A green one-pot three-component synthesis of tetrahydrobenzo[b]pyran and 3,4-dihydropyrano[c]chromene derivatives using a Fe3O4@SiO2–imid–PMAn magnetic nanocatalyst under ultrasonic irradiation or reflux conditions

Mohsen Esmaeilpour*a, Jaber Javidi*bc, Farzaneh Dehghania and Fatemeh Nowroozi Dodejib
aChemistry Department, College of Science, Shiraz University, Shiraz, Iran. E-mail: m1250m551085@yahoo.com; Fax: +98 7112286008; Tel: +98 7116137738
bDepartment of Pharmaceutics, School of Pharmacy, Shahid Beheshti University of Medical Sciences, Tehran, Iran. E-mail: JaberJavidi@gmail.com
cStudents Research Committee, School of Pharmacy, Shahid Beheshti University of Medical Sciences, Tehran, Iran

Received 18th January 2015 , Accepted 9th March 2015

First published on 9th March 2015


Abstract

An efficient and environmentally benign procedure for the synthesis of tetrahydrobenzo[b]pyran and 3,4-dihydropyrano[c]chromene derivatives has been developed by a one-pot three-component reaction of various aldehydes, malononitrile, and dimedone or hydroxycoumarin in the presence of Fe3O4@SiO2–imid–PMAn nanoparticles as magnetic catalysts under ultrasonic irradiation or reflux conditions in water. This new procedure has notable advantages such as operational simplicity, excellent yields, short reaction time, and absence of any tedious workup or purification. In addition, the excellent catalytic performance in a water medium and the easy preparation, thermal stability and separation of the catalyst make it a good heterogeneous system and a useful alternative to other heterogeneous catalysts. Also, the catalyst can be easily recovered by a magnetic field and reused for eight consecutive reaction cycles without significant loss of activity.


1. Introduction

Today the efficiency of a chemical synthesis can be measured, not only by parameters such as selectivity and overall yield but also by its raw material, time, human resources, and energy requirements, as well as the toxicity and hazards of the chemicals and the protocols involved.1 Multi-component reactions (MCRs) have developed as an efficient and powerful tool in modern synthetic organic chemistry because the synthesis of complex organic molecules from simple and readily available substrates can be achieved in a very fast and efficient manner without the isolation of any intermediate.2 In addition, the employment of several transformations in a single manipulation is highly compatible with the goals of sustainable and “green” chemistry.3 Developing MCR procedures in an aqueous medium is an active area of research in this direction that have many advantages, such as devoid of any carcinogenic effects, reduced pollution, lower cost, and simplicity in processing, which are valuable to the industry as well as to the environment.4 On the other hand, the use of ultrasound irradiation represents very powerful green chemical protocols from both the economic and synthetic point of view.5 The use of ultrasonic irradiation accelerates an organic transformation at ambient conditions which otherwise require harsh conditions of temperature and pressure.6 The interaction between molecules and ultrasound isn't direct but the energy of theses long wavelength can cause cavitation which makes the reaction faster.7

Pyran derivatives have received significant attention due to their important biological and pharmacological properties.8 Among the pyrans, the substituted tetrahydrobenzo[b]pyran have a special importance among the 6-membered oxygen-containing heterocycles as they have been utilized in synthesis of blood anticoagulant warfarin,9 they have been also used as anticancer and antimicrobial agents,10 and photoactive materials.11

The conventional synthetic method for the preparation of tetrahydrobenzo[b]pyran derivatives is a three-component reaction of cyclic 1,3-diketones, aryl aldehydes, and activated methylene compounds under various reaction conditions, e.g., in the presence of piperidine and triethylamine in acetic acid or DMF.12 Other methods include use of microwaves,13 ultrasonic radiation,14 and electrogenerated base15 or using some catalysts such as triethylbenzylammonium chloride (TEBA),16 perfluorooctanoate,17 (S)-proline,18 silica bonded n-propyl-4-aza-1-azoniabicyclo[2.2.2]octane chloride,19 n-TiO2/H14[NaP5W30O110],20 nanozeoliteclinoptilolite,21 SO3H-bearing carbonaceous solid catalyst (PEG–SAC),22 multi-walled carbon nanotube supported Fe3O4 nanoparticles,23 free-ZnO nanoparticles,24 acetic acid functionalized imidazolium salts,25 Nano α-Al2O3 supported ammonium dihydrogenphosphate (NH4H2PO4/Al2O3)26 and amino functionalized ionic liquids.27 However, these methods show varying degrees of success as well as limitations such as long reaction times, low yields, and use of toxic solvents. Thus, the development of an alternate milder and cleaner procedure, which surpasses those limitations, is very much relevant for the synthesis of tetrahydrobenzo[b]pyran.

Also, Dihydropyrano[3,2-c]chromene and its derivatives are very useful compounds in various fields of chemistry, biology and pharmacology.28 Some of these compounds exhibit spasmolytic, diuretic, anticoagulant, anticancer, and antianaphylactic activity.28 Moreover they can be used as cognitive enhancers, for the handling of neurodegenerative diseases, including amyotrophic lateral sclerosis, Alzheimer's disease, Parkinson's disease, Huntington's disease, AIDS associated dementia and Down's syndrome as well as for the treatment of schizophrenia and myoclonus.29 Despite their importance from pharmacological, industrial and synthetic point of views, comparatively few methods for the preparation of dihydropyrano[3,2-c]chromene derivatives have been reported.30 Some of the reported procedures require long reaction times; afford products with only modest yields and non-reusability of the catalyst. Therefore, the development of more effective methods for their preparation is still necessary.

In recent years, Brønsted acids such as Keggin-type heteropolyacids (HPAs) have been used as efficient catalysts for a variety of organic reactions because of their redox properties and superacidic, high thermal stability, ease of handling, high proton mobility, stronger acids than homogeneous acid catalysts, low toxicity, development of clean technologies and low cost.31 Although HPAs are versatile compounds in their acidic form, their main disadvantages are high solubility in polar solvents and low surface area (<10 m2 g−1). Therefore, in a homogeneous reaction the isolation of the products and the reuse of the catalyst after reaction become difficult.32 Therefore, in order to overcome this problem, these materials disperse on supports (such as active carbon, silica, acidic ion-exchange resins and etc.) which possess large surface area. The use of support allows the heteropolyacids to be dispersed over a large surface area and increases their catalytic activity.33 In previous work,34 we introduce a simple, repaid, inexpensive and one step method, solvothermal, for synthesis of H3PMo12O4 nanoparticles (PMAn) from H3PMo12O4 bulk particles (PMAb). Acidity of as-prepared nanoparticles was investigated by pyridine adsorption method. Results showed acidity of PMA rise by declining particle size.

Magnetic nanoparticles (MNPs)such as magnetite (Fe3O4) have attracted great interest because of magnetic and electrical properties, high specific surface area, their unique catalytic and their wide applications, including drug delivery systems, targeted gene therapy, ion exchange separation, magnetic resonance imaging, biosensors, magnetic data storage, and environmental remediation and catalysis.35 For many applications, magnetic nanoparticles are suitable to be chemically stable and uniform in size. But the magnetic nanoparticles trend to aggregate due to their nanoscale and strong interaction between each other. This problem can be solved by coating of magnetite nanoparticles with a silica layer as the stabilizer, which prevents direct contact between the nanoparticles. Furthermore, the abound hydroxyl groups on the surface of composite particles provide the opportunity to conjugate various function molecules for many special applications.36 Recently, a number of functionalized Fe3O4 nanoparticles have been employed in a range of organic transformations, and the studies on immobilization of organo catalysts on silica coated iron oxide nanoparticles have been reported.37

Therefore, as part of our continuing interest in developing newer methods for the synthesis of useful compounds, we have recently successfully developed Fe3O4@SiO2–imid–PMAn as recyclable catalyst for the one-pot synthesis of 1-amidoalkyl-2-naphthols.38 Compared to other substrates (silica, acidic ion-exchange resins, active carbon and nanotitania), Fe3O4@SiO2–imid nanoparticles have various advantages such as high loading capacity, low leaching and simple and efficient recovery procedure. Fig. 1 presents the procedure for the preparation of Fe3O4@SiO2–imid–PMAn stepwise.


image file: c5ra01021g-f1.tif
Fig. 1 Process for preparation of Fe3O4@SiO2–imid–PMAn nanoparticles and characterization by TEM, SEM and DLS technique.

Encouraged by these efforts and aiming to show the efficiency and generality of Fe3O4@SiO2–imid–PMAn as catalysts for further, we have utilized this novel catalyst for the synthesis of tetrahydrobenzo[b]pyrans and dihydropyrano[3,2-c]chromenes from simple and easily available starting materials under much milder reaction conditions (Scheme 1).


image file: c5ra01021g-s1.tif
Scheme 1 Fe3O4@SiO2–imid–PMAn catalyzed synthesis of tetrahydrobenzo[b]pyran and dihydropyrano[3,2-c]chromene derivatives under ultrasonic irradiation or reflux conditions.

2. Results and discussion

In our previous work,38 the Fe3O4, Fe3O4@SiO2 and Fe3O4@SiO2–imid–PMAn nanocatalysts were characterized by various methods such as transmission electron microscopy (TEM), scanning electron microscopy (SEM), dynamic light scattering (DLS), Fourier transform infrared (FT-IR), vibrating sample magnetometer (VSM) and etc. As shown in Fig. 1 Fe3O4@SiO2–imid–PMAn nanoparticles have spherical shapes with approximately 50 nm diameters. The size distribution of these is centered at a value of 55 nm. The magnetic properties of Fe3O4, Fe3O4@SiO2, Fe3O4@SiO2–imid–PMAn nanoparticles were measured by VSM at room temperature. All the samples show a typical superparamagnetic behavior. Hysteresis phenomenon was not found and the magnetization and demagnetization curves were coincident. The saturation magnetization of Fe3O4, Fe3O4@SiO2 and Fe3O4@SiO2–imid–PMAn is 63.4, 39.7, 33.2 emu g−1, respectively.

Due to ability of Fe3O4@SiO2–imid–PMAn as a mild and efficient acid catalyst, we decided to apply this catalyst for synthesis of tetrahydrobenzo[b]pyrans. At the first stage, to obtain the best reaction conditions, the reaction of dimedone, benzaldehyde and malononitrile, as a model reaction was chosen. The model reaction was refluxed in the presence of 0.02 g of Fe3O4@SiO2–imid–PMAn and a variety of solvents such as H2O, EtOH, MeOH, EtOAc, CHCl3and CH3CN. The represented data in Table 1, showed the reaction proceeded efficiently in refluxing H2O and resulted in high yields of the desired product (Table 1, entry 6). This three-component condensation was also accomplished in protic solvents such as EtOH and MeOH under reflux and the corresponding products were obtained in 87% and 79% yield, respectively (Table 1, entries 1 and 3). Aprotic solvents such as CHCl3, EtOAc and CH3CN afforded the desired product in lower yields and longer reaction times (Table 1, entries 2, 4 and 5). Moreover, the model reaction was examined under solvent-free conditions at 100 °C and gave the corresponding product in 71% yield after 70 min (Table 1, entry 7).

Table 1 Optimization of the amount of catalyst, solvent and temperature in a one-pot synthesis of the model reactiona
Entry Catalyst amount (g) Solvent Condition Time (min) Yieldb (%)
a Reaction conditions: benzaldehyde (1 mmol), malonitrile (1.2 mmol), dimedone (1 mmol), Fe3O4@SiO2–imid–PMAn catalyst and solvent (5 mL).b Isolated yield.
1 0.02 EtOH Reflux 20 87
2 0.02 EtOAc Reflux 80 43
3 0.02 MeOH Reflux 30 79
4 0.02 CHCl3 Reflux 100 Trace
5 0.02 CH3CN Reflux 60 28
6 0.02 H2O Reflux 20 94
7 0.02 Solvent-free 100 °C 70 71
8 None H2O Reflux 360
9 0.005 H2O Reflux 60 33
10 0.01 H2O Reflux 60 67
11 0.015 H2O Reflux 30 87
12 0.025 H2O Reflux 20 92
13 0.02 H2O r.t 120 46
14 0.02 H2O 60 °C 90 77
15 0.02 H2O 80 °C 40 85
16 None H2O Sonication (40 kHz)/r.t 20
17 0.005 H2O Sonication (40 kHz)/r.t 15 54
18 0.01 H2O Sonication (40 kHz)/r.t 10 88
19 0.015 H2O Sonication (40 kHz)/r.t 7 96
20 0.02 H2O Sonication (40 kHz)/r.t 7 93


After that we checked the model reaction in presence of H2O as a green solvent under reflux conditions and various amount of catalyst. As it was shown in Table 1 the best result was obtained when we carried out the model reaction in the presence of 0.02 g of catalyst. This condensation was carried out with low amounts of Fe3O4@SiO2–imid–PMAn of 0.005, 0.01 and 0.015 g and the corresponding products were obtained in 33%, 67% and 87% yield, respectively (Table 1, entries 9–11). The effect of temperature was also studied by carrying out the model reaction in the presence of water as a solvent and 0.02 g of catalyst at room temperature, 60 °C and 80 °C. It was observed that the yield was increased when the reaction temperature was increased (Table 1, entries 13–15).

The reaction was also checked without the catalyst in which the reaction did not proceed even after 6 h (Table 1, entry 8). These observations established the crucial rule of Fe3O4@SiO2–imid–PMAn for the expedition of the reaction time and the product yield.

For investigation of the ability of ultrasonic irradiation for the acceleration of organic reactions, we examined the model reaction under ultrasonic irradiation at room temperature in presence of H2O as a green solvent and various amount of catalyst. As it is reveal from Table 1, the best results were obtained in the presence of 0.015 g of catalyst (Table 1, entry 19). However, synthesis of organic compounds under ultrasound irradiation has been limited by the need for a specialized apparatus that may not be available in many laboratories. Because of this limitation, herein we report both ultrasonic irradiation in H2O in presence of 0.015 g of catalyst at room temperature and also refluxing water in presence of 0.02 g of catalyst (Table 1, entries 6 and 19) for the synthesis of tetrahydrobenzo[b]pyrans.

After optimizing the reaction conditions, the generality of this catalytic system was confirmed by the employment of a series of aldehydes, dimedone and malononitrile to obtain desired products under the optimized conditions. The results are summarized in Table 2.

Table 2 Synthesis of derivatives of 2-amino-5-oxo-5,6,7,8-tetrahydro-4H-benzo[b]pyran in the presence of Fe3O4@SiO2–imid–PMAna

image file: c5ra01021g-u1.tif

Entry Ar Product H2O/reflux Ultrasonic/H2O/r.t M.p. °C (Lit.)
Time (min) Yieldb (%) Time (min) Yieldb (%)
a Reaction conditions: aldehyde (1 mmol), dimedone (1 mmol), malononitrile (1.2 mmol), water (5 mL, reflux), Fe3O4@SiO2–imid–PMAn (0.02 g) or water (5 mL, ultrasonic irradiation (40 kHz), r.t), Fe3O4@SiO2–imid–PMAn (0.015 g).b The yields refer to isolated products.
1 C6H5 4a 20 94 7 96 238–240 (227–239)39
2 4-Me-C6H4 4b 30 85 10 91 212–214 (214–216)39
3 2-Me-C6H4 4c 40 82 12 92 210–211 (210–212)26
4 4-MeO-C6H4 4d 25 87 10 94 201–203 (198–202)21
5 3-MeO-C6H4 4e 30 88 10 92 195–197 (196–198)26
6 4-Me2N-C6H4 4f 25 90 8 95 212–214 (210–213)21
7 3,4-(MeO)2-C6H3 4g 30 84 12 93 229–231 (227–229)24
8 2,3-(MeO)2-C6H3 4h 50 85 12 90 216–218 (216–217)21
9 4-OH-C6H4 4i 30 92 9 96 225–227 (224–226)12
10 4-OH-3-MeO-C6H3 4j 40 88 12 91 230–231 (227–229)24
11 4-Cl-C6H4 4k 10 95 5 97 212–213 (213–214)21
12 2-Cl-C6H4 4l 18 90 7 93 213–214 (212–213)21
13 4-NO2-C6H4 4m 10 96 4 97 179–181 (180–182)26
14 3-NO2-C6H4 4n 15 93 7 95 209–211 (210)4
15 2-NO2-C6H4 4o 10 93 5 92 237–238 (238–239)24
16 4-Cl-3-NO2-C6H3 4p 5 97 5 96 214–216 (216–217)24
17 4-CN-C6H4 4q 10 94 5 97 227–228 (225–228)26
18 4-F-C6H4 4r 12 96 5 95 188–189 (190–191)26
19 2,4-(Cl)2-C6H4 4s 10 95 4 97 192–194 (190–192)26
20 3-OPh-C6H4 4t 30 89 12 90 191–193 (193–194)19
21 2-Naphthalene 4u 25 92 8 93 232–233 (232–234)8
22 4-C3H7-C6H4 4v 40 90 14 91 190–191 (187–189)8
23 2-Thionyl 4w 35 92 10 94 221–222 (223–225)21
24 2-Furyl 4x 30 90 10 89 219–221 (220–223)21


As shown in Table 2, we investigated the reaction with a wide range of aromatic aldehydes with electron donating and electron withdrawing groups. Both electron-rich and electron-deficient aldehydes worked well and give high yields of products under both refluxing H2O and ultrasonic irradiation.

Electron-deficient aldehydes furnished excellent yields of the corresponding products in the shorter reaction times (Table 2, entries 11–19), but the aromatic aldehydes with electron-donating substituents carried out the reaction at longer reaction times in lower yields of the corresponding tetrahydrobenzo[b]pyrans (Table 2, entries 2–10). Moreover, it is clear that, under the same reaction conditions, reactions under ultrasonic irradiation led to shorter reaction times.

Also both of our methodologies have been successfully used for heteroaromatic aldehydes that are acid sensitive species such as thiophene-2-carbaldehyde and furan-2-carbaldehyde and the corresponding tetrahydrobenzo[b]pyrans were obtained in excellent yields without the formation of any byproduct (Table 2, entries 23 and 24).

Encouraged by these results, we replaced the dimedone with 4-hydroxycoumarin to explore its further applications toward the synthesis of dihydropyrano[3,2-c]chromene derivatives. Initially, we optimized the reaction conditions. For this purpose, the reaction of 4-hydroxycoumarin, benzaldehyde and malononitrile was chosen as a model reaction then the model reaction was refluxed in the presence of 0.025 g of Fe3O4@SiO2–imid–PMAn and a variety of solvents such as H2O, EtOH, MeOH, EtOAc, CHCl3, CH2Cl2, THF and CH3CN. As shown in Table 3, the best result in terms of the reaction time (15 min) and yield (96%) was obtained when the reaction was carried out in refluxing H2O (Table 3, entry 8). This condensation was also accomplished efficiently in refluxing EtOH and MeOH and resulted in high yields of the desired products (Table 3, entries 1, 2). The model reaction was also examined under solvent-free conditions at 100 °C and gave the corresponding product in 81% yield after 60 min (Table 3, entry 9). Due to green nature and cost effective of the reactions in H2O, we decided to choose H2O as a solvent in this condensation and checked the model reaction in presence of H2O as a solvent under reflux and various amount of catalyst. As shown in Table 3 the best result was obtained when the model reaction carried out in the presence of 0.025 g of catalyst. The results show clearly that the catalyst is effective for this transformation and in its absence; the reaction did not take place even after higher reaction time (Table 3, entry 13). Use of a higher amount of catalysts did not improve the yield (Table 1, entry 17) while a decrease in the amount of catalysts decreases the yield (Table 3, entry 13).

Table 3 Effect of solvent, temperature and catalyst amount on the condensation of benzaldehyde, malononitrile, and 4-hydroxycoumarina
Entry Catalyst amount (g) Solvent Condition Time (min) Yieldb (%)
a Reaction conditions: benzaldehyde (1 mmol), malonitrile (1.2 mmol), 4-hydroxycoumarin (1 mmol), Fe3O4@SiO2–imid–PMAn catalyst and solvent (5 mL).b Isolated yield.
1 0.025 EtOH Reflux 15 92
2 0.025 MeOH Reflux 25 87
3 0.025 CHCl3 Reflux 90 22
4 0.025 CH2Cl2 Reflux 90 Trace
5 0.025 CH3CN Reflux 60 48
6 0.025 THF Reflux 90 35
7 0.025 EtOAc Reflux 60 52
8 0.025 H2O Reflux 15 96
9 0.025 Solvent-free 100 °C 60 81
10 0.025 H2O r.t 100 56
11 0.025 H2O 60 °C 60 75
12 0.025 H2O 80 °C 30 88
13 None H2O Reflux 300
14 0.01 H2O Reflux 60 22
15 0.015 H2O Reflux 40 44
16 0.02 H2O Reflux 15 81
17 0.03 H2O Reflux 15 94
18 None H2O Sonication (40 kHz)/r.t 20
19 0.01 H2O Sonication (40 kHz)/r.t 15 44
20 0.015 H2O Sonication (40 kHz)/r.t 10 77
21 0.02 H2O Sonication (40 kHz)/r.t 8 94
22 0.025 H2O Sonication (40 kHz)/r.t 6 97
23 0.03 H2O Sonication (40 kHz)/r.t 8 96


We also checked the model reaction in presence of water as a solvent and 0.025 g of catalyst at room temperature, 60 °C and 80 °C and observed that the corresponding product was formed in longer reaction time and lower yield (Table 3, entries 10–12).

Because of the ability of ultrasonic irradiation for the synthesis of tetrahydrobenzo[b]pyrans, we examined the model reaction under ultrasonic irradiation at room temperature in presence of H2O as a green solvent and various amount of catalyst at room temperature. As shown in Table 3, the best results were obtained in the presence of 0.025 g of catalyst (Table 3, entry 22).

So, the best results were obtained when the reaction was carried out under ultrasonic irradiation in H2O at room temperature and also refluxing water in presence of 0.025 g of catalyst (Table 3, entries 8 and 22).

In order to develop the scope of these reactions, we conducted the reaction with a series of aromatic aldehydes bearing different substituent groups, heteroaromatic and aliphatic aldehydes under the determined optimized conditions. The results are summarized in Table 4.

Table 4 Synthesis of dihydropyrano[3,2-c]chromene derivatives using Fe3O4@SiO2–imid–PMAn as catalysta

image file: c5ra01021g-u2.tif

Entry Ar Product H2O/reflux Ultrasonic/H2O/r.t M.p. °C (Lit.)
Time (min) Yieldb (%) Time (min) Yieldb (%)
a Reaction conditions: aldehyde (1 mmol), 4-hydroxycoumarin (1 mmol), malononitrile (1.2 mmol), water (5 mL, reflux), or water (5 mL, ultrasonic irradiation (40 kHz), r.t), Fe3O4@SiO2–imid–PMAn (0.025 g).b The yields refer to isolated products.
1 C6H5 6a 15 96 6 97 260–261 (256–258)40
2 4-MeO-C6H4 6b 25 88 8 93 244–246 (247–249)19
3 4-Me-C6H5 6c 25 90 8 95 253–255 (254–255)19
4 4-OH-C6H5 6d 25 93 10 96 261–263 (258–260)41
5 3-OH-C6H5 6e 30 95 7 95 268–270 (269–270)42
6 4-OH-3-MeO-C6H5 6f 30 86 12 92 254–256 (256–257)41
7 4-Me2N-C6H4 6g 20 91 10 94 224–226 (224–225)19
8 3,4-(MeO)2-C6H3 6h 30 86 10 90 225–227 (228–230)43
9 3,4,5-(MeO)3-C6H2 6i 40 89 12 88 234–236 (236–238)42
10 4-Br-C6H4 6j 15 92 6 95 250–251 (252–254)40
11 4-Cl-C6H5 6k 12 95 5 96 265–267 (264–266)22
12 2-Cl-C6H5 6l 15 88 8 90 267–269 (266–268)44
13 2,4-(Cl)2-C6H4 6m 8 93 5 97 260–261 (258–259)40
14 4-F-C6H4 6n 10 93 6 91 260 (259–261)19
15 4-CN-C6H4 6o 12 96 6 95 281–283 (284–286)30
16 4-NO2-C6H5 6p 5 96 5 97 252–254 (250–252)41
17 3-NO2-C6H5 6q 10 91 6 93 266–267 (263–265)41
18 2-NO2-C6H5 6r 10 89 5 96 257–259 (258–260)42
19 2-Furyl 6s 30 87 12 92 255–256 (253–255)41
20 2-Thionyl 6t 30 92 10 90 225–227 (228–229)17
21 CH3(CH2)2 6u 40 90 12 93 240–242 (243–245)40
22 (CH3)2CH 6v 35 86 12 89 251–253 (250–252)19
23 Cyclohexyl 6w 40 91 12 94 283–285 (281–283)17


According to Table 4, aromatic aldehydes bearing electron-withdrawing groups (Table 4, entries 10–18) generally exhibit higher reactivity in this reaction compared to those bearing electron-releasing groups (Table 4, entries 2–9). In contrast, as seen in Table 4, the present method is not only suitable for aromatic aldehydes but can also successfully be applied to aliphatic and heteroaromatic aldehydes.

Heteroaromatic aldehydes such as thiophene-2-carbaldehyde and furan-2-carbaldehyde were also converted to the corresponding products in excellent yields under both of our methodologies (Table 4, entries 19 and 20).

Moreover, it is clear that, about all of above results, reactions under ultrasonic irradiation led to shorter reaction times. The generality of this catalytic system was also checked for aliphatic aldehydes and a high yield of desired product was obtained (Table 4, entry 21–23).

A proposed mechanism for the synthesis of tetrahydrobenzo[b]pyrans is outlined in ESI. Based on this mechanism, as shown in S1, the intermediate (I) is produced upon initial condensation of aldehyde with malononitrile under the catalytic activity of Fe3O4@SiO2–imid–PMAn.

Subsequent nucleophilic addition of dimedone to the intermediate (I) followed successively by intramolecular cyclization to the intermediate (II), and rearrangement to furnish the corresponding product. A similar mechanism may occur for the formation of dihydropyrano[3,2-c] chromene derivatives.

To compare the reactivity of the Fe3O4@SiO2–imid–PMAn with previously reported catalysts/reagents a comparative chart is presented in S2. Although, all the catalysts listed in S2 were able to produce good yields of corresponding products, however some of these reactions were carried out under harsh reaction conditions such as toxic solvents. Moreover, some of the catalysts listed in S2 are not reusable and thus, our reusable catalyst and green methodologies in water have been established as a better alternative compared to the reported methods for the fabrication of corresponding products.

We also compared the catalytic activity of Fe3O4@SiO2–imid–PMAn with H3PMo12O40 (PMAb) and nano H3PMo12O40 (PMAn) in these condensations and we observed the best results in terms of the reaction times and yields were obtained when the reaction was carried out in the presence of Fe3O4@SiO2–imid–PMAn as a catalyst (S2, entries 17–19 and 35–37).

Fe3O4@SiO2–imid–PMAn magnetic catalyst dispersed in H2O can be easily separated by external magnetic field within several minutes without the need for a centrifugation or filtration step or a tedious workup of the final reaction mixture. However, magnetic separation performance makes the nanoparticles more effective and convenient in applications.

The reusability and recovery of the catalyst are important issues, especially when the reactions use solid catalysts. Thus, the recovery and reusability of catalyst were investigated for the preparation of 2-amino-3-cyano-4-(4-chlorophenyl)-7,7-dimethyl-5-oxo-4H-5,6,7,8-tetrahydrobenzo[b]pyran (4k) and 2-amino-4-(4-chlorophenyl)-5-oxo-4H, 5H-pyrano[3,2-c]chromene-3-carbonitrile (6k) at reflux conditions or ultrasonic irradiation under optimized conditions in water (Table 5). The catalyst was recovered by a magnetic field and the remaining solid was washed with hot ethanol (2 × 10 mL). Then, the recovered catalyst was dried under vacuum at 60 °C. The performance of the recycled catalyst in reaction up to eight successive runs was shown in Table 5. These bar diagrams clearly suggest that the desired products were obtained in high yields without distinct deterioration in catalytic activity.

Table 5 Recyclability of Fe3O4@SiO2–imid–PMAn in the synthesis of 4k and 6k under the optimized conditions and PMAn leaching (%) in each reaction cyclea
Run 1 2 3 4 5 6 7 8
a (a) Reaction conditions: 4-chlorobenzaldehyde (1 mmol), dimedone (1 mmol), malononitrile (1.2 mmol), water (5 mL), reflux, Fe3O4@SiO2–imid–PMAn (0.02 g), 10 min. (b) Reaction conditions: 4-chlorobenzaldehyde (1 mmol), dimedone (1 mmol), malononitrile (1.2 mmol), water (5 mL), ultrasonic irradiation (40 kHz), Fe3O4@SiO2–imid–PMAn (0.015 g), r.t, 5 min. (c) Reaction condition: 4-chlorobenzaldehyde (1 mmol), 4-hydroxycoumarin (1 mmol), malononitrile (1.2 mmol), water (5 mL), reflux, Fe3O4@SiO2–imid–PMAn (0.025 g), 15 min. (d) Reaction condition: 4-chlorobenzaldehyde (1 mmol), 4-hydroxycoumarin (1 mmol), malononitrile (1.2 mmol), water (5 mL),ultrasonic irradiation (40 kHz), Fe3O4@SiO2–imid–PMAn (0.025 g), 5 min.
Yield (%)
a 95 95 94 94 94 92 92 92
b 97 97 96 95 95 95 92 92
c 95 95 95 93 93 92 92 92
d 96 96 95 94 94 93 93 93
[thin space (1/6-em)]
Leaching (%)
a 0.43 0.61 0.85 1.47 1.92 2.42 3.16 3.71
b 0.27 0.43 0.78 1.15 1.89 2.12 2.57 2.91
c 0.63 1.17 1.57 1.88 2.70 2.98 3.63 4.11
d 0.51 0.77 1.07 1.20 1.50 1.62 1.96 2.33


To determine the exact PMAn species responsible for the observed reactions and to measure the extent of molybdenum (Mo) leaching after the reactions, we have used the hot filtration test.45 For this aim, we have studied the model reaction between 4-chlorobenzaldehyde, malononitrile and dimedone under reflux conditions in water. The hot reaction mixture was filtered after 27% conversion (GC) to remove the catalyst. Continuation of the reaction upon the resulting filtrate under the same conditions showed 32% conversion (GC) after 6 h. This result shows that the amount of leaching of the catalyst into the reaction mixture should be low and confirms that the catalyst acts heterogeneously in the reaction.

To determine the degree of leaching of the metal from the heterogeneous catalyst, the catalyst was removed by using a magnetic field and the molybdenum amount in reaction medium after each reaction cycle was measured through Inductively Coupled Plasma (ICP) analyzer. The analysis of the reaction mixture by the ICP technique showed that the leaching of H3PMo12O40 was negligible (Table 5).

SEM and DLS images of the catalyst after the first recycle have been represented in Fig. 1. As shown in Fig. 1, Fe3O4@SiO2–imid–PMAn nanoparticles had an average diameter of 70 nm, were of uniform size, and showed good dispersity. Additionally, the hydrodynamic diameter of catalyst was investigated by DLS technique. This size distribution is centered at a value of 70 nm. Generally, leaching of H3PMo12O40 and increasing of catalyst size lead to decreases the yield after each cycle.

3. Conclusion

In conclusion, we have developed an easy, highly efficient and green methodology for the synthesis of tetrahydrobenzo[b]pyran and dihydropyrano[3,2-c]chromene derivatives via one-pot three-component reaction in the presence of Fe3O4@SiO2–imid–PMAn nanoparticles as efficient and magnetic catalyst under ultrasonic irradiation or reflux conditions in water.

This method gives notable advantages such as easy preparation, heterogeneous nature, thermal stability and easy separation of the catalyst, clean and simple procedure, excellent yields, short reaction time, easy product separation and purification, lower loading of catalyst compared with the other methods, and avoidance of using hazardous organic solvents that makes this method an instrumental alternative to the previous methodologies for the scale up of these one-pot three-component reactions. Furthermore, the catalyst is magnetically separable and eliminates the requirement of catalyst filtration after completion of the reaction, which represents a major advantage for reactions from an economic and environmental point of view. In addition, the catalyst used is easily recovered by using a permanent magnet and reused without any noticeable loss of activity after at least eight times.

4. Experimental

Chemicals were purchased from Merck and Aldrich Chemical Companies. All the solvents were distilled, dried and purified by standard procedures. Determination of the purity of the substrate and monitoring of the reaction were accomplished by thin-layer chromatography (TLC) on a silica-gel polygram SILG/UV 254 plates. The NMR spectra were recorded on a Bruker Avance DPX 250 MHz spectrometer (using DMSO-d6 or CDCl3 with TMS as the standard). Fourier transform infrared spectroscopy (FT-IR) analysis of the samples was taken on a Shimadzu FT-IR 8300 spectrophotometer and the sample and KBr were pressed to form a tablet. Magnetic characterization was carried out on a vibrating sample magnetometer (Meghnatis Daghigh Kavir Co., Iran) at room temperature. Dynamic light scatterings (DLS) were recorded on a HORIBA-LB550. Scanning electron microscopy (SEM) image was obtained on Philips XL-30ESEM. Sonication was performed using an ultrasound cleaning bath (KQ-250B, China) with a frequency of 40 Hz and voltage of 220 V. The C, H, N and S elemental analyses were carried out by the using a Thermo finigan Flash EA-1112 CHNSO rapid elemental analyzer. Melting points were determined by open capillary method and were uncorrected. Therefore, all of the products were characterized by FT-IR, 1H NMR and 13C NMR, and also by comparison with authentic samples.

4.1. General procedure

4.1.1. Preparation of Fe3O4@SiO2 core–shell. The core–shell Fe3O4@SiO2 nanospheres were prepared by a modified Stober method in our previous work (Fig. 1).46
4.1.2. Preparation of Fe3O4@SiO2–imid. Fe3O4@SiO2 (1 g) was added to the solution of 3-chloro triethoxy propylsilane (1 mmol, 0.241 g) and imidazole (1 mmol, 0.0680 g) in p-xylene (20 mL) and the resultant mixture was under reflux for 24 h under nitrogen atmosphere. After refluxing for about 24 h, the mixture was cooled to room temperature, filtered by an external magnet and the product was washed with p-xylene to remove any reacted species and dried at 70 °C for 6 h.38
4.1.3. Preparation of H3PMo12O40 nanoparticles (PMAn). PMAn nanoparticles were prepared in our previous work.34 In a typical procedure, 5 mmol of bulk H3PMo12O40 (PMAb) was dispersed in 50 mL n-octane and the resulting dispersion was stirred vigorously for 30 min at room temperature to form a homogeneous dispersion. This dispersion was transferred into a Teflon-lined stainless autoclave filling 80% of the total volume. The autoclave was sealed and maintained at 150 °C for 12 h. The autoclave was then cooled to room temperature. Finally, the resulted powder was filtered and washed for several times by octane, and dried in a vacuum at 80 °C for 12 h (Fig. 1).
4.1.4. Preparation of Fe3O4@SiO2–imid–PMAn. Fig. 1 represents the anchoring of PMAn onto Fe3O4@SiO2–imid. Fe3O4@SiO2–imid (1.0 g) was added to an acetonitrile solution of PMAn (1.0 mmol) in 20 mL was taken in a round-bottom flask. The mixture was refluxed for 24 h under nitrogen atmosphere. After 24 h, the mixture was filtered by an external magnet, washed with acetonitrile and dichloromethane, and dried at 70 °C for 6 h.38
4.1.5. General procedure for the synthesis of tetrahydrobenzo[b]pyrans. A mixture of aromatic aldehyde (1 mmol), malononitrile (1.2 mmol), 5,5-dimethyl-1,3-cyclohexanedione (1 mmol) and Fe3O4@SiO2–imid–PMAn catalyst (0.02 or 0.015 g) in water (5 mL) was refluxed in an oil bath or sonicated at room temperature for an appropriate period of time as indicated in Table 2. During the procedure, the reaction was monitored by TLC. Upon completion, the catalyst was removed by using a magnetic field. Then, the resulting solid product was recrystallized from ethanol to give pure products in high yields.
4.1.6. General procedure for the synthesis of 3,4-dihydropyrano[c]chromene derivatives. A mixture of malononitrile (1.2 mmol), aromatic aldehyde (1 mmol),4-hydroxycoumarin (1 mmol) and Fe3O4@SiO2–imid–PMAn catalyst (0.025 g) in water (5 mL)was refluxed in an oil bath or sonicated at room temperature for an appropriate time as mentioned in Table 4. The progress of the reaction was monitored by TLC, for disappearance of aldehyde. After completion of the reaction, the catalyst was removed by using a magnetic field. Then, for the purification of the product, the precipitates were washed with cold aqueous ethanol and recrystallized from ethanol to give the pure product.

Acknowledgements

Authors are grateful to the council of Iran National Science Foundation and University of Shiraz for their unending effort to provide financial support to undertake this work.

References

  1. N. Isambert, M. M. S. Duque, J.-C. Plaquevent, Y. Genisson, J. Rodriguez and T. Constantieux, Chem. Soc. Rev., 2011, 40, 1347–1357 RSC.
  2. (a) D. Strubing, H. Neumann, S. Klaus, S. Hubner and M. Beller, Tetrahedron, 2005, 61, 11333 CrossRef PubMed; (b) A. Davoodnia, M. M. Heravi, L. Rezaei-Daghigh and N. Tavakoli-Hoseini, Monatsh. Chem., 2009, 140, 1499 CrossRef CAS PubMed; (c) L. Yu, B. Chen and X. Huang, Tetrahedron Lett., 2007, 48, 925 CrossRef CAS PubMed; (d) A. Davoodnia, M. Bakavoli, R. Moloudi, M. Khashi and N. Tavakoli-Hoseini, Monatsh. Chem., 2010, 141, 867 CrossRef CAS.
  3. (a) B. M. Trost, Science, 1991, 254, 1471 CAS; (b) R. A. Sheldon, Pure Appl. Chem., 2000, 72, 1233 CrossRef CAS.
  4. F. M. Alaa, A. El-Latif, F. F. Ahmed and M. Amira, Chin. J. Chem., 2010, 28, 91 CrossRef.
  5. (a) N. M. A. El-Rahman, T. S. Saleh and M. F. Mady, Ultrason. Sonochem., 2009, 16, 70 CrossRef PubMed; (b) H. Zang, Y. Zhang, Y. Zhang and B. W. Cheng, Ultrason. Sonochem., 2010, 17, 495 CrossRef CAS PubMed; (c) D. Venzke, A. F. C. Flores, F. H. Quina, L. Pizzuti and C. M. P. Pereira, Ultrason. Sonochem., 2011, 18, 370 CrossRef CAS PubMed; (d) M. Shekouhy and A. Hasaninejad, Ultrason. Sonochem., 2012, 19, 307 CrossRef CAS PubMed.
  6. (a) A. E. Torkamani, P. Juliano, S. Ajlouni and T. K. Singh, Ultrason. Sonochem., 2014, 21, 951 CrossRef CAS PubMed; (b) M. Salavati-Niasaria, J. Javidi, F. Davar and A. Amini Fazl, J. Alloys Compd., 2010, 503, 500 CrossRef PubMed; (c) M. Salavati-Niasari and J. Javidi, J. Cluster Sci., 2012, 23, 1019 CrossRef CAS.
  7. (a) A. Kumar and R. A. Maurya, Synlett, 2008, 883 CrossRef CAS PubMed; (b) M. Salavati-Niasari, J. Javidi and F. Davar, Ultrason. Sonochem., 2010, 17, 870 CrossRef CAS PubMed.
  8. G. R. Green, J. M. Evans and A. K. Vong, in “In Comprehensive Heterocyclic Chemistry II”, ed. A. R. Katritsky, C. Rees and E. F. V. Scriven, Pergamon Press, Oxford, 1995, p. 469 Search PubMed.
  9. C. Wiener, C. H. Schroeder, B. D. West and K. P. Link, J. Org. Chem., 1962, 27, 3086 CrossRef CAS.
  10. (a) W. Kemnitzer, J. Drewe, S. Jiang, H. Zhang, Y. Wang, J. Zhao, S. Jia, J. Herich, D. Labreque, R. Storer, K. Meerovitch, D. Bouffard, R. Rej, R. Denis, C. Blais, S. Lamothe, G. Attardo, H. Gourdeau, B. Tseng, S. Kasibhatla and S. X. Cai, J. Med. Chem., 2004, 47, 6299 CrossRef CAS PubMed; (b) M. A. Al-Haiza, M. S. Mostafa and M. Y. El-Kady, Molecules, 2003, 8, 275 CrossRef CAS PubMed.
  11. K. Shibata, S. Takegawa, N. Koizumi, N. Yamakoshi, E. Shimazawa and I. Antiandrogen, Chem. Pharm. Bull., 1992, 40, 935 CrossRef CAS.
  12. (a) S. Gao, C. H. Tsai, C. Tseng and C.-F. Yao, Tetrahedron, 2008, 64, 9143 CrossRef CAS PubMed; (b) X. S. Wang, D. Q. Shi, S. J. Tu and C. S. Yao, Synth. Commun., 2003, 33, 119 CrossRef CAS PubMed.
  13. I. Devi and P. J. Bhuyan, Tetrahedron Lett., 2004, 45, 8625 CrossRef CAS PubMed.
  14. (a) K. Azizi, M. Karimi, H. R. Shaterian and A. Heydari, RSC Adv., 2014, 4, 42220 RSC; (b) S. J. Tu, H. Jiang, Q. Y. Zhang, C. B. Miu, D. Q. Shi, X. S. Wang and Y. Gao, Chin. J. Org. Chem., 2003, 23, 488 CAS.
  15. L. Fotouhi, M. M. Heravi, A. Fatehi and K. Bakhtiari, Tetrahedron Lett., 2007, 48, 5379 CrossRef CAS PubMed.
  16. L. M. Wang, J. H. Shao, H. Tian, Y. H. Wang and B. Liu, J. Fluorine Chem., 2006, 127, 97 CrossRef CAS PubMed.
  17. D. Q. Shi, S. Zhang, Q. Y. Zhuang, S. J. Tu and H. W. Hu, Chin. J. Org. Chem., 2003, 23, 877 CAS.
  18. S. Balalaie, M. Bararjanian, A. M. Amani and B. Movassagh, Synlett, 2006, 263 CrossRef CAS PubMed.
  19. A. Hasaninejad, M. Shekouhy, N. Golzar, A. Zare and M. M. Doroodmand, Appl. Catal., A, 2011, 402, 11 CrossRef CAS PubMed.
  20. D. Azarifar, S. M. Khatami and R. Nejat-Yami, J. Chem. Sci., 2014, 126, 95 CrossRef CAS.
  21. S. M. Baghbanian, N. Rezaeia and H. Tashakkorian, Green Chem., 2013, 15, 3446 RSC.
  22. S. Paul, S. Ghosh, P. Bhattacharyy and A. R. Das, RSC Adv., 2013, 3, 14254 RSC.
  23. A. F. Shojaei, K. Tabatabaeian, F. Shirini and S. Z. Hejazi, RSC Adv., 2014, 4, 9509 RSC.
  24. S. Banerjee and A. Saha, New J. Chem., 2013, 37, 4170 RSC.
  25. A. R. Moosavi-Zare, M. A. Zolfigol, O. Khaledian, V. Khakyzadeh, M. Darestani Farahanic and H. G. Kruger, New J. Chem., 2014, 38, 2342 RSC.
  26. B. Maleki and S. S. Ashrafi, RSC Adv., 2014, 4, 42873 RSC.
  27. Y. Peng and G. Song, Catal. Commun., 2007, 8, 111 CrossRef CAS PubMed.
  28. L. Bonsignore, G. Loy and D. Secci, Eur. J. Med. Chem., 1993, 28, 517 CrossRef CAS.
  29. C. S. Konkoy, D. B. Fick, S. X. Cai, N. C. Lan and J. F. W. Keana, PCT Int Appl WO 0075123, 2000transChem. Abstr., 2001, 134, 29313a Search PubMed.
  30. M. E. Sedaghat, M. Rajabpour Booshehri, M. R. Nazarifar and F. Farhadi, Appl. Clay Sci., 2014, 95, 55 CrossRef CAS PubMed.
  31. L. T. Aany Sofia, A. Krishnan, M. Sankar, N. K. Kala Raj, P. Manikandan, P. R. Rajamohanan and T. G. Ajithkumar, J. Phys. Chem. C, 2009, 113, 21114 Search PubMed.
  32. J. Javidi, M. Esmaeilpour and F. Nowroozi Dodeji, RSC Adv., 2015, 5, 308 RSC.
  33. M. Esmaeilpour, J. Javidi, F. Dehghani and F. Nowroozi Dodeji, New J. Chem., 2014, 38, 5453 RSC.
  34. (a) J. Javidi, M. Esmaeilpour, Z. Rahiminezhad and F. Nowroozi Dodeji, J. Cluster Sci., 2014, 25, 1511 CrossRef CAS; (b) J. Jaber and E. Mohsen, Colloids Surf., B, 2013, 102, 265 CrossRef CAS PubMed.
  35. M. Esmaeilpour, J. Javidi, F. Nowroozi Dodeji and M. Mokhtari Abarghoui, Transition Met. Chem., 2014, 39, 797 CrossRef CAS.
  36. M. Esmaeilpour, J. Javidi, F. Nowroozi Dodeji and M. Mokhtari Abarghoui, J. Mol. Catal. A: Chem., 2014, 393, 18 CrossRef CAS PubMed.
  37. M. Esmaeilpour, A. R. Sardarian and J. Javidi, Appl. Catal., A, 2012, 445–446, 359 CrossRef CAS PubMed.
  38. M. Esmaeilpour, J. Javidi and M. Zandi, Mater. Res. Bull., 2014, 55, 78 CrossRef CAS PubMed.
  39. T. S. Jin, A. Q. Wang, X. Wang, J. S. Zhang and T. S. Li, Synlett, 2004, 5, 871 CrossRef PubMed.
  40. M. Jitender Khurana, B. Nand and P. Saluja, Tetrahedron, 2010, 66, 5637 CrossRef PubMed.
  41. M. G. Dekamin, M. Eslami and A. Maleki, Tetrahedron, 2013, 69, 1074 CrossRef CAS PubMed.
  42. K. Niknam and A. Jamali, Chin. J. Catal., 2012, 33, 1840 CrossRef CAS.
  43. H. Mehrabi and H. Abusaidi, J. Iran. Chem. Soc., 2010, 7, 890 CrossRef CAS.
  44. J. Wang, D. Q. Shi, Q. Y. Zhuang, X. S. Wang and H. W. Hu, J. Chem. Res., 2004, 2004, 818 CrossRef.
  45. M. Esmaeilpour, A. R. Sardarian and J. Javidi, J. Organomet. Chem., 2014, 749, 233 CrossRef CAS PubMed.
  46. F. Dehghani, A. R. Sardarian and M. Esmaeilpour, J. Organomet. Chem., 2013, 743, 87 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra01021g

This journal is © The Royal Society of Chemistry 2015