Microdroplets as Microreactors for Fast Synthesis of Ketoximes and

Dec 24, 2018 - In addition, the synthesis in microdroplets also saves reactants and solvents and reduces the waste amounts. All of these merits indica...
0 downloads 0 Views 516KB Size
Subscriber access provided by RMIT University Library

Article

Microdroplets as microreactors for fast synthesis of ketoximes and amides Wenwen Zhang, Shiwei Yang, Qiuyu Lin, Heyong Cheng, and Jinhua Liu J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.8b02669 • Publication Date (Web): 24 Dec 2018 Downloaded from http://pubs.acs.org on January 1, 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 20 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

Title:

Microdroplets as microreactors for fast synthesis of ketoximes and amides Author names and affiliations: Wenwen Zhang a, Shiwei Yang a, Qiuyu Lin a, Heyong Cheng a, c*, Jinhua Liu b, c* a

College of Material Chemistry and Chemical Engineering, Hangzhou Normal University, Hangzhou 311121,

China b

Qianjiang College, Hangzhou Normal University, Hangzhou 310036, China

c

Key Laboratory of Organosilicon Chemistry and Material Technology, Hangzhou Normal University, Hangzhou

311121, China

* Corresponding author: Tel: +86-571-28866903.

Fax: +86-571-28866903.

E-mail: [email protected] (H.-Y. Cheng);

[email protected] (J.-H. Liu)

1

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

Page 2 of 20

Abstract: The formation of amide bonds is one of the most valuable transformations in organic synthesis. Beckmann rearrangement is a well-known method for producing secondary amides from ketoximes. This study demonstrates the rapid synthesis of ketoximes and amides in microdroplets. Many factors are found to affect the yield, such as microdroplet generation devices, temperature, catalysts, concentration of reactants. In particular, the temperature has a great influence on the synthesis of amide, which is demonstrated by a sharp ascendance to the yield when the temperature was increased to 45 °C. The best amide yield (93.3%) can be obtained by using a coaxial flowing devices and a sulfonyl chloride compound as a catalyst, and heating to 55 °C in microdroplets. The yields can reach 78.7-91.3% for benzoylaniline and 87.2-93.4% for benzophenone oximes in several seconds in microdroplets compared to 10.1-66.1% and 82.5-93.3% in several hours in bulk phase. Apart from the dramatically decreased reaction time and enhanced reaction yields, the microdroplet synthesis is also free of severe reaction environments (anhydrous and anaerobic conditions). In addition, the synthesis in microdroplets also saves reactants and solvents, and reduces the waste amounts. All these merits indicate the microdroplet synthesis is a high-efficiency green methodology. Keywords: Amides; Oximes; Beckmann rearrangement; microdroplet synthesis; green chemistry

2

ACS Paragon Plus Environment

Page 3 of 20 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

Introduction Due to the widespread presence of amides in drugs, natural products and biologically active compounds, amide synthesisis one of the most valuable transformations in organic chemistry1. The most common methods for producing amides need the activation of carboxylic derivatives such as acyl chlorides, anhydrides and esters. Alternatively, carboxylic acid is reacted directly with amine assisted by a coupling agent of stoichiometric amount, such as a carbodiimide or a 1H-benzotriazole derivative2, 3. However, these classical methods have low atomic efficiency, leading to environmental pollution. Therefore, the Pharmaceutical Roundtable identified ‘‘amide formation adopting high atom economy reactants as one of the most challengeable tasks in organic chemistry4. New effective and sustainable methods to synthesize important compounds are desirable5. In the search for more efficient, atomic and economical methods, the appearance of metal catalytic conversions in recent years provides a new synthetic route and expands the previous synthetic substrates6. With the aid of transition metals, many functional groups such as nitriles, aldehydes, ketones, oximes, primary alcohols and amines can be conveniently used as starting materials for the construction of amide bonds. The use of oximes as starting materials to obtain amides by rearrangement has a long history, the so-called Beckmann rearrangement reaction. The reaction was first discovered by the German chemist Beckmann in 18867, 8. For example, ε-caprolactam which is a precursor to produce nylon-6 is synthesized by Beckmann rearrangement of cyclohexanone oxime9. The Beckmann rearrangement of oximes involves the translocation of a group which locates in the trans position of hydroxy groups from carbon to nitrogen atom. This process usually requires a Brønsted or Lewis acid as a catalyst and is completed under severe temperature conditions10, 11. In the last five decades, chemists have been working on the screening of this catalytic reaction system. Diverse catalytic systems have been developed including liquid phase systems12, vapor phase systems13, supercritical water systems14, ionic liquid systems15, etc. The advantages of mild conditions, ease of post-processing and industrial applicability have stimulated many researchers to learn about liquid phase catalysis of Beckmann rearrangement. They have developed a lot of catalysts such as inorganic and organic catalysts, and metal Lewis acids ([RhCl(cod)]216,Y(OTf)317, Ga(OTf)318, FeCl319 , AlCl320 , and HgCl221).

However, these methods suffered from several drawbacks such as use of

toxic/costly solvents, co-catalysts, expensive reagents, production of considerable amount of byproducts, low yields and long reaction time. In the Beckmann rearrangement reaction, organic sulfonyl chlorides are widely used reagents for stoichiometric 3

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

Page 4 of 20

dehydrogenation. p-Toluenesulfonyl chloride (TsCl) can efficiently catalyze the conversion of various oximes to corresponding amides under mild conditions with excellent yields22. The mechanism of this catalyst has been proposed by a new autonomous cycle, where TsCl initiates Beckmann rearrangement by producing an azayne cation intermediate23. Scientists also explored the use of ZnCl2 as a co-catalyst where the amide yield was improved24. Generally, the rearrangement reaction catalyzed by p-toluenesulfonyl chloride needs to be carried out under severe conditions including anhydrous, anaerobic, and at a certain high temperature for several hours. Many researches have demonstrated that microdroplet reactions can be accelerated by orders of magnitude in comparison with bulk reactions25,26. Although the study of ultra-fast reactions in microdroplets is in the early stage, it is worth noting that important applications have emerged27-29. The synthesis of large quantities of isoquinoline using the Pomeranz-Fritsch reaction is taken as a typical example. It takes a long time (a few days) and requires a very high concentration of acid in bulk solution30. The same reaction occurring in microdroplets from an electrospray ionization source, however, can be completed in milliseconds with no need of any external acid31. Green chemistry has increasingly attracted the interest of chemical workers in recent years32,33. Scientists are looking for a gentle, efficient and green method for the synthesis of ketoximes and amides. In this paper, we explored synthesize ketoximes and amides with different substituents in microdroplets (which were produced by homemade sheath-gas-assisted spray emitters without applying any high voltage in Figure 1). The products in microdroplets were collected and then analyzed by high performance liquid chromatography for yields. Microdroplet synthesis accelerates the reaction and meets the expectations of green chemistry.

Figure 1 (A) Synthetic ketoximes in microdroplets with the online mixing device; (B) Synthetic amides in microdroplets with the coaxial flowing device 4

ACS Paragon Plus Environment

Page 5 of 20 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

Results and discussion Screening of optimal conditions for ketoxime synthesis Hydroxylamine hydrochloride is commonly used in the preparation of oximes (Scheme 1) because hydroxylamine is stable under acidic conditions. In the preliminary experiment, we tried to use hydroxylamine hydrochloride as a reactant for synthesizing benzophenone oxime. We observed salt deposition when hydroxylamine hydrochloride and sodium hydroxide were mixed, which led to tip clogging of capillaries of small internal diameters. Thereafter hydroxylamine selected as a reactant. From Figure 2a, it can be seen that good yield of benzophenone oxime is obtained with 5-10 mol L-1 NH2OH, indicating a selection of 5 mol L-1 NH2OH for the following experiment.

HO

O

N

NaOH NH2OH R2 R1

R1

R2

OM1: R1, R2 = H

OM2: R1, R2 = CH3 OM3: R1, R2 = Cl

OM4: R1, R2 = F

OM5: R1, R2 = Br

OM7: R1 = H, R2 = CH3

OM6: R1, R2 = OCH3

OM8: R1 = F, R2 = OCH3

Scheme 1 Synthesis of ketoximes from ketones in microdroplets

Figure 2 Production of benzophenone oxime in microdroplets using different NH2OH concentrations (a) and sodium hydroxide concentrations (b) at different temperatures

The reaction mechanism of ketones to form oximes is the nucleophilic addition. Briefly, the lone electron pair on the nitrogen of hydroxylamine attacks the carbonyl to make the formation of a hydroxyl group between the hydrogen hydroxylamine and the carbonyl oxygen. The -CH=NOH group is then generated by the carbon dehydroxylation and nitrogen dehydrogenation under the alkaline conditions. From Figure 2b and Figures S1 and S2, we can see at low temperatures (25-45 °C), the yield of benzophenone oxime increases gradually when the sodium hydroxide concentration increases from 1.0 to 2.0 mol L-1, but the yield was stable by using 2.0-3.0 mol 5

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

Page 6 of 20

L-1 NaOH. However, when the reaction temperature is higher than 45 °C, the yield decreases dramatically with the increase of the sodium hydroxide concentration. Sodium hydroxide of high concentrations is apt to crystallize at high temperatures in capillaries of small internal diameters to clog the capillary tip, leading to a failure in microdroplet generation. It is also observable from Figure 2b that the yield of benzophenone oxime using low NaOH concentrations gradually increased on increasing the temperature from 25 to 45 °C, but was apparently reduced on further increasing the temperature to 65 °C. When NaOH of high concentrations (2.0-3.0 M NaOH), the yield was reversely drastically decreased with the temperature increase. Moderately increased temperature facilitates the solvent evaporation of microdroplets, leading to highly concentrated reactants to enhance the reaction rate (not rate constant) in microdroplets. However, the wet microdroplets may be converted into dry particles upon the evaporation of the solvents at high temperatures. A 2.5 mol L-1 NaOH was finally selected for the oxime synthesis at a temperature of 35 °C. In addition, a deep checking to Figure 2 demonstrates that the reactant concentrations had more obvious influence on the conversion than the reaction temperature. As is well-known, the conversion of ketone to ketoxime initiates with a hydroxylamine attack toward carbonyl, which can be reasonably promoted by more hydroxylamines (high concentrations in microdroplets). Similarly, the reaction ends with carbon dehydroxylation and nitrogen dehydrogenation, which can also be boosted under strong alkaline conditions (high NaOH concentrations in microdroplets). Under the optimal conditions, several aromatic oximes were synthesized in microdroplets by the online mixing device (Figure 1A) from their corresponding ketones. It is observable from Table 1 that the yields in microdroplets were comparable to those in bulk solution. However, the reaction in microdroplets can be completed within several seconds whereas the synthesis in bulk solution typically takes one hour around. Table 1 also demonstrates high yields around 90% for all tested benzophenones in microdroplets or bulk phase no matter if activating or deactivating substituents are attached to the benzene ring because carbonyl is the target group in the synthesis of benzophenone ketoximes. The crude oxime products after 6 h collection were purified by column chromatography for structure certification by NMR spectra (Figure S12-19). Table 1 Synthesis of various ketoximes in microdroplets and bulk solution Reagent O

Product HO

Yielda (%)

Yieldb (%)

90.4%

92.5%

N

6

ACS Paragon Plus Environment

Page 7 of 20 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

HO

O

F

F

F HO

OCH3

N

N

F a

OCH3

H3CO

93.4%

86.5%

81.2%

79.9%

95.3%

94.6%

Br OH

HO

N

+ O

91.6%

N

Br

O

91.1% Cl HO

Br

93.3%

N

Cl

O

Br

88.2% CH3

HO

Cl

82.5%

N

H 3C

O

Cl

93.5% OCH3

HO

CH3

91.9%

N

H3CO

O

H 3C

87.2% F

O

H3CO

N

OH

N

F

+F

OH

OCH3

synthesis of ketoxime in microdroplets; b synthesis of ketoxime in bulk solution.

Screening of optimal conditions for amide synthesis Following the synthesis of oximes in microdroplets, we further investigated the feasibility of the transformation of oximes to amides in microdroplets (Scheme 2). Benzophenone oxime was taken as an example to explore the effects of different synthetic conditions on the Beckmann rearrangement. We preliminarily tested the reaction in microdroplets produced by the same device used in the synthesis of oximes (Figure 1A). Conversion ratios no more than 67.6% were obtained no matter how the reaction temperature was optimized (Figure S3). We then tried to offline mix benzophenone ketoxime with TsCl ahead of the pneumatic spraying, which demonstrates a slightly improved yield (< 70.3%). It indicated that online mixing was as efficient as offline mixing and the transformation was not merely dependent on the mixing degree. We further tested the coaxial flowing device in Figure 1B (ketoxime surrounding catalysts for the setup b and ketoxime surrounded by catalysts for the setup a). The conversion ratios by using the two coaxial flow devices (85.1-88.7%) were significantly superior over those by the offline and/or online mixing device (Figure S3). The yield was slightly enhanced when the catalyst (TsCl) flowed in the inner capillary (the coaxial flowing device b) instead of the outer capillary (the coaxial flowing device a). Therefore the coaxial flow device a was selected for microdroplet-based synthesis of amides. 7

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

HO

N

Page 8 of 20

R2

O Catalyst

N

CH3CN R1 R2

R1

H

AD1: R1, R2 = H

AD2: R1, R2 = CH3 AD3: R1, R2 = Cl

AD4: R1, R2 = F

AD5: R1, R2 = Br

AD7: R1 = H, R2 = CH3

AD6: R1, R2 = OCH3

AD8: R1 = F, R2 = OCH3

Scheme 2 Catalytic synthesis of amides by oximes

It was observable that Beckmann rearrangement of ketoxime is highly relative to the reaction temperature (Figure 3a). Ketoximes are hardly converted into amides when the reaction temperature was below 25°C. A sharp ascendance to the yield is observed when the temperature ramped to 45°C (Figure 3b). However, the yield was drastically reduced on further increasing the temperature from 55°C to 75°C. It was well-known that p-toluenesulfonyl chloride is an initiator in this reaction34. A comparison experiment was performed to ensure if the reaction was irradiatively initiated, where the reaction took place in microdroplets at room temperature with and without UV lighting. The experimental results demonstrated no apparent difference to the yields on the both occasions. All these results proved the thermally initiated conversion of oximes to amides. The reaction temperature was chosen at 55°C considering the best yield. However, we observed that more ketoximes were converted into corresponding amides at higher temperature for the bulk-phase reaction (Figure 3c). Such a different temperature effect on the Beckmann rearrangement from the microdroplet synthesis can be ascribed by rapid solvent evaporation of microdroplets. Solvent evaporation of microdroplets at moderately high temperatures increases reagents concentrations to promote the reaction (not rate constant). However, microdroplets may be transformed into dry particles to inhibit the liquid reaction owing to excessive evaporation at extremely high temperatures. In comparison with microdroplets, solvent evaporation of the bulk solution was very mild to maintain the liquid reaction in bulk phase.

Figure 3 HPLC chromatograms (a) and yields (b) showing the microdroplet synthesis of amides at different temperatures, and 8

ACS Paragon Plus Environment

Page 9 of 20 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

dependence of amide yields upon reaction temperature for bulk-phase Beckmann rearrangement (c). Peak identity: 1, Benzoylaniline; 2, Benzophenone oxime; 3, Tosyl chloride.

It was reported that the Beckmann rearrangement reaction can be catalyzed by acids (citric acid, trifluoroacetic acid, etc.), salts (anhydrous aluminum, FeCl3, etc.) and sulfonyl chlorides19,20,22,35. Therefore we explored the catalyzed actions of the above compounds for the synthesis of amides in microdroplets. It was observable that benzophenone oxime cannot be converted into the corresponding amide under the catalysis of citric acid, CF3COOH and AlCl3 except sulfonyl chlorides, which was ascribed by harsh reaction conditions required by these catalysts (high temperature without solvents for citric acid, anhydrous environment for AlCl3, etc.). We then investigated several sulfonyl chlorides (2,4,6-trimethylbenzenesulfonyl chloride, 4-methoxybenzenesulfonyl chloride,

4-trifluoromethylbenzenesulfonyl

chloride

and

D(+)-10-camphorsulfonyl

chloride)

for

the

microdroplet-based synthesis of amides. The results in Table 2 and Figure S4 show superior yields in microdroplets over those in bulk solution for all tested sulfonyl chlorides, proving ultra-fast reaction rates (not rate constant) by microdroplets based reactors. It also demonstrates that the best yield was obtained with tosyl chloride. Therefore tosyl chloride was selected as the catalyst. Table 2 Yields of benzoylaniline by different catalysts in microdroplets and bulk solution Catalyst Tosyl chloride

Structural formula

SO2Cl

H3C

Yielda (%)

Yieldb (%)

91.3%

77.0%

89.5%

45.4%

87. 0%

33.8%

88.7%

58.5%

53.9%

15.2%

SO2Cl

2,4,6-trimethylbenzenesulfonyl chloride

CH3

H3C CH3

4-methoxybenzenesulfonyl chloride

SO2Cl

H3CO

4-trifluoromethylbenzenesulfony l chloride

D(+)-10-Camphorsulfonyl

SO2Cl

F3C

O

SO2Cl

chloride

H a

synthesis of benzoylaniline in microdroplets; b synthesis of benzoylaniline in bulk solution.

We further extended the Beckmann rearrangement in microdroplets to benzophenone oximes with different substituents (4,4'-dichlorobenzophenone oxime, 4,4'-dimethylbenzophenone oxime, 4,4'-dimethoxy benzophenone 9

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

Page 10 of 20

oxime etc.) with p-trifluoromethylbenzenesulfonyl chloride as a catalyst and zinc chloride as a co-catalyst. The yields of these substituted amides produced in microdroplets and in bulk solution are listed in Table 3. HPLC chromatograms of the crude amides are shown in Figures S5-S11, and 1H NMR spectra and mass spectra of the amides are shown in Figures S20-S35 after 6 h collection. It demonstrated that the microdroplet can significantly enhance the conversion yields of oximes to amides in comparison with the reaction in bulk solution. It was observed from Table 3 that the yield difference between the microdroplet synthesis and the bulk reaction was significant in the presence of substitutes attached to benzene rings, which can be ascribed by steric hindrance of all substituted groups. In bulk solution, the steric hindrance was apparent to prohibit the formation of the dimer-like cation intermediate23, and thus the Beckmann rearrangement was retarded. However, such steric hindrance may be negligible because of surface effect and small size effect in microdroplets of micrometer to nanometer sizes. The Beckmann rearrangement was also accelerated sharply in several seconds under mild experimental conditions in microdroplets versus several hours under harsh conditions (anhydrous and anaerobic) in bulk solution. In addition, the microdroplets as microreactors saved solvents and reactants, and decreased the waste amounts dramatically. All these figures meet the expectations of green chemistry. Apart from the above merits, the product microdroplets in microdroplet synthesis can be conveniently introduced into a commercial mass spectrometry for mechanism confirmation. We used an LTQ 133 Orbitrap XL hybrid mass spectrometer to probe residual reactants and products and mainly unstable intermediates for this aim. Two transient intermediates (IM1 and IM2 in Scheme S1) at m/z 180.0807 (theoretical m/z 180.0808, error -0.56 ppm) and 377.1643 (theoretical m/z 377.1648, error -1.3 ppm) can be observed from the mass spectrum in Figure S36a. The intermediate IM1 at m/z 180.0807 may form a fragment ion at m/z 77.0387 (C6H5+, theoretical m/z 77.0386, error 1.3 ppm) upon collision induced dissolution (CID) mass spectrometry (Figure S36b) to confirm its structure whereas the intermediate IM2 cannot be confirmed by MS upon CID because of low intensity. The main peak at m/z 198.0917 (theoretical m/z 198.0913, error 2.0 ppm) in Figure S36a corresponded to the residual benzophenone oxime and the synthetic benzoylaniline, which can be confirmed by the mass spectrum in Figure S36c upon CID. By comparing with the mass spectra of the benzophenone oxime and benzoylaniline standard solutions upon CID (Figures S36d and 36e), the fragment ion at m/z 180.0810 (theoretical m/z 180.0813, error -1.7 ppm) belonged to the only precursor ion of the protonated benzophenone oxime by the loss of H2O while the fragment ions at m/z 92.0500 (theoretical m/z 92.0500) and 105.0340 (theoretical m/z 105.0340) were attributed from the only precursor ion of the protonated benzoylaniline by the losses of C7H6O and C6H7N, respectively. In addition, the minor peaks at m/z 295.9814, 457.0891 and 493.0660 corresponded to the three adducts of 10

ACS Paragon Plus Environment

Page 11 of 20 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

benzophenone oxime and benzoylaniline with ZnCl2 ([M+H+ZnCl]+ (theoretical m/z 295.9821, error -2.4 ppm), [2M+Zn]+ (theoretical m/z 457.0894, error -0.66 ppm) and [2M+ZnCl]+ (theoretical m/z 493.0661, error -0.2 ppm)). All the MS information proved the synthesis of benzoylaniline from benzophenone oxime following the reported mechanism (Scheme S1). However, MS can only detect the residual benzophenone at m/z 183.0808 (theoretical m/z 183.0804, error 2.2 ppm) and the synthetic benzophenone oxime at m/z 198.0918 (theoretical m/z 198. 0913, error 2.5 ppm) from the mass spectra in Figure S37. We cannot find any intermediate using MS because the intermediates were so transient to probe in the ultra-fast transformation of benzophenone to benzophenone oxime. Overall, the microdroplet synthesis was feasible to couple with MS for mechanism confirmation. Last but not the least, the microdroplet synthesis coupled with MS detection was invaluable to pharmaceutical discovery because they facilitated reaction screening and optimization by reducing the analysis time36. In particular, upon the use of high throughput array design, the microdroplet synthesis will guide scalable syntheses in bulk continuous-flow reactions36. It must be noted that the obtained amounts of all products were too small (less than 50 μg min-1) to limit its application in large scale synthesis (preparative or industrial production). By combing with multiplex and array sprayers36-38 or high flux devices using the array internal-mix nozzles29 for microdroplets generation, the scaling up microdroplet synthesis can achieve high rates of several tens of milligram per minute for isolated products, which may be comparable to the bulk phase reaction. Table 3 Synthesis of various amides in microdroplets and bulk solution Reagent HO

Product

Yielda (%)

Yieldb (%)

91.3%

77.0%

78.7%

37.2%

81.1%

32.5%

87.8%

66.1%

O

N

N H

HO

F

O

N

N F

F HO

H

F

N

OCH3

O N

H3CO

OCH3 HO

H

H3CO

N

CH3

O N

H3C

CH3

H3C

H

11

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

HO

N

Page 12 of 20

Cl

O N

Cl HO

N

N

Br HO

OH

OH

N

N

N

N

O

O

93.0%

88.2%

H N

a

13.5%

OH H

F

77.6%

H

H

OCH3

F

H3CO

50.7%

H

Br

+ N

83.1%

Br

O

N

Br

10.1%

H

Cl

Cl

85.6%

+F

O

OCH3

N

F

H3CO O

synthesis of various amides in microdroplets; b synthesis of various amides in bulk solution

Conclusion In summary, we synthesized ketoximes and amides in microdroplets for the first time. Compared with the bulk-phase counterpart, the yields of amides in microdroplets were improved significantly and the reaction time was greatly shortened within several seconds. The microdroplets as microreactors saved organic solvents, the experiment operation is simple, conditions are mild, and decreased the waste amounts dramatically. It shows that the microdroplet synthesis method is suitable for the synthesis of ketoximes and amides, which also meets the expectations of green chemistry. Furthermore, the microdroplet synthesis was feasible to couple with MS for mechanism confirmation, and it also promoted reaction screening and optimization to guide scalable syntheses in pharmaceutical discovery. However, the formation of amides from ketoximes based on Beckmann rearrangement in this work is just one but not general method because alcohols, aldehydes/ketones and unsaturated hydrocarbons typically react with organic/inorganic amines for the corresponding amides39. The obtained amounts of ketoximes and amides should be further enhanced from micrograms to several tens of milligrams per minute by coupling with array sprayers36-38 or high flux devices29 for generating microdroplets.

EXPERIMENTAL SECTION General methods Benzophenone,

sodium

4,4'-dimethoxybenzophenone,

hydroxide,

4,4'-difluorobenzophenone,

4,4'-dibromobenzophenone,

4,4'-dimethylbenzophenone,

4,4'-dichlorobenzophenone,

zinc

chloride,

4-methylbenzophenone and 4-fluoro-4'-methoxybenzophenone were obtained from Energy Chemical (Shanghai, 12

ACS Paragon Plus Environment

Page 13 of 20 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

China). Hydroxylamine hydrochloride, citric acid, hydroxylamine, 4-methoxybenzenesulfonyl chloride, benzophenone oxime, 4-trifluoromethylbenzenesulfonyl chloride, 2,4,6-trimethylbenzenesulfonyl chloride, p-toluenesulfonyl chloride and

D-(+)-10-Camphorsulfonyl chloride were purchased from Sinopharm Chemical

Reagent (Shanghai, China). Methanol and acetonitrile of chromatographic grade were purchased from J&K (Shanghai, China). Benzoylanilide was purchased from TCI (Shanghai, China). Fused-silica capillaries of 50 μm i.d. × 150 μm o.d., 150 μm i.d. × 365 μm o.d. and 200 μm i.d. × 365 μm o.d. were obtained from Yongnian Optic Fiber Factory (Hebei, China). A Model 35-2226 syringe pump from Harvard Apparatus (Harvard Apparatus Inc., MA, USA) equipped with two 1.0 mL syringes was utilized for solution feeding. A Shimadzu LC-20AT HPLC system (Shimadzu (Suzhou) Co., Ltd., Suzhou, China) was used for all yield measurements by high performance liquid chromatography (HPLC). It consisted of two double plunger pumps, a mixer with two inlets and one outlet, a manual six-port injection valve with a 20 μL sample loop, an Agilent Zorbax C18 column (4.6 mm i.d. × 15 cm × 5 μm) and a 190 ~ 700 nm variable wavelength ultraviolet-visible detector using mobile phases in Table S1. A Bruker Avance ( Bruker, Germany) was operated at a frequency of 500 MHz for the collection of 1H NMR spectra using tetramethylsilane as an internal standard. A Hewlett-Packard 5989A mass spectrometer (Agilent, America) was operated on the ESI mode for recording mass spectra.

Microdroplet synthesis of ketoximes from benzophenone with basic catalyst For the microdroplet synthesis of ketoximes, benzophenone (0.02 mol L-1) in the methanol and hydroxylamine solution (5.0 mol L-1) in 2.5 mol L-1 NaOH were individually loaded into two airtight glass syringe. The syringe pump ran at 10 μL min-1 to deliver the solutions to the 150 μm i.d. capillary, and the capillary was inserted into a spray emitter using dry nitrogen gas at 100 psi (Figure 1A). The microdroplets were introduced into an electrically heated single pass chamber by the sheathing gas with temperature monitoring by a thermocouple thermometer. The merged plumes from two sprayers flowed through a glass surface for product collection. After 10 min collection (200 μL collected volume), 2 mL of a water-methanol solution (v:v = 1:1) was used to dissolve the product. After filtration, the sample was diluted by ten-fold and then analyzed by HPLC and the yield was calculated based on peak area.

Microdroplet synthesis of amides from ketoximes with catalysts For the synthesis of amides catalyzed by TsCl, an inner capillary of 50 μm i.d. and 150 μm o.d. was surrounded by a concentric outer capillary of 200 μm i.d. with the inner capillary tip keeping at 0.1 mm outside the outer capillary). The outer capillary was then inserted into a spray emitter with a stream of dry nitrogen gas at 100 psi. The syringe pump fed two solutions of ketoxime (0.02 mol L-1) in acetonitrile and TsCl + ZnCl2 (each 0.02 mol 13

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

Page 14 of 20

L-1) to the inner and outer capillaries individually at 10 μL min-1 (Figure 1B). The microdroplets were captured by 10 mL methanol in a 25-mL glass beaker. After 10 min collection (200 μL collected volume), the product was determined by HPLC after ten-fold dilution with methanol.

Bulk reaction synthesis of ketoximes and amides Ketoximes: A 100 μL benzophenone solution (0.02 mol L-1) in methanol and a 100 μL hydroxylamine solution (5.0 mol L-1) in 2.5 mol L-1 NaOH were mixed and allowed for 10 min reaction at 40 oC. Upon completion of the reaction, 2 mL of a water-methanol solution (v:v = 1:1) was used to dissolve the product. After filtration, the sample was diluted by ten-fold and then analyzed by HPLC and the yield was calculated based on peak area.

Amides: A mixture solution of ketoxime (0.02 mol L-1) and TsCl + ZnCl2 (each 0.02 mol L-1) in the acetonitrile (200 μL) was allowed for 10 min reaction at 70 oC. The product was measured by HPLC after twenty-fold dilution with methanol.

Online monitoring of microdroplet synthesis by mass spectrometry A Thermo Fisher LTQ-Orbitrap mass spectrometer (Thermo Scientific, San Jose, CA, USA) was used for mechanism confirmation. Two sprayers for microdroplet generation in Figure 1 were directly placed at 5 cm away from the inlet capillary of the mass spectrometer with high voltages applied upon the reactant solutions (ketone and ketoxime). Considering high sensitivity of mass spectrometry, 0.2 mmol L-1 reactants (benzophenone and benzophenone oxime), 0.2 mmol L-1 catalysts (TsCl and ZnCl2) and other reagents (0.5 mol L-1 hydroxylamine and 0.1 mol L-1 NaOH) were used to avoid interface contamination to the mass spectrometer. The mass spectrometer worked in the positive mode using the following parameters: spray voltage, +3.5 kV; capillary temperature, 275 °C; capillary voltage, 30 V; tube lens, 100 V; resolution, 30000. Mass spectra of product ions upon CID were collected with a 1 m/z isolation window and 26-35 eV collision energies applied to selected precursor ions.

Benzophenone oxime: (white solid, 12.83 mg, 90.4% yield); 1H NMR (500 MHz, CDCl3) δ 9.23 (s, 1H), 7.40-7.34 (m, 6H), 7.28-7.22 (m, 3H). 4,4'-Dimethylbenzophenone oxime: (white solid, 14.3 mg, 88.2% yield); 1H NMR (500 MHz, CDCl3) δ 8.73 (s, 1H), 7.31-7.15 (m, 6H), 7.05 (d, J = 7.5 Hz, 2H), 2.33 (s, 3H), 2.27 (s, 3H). 4,4'-Difluorobenzophenone oxime: (white solid, 14.6 mg, 87.2% yield); 1H NMR (500 MHz, CDCl3) δ 8.46 (s, 1H), 7.38-7.32 (m, 4H), 7.11-7.04 (m, 2H), 6.99-6.90 (m, 2H). 4,4'-Dichlorobenzophenone

oxime: (white solid, 17.5 mg, 91.1% yield); 1H NMR (500 MHz, CDCl3) δ 8.19 (s, 1H), 7.38-7.37 (m, 2H), 7.32-7.30 (m, 2H), 7.27-7.23 (m, 4H). 4,4'-Dibromobenzophenone oxime: (brown solid, 23.8 mg, 93.4% 14

ACS Paragon Plus Environment

Page 15 of 20 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

yield); 1H NMR (500 MHz, CDCl3) δ 7.77 (s, 1H), 7.55-7.52 (m, 2H), 7.42-7.39 (m, 2H), 7.26-7.24 (m, 2H), 7.20-7.18 (m, 2H). 4,4'-Dimethoxybenzophenoneoxime: (white solid, 17.3 mg, 93.5% yield); 1H NMR (500 MHz, CDCl3) δ 7.33 (d, J = 8.5 Hz, 4H), 6.90 (d, J = 8.5 Hz, 2H), 6.81-6.72 (d, J = 9.0 Hz, 2H), 3.78 (s, 3H), 3.74 (s, 3H). 4-Methylbenzophenoneoxime: (white solid, 12.3 mg, 81.2% ); 1H NMR (500 MHz, CDCl3) δ 2.28 (s, 3H), 2.34 (s, 3H), 7.05 (s, 3H), 7.06 (s, 3H), 7.19-7.40 (m, 18H), 8.50 (s, 2H).

4-Fluoro-4'-methoxybenzophenoneoxime: (white solid, 16.8 mg, 95.3% ); 1H NMR (500 MHz, CDCl3) δ 3.73 (s, 3H), 3.78 (s, 3H), 6.77-7.36 (m, 18H), 8.93 (s, 2H)

Benzoylanilide: (white solid, 13.0 mg, 91.3% yield); 1H NMR (500 MHz, CDCl3) δ 7.87 (s, 1H), 7.78 (d, J = 7.0 Hz, 2H), 7.7 (d, J = 8.0 Hz,, 2H), 7.47 (t, J = 7.5 Hz, 1H), 7.39 (t, J = 7.5 Hz, 2H), 7.29 (t, J = 8.0 Hz, 2H), 7.08 (t, J = 7.5 Hz, 1H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C13H12NO, 198.0913; found, 198.0912.

4,4'-Dimethylbenzanilide: (white solid, 14.2 mg, 87.8% yield); 1H NMR (500 MHz, CDCl3) δ 7.69 (d, J = 8.0 Hz, 2H), 7.44 (d, J = 8.5 Hz, 2H), 7.21-7.19 (m, 2H), 7.09 (d, J = 8.5 Hz, 2H), 2.35 (s, 3H), 2.26 (s, 3H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C15H16NO, 226.1226; found, 226.1239. 4,4'-Difluorobenzanilide: (white solid, 13.2 mg, 78.7% yield); 1H NMR (500 MHz, CDCl3) δ 7.82-7.79 (m, 2H), 7.68( s, 1H), 7.52-7.49 (m, 2H), 7.11 (d, J = 8.5 Hz, 2H), 7.01 (d, J = 9.0 Hz, 2H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C13H10NOF2, 234.0725; found, 234.0730. 4,4'-Dichlorobenzanilide: (white solid, 16.4 mg, 85.6% yield); 1H NMR (500

MHz, CDCl3) δ 8.63 (s, 1H), 7.38-7.36 (m, 2H), 7.31-7.23 (m, 6H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C13H10NOCl2, 266.0134; found, 266.0135. 4,4'-Dibromobenzanilide: (brown solid, 21.3 mg, 83.1% yield); 1H

NMR (500 MHz, CDCl3) δ 8.18(s, 1H), 7.54 (d, J = 8.0 Hz, 2H), 7.40 (d, J = 8.5 Hz, 2H), 7.25-7.19 (m, 4H);

HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C13H10NOBr2, 353.9124; found, 355.9102.

4,4'-Dimethoxybenzanilide: (white solid, 15.0 mg, 81.1% yield); 1H NMR (500 MHz, CDCl3) δ 7.76 (d, J = 8.5 Hz, 2H), 7.59 (s, 1H), 7.45 (d, J = 9.0 Hz, 2H), 6.90 (d, J = 9.0 Hz, 2H), 6.84(d, J = 9.0 Hz, 2H), 3.80 (s, 3H), 3.74 (s, 3H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C15H16NO3, 258.1125; found, 258.1123.

4-Fluoro-4'-methoxybenzanilide. (white solid, 16.4 mg, 93.0% yield); 1H NMR (500 MHz, CDCl3) δ 7.81-7.75 (m, 5H), 7.70 (s, 1H), 7.65 (s, 1H), 7.52-7.43 (m, 4H), 7.09-6.82 (m, 9H), 3.80 (s, 3H), 3.74 (s, 3H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C14H12FNO2, 246.0925; found, 246.0929. 4-Methylbenzanilide. (white solid, 11.8 mg, 77.6% yield); 1H NMR (500 MHz, CDCl3) δ 7.92-7.94 (m, 2H), 7.76 (d, J = 7.5 Hz, 2H), 7.76 (d, J = 7.5 Hz, 2H), 7.67 (d, J = 8 Hz, 2H), 7.56 (d, J = 8 Hz, 2H), 7.44-7.41 (m, 3H), 7.36-7.33 (m, 15

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

Page 16 of 20

2H),7.27-7.24 (m, 2H), 7.17-7.14 (m, 2H), 7.06-7.03 (m, 3H), 2.25 (s, 3H), 2.32 (s, 3H); HRMS (ESI-TOF) m/z: [M+H]+ Calcd. for C14H13NO, 212.1082; found, 212.1076.

Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: XXXXXXXX. HPLC mobile phases, HPLC chromatograms of oximes and amides under different conditions, 1H NMR spectra and high-resolution mass spectra of oximes and amides and mechanism of Beckmann rearrangement of ketoxime.

Acknowledgements The authors acknowledge the financial support from the National Natural Science Foundation of China under project Nos. 21675037 and 21405030, and Natural Science Foundation of Zhejiang Province under project No. LGN19B050001. We also thank Dr. Jiang in our institute for his kind help in the collection of mass spectra. Reference: (1) Crochet, P.; Cadierno, V. Catalytic synthesis of amides via aldoximes rearrangement. Chem. Commun. 2015, 51, 2495. (2) Montalbetti, C. A. G. N.; Falque, V. Amide bond formation and peptide coupling. Tetrahedron 2005, 61, 10827. (3) Valeur, E.; Bradley, M. Amide bond formation: beyond the myth of coupling reagents. Chem. Soc. Rev. 2009, 38, 606. (4) Constable, D. J. C.; Dunn, P. J.; Hayler, J. D.; Humphrey, G. R.; Leazer, J. J. L.; Linderman, R. J.; Lorenz, K.; Manley, J.; Pearlman, B. A.; Wells, A.; Zaks, A.; Zhang, T. Y. Key green chemistry research areas-a perspective from pharmaceutical manufacturers. Green Chem. 2007, 9, 411. (5) Carey, J. S.; Laffan, D.; Thomson, C.; Williams, M. T. Analysis of the reactions used for the preparation of drug candidate molecules. Org. Biomol. Chem. 2006, 4, 2337. (6) Chitra, S.; Vishal, K.; Upendra, S.; Neeraj, K.; Bikram, S. Emerging Catalytic Methods for Amide Synthesis&dagger. Curr. Org. Synth. 2013, 10, 241. (7) Blatt, A. H. The Beckmann Rearrangement. Chem. Rev. 1933, 12, 215. (8) Aricò, F.; Quartarone, G.; Rancan, E.; Ronchin, L.; Tundo, P.; Vavasori, A. One-pot oximation–Beckmann rearrangement of ketones and aldehydes to amides of industrial interest: Acetanilide, caprolactam and acetaminophen. Catal. Commun. 2014, 49, 47. 16

ACS Paragon Plus Environment

Page 17 of 20 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

(9) Chu, Y.; Ji, P.; Yi, X.; Li, S.; Wu, P.; Zheng, A.; Deng, F. Strong or weak acid, which is more efficient for Beckmann rearrangement reaction over solid acid catalysts? Catal.Sci. Technol. 2015, 5, 3675. (10) Gregory, B. J.; Moodie, R. B.; Schofield, K. Kinetics and mechanism of the Beckmann rearrangement of acetophenone oximes in sulphuric acid. J. Chem. Soc. B 1970, 338. (11) C., M. N.; L., R.; C., T.; O., T.; R., B. On the Beckmann rearrangement of ketoximes: A kinetic study in trifluoromethanesulfonic acid. Int. J. Chem. Kinet. 2004, 36, 417. (12) Wang, B.; Gu, Y.; Luo, C.; Yang, T.; Yang, L.; Suo, J. Sulfamic acid as a cost-effective and recyclable catalyst for liquid Beckmann rearrangement, a green process to produce amides from ketoximes without waste. Tetrahedron Lett. 2004, 45, 3369. (13) Ghiaci, M.; Abbaspur, A.; Kalbasi, R. J. Vapor-phase Beckmann rearrangement of cyclohexanone oxime over H3PO4/ZrO2–TiO2. Appl. Catal. A-Gen.2005, 287, 83. (14) Ikushima, Y.; Hatakeda, K.; Sato, M.; Sato, O.; Arai, M. Innovation in a chemical reaction process using a supercritical water microreaction system: environmentally friendly production of ε-caprolactam. Chem. Commun. 2002, 2208. (15) Kore, R.; Srivastava, R. A simple, eco-friendly, and recyclable bi-functional acidic ionic liquid catalysts for Beckmann rearrangement. J. Mol. Catal. A-Chem. 2013, 376, 90. (16) Arisawa, M.; Yamaguchi, M. Rhodium-Catalyzed Beckmann Rearrangement. Org. Lett. 2001, 3, 311. (17) De, S. K. Facile Becmann rearrangement of ketoximes mediated by yttrium triflate. . Org. Prep. Proced. Int. 2004, 36, 383. (18) Yan, P.; Batamack, P.; Surya Prakash, G. K.; Olah, G. A. Gallium (III) Triflate Catalyzed Beckmann Rearrangement. Catal. Lett. 2005, 103, 165. (19) Mahajan, S.; Sharma, B.; Kapoor, K. K. A solvent-free one step conversion of ketones to amides via Beckmann rearrangement catalysed by FeCl3·6H2O in presence of hydroxylamine hydrochloride. Tetrahedron Lett. 2015, 56, 1915. (20) Liu, L.-F.; Liu, H.; Pi, H.-J.; Yang, S.; Yao, M.; Du, W.; Deng, W.-P. Facile AlCl3-Promoted Catalytic Beckmann Rearrangement of Ketoximes. Synthetic Commun. 2011, 41, 553. (21) Ramalingan, C.; Park, Y.-T. Mercury-Catalyzed Rearrangement of Ketoximes into Amides and Lactams in Acetonitrile. J. Org. Chem. 2007, 72, 4536. (22) Pi, H.-J.; Dong, J.-D.; An, N.; Du, W.; Deng, W.-P. Unexpected results from the re-investigation of the Beckmann rearrangement of ketoximes into amides by using TsCl. Tetrahedron 2009, 65, 7790. 17

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

Page 18 of 20

(23) An, N.; Tian, B.-X.; Pi, H.-J.; Eriksson, L. A.; Deng, W.-P. Mechanistic Insight into Self-Propagation of Organo-Mediated Beckmann Rearrangement: A Combined Experimental and Computational Study. J. Org. Chem. 2013, 78, 4297. (24) Li, J.-T.; Meng, X.-T.; Yin, Y. Efficient Procedure for Beckmann Rearrangement of Ketoximes Catalyzed by Sulfamic Acid/Zinc Chloride. Synthetic Commun. 2010, 40, 1445. (25) Yan, X.; Bain, R. M.; Cooks, R. G. Organic Reactions in Microdroplets: Reaction Acceleration Revealed by Mass Spectrometry. Angew. Chem. Int. Edit. 2016, 55, 12960. (26) Cooks, R. G.; Yan, X. Mass Spectrometry for Synthesis and Analysis. Ann. Rev. Anal. Chem. 2018, 11, 1. (27) Ioannou, G. I.; Montagnon, T.; Kalaitzakis, D.; Pergantis, S. A.; Vassilikogiannakis, G. Synthesis of cyclopent-2-enones from furans using a nebulizer-based continuous flow photoreactor. Org. Biomol. Chem. 2017, 15, 10151. (28) Ioannou, G. I.; Montagnon, T.; Kalaitzakis, D.; Pergantis, S. A.; Vassilikogiannakis, G. One-Pot Synthesis of Diverse γ -Lactam Scaffolds Facilitated by a Nebulizer-Based Continuous Flow Photoreactor. ChemPhotoChem 2018, 2, 860. (29) Yan, X.; Lai, Y.-H.; Zare, R. N. Preparative microdroplet synthesis of carboxylic acids from aerobic oxidation of aldehydes. Chem.Sci. 2018, 9, 5207. (30) Gensler, W. J. The Synthesis of Isoquinolines by the Pomeranz-Fritsch Reaction. Org. React. 2011. (31) Banerjee, S.; Zare Richard, N. Syntheses of Isoquinoline and Substituted Quinolines in Charged Microdroplets. Angew. Chem. Int. Edit. 2015, 54, 14795. (32) Poliakoff, M.; Anastas, P. A principled stance. Nature 2001, 413, 257. (33) DeSimone, J. M. Practical Approaches to Green Solvents. Science 2002, 297, 799. (34) Vanos, C. M.; Lambert, T. H. Cyclopropenium-activated Beckmann rearrangement. Catalysis versus self-propagation in reported organocatalytic Beckmann rearrangements. Chem.Sci. 2010, 1, 705. (35) Thopate, S. R.; Kote, S. R.; Rohokale, S. V.; Thorat, N. M. Citric acid catalysed Beckmann rearrangement, under solvent free conditions. J. Chem. Res. 2011, 35, 124. (36) Wleklinski, M.; Loren, B. P.; Ferreira, C. R.; Jaman, Z.; Avramova, L.; Sobreira, T. J. P.; Thompson, D. H.; Cooks, R. G. High throughput reaction screening using desorption electrospray ionization mass spectrometry. Chem.Sci. 2018, 9, 1647. (37) Müller, T.; Badu-Tawiah, A.; Cooks, R. G. Accelerated CarbonCarbon Bond-Forming Reactions in 18

ACS Paragon Plus Environment

Page 19 of 20 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

Preparative Electrospray. Angew. Chem. Int. Edit. 2012, 51, 11832. (38) Yan, X.; Cheng, H.; Zare Richard, N. Two‐Phase Reactions in Microdroplets without the Use of Phase‐Transfer Catalysts. Angew. Chem. Int. Edit. 2017, 56, 3562. (39) Dong Hao, H. M. Recent Progress in Synthesis of Amides. Chin. J. Org. Chem. 2017, 37, 267.

19

ACS Paragon Plus Environment

The Journal of Organic Chemistry

Page 20 of 20

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 20

ACS Paragon Plus Environment