Synthesis of a copper(II) complex covalently anchoring a (2-iminomethyl)phenol moiety supported on HAp-encapsulated-α-Fe2O3 as an inorganic–organic hybrid magnetic nanocatalyst for the synthesis of primary and secondary amides

M. Mamaghani*, F. Shirini, M. Sheykhan and M. Mohsenimehr
Department of Chemistry, Faculty of Sciences, University of Guilan, P. O. Box 41335-1914, Rasht, Iran. E-mail: m-chem41@guilan.ac.ir; Fax: +98 13 33333262; Tel: +98 13 33770899

Received 6th March 2015 , Accepted 8th May 2015

First published on 8th May 2015


Abstract

A novel hydroxyapatite-encapsulated-α-Fe2O3-based Cu(II) organic–inorganic hybrid (interphase) catalyst was prepared. The prepared nanocatalyst provided an efficient, useful and green method for the oxidative amidation of aromatic aldehydes with ammonium hydrochloride and aniline hydrochloride, in short reaction times and good yields. The magnetic nature of the catalyst led to its easy recovery by an external magnetic field and convenient reuse.


Introduction

Homogeneous catalysts have some advantages due to their uniform and well defined reactive centers which lead to their high activity and selectivity. Also, homogeneous catalysts allow for more effective mixing of the catalyst with the reaction mixture (occupying the same phase as the reactants), which in turn creates problems such as catalyst separation from the reaction mixture and therefore the catalyst is difficult to reuse. To solve this problem, several methods have been checked out to combine the advantages of homogeneous and heterogeneous catalysis.1 In this respect, nanoparticles (NPs), especially magnetic NPs (MNPs) have appeared as a bridge between homogeneous and heterogeneous catalysts. The magnetic nature of these particles allows for easy recovery (green chemistry) and recycling of the catalysts by an external magnestic field.2 To improve the chemical stability as well as to achieve some advantages as the versatility in surface modification, MNPs commonly coat with different organic or inorganic compounds, precious metals and recently calcium hydroxyapatite Ca10(PO4)6(OH)2 (HAp).3

Covalently combining an active site onto a large surface area solid through a flexible spacer to create an organic–inorganic hybrid (interphase) catalyst is known as a way to simulate homogeneous reaction conditions together with having the advantage of recyclability.1,4,5

On the other hand, amide bonds have been found in many natural important polymers such as peptides and proteins and in a wide range of biological compounds, such as pharmaceuticals and is also employed as synthetic intermediates for fine chemical industry and agrochemicals.6,7 Commonly, the amide bond is synthesized by acylation of amines with either activated carboxylic acids or coupling with carboxylic acids mediated by a coupling reagent.8,9 However, these methods have the innate problems of producing waste products, use of hazardous reagents and poor atom economy.10 To solve these problems in amide synthesis, novel methods were developed, such as direct amide synthesis from alcohols with amines or nitroarenes,11 amino carbonylation of haloarenes,12 hydroamination of alkynes, amidation of thioacids with azides,13 transamidation of primary amides,14 and C–H oxidative amidation.15

Among these, oxidative amidation of aldehydes with amine salts is synthetically useful and less hazardous than those traditionally used such as acid halides, which was first reported by Nakagawa et al. in 1966 using stoichiometric amounts of nickel peroxide as the oxidant.16 Subsequently, other researchers have used new oxidants and novel catalysts for the direct conversion of aldehydes to amides.

More recently, Yoo and Li introduced a copper-catalyzed procedure for oxidative amidation of aldehydes with amine hydrochlorides using CuI–AgIO3 as the catalyst and tert-butyl hydroperoxide as the oxidant.17

In the years since, other copper catalysts such as Cu(OAc)2 and CuSO4·5H2O have been reported for the amidation reactions.18 Silica-coated magnetic carbon nanotubes (MagCNTs@SiO2)-immobilized imine–Cu(I) was also introduced as a magnetic heterogeneous catalyst for the synthesis of different types of amides.19 Although, significant advances have been achieved in this field, but introduction of an efficient, economical and greener catalyst is still desirable.

On the basis of the above concerns, we report herein a novel copper(II) complex covalently anchoring (2-iminomethyl)phenol moiety supported on HAp-encapsulated-α-Fe2O3 (denoted as catalyst 4) (Scheme 1) as an inorganic–organic hybrid magnetic nanocatalyst for efficient oxidative amidation of aromatic aldehydes with amine hydrochloride and aniline hydrochloride.


image file: c5ra03977k-s1.tif
Scheme 1 (i) Ca(NO3)2, (NH4)2HPO4, NH4OH, (ii) calcination, (iii) 3-aminopropyltrimethoxysilane, toluene, reflux, 48 h, (iv) 2-hydroxy benzaldehyde, dry EtOH, reflux, 24 h, (v) CuCl2, EtOH, r.t., 12 h.

Results and discussion

In continuation of our research devoted to the development of new benign methodologies for the preparation of magnetic nanoparticles,3a,20 we were interested in the synthesis of amide derivatives using nanocatalyst 4 (Scheme 1). The magnetic nanoparticles (Fe3O4) were synthesized by a chemical co-precipitation technique using ferric and ferrous ions. Based on our previous report,3a after coating a layer of hydroxyapatite on the surface of the Fe3O4 nanoparticles and its calcination at 450 °C, it was functionalized by 3-aminopropyltrimethoxysilane to produce an organic–inorganic hybrid. Then, the prepared hybrid was reacted with 2-hydroxybenzaldehyde to obtain a bi-dentate ligand functionalized core–shell magnetic nanocatalyst (named as catalyst 3). Finally, surface-bound ligands were complexed with copper(II) by using CuCl2 (Scheme 1). The nanocatalyst 4 was characterized by various microscopic and spectroscopic techniques such as XRD, FT-IR, ICP, SEM-EDX, VSM. In the FT-IR spectrum (Fig. 1), the absorption peak at 464 cm−1 can be assigned to the Fe–O absorption in the hematite.21 The O–P–O surface phosphate groups in the hydroxyapatite shell appeared at 570 and 604 cm−1 which were in overlap with Fe–O stretching. Also, the P–O stretching appeared at 1040 cm−1. Signals at around 1419 and 1458 cm−1 related to the aromatic C[double bond, length as m-dash]C stretching as well as a signal at 1629 cm−1 produced by C[double bond, length as m-dash]N double bond confirm that the organic linker was properly immobilized onto the surface. Note that the C[double bond, length as m-dash]N signal is shifted to a lower wavenumber than the C[double bond, length as m-dash]N stretching frequency of 3 (1629 cm−1 rather than 1634 cm−1). It is implying that C[double bond, length as m-dash]N bond is coordinated to copper through the lone pair of nitrogen and oxygen. In XRD analysis (Fig. 2), the resulted patterns can be readily referred to α-Fe2O3 (JCPDS no. 33-0664), and hydroxyapatite (JCPDS no. 24-0033). Diffraction peaks at around 28.0°, 38.6°, 41.6°, 47.8°, 58.0°, 63.6°, 66.8°, 68.0°, 72.8° and 74.2° correspond to the (012), (104), (110), (113), (024), (116), (122), (018), (214) and (300) reflections patterns of hematite. No other phase except the hematite and hydroxyapatite is detectable. The measurements were carried out on a Philips X'Pert diffractometer with CoKα radiation. The ICP analysis showed that 4.9% of Cu was anchored on catalyst 4. It means that there is about 0.8 mmol (Cu) per gram of the catalyst.
image file: c5ra03977k-f1.tif
Fig. 1 FT-IR spectrum of catalysts 3 and 4.

image file: c5ra03977k-f2.tif
Fig. 2 XRD spectra of catalyst 4.

The morphology and size details were investigated by SEM measurement (Fig. 3). The SEM image of the catalyst clearly showed the nano size of the particles (with an average about 50 nm) and their morphology is spherical. The SEM image of the catalyst taken after the fifth cycle of the reaction did not show significant change in the morphology or in the size of the catalyst nanoparticles (58 nm) (Fig. 4), which indicates the retention of the catalytic activity after recycling.


image file: c5ra03977k-f3.tif
Fig. 3 SEM image of catalyst 4.

image file: c5ra03977k-f4.tif
Fig. 4 SEM image (after the fifth cycle of the reaction) of catalyst 4.

The EDX spectrum of the nanocatalyst is presented in Fig. 5, atoms such as Fe, Si, Ca, P, Cu and Cl related to the catalyst structure, are seen in the spectrum. Based on the EDX spectrum the copper[thin space (1/6-em)]:[thin space (1/6-em)]chlorine[thin space (1/6-em)]:[thin space (1/6-em)]nitrogen content of the catalyst were found to be in ratio of 1.2[thin space (1/6-em)]:[thin space (1/6-em)]2.03[thin space (1/6-em)]:[thin space (1/6-em)]1.0.


image file: c5ra03977k-f5.tif
Fig. 5 EDX spectra of catalyst 4.

Magnetic characterization of the nanoparticles, which accounts for its easy recovery, was done using a vibrating sample magnetometer (VSM) at room temperature, with the field sweeping from −8500 to +8500 Oersted. The M(H) hysteresis loop for the sample was completely reversible and the ratio of magnetic remanence (Mr)/magnetic saturation (Ms) is about 0.0005, Ms = 3.71 emu g−1, showing that the nanoparticles exhibit superparamagnetic characteristics (Fig. 6).


image file: c5ra03977k-f6.tif
Fig. 6 VSM curve of catalyst 4.

After characterization, catalytic activity of the catalyst was tested in the oxidative coupling of the aromatic aldehydes with ammonium chloride and aniline hydrochloride. In a preliminary experiment the reaction of p-chlorobenzaldehyde and ammonium chloride was chosen as a model reaction for amidation of aromatic aldehydes and the reaction conditions, such as temperature, reaction time, oxidant and molar ratio of the catalyst to substrate was optimized (Table 1). We found that the reaction of p-chlorobenzaldehyde (1 mmol), ammonium chloride (2 mmol) and tert-butylhydroperoxide (TBHP) (0.4 mL) as oxidant in CH3CN (1 mL) in the presence of nanocatalyst (10 mg, 0.8 mol%) and Na2CO3 (2 mmol) as base, at 50 °C under Ar atmosphere, produce the desired primary amide 5a in 40 min and 78% yield (Table 1, entry 1). As shown in Table 1, entry 1, TBHP, based on the reaction time and yield, was proved as selected oxidant at 50 °C. In comparison, the reaction at room temperature resulted in a pronounced reduction in the yield of product 5a and increasing the reaction time did not change the efficiency of the reaction significantly (Table 1, entry 4).

Table 1 Optimization of reaction conditionsb for synthesis of 4-chlorobenzamide in the presence of nanocatalyst 4
Entry Oxidant Solvent Temperature (°C) Yielda (%) Time (min)
a Isolated yield.b Reaction conditions: p-chlorobenzaldehyde (1 mmol), ammonium chloride (2 mmol), nanocatalyst 4 (10 mg, 0.8 mol%), solvent (1 mL), Na2CO3 (2 mmol), oxidant (0.4 mL), Ar atmosphere.
1 TBHP CH3CN 50 78 40
2 H2O2 CH3CN 50 50 40
3 TBHP CH3CN r.t. 40 40
4 TBHP CH3CN r.t. 45 90
5 TBHP CH3CN r.t. 45 120
6 TBHP H2O + CH3CN (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 50 39 40
7 TBHP THF 50 47 40
8 TBHP DMF 50 24 40
9 TBHP Toluene 50 54 40


The oxidation efficiency of TBHP was also compared against H2O2 (Table 1, entry 2), and the effect of various solvents was examined (Table 1). The results revealed that TBHP in CH3CN produce a better result and in most of the cases it is not necessary to use the chromatographic methods for separation of the products.

The procedure led to a green method for the synthesis of primary and secondary amides (5a–k, 6a–e, Scheme 2, Table 2).


image file: c5ra03977k-s2.tif
Scheme 2 Synthesis of amide derivatives in the presence of nanocatalyst 4.
Table 2 Synthesis of amide derivatives
Entry Ar Product Yielda,b (%) M.p.c
a Isolated yield.b Reaction times 40–42 min.c [ ] reported references.
1 4-ClC6H4 5a 78 177–179 [ref. 16]
2 4-HOCC6H4 5b 76 178–181
3 Thiophen-2-yl 5c 77 126–128 [ref. 22]
4 4-CNC6H4 5d 76 146–148 [ref. 22]
5 Cinnamaldehyde 5e 75 147–149 [ref. 22]
6 4-FC6H4 5f 77 138–140 [ref. 22]
7 4-BrC6H4 5g 78 192–194 [ref. 18]
8 4-H3COC6H4 5h 75 165–167 [ref. 18]
9 3-O2NC6H4 5i 75 141–143 [ref. 22]
10 C6H5 5j 77 128–130 [ref. 24]
11 4-CF3C6H4 5k 74 189–191 [ref. 18]
12 4-ClC6H4 6a 83 192–193 [ref. 23]
13 4-BrC6H4 6b 84 204–206 [ref. 23]
14 C6H5 6c 81 164–166 [ref. 23]
15 Thiophen-2-yl 6d 79 178–180
16 2,4-Cl2C6H3 6e 80 199–201


As shown in the Table 2, aromatic aldehydes having both electron-donating and electron-withdrawing groups were converted to the related amides in good yields. Also in some cases the yield of secondary amides is slightly higher than the primary amides.

Reusability of the catalyst was also re-checked in the preparation of 5a. After completion of the reaction, the nanocatalyst was separated from the reaction medium simply by an external magnetic field, washed with ethyl acetate, dried under vacuum and reused for the next run. After 5 successive runs the catalytic activity of the catalyst was almost remained unchanged (Fig. 7).


image file: c5ra03977k-f7.tif
Fig. 7 The reusability of the catalyst 4 in five runs for the preparation of compound 5a.

On the basis of the reaction conditions, we propose that in the first step (Scheme 3) catalyst 4 reacts with TBHP generating a tert-butylperoxy radical following the mechanism proposed in literature25 and a Cu(I)-anchored on the catalyst surface is re-oxidized by another TBHP. The aniline hydrochloride is then converted into an amino radical by the resulting tert-butyloxy radical. The reaction between aldehyde and tert-butylperoxy lead to an acyl radical, as reported previously.26 Finally, a terminative coupling reaction occurs between the acyl radical and the amino radical to form the desired amide. To confirm the radical pathway of the reaction, it was repeated in the air (absence of argon atmosphere). In these conditions, no product was obtained after 24 h.


image file: c5ra03977k-s3.tif
Scheme 3 The postulated mechanism for the catalytic amidation in the presence of the catalyst 4.

In order to show the efficiency of present protocol, we compared our results for amidation of p-chlorobenzaldehyde with previously reported methods and the results are shown in Table 3. As it is evident from the results this method avoids the disadvantages of some other methods such as longer reaction time and harsh reaction conditions.

Table 3 Comparison of various catalysts for the amidation of p-chlorobenzaldehyde with ammonium chloride
Entry Yield (%) Time Catalyst Reference
1 89.4 4 h Nickel peroxide 16
2 48 4 h Cu2O 18
3 52 6 h MagCNTs@SiO2–linker–CuI 19
4 46 5 h CuCl2 This work
4 78 40 min Catalyst 4 This work


Experimental section

Material and method

Melting points were measured on an Electrothermal 9100 apparatus. FT-IR spectra were determined on a Bruker Alfa FT-IR spectrometer. 1H NMR (400 MHz) and 13C NMR (100 MHz) spectra were recorded on a Bruker DRX-400 in DMSO-d6 as solvent and TMS as an internal standard. Chemical shifts on 1H and 13C NMR were expressed in ppm downfield from tetramethylsilane. Elemental analyses were performed on a Carlo-Erba EA1110CNNO-S analyzer and agreed (within 0.30) with the calculated values. XRD was carried out on a Philips X-Pert diffractometer using Co tube. Scanning electron microphotographs (SEM-EDX) were obtained on a PHILIPS XL30 electron microscope. ICP analysis was accomplished using a VISTA-PRO, CCD simultaneous ICP analyzer. Magnetic properties of catalyst were obtained by vibrating sample magnetometer/Alternating Gradient Force Magnetometer (VSM/AGFM, MDK Co, Ltd, Iran). All the chemicals were purchased from Merck and used without further purification. All solvents used were dried and distilled according to standard procedures.

General procedure for the synthesis of catalyst 4

[α-Fe2O3@HAp–Si–(CH2)3–NH2] was synthesized according to the ref. 3a and 27 with a few modification and then was allowed to react with equimolar amount of 2-hydroxybenzaldehyde in super dry ethanol for 24 h under Ar atmosphere. Then, the solid product was magnetically separated by an external magnet and washed with dry ethanol to produce the catalyst 3 (Scheme 1). A stirring mixture of catalyst 3 and CuCl2 in ethanol were heated at room temperature for 12 h. The resultant light brown precipitate formed was filtered, washed repeatedly with ethanol, and air-dried under vacuum at room temperature to produce catalyst 4.

General procedure for the synthesis of amide derivatives (5a–k, 6a–e)

A mixture of ammonium chloride or aniline hydrochloride (2 mmol), aryl aldehyde (1 mmol), Na2CO3 (0.21 g, 2 mmol) and nanocatalyst (0.01 g, 0.8 mol%) in acetonitrile (1 mL) was stirred at room temperature. Then tert-butyl hydroperoxide (TBHP) (0.4 mL) was added under argon atmosphere. The reaction mixture was stirred at 50 °C for 40 min. The progress of the reaction was monitored by TLC (EtOAc[thin space (1/6-em)]:[thin space (1/6-em)]n-hexane, 1[thin space (1/6-em)]:[thin space (1/6-em)]1). After completion of the reaction, the reaction mixture was diluted with ethyl acetate and the catalyst was easily separated by an external magnet and was washed several times with ethyl acetate. The mixed organic solvents were concentrated under vacuum to furnish the desired products (5a–k, 6a–e). All isolated products were identified by spectroscopic analysis (FT-IR, 1H NMR and 13C NMR). The spectral data and melting points of known compounds were compared with those reported in the literature.
4-Formylbenzamide (5b). White powder, IR (KBr) νmax/cm−1: 3375, 3178 (NH2), 1664 (CONH); 1H NMR (400 MHz, DMSO-d6) δH: 10.09 (s, br, 1H, CHO), 8.24 (s, br, 2H, NH), 8.07 (d, J = 8.0 Hz, 1H), 7.98 (d, J = 8.0 Hz, 1H); 13C NMR (100 MHz, DMSO-d6) δC: 192.9, 167.6, 132.3, 129.3, 128.1, 127.3; anal. cald for C8H7NO2 (149.1): C, 64.42; H, 4.73; N, 9.39; found: C, 64.51; H, 4.55; N, 9.20%.
N-Phenylthiophene-2-carboxamide (6d). Yellow powder, IR (KBr) νmax/cm−1: 3303 (N–H), 1633 (CONH); 1H NMR (400 MHz, DMSO-d6) δH: 10.23 (s, br, 1H, NH), 8.03 (d, J = 3.6 Hz, 1H), 7.87 (d, J = 5.2 Hz, 1H), 7.72 (d, J = 7.6 Hz, 2H), 7.36 (t, J = 7.6 Hz, 2H), 7.23 (dd, J = 5.2, 3.6 Hz, 2H), 7.12 (t, J = 7.2 Hz, 1H); 13C NMR (100 MHz, DMSO-d6) δC: 159.8, 140.0, 138.7, 131.8, 129.1, 128.6, 128.0, 123.7, 120.3; anal. cald for C11H9NOS (203.3): C, 65.00; H, 4.46; N, 6.89; found: C, 65.16; H, 4.35; N, 6.75%.
2,4-Dichloro-N-phenylbenzamide (6e). Yellow powder, IR (KBr) νmax/cm−1: 3342 (N–H), 1659 (CONH); 1H NMR (400 MHz, DMSO-d6) δH: 10.54 (s, br, 1H, NH), 7.78 (d, J = 2.0 Hz, 1H), 7.70 (d, J = 8.0 Hz, 1H), 7.64 (d, J = 8.0 Hz, 2H), 7.56 (dd, J = 8.2, 2.0 Hz, 1H), 7.35 (m, 2H), 7.13 (t, J = 7.4 Hz, 1H); 13C NMR (100 MHz, DMSO-d6) δC: 164.0, 138.8, 135.8, 134.8, 130.3, 129.2, 128.8, 128.7, 127.4, 124.0, 119.5; anal. cald for C13H9Cl2NO (266.1): C, 58.67; H, 3.41; N, 5.26; found: C, 58.77; H, 3.52; N, 5.11%.

Conclusion

In conclusion, a novel heterogeneous HAp-encapsulated-α-Fe2O3-based Cu(II) organic–inorganic hybrid nanocatalyst was successfully produced and its catalytic activity was explored in the synthesis of amide derivatives from aromatic aldehydes with ammonium chloride and aniline hydrochloride in good yield and lower reaction time. The magnetic nature of this nanocatalyst allows its easy recovery and recycling by an external magnetic field.

Acknowledgements

The authors are grateful to the Research Council of University of Guilan for the financial support of this research work.

Notes and references

  1. Z.-L. Lu, E. Lindner and H. A. Mayer, Chem. Rev., 2002, 102, 3543–3578 CrossRef CAS PubMed.
  2. M. Kidwa, J. Arti and S. Bhardwaj, Mol. Diversity, 2012, 16, 121–128 CrossRef PubMed.
  3. (a) L. Ma'mani, M. Sheykhan, A. Heydari, M. Faraji and Y. Yamini, Appl. Catal., A, 2010, 377, 64–69 CrossRef PubMed; (b) W. Teunissen, A. A. Bol and J. W. Geus, Catal. Today, 1999, 48, 329–336 CrossRef CAS; (c) T.-J. Yoon, W. Lee, Y.-S. Oh and J.-K. Lee, New J. Chem., 2003, 27, 227–229 RSC; (d) H. Yoon, S. Ko and J. Jang, Chem. Commun., 2007, 1468–1470 RSC; (e) H.-H. Yang, S.-Q. Zhang, X.-L. Chen, Z.-X. Zhuang, J.-G. Xu and X.-R. Wang, Anal. Chem., 2004, 76, 1316 CrossRef CAS PubMed; (f) D. Lee, J. Lee, H. Lee, S. Jin, T. Hyeon and B. M. Kim, Adv. Synth. Catal., 2006, 348, 41–46 CrossRef CAS PubMed; (g) Y. Zhang, Z. Li, W. Sun and C. Xia, Catal. Commun., 2008, 10, 237–242 CrossRef CAS PubMed; (h) J. Deng, L.-P. Mo, F.-Y. Zhao, L.-L. Hou, L. Yang and Z.-H. Zhang, Green Chem., 2011, 13, 2576–2584 RSC; (i) Y.-H. Liu, J. Deng, J.-W. Gao and Z.-H. Zhang, Adv. Synth. Catal., 2012, 354, 441–447 CrossRef CAS PubMed.
  4. (a) C. Baleizão, A. Corma, H. García and A. Leyva, Chem. Commun., 2003, 606–607 RSC; (b) D. Wang and D. Astruc, Chem. Rev., 2014, 114, 6949–6985 CrossRef CAS PubMed; (c) P. D. Stevens, G. Li, J. Fan, M. Yen and Y. Gao, Chem. Commun., 2005, 4435–4437 RSC; (d) C. Ó. Dálaigh, S. A. Corr, Y. Gun'ko and S. J. Connon, Angew. Chem., Int. Ed., 2007, 46, 4329–4332 CrossRef PubMed.
  5. A. P. Wight and M. E. Davis, Chem. Rev., 2002, 102, 3589–3614 CrossRef CAS PubMed.
  6. V. R. Pattabiraman and J. W. Bode, Nature, 2011, 480, 471–479 CrossRef CAS PubMed.
  7. J. W. Bode, Curr. Opin. Drug Discovery Dev., 2006, 9, 765–775 CAS.
  8. C. A. G. N. Montalbetti and V. Falque, Tetrahedron, 2005, 61, 10827–10852 CrossRef CAS PubMed.
  9. P. Starkov and T. D. Sheppard, Org. Biomol. Chem., 2011, 9, 1320–1323 CAS.
  10. B. Cornils and W. A. Herrmann, Applied Homogeneous Catalysis with Organometallic Compounds: A Comprehensive Handbook, VCH, Weinheim, 1996 Search PubMed.
  11. (a) C. Gunanathan, Y. Ben-David and D. Milstein, Science, 2007, 317, 790–792 CrossRef CAS PubMed; (b) L. U. Nordstom, H. Vogt and R. Madsen, J. Am. Chem. Soc., 2008, 130, 17672–17673 CrossRef PubMed; (c) S. Yaragorla, G. Singh, P. L. Saini and M. K. Reddy, Tetrahedron Lett., 2014, 55, 4657–4660 CrossRef CAS PubMed.
  12. L. M. Jin, H. Lu, Y. Cui, C. L. Lizardi, T. N. Arzua, L. Wojtas, X. Cui and X. P. Zhang, Chem. Sci., 2014, 5, 2422–2427 RSC.
  13. S. H. Cho, E. J. Yoo, I. Bae and S. Chang, J. Am. Chem. Soc., 2005, 127, 16046–16047 CrossRef CAS PubMed.
  14. T. A. Dineen, M. A. Zajac and A. G. Myers, J. Am. Chem. Soc., 2006, 128, 16046–16047 CrossRef PubMed.
  15. (a) W. P. Mai, G. Song, J. W. Yuan, L. R. Yang, G. C. Sun, Y. M. Xiao, P. Mao and L. B. Qu, RSC Adv., 2013, 3, 3869–3872 RSC; (b) C. Roberta, P. Andrea, G. Giampaolo and L. D. Lidia, Org. Lett., 2012, 14, 5014 CrossRef PubMed.
  16. K. Nakagawa, H. Inoue and K. Minami, Chem. Commun., 1966, 17–18 RSC.
  17. W. J. Yoo and C. J. Li, J. Am. Chem. Soc., 2006, 128, 13064–13065 CrossRef CAS PubMed.
  18. S. C. Ghosh, J. S. Y. Ngiam, A. M. Seayad, D. T. Tuan, C. L. L. Chai and A. Chen, J. Org. Chem., 2012, 77, 8007–8015 CrossRef CAS PubMed.
  19. D. Saberi and A. Heydari, Appl. Organomet. Chem., 2013, 28, 101–108 CrossRef PubMed.
  20. (a) M. Sheykhan, M. Mohammadquli and A. Heydari, J. Mol. Struct., 2012, 1027, 156–161 CrossRef CAS PubMed; (b) M. Mohsenimehr, M. Mamaghani, F. Shirini, M. Sheykhan and F. Azimian Moghaddam, Chin. Chem. Lett., 2014, 25, 1387–1391 CrossRef CAS PubMed; (c) L. Ma'mani, M. Sheykhan and A. Heydari, Appl. Catal., A, 2011, 395, 34–38 CrossRef PubMed; (d) M. Sheykhan, L. Ma'mani, A. Ebrahimi and A. Heydari, J. Mol. Catal. A: Chem., 2011, 335, 253–261 CrossRef CAS PubMed.
  21. S. Thomas, K. Joseph, S. K. Malhotra, K. Goda and M. S. Sreekala, Polymer Composites, Nanocomposites, Wiley-VCH, 2013, vol. 2, p. 34 Search PubMed.
  22. Chemical Book, http://www.chemicalbook.com/.
  23. D. R. Lide and G. W. A. Milne, Handbook of Applied Data on Organic Compounds, CRC Press, Inc., Boca Raton, FL., 3rd edn, 1994, vol. 1 Search PubMed.
  24. J. Wang, X. Yin, J. Wu, D. Wu and Y. Pan, Tetrahedron, 2013, 69, 10463–10469 CrossRef CAS PubMed.
  25. R. Cadoni, A. Porcheddu, G. Giacomelli and L. D. Luca, Org. Lett., 2012, 14, 5014–5017 CrossRef CAS PubMed.
  26. (a) Z. Liu, J. Zhang, S. Chen, E. Shi, Y. Xu and X. Wan, Angew. Chem., Int. Ed., 2012, 51, 3231–3235 CrossRef CAS PubMed; (b) K. Xu, Y. Hu, S. Zhang, Z. Zha and Z. Wang, Chem.–Eur. J., 2012, 18, 9793–9797 CrossRef CAS PubMed.
  27. M. Khoobi, L. Ma'mani, F. Rezazadeh, Z. Zareie and A. Foroumadi, J. Mol. Catal. A: Chem., 2012, 359, 74–80 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015