xuan, J.; Zhang, Z.-G.; Xiao, W.-J. Visible-light-induced decarboxylative functionalization of carboxylic acids and their derivatives. Angew. Chem., Int. Ed. 2015, 54, 15632. (f) Huang, H.; Jia, K.; Chen, Y. Radical decarboxylative functionalizations enabled by dual photoredox catalysis. ACS Catal. 2016, 6, 4983. (g) Fang, G.; Cong, X.; Zanoni, G.; Liu, Q.; Bi, X. Silver-based radical reactions: development and insights. Adv. Synth. Catal. 2017, 359, 1422. (2) See references 1e-g for review articles on RDFs; for selected examples, see: (a) Wang, Z.; Zhu, L.; Yin, F.; Su, Z.; Li, Z.; Li, C. Silver-catalyzed decarboxylative chlorination of aliphatic carboxylic acids. J. Am. Chem. Soc. 2012, 134, 4258. (b) Liu, X.; Wang, Z.; Cheng, X.; Li, C. Silvercatalyzed decarboxylative alkynylation of aliphatic carboxylic acids in aqueous solution. J. Am. Chem. Soc. 2012, 134, 14330. (c) Liu, C.; Wang, X.; Li, Z.; Cui, L.; Li, C. Silver-catalyzed decarboxylative radical azidation of aliphatic carboxylic acids in aqueous solution. J. Am. Chem. Soc. 2015, 137, 9820. (d) Tan, X.; Liu, Z.; Shen, H.; Zhang, P.; Zhang, Z.; Li, C. Silvercatalyzed decarboxylative trifluoromethylation of aliphatic carboxylic acids. J. Am. Chem. Soc. 2017, 139, 12430. (e) Zhu, Y.; Wen, X.; Song, S.; Jiao, N. Silver-catalyzed radical transformation of aliphatic carboxylic acids to oxime ethers. ACS Catal. 2016, 6, 6465. (f) Zuo, Z.; Ahneman, D. T.; Chu, L.; Terrett, J. A.; Doyle, A. G.; MacMillan, D. W. C. Merging photoredox with nickel catalysis: coupling of α-carboxyl sp3-carbons with aryl halides. Science 2014, 345. 437. (g) Chu, L.; Ohta, C.; Zuo, Z.; MacMillan, D. W. C. Carboxylic acids as a traceless activation group for conjugate additions: a three-step synthesis of (±)-pregabalin. J. Am. Chem. Soc. 2014, 136, 10886. (h) Johnston, C. P.; Smith, R. T.; Allmendinger, S.; MacMillan, D. W. C. Metallaphotoredox-catalysed sp3-sp3 cross-coupling of carboxylic acids with alkyl halides. Nature 2016, 536, 322. (i) Liang, Y.; Zhang, X.; MacMillan, D. W. C. Decarboxylative sp3 C–N coupling via dual copper and photoredox catalysis. Nature 2018, 559, 83. (j) Song, H.T.; Ding, W.; Zhou, Q.-Q.; Liu, J.; Lu, L.-Q.; Xiao, W.-J. Photocatalytic decarboxylative hydroxylation of carboxylic acids driven by visible light and using molecular oxygen. J. Org. Chem. 2016, 81, 7250. (k) Edwards, J. T.; Merchant, R. R.; McClymont, K. S.; Knouse, K. W.; Qin, T.; Malins, L. R.; Vokits, B.; Shaw, S. A.; Bao, D.-H.; Wei, F.-L.; Zhou, T.; Eastgate, M.

ACS Paragon Plus Environment

Page 4 of 6

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

Journal of the American Chemical Society D.; Baran, P. S. Decarboxylative alkenylation. Nature 2017, 545, 213. (l) Li, C.; Wang, J.; Barton, L. M.; Yu, S.; Tian, M.; Peters, D. S.; Kumar, M.; Yu, A. W.; Johnson, K. A.; Chatterjee, A. K.; Yan, M.; Baran, P. S. Decarboxylative borylation. Science 2017, 356, eaam7355. (m) Fawcett, A.; Pradeilles, J.; Wang, Y.; Mutsuga, T.; Myers, E. L.; Aggarwal, V. K. Photoinduced decarboxylative borylation of carboxylic acids. Science 2017, 357, 283. (n) Hu, D. W.; Wang, L. H.; Li, P. F. Decarboxylative borylation of aliphatic esters under visible-light photoredox conditions. Org. Lett. 2017, 19, 2770. (o) Zhao, W.; Wurz, R. P.; Peters, J. C.; Fu, G. C. Photoinduced, copper-catalyzed decarboxylative C−N coupling to generate protected amines: an alternative to the curtius rearrangement. J. Am. Chem. Soc. 2017, 139, 12153. (3) (a) Kürti, L.; Czakó, B. Strategic applications of named reactions in organic synthesis: background and detailed mechanisms; Elsevier: Boston, 2005, 272-273. (b) Blanchet, J.; Baudoux, J.; Amere, M.; Lasne, M.-C.; Rouden, J. Asymmetric malonic and acetoacetic acid syntheses – a century of enantioselective decarboxylative protonations. Eur. J. Org. Chem. 2008, 5493. (4) Blankenship, C.; Paquette, L. A. Double Hunsdiecker reactions. Convenient preparation of 1,1-dibromocyclopropane and 1,1dibromocyclobutane. Synth. Commun. 1984, 14, 983. (5) (a) Tufariello, J. J.; Kissel, W. J. A new method for the oxidative decarboxylation of disubstituted malonic acids using lead tetraacetate. Tetrahedron Lett. 1966, 7, 6145. (b) Nokami, J.; Yamamoto, T.; Kawada, M.; Izumi, M.; Ochi, N.; Okawara, R. New synthetic reaction by electrolysis malonic acid as a ketone synthon. Tetrahedron Lett. 1979, 20, 1047. (c) Ma, X.; Luo, X.; Dochain, S.; Mathot, C.; Markò, I. E. Electrochemical oxidative decarboxylation of malonic acid derivatives: a method for the synthesis of ketals and ketones. Org. Lett. 2015, 17, 4690. (6) Griffin, J. D.; Zeller, M. A.; Nicewicz, D. A. Hydrodecarboxylation of carboxylic and malonic acid derivatives via organic photoredox catalysis: substrate scope and mechanistic insight. J. Am. Chem. Soc. 2015, 137, 11340. (7) (a) Müller, K.; Faeh, C.; Diederich, F. Fluorine in pharmaceuticals: looking beyond intuition. Science 2007, 317, 1881. (b) O’Hagan, D. Understanding organofluorine chemistry. An introduction to the C–F bond. Chem. Soc. Rev. 2008, 37, 308. (8) (a) Chen, B.; Vicic, D. A. Transition-metal-catalyzed difluoromethylation, difluoromethylenation, and polydifluoromethylenation reactions. Top. Organomet. Chem. 2014, 52, 113. (b) Xu, P.; Guo, S.; Wang, L.; Tang, P. Recent advances in the synthesis of difluoromethylated arenes. Synlett 2015, 26, 36. (c) Ni, C.; Zhu, L.; Hu, J. Advances in transition-metal-mediated di- and monofluoroalkylations. Acta Chim. Sin. 2015, 73, 90. (d) Feng, Z.; Xiao, Y.-L.; Zhang, X. Transition-metal (Cu, Pd, Ni)-catalyzed difluoroalkylation via cross-coupling with difluoroalkyl halides. Acc. Chem. Res. 2018, 51, 2264. (9) (a) Markovskij, L. N.; Pashinnik, V. E.; Kirsanov, A. V. Application of dialkylaminosulfur trifluorides in the synthesis of fluoroorganic compounds. Synthesis 1973, 787. (b) Middleton, W. J. New fluorinating reagents. Dialkylaminosulfur fluorides. J. Org. Chem. 1975, 40, 574. (c) Singh, R. P.; Shreeve, J. M. One-pot route to new α,α-difluoroamides and αketoamides. J. Org. Chem. 2003, 68, 6063. (10) (a) Olah, G. A.; Welch, J. T.; Vankar, Y. D.; Nojima, M.; Kerekes, I.; Olah, J. A. Synthetic methods and reactions. 63. Pyridinium poly(hydrogen fluoride) (30% pyridine-70% hydrogen fluoride): a convenient reagent for organic fluorination reactions. J. Org. Chem. 1979, 44, 3872. (b) Okoromoba, O. E.; Han, J.; Hammond, G. B.; Xu, B. Designer HF-based fluorination reagent: highly regioselective synthesis of fluoroalkenes and gem-difluoromethylene compounds from alkynes. J. Am. Chem. Soc. 2014, 136, 14381. (11) (a) Chernov, G. N.; Levin, V. V.; Kokorekin, V. A.; Struchkova, M. I.; Dilman, A. D. Interaction of gem-difluorinated iodides with silyl enol ethers mediated by photoredox catalysis. Adv. Synth. Catal. 2017, 359, 3063. (b) Supranovich, V. I.; Levin, V. V.; Struchkova, M. I.; Hu, J.; Dilman, A. D. Visible light-mediated difluoroalkylation of electron-deficient alkenes. Beilstein J. Org. Chem. 2018, 14, 1637. (12) (a) Gu, J.-W.; Min, Q.-Q.; Yu, L.-C.; Zhang, X. Tandem difluoroalkylation-arylation of enamides catalyzed by nickel. Angew. Chem. Int. Ed. 2016, 55, 12270. (b) An, L.; Xu, C.; Zhang, X. Highly selective nickel-catalyzed gem-difluoropropargylation of unactivated alkylzinc reagents. Nat. Commun. 2017, 8, 1460. (13) (a) Yin, F.; Wang, Z.; Li, Z.; Li, C. Silver-catalyzed decarboxylative fluorination of aliphatic carboxylic acids in aqueous solution. J. Am. Chem. Soc. 2012, 134, 10401. (b) Rueda-Becerril, M.; Sazepin, C. C.; Leung, J. C.

T.; Okbinoglu, T.; Kennepohl, P.; Paquin, J.-F.; Sammis, G. M. Fluorine transfer to alkyl radicals. J. Am. Chem. Soc. 2012, 134, 4026. (c) RuedaBecerril, M.; Mahé, O.; Drouin, M.; Majewski, M. B.; West, J. G.; Wolf, M. O.; Sammis, G. M.; Paquin, J. F. Direct C-F bond formation using photoredox catalysis. J. Am. Chem. Soc. 2014, 136, 2637. (d) Ventre, S.; Petronijevic, F. R.; MacMillan, D. W. C. Decarboxylative fluorination of aliphatic carboxylic acids via photoredox catalysis. J. Am. Chem. Soc. 2015, 137, 5654. (e) Huang, X.; Liu, W.; Hooker, J. M.; Groves, J. T.; Targeted fluorination with the fluoride ion by manganese-catalyzed decarboxylation. Angew. Chem. Int. Ed. 2015, 54, 5241. (14) The hydroxyl group of 2s was oxidized to form a ketone. (15) 4-(1,3-Dioxoisoindolin-2-yl)-2,2-difluorobutanoic acid was isolated as a side product in a 28% isolated yield. (16) To facilitate purification, the α-fluorocarboxylic acids have been converted to the corresponding methyl esters by treatment with trimethylsilyldiazomethane. (17) (a) Rozen, S.; Hagooly, A.; Harduf, R. Synthesis of αfluorocarboxylates from the corresponding acids using acetyl hypofluorite. J. Org. Chem. 2001, 66, 7464. (b) Lee, S. Y.; Neufeind, S.; Fu, G. C. Enantioselective nucleophile-catalyzed synthesis of tertiary alkyl fluorides via the α-fluorination of ketenes: synthetic and mechanistic studies. J. Am. Chem. Soc. 2014, 136, 8899. (c) McDonald, I. M.; Mate, R. A.; Zusi, F. C.; Huang, H.; Post-Munson, D. J.; Ferrante, M. A.; Gallagher, L.; Bertekap, R. L., Jr.; Knox, R. J.; Robertson, B. J.; Harden, D. G.; Morgan, D. G.; Lodge, N. J.; Dworetzky, S. I.; Olson, R. E.; Macor, J. E. Discovery of a novel series of quinolone α7 nicotinic acetylcholine receptor agonists. Bioorg. Med. Chem. Lett. 2013, 23, 1684. (18) Jiang, X.; Sakthivel, S.; Kulbitski, K.; Nisnevich, G.; Gandelman, M. Efficient synthesis of secondary alkyl fluorides via Suzuki cross-coupling reaction of 1-halo-1-fluoroalkanes. J. Am. Chem. Soc. 2014, 136, 9548. (19) Pacheco, M. C.; Purser, S.; Gouverneur, V. The chemistry of propargylic and allylic fluorides. Chem. Rev. 2008, 108, 1943. (20) (a) Patel, N. R.; Flowers, R. A. Mechanistic study of silver-catalyzed decarboxylative fluorination. J. Org. Chem. 2015, 80, 5834. (b) Zhang X. Insights into the mechanism of silver-catalyzed decarboxylative fluorination. Comput. Theor. Chem. 2016, 1082, 11. (21) Jaros, S. W.; Guedes da Silva, M. F. C.; Florek, M.; Smoleński, P.; Pombeiro, A. J. L.; Kirillov, A. M. Silver(I) 1,3,5-triaza-7phosphaadamantane coordination polymers driven by substituted glutarate and malonate building blocks: self-assembly synthesis, structural features, and antimicrobial properties. Inorg. Chem. 2016, 55, 5886. (22) This was further confirmed by the time course study of the decarboxylative gem-difluorination of substrate 1p, see supporting information for details. (23) Since the chemoselectivity was mainly caused by the different reactivities of α-fluorocarboxylic acid 3 (Scheme 3) in decarboxylative mono- and gem-difluorination, we speculate that the contrasting reactivity behavior of 3 under different conditions might be due to different existing forms of corresponding silver α-fluorocarboxylate salt, in which the number of coordinated carboxylate ligands and different coordinated solvents with silver might affect the redox potential of the Ag(I)/Ag(II) couple, see: Kumar, A.; Neta, P. Complexation and oxidation of glycine and related compounds by silver(II). J. Am. Chem. Soc. 1980, 102, 7284.

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Insert Table of Contents artwork here F HOOC

COOH

R1 R2 1 malonic acid derivative R1= Ar, Alkyl; R2 = Alkyl

AgNO3 (cat.) Selectfluor solvent, base

HOOC

F

R1 R2 2 2/3 = 90->95:10-5 35-85% yields

R1 R2

-carboxylic acid radical key intermediate

F

COOH

R1 R2 3 3/2 = 78->95:22-5 30-92% yields

ACS Paragon Plus Environment

Page 6 of 6