Trichloromethanesulfonyl Chloride: A Chlorinating Reagent for

Water in asymmetric organocatalytic systems: a global perspective. Ciril Jimeno. Organic & Biomolecular Chemistry 2016 14 (26), 6147-6164 ...
0 downloads 0 Views 416KB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Note

Trichloromethanesulfonyl chloride: a chlorinating reagent for aldehydes Ciril Jimeno, Lidong Cao, and Philippe Renaud J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.5b02543 • Publication Date (Web): 08 Jan 2016 Downloaded from http://pubs.acs.org on January 14, 2016

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.

The Journal of Organic Chemistry 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 14

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

The Journal of Organic Chemistry

Trichloromethanesulfonyl chloride: a chlorinating reagent for aldehydes

Ciril Jimeno,a Lidong Caob and Philippe Renaudb*

a

Department of Biological Chemistry and Molecular Modelling. IQAC-CSIC. c/Jordi Girona 18-

24. E08034 Barcelona, Spain. b

Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012

Bern, Switzerland.

E-mail: [email protected]

Abstract. Trichloromethanesulfonyl chloride (CCl3SO2Cl), a commercially available reagent, has been found to perform efficiently in the α-chlorination of aldehydes, including its catalytic asymmetric version, under very mild reaction conditions. Under our reaction conditions, this compound outperforms typical chlorinating reagents for organic synthesis, facilitates work-up and purification of the product, and minimizes the formation of toxic, chlorinated organic waste.

α-Chlorinated aldehydes are very useful building blocks for synthesis of a of pharmaceutically important compounds.1 They can be prepared via a variety of reactions, among them, the direct catalytic α-chlorination of ketones and aldehydes is the most established procedure (Scheme 1). The catalysts used for this transformation are metal derivatives and secondary amines. By using chiral secondary amine catalysts, enantioselective α-chlorinations of aldehydes and ketones have been reported.

ACS Paragon Plus Environment

The Journal of Organic Chemistry

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

Scheme 1. General representation of the α-chlorination of ketone and aldehydes.

The selection of the chlorinating agent is also a key element of the process since it determines the nature of the by-products. It also strongly influences the mechanism of the chlorination reaction and therefore the yield as well as the level of enantioselectivity of the process.2 Numerous different chlorinating reagents have been used, ranging from inorganic compounds to polychlorinated organic molecules. Molecular chlorine may be considered the most atom economical, but its high reactivity and the difficulties associated with handling of this gas make it often impractical. No asymmetric version is known using gaseous chlorine, but a clever approach using lithium chloride combined with a strong oxidant like sodium persulfate and copper(II) salts has been described.5 Enantioselective catalytic chlorination are typically run with a chlorinated organic compound such as N-chlorosuccinimide (NCS) and a metal-6-10 or an organocatalyst.11-18 Other useful reagents include hypervalent iodine compounds,19 N-chlorophthalimide,20 2,2,6,6-tetrachlorocyclohexanone,21 hexachlorocyclohexadienone,22-25 and trichloroquinolinone,26 They have been widely used in combination with organocatalysts or metal catalysts. In any case, organic waste in the form of succinimide, phthalimide, chlorocyclohexanones and chlorophenols are generated requiring a careful purification of the products. Sulfuryl chloride27 and p-toluenesulfonyl chloride (TsCl)28 have also been employed with some success. Trichloromethylsulfonyl chloride had been also used as a chlorinating reagent of preformed silyl enol ethers in a ruthenium-catalyzed process.29 Interestingly, trifluoromethanesulfonyl fluoride (CF3SO2F) have been used in both non-catalyzed30 and asymmetric metal-catalyzed fluorination reactions.31 Recently, we described how the commercially available trichloromethanesulfonyl chloride, was an extremely efficient reagent for radical carbochlorination reactions.32,33 When trying to extend this reaction to the carbochlorination of enamines with trichloromethylsulfonyl chloride, we discovered that instead of the desired reaction, an electrophilic chlorination was taking place. Based on this initial observation, we report here a fast and clean method for the α-monochlorinations of aldehydes using pyrrolidine-type catalysts and trichloromethanesulfonyl chloride.34-38 This reagent proved to behave also very well in an enantioselective version of this reaction outperforming other well-established chlorinating

ACS Paragon Plus Environment

Page 2 of 14

Page 3 of 14

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

The Journal of Organic Chemistry

agents. The reaction conditions were optimized for the chlorination of 3-phenylpropanal (1a) using pyrrolidine as a catalyst in different solvents in the presence and in the absence of a base and water (eq. 1). Results are summarized in Table 1. Reaction in dry CH2Cl2 without base was disappointing, the conversion just reached only 31% and a mixture of mono- (2a) and dichlorinated (2a') aldehydes was obtained (entry 1, Table 1). In 1,2-dimethoxyethane (DME), the conversion was even lower (entry 2, Table 1). Gratifyingly, when two equivalents of a base, 2,6-lutidine, was added, the conversion reached 98% but 1:1 mixture of mono- and dichlorinated products were obtained. When three equivalents H2O was added in the absence of a base, only the monochlorinated aldehyde 2a was detected, but the reaction was very slow (entry 4, Table 1). The combination of 2,6-lutidine (2 equivalents) and H2O (3 equivalents) gave the best results (Table 1, entry 5). After reduction of the crude chlorinated aldehyde 2a with NaBH4, the stable 2-chloroalcohol 3a was obtained in 95% yield (entry 5, Table 1). The use of dichloromethane in the presence of 2,6-lutidine and H2O led to full conversion but lower chemoselectivity (Table 1, entry 6), which, after reductive work up, led to alcohol 3a in lower yield.

Table 1. Optimization of the chlorination reaction of 1a according to Equation 1.

Entry

Solvent

2,6-Lutidine

H2O

Conversiona

2a/2a'

Yield 3ab

1

CH2Cl2

-

-

31

1:1

-

2

DME

-

-

98

1:1

-

4

DME

-

3 equiv

49

>99:1

-

5

DME

2 equiv

3 equiv

>99

>99:1

95%

ACS Paragon Plus Environment

The Journal of Organic Chemistry

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

6 a

CH2Cl2

2 equiv

3 equiv

>99c

4:1

Page 4 of 14

74%

Determined by 1H-NMR on the crude mixture. b Isolated yield after flash chromatography.

The scope and limitation of the method was then examined with a range of aliphatic aldehydes (Scheme 2). Unsubstituted aldehydes 1a–1d provided the desired β-chloro alcohols 3a–3d in ≥89% yield. Substrates containing double bonds such as 1d and 1f reacted smoothly too. Phenylacetaldehyde 1g provided the desired product 3g in moderate yield together with some dichlorinated product. This result may be explained by the increased acidity of this system that favors a rapid isomerization of the α-chloroiminium ion to the chloroenamine competing with the hydrolysis step. Then, sterically more demanding αsubstituted aldehydes were tested. Reaction of 2-phenylpropanal 1h was slower, and the amount of CCl3SO2Cl had to be increased to 1.4 equivalents and the temperature to 60 °C. Under these conditions, the desired aldehyde was isolated in 61% yield. In the course of the reaction, a white precipitate appeared that we identified as the result of reaction of 2,6lutidine with CCl3SO2Cl. The structure of this precipitate could not be characterized but it showed a complex oligomeric nature. We rationalized that the rate of enamine formation from aldehydes and pyrrolidine decreased with α-substitution of the aldehydes and therefore the reaction between the base and the trichloromethylsulfonyl chloride can proceed, thus reducing the efficacy of the process. To avoid this competing reaction, the bulkier 2,6-di-tert-butylpyridine was used as a base and the yield increased to 80% after 5 hours. This procedure was then applied with success to other α-disubstituted aldehydes such as 1i and 1j.

ACS Paragon Plus Environment

Page 5 of 14

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

The Journal of Organic Chemistry

Scheme 2. α-Chlorination of aldehydes with CCl3SO2Cl and pyrrolidine as a catalyst. a) 2,2Dichloro-2-phenyl-ethan-1-ol (16%) is also formed. b) Using 2,6-di-tert-butylpyridine instead of 2,6-lutidine and 1.3 equivalents of CCl3SO2Cl.

Next, the chlorination of cyclohexanone 4 was examined (eq. 3). When 1.1 equivalents of CCl3SO2Cl were used under the optimized conditions developed for aldehydes (see above), the major product was the dichlorinated cyclohexanone. However, when using cyclohexanone in excess and longer reaction time (36 h), the desired α-chlorocyclohexanone 5 was obtained in a modest 43% yield together with some polychlorinated products. All attempts to improve this yield by using bases such as 2,6-di-tert-butylpyridine were unsuccessful (Scheme 3).

Scheme 3. α-Chlorination of cyclohexanone.

The efficient pyrrolidine catalyzed α-chlorination of aldehydes using CCl3SO2Cl as an electrophilic source of chlorine atom offers the possibility of developing an organocatalytic asymmetric version of this reaction. The α-chlorination of 3-phenylpropanal 1a and n-octanal 1b was investigated (equation 4, Table 2). 2,6-Lutidine was used as a base and DME as a

ACS Paragon Plus Environment

The Journal of Organic Chemistry

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

solvent.2 A rapid initial screening of commercial asymmetric organocatalysts showed that the simple diarylprolinol silyl ether 6 is a suitable catalyst for this reaction.39,40 Surprisingly, this commercially available and general catalyst has not been used in chlorination reaction at the exception of the mechanistic work of Blackmond.2 The addition of a small amount of water (3 equiv.) was essential for the reaction to proceed efficiently.41,42 To avoid racemization, the chlorinated aldehydes (S)-2a and (S)-2b were not isolated but directly reduced with sodium borohydride to the corresponding alcohol (S)-3a and (S)-3b. The alcohol (S)-3a was obtained in 74% yield and 81% ee (Table 2, entry 1). Aldehyde 1b was converted to (S)-3b in 98% yield and 93% ee under the same reaction conditions (Table 2, entry 5). The absolute configuration of (S)-3a and (S)-3b were attributed based on comparison of measured optical rotations with literature values (See SI).5 Both yields and ee's compare well with the results described in the literature.5,14,17,24 Higher ee's were only reported by Jorgensen14 (2 × 95% ee) and MacMillan24 (92% and 52% ee) for the chlorination of aldehydes 1a and 1b, respectively. The ee's and yields obtained under our conditions are noticeably higher than the one obtained by Blackmond with the same catalyst and NCS for the chlorination of 3methylpropanal.2 For comparison purposes, the chlorination of 1a with NCS was performed with and without 2,6-lutidine. Both reaction conditions afforded low enantioselectivities and low to moderate yields (Table 2, entries 2 and 3). Other chlorinating agents were also tested under our reaction conditions: hexachlorocyclohexadienone and trifluoromethanesulfonyl chloride gave with both substrates lower yields and enantioselectivities, (Table 2, entries 4, 6 and 7) but always the same absolute configuration of product.

Table 2. Enantioselective chlorination of aldehydes according to equation 4. Entry

Aldehyde 1

Reagent

Yield

Ee's

(S)-3a/3bb (S)-3a/3bc

ACS Paragon Plus Environment

Page 6 of 14

Page 7 of 14

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

The Journal of Organic Chemistry

1

1a (R = Ph)

CCl3SO2Cl

74%

81% ee

2

1a (R = Ph)

NCS