Cu–EDTA-modified APTMS-Fe3O4@SiO2 core–shell nanocatalyst: a novel magnetic recoverable catalyst for the Biginelli reaction

Mehdi Sheykhan*, Asieh Yahyazadeh* and Zahra Rahemizadeh
Chemistry Department, University of Guilan, P.O. Box 41335-1914, Rasht, Iran. E-mail: sheykhan@guilan.ac.ir; Fax: +981333367262

Received 26th January 2016 , Accepted 24th March 2016

First published on 30th March 2016


Abstract

A novel copper–ethylenediamine tetracarboxylate modified core–shell magnetic catalyst is introduced. The prepared catalyst was fully characterized by various spectroscopic analyses such as XRD, SEM, FT-IR, EDX, ICP, and CHNOS. After characterization, its activity was evaluated as a supported transition metal catalyst in the multi-component Biginelli reaction. The novel catalyst acts as an efficient heterogenized catalyst for synthesis of 3,4-dihydropyrimidin-2(1H)-one/thione derivatives in solvent-free conditions. A wide range of biologically active dihydropyrimidin-2(1H)-one/thiones were synthesized in the presence of the novel catalyst in 10–15 minutes with high yields (85–98% isolated yields). In addition, the reusability of the catalyst was tested by an external magnet. The investigation showed that no notable reduction of yields was observed after reusing over ten runs, proving its stability during recycling processes. More importantly, very small amounts (0.35 mol%) of the novel catalyst were required to result in the maximum turnover frequency of the Biginelli reaction obtained to date (TOF about 1000–1680 h−1 and total TOF over 14[thin space (1/6-em)]000 h−1).


Introduction

Because of their outstanding properties, magnetic nanoparticles have received striking interest in various disciplines such as magnetic drug delivery, information storage, ferrofluids, separations, magnetic resonance imaging, and tumor hyperthermia treatment.1 In catalysis, magnetic nanoparticles are attractive catalysts since they can be separated from the reaction medium after magnetization by an external magnet. For industrial applications, where the total cost of chemical processes is of most important factors, magnetic separation is an intriguing alternative to filtration or centrifugation because it prevents the loss of catalyst and enhances reusability.2 Therefore, the development of novel magnetic nanoparticles with tunable catalytic activity is of great significance for both academia and industry.

As a challenging criterion in organic synthesis, pharmaceutical and therapeutic chemicals syntheses have been extensively explored in recent decades.3 Meanwhile, biologically active 3,4-dihydropyrimidin-2(1H)-one/thiones (DHPMs) due to important roles in live systems such as calcium channel blocking, anti-inflammation, anti-hypertension, anti-tumor, and acting as neuropeptide antagonists, mitotic kinesin inhibitors, anti-virals, and others, as reviewed elsewhere have been of the most important synthetic targets.4 They can be prepared by the three component condensation of an aldehyde, β-ketoester and urea/thiourea in the presence of a strong acid, through the Biginelli reaction.5 Owing to the considerable attention to DHPMs, the synthetic procedure of Biginelli reaction has been repeatedly modified, thus, to date, marked improvements including the use of Bronsted acid catalysts,6 Lewis acid catalysts,7 ionic liquids,8 magnetic catalysts9 and magnetic ionic liquids10 have been reported on it.

Among the aforementioned methods, heterogeneous catalysts play an efficient role in amelioration of the conditions.11 As examples, the use of Indion-130, Nafion-H, Nafion-NR-50, Amberlyst-70,12 supported catalysts based on resins, silica gel, alumina or PEG,13 bioglycerol-based carbon catalyst as one of the carbon-based solid acids14 has been reported.15 Most of the reported methods are worthwhile; however, many of them have drawbacks such as: tedious workup of the reaction mixture, difficult separation and recovery of the catalyst, toxic and moisture sensitive reaction conditions, low yields and long reaction times. Therefore, investigations for development of more efficient, simpler and milder catalytic systems are still needed.

With the “greening” of global chemical processes16 in mind, ‘Heterogenization’ of homogeneous catalysts is a general trend in catalysis science.17 As an example, one can find Fe3O4@SiO2 core–shell based heterogenized ionic liquid catalysts such as phosphomolybdic acid9b and HSO4 immobilized catalysts10 for the Biginelli reaction. The prepared heterogenized catalysts can now easily separate from the reaction mixture, converting them to the reusable catalysts. The only restriction during heterogenization, is the lower activity/selectivity due to the increasing of mass-diffusion to the catalyst sites.18 Nowadays, the problem fixed both by the use of porous compounds19 and the synthesis of “inorganic–organic hybrids” by attachment of organic moieties with pendant attached chains on the inorganic heterogeneous surfaces.20 The latter leads to the combination of the two complementary properties: the inorganic properties like mechanical/thermal/structural stability and the properties of organic pendant moieties such as flexibility in solution (like homogeneous catalysts) and therefore, high reactivity of the catalyst.21 Herein, we report a convenient preparation and structural characterization of an Cu–EDTA-functionalized core–shell magnetic compound as a supported transition metal catalyst in the multi-component Biginelli reaction.

Results and discussion

The catalyst characterization

The core–shell support was prepared according to the previously reported procedure.22 The prepared nanoparticles of Fe3O4@SiO2, 1, were then functionalized with 3-aminopropyltrimethoxysilane (APTMS) to produce an organic–inorganic hybrid (Fe3O4@SiO2–(CH2)3–NH2, 2). Then, the prepared hybrid was affected by basic solution of EDTA–Cu(II) complex. After 24 h stirring at 75 °C, the brown powder washed by deionized water and dichloromethane and dried under vacuum which led to the formation of the EDTA-modified magnetic core–shell compound named as (Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu, 3), as illustrated in Scheme 1. The prepared (Fe3O4@SiO2–(CH2)3–NH2, 2) and (Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu, 3) were fully characterized by XRD (Fig. 1a and b), SEM and histogram analyses (Fig. 1c–f), FT-IR (Fig. 2a and b), EDX (Fig. 3) and CHNOS. In addition, the organic loading percent and magnetic properties of 3 were characterized by TGA and VSM analyses, respectively (Fig. 4a and b).
image file: c6ra02415g-s1.tif
Scheme 1 Preparation of the catalyst.

image file: c6ra02415g-f1.tif
Fig. 1 X-ray powder diffraction patterns of the prepared 1 (a) and 3 (b), scanning electron microscopy of 1 (c) and 3 (d), histogram size distribution of 1 (e) and 3 (f).

image file: c6ra02415g-f2.tif
Fig. 2 FTIR spectra of the prepared 1 ((a), blue-line), 2 ((a), red-line) and 3 (b).

image file: c6ra02415g-f3.tif
Fig. 3 Energy-dispersive X-ray spectroscopy microanalysis (EDAX) of compound 3.

image file: c6ra02415g-f4.tif
Fig. 4 Thermo-gravimetric analysis (a) and vibrating sample magnetometer curve of 3 at r.t (b).

The Fe3O4@SiO2 and Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu synthesized were subjected to structural characterization with XRD, Fig. 1a and b. Diffraction peaks related to the (111), (220), (311), (400), (333) and (440) planes were clearly observed. The diffraction peaks are in agreement with that of the cubic structure of Fe3O4 (magnetite) with Fd[3 with combining macron]m space group (ICDD card no. 75-1372) in both patterns. Also, the broad diffraction peak at 23.5° is the characteristic peak of SiO2 shell. No other phase was detectable. In addition, there was no copper phase, proving that metallic Cu was not formed. Furthermore, competing two X-ray diffraction patterns proved that no clear loss of crystallinity appeared after the modification of the surface. The measurements were carried out on a Philips X'Pert diffractometer with CuKα radiation (λ = 0.154056 nm).

The morphologies of prepared compounds were identified by the scanning electron microscopy (Fig. 1c and d). SEM photographs of Fe3O4@SiO2 and Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu indicated that both synthetic compounds were present as uniform nanoparticles. Histogram analysis of the SEM images (Fig. 1e and f) showed that the size of the nanoparticles of Fe3O4@SiO2 and Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu is about 12 nm and 18 nm, respectively.

The FT-IR spectra of Fe3O4@SiO2, Fe3O4@SiO2–(CH2)3–NH2 and Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu compounds were recorded in the range 400–4000 cm−1 (Fig. 2a and b). For Fe3O4@SiO2, the H–O–H bending vibrations are observed at about 1000–1650 cm−1, typical of the adsorbed H2O. In addition, the band at 900–1000 cm−1 corresponds to bending vibration of O–H bond. The O–H in plane and out of plane vibrations appear at 1583–1481 and 935–838 cm−1, respectively. The bands at 400–660 cm−1, are corresponding to the stretching of Fe–O bonds in the crystalline lattice of Fe3O4. They are characteristically pronounced for all spinel structures and for ferrites in particular. The broader IR absorption band in the 2800–3700 cm−1 region is ascribed to Si–OH groups. Stretching vibration modes of Si–O bond are observed at 1120 cm−1 and 1180 cm−1. In the FTIR of Fe3O4@SiO2–(CH2)3–NH2 all of the mentioned bands are present. In addition, a characteristic band due to the stretching of C–H bonds is appeared at 2938 cm−1 (red highlighted dotted-line) in which proves the modification of the surface of core–shell Fe3O4@SiO2 is successful (Fig. 2a). The FTIR spectrum of the compound named Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu shows all of the mentioned bands. It seems that there is no detectable change associated by introducing EDTA–Cu except the band in 1646 cm−1 that shift to 1638 cm−1 after modification (Fig. 2b). We consider the new band (1638 cm−1) to the stretching vibration of C[double bond, length as m-dash]O bond in EDTA modified compound.

The presence of EDTA and more importantly Cu is confirmed by the elemental analysis of the compound 3. Energy-dispersive X-ray spectroscopy microanalysis (EDAX) was recorded for this compound (Fig. 3). As shown in Fig. 3, the compound 3 has carbon, nitrogen, and copper as well as iron and silicon.

Quantitative elemental analysis of 3, showed 8.81% C (equal to 7.3 mmol C g−1) and 2.72% N (equal to 1.9 mmol N g−1). Therefore, the ratio of C/N resulted from EDAX is about 3.84. To characterize more precisely, the ratio of ligand/metal for the complex must be calculated by this method. Calculation of the N/Cu resulted in 2.7. Interestingly, the resulted 2.7 ratio is in agreement with the proposed structure of 3 (Scheme 2), in which there are 3 nitrogen atoms for each copper ion.


image file: c6ra02415g-s2.tif
Scheme 2 The proposed structure of 3.

In addition, the C/N ratio of 3.84 is approximately in agreement with this structure because it must have 14 C atoms and 3 N atoms. So, it must have the C/N ratio of 4.6.

The loading of the active site on the magnetic core–shell is determined by considering the presence of 3 N atoms in each active site and is about 0.64 mmol per gram of the catalyst.

There was just one uncertainty on the basis of the results of EDAX analysis. Whether the structure is as illustrated in Scheme 2 or the directly complexed Cu–amino groups structure without the presence of EDTA (Scheme 3). However, if this was the case, the N/Cu ratio must be 2 instead of calculated 2.7, but the calculated C/N ratio of 3.8 is more near to 4 than 4.6.


image file: c6ra02415g-s3.tif
Scheme 3 The ruled out directly complexed Cu–amino structure of 3 without the presence of EDTA.

However, the formation equilibrium constant (Kf) for EDTA–Cu is higher than Kf for a bidentate N–Cu complex, for more exact confirmation of the proposed structure and to rule out the structure shown in Scheme 3, the compound 3 was subjected to the elemental CHNOS analysis.

The CHNOS analysis showed 6.23% C, 1.49% N and 1.55% H in the compound 3. Clearly, it means there are 5.19 mmol C g−1 and 1.07 mmol N g−1 in the prepared structure. On the basis of the results, C/N ratio is about 4.8, the number which is more near to 4.6 than 4. Therefore, the presence of EDTA and subsequently the structure proposed in Scheme 2 is confirmed.

Considering 3 N atoms in each active site, the loading of compound is calculated about 0.35 mmol g−1 by CHNOS. Since EDAX analyzes a dot-non-uniform region of structure and because in the synthetic procedure, 0.5 mmol EDTA–Cu complex was affected by 1 g of the APTMS-modified core–shell compound, the results obtained from CHNOS must be more realistic than results of EDAX (0.64 mmol g−1). So, the loading is considered as is determined by CHNOS (0.35 mmol g−1) for the compound 3.

In contrast to a homogeneous catalyst which has well-defined active sites, the active sites of heterogeneous catalysts have remained obscure. Therefore, many reports are present in the literature that consider the modification by APTES and APTMS as a continuous and distinct shell around the SiO2 layer, whereas, others consider just a local functionalization on the surface. Among those which showed local functionalization, many indicated a tri-bridged O–Si bond between the linker and surface of SiO2 (ref. 23) but some others demonstrated a two-bridged O–Si bond between the linker and surface of SiO2 (ref. 24) as did we.

In Scheme 2, the sodium comes from EDTA·2Na. XRD spectroscopic analysis rules out the reduction of Cu(II) to metallic Cu. Therefore, it is reasoned that copper was in its +2 oxidation state and must have had a counter ion in the structure of its complex. According to the materials used, there are just two probabilities, the presence of (1) sodium ion (from the EDTA·2Na) and (2) potassium ion (from the K2CO3). Based on EDAX spectroscopy, there is no K element in the structure. Hence, sodium must be the counter ion present in the structure of compound 3. However, the reason that sodium was not observed in EDAX spectrum is that Na element is lighter than being detected by EDAX.

Further quantitative determination of the organic group loaded on the surface of compound 3 was performed by using thermo-gravimetric analysis (TGA) (Fig. 4a). Three weight losing steps were observed in the analysis. First step which is due to the evaporation of adsorbed water appeared at about 108 °C. The peak is followed by a weight loss of 8.0% at about 261 °C, corresponding to the loss of EDTA–Cu complex. This proves a loading of about 0.32 mmol g−1. As the third step, a weight loss of 11.6% at 408 °C is observed which is corresponded to the complete loss of organic linker from the surface of compound 3. The result is in agreement with those of CHNOS analyses.

The magnetic feature of compound 3 is also measured in an applied magnetic field at r.t, with the field sweeping from −8000 to +8000 Oersted (Fig. 4b). The ‘retentivity/magnetic saturation’ ratio for the compound is about 0.001, proving that 3 has superparamagnetic nature. Its M (H) hysteresis loop is completely reversible and the mentioned reversibility confirms that no aggregation occurred in the magnetic fields. In addition, the magnetic saturation value of 3 is 5.00 emu g−1 at r.t. Its high permeability in magnetization as well as good magnetic saturation is sufficient for magnetic separation of it with a conventional magnet.

Catalytic activity

After detailed characterization of the prepared compound 3, its catalytic activity was investigated in the Biginelli reaction for the formation of various dihydropyrimidin-2(1H)-one or thione compounds.

A mixture of benzaldehyde (1.0 mmol), ethyl acetoacetate (1.0 mmol) and urea (1.2 mmol) was allowed to react in the presence of catalyst 3 as the model reaction. Systematic screening of the reaction conditions was done precisely in the presence of various catalyst amounts, different times, different solvents and various reaction temperatures (Table 1).

Table 1 Screening the reaction conditionsa
Entry Catalyst Time (min) Temperature (°C) Solvent Yield %
a Reaction conditions: benzaldehyde (1 mmol), ethyl acetoacetate (1 mmol), urea (1.2 mmol).
1 3 mg (0.11 mol%) 30 100 ACN 55
2 7 mg (0.25 mol%) 30 100 ACN 72
3 10 mg (0.35 mol%) 30 100 ACN 89
4 15 mg (0.55 mol%) 30 100 ACN 87
5 10 mg (0.35 mol%) 10 100 ACN 78
6 10 mg (0.35 mol%) 15 100 ACN 80
7 10 mg (0.35 mol%) 25 100 ACN 83
8 10 mg (0.35 mol%) 35 100 ACN 81
9 10 mg (0.35 mol%) 10 75 ACN 61
10 10 mg (0.35 mol%) 10 120 ACN 83
11 10 mg (0.35 mol%) 10 100 88
12 10 mg (0.35 mol%) 10 100 DMF 75
13 10 mg (0.35 mol%) 10 100 EtOH 80
14 10 mg (0.35 mol%) 10 100 DMSO 71
15 10 mg (0.35 mol%) 10 100 NMP 76


After optimization of reaction conditions, it was proved that the reaction in the presence of 10 mg of the catalyst (0.35 mol%) under solvent-free conditions and at 100 °C resulted in the corresponding product 4a in excellent yield in 10 minutes. On the basis of the optimal conditions established, the Biginelli reaction of various aldehydes, 1,3-dicarbonyl compounds and urea/thiourea in solvent-free conditions were examined. As shown in Table 2 the reactions proceed smoothly and corresponding 3,4-dihydropyrimidin-2(1H)-one/thiones could be obtained in high yields.

Table 2 The synthesis of various 3,4-dihydropyrimidin-2(1H)-one/thiones in the presence of catalyst 3aimage file: c6ra02415g-u1.tif
Entry Aldehyde R2/X Product Yields (%) Time (min) TONb TOFc (h−1) Ref.
a Reaction conditions: aldehydes (1 mmol), 1,3-dicarbonyl compounds (1 mmol), urea/thiourea (1.2 mmol), free-solvent, 100 °C. Just the isolated yields are reported.b Number of moles of product produced from 1 mole of catalyst.c TON per unit of time.
4a image file: c6ra02415g-u2.tif OEt/O image file: c6ra02415g-u3.tif 88 10 251 1506 25
4b image file: c6ra02415g-u4.tif OEt/O image file: c6ra02415g-u5.tif 98 10 280 1680 25
4c image file: c6ra02415g-u6.tif OEt/O image file: c6ra02415g-u7.tif 92 10 262 1572 26
4d image file: c6ra02415g-u8.tif OEt/O image file: c6ra02415g-u9.tif 85 10 242 1452 25
4e image file: c6ra02415g-u10.tif OEt/O image file: c6ra02415g-u11.tif 90 10 257 1542 27
4f image file: c6ra02415g-u12.tif OEt/O image file: c6ra02415g-u13.tif 91 10 260 1560 28
4g image file: c6ra02415g-u14.tif OEt/O image file: c6ra02415g-u15.tif 95 10 271 1628 25
4h image file: c6ra02415g-u16.tif OEt/O image file: c6ra02415g-u17.tif 92 10 262 1572 29
4i image file: c6ra02415g-u18.tif OEt/S image file: c6ra02415g-u19.tif 92 10 262 1572 30
4j image file: c6ra02415g-u20.tif OEt/S image file: c6ra02415g-u21.tif 89 15 254 1016 31
4k image file: c6ra02415g-u22.tif OEt/S image file: c6ra02415g-u23.tif 94 10 268 1608 32
4l image file: c6ra02415g-u24.tif OEt/S image file: c6ra02415g-u25.tif 92 15 262 1048 33
4m image file: c6ra02415g-u26.tif OEt/S image file: c6ra02415g-u27.tif 88 10 251 1506 30
4n image file: c6ra02415g-u28.tif OEt/S image file: c6ra02415g-u29.tif 87 15 248 992 27
4o image file: c6ra02415g-u30.tif Me/S image file: c6ra02415g-u31.tif 85 10 242 1452 29


At the end of the reaction, reusability of the catalyst was evaluated by decanting the vessel using an external magnet and washing the retained catalyst with dichloromethane, drying, and using in a subsequent reaction (Fig. 5). The reaction of benzaldehyde, ethyl acetoacetate and urea resulted in the corresponding 3,4-dihydropyrimidin-2(1H)-one 4a in 88% isolated yield. After ten consecutive reactions, the isolated yield remained similar to the first run and no detectable loss was obtained. The progress was made with 82.6% average yield of the reaction and the total turnover number of up to 14[thin space (1/6-em)]000 h−1.


image file: c6ra02415g-f5.tif
Fig. 5 The reusing of catalyst in synthesis of 4a; carried out in 100 °C for 10 minutes.

To check leaching of the catalyst into solution, after 5 minutes from starting, the vessel was magnet decanted and observed that the reaction in the supernatant did not complete even after 6 h. The experiment was repeated and this time instead of 5 minutes, the catalyst was separated after 30 minutes by magnetic decantation and the supernatant was tested by inductive coupled plasma spectroscopy (ICP-AES). ICP-AES result supported that no detectable amounts of Cu were found in the supernatant proving there is no contribution of homogeneous catalysis (via leached catalyst) in the course of reaction. The novel catalytic procedure in comparison with some other ones is presented in Table 3.

Table 3 Comparison of this work with the previously reported catalysts for Biginelli reaction of 4-nitrobenzadehyde, ethyl acetoacetate and ureaa
Entry Catalyst Time (h) Yield% TOF (h−1) Ref.
a Because there was not reaction data of 4-nitrobenzaldehyde, the data of 3-nitrobenzaldehyde is presented here.
1 H5PW10V2O40/Pip-SBA-15 (0.6 g, 2 mol%) 0.75 80 53 25
2 H3PMo12O40 nanoparticles on imidazole functionalized Fe3O4@SiO2 (0.03 g, 0.3 mol%) 0.33 94 939 9b
3 HSO4 imidazole functionalized Fe3O4@SiO2 (0.05 g, 1.1 mol%) 0.5 97 176 10a
4 NH4H2PO4/MCM-41 (0.04 g, 5 mol%) 2.5 85 6.8 26
5 [Et3NH][HSO4] (3 equiv.) 1.33 76 34
6 SBSSA (0.05 g) 1 93 35
7 Nano-ZnO (5 mol%) 12 65 1.1 36
8 Fe3O4 (20 mol%) 0.4 71 8.9 29
9 Fe3O4/PAA–SO3H (22.3 mol%) 1.9 84 2 37
10 Ce(LS)3 (20 mol%) 8 91 0.6 38
11 Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu(II) (0.35 mol%) 0.17 98 1680 Current work


Conclusion

In summary, the preparation and characterization of a copper-EDTA-functionalized core–shell magnetic compound as a supported transition metal catalyst in the multi-component Biginelli reaction was described. The novel catalyst acts as a powerful heterogenized catalyst for synthesis of 3,4-dihydropyrimidin-2(1H)-one/thione derivatives in solvent-free conditions. In addition, in competition with other reported catalysts, the new magnetic catalyst has higher activity. Appropriate distance between catalytic active sites (indeed copper ion) from solid surface reduces the steric hindrance and therefore eliminates mass transport limitations. More importantly, the novel catalyst acts its catalytic role in very small amounts (0.35 mol%) and therefore shows the maximum turnover frequency of Biginelli reaction obtained to date (TOF about 1000–1680 h−1 and total TOF over 14[thin space (1/6-em)]000 h−1). Reusing the catalyst over ten subsequent reactions was accomplished without notable reduction of yields and proved its stability over recycling processes.

Experimental section

The preparation of catalyst

A mixture of FeCl2·4H2O (2.5 mmol) and FeCl3·6H2O (5 mmol) in de-ionized water (20 mL) was added to a two-necked balloon. After stirring for ten minutes ammonia (25%, 40 mL) was added drop-wise (1 mL min−1) until the solution pH rise to about 11. Then, TEOS (20 mL) added to above mentioned suspension drop-wise under reflux and stirred for 12 hour. The dark-brown solid was washed with water (3 times) and ethanol (3 times) and oven-dried at 200 °C to produce Fe3O4@SiO2 (compound 1) as a brown solid. 1 gram of 1 was dispersed in 50 mL toluene and then 3-(trimethoxysilyl)propylamine (20 mmol, 4.5 mL) was added under reflux and inert atmosphere. The mixture stirred for 8 hours and then cooled to r.t, washed with water and dichloromethane and oven-dried at 50 °C to yield Fe3O4@SiO2–Si–(CH2)3–NH2 (compound 2).

A mixture of CuCl2 (anhydrous, 0.5 mmol), EDTA·2Na (0.5 mmol) and K2CO3 (1 mmol) in water (5 mL) was stirred at r.t for 4 hours resulted in the production of EDTA–Cu(II) complex.

The EDTA–Cu(II) complex solution was added to 1 gram of 2 in water (25 mL) and stirred at 75 °C for 24 hours. The resulted solid was washed with water and oven-dried at 100 °C to produce Fe3O4@SiO2–(CH2)3–NH–EDTA–Cu, (compound 3).

General procedure for the synthesis of 3,4-dihydropyrimidin-2(1H)-one/thiones in the presence of catalyst 3

A mixture of aldehyde (1 mmol), 1,3-dicarbonyl compound (1 mmol), urea/thiourea (1.2 mmol), and catalyst 3 (10 mg, 0.35 mol%) was stirred at 100 °C in solvent-free conditions. At the end of reaction (determined by TLC), ethanol (10 mL) was added and the catalyst separated by external magnet. The products was obtained and purified via recrystallization from ethanol and chromatographic purification. The structures of products were proved by the comparison of their mp, FTIR and 1HNMR by related literature.

Acknowledgements

This work was supported by research council of the University of Guilan.

Notes and references

  1. (a) D. E. Speliotis, J. Magn. Magn. Mater., 1999, 193, 29 CrossRef CAS; (b) C. Corot, P. Robert, J. M. Idee and M. Port, Adv. Drug Delivery Rev., 2006, 58, 1471 CrossRef CAS PubMed.
  2. (a) W. Teunissen, F. M. F. De Groot, J. Geus, O. Stephan, M. Tence and C. Colliex, J. Catal., 2001, 204, 169 CrossRef CAS; (b) P. D. Stevens, J. Fan, H. M. R. Gardimalla, M. Yen and T. Gao, Org. Lett., 2005, 7, 2085 CrossRef CAS PubMed; (c) H. M. R. Gardimalla, D. Mandal, P. D. Stevens, M. Yen and Y. Gao, Chem. Commun., 2005, 4432 RSC; (d) A.-H. Lu, W. Schmidt, N. Matoussevitch, H. Bönnemann, B. Spliethoff, B. Tesche, E. Bill, W. Kiefer and F. Schuth, Angew. Chem., Int. Ed., 2004, 43, 4303 CrossRef CAS PubMed; (e) Y. Zheng, P. D. Stevens and Y. Gao, J. Org. Chem., 2006, 71, 537 CrossRef CAS PubMed; (f) H. Yang, G. Li and Z. Ma, J. Mater. Chem., 2012, 22, 6639 RSC; (g) H. Yang, Y. Wang, Y. Qin, Y. Chong, Q. Yang, G. Li, L. Zhang and W. Li, Green Chem., 2011, 13, 1352 RSC; (h) R. B. N. Baig and R. S. Varma, ACS Sustainable Chem. Eng., 2014, 2, 2155 CrossRef; (i) S.-W. Chen, Z.-C. Zhang, M. Ma, C.-M. Zhong and S.-G. Lee, Org. Lett., 2014, 16, 4969 CrossRef CAS PubMed; (j) C. Sun, R. Zhou, J. E. J. Sun, Y. Su and H. Ren, RSC Adv., 2016, 6, 10633 RSC; (k) B. G. Wang, B. C. Ma, Q. Wang and W. Wang, Adv. Synth. Catal., 2010, 352, 2923 CrossRef CAS; (l) M. Ferré, R. Pleixats, M. Wong Chi Man and X. Cattoën, Green Chem., 2016, 18, 881–922 RSC.
  3. (a) H. C. Kolb, M. G. Finn and K. B. Sharpless, Angew. Chem., Int. Ed., 2001, 40, 2004 CrossRef CAS; (b) E. A. Couladouros and A. T. Strongilos, Angew. Chem., Int. Ed., 2002, 41, 3677 CrossRef CAS; (c) C. Chen, X. Li, C. S. Neumann, M. M.-C. Lo and S. L. Schreiber, Angew. Chem., Int. Ed., 2005, 44, 2249 CrossRef CAS PubMed; (d) D. M. Swanson, D. R. Shah, B. Lord, K. Morton, L. K. Dvorak, C. Mazur, R. Apodaca, W. Xiao, J. D. Boggs, M. Feinstein, S. J. Wilson, A. J. Barbier, P. Bonaventure, T. W. Lovenberg and N. I. Carruthers, Eur. J. Med. Chem., 2009, 44, 4413 CrossRef CAS PubMed.
  4. (a) A. K. Sharma, S. Jayakumar, M. S. Hundal and M. P. Mahajan, J. Chem. Soc., Perkin Trans. 1, 2002, 774 RSC; (b) S. Sasaki, N. Cho, Y. Nara, M. Harada, S. Endo and N. Suzuki, J. Med. Chem., 2003, 46, 113 CrossRef CAS PubMed; (c) S. Bartolini, A. Mai, M. Artico, N. Paesano, D. Rotili and C. Spadafora, J. Med. Chem., 2005, 48, 6776 CrossRef CAS PubMed; (d) A. R. Khosropour, I. Mohammadpoor-Baltork and H. Ghorbankhani, Catal. Commun., 2006, 7, 713 CrossRef CAS; (e) C. O. Kappe, Eur. J. Med. Chem., 2000, 35, 1043 CrossRef CAS PubMed; (f) C. O. Kappe, O. V. Shishkin, G. Uray and P. Verdino, Tetrahedron, 2000, 56, 1859 CrossRef CAS; (g) W. Jie-Ping and P. Yuan-Jiang, Chem. Commun., 2009, 2768 Search PubMed; (h) B. L. Nilsson and L. E. Overman, J. Org. Chem., 2006, 71, 7706 CrossRef CAS PubMed; (i) A. Csampai, A. Z. Gyor, G. I. Turos and P. Sohar, J. Organomet. Chem., 2009, 694, 3667 CrossRef CAS; (j) M. B. Deshmukh, S. M. Salunkhe, D. R. Patil and P. V. Anbhule, Eur. J. Med. Chem., 2009, 44, 2651 CrossRef CAS PubMed; (k) J. E. Biggs-Houck, A. Younai and J. T. Shaw, Curr. Opin. Chem. Biol., 2010, 14, 371 CrossRef CAS PubMed; (l) P. Slobbe, E. Ruijter and R. V. A. Orru, MedChemComm, 2012, 3, 1189 RSC.
  5. (a) D. Shobha, M. A. Chari, A. Mano, S. T. Selvan, K. Mukkanti and A. Vinu, Tetrahedron, 2009, 65, 10608 CrossRef CAS; (b) G. C. Tron, A. Minassi and G. Appendino, Eur. J. Org. Chem., 2011, 5541 CrossRef CAS.
  6. (a) F. Bigi, S. Carloni, B. Frullanti, R. Maggi and G. Sartori, Tetrahedron Lett., 1999, 40, 3465 CrossRef CAS; (b) V. R. Rani, N. Srinivas, M. R. Krishnan, S. J. Kulkarni and K. V. Raghavan, Green Chem., 2001, 3, 305 RSC; (c) Q. Zhang, X. Wang, Z. Li, W. Wu, J. Liu, H. Wu, S. Cui and K. Guo, RSC Adv., 2014, 4, 19710 RSC; (d) M. Nasr-Esfahani and M. Taei, RSC Adv., 2015, 5, 44978 RSC; (e) E. Kolvari, N. Koukabi, M. M. Hosseini, M. Vahidian and E. Ghobadi, RSC Adv., 2016, 6, 7419 RSC.
  7. (a) O. M. Singh, S. J. Singh, M. B. Devi, L. Alini Devi, N. I. Singh and S. G. Lee, Bioorg. Med. Chem. Lett., 2008, 18, 6462 CrossRef CAS PubMed; (b) J. Lu and H. Ma, Synlett, 2000, 63 CAS; (c) J. Lu, Y. Bai, Z. Wang, B. Yang and H. Ma, Tetrahedron Lett., 2000, 41, 9075 CrossRef CAS; (d) D. S. Bose, L. Fatima and H. B. Mereyala, J. Org. Chem., 2003, 68, 587 CrossRef CAS PubMed; (e) Y. Ma, C. Qian, L. Wang and M. Yang, J. Org. Chem., 2000, 65, 3864 CrossRef CAS PubMed; (f) B. C. Ranu, A. Hajra and U. Jana, J. Org. Chem., 2000, 65, 6270 CrossRef CAS PubMed; (g) S. Khademinia, M. Behzad and H. Samari Jahromi, RSC Adv., 2015, 5, 24313 RSC; (h) S. Rostamnia and A. Morsali, RSC Adv., 2014, 4, 10514 RSC.
  8. (a) N. Sharma, U. K. Sharma, R. Kumar and A. K. Sinha, RSC Adv., 2012, 2, 10648 RSC; (b) A. Khazaei, M. A. Zolfigol, S. Alaie, S. Baghery, B. Kaboudin, Y. Bayat and A. Asgari, RSC Adv., 2016, 6, 10114 RSC.
  9. (a) J. Safari and Z. Zarnegar, RSC Adv., 2013, 3, 17962 RSC; (b) J. Javidi, M. Esmaeilpour and F. Nowroozi Dodeji, RSC Adv., 2015, 5, 308 RSC.
  10. J. Safari and Z. Zarnegar, New J. Chem., 2014, 38, 358 RSC.
  11. L. Peng, R. Sridhar, R. J. Butcher, H. D. Arman, Z. Chen, S. Xi, B. Chen and C.-G. Zhao, Tetrahedron Lett., 2011, 52, 6220 CrossRef PubMed.
  12. (a) H. Lin, Q. Zhao, B. Xu and X. Wang, J. Mol. Catal. A: Chem., 2007, 268, 221 CrossRef CAS; (b) V. Polshettiwar and R. S. Varma, Tetrahedron Lett., 2007, 48, 7343 CrossRef CAS; (c) M. Tajbakhsh, B. Mohajerani, M. M. Heravi and A. N. Ahmadi, J. Mol. Catal. A: Chem., 2005, 236, 216 CrossRef CAS; (d) C. O. Kappe, Synlett, 1998, 718 CrossRef CAS; (e) J. K. Joseph, S. L. Jain and B. Sain, J. Mol. Catal. A: Chem., 2006, 247, 99 CrossRef CAS; (f) H. S. Chandak, N. P. Lad and P. P. Upare, Catal. Lett., 2009, 131, 469 CrossRef CAS.
  13. (a) A. Dondoni and A. Massi, Tetrahedron Lett., 2001, 42, 7975 CrossRef CAS; (b) M. A. Chari and K. Syamasundar, J. Mol. Catal. A: Chem., 2004, 221, 137 CrossRef CAS; (c) J. Mondal, T. Senb and A. Bhaumik, Dalton Trans., 2012, 41, 6173 RSC; (d) Z.-J. Quan, Y.-X. Da, Z. Zhang and X.-C. Wang, Catal. Commun., 2009, 10, 1146 CrossRef CAS.
  14. (a) M. Hara, T. Yoshida, A. Takagaki, T. Takata, J. N. Kondo, S. Hayashi and K. Domen, Angew. Chem., Int. Ed., 2004, 43, 2955 CrossRef CAS PubMed; (b) M. Toda, A. Takagaki, M. Okamura, J. N. Kondo, S. Hayashi, K. Domen and M. Hara, Nature, 2005, 438, 178 CrossRef CAS PubMed; (c) A. Takagaki, M. Toda, M. Okamura, J. N. Kondo, S. Hayashi, K. Domen and M. Hara, Catal. Today, 2006, 116, 157 CrossRef CAS; (d) M. H. Zong, Z. Q. Duan, W. Y. Lou, T. J. Smith and H. Wu, Green Chem., 2007, 9, 434 RSC.
  15. K. Konkala, N. M. Sabbavarapu, R. Katla, N. Y. Durga, V. K. Reddy, B. Devi and R. Prasad, Tetrahedron Lett., 2012, 53, 1968 CrossRef CAS.
  16. (a) N. Winterton, Chemistry for sustainable technologies: A foundation, RSC Publisher, 2010 Search PubMed; (b) M. Lancaster, Green chemistry: An introductory text, RSC Publisher, 2010 Search PubMed.
  17. (a) A. Corma and H. Garcia, Chem. Rev., 2003, 103, 4307 CrossRef CAS PubMed; (b) G. A. Somorjai and K. McCrea, Appl. Catal., A, 2001, 222, 3 CrossRef CAS; (c) B. W. Wojciechowski and A. Corma, Catalytic Cracking, Catalysts Kinetics and Mechanisms, Marcel Dekker, New York, 1984 Search PubMed; (d) H. U. Blaser and E. Schmidt, Asymmetric Catalysis on Industrial Scale: Challenges Approaches and Solutions, Wiley-VCH, Weinheim, 2004 Search PubMed; (e) E. N. Jacobsen, A. Pfaltz and H. Yamamoto, Comprehensive Asymmetric Catalysis, Springer, Heidelberg, 1999 Search PubMed; (f) S. V. Malhotra, Methodologies in Asymmetric Catalysis, ACS, Orlando, FL, 2004, vol. 880 Search PubMed; (g) R. Noyori, Angew. Chem., Int. Ed., 2002, 41, 2008 CrossRef CAS; (h) K. B. Sharpless, Angew. Chem., Int. Ed., 2002, 41, 2024 CrossRef CAS; (i) E. N. Jacobsen, W. Zhang, A. R. Muci, J. R. Ecker and L. Deng, J. Am. Chem. Soc., 1991, 113, 7063 CrossRef CAS; (j) P. T. Anastas and M. M. Kirchhoff, Acc. Chem. Res., 2002, 35, 686 CrossRef CAS PubMed; (k) M. Poliakoff, J. M. Fitzpatrick, T. R. Farren and P. T. Anastas, Science, 2002, 297, 807 CrossRef CAS PubMed; (l) A. Abad, P. Concepcio, A. Corma and H. Garcia, Angew. Chem., Int. Ed., 2005, 44, 4066 CrossRef CAS PubMed; (m) A. Corma and H. Garcia, Chem. Rev., 2002, 102, 3837 CrossRef CAS PubMed.
  18. B. Cornils, W. A. Herrmann, P. Panster and S. Wieland, Applied Homogeneous Catalysis with Organometallic Compounds, Wiley/VCH, Weinheim, 1996, p. 576 Search PubMed.
  19. (a) K. K. Bando, K. Soga, K. Kunimori and H. Arakawa, Appl. Catal., A, 1998, 175, 1 CrossRef; (b) K. Yuzaki, T. Yarimizu, S. Ito and K. Kunimori, Catal. Lett., 1997, 47, 3 CrossRef; (c) H. T. Ma, Z. Y. Yuan, Y. Wang and X. H. Bao, Surf. Interface Anal., 2001, 32, 224 CrossRef CAS; (d) A. J. Sandee, J. N. Reek, P. C. Kamer and P. W. Van Leeuwen, J. Am. Chem. Soc., 2001, 123, 8468 CrossRef CAS PubMed; (e) T. Tago, T. Hanaoka, P. Dhupatemiya, H. Hayashi, M. Kishida and K. Wakabayashi, Catal. Lett., 2000, 64, 27 CrossRef CAS.
  20. O. G. Da Silva, E. C. da Silva Filho, M. G. da Fonseca, L. N. H. Arakaki and C. Airoldi, J. Colloid Interface Sci., 2006, 302, 485 CrossRef CAS PubMed.
  21. A. Stein, B. J. Melde and R. C. Schroden, Adv. Mater., 2000, 12, 1403 CrossRef CAS.
  22. (a) Y. L. Gu, Green Chem., 2012, 14, 2091 RSC; (b) M. S. Singh and S. Chowdhury, RSC Adv., 2012, 2, 4547 RSC.
  23. (a) B. Karimi, M. Ghoreishi-Nezhad and J. H. Clark, Org. Lett., 2005, 7, 625 CrossRef CAS PubMed; (b) B. Karimi and D. Enders, Org. Lett., 2006, 8, 1237 CrossRef CAS PubMed; (c) C. S. Gill, W. Long and C. W. Jones, Catal. Lett., 2009, 131, 425 CrossRef CAS.
  24. (a) A. Corma and H. Garcia, Adv. Synth. Catal., 2006, 348, 1391 CrossRef CAS; (b) A. Taher, J.-B. Kim, J.-Y. Jung, W.-S. Ahn and M.-J. Jin, Synlett, 2009, 2477 CAS.
  25. R. Tayebee, M. M. Amini, M. Ghadamgahi and M. Armaghan, J. Mol. Catal. A: Chem., 2013, 366, 266 CrossRef CAS.
  26. R. Tayebee and M. Ghadamgahi, Arabian J. Chem. DOI:10.1016/j.arabjc.2012.12.001.
  27. M. Moosavifar, C. R. Chim., 2012, 15, 444 CrossRef CAS.
  28. A. Debache, M. Amimour, A. Belfaitah, S. Rhouati and B. Carboni, Tetrahedron Lett., 2008, 49, 6119 CrossRef CAS.
  29. E. Ramu, V. Kotra, N. Bansal, N. Bansal, R. Varala and S. R. Adapa, J. Chem., 2008, 1, 188 CAS.
  30. S. R. Narahari, B. R. Reguri, O. Gudaparthi and K. Mukkanti, Tetrahedron Lett., 2013, 53, 266 Search PubMed.
  31. B. Ahmed, R. A. Khan, Habibullah and M. Keshari, Tetrahedron Lett., 2009, 50(24), 2889–2892 CrossRef CAS.
  32. Y. Yu, D. Liu, C. Liu and G. Luo, Bioorg. Med. Chem. Lett., 2007, 17, 3508 CrossRef CAS PubMed.
  33. E. S. Putilova, G. V. Kryshtal, G. M. Zhdankina, N. A. Troitskii and S. G. Zlotin, Russ. J. Org. Chem., 2005, 41, 512 CrossRef CAS.
  34. H. Khabazzadeh, E. T. Kermani and T. Jazinizadeh, Arabian J. Chem., 2012, 5, 485 CrossRef CAS.
  35. M. Tajbakhsh, Y. Ranjbary, A. Masuodi and S. Khaksar, Chin. J. Catal., 2012, 33, 1542 CrossRef CAS.
  36. F. Tamaddon and S. Moradi, J. Mol. Catal. A: Chem., 2013, 117, 370 Search PubMed.
  37. H. Khabazzadeh, K. Saidi and H. Sheibani, ACS Med. Chem. Lett., 2008, 18, 278 CrossRef CAS PubMed.
  38. F. Shirini, M. Abedini and R. Pourhassan-Kisomi, Chin. Chem. Lett., 2014, 25, 111 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra02415g

This journal is © The Royal Society of Chemistry 2016