Stereoselective synthesis of vic-halohydrins and an

2 downloads 0 Views 2MB Size Report
Jul 20, 2015 - Knoevenagel product from an organocatalyzed aldol reaction: A non-enamine ... The condensation reaction of carbonyl compounds with an.
2015

2015年 第36卷 第7期

CHINESE JOURNAL OF CATALYSIS

Vol. 36 No. 7

In This Issue 封面: 朱秋桦等采用溶胶凝胶法制备了 La0.6Sr0.4NixCo1-xO3 钙钛矿 型催化剂, 该催化剂在焦炉煤气干重整反应中生成了活性金属 Ni、Co 颗 粒和 La2O2CO3, 这些新的物相对催化剂的活性、稳定性和抗积碳能力起关 键性的作用. 见本期第 915–924 页. Cover: Zhu and co-workers in their Article on pages 915–924 reported a series of La0.6Sr0.4NixCo1-xO3 perovskite-type catalysts synthesized by sol-gel method, and used for producing syngas from coke oven gas. After reaction, the catalysts show the formation of Ni0, Co0 and La2O2CO3, the new phase play a crucial role on the catalyst activity, stability and resistance to carbon deposition.

About the Journal Chinese Journal of Catalysis is an international journal published monthly by Chinese Chemical Society, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, and Elsevier. The journal publishes original, rigorous, and scholarly contributions in the fields of heterogeneous and homogeneous catalysis in English or in both English and Chinese. The scope of the journal includes:  New trends in catalysis for applications in energy production, environmental protection, and production of new materials, petroleum chemicals, and fine chemicals;  Scientific foundation for the preparation and activation of catalysts of commercial interest or their representative models;  Spectroscopic methods for structural characterization, especially methods for in situ characterization;  New theoretical methods of potential practical interest and impact in the science and applications of catalysis and catalytic reaction;  Relationship between homogeneous and heterogeneous catalysis;  Theoretical studies on the structure and reactivity of catalysts.  The journal also accepts contributions dealing with photo-catalysis, bio-catalysis, and surface science and chemical kinetics issues related to catalysis. Types of Contributions

Impact Factor

 Reviews deal with topics of current interest in the areas covered by this journal. Re-

2013 SCI Impact Factor: 1.552 2013 SCI 5-Year Impact Factor: 1.180 2013 ISTIC Impact Factor: 1.139



 





views are surveys, with entire, systematic, and important information, of recent progress in important topics of catalysis. Rather than an assemblage of detailed information or a complete literature survey, a critically selected treatment of the material is desired. Unsolved problems and possible developments should also be discussed. Authors should have published articles in the field. Reviews should have more than 80 references. Communications rapidly report studies with significant innovation and major academic value. They are limited to four Journal pages. After publication, their full-text papers can also be submitted to this or other journals. Articles are original full-text reports on innovative, systematic and completed research on catalysis. Highlights describe and comment on very important new results in the original research of a third person with a view to highlight their significance. The results should be presented clearly and concisely without the comprehensive details required for an original article. Perspectives are short reviews of recent developments in an established or developing topical field. The authors should offer a critical assessment of the trend of the field, rather than a summary of literatures. Viewpoints describe the results of original research in general in some area, with a view to highlighting the progress, analyzing the major problems, and commenting the possible research target and direction in the future.

Abstracting and Indexing Abstract Journals (VINITI) Cambridge Scientific Abstracts (CIG) Catalysts & Catalysed Reactions (RSC) Current Contents/Engineering, Computing and Technology (Thomson ISI) Chemical Abstract Service/SciFinder (CAS) Chemistry Citation Index (Thomson ISI) Japan Information Center of Science and Technology Journal Citation Reports/Science Edition (Thomson ISI) Science Citation Index Expanded (Thomson ISI) SCOPUS (Elsevier) Web of Science (Thomson ISI)

2015年 第36卷 第7期

2015 Vol. 36 No. 7

CHINESE JOURNAL OF CATALYSIS 《催化学报》第五届编辑委员会

月刊 SCI 收录 1980 年 3 月创刊 中国化学会催化学会会刊 2015年7月20日出版

The Fifth Editorial Board of Chinese Journal of Catalysis 顾问 (Advisors)

主管 中国科学院 主办 中国化学会 中国科学院大连化学物理研究所 主编 李 灿 张 涛 编辑 《催化学报》编辑委员会 出版

国内统一连续出版物号 CN 21-1195/O6 国际标准连续出版物号 ISSN 0253-9837 CODEN THHPD3 广告经营许可证号 2013003

总发行 北京东黄城根北街 16 号 邮编: 100717 电话: (010) 64017032 E-mail: [email protected] 国内订购 全国各地邮政局 邮发代号 8-93 国外订购 中国国际图书贸易总公司 北京 399 信箱 邮编 100044 国外发行代号 M417 印刷 大连海大印刷有限公司 定价 50 元

Publication Monthly (12 issues) Started in March 1980 Transaction of the Catalysis Society of China Superintended by Chinese Academy of Sciences Sponsored by Chinese Chemical Society and Dalian Institute of Chemical Physics of CAS Editors-in-Chief Can Li, Tao Zhang Edited by Editorial Board of Chinese Journal of Catalysis Published by Science Press

Distributed by Science Press 16 Donghuangchenggen North Street, Beijing 100717, China Tel: +86-10-64017032 E-mail: [email protected] Subscription Agents Domestic All Local Post Offices in China Foreign China International Book Trading Corporation, P.O.Box 399, Beijing 100044, China Printed by Dalian Haida Printing Company, Limited Price $50

公 开 发 行

Alexis T. Bell (美国) Jürgen Caro (德国) Gabriele Centi (意大利) Michel Che (法国) 陈 懿 (Yi Chen) Avelino Corma (西班牙) 高 滋 (Zi Gao)

Masatake Haruta (日本) 何鸣元 (Mingyuan He) Graham J. Hutchings (英国) Johannes A. Lercher (德国) 闵恩泽 (Enze Min) S. Ted. Oyama (日本) Daniel E. Resasco (美国)

Rutger A. van Santen (荷兰) Ferdi Schüth (德国) 万惠霖 (Huilin Wan) 谢有畅 (Youchang Xie) 辛 勤 (Qin Xin) 郑小明 (Xiaoming Zheng)

荣誉主编 (Honorary Editor-in-Chief) 林励吾 (Liwu Lin) 主编 (Editors-in-Chief) 李

灿 (Can Li)



涛 (Tao Zhang)

副主编 (Associate Editors-in-Chief) 李兴伟 (Xingwei Li) 刘海超 (Haichao Liu)

Roel Prins (瑞士) 唐军旺 (Junwang Tang, 英国)

吴 鹏 (Peng Wu) 杨启华 (Qihua Yang)

刘昌俊 (Changjun Liu) 刘中民 (Zhongmin Liu) 卢冠忠 (Guanzhong Lu) Marcel Schlaf (加拿大) Susannah L. Scott (美国) 沈俭一 (Jianyi Shen) 申文杰 (Wenjie Shen) 宋春山 (Chunshan Song, 美国) 苏宝连 (Baolian Su, 比利时) 苏党生 (Dangsheng Su) 田志坚 (Zhijian Tian) 万 颖 (Ying Wan) 王爱琴 (Aiqin Wang) 王德峥 (Dezheng Wang) 王 峰 (Feng Wang) 王建国 (Jianguo Wang) 王 野 (Ye Wang) Yong Wang (美国)

魏迎旭 (Yingxu Wei) 吴自力 (Zili Wu, 美国) 夏春谷 (Chungu Xia) 肖丰收 (Fengshou Xiao) 肖建良 (Jianliang Xiao, 英国) 谢在库 (Zaiku Xie) 徐柏庆 (Boqing Xu) 徐 杰 (Jie Xu) 徐龙伢 (Longya Xu) 严玉山 (Yushan Yan, 美国) 杨为民 (Weimin Yang) 杨维慎 (Weishen Yang) 尹双凤 (Shuangfeng Yin) 余家国 (Jiaguo Yu) 袁友珠 (Youzhu Yuan) 张宗超 (Zongchao Zhang) 赵 震 (Zhen Zhao) 周永贵 (Yonggui Zhou)

编委 (Members) 包信和 (Xinhe Bao) 曹 勇 (Yong Cao) 陈 德 (De Chen, 挪威) 陈经广 (Jingguang G. Chen,美国) 丁维平 (Weiping Ding) 丁云杰 (Yunjie Ding) 关乃佳 (Naijia Guan) 郭新闻 (Xinwen Guo) 韩洪宪 (Hongxian Han) 贺鹤勇 (Heyong He) 贺 泓 (Hong He) Emiel J. M. Hensen (荷兰) George W. Huber (美国) 景欢旺 (Huanwang Jing) Alexander Katz (美国) 李 隽 (Jun Li) 李微雪 (Weixue Li) 李永丹 (Yongdan Li)

编辑部联系方式 (Editorial Office Address) 地址: 大连市中山路 457 号 中国科学院大连化学物理研究所 邮编: 116023 电话: (0411)84379240 传真: (0411)84379543 电子信箱: [email protected]

Add.: Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, Liaoning, China Tel.: +86-411-84379240 Fax: +86-411-84379543 E-mail: [email protected]

中文主页 http://www.chxb.cn 国际版主页 http://www.elsevier.com/locate/chnjc 国际版全文 http://www.sciencedirect.com/science/journal/18722067

(CUIHUA XUEBAO)

CHINESE JOURNAL OF CATALYSIS 中国科学院科学出版基金资助出版

月刊

SCI 收录

2015 年 7 月 第 36 卷 第 7 期

目 论



915 (英/封面) 焦炉煤气二氧化碳重整制合成气中 La0.6Sr0.4NixCo1-xO3 钙钛 矿催化剂的合成、表征和催化性能研究 朱秋桦, 程红伟, 邹星礼, 鲁雄刚, 许茜, 周忠福 925 (英) 三价铑催化硝酮与炔的碳氢活化偶联合成二氢吲哚 孔令恒, 谢芳, 于松杰, 戚自松, 李兴伟 933 (英) 无铬 Co-Cu/SBA-15 催化剂催化生物质衍生 α-, β-不饱和醛 加氢制醇 Sanjay Srivastava, Pravakar Mohanty, Jigisha K. Parikh, Ajay K. Dalai, S. S. Amritphale, Anup K. Khare 943 (英) 聚苯胺对钯催化甲酸电氧化反应的促进作用 马先斌, 冯媛媛, 李扬, 韩运石, 鹿国萍, 杨海芳, 孔德生 952 (英) 催化湿式共氧化法同时去除硝基苯和苯酚 付冬梅, 章飞芳, 王联芝, 杨帆, 梁鑫淼 957 (英) Ru/C 催化剂上香豆素加氢制八氢香豆素 Dana Bílková, Petr Jansa, Iva Paterová, Libor Červený 961 (英) 超顺磁性氧化石墨烯复合材料固定辣根过氧化物酶催化去 除氯酚 常青, 江国栋, 唐和清, 李娜, 黄佳, 吴来燕 969 (英) Bi 掺杂 NaTaO3 中 Bi 的化学价态对其光催化性能的影响 崔华楠, 石建英, 刘鸿 975 (英) Ti/α-PbO2/β-PbO2 电极电化学降解 2-氯酚 张钱丽, 郭新艳, 曹晓丹, 王东田, 魏杰 982 (英) 双子型 Brönsted 酸性离子液体催化高级脂肪酸与醇反应合 成生物柴油 常涛, 何乐芹, 张晓婧, 袁明霞, 秦身钧, 赵继全 987 (英) 异质结构 AgCl/Bi2WO6 微米球制备、表征及其光催化性能 李家德, 余长林, 方稳, 朱丽华, 周晚琴, 樊启哲

次 994 (英) Ni 对 Pd/Al2O3 密偶催化剂催化性能的影响 方瑞梅, 崔亚娟, 史忠华, 龚茂初, 陈耀强 1001 (英) Al3+ 离子介入提升 (NH4)2SiF6 对 SBA-15 介孔材料的水热稳 定化作用 邹成龙, 沙观宇, 黄曜, 牛国兴, 赵东元 1009 (英) g-C3N4/rGO 杂化催化剂的简易合成及其对罗丹明 B 的光催 化降解作用 原博, 魏江霞, 胡天娇, 姚海波, 蒋振华, 方志薇, 楚增勇 1017 (英) SAPO-34 分子筛中多甲基苯分子与卤代甲烷偕甲基化反应 的密度泛函理论研究 孔令涛, 沈本贤 1023 (英) 非对映选择性一锅八组分反应高效绿色合成二螺氢喹啉 Sajjad Salahi, Malek Taher Maghsoodlou, Nourallah Hazeri, Mojtaba Lashkari, Santiago Garcia-Granda, Laura Torre-Fernandez 1029 (英) 多孔纳米结构 Fe/Pt-Fe 电极对碱性介质中甲醇电氧化的催 化活性 Javad Hosseini, Mehdi Abdolmaleki, Hamid Reza Pouretedal, Mohammad Hossein Keshavarz 1035 (英) 氨功能化介孔 MCM-41 固载锰卟啉作为萘羟基化多相催化 剂 杨福, 高树英, 熊翠蓉, 王海青, 陈进, 孔岩 1042 (英) 柴桂叶提取物介导的银纳米粒子的绿色合成及其在吡喃并 吡唑衍生物合成中的应用 Sneha Yadav, Jitender M. Khurana 1047 (英) Fe3O4/油酸担载肟衍生环钯配合物中钯粒子作为催化剂催化 无铜 Sonogashira 交联反应 Kazem Karami, Samaneh Dehghani Najvani, Nasrin Haghighat Naeini, Pablo Hervés 1054 (英) 无溶剂条件下以 10-钼-2-钒磷酸为高效、可重复使用催化剂 一锅法合成 2,4,5-三取代咪唑衍生物 Laxmikant D. Chavan, Sunil G. Shankarwar

1060 (英) V2O5/MO-Al2O3 (M = Mg, Ca, Sr, Ba) 催化剂用于正丁烷催 化氧化脱氢反应 徐兵, 朱雪峰, 曹中卫, 杨丽娜, 杨维慎

1117 (英) 自下而上法制备金-介孔二氧化硅复合纳米管用作还原4-硝 基苯酚的催化剂 彭永胜, 冷文光, 董彬, 格日乐, 段洪东, 高艳安

1068 (英) 微生物燃料电池处理酿酒厂废水过程中的产电性能及其膜 性能衰减 Afşin Y. Çetinkaya, Emre Oğuz Köroğlu, Neslihan Manav Demir, Derya Yılmaz Baysoy, Bestamin Özkaya, Mehmet Çakmakçı

1124 (英) 铋三氟甲磺酸酯: 用于一锅多组分反应合成具有生物活性的 香豆素化合物的高效催化剂 Mahmoud. Abd El Aleem. Ali. Ali. El-Remaily

1077 (英) 一步温和水热法制备具有改善光催化活性和稳定性的碳包 覆 CdS 纳米粒子 邹帅, 伏再辉, 向超, 吴文锋, 汤森培, 刘亚纯, 尹笃林 1086 (英) V@CN 催化的芳烃的氧气羟基化反应 李岩, 李冰, 陈婷, 周志成, 王军, 黄军 1093 (英) 非烯胺途径有机催化醛醇缩合反应立体选择性合成邻卤代 醇和 Knoevenagel 产物 P. B. Thorat, S. V. Goswami, V. P. Sondankar, S. R. Bhusare 1101 (英) 2-吡嗪羧酸 Ni(II), Co(II) 和 Cu(II) 配合物的合成、表征、电 化学测试及其催化合成 2H-吲哚 [2,1-b] 酞嗪三酮 Behnaz Afzalian, Joel T. Mague, Maryam Mohamadi, S. Yousef Ebrahimipour, Behjat Pour amiri, Esmat Tavakolinejad Kermani 1109 (英) 磷酸锆镍上的醇选择性氧化 Abdol R. Hajipour, Hirbod Karimi, Afshin Koohi

1131 (英) 核壳结构催化剂 Cr-Zn@SiO2@SAPO-34 催化合成气直接转 化制低碳烃的性能 李津京, 潘秀莲, 包信和 1136 (英) 低压下碱金属碳酸盐催化一步法合成碳酸二甲酯 刘春, 张绍科, 蔡宝仪, 金子林 1142 (英/中) Pd-O/CeO2 纳米管催化苯酚氧化羰基化反应 袁烨, 王志苗, 安华良, 薛伟, 王延吉

相关信息 1155

作者索引

英文全文电子版(国际版)由Elsevier出版社在ScienceDirect上出版 http://www.sciencedirect.com/science/journal/18722067 http://www.elsevier.com/locate/chnjc http://www.chxb.cn 在线投审稿网址 https://mc03.manuscriptcentral.com/chinjcatal

Chinese Journal of Catalysis Vol. 36, No. 7, July 2015

催化学报 2015年 第36卷 第7期 | www.chxb.cn

a v a i l a b l e a t w w w. s c i e n c e d i r e c t . c o m

j o u r n a l h o m e p a g e : w w w . e l s e v i e r. c o m / l o c a t e / c h n j c

Chinese Journal of Catalysis Graphical Contents

Articles Chin. J. Catal., 2015, 36: 915–924

doi: 10.1016/S1872-2067(14)60303-X

Synthesis, characterization, and catalytic performance of La0.6Sr0.4NixCo1−xO3 perovskite catalysts in dry reforming of coke oven gas Qiuhua Zhu, Hongwei Cheng *, Xingli Zou, Xionggang Lu *, Qian Xu, Zhongfu Zhou Shanghai University, China; Aberystwyth University, United Kingdom

CO2 is adsorbed on a La2O3-SrO, matrix producing SrCO3 and La2O2CO3, the active species for CO2 reforming of COG; these species inhibit carbon formation by reacting with CH4 to reform CO and H2. The reaction of carbonate species with surface carbon species occurs at the metal sites, converting the carbon species to CO. COG containing abundant H2 can also reduce the formation of carbon deposits by inhibiting decomposition of CH4. Chin. J. Catal., 2015, 36: 925–932

doi: 10.1016/S1872-2067(15)60866-X

Rh(III)-catalyzed coupling of nitrones with alkynes for the synthesis of indolines Lingheng Kong, Fang Xie, Songjie Yu, Zisong Qi, Xingwei Li * Dalian Institute of Chemical Physics, Chinese Academy of Sciences

Rh-catalyzed redox-neutral coupling between N-aryl nitrones and alkynes has been achieved under relatively mild conditions. The reaction proceeded via C-H activation at the N-aryl ring with subsequent O-atom transfer, affording trisubstituted indolines in good chemoselectivity and moderate to good diasteroselectivity.

vi

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 933–942

doi: 10.1016/S1872-2067(15)60870-1

Cr-free Co–Cu/SBA-15 catalysts for hydrogenation of biomass-derived α-, β-unsaturated aldehyde to alcohol Sanjay Srivastava, Pravakar Mohanty, Jigisha K. Parikh *, Ajay K. Dalai, S. S. Amritphale, Anup K. Khare S. V. National Institute of Technology, India; Indian Institute of Technology, India; Sardar Patel Renewable Energy Research Institute (SPRERI), India; University of Saskatchewan, Saskatoon, Saskatchewan, Canada; Council of Scientific & Industrial Research (CSIR), India

O

Methyl f uran (MF)

Conversion or selectivity (%)

Hydrogenation, 170 °C, 20 bar 100 80 60

FFR FOL MF CPL

40 20 0

150

160 170 180 o Temperature ( C)

190

This work presents the hydrogenation of furfural in liquid phase, as a direct step for bio-oil upgradation, using bi-metallic Co–Cu catalysts supported on SBA-15 (Cu 10 wt%; Co 2.5, 5, 10 wt%). Chin. J. Catal., 2015, 36: 943–951

doi: 10.1016/S1872-2067(15)60863-4

Promoting effect of polyaniline on Pd catalysts for the formic acid electrooxidation reaction Xianbin Ma, Yuanyuan Feng *, Yang Li, Yunshi Han, Guoping Lu, Haifang Yang, Desheng Kong Qufu Normal University

HCOOH

4 0.20

PANI

Pd

3

0.10

IA (A/m2)

0.15 2

0.05

1

0.00

0

10 PA 15 NI/ PA Pd 20 NI PA /Pd 30 NI/ PA Pd N I/P d Pd Pd /C Pd /5P /C A Pd /10 NI /C PA Pd /20P NI /C A N / Pd 30P I /C A /4 NI 0P A NI Pd /C Pd / Pd C/5 /C PA Pd /10 NI /C PA Pd /20P NI /C A Pd /30 NI /C PA /4 N 0P I AN I Pd /C

MSA (A/mg)

CO2 + H2O

For both nPANI/Pd and Pd/C/nPANI catalysts, polyaniline (PANI) shows significant promoting effect on Pd catalyst toward formic acid electrooxidation reaction.

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 952–956

vii

doi: 10.1016/S1872-2067(15)60835-X

Simultaneous removal of nitrobenzene and phenol by homogenous catalytic wet air oxidation Dongmei Fu *, Feifang Zhang, Lianzhi Wang, Fan Yang, Xinmiao Liang Dalian Institute of Chemical Physics, Chinese Academy of Sciences; East China University of Science and Technology; Hubei University for Nationalities

Batch-wise addition of phenol together with the use of homogeneous catalysts significantly improves the conversion of nitrobenzene. This oxidation process should provide an alternative and effective means of treating other organic pollutants in the environment.

Chin. J. Catal., 2015, 36: 957–960

doi: 10.1016/S1872-2067(15)60860-9

Hydrogenation of coumarin to octahydrocoumarin over a Ru/C catalyst Dana Bílková *, Petr Jansa, Iva Paterová, Libor Červený Institute of Chemical Technology Prague, Czech Republic; Aroma Praha a.s., Czech Republic SIMPLIFIED SCHEME OF HETEROGENEOUS COUMARIN HYDROGENATION

COUMARIN

OCTAHYDROCOUMARIN DIHYDROCOUMARIN

!!! TOXIC !!!

!!! NON TOXIC !!!

The hydrogenation of toxic coumarin to octahydrocoumarin over a Ru/C catalyst and optimizing the reaction conditions (pressure, solvent and coumarin concentration) were investigated. Chin. J. Catal., 2015, 36: 961–968

doi: 10.1016/S1872-2067(15)60865-7

Enzymatic removal of chlorophenols using horseradish peroxidase immobilized on superparamagnetic Fe3O4/graphene oxide nanocomposites Qing Chang, Guodong Jiang *, Heqing Tang, Na Li, Jia Huang, Laiyan Wu South-Central University for Nationalities; Hubei University of Technology

100

2,4-dichlorophenol 2-chlorophenol 4-chlorophenol

80

NH3·H2O

c/c0

60

EDC HRP

40 20 0

Fe2+

Fe3+

Fe3O4

HRP

0

30

60

90 120 150 180 210 t (min)

Horseradish peroxidase (HRP) was immobilized on superparamagnetic Fe3O4/graphene oxide and used for enzymatic removal of chlorophenols. The immobilized enzyme had high stability and was recovered by magnetic separation.

viii

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 969–974

doi: 10.1016/S1872-2067(15)60858-0

Influence of Bi chemical state on the photocatalytic performance of Bi-doped NaTaO3 Huanan Cui, Jianying Shi *, Hong Liu * Sun Yat-sen University

H2 evolution (mmol)

800

Na+

V-0.02

600

Bi5+

NaTaO3

400 200

Bi3+ TaO6

0 0

1

2

3

4

Time (h)

5

Doping of Bi5+ and Bi3+ in NaTaO3 substituting for Na+promoted the separation of charge carriers, and the photocatalytic for H2 evolution was enhanced. Chin. J. Catal., 2015, 36: 975–981

doi: 10.1016/S1872-2067(15)60851-8

Facile preparation of a Ti/α-PbO2/β-PbO2 electrode for the electrochemical degradation of 2-chlorophenol Qianli Zhang, Xinyan Guo, Xiaodan Cao, Dongtian Wang, Jie Wei * Suzhou University of Science and Technology 100

 (%)

80 60 40

Ti/-PbO2/ -PbO2 Ti/-PbO2

20 0 0

50

100

150

200

250

300

Time (min)

A cauliflower-structured Ti/α-PbO2/-PbO2 anode displays high activity toward the degradation of 2-chlorophenol, with a long active service life of 68.4 years at a current density of 20 mA/cm2. Chin. J. Catal., 2015, 36: 982–986

doi: 10.1016/S1872-2067(15)60852-X

Brönsted acid surfactant-combined dicationic ionic liquids as green catalysts for biodiesel synthesis from free fatty acids and alcohols Tao Chang, Leqin He, Xiaojing Zhang, Mingxia Yuan, Shenjun Qin *, Jiquan Zhao * Hebei University of Engineering; Hebei University of Technology

FFAs Alcohols

Cooling

Biodiesel BASDILs

BASDILs Recyling During the reaction

SO3H

BASDILs

H2n+1Cn

N

HO3S

N

CnH2n+1 2 X

After the reaction

Brönsted acid surfactant-combined dicationic ILs were synthesized and gave high activity for biodiesel synthesis from free fatty acids and alcohols. The structure of ILs influenced the activity.

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 987–993

doi: 10.1016/S1872-2067(15)60849-X

Preparation, characterization and photocatalytic performance of heterostructured AgCl/ Bi2WO6 microspheres

Light

Jia-de Li, Chang-lin Yu *, Wen Fang, Li-hua Zhu, Wan-qin Zhou, Qi-zhe Fan Jiangxi University of Science and Technology

e - e- e0.17 eV

O2 •O2-

CB e-

e-

CB

0.54 eV

Bi2WO6

H 2O

The formation of AgCl/Bi2WO6 heterostructures could effectively separate its photo-generated electron (e–) and hole (h+) pairs, then increasing its photocatalytic activity.

Chin. J. Catal., 2015, 36: 994–1000

ix

AgCl 2.07 eV

2.84 eV

2.61 eV h+ h+ VB

3.01 eV h+ h+ h+ VB

•OH

doi: 10.1016/S1872-2067(15)60850-6

Promotion of a Pd/Al2O3 close-coupled catalyst by Ni Ruimei Fang, Yajuan Cui, Zhonghua Shi *, Maochu Gong, Yaoqiang Chen * Sichuan University C 3 H8 CO NOx

C 3 H8 CO NOx

CO2 H2 O N2

CO2 H2 O N2 Al2O3

Addition of Ni

PdOx NiAl2O4 NiO

Addition of Ni to a Pd/Al2O3 close-coupled catalyst improved the catalytic activity for C3H8 conversion and increased the amount of active PdOx species. The figure describes how C3H8 conversion over the aged Pd/Al2O3 and PdNi/Al2O3 close-coupled catalysts occurs. Chin. J. Catal., 2015, 36: 1001–1008 Incorporation of Al3+ ions to

SBA-15 zeolite

doi: 10.1016/S1872-2067(15)60855-5

promote the stabilization effect of (NH4)2SiF6 treatment on hydrothermal stability of mesoporous

Chenglong Zou, Guanyu Sha, Yao Huang, Guoxing Niu *, Dongyuan Zhao Fudan University

An improved (NH4)2SiF6 treatment by pre-incorporating with Al3+, which can capture F– and reduce its etching into SBA-15 framework, finally promote the stabilization effect of (NH4)2SiF6 treatment on SBA-15 zeolite.

x

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1009–1016

doi: 10.1016/S1872-2067(15)60844-0

Simple synthesis of g-C3N4/rGO hybrid catalyst for the photocatalytic degradation of rhodamine B e

Bo Yuan, Jiangxia Wei, Tianjiao Hu, Haibo Yao, Zhenhua Jiang, Zhiwei Fang, Zengyong Chu * National University of Defense Technology; Changsha University

e

e

e

H+

e

H+

e e

e−

H+

H+

e

h+ h+ h+ h+

A hybrid catalyst of g-C3N4/rGO was prepared by directly heating a mixture of melamine and graphene oxide in air, which showed superior photocatalytic activity in the degradation of RhB under acidic conditions. Chin. J. Catal., 2015, 36: 1017–1022

H+

H+

RhB

H+

H+

doi: 10.1016/S1872-2067(15)60842-7

Theoretical study on the geminal methylation of methylbenzene by halomethane over SAPO-34 molecular sieve Lingtao Kong, Benxian Shen * East China University of Science & Technology

0.1597 0.2219

0.3967

0.4041 0.2488

0.1815 0.1004 0.2080

0.2947

0.1121 0.1680

The geminal methylation of different methylbenzenes has been investigated within the framework of a HSAPO-34 catalyst using periodic density functional theory calculations. Hexamethylbenzene showed the lowest energy barriers for its geminal methylation reactions with CH3Cl and CH3Br. Chin. J. Catal., 2015, 36: 1023–1028

doi: 10.1016/S1872-2067(15)60846-4

An efficient green synthesis of dispirohydroquinolines via a diastereoselective one-pot eight-component reaction Sajjad Salahi, Malek Taher Maghsoodlou *, Nourallah Hazeri, Mojtaba Lashkari, Santiago Garcia-Granda, Laura Torre-Fernandez University of Sistan and Baluchestan, Iran; University of Oviedo-CINN, Spain

Ar O

O

Ar O

O

O O

O Ar

O

O

O

O O

O O NH2 Ar'

O

O Ar

The one-pot eight-component reaction between Meldrum’s acid, an aromatic aldehyde and an arylamine was achieved in the presence of citric acid as the catalyst. The corresponding dispirohydroquinoline was obtained in good yield with excellent diastereoselectivity. This method is a combination of the Knoevenagel and Michael reactions.

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1029–1034

xi

doi: 10.1016/S1872-2067(15)60841-5

Electrocatalytic activity of porous nanostructured Fe/Pt-Fe electrode for methanol electrooxidation in alkaline media Javad Hosseini, Mehdi Abdolmaleki, Hamid Reza Pouretedal *, Mohammad Hossein Keshavarz Malek-ashtar University of Technology, Iran; Sayyed Jamaleddin Asadabadi University, Iran

Current density (µA/cm2)

14900

A porous Fe/Pt-Fe nanostructured electrode prepared by galvanic replacement exhibits higher electrocatalytic activity towards methanol oxidation than flat Pt and smooth Fe electrodes, which can be attributed to its high surface area.

Chin. J. Catal., 2015, 36: 1035–1041

a: smooth Fe electrode (50 mV/s) b: flat Pt electrode (50 mV/s) c: porous Fe/Pt-Fe electrode (50 mV/s)

16900

12900 10900 8900 6900

c

4900 2900

b

900

a

-1100 -3100 -1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

Potential (V vs Ag/AgCl)

doi: 10.1016/S1872-2067(15)60836-1

Coordination of manganese porphyrins on amino-functionalized MCM-41 for heterogeneous catalysis of naphthalene hydroxylation Fu Yang, Shuying Gao, Cuirong Xiong, Haiqing Wang, Jin Chen, Yan Kong * Nanjing Tech University

TF20PPMnCl

+ MCM-41

Mn(TF5PP)-MCM-41

N-MCM-41

Polyhalogenated manganese porphyrins with fluorine substituents were anchored on amino-functionalized MCM-41 to produce a heterogeneous catalyst, which exhibited remarkable catalytic activity and desirable reusability in the hydroxylation of naphthalene. Chin. J. Catal., 2015, 36: 1042–1046

doi: 10.1016/S1872-2067(15)60853-1

Cinnamomum tamala leaf extract-mediated green synthesis of Ag nanoparticles and their use in pyranopyrazles synthesis Sneha Yadav, Jitender M. Khurana * University of Delhi, India O Ar

+ CN

O

N N Ph

O

NH 2

O

O

O

+ PhNHNH2 +

CN

ArCHO Ar

+ Ag NPs, H2O

O

CN N 2H 4.H 2O +

CN HN N

O

NH2

CN CN

A novel, biochemical, and eco-friendly method was developed for the synthesis of Ag nanoparticles, using an aqueous leaf extract of Cinnamomum tamala as reducing and stabilizing agents. The Ag nanoparticles efficiently catalyzed pyranopyrazole synthesis.

xii

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1047–1053

doi: 10.1016/S1872-2067(15)60837-3

Palladium particles from oxime-derived palladacycle supported on Fe3O4/oleic acid as a catalyst for the copper-free Sonogashira cross-coupling reaction

X R

Kazem Karami *, Samaneh Dehghani Najvani, Nasrin Haghighat Naeini, Pablo Hervés Isfahan University of Technology, Iran; University of Vigo, Campus Academic, Spain

C C C O O OO O O O C O C Fe3O4 O O O O O C O C O O OO C C

This review presents a Pd catalyst on the surface of Fe3O4/oleic acid for copper-free Sonogashira cross-coupling. Very small amounts of the catalyst catalyzed the Sonogashira reaction in ethanol and water-organic solvent mixtures. Chin. J. Catal., 2015, 36: 1054–1059

+ Solvent, K2CO3

0.0005 mmol

R

doi: 10.1016/S1872-2067(15)60830-0

KSF supported 10-molybdo-2-vanadophosphoric acid as an efficient and reusable catalyst for the one-pot synthesis of 2,4,5-trisubstituted imidazole derivatives under solvent-free condition Laxmikant D. Chavan, Sunil G. Shankarwar * Dr. Babasaheb Ambedkar Marathwada University, India

CHO O

20% H5PMo10V2O4 0 / KSF NH4OAC

O

R

Solvent-free 110 oC

N R

N H

Twenty percent 10-molybdo-2-vanadophosphoric acid supported on KSF clay acts as an efficient, green and reusable solid acid catalyst for one-pot synthesis of 2,4,5-trisubstituted imidazole derivatives. Chin. J. Catal., 2015, 36: 1060–1067

doi: 10.1016/S1872-2067(15)60839-7

Catalytic oxidative dehydrogenation of n-butane over V2O5/MO-Al2O3 (M = Mg, Ca, Sr, Ba) catalysts

n-C4H10 + O2

C4 (C4H8+ C4H6)

Catalyst COx

Conversion or selectivity (%)

Bing Xu, Xuefeng Zhu *, Zhongwei Cao, Lina Yang *, Weishen Yang Liaoning Shihua University; Dalian Institute of Chemical Physics, Chinese Academy of Sciences 70 60

X (C 4 H 10 ) S(C 4 = ,tot)

50 40 30 20 10 0

3 3 3 O3 3 A l 2O A l 2O Al 2 A l 2O A l 2O 5OOOOa g r a V 2O S C B M 5/ 5/ 5/ 5/ V 2O V 2O V 2O V 2O

V2O5/MgO-Al2O3 catalysts prepared by a modified sol-gel and wet impregnation method showed high catalytic activity and selectivity for the oxidative dehydrogenation of n-butane. The high dispersion of VOx species and the existence of the MgO crystalline phase were important for the good catalytic performance.

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1068–1076

xiii

doi: 10.1016/S1872-2067(15)60833-6

Electricity production by a microbial fuel cell fueled by brewery wastewater and the factors in its membrane deterioration Afşin Y. Çetinkaya *, Emre Oğuz Köroğlu, Neslihan Manav Demir, Derya Yılmaz Baysoy, Bestamin Özkaya, Mehmet Çakmakçı Yildiz Technical University, Turkey; Tampere University of Technology, Finland

Municipal Wastewater Treatment Plant

Brewery treated wastewater

MFC wastewater Anaerobic Digester

Electricity production by microbial fuel cells (MFCs) using wastewater from a brewing plant and the influence of the hydraulic retention time (HRT) and effect of chemical oxygen demand (COD) removal on MFC performance, the microbial population, surface morphology, and functional groups were investigated. Chin. J. Catal., 2015, 36: 1077–1085

doi: 10.1016/S1872-2067(15)60827-0

Mild, one-step hydrothermal synthesis of carbon-coated CdS nanoparticles with improved photocatalytic activity and stability Shuai Zou, Zaihui Fu *, Chao Xiang, Wenfeng Wu, Senpei Tang, Yachun Liu, Dulin Yin Hunan Normal University

Cd(CH3COO)2 (NH2)2CS

Cd(CH3COO)2((NH2)2CS)2 Decompose

Glucose

Polymerize

This paper describes carbon-coated CdS (CdS@C) nanoparticles, prepared via a convenient one-step hydrothermal carbonization method, that exhibit enhanced photocatalytic activity and stability for visible light-triggered oxidative degradation of methyl orange in aqueous solution. Chin. J. Catal., 2015, 36: 1086–1092

CdS nuclei

n tio n sa tio en ea nd ucl Co d n an

Soluble polymer

Aggregation

Carbonization

CdS@C

Carbon colloid and CdS nuclei

doi: 10.1016/S1872-2067(14)60516-0

Direct hydroxylation of arenes with O2 catalyzed by V@CN catalyst Yan Li, Bing Li, Ting Chen, Zhicheng Zhou, Jun Wang, Jun Huang * Nanjing Tech University

Benzenes with electron-withdrawing groups were oxygenated to the corresponding phenols in considerable yields, and the V@CN catalyst was also applicable for the hydroxylation of aromatic halides (F, Cl, and Br) with O2.

xiv

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1093–1100

doi: 10.1016/S1872-2067(14)60317-X

Stereoselective synthesis of vic-halohydrins and an unusual Knoevenagel product from an organocatalyzed aldol reaction: A non-enamine mode P. B. Thorat, S. V. Goswami, V. P. Sondankar, S. R. Bhusare * Dnyanopasak College, India

Stereoselective synthesis of vic-halohydrins and an unusual Knoevenagel product were obtained using a newly synthesized chiral organocatalyst and triethylamine. The reaction gave the products in excellent yield with high anti selectivity and enantioselectivity. Chin. J. Catal., 2015, 36: 1101–1108

doi: 10.1016/S1872-2067(14)60318-1

Ni(II), Co(II), and Cu(II) complexes incorporating 2-pyrazinecarboxylic acid: Synthesis, characterization, electrochemical evaluation, and catalytic activity for the synthesis of 2H-indazolo[2,1-b]phthalazine-triones Behnaz Afzalian, Joel T. Mague, Maryam Mohamadi, S. Yousef Ebrahimipour *, Behjat Pour amiri, Esmat Tavakolinejad Kermani Payam Noor University, Iran; Tulane University, USA; Shahid Bahonar University of Kerman, Iran Ni(II), Co(II), Cu(II)

Three complexes containing pyrazine carboxylic acid, with the formulas [Ni(pzca)2(H2O)2], [Co(pzca)2(H2O)2], and [Cu(pzca)2(H2O)2], have been synthesized and fully characterized. These complexes were also evaluated as catalysts for the efficient one-pot four-component synthesis of 2H-indazolo[2,1-b]phthalazine-triones. Chin. J. Catal., 2015, 36: 1109–1116

doi: 10.1016/S1872-2067(14)60315-6

Selective oxidation of alcohols over nickel zirconium phosphate Abdol R. Hajipour *, Hirbod Karimi, Afshin Koohi Isfahan University of Technology, Iran; University of Wisconsin, USA; Islamic Azad University, Iran

Nickel zirconium phosphate was prepared and used as an efficient catalyst for the selective oxidation of various alcohols to the corresponding aldehydes and ketones, in good yields and with excellent selectivity.

(1) Washed with ethanol and water; (2) Dried at 110 oC; (3) Activated at 450 oC.

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1117–1123

xv

doi: 10.1016/S1872-2067(14)60310-7

Bottom-up preparation of gold nanoparticle-mesoporous silica composite nanotubes as a catalyst for the reduction of 4-nitrophenol Yongsheng Peng, Wenguang Leng, Bin Dong, Rile Ge, Hongdong Duan *, Yan’an Gao * Qilu University of Technology; Dalian Institute of Chemical Physics, Chinese Academy of Sciences

Au NPs

PDDA

TEOS

Calcination

C18TMS

ONW-Au composites

PDDA-modified ONWs

ONWs

ONW-Au-SiO2 composites

Au-mSiO2 composite nanotubes

Composite nanotubes consisting of a mesoporous silica shell with gold nanoparticles anchored to their inner wall were prepared by a bottom-up approach. This composite exhibited high efficiency as a catalyst in the reduction of 4-nitrophenol. Chin. J. Catal., 2015, 36: 1124–1130

doi: 10.1016/S1872-2067(14)60308-9

Bismuth triflate: A highly efficient catalyst for the synthesis of bio-active coumarin compounds via one-pot multi-component reaction Mahmoud. Abd El Aleem. Ali. Ali. El-Remaily * Sohag University- 82524, Sohag, Egypt O O

CHO

O

+

DCM/ 5 mol% Bi(OTf)3 OR'

2

OH 1

+

50 oC, 15-30 min

O

O

R

5

R'= CH 3 or C 2H5

96 - 88 % Yield CHO R

R 3

NH2 H2N

6

NH N

DCM/ 55mol% mol%Bi(OTf) Bi(OTf)3 3 2h - 4h

O

O 7

95 - 86 % Yield

A series of coumarin-chalcone hybrid compounds and coumarins linked to pyrazoline was synthesized in good yields and short time using a simple and efficient method. This method involved the one-pot reaction of salicylaldehyde, an α-ketoester and an aromatic aldehyde (in the case of the coumarin-chalcone derivatives) in addition to hydrazine hydrate (in the case of the pyrazolyl coumarins) in the presence of a catalytic amount of bismuth triflate [Bi(OTf)3, 5 mol%]. The synthesized compounds showed scavenging activity towards the free radical 2,2-diphenyl-1-picrylhydrazyl. All compounds were characterized by IR, 1H NMR and 13C NMR spectroscopy. Chin. J. Catal., 2015, 36: 1131–1135

doi: 10.1016/S1872-2067(14)60297-7

Direct conversion of syngas into hydrocarbons over a core–shell Cr-Zn@SiO2@SAPO-34 catalyst

CO + H2

Jinjing Li, Xiulian Pan *, Xinhe Bao Dalian Institute of Chemical Physics, Chinese Academy of Sciences

An alternative route to the Fischer-Tropsch process for the one-step conversion of syngas into hydrocarbons is proposed using core–shell Cr-Zn@SiO2@SAPO-34 catalyst that enabled selective formation of C2–C4 hydrocarbons.

Cr-Zn

Cr-Zn@SiO2

C2−C4 hydrocarbons

Cr-Zn@SiO2@SAPO-34

xvi

Graphical Contents / Chinese Journal of Catalysis Vol. 36, No. 7, 2015

Chin. J. Catal., 2015, 36: 1136–1141

doi: 10.1016/S1872-2067(14)60309-0

Low pressure one-pot synthesis of dimethyl carbonate catalyzed by an alkali carbonate Chun Liu *, Shaoke Zhang, Baoyi Cai, Zilin Jin Dalian University of Technology

This work developed a protocol of alkali carbonate-catalyzed one-pot synthesis of dimethyl carbonate (DMC) from epoxide, CO2, and methanol. 63.5% yield of DMC was achieved with an initial CO2 pressure of 0.5 MPa and ethylene oxide as reactant. Chin. J. Catal., 2015, 36: 1142–1154

doi: 10.1016/S1872-2067(14)60312-0

Oxidative carbonylation of phenol with a Pd-O/CeO2-nanotube catalyst Ye Yuan, Zhimiao Wang, Hualiang An, Wei Xue *, Yanji Wang * Hebei University of Technology

CeO2 nanotube

CeO2 nanotubes (CeO2-NT) as the support for a Pd catalyst, Pd-O/CeO2-NT, gave high phenol conversion with high dipenyl carbonate selectivity in the oxidative carbonylation of phenol.

Chinese Journal of Catalysis 36 (2015) 1093–1100

催化学报 2015年 第36卷 第7期 | www.chxb.cn

available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/chnjc

Article

Stereoselective synthesis of vic-halohydrins and an unusual Knoevenagel product from an organocatalyzed aldol reaction: A non-enamine mode P. B. Thorat, S. V. Goswami, V. P. Sondankar, S. R. Bhusare * Department of Chemistry, Dnyanopasak College, Parbhani-431401, MS, India

A R T I C L E

I N F O

Article history: Received 5 January 2015 Accepted 10 February 2015 Published 20 July 2015 Keywords: Aldol reaction Asymmetric synthesis Chloroacetone Diastereoselectivity Hydroxy propanoate Knoevenagel reaction Pyrrolidine derivative

A B S T R A C T

Stereoselective synthesis by an aldol reaction between chloroacetone and aldehyde was studied using a synthesized chiral organocatalyst and triethylamine. The reaction gave α-chloro-β-hydroxy ketones in excellent yield with high anti selectivity and enantioselectivity. The chiral organocatalyst was also used in the Knoevenagel reaction, which gave α-cyano-β-hydroxy ketones at a low temperature and the usual Knoevenagel product at a high temperature. Both products were obtained in good to moderate yield with good anti selectivity in the case of α-cyano-β-hydroxy ketone derivatives. © 2015, Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved.

1. Introduction Vic-halohydrins are important intermediates for the synthesis of biologically active compounds and natural products such as bagonists [1], substituted pyrrolidines [2], polychlorosulfolipid [3], and insect sex pheromones [4]. These molecules are building blocks for various bioactive products and natural compounds. Thus efforts have been made to develop effective techniques to synthesize these compounds. The ring opening of enantiomerically pure epoxides is one of the most popular process [5] but suffers from the disadvantage of the formation of regioisomers. The asymmetric reduction of prochiral α-halo ketones by chiral catalysts such as oxazaborolidine with borane [1–2,6], Ru [1,7], or Rh [7,8] by asymmetric hydrogenation has shown good enantioselectivity. The condensation reaction of carbonyl compounds with an

active methylene compound in the presence of a base is known as Knoevenagel condensation [8]. The Knoevenagel reaction is an important C–C bond forming reaction because the synthesized alkenes are very useful intermediates in organic synthesis [9]. For example, Knoevenagel condensation has been successfully combined with hetero-Diels-Alder reactions, Michael addition reactions, ene reactions, and sigmatropic rearrangement for the synthesis of highly functionalized molecules [10–13]. Asymmetric organocatalysis has become an important area of research in organic synthesis. The development of organocatalyzed reactions for the stereoselective construction of C–C bonds has been intensively investigated. Since the discovery by List et al. [14] that L-proline can mimic enantioselectively catalyzed intermolecular aldol reactions, many organocatalysts have been synthesized with the aim of increasing their activity and stereoselectivity [15–17]. These reactions are typical of

* Corresponding author. Tel: +91-2452-242411; Fax: +91-2452-242493; E-mail: [email protected] DOI: 10.1016/S1872-2067(14)60317-X | http://www.sciencedirect.com/science/journal/18722067 | Chin. J. Catal., Vol. 36, No. 7, July 2015

1094

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

enamine based organocatalysis and proceeds via the reversible condensation of the catalytic amine with a ketone to provide a nucleophilic enamine intermediate. In these reactions, the carboxylic acid functionality on proline is important, and it is postulated that it activates and orients the aldehyde acceptor via a hydrogen bonding interaction. Alternatively, the utility of a N-sulfinyl group [18–22] as both a chiral directing group and acidifying element in hydrogen-bonded organocatalysts has been demonstrated. In these reactions, the sulfinyl N–H bond is postulated to activate the substrate by the formation of a key hydrogen bonding interaction. The inductive electron-withdrawing effect of the sulfinyl group acidifies the N–H bond, which serves to modulate the hydrogen bonding interaction. In addition, the close proximity of the stereogenic sulfur to the active site of the catalyst contributes to high stereoselectivity control in these reactions. On the basis of the success of hydrogen bonding organocatalysis, we studied pyrrolidine based organocatalysis in the absence of a Lewis basic primary or secondary amine functionality, which occurs exclusively via hydrogen bonding and van der Waals interactions. We restricted enamine/imine formation by introducing an acetyl group onto the active nitrogen site in the pyrrolidine ring and replaced the carboxylic group by an amide and sulfinyl group. To perform hydrogen bonding organocatalysis, we synthesized the organocatalysts shown in Fig. 1. Earlier results have been reported in a previous communication [23]. 2. Experimental 2.1. General details All solvents used were commercial anhydrous grade used without further purification. Aluminium sheets 20 cm  20 cm, silica gel 60 F254, Merck grade was used for thin layer chromatography to determine the progress of the reaction. Column chromatography was carried out over silica gel (80–120 mesh). The optical rotation was measured on a Polax-2L digital polarimeter. The melting point was determined in an open capillary tube and was uncorrected. 1H and 13C NMR spectra were recorded on an Avance 300 MHz spectrometer using CDCl3 solvent. Mass spectra were obtained on a Polaris-Q Thermoscientific GC-MS. Elemental analyses were obtained using a flash EA 1112. Enantiomeric purity was determined on a PerkinElmer Series 200 HPLC System. O CONH2

N 1

N

O

O O N HN S O O 3

O HN S Me O O 2 O

Me

N

O HN S O O 4

Fig. 1. Pyrrolidine based organocatalysts.

NO2

2.2. Preparation of organocatalysts 1–4 The synthesis and characterization of the organocatalyst was described in our previous communication [23]. 2.3. General procedure for the aldol reaction for the synthesis of 7a–7f To a solution of chloroacetone 6 (1.5 eq.) dissolved in ethanol, a few drops of triethylamine (3–4 drops) was added and the reaction mixture was stirred for 15–20 min. To this stirred solution, an aromatic aldehyde 5(a–f) (1.0 eq.) and a catalytic amount of the organocatalyst (S)-1-acetyl-N-tosylpyrrolidine2-carboxamide 3 was added and reacted and stirred for an appropriate time (Table 2). The progress of reaction was monitored by thin layer chromatography. After completion of the reaction, the solvent was evaporated under vacuum. The crude product was partitioned between ethyl acetate and water. The organic layer was collected, and the aqueous phase was extracted with ethyl acetate. The collected organic layer was washed with saturated brine solution, and the product was purified using column chromatography using silica gel (80–120 mesh). 3-Chloro-4-hydroxy-4-(4-nitrophenyl)butan-2-one (7a). Light yellow solid (182 mg, 90%), M.P. 142–144 °C [19]; 1H NMR (300 MHz, CDCl3): δ 7.40–7.52 (m, 2H), 7.12–7.28 (m, 2H), 6.01 and 5.76 (bs, 1H, OH for syn and anti), 5.21 (s, 1H) syn, 4.96 (d, 1H, J = 8.7 Hz) anti, 4.29 (s, 1H) anti, 3.95 (s, 1H) syn, 2.08 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 209.9, 154.5, 149.8, 129.7, 123.2, 68.9, 59.1, 28.1; GC-MS m/z 243 (M+); Elemental analysis: Anal. Calcd for C10H10ClNO4: C, 49.30; H, 4.14; Cl, 14.55; N, 5.75; O, 26.27; Found C, 49.31; H, 4.17; Cl, 14.57; N, 5.76; O, 26.29. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/ hexane/ HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 17.2 min, tR(minor) = 25.5 min, ee = 92%. 3-Chloro-4-(4-fluorophenyl)-4-hydroxybutan-2-one (7b). Colorless solid, (158 mg, 92%), M.P. 77–79 °C; 1H NMR (300 MHz, CDCl3): δ 7.56–7.66 (m, 2H), 7.23–7.34 (m, 2H), 5.79 and 5.46 (bs, 1H, OH for syn and anti), 5.21 (s, 1H) syn, 4.88 (d, 1H, J = 8.2 Hz) anti, 4.33 (s, 1H) anti, 3.89 (s, 1H) syn, 1.98 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 208.4, 151.8, 147.6, 128.9, 124.5, 70.4, 57.9, 29.1; GC-MS m/z 246 (M+); Elemental analysis: Anal. Calcd for C11H12ClFO3: C, 53.56; H, 4.90; Cl, 14.37; F, 7.70; O, 19.46; Found C, 53.54; H, 4.91; Cl, 14.40; F, 7.73; O, 19.44. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 23.2 min, tR(minor) = 27.1 min, ee = 92%. 4-(2-Chloro-1-hydroxy-3-oxobutyl)benzonitrile (7c). Viscous oil (171 mg, 88%) [19]; 1H NMR (300 MHz, CDCl3): δ 7.56–7.94 (m, 4H), 6.14 and 5.82 (bs, 1H, OH for syn and anti), 5.61 (s, 1H) syn, 4.37 (d, 1H, J = 9.3 Hz) anti, 4.45 (s, 1H) anti, 3.95 (s, 1H) syn, 2.17 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 205.1, 149.1, 132.7, 128.3, 112.2, 106.6, 81.1, 64.4, 23.4; GC-MS m/z 223 (M+); Elemental analysis: Anal. Calcd for C11H10ClNO2: C, 59.07; H, 4.51; Cl, 15.85; N, 6.26; O, 14.31; Found C, 59.10; H,

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

4.55; Cl, 15.87; N, 6.29; O, 14.33. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 24.2 min, tR(minor) = 31.4 min, ee = 90%. 3-Chloro-4-(4-chlorophenyl)-4-hydroxybutan-2-one (7d). White sticky solid (165 mg, 85%); 1H NMR (300 MHz, CDCl3): δ 7.26–7.36 (m, 4H), 6.03 and 5.79 (bs, 1H, OH for syn and anti), 5.34 (s, 1H) syn, 4.41 (d, 1H, J = 8.4 Hz) anti, 4.36 (s, 1H) anti, 4.03 (s, 1H) syn, 2.14 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 203.4, 153.3, 147.1, 129.2, 126.7, 71.0, 59.8, 29.5; GC-MS m/z 263 (M+); Elemental analysis: Anal. Calcd for C11H12Cl2O3: C, 50.21; H, 4.60; Cl, 26.95; O, 18.24; Found C, 50.24; H, 4.63; Cl, 26.91; O, 18.22. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/ hexane/ HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 20.2 min, tR(minor) = 27.3 min, ee = 88%. 4-(2-Bromophenyl)-3-chloro-4-hydroxybutan-2-one (7e). Brownish oil (178 mg, 77%) [24]; 1H NMR (300 MHz, CDCl3): δ 7.26–7.35 (m, 2H), 7.19–7.25 (m, 1H), 7.05–7.12 (m, 1H), 5.74 and 5.51 (bs, 1H, OH for syn and anti), 5.19 (s, 1H) syn, 4.93 (d, 1H, J = 7.2 Hz) anti, 4.40 (s, 1H) anti, 3.85 (s, 1H) syn, 2.11 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 210.3, 153.6, 148.3, 127.2, 126.5, 71.8, 59.4, 30.4; GC-MS m/z 307 (M+); Elemental analysis: Anal. Calcd for C11H12BrClO3: C, 42.96; H, 3.93; Br, 25.98; Cl, 11.53; O, 15.61; Found C, 42.96; H, 3.93; Br, 25.98; Cl, 11.53; O, 15.61. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/ HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 23.4 min, tR(minor) = 25.1 min, ee = 84%. 3-Chloro-4-hydroxy-4-phenylbutan-2-one (7f). Colorless solid (120 mg, 73%), M.P. 110–112 °C [24]; 1H NMR (300 MHz, CDCl3): δ 6.97–7.43 (m, 4H), 5.84 and 5.51 (bs, 1H, OH for syn and anti), 5.17 (s, 1H) syn, 4.92 (d, 1H, 7.5 Hz) anti, 4.27 (s, 1H) anti, 4.01 (s, 1H) syn, 2.04 (s, 3H); 13C NMR (300 MHz, CDCl3): δ 203.4, 153.3, 145.5, 123.2, 123.1, 71.4, 60.0, 31.1; GC-MS m/z 228 (M+); Elemental analysis: Anal. Calcd for C11H13ClO3: C, 57.78; H, 5.73; Cl, 15.50; O, 20.99; Found C, 57.80; H, 5.74; Cl, 15.52; O, 20.96. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/ hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 27.4 min, tR(minor) = 28.2 min, ee = 86%. 2.4. General procedure for the Knoevenagel condensation reaction (9a–9j) To a solution of ethylcyanoacetate 8 (1.5 eq.) dissolved in ethanol, a few drops of triethylamine (3–4 drops) was added and the reaction mixture was stirred for 15–20 min. To this stirred solution, an aromatic aldehyde 5(a–f) (1.0 eq.) and a catalytic amount of the organocatalyst (S)-1-acetyl-N-tosylpyrrolidine-2-carboxamide 3 in ethanol was added. The reaction mixture was stirred at 50 °C for an appropriate time (Table 4). The progress of reaction was monitored by thin layer chromatography. After completion of the reaction, the mixture was cooled and the solvent was evaporated under vacuum. The crude product was partitioned between ethyl acetate and water. The organic layer was collected and the aqueous phase was

1095

extracted with ethyl acetate. The collected organic layer was washed with saturated brine solution and the product was purified using column chromatography using silica gel (80–120 mesh). Ethyl 2-cyano-3-(4-nitrophenyl)acrylate (9a). Light yellow solid (261 mg, 79%), M.P. 168–170 °C [25]; 1H NMR (300 MHz, CDCl3): δ 8.27 (s, 1H), 7.12–7.28 (m, 2H), 7.69–7.79 (m, 2H), 3.98 (q, 2H), 1.38 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 165.03, 151.00, 146.65, 129.71, 122.58, 119.74, 117.18, 101.69, 59.21, 18.56; GC-MS m/z 246 (M+); Elemental analysis: Anal. Calcd for C12H10N2O4: C, 58.54; H, 4.09; N, 11.38; O, 25.99; Found C, 58.53; H, 4.10; N, 11.36; O, 25.97. Ethyl 2-cyano-3-(3-nitrophenyl)acrylate (9b). Yellow solid (240 mg, 78%), M.P. 133–135 °C [26]; 1H NMR (300 MHz, CDCl3): δ 8.19 (s, 1H), 8.21–8.26 (m, 1H), 7.91–7.98 (m, 1H), 7.81–7.89 (m, 2H), 4.12 (q, 2H), 1.25 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 162.6, 152.30, 147.33, 132.91, 128.25, 126.81, 126.11, 122.61, 117.18, 105.16, 61.42, 14.87; GC-MS m/z 246 (M+); Elemental analysis: Anal. Calcd for C12H10N2O4: C, 58.54; H, 4.09; N, 11.38; O, 25.99; Found C, 58.51; H, 4.11; N, 11.35; O, 26.02. Ethyl 2-cyano-3-(4-fluorophenyl)acrylate (9c) (CAS No. 18861-57-9). Yellow crystalline (135 mg, 79%), M.P. 96–98 °C; 1H NMR (300 MHz, CDCl3): δ 8.42 (s, 1H), 7.01–7.19 (m, 2H), 7.82–8.10 (m, 2H), 4.26 (q, 2H), 1.60 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 163.51, 162.24, 153.33, 126.03, 119.68, 117.27, 98.64, 55.87, 22.31; GC-MS m/z 219 (M+); Elemental analysis: Anal. Calcd for C12H10FNO2: C, 65.75; H, 4.60; F, 8.67; N, 6.39; O, 14.60; Found C, 65.71; H, 4.63; F, 8.68; N, 6.40; O, 14.61. Ethyl 2-cyano-3-(4-cyanophenyl)acrylate (9d). Colorless solid (244 mg, 80%), M.P. 133–135 °C; 1H NMR (300 MHz, CDCl3): δ 8.34 (s, 1H), 7.67–7.73 (m, 2H), 7.42–8.61 (m, 2H), 4.23 (q, 2H), 1.34 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 161.54, 157.43, 138.61, 130.91, 129.17, 117.34, 117.02, 109.76, 99.97, 63.12, 15.43; GC-MS m/z 226 (M+); Elemental analysis: Anal. Calcd for C13H10N2O2: C, 69.02; H, 4.46; N, 12.38; O, 14.14; Found C, 68.99; H, 4.49; N, 12.36; O, 14.12. Ethyl 3-(4-chlorophenyl)-2-cyanoacrylate (9e). Brownish solid (240 mg, 76%), M.P. 92–94 °C [27]; 1H NMR (300 MHz, CDCl3): δ 8.19 (s, 1H), 7.82–7.88 (m, 2H), 7.12–8.23 (m, 2H), 4.18 (q, 2H), 1.28 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 162.71, 152.13, 135.37, 134.89, 131.29, 130.27, 118.35, 101.49, 62.94, 14.99; GC-MS m/z 235 (M+); Elemental analysis: Anal. Calcd for C12H10ClNO2: C, 61.16; H, 4.28; Cl, 15.04; N, 5.94; O, 13.58; Found C, 61.14; H, 4.31; Cl, 15.02; N, 5.96; O, 13.61. Ethyl 3-(2-chlorophenyl)-2-cyanoacrylate (9f). Reddish solid (231 mg, 73%), M.P. 54–55 °C [27]; 1H NMR (300 MHz, CDCl3): δ 8.11 (s, 1H), 7.54–7.49 (m, 1H), 7.31–8.44 (m, 2H), 4.23 (q, 2H), 1.32 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 162.71, 150.84, 134.76, 130.41, 129.65, 129.01, 128.77, 116.97, 104.42, 63.24, 15.43; GC-MS m/z 235 (M+); Elemental analysis: Anal. Calcd for C12H10ClNO2: C, 61.16; H, 4.28; Cl, 15.04; N, 5.94; O, 13.58; Found C, 61.13; H, 4.29; Cl, 15.03; N, 5.93; O, 13.60. Ethyl 2-cyano-3-phenylacrylate (9g). White solid (210 mg, 72%), M.P. 51–53 °C [28]; 1H NMR (300 MHz, CDCl3): δ 8.24 (s, 1H), 7.33–7.39 (m, 1H), 7.33–8.10 (m, 4H), 4.20 (q, 2H), 1.27 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 160.21, 154.11, 133.54,

1096

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

132.87, 129.75, 129.22, 128.77, 117.16, 101.59, 60.89, 14.87; GC-MS m/z 201 (M+); Elemental analysis: Anal. Calcd for C12H11NO2: C, 71.63; H, 5.51; N, 6.96; O, 15.90; Found C, 71.63; H, 5.51; N, 6.96; O, 15.90. Ethyl 2-cyano-3-p-tolylacrylate (9h). White solid (203 mg, 74%), M.P. 90–92 °C [29]; 1H NMR (300 MHz, CDCl3): δ 8.17 (s, 1H), 7.76–8.88 (m, 2H), 7.28–7.39 (m, 2H), 4.01 (q, 2H), 2.40 (s, 3H), 1.31 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 161.91, 151.77, 142.81, 122.72, 122.68, 119.27, 99.99, 60.08, 23.19, 16.29; GC-MS m/z 215 (M+); Elemental analysis: Anal. Calcd for C13H13NO2: C, 72.54; H, 6.09; N, 6.51; O, 14.87; Found C, 72.56; H, 6.06; N, 6.53; O, 14.85. Ethyl 2-cyano-3-(4-hydroxyphenyl)acrylate (9i). Reddish solid (203 mg, 69%), M.P. 172–174 °C [30]; 1H NMR (300 MHz, CDCl3): δ 9.28 (bs, 1H, OH), 8.68 (s, 1H), 8.21–8.40 (m, 2H), 7.28–7.71 (m, 2H), 4.08 (q, 2H), 1.26 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 163.47, 159.83, 150.01, 138.30, 127.61, 121.89, 117.26, 108.05, 61.87, 23.24; GC-MS m/z 217 (M+); Elemental analysis: Anal. Calcd for C12H11NO3: C, 66.35; H, 5.10; N, 6.45; O, 22.10; Found C, 66.33; H, 5.12; N, 6.43; O, 22.11. Ethyl 2-cyano-3-(3,4-dimethoxyphenyl)acrylate (9j). Brownish solid (234 mg, 67%), M.P. 56–58 °C [31]; 1H NMR (300 MHz, CDCl3): δ 8.23 (s, 1H), 7.23–7.10 (m, 2H), 7.01–7.06 (m, 1H), 4.16 (q, 2H), 3.90 (s, 6H), 1.31 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 164.09, 151.85, 150.31, 149.78, 129.65, 123.76, 117.08, 116.98, 116.31, 102.37, 60.45, 58.93, 17.62; GC-MS m/z 261 (M+); Elemental analysis: Anal. Calcd for C14H15NO4: C, 64.36; H, 5.79; N, 5.36; O, 24.49; Found C, 64.34; H, 5.81; N, 5.39; O, 24.47. 2.5. General procedure for the synthesis of hydroxy propanoate derivatives (10a–10j) To a solution of ethylcyanoacetate 8 (1.5 eq.) dissolved in ethanol, a few drops of triethylamine (3–4 drops) was added and the reaction mixture was stirred for 15–20 min. To this stirred cold solution, an aromatic aldehyde 5(a–f) (1.0 equivalent) and a catalytic amount of the organocatalyst (S)-1-acetylN-tosylpyrrolidine-2-carboxamide 3 in ethanol was added. The reaction mixture was cooled to –78 °C and stirred for an appropriate time (Table 4). The progress of reaction was monitored by thin layer chromatography. After completion of the reaction, crushed ice was added to this cold solution and the mixture was allowed to stand. The reaction mixture was extracted with ethyl acetate. The collected organic layer was washed with saturated brine solution and the product was purified using column chromatography using silica gel (80–120 mesh). Ethyl 2-cyano-3-(4-nitrophenyl)-3-hydroxypropanoate (10a). Yellow viscous oil (274 mg, 89%); 1H NMR (300 MHz, CDCl3): δ 7.48–60 (m, 2H), 7.19–7.30 (m, 2H), 5.17 (bs, 1H, OH), 4.44 (d, 1H), 4.31 (q, 2H), 3.87 (d, 1H), 1.32 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 163.85, 149.97, 148.69, 130.78, 125.63, 117.11, 74.83, 58.23, 39.00, 16.91; GC-MS m/z 264 (M+); HRMS: Calculated 264.0746; Found 264.0744; Elemental analysis: Anal. Calcd for C12H12N2O5: C, 54.55; H, 4.58; N, 10.60; O, 30.28; Found C, 54.55; H, 4.58; N, 10.60; O, 30.28. Enantiomeric purity was de-

termined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 7.7 min, tR(minor) = 15.2 min, ee = 81%. Ethyl 2-cyano-3-(3-nitrophenyl)-3-hydroxypropanoate (10b). Yellow viscous oil (261 mg, 85%); 1H NMR (300 MHz, CDCl3): δ 7.27–7.35 (m, 2H), 7.21–7.26 (m, 1H), 7.05–7.17 (m, 1H), 5.19 (bs, 1H, OH), 4.39 (d, 1H), 4.37 (q, 2H), 3.91 (d, 1H), 1.37 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 164.33, 150.02, 143.32, 131.78, 130.65, 125.63, 123.81, 113.01, 72.64, 59.44, 41.21, 18.92; GC-MS m/z 264 (M+); HRMS: Calculated 264.0746; Found 264.0747; Elemental analysis: Anal. Calcd for C12H12N2O5: C, 54.55; H, 4.58; N, 10.60; O, 30.28; Found C, 54.55; H, 4.58; N, 10.60; O, 30.28. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/ hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 9.0 min, tR(minor) = 15.7 min, ee = 84%. Ethyl 2-cyano-3-hydroxy-3-(4-flurophenyl) propanoate (10c). Colorless oil (238 mg, 87%); 1H NMR (300 MHz, CDCl3): δ 7.94–8.29 (m, 2H), 7.18–7.29 (m, 2H), 5.28 (bs, 1H, OH), 4.90 (d, 1H), 4.30 (q, 2H), 3.31 (d, 1H), 1.29 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 166.19, 165.51, 130.11, 127.03, 126.02, 110.13, 71.58, 66.70, 46.42, 14.59; GC-MS m/z 237 (M+); HRMS: Calculated 237.0801; Found 237.0802; Elemental analysis: Anal. Calcd for C12H12FNO3: C, 60.76; H, 5.10; F, 8.01; N, 5.90; O, 20.23; Found C, 60.74; H, 5.13; F, 7.99; N, 5.93; O, 20.26. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 13.3 min, tR(minor) = 15.3 min, ee = 82%. Ethyl 2-cyano-3-(4-cyanophenyl)-3-hydroxypropanoate (10d). Colorless oil (249 mg, 89%); 1H NMR (300 MHz, CDCl3): δ 7.32–7.57 (m, 4H), 5.32 (bs, 1H, OH), 4.87 (d, 1H), 4.35 (q, 2H), 3.31 (d, 1H), 1.31 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 165.66, 147.39, 133.87, 127.03, 124.38, 120.57, 112.59, 71.56, 68.93, 42.67, 16.74; GC-MS m/z 244 (M+); HRMS: Calculated 244.0848; Found 244.0850; Elemental analysis: Anal. Calcd for C13H12N2O3: C, 63.93; H, 4.95; N, 11.47; O, 19.65; Found C, 63.91; H, 4.98; N, 11.44; O, 19.63. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 19.1 min, tR(minor) = 19.9 min, ee = 79%. Ethyl 3-(4-chlorophenyl)-2-cyano-3-hydroxypropanoate (10e). Brown semisolid (247 mg, 84%); 1H NMR (300 MHz, CDCl3): δ 7.16–7.24 (m, 2H), 7.02–7.11 (m, 2H), 5.35 (bs, 1H, OH), 4.89 (d, 1H), 4.27 (q, 2H), 3.29 (d, 1H), 1.31 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 164.22, 145.51, 137.23, 126.02, 124.32, 117.71, 69.55, 59.49, 39.41, 16.94; GC-MS m/z 253 (M+); HRMS: Calculated 253.0506; Found 253.0508; Elemental analysis: Anal. Calcd for C12H12ClNO3: C, 56.81; H, 4.77; Cl, 13.98; N, 5.52; O, 18.92; Found C, 56.83; H, 4.78; Cl, 13.95; N, 5.55; O, 18.93. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 11.6 min, tR(minor) = 16.4 min, ee = 83%. Ethyl 3-(2-chlorophenyl)-2-cyano-3-hydroxypropanoate (10f). Reddish gum (238 mg, 81%); 1H NMR (300 MHz, CDCl3): δ 7.51–7.58 (m, 1H), 7.22–7.46 (m, 3H), 5.29 (bs, 1H, OH), 4.89

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

(d, 1H), 4.31 (q, 2H), 3.35 (d, 1H), 1.29 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 164.62, 144.45, 133.28, 127.34, 127.27, 122.63, 118.84, 63.67, 60.34, 40.27, 15.92; GC-MS m/z 237 (M+); HRMS: Calculated 253.0506; Found 253.0507; Elemental analysis: Anal. Calcd for C12H12ClNO3: C, 56.81; H, 4.77; Cl, 13.98; N, 5.52; O, 18.92; Found C, 56.79; H, 4.79; Cl, 13.96; N, 5.55; O, 18.93. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/ HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 14.0 min, tR(minor) = 20.3 min, ee = 77%. Ethyl 2-cyano-3-hydroxy-3-phenylpropanoate (10g). Colorless oil (214 mg, 79%); 1H NMR (300 MHz, CDCl3): δ 7.04–7.48 (m, 4H), 5.33 (bs, 1H, OH), 4.94 (d, 1H), 4.33 (q, 2H), 3.30 (d, 1H), 1.31 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 164.42, 141.32, 126.24, 125.25, 124.89, 117.83, 61.37, 60.07, 40.33, 15.43; GC-MS m/z 237 (M+); HRMS: Calculated 219.0895; Found 219.0893; Elemental analysis: Anal. Calcd for C12H13NO3: C, 65.74; H, 5.98; N, 6.39; O, 21.89; Found C, 65.71; H, 6.00; N, 6.41; O, 21.87. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/ hexane/ HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 10.1 min, tR(minor) = 11.0 min, ee = 78%. Ethyl 2-cyano-3-hydroxy-3-p-tolylpropanoate (10h). White amorphous (201 mg, 80%); 1H NMR (300 MHz, CDCl3): δ 7.34–8.45 (m, 2H), 7.14–7.26 (m, 2H), 5.30 (bs, 1H, OH), 4.95 (d, 1H), 4.31 (q, 2H), 3.83 (d, 1H), 2.68 (s, 3H), 1.25 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 162.74, 147.75, 147.26, 122.09, 120.16, 117.61, 64.32, 58.78, 40.69, 21.73, 13.58; GC-MS m/z 233 (M+); HRMS: Calculated 235.0845; Found 235.0844; Elemental analysis: Anal. Calcd for C13H15NO3: C, 66.94; H, 6.48; N, 6.00; O, 20.58; Found C, 66.97; H, 6.45; N, 5.97; O, 20.59. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 14.6 min, tR(minor) = 17.5 min, ee = 75%. Ethyl 2-cyano-3-hydroxy-3-(4-hydroxyphenyl) propanoate (10i). Reddish viscous oil (207 mg, 77%); 1H NMR (300 MHz, CDCl3): δ 7.94–8.29 (m, 2H), 7.18–7.29 (m, 2H), 5.29 (bs, 1H, OH), 4.96 (d, 1H), 4.28 (q, 2H), 3.73 (d, 1H), 1.26 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 165.03, 72.34, 157.62, 140.78, 122.09, 116.28, 115.55, 60.12, 57.08, 36.93, 12.96; GC-MS m/z 235 (M+); HRMS: Calculated 279.1107; Found 279.1106; Elemental analysis: Anal. Calcd for C12H13NO4: C, 61.27; H, 5.57; N, 5.95; O, 27.21; Found C, 61.28; H, 5.59; N, 5.92; O, 27.24. Enantiomeric purity was determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 12.3 min, tR(minor) = 19.2 min, ee = 76%. Ethyl 2-cyano-3-hydroxy-3-(3,4-dimethoxyphenyl) propanoate (10j). Brownish oil (233 mg, 73%); 1H NMR (300 MHz, CDCl3): δ 7.57–8.69 (m, 2H), 7.24–7.33 (m, 2H), 5.20 (bs, 1H, OH), 4.77 (d, 1H), 4.31 (q, 2H), 3.89 (s, 6H), 3.43 (d, 1H), 1.32 (t, 3H); 13C NMR (300 MHz, CDCl3): δ 163.09, 157.51, 157.13, 130.91, 118.83, 109.20, 106.00, 105.01, 78.25, 73.14, 62.00, 49.23, 20.50; GC-MS m/z 279 (M+); Elemental analysis: Anal. Calcd for C14H17NO5: C, 60.21; H, 6.14; N, 5.02; O, 28.64; Found C, 60.19; H, 6.17; N, 5.04; O, 28.66. Enantiomeric purity was

1097

determined by chiral HPLC using Whelk-O1 (25 cm  4.6 mm), 20:80:0.5 IPA/hexane/HOAc, flow rate 1.0 mL/min,  = 254 nm; tR(major) = 11.0 min, tR(minor) = 11.4 min, ee = 76%. 3. Results and discussion In a continuation of our work on organocatalysis [32–36], we explored these chiral organocatalysts in the asymmetric aldol reaction. The aldol reaction of chloroacetone (6) and p-nitrobenzaldehyde (5a) to afford 7a was performed in ethanol using 10 mol% of the synthesized organocatalysts 1–4 bearing different sulfonamide groups in the presence of triethylamine (Scheme 1). Initially, when just N-acylated prolinamide 1 was used as the catalyst in the aldol reaction, after stirring for 48 h, we obtained only 29% yield. The diastereoselectivity and enantioselectivity were not determined for the reaction (Table 1, entry 1). The methanesulfonamide catalyst 2 afforded some improved yield (40%), but the diastereoselectivity and enantiomeric excess were very poor. The best result with respect to yield, diastereoselectivity, and enantioselectivity was with catalyst 3 and triethylamine. The reaction with organocatalyst 3 gave 7a in good yield with good diastereoselectivity and improved enantioselectivity 69% ee for anti-7a (Table 1, entry 3). Catalyst 4 catalyzed the aldol reaction affording 59% yield and decreased diastereoselectivity and enantiomeric excess (Table 1, entries 2–4). Although organocatalysts 3 and 4 only showed moderate stereoselectivity, the reaction time was short, esp. with organocatalyst 3, which completed the reaction within 20 h. Interestingly, the reaction gave enhanced stereoselectivity on increasing the catalyst loading (Table 1, entries 7–8). When the loading of organocatalyst 3 was increased to 12 mol%, the anti-aldol product was obtained with good enantioselectivity (74% ee). The breakthrough was achieved when 15 mol% of catalyst 3 was employed with triethylamine as the co-catalyst. O

O

O

OH

Organocatalyst H + O2N

Cl 5a

Et3N, ethanol

Cl

NO2 7a

6

Scheme 1. Aldol reaction of chloroacetone with p-nitrobenzaldehyde. Table 1 Optimization of the aldol reaction between chloroacetone and p-nitro benzaldehyde. Entry Catalyst

Catalyst (mol%) 10 10 10 10 12 15 20 15

Time (h) 48 48 20 24 18 15 26 19

Yield a (%) 29 40 68 59 75 90 53 61

1 1 2 2 3 3 4 4 5 3 6 3 7 3 8 4 a Isolated yield. b Determined by 1H NMR. c Determined by chiral-phase HPLC analysis. d Not determined.

anti/syn b Nd d 2:1 7:1 5:1 7:1 8:1 3:1 4:1

anti ee c (%) Nd d 36 69 57 74 79 45 66

1098

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

Table 2 Aldol reaction of chloroacetone with various aldehydes.

Table 3 Knoevenagel reaction between ethylcyanoacetate and p-nitrobenzaldehyde at different temperatures in the presence of organocatalyst 3b (reaction time 20 h). Entry

Entry

Ar

Time (h) 15 15 16 18 20 24

Product

Yield a anti/syn b (%) 90 8:1 88 7:1 92 8:1 85 7:1 77 6:1 73 6:1

1 p-NO2–C6H4 7a 2 p-F–C6H4 7b 7c 3 p-CN–C6H4 4 p-Cl–C6H4 7d 7e 5 o-Br–C6H4 7f 6 C6H5 a Isolated yield. b Determined by 1H NMR. c Determined by chiral-phase HPLC analysis.

anti ee c (%) 92 92 90 88 84 86

Up to 90% of the desired product 7a was obtained with a higher selectivity (anti/syn was 8:1). The anti-aldol product was obtained with the excellent enantioselectivity of 92%. Further increase in the loading of catalyst 3 in the presence of triethylamine gave no more improvement, and the result showed significant deviation from the previous trend of optimization. The aldol reaction under the optimized condition using organocatalyst 4 in ethanol was also examined. This catalyst showed lower diastereoselectivity and enantioselctivity with decreased yield of the product 7a (Table 1, entries 5–7), even with an extended reaction time. The scope of this class of aldol reactions using organocatalyst 3 and co-catalyst triethylamine in ethanol was examined with a series of arylaldehydes 5 (Table 2) and chloroacetones 6. Aromatic aldehydes with an electron-withdrawing substituent were excellent substrates. The reaction proceeded smoothly to afford the desired products 7a–7f. In all cases, the reactions afforded the anti-aldol products in high yield with excellent enantioselectivity. The scope of the catalyst system of chiral compound 3 and triethylamine was also assessed for the Knoevenagel condensation reaction between p-nitrobenzaldehyde 5a and ethylcyanoacetate 8 using ethanol as solvent (Scheme 2). The base-mediated reaction at room temperature afforded the Knoevenagel adduct 9a in 69% yield. A good increase in the product yield was observed with a slight raise in the temperature to 40 °C. At the still higher temperature condition of 50 °C, the best yield for the Knoevenagel product was achieved. No significant change in yield was observed by a further increase in the temperature from 50 °C to 60 °C. The influence of temperature on the Knoevenagel reaction is shown in Table 3. Surprisingly, while experimenting on the same reaction at cold conditions, a byproduct was observed in very low percent yield. The byproduct was isolated by column chromatography

Scheme 2. Knovenagel reaction between p-nitrobenzaldehyde 5a and ethylcyanoacetate 8 using organocatalyst 3 and triethylamine.

Temperature (°C)

1 2 3 4 5 6 7 a Isolated yield.

RT 40 50 60 0 –20 –78

Yield a (%) 9a 69 78 89 84 54 23 18

10a — — — — 26 40 73

and analyzed for its structure. Our estimate was that the byproduct obtained could be 2-cyano-3-hydroxy-3-(4-nitrophenyl) propanoate formed during the reaction (Scheme 3). This was confirmed by the spectral data: proton NMR gave a broad singlet at 6.19 ppm attributed to the presence of a hydroxyl group which was not found in the proton NMR of ethyl2-cyano-3-(4-nitrophenyl) acrylate. Further confirmation was provided by 13C NMR which showed a signal at 74.83 ppm for the carbon atom bearing the hydroxyl group. In the Knoevenagel product, the signal for the same carbon atom appeared at 151 ppm. The stability of product 10a decreased with increase in temperature. The yield of product 10a was observed to increase with decrease in temperature. In the Knoevenagel reaction at a low temperature, 2-cyano-3-hydroxy-3-(4-nitrophenyl) propanoate is kinetically stable whereas at a higher temperature ethyl2-cyano-3-(4-nitrophenyl) acrylate is thermodynamically stable. After stirring for 20 h, the reaction yielded 54% Knoevenagel adduct and 26% of product 10a at 0 °C. At the lower temperature of –20 °C, only 23% of the Knoevenagel product 9a was observed with an increased yield of compound 10a. When the temperature was decreased to –78 °C, the percentage of kinetic stable product 10a was 73%. However, at elevated temperatures, compound 2-cyano-3-hydroxy-3-(4nitrophenyl) propanoate was not observed. At these temperatures, only the Knoevenagel product was observed, which is thermodynamically more stable. To check the necessity of the catalyst, the reaction was carried out in the absence of the catalyst. The reaction did not progress at a low temperature even after 48 h of stirring. However, at an increased temperature, the reaction showed some progress to give exclusively the Knoevenagel product. We also performed the reaction in the absence of the base triethylamine. The catalyst was able to promote the reaction to give a trace amount of product 9a only at high temperature. When the reaction was carried out with the replacing of ethyl-

Scheme 3. Knoevenagel reaction between p-nitrobenzaldehyde 5a and ethylcyanoacetate 8 using organocatalyst 3 and triethylamine at low temperature.

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

cyanoacetate by malononitrile or diethyl malonate, the reaction went to completion only at a high temperature to give only the Knoevenagel product. We extended this protocol to other aldehydes to ensure the reproducibility of the results (Table 4). The reaction proceeded smoothly to give both the products with the hydroxy propanoate derivatives as the major product at –78 °C. When reaction was performed at room temperature, only the Knoevenagel product was obtained. The reaction was effective with electron withdrawing groups present on the aromatic ring while an electron donating substituent gave a moderate performance. Analysis by 1H NMR spectra of the ethyl 2-cyano-3-hydroxy-3(aryl) propanoate derivatives showed that only the anti product was formed. We also determined the enantioselectivity to give the anti isomer of all products. The enantioselectivities were good to moderate. Figure 2 shows a possible mechanism for the reaction, which explains the results. Triethylamine converts the keto form into the enol form by abstracting a hydrogen. The carbonyl group in p-nitrobenzaldehyde is activated by strong hydrogen bonding, which weakens the π bond interaction between the carbon and oxygen of the carbonyl group and increases its electrophilicity. The enol attacks the Re-face of the carbon of the aldehyde, which leads to the formation of the favored transition state I shown in Fig. 2. The sulfur from the sulfonyl group contributes by the formation of van der Waals interaction, which plays a vital role in the transition state [32–36]. The hydroxy propanoate derivatives formed usually get dehydrated to give the Knoevenagel product. Due to the mild basicity, at low temperatures, the product does not get dehydrated, and hydroxy propanoate derivatives were obtained. However, at a higher temperature, the conditions were enough for the dehydration of the product formed in the presence of the base triethylamine. Table 4 Aldol reaction of ethylcyanoacetate with various aldehydes.

1099

O

O NH(Et)3 N(Et)3 Y

X O

X

O N

O

O N S O H OO X

X = Cl, CN Y = Me, OEt

Y

Y

O N

NH(Et)3

I Favored (anti)

O N S O H X OO Y NH(Et)3

II Disfavored (syn)

Fig. 2. Suggested transition state for the organocatalysed asymmetric aldol reaction.

4. Conclusions We successfully used an organocatalyst in an asymmetric aldol reaction. Our hypothesis on hydrogen bonding catalysis using triethylamine and an organocatalyst having a pyrrolidine ring as the backbone was verified. The organocatalyst (S)-1acetyl-N-tosylpyrrolidine-2-carboxamide 3 was an excellent catalyst. Extending the scope of our organocatalyst 3, we also studied the Knoevenagel reaction, which gave the usual Knoevenagel product at elevated temperatures and hydroxy propanoate derivatives at the low temperature of –78 °C. Our results also demonstrated that hydrogen bonding and van der Waals interactions are important in catalysis and can catalyze the reaction by stabilizing the intermediate. This protocol offers non-hazardous and eco-friendly reaction conditions with good yield and enantioselectivity. Acknowledgments We acknowledge Dr. P. L. More, Principal, Dnyanopasak College, Parbhani, for providing necessary facilities and IICT Hyderabad for providing spectral data. References [1] Lu C J, Luo Z H, Huang L, Li X S. Tetrahedron: Asymmetry, 2011, 22:

722 Entry

Ar

Time (h)

Yield a (%) 9 10 89 73 85 72 87 79 88 75 84 69 81 67 79 68 80 61 77 63 73 59

anti ee c (%) 81 84 82 79 83 77 78 75 76 76

1 p-NO2–C6H4 (5a) 20 2 m-NO2–C6H4 (5b) 21 3 p-F–C6H4 (5c) 19 22 4 p-CN–C6H4 (5d) 26 5 p-Cl–C6H4 (5e) 6 o-Cl–C6H4 (5f) 25 30 7 C6H5 (5g) 30 8 p-Me–C6H4 (5h) 9 p-OH–C6H4 (5i) 31 31 10 m,p-OCH3–C6H3 (5j) a Isolated yield at 50 °C. b Isolated yield at –78 °C c Determined by chiral-phase HPLC analysis for compounds 10a–10j.

[2] Draper J A, Britton R. Org Lett, 2010, 12: 4034 [3] Yoshimitsu T, Nakatani R, Kobayashi A, Tanaka T. Org Lett, 2011,

13: 908 [4] Kang B, Britton R. Org Lett, 2007, 9: 5083 [5] Pajkert R, Kolomeitsev A A, Milewska M, Rӧschenthaler G V, Koro-

niak H. Tetrahedron Lett, 2008, 49: 6046 [6] Corey E J, Helal C J. Angew Chem Int Ed, 1998, 37: 1986 [7] Ohkuma T, Tsutsumi K, Utsumi N, Arai N, Noyori R, Murata K. Org

Lett, 2007, 9: 255 Knoevenagel E. Chem Ber, 1896, 29: 172 Jones G. Org React, 1967, 15: 204 Ramachary D B, Kishor M. J Org Chem, 2007, 72: 5056 Ramachary D B, Ramakumar K, Narayana V V. J Org Chem, 2007, 72: 1458 [12] Kumar A, Maurya R A. Tetrahedron, 2007, 63: 1946 [8] [9] [10] [11]

1100

P. B. Thorat et al. / Chinese Journal of Catalysis 36 (2015) 1093–1100

Graphical Abstract Chin. J. Catal., 2015, 36: 1093–1100

doi: 10.1016/S1872-2067(14)60317-X

Stereoselective synthesis of vic-halohydrins and an unusual Knoevenagel product from an organocatalyzed aldol reaction: A non-enamine mode P. B. Thorat, S. V. Goswami, V. P. Sondankar, S. R. Bhusare * Dnyanopasak College, India

Stereoselective synthesis of vic-halohydrins and an unusual Knoevenagel product were obtained using a newly synthesized chiral organocatalyst and triethylamine. The reaction gave the products in excellent yield with high anti selectivity and enantioselectivity.

[13] Ramachary D B, Anebouselvy K, Chowdari N S, Barbas C F III. J Org [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25]

Chem, 2004, 69: 5838 List B, Lerner R A, Barbas C F III. J Am Chem Soc, 2000, 122: 2395 Guillena G, Najera C, Ramon D J. Tetrahedron: Asymmetry, 2007, 18: 2249 Seayad J, List B. Org Biomol Chem, 2005, 3: 719 Dondoni A, Massi A. Angew Chem Int Ed, 2008, 47: 4638 Ferreira F, Botuha C, Chemla F, Perez-Luna A. Chem Soc Rev, 2009, 38: 1162 Morton D, Stockman R A. Tetrahedron, 2006, 62: 8869 Robak M T, Herbage M A, Ellman J A. Chem Rev, 2010, 110: 3600 Zhou P, Chen B C, Davis F A. Tetrahedron, 2004, 60: 8003 Robak M T, Trincado M, Ellman J A. J Am Chem Soc, 2007, 129: 15110 Thorat P B, Goswami S V, Khade B C, Bhusare S R. Tetrahedron: Asymmetry, 2012, 23: 1320 Umehara A, Kanemitsu T, Nagata K, Itoh T. Synlett, 2012, 23: 453 Liu Y, Liang J, Liu X H, Fan J C, Shang Z C. Chin Chem Lett, 2008, 19: 1043

[26] Sun Q, Shi L X, Ge Z M, Cheng T M, Li R T. Chin J Chem, 2005, 23:

745 [27] Yue C B, Mao A Q, Wei Y Y, Li M J. Catal Commun, 2008, 9: 1571 [28] Yadav J S, Reddy B V S, Basak A K, Visali B, Narsaiah A V, Nagaiah

K. Eur J Org Chem, 2004: 546 [29] Narsaiah A V, Nagaiah K. Synth Commun, 2003, 33: 3825 [30] Wang S, Ren Z, Cao W, Tong W. Synth Commun, 2001, 31: 673 [31] Lai Y F, Zheng H, Chai S J, Zhang P F, Chen X Z. Green Chem, 2010,

12: 1917 [32] Thorat P B, Goswami S V, Magar R L, Patil B R, Bhusare S R. Eur J

Org Chem, 2013, 2013: 5509 [33] Thorat P B, Goswami S V, Jadhav W N, Bhusare S R. Aust J Chem,

2013, 66: 661 [34] Thorat P B, Goswami S V, Magar R L, Bhusare S R. J Chem Sci, 2013,

125: 1109 [35] Thorat P B, Goswami S V, Bhusare S R. Tetrahedron: Asymmetry,

2013, 24: 1324 [36] Thorat P B, Goswami S V, Khade B C, Bhusare S R. Tetrahedron

Lett, 2012, 53: 6083