nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

6
Short Communication Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane dichloride silica as a novel nanomagnetic phase-transfer catalyst for the aqueous synthesis of benzyl acetates and thiocyanates Jamal Davarpanah, Ali Reza Kiasat Chemistry Department, College of Science, Shahid Chamran University, Ahvaz 61357-4-3169, Iran abstract article info Article history: Received 21 June 2013 Received in revised form 24 July 2013 Accepted 26 July 2013 Available online 2 August 2013 Keywords: Organicinorganic hybrid silica Magnetite composite Double-charged hybrid silica nucleophilic substitution reactions nucleophilic substitution reactions Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane dichloride silica hybrid (Fe 3 O 4 @SiO 2 /DABCO) was used as an efcient and magnetically recoverable phase-transfer catalyst (PTC) for nucleophilic substitution reac- tions of benzyl halides for the synthesis of benzyl acetates and thiocyanates in good to excellent yields at 100 °C in water. No evidence for the formation of by-products, for example, isothiocyanate or benzyl alcohol was observed and the products were obtained in pure form without further purication. The catalyst was easily separated with the assistance of an external magnetic eld from the reaction mixture and reused for several consecutive runs without signicant loss of its catalytic efciency. © 2013 Elsevier B.V. All rights reserved. 1. Introduction Hybrid xerogel materials, where the organic component is bonded to a polymeric silica skeleton framework, have attracted signicant at- tention over the last decade. In this context, single- and double-charged silica-based hybrid xerogels containing the 1-azonia-4-azabicyclo [2.2.2]octane, DABCO, chloride group have recently drawn particular interest [14]. Ionic liquids (ILs) have been widely used in organic synthesis and cat- alytic reactions as solvents and catalysts. The practical utility of ILs as a catalyst in aqueous media could be extended further if it can be rendered water insoluble. The chemical industry still prefers to use heterogeneous catalyst systems because of the ease of separation, thus supported ILs are highly desirable [5]. Immobilization of ILs on solid supports has been suggested as an alternative way of getting around the problems. Due to the high surface area to volume ratios, superparamagnetic properties and biocompatibility, magnetic nanoparticles (MNPs) can effectively improve the loading and catalytic efciency of immobilized catalysts. They often encompass the desirable features of both organic and inorganic compounds [611]. More than these, they can be easily separated by the simple application of an external magnetic eld, which may optimize operational costs and enhance a products purity [12]. Iron oxides, magnetite (Fe 3 O 4 ), and maghemite (γ-Fe 2 O 3 ) are by far the most used magnetic nanoparticles because they are much less toxic than their metallic counterparts, and they still have high saturation magnetization and superparamagnetic behavior, among which magne- tite is a very promising choice due to its already proven biocompatibility. In addition, previous investigations have shown that magnetite has rel- atively high LD50 values (400 mg/kg in rats) and polymer-coated mag- netite has not shown any acute or subacute toxicity in animal studies [13,14]. Phase-transfer catalysis (PTC) is a very useful technique, which is widely used in the synthesis of dyes, pharmaceutics/perfumes, avorants, agricultural chemicals, monomers, and polymers, and it has been widely used for organic synthesis, particularly for nucleo- philic substitution [1518]. The design of efcient and recoverable phase-transfer catalysts has become an important issue for reasons of economic and environmental impact, in recent years. In particular, immobilization of the phase-transfer catalyst on a support has con- siderable advantages, including easy catalyst recovery and product isolation, and employment of a continuous ow method owing to the three-phase nature of the system, which make the technique at- tractive for industrial applications. Although many phase-transfer catalysts are known, quaternary salts formed from ammonia are practically important and used in many of organic reactions. Its high polarity and ability to solubilize both organic and inorganic compounds can result in enhanced reaction rates and can provide higher selectivity compared with conventional methods [1922]. Considering these facts and taking all this into account, we decided to use supported nanomagnetic double-charged diazoniabicyclo[2.2.2] octane dichloride silica hybrid (Fe 3 O 4 @SiO 2 /DABCO) as an efcient nanomagnetic phase transfer system for organic transformation. Catalysis Communications 42 (2013) 98103 Corresponding author. Tel./fax: +98 611 3331746. E-mail address: [email protected] (A.R. Kiasat). 1566-7367/$ see front matter © 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.catcom.2013.07.040 Contents lists available at ScienceDirect Catalysis Communications journal homepage: www.elsevier.com/locate/catcom

Upload: others

Post on 08-Jan-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

Catalysis Communications 42 (2013) 98–103

Contents lists available at ScienceDirect

Catalysis Communications

j ourna l homepage: www.e lsev ie r .com/ locate /catcom

Short Communication

Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane dichloridesilica as a novel nanomagnetic phase-transfer catalyst for the aqueoussynthesis of benzyl acetates and thiocyanates

Jamal Davarpanah, Ali Reza Kiasat ⁎Chemistry Department, College of Science, Shahid Chamran University, Ahvaz 61357-4-3169, Iran

⁎ Corresponding author. Tel./fax: +98 611 3331746.E-mail address: [email protected] (A.R. Kiasat).

1566-7367/$ – see front matter © 2013 Elsevier B.V. All rihttp://dx.doi.org/10.1016/j.catcom.2013.07.040

a b s t r a c t

a r t i c l e i n f o

Article history:Received 21 June 2013Received in revised form 24 July 2013Accepted 26 July 2013Available online 2 August 2013

Keywords:Organic–inorganic hybrid silicaMagnetite compositeDouble-charged hybrid silicanucleophilic substitution reactionsnucleophilic substitution reactions

Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane dichloride silica hybrid (Fe3O4@SiO2/DABCO) wasused as an efficient andmagnetically recoverable phase-transfer catalyst (PTC) for nucleophilic substitution reac-tions of benzyl halides for the synthesis of benzyl acetates and thiocyanates in good to excellent yields at 100 °C inwater. No evidence for the formation of by-products, for example, isothiocyanate or benzyl alcohol was observedand the products were obtained in pure formwithout further purification. The catalyst was easily separatedwiththe assistance of an external magnetic field from the reaction mixture and reused for several consecutive runswithout significant loss of its catalytic efficiency.

© 2013 Elsevier B.V. All rights reserved.

1. Introduction

Hybrid xerogel materials, where the organic component is bondedto a polymeric silica skeleton framework, have attracted significant at-tention over the last decade. In this context, single- and double-chargedsilica-based hybrid xerogels containing the 1-azonia-4-azabicyclo[2.2.2]octane, DABCO, chloride group have recently drawn particularinterest [1–4].

Ionic liquids (ILs) have beenwidely used in organic synthesis and cat-alytic reactions as solvents and catalysts. The practical utility of ILs as acatalyst in aqueousmedia could be extended further if it can be renderedwater insoluble. The chemical industry still prefers to use heterogeneouscatalyst systems because of the ease of separation, thus supported ILs arehighly desirable [5]. Immobilization of ILs on solid supports has beensuggested as an alternative way of getting around the problems.

Due to the high surface area to volume ratios, superparamagneticproperties and biocompatibility, magnetic nanoparticles (MNPs) caneffectively improve the loading and catalytic efficiency of immobilizedcatalysts. They often encompass the desirable features of both organicand inorganic compounds [6–11]. More than these, they can be easilyseparated by the simple application of an external magnetic field,which may optimize operational costs and enhance a product’s purity[12]. Iron oxides, magnetite (Fe3O4), and maghemite (γ-Fe2O3) are byfar the most used magnetic nanoparticles because they are much less

ghts reserved.

toxic than theirmetallic counterparts, and they still have high saturationmagnetization and superparamagnetic behavior, among which magne-tite is a very promising choice due to its already proven biocompatibility.In addition, previous investigations have shown that magnetite has rel-atively high LD50 values (400 mg/kg in rats) and polymer-coated mag-netite has not shown any acute or subacute toxicity in animal studies[13,14].

Phase-transfer catalysis (PTC) is a very useful technique, which iswidely used in the synthesis of dyes, pharmaceutics/perfumes,flavorants, agricultural chemicals, monomers, and polymers, and ithas been widely used for organic synthesis, particularly for nucleo-philic substitution [15–18]. The design of efficient and recoverablephase-transfer catalysts has become an important issue for reasonsof economic and environmental impact, in recent years. In particular,immobilization of the phase-transfer catalyst on a support has con-siderable advantages, including easy catalyst recovery and productisolation, and employment of a continuous flow method owing tothe three-phase nature of the system, which make the technique at-tractive for industrial applications. Although many phase-transfercatalysts are known, quaternary salts formed from ammonia arepractically important and used in many of organic reactions. Itshigh polarity and ability to solubilize both organic and inorganiccompounds can result in enhanced reaction rates and can providehigher selectivity compared with conventional methods [19–22].

Considering these facts and taking all this into account, we decidedto use supported nanomagnetic double-charged diazoniabicyclo[2.2.2]octane dichloride silica hybrid (Fe3O4@SiO2/DABCO) as an efficientnanomagnetic phase transfer system for organic transformation.

Page 2: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

Table 1Nucleophilic substitution reaction of benzyl halides with acetate anions in water catalyzed by Fe3O4@SiO2/DABCO hybrid.

Entry Benzyl halide Product Time (h) Yield (%)

1 1.5 85

2 1.5 86

3 2 83

4 1.5 83

5 1.5 83

6 2 78

7 4 74

8 4 76

9 2.5 80

99J. Davarpanah, A.R. Kiasat / Catalysis Communications 42 (2013) 98–103

2. Experimental

2.1. General

Iron (II) chloride tetrahydrate (99%), iron (III) chloride hexahydrate(98%), benzyl halides and other chemical materials were purchasedfrom Fluka andMerck companies and usedwithout further purification.Products were characterized by comparison of their physical data, suchas IR and 1H NMR and 13C NMR spectra, with known samples. NMRspectra were recorded in CDCl3 on a Bruker Advance DPX 400 MHzinstrument spectrometer using TMS as an internal standard. IR spectrawere recorded on a BOMEM MB-Series 1998 FT-IR spectrometer. Thepurity determination of the products and reaction monitoring wereaccomplished by TLC on silica gel PolyGram SILG/UV 254 plates. Theparticlemorphologywas examined by SEM(Philips XL30 scanning elec-tron microscope) and TEM (Zeiss; EM10C; 80 kV).

2.2. Preparation of Fe3O4 superparamagnetic nanoparticles

Superparamagnetic nanoparticles (MNPs) were prepared via im-proved chemical coprecipitation method [23]. According to this meth-od, FeCl2·4H2O (6.346 g, 31.905 mmol) and FeCl3·6H2O (15.136 g,55.987 mmol) were dissolved in 640 mL of deionized water. The

mixed solution was stirred under N2 at 90 °C for 1 h. Eighty millilitersof NH3.H2O (25%) was injected into the reactionmixture rapidly, stirredunder N2 for another 1 h and then cooled to room temperature. Theprecipitated particles were washed five times with hot water and sepa-rated by magnetic decantation. Finally, magnetic NPs were dried undervacuum at 70 °C.

2.3. Synthesis of bis(n-propyltrimethoxysilane)-1,4-diazoniabicycle [2.2.2]octane chloride, BPTDABCOCl

BPTDABCOCl was prepared according to the procedure described byArenas et al. [3]. Initially, the previously sublimed DABCO (0.897 g,8.0 mmol) was dissolved in DMF (5 mL). To this solution, 16 mmol of3-chloropropyltrimethoxysilane (CPTMS) was added. The mixture wasstirred for 72 h, under argon atmosphere at 90 °C. The white solid,BPTDABCOCl, was filtered and washed with methanol and then driedfor 2 h in an oven at 90 °C.

2.4. Preparation of magnetic double-charged diazoniabicyclo[2.2.2]octanedichloride silica hybrid, Fe3O4@SiO2/DABCO

Fe3O4@SiO2/DABCO was prepared using a sol–gel process. The syn-thesized magnetite nanoparticles (2.0 g) was first diluted with water

Page 3: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

Scheme 1. Synthesis of Fe3O4@SiO2/DABCO.

100 J. Davarpanah, A.R. Kiasat / Catalysis Communications 42 (2013) 98–103

(40 mL), absolute ethanol (120 mL) and 3.0 mL ammonia aqueous(25%). This suspension was well-dispersed by ultrasonic vibration for15 min. To this dispersed suspension, BPTDABCOCl (5 g), previouslydissolved in DMF (7 mL), was added. Then under continuous mechani-cal stirring, 1.5 mLof TEOSdiluted in ethanol (40 mL)was slowly addedto this dispersion, and after stirring for 48 h, the obtained magneticnanocomposite, Fe3O4@SiO2/DABCO silica was collected by magneticseparation and washed with ethanol.

2.5. Typical procedure for the conversion of benzyl halides to the corre-sponding benzyl acetates and thiocyanates

To a suspension of alkyl halide (1 mmol) and nucleophilic reagents(NaOAc or KSCN) (2 mmol) in water (5 mL), Fe3O4@SiO2/DABCO(0.225 g) was added and the mixture was stirred at 90 °C for thelengths of time shown in Table 1. After completion of the reaction as in-dicated by TLC [using n-hexane/ethyl acetate (5:1)], the reaction wasallowed to cool to room temperature and the magnetic catalyst wasconcentrated on the sidewall of the reaction vessel using an externalmagnet. The reaction mixture residue was poured into water (10 mL)

Scheme 2. Postulated roles of Fe3O4@SiO2/DABCO in the nucleophilic substitution.

and extracted with EtOAc (3 × 10 mL). The combined organic layerswere dried over anhydrous Na2SO4, filtered and evaporated in vacuoto give the product in 74% to 91% isolated yields.

3. Results and discussion

Recently, we focused on developing novel PTC systems for organ-ic transformations [24–28] and also successfully performed the one-pot synthesis of pyran annulated heterocyclic compounds in waterwith nanomagnetic double-charged diazoniabicyclo[2.2.2]octanedichloride silica hybrid (Fe3O4@SiO2/DABCO). Supported nanocom-posite (Fe3O4@SiO2/DABCO) was prepared using Fe3O4 spheres asthe core and the positively double-charged organic-inorganic hybridsilica as the shell (Scheme 1) [29].

In the continuation of our study about developing novel PTCsystems for organic transformations, herein, we decided to useFe3O4@SiO2/DABCO as a new magnetic phase transfer system forthe facile preparation of benzyl acetates and thiocyanates in waterby nucleophilic substitution reaction (Scheme 2).

10

20

30

40

50

400900140019002400290034003900

Tra

nsm

issi

on %

Wavenumber (cm-1)

Fig. 1. The FT-IR spectra of Fe3O4, Fe3O4@SiO2/DABCO and BPTDABCOCl.

Page 4: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

Table 2Nucleophilic substitution reaction of benzyl halides with thiocyanate anions in water catalyzed by Fe3O4@SiO2/DABCO hybrid.

Entry Benzyl halide Product Time (h) Yield (%)

1 1 88

2 1 91

3 1.5 84

4 1.5 83

5 1 85

6 1 85

7 3 83

8 3 80

9 2 81

101J. Davarpanah, A.R. Kiasat / Catalysis Communications 42 (2013) 98–103

First, the Fe3O4@SiO2/DABCO was prepared according to our report-ed method as shown in Scheme 1. The catalyst has been characterizedby scanning electron microscope (SEM), vibrating sample magnetome-ter (VSM), thermogravimetric analysis (TGA) and differential thermalanalysis (DTA) (see the Supplementary Materials for Figs. S1–S3), Fou-rier transform infrared spectroscopy (FT-IR) (Fig. 1), X-ray diffraction(XRD) (Fig. 2), transmission electron microscopy (TEM) (Fig. 3) andby their comparisons with that of authentic sample.

The loading of the catalyst was calculated by TGA analysis(Fig. S3a). The amount of DABCO is 0.5 mmol per gram in Fe3O4@SiO2/DABCO. The FT-IR analysis of Fe3O4@SiO2/DABCO exhibitsbasic characteristic peak at approximately 580 cm−1, which was at-tributed to the presence of Fe–O stretching vibrations. The presenceof peaks at 440, 780 and 990 to 1200 cm−1 were most probably dueto the symmetric and asymmetric stretching vibrations of frame-work and terminal Si–O groups. The C–H stretching peaks at 1200to 1500 and 2870 to 3040 cm−1 in the Fe3O4@SiO2/DABCO spectraindicate that the magnetic particles are successfully coated bydouble-charged diazoniabicyclo[2.2.2]octane chloride silica hybrid.A band at 1465 cm−1 is characteristic for the tertiary amine group.

The degree of crystallinity of magnetic iron oxide (Fe3O4) in the syn-thesized Fe3O4@SiO2/DABCO was obtained from XRD measurement(Fig. 2). According to the database of the Joint Committee on PowderDiffraction Standards (JCPDS; JCPDs, 19-629), the XRD pattern of a

standard Fe3O4 crystal with spinel structure has six characteristicpeaks at 2 = 30.5°, 35.8°, 43.4°, 53.8°, 57.5°, and 63.1° [30]. The patternof Fe3O4@SiO2/DABCO displays aremost intense at 2 = 35.83. This line,which corresponds to pure Fe3O4 [31], confirms its presence. It is alsoapparent that the results of the analysis of Fe3O4@SiO2/DABCO fit thepattern exhibited by standardmagnetite. Therefore, it can be concludedthat the obtained Fe3O4@SiO2/DABCO show spinel structure and thesemodifications don’t cause a phase change in Fe3O4 [14].

TEMmicrographs providemore accurate information on the particlesize andmorphology of immobilization silica parts on themagnetic sur-face. An image of the composite displays darkMNP cores surrounded byan amorphous silica layer (Fig. 3).

In order to investigate the possible catalytic properties of Fe3O4@SiO2/DABCO in the nucleophilic substitution reaction, the reaction of benzylhalides with acetate anion in water was chosen. Examination of thecatalytic activity catalyst indicated that the use of 2 equiv of NaOAc inthe presence of Fe3O4@SiO2/DABCO (0.225 g) in water at 100 °C is thebest condition for the conversion of benzyl halides to the correspondingbenzyl acetates (Scheme 2; Table 1). 1H NMR spectra of the crude prod-ucts clearly showed the formation of acetates and no evidence for thehydrolysis of benzyl halides to the alcohols was observed, which provedthat the reactions proceeded cleanly. On the other hand, in the absence ofFe3O4@SiO2/DABCO, the reaction was sluggish and, even after prolongedreaction time, a considerable amount of starting material remained.

Page 5: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

Fig. 2. XRD pattern of Fe3O4@SiO2/DABCO.

102 J. Davarpanah, A.R. Kiasat / Catalysis Communications 42 (2013) 98–103

Moreover, the reactionmixture was contaminated with alcohol. This ob-servation was confirmed with the presence of alcohol on the TLC plate.

With these promising results in hand and establishing the advan-tages of Fe3O4@SiO2/DABCO as a phase-transfer catalyst, we focusedour attention on another nucleophilic substitution reaction, the conver-sion of benzyl halides to the corresponding thiocyanates (Scheme 2;Table 2).

This catalyst acted very efficiently and, in all cases, a very clean reac-tion was observed. It is noteworthy that no evidence for the formationof by-products such as alcohols or isothiocyanates was observed andthe products were obtained in pure form without further purification.13C resonance of the −SCN and −NCS groups at ~111 and ~145 ppm,respectively, are very characteristic for thiocyanate and isothiocyanatefunctionalities [14].

The structures of all of the benzyl acetate and thiocyanate productswere determined from their analytical and spectral (IR, 1H and 13CNMR) data and by direct comparison with authentic samples (Fig. S4).

The high polarity of the catalysts (double-charged silica-basedhybrid xerogel containing the 1-azonia-4-azabicyclo[2.2.2]octane,DABCO, chloride group) and the ability to solubilize both organicand inorganic compounds can result in enhanced reaction rates andprovide higher selectivity compared with conventional methods.

As shown in Tables 1 and 2, the compounds with both electron do-nating and withdrawing groups approximately participated in the nu-cleophilic reaction with equal efficiency. Thus, the nature and positionof substitution in the aromatic ring did not affect the reactions much.Themeta-substituted aromatic, as well as the sterically hindered ortho-

Fig. 3. TEM image of Fe3O4@SiO2/DABCO.

substituted aromatic aldehydes, both undergo condensation reactionwithout any difficulty.

It is worthy to note that Fe3O4@SiO2/DABCO does not suffer fromextensive mechanical degradation after operating and could be quanti-tatively recovered without filtration since the Fe3O4@SiO2/DABCO wasrapidly concentrated as soon as an external magnet was set close tothe sidewall of the reaction vessel. The residual catalyst was washedwith water and methanol and dried, then immediately reused for thenext run. The recovered resinwas reused four times for the nucleophilicsubstitution reaction of benzyl bromide with thiocyanate anion. Theresults showed that the catalyst does not reveal any loss in its activityand produced benzyl thiocyanate in 88%, 85%, 84% and 84% yield,respectively.

4. Conclusion

In the present work, the performance of Fe3O4@SiO2/DABCO as asolid–liquid phase-transfer catalyst for nucleophilic substitution reac-tions of benzyl halides in water was investigated. The catalytic systemcan combine the advantages of homogeneous and heterogeneouscatalysts and therefore they can be selective, reactive and recyclable.In conclusion, we have developed an easy to operate, safe and cost-effective method for the preparation of benzyl acetates and thiocya-nates in water by nucleophilic substitution.

Acknowledgments

We are grateful to the Research Council of Shahid Chamran Univer-sity for financial support.

Appendix A. Supplementary data

Supplementary data to this article can be found online at http://dx.doi.org/10.1016/j.catcom.2013.07.040.

References

[1] L.T. Arenas, T.A.S. Aguirre, A. Langaro, Y. Gushikem, E.V. Benvenutti, T.M.H. Costa,Polymer 44 (2003) 5521.

[2] L.T. Arenas, S.L.P. Dias, C.C. Moro, T.M.H. Costa, E.V. Benvenutti, A.M.S. Lucho, Y.Gushikem, J. Colloid Interface Sci. 297 (2006) 244.

[3] L.T. Arenas, E.C. Lima, A.A. dos Santos Jr, J.C.P. Vaghetti, T.M.H. Costa, E.V. Benvenutti,Colloids Surf. A 297 (2007) 240.

[4] L.T. Arenas, D.S.F. Gay, C.C. Moro, S.L.P. Dias, D.S. Azambuja, T.M.H. Costa, E.V.Benvenutti, Y. Gushikem, Microporous Mesoporous Mater. 112 (2008) 273.

[5] Y. Zhang, Y. Zhao, C. Xia, J. Mol. Catal. A Chem. 306 (2009) 107.[6] S. Ghosh, A.Z.M. Badruddoza, M.S. Uddin, K. Hidajat, J. Colloid Interface Sci. 354

(2011) 483.[7] Z. Lei, Y. Li, X. Wei, J.Solid State Chem. 181 (2008) 480.[8] R. Cano, D.J. Ramon, M. Yus, Tetrahedron 67 (2011) 5432.

Page 6: Nanomagnetic double-charged diazoniabicyclo[2.2.2]octane

103J. Davarpanah, A.R. Kiasat / Catalysis Communications 42 (2013) 98–103

[9] M.B. Gawande, A.K. Rathi, P.S. Branco, I.D. Nogueira, A. Velhinho, J.J. Shrikhande, U.U.Indulkar, R.V. Jayaram, C.A. Ghumman, N. Bundaleski, O.M.N.D. Teodoro, Chem. Eur.J. 18 (2012) 12628.

[10] M.B. Gawande, P.S. Branco, I.D. Nogueira, C.A. Ghumman, N. Bundaleski, A. Santos,O.M.N.D. Teodoro, R. Luque, Green Chem. 15 (2013) 682.

[11] M.B. Gawande, P.S. Branco, R.S. Varma, Chem. Soc. Rev. 42 (2013) 3371.[12] L. Ma’mani, M. Sheykhan, A. Heydari, M. Faraji, Y. Yamini, Appl. Catal. Gen. 377

(2010) 64.[13] D.J. Widder, W.L. Greif, K.J. Widder, R.R. Edelman, T.J. Brady, Am. J. Roentgenol. 148

(1987) 399.[14] A.R. Kiasat, S. Nazari, J. Mol. Catal. A Chem. 365 (2012) 80.[15] H. Mahdavi, B. Tamami, React. Funct. Polym. 64 (2005) 179.[16] S. Xue, S. Ke, T. Wei, L. Duan, Y. Guo, J. Chin. Chem. Soc. 51 (2004) 983.[17] E. Aly, M. Abdo, A. El-Gharably, J. Chin. Chem. Soc. 51 (2004) 1013.[18] H. Mahdavi, E. Haghani, B. Malakian, J. Chin. Chem. Soc. 66 (2006) 1033.

[19] Z.H. Weng, J.Y. Wang, X.G. Jian, Chin. Chem. Lett. 18 (2007) 936.[20] M.J.O. Donnell, Asymmetric phase transfer reactions, in: I. Ojima (Ed.), Catalytic

Asymmetric SynthesisVerlag Chemie, New York, 2000.[21] S. Mallakpour, Z. Rafiee, Eur. Polym. J. 43 (2007) 1510.[22] B.X. Tang, F. Wang, J.H. Li, Y.X. Xie, M.B. Zhang, J. Org. Chem. 72 (2007) 6294.[23] K.D. Kim, S.S. Kim, Y.H. Choa, H.T. Kim, J. Ind. Eng. Chem. 13 (2007) 1137.[24] A.R. Kiasat, M. Fallah Mehrjardi, Catal. Commun. 9 (2008) 1497.[25] A.R. Kiasat, R. Badri, B. Zargar, S. Sayyahi, J. Org. Chem. 73 (2008) 8382.[26] A.R. Kiasat, M. Zayadi, Catal. Commun. 9 (2008) 2063.[27] A.R. Kiasat, S. Sayyahi, Mol. Divers. 14 (2010) 155.[28] A.R. Kiasat, S. Nazari, Catal. Commun. 18 (2012) 102.[29] J. Davarpanah, A.R. Kiasat, S. Noorizadeh, M. Ghahremani, J. Mol. Catal. A 376

(2013) 78.[30] Y. Lin, H. Chen, K. Lin, B. Chen, C. Chiou, J. Environ. Sci. 23 (2011) 44.[31] E.V. Stoyanov, I.C. Ivanov, D. Heber, Molecules 5 (2000) 19.