A new type of SO3H-functionalized magnetic-titania as a robust magnetically-recoverable solid acid nanocatalyst for multi-component reactions

Elham Tabrizian and Ali Amoozadeh*
Department of Organic Chemistry, Faculty of Chemistry, Semnan University, Semnan 35131-19111, Iran. E-mail: aamozadeh@semnan.ac.ir

Received 21st August 2016 , Accepted 29th September 2016

First published on 30th September 2016


Abstract

SO3H-functionalized magnetic-titania nanoparticles (Fe3O4@TDI@TiO2–SO3H) have been synthesized by a two-step procedure, involving the covalent grafting of n-TiO2 to n-Fe3O4 via 2,4-toluene diisocyanate as a regioselective linker (n-Fe3O4@TDI@TiO2) and subsequent sulfonation using chlorosulfonic acid. The as-prepared nanocatalyst was characterized by Fourier transform infrared spectroscopy (FT-IR), thermogravimetric analysis (TGA), X-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM) and vibrating sample magnetometry (VSM). The catalytic activity of the nanocatalyst was assessed for the synthesis of benzimidazoquinazolinones and polyhydroquinolines, in which the reaction conditions were optimized by applying central composite design (CCD) through response surface methodology. The nanocatalyst could be separated from the reaction mixture easily by magnetic decantation and reused at least six times without a noticeable degradation in catalytic activity. To the best of our knowledge, there are no literature reports on applying experimental design to optimize the reaction conditions for benzimidazoquinazolinones synthesis.


1. Introduction

Catalysis is one of the most important fields in the realm of organic synthesis and has received extensive attention from both academia and industry.1 Homogeneous Brønsted acid catalysts such as H2SO4, H3PO4, HF, HCl, HBr, CH3COOH and CF3COOH are widely used in the large-scale synthesis of industrial and fine chemicals as well as in organic transformations.2,3 Unfortunately, these acids suffer some disadvantages due to handling problems, difficulty in their work-up and separation procedures, waste neutralization issues and so on. To address environmental and safety concerns, the replacement of corrosive and hazardous mineral acids with heterogeneous ones is one of the main tasks assigned to organic chemists for the sustainable production of chemicals.4–6

In recent years, a great deal of attention has been paid to the preparation of heterogeneous catalysts by immobilizing homogeneous ones on various solid supports like organic polymers,7–9 inorganic silica,5,10,11 metal oxides12–14 and natural supports.15–17 However, this immobilization usually decreases the selectivity and catalytic activity.18 This drawback can be overcome by using nanomaterials as heterogeneous supports. When the size of the support is decreased to the nanometer-scale, the surface-to-volume ratio is significantly increased, which results in high accessibility for the surface-bound catalytic sites per unit area.19

Among numerous solid supports, transition-metal oxide nanoparticles have great potential due to their large amount of surface hydroxyl groups that can be used as active sites for the preparation of solid acid catalysts.20,21 Nevertheless, separation of these nanoparticles from the reaction mixture is almost as difficult as separation using conventional methods.22 This problem is surmounted by using magnetic nanoparticles as heterogeneous supports, which can be readily separated by employing an external magnetic field. This technique is typically more effective than centrifugation or filtration as it prevents loss of the catalyst.23–26

In our research towards designing and synthesizing novel heterogeneous nanocatalysts for organic reactions,13,27–29 we had previously reported TiO2-coated magnetite nanoparticle-supported sulfonic acid as an efficient, magnetically separable heterogeneous solid acid catalyst,30 which was prepared through immobilizing SO3H groups on the surface of Fe3O4–TiO2 core–shell nanostructures. However, very recently, our group uncovered a unique process for the magnetization of metal oxides,31 in which magnetic nanoparticles were bound to nano-titania through the use of 2,4-toluene diisocyanate (TDI) as a covalent linker, due to regioselectivity of its two different isocyanate groups. The present development builds on our prior study, and aims to immobilize chlorosulfonic acid on magnetic-titania for the preparation of a non-toxic, low cost, highly efficient, stable, and magnetically recyclable Brønsted solid acid nanocatalyst. The performance of the prepared catalyst was examined in the synthesis of benzimidazoquinazolinones and polyhydroquinolines.

2. Results and discussion

The general synthetic route for the preparation of the Fe3O4@TDI@TiO2–SO3H solid acid nanocatalyst is schematically outlined in Scheme 1. Initially, Fe3O4 nanoparticles were prepared via a co-precipitation method.32 In the second step, Fe3O4 nanoparticles were treated with excessive TDI, in which the highly electrophilic carbon of the para-isocyanate group readily undergoes a nucleophilic reaction with the accessible surface hydroxyls of n-Fe3O4 by forming a urethane bond.33,34 The ortho-isocyanate group is left unreacted due to the different reactivity of the two isocyanate groups, which together lead to steric hindrance in TDI.35 Subsequently, the as-synthesized nano-TiO2 (ref. 36) was added. In the same way, the surface hydroxyl groups of n-TiO2 reacted with the preserved ortho-isocyanate group of n-Fe3O4@TDI, under a higher temperature and longer reaction time to prepare magnetic-titania. Ultimately, the functionalization of the remaining surface hydroxyl groups of n-TiO2 with chlorosulfonic acid led to the formation of Fe3O4@TDI@TiO2–SO3H, a novel type of solid acid nanocatalyst. The catalyst was then characterized by FT-IR, TGA, XRD, FE-SEM, VSM and acid–base titration.
image file: c6ra21048a-s1.tif
Scheme 1 Preparation of the Fe3O4@TDI@TiO2–SO3H nanocatalyst.

Characterization of Fe3O4@TDI@TiO2–SO3H nanoparticles using FT-IR spectroscopy

Fourier transform infrared spectroscopy (FT-IR) is one of the most effective techniques for characterizing the functionalization of nanoparticles. Therefore, the FT-IR spectrum for every step of the catalyst synthesis was analyzed (Fig. 1). According to our previous report,31 the deposition of TiO2 nanoparticles on n-Fe3O4@TDI is successfully achieved (curve e) due to the disappearance of the isocyanate peak at 2262 cm−1 and the differences in the fingerprinting of the IR bands in the range of 500–700 cm−1 compared to curve d. Immobilization of chlorosulfonic acid on the n-TiO2 surface is verified by the very broad band of the acidic group from 2800 to 3500 cm−1 and in the absorption range of 1034–1267 cm−1, which is related to O[double bond, length as m-dash]S[double bond, length as m-dash]O asymmetric and symmetric stretching.13,29
image file: c6ra21048a-f1.tif
Fig. 1 FT-IR spectra of (a) n-Fe3O4, (b) n-TiO2, (c) TDI, (d) n-Fe3O4@TDI, (e) n-Fe3O4@TDI@TiO2, (f) n-Fe3O4@TDI@TiO2–SO3H.

Thermogravimetric analysis

The thermogravimetric analysis (TGA) curve suggests the grafting of organic groups onto the surface of the supports through mass loss upon heating. As shown in Fig. 2, the TGA curves of both Fe3O4@TDI@TiO2 (curve a) and Fe3O4@TDI@TiO2–SO3H (curve b) are similar, but differ in the amount of mass loss and the temperature range. The TGA curve of Fe3O4@TDI@TiO2–SO3H in the temperature range of 30–800 °C exhibits three mass losses. The initial mass loss up to 150 °C is caused by the removal of physically trapped solvents and water. The mass loss of about 10.74% in the range of 150–460 °C and that of 6.32% in the range of 460–650 °C, are attributed to the thermal decomposition of TDI and sulfonic acid groups, and are similar to those reported in the literature.13,37
image file: c6ra21048a-f2.tif
Fig. 2 TGA curves of (a) n-Fe3O4@TDI@TiO2, (b) n-Fe3O4@TDI@TiO2–SO3H.

X-ray diffraction

The phase purity and crystal structure of the catalyst were analyzed via XRD. The cubic Fe3O4 structure is confirmed by the peaks at values of 2θ equal to 30.4 (220), 35.7 (311), 43.4 (400), 54.1 (422), 57.1 (511) and 62.9 (440), in accordance with JCPD 79-0417. The anatase crystalline phase of TiO2 is matched with JCPD 89-4921 by the diffraction signals located at 2θ values equal to 25.2 (101), 37.3 (103), 37.8 (004), 38.7 (112), 48.2 (200), 54.0 (105), 55.1 (211), 62.8 (204), 68.9 (116), 70.4 (220), 75.1 (215) and 83.1 (224). The same set of characteristic peaks for Fe3O4 and TiO2 is detected in the XRD pattern of the catalyst (Fig. 3d), indicating that modification has occurred and the crystallite phases are retained after treatment with acidic reagent.
image file: c6ra21048a-f3.tif
Fig. 3 The X-ray diffraction patterns of (a) n-Fe3O4, (b) n-TiO2, (c) n-Fe3O4@TDI@TiO2, (d) n-Fe3O4@TDI@TiO2–SO3H.

Field emission scanning electron microscopy

The surface morphology, particle shape and size distribution features of n-Fe3O4, n-TiO2, n-Fe3O4@TDI@TiO2 and n-Fe3O4@TDI@TiO2–SO3H were examined using FE-SEM (Fig. 4). The micrographs illustrate that the catalyst particles are quasi-spherical with a larger average diameter in comparison to those of n-Fe3O4 and n-TiO2. However, aggregation of the nanoparticles is observed, which occurs during the functionalization process. The results show the presence of the catalyst on the nanometer-sized particles.
image file: c6ra21048a-f4.tif
Fig. 4 FE-SEM images of (a) n-Fe3O4, (b) n-TiO2, (c) n-Fe3O4@TDI@TiO2, (d) n-Fe3O4@TDI@TiO2–SO3H.

Vibrating sample magnetometry

The magnetic properties of the catalyst were examined at room temperature by using a vibrating sample magnetometer (VSM). The saturation magnetization value of n-Fe3O4@TDI@TiO2–SO3H is 31.9 emu g−1, lower than that of the bare magnetic nanoparticles (68.4 emu g−1), due to presence of a non-magnetic layer on the particle surface. However, this value still reveals its magnetic behavior, and confirms that the magnetization is still large enough for easy catalyst separation by using a magnetic stirring bar in the reaction mixture (Fig. 5).
image file: c6ra21048a-f5.tif
Fig. 5 Magnetization curves of the (a) n-Fe3O4, (b) n-Fe3O4@TDI@TiO2, (c) n-Fe3O4@TDI@TiO2–SO3H measured at room temperature.

Acid–base titration

The quantity of sulfonic acid groups on the surface of n-Fe3O4@TDI@TiO2–SO3H was determined using ion-exchange pH analysis. To do this, 100 mg of the catalyst was added to 10 mL of a 1 M NaCl solution and was stirred continuously for 24 h at room temperature. The catalyst was separated using an external magnet, and the NaCl solution was decanted and recovered. Then, two drops of a phenolphthalein solution were added to the NaCl solution and the solution was titrated against 0.01 M NaOH solution to neutrality. According to the obtained results, the loading of acidic sites was about 2.70 mmol H+ per g of catalyst.

Evaluating the catalytic activity of the Fe3O4@TDI@TiO2–SO3H nanoparticles

Nitrogen-containing heterocycles are of considerable interest in the agrochemical and pharmaceutical industries due to their diverse range of physiological properties. Benzimidazoquinazolinones, including both biodynamic benzimidazole and quinazoline heterosystems, are an important class of N-heterocycles that exhibit a broad spectrum of biological activities, such as antineoplastic, antihypertensive, anticancer, antihistaminic, anti-HIV and analgesic properties.38–40 In view of the importance of this type of heterocycle, the performance of n-Fe3O4@TDI@TiO2–SO3H as a nanocatalyst was probed for the synthesis of benzimidazoquinazolinones via the condensation reaction between benzaldehyde, dimedone and 2-aminobenzimidazole under solvent-free conditions as a model reaction (Scheme 2).
image file: c6ra21048a-s2.tif
Scheme 2 Synthesis of benzimidazoquinazolinones using n-Fe3O4@TDI@TiO2–SO3H as a nanocatalyst.

In order to reach optimal conditions, a response surface model (RSM) was employed and the effects of the catalyst amount (X1), temperature (X2) and reaction time (X3) on the reaction yield were investigated. For each factor, 5 levels were defined. These values and their coded values are shown in Table 1.

Table 1 Parameter levels and coded values of CCD
Independent variables Levels
Lowest (−1.68) Low (−1) Central (0) High (+1) Highest (+1.68)
X1: catalyst amount (mg) 0 5 10 15 20
X2: temperature (°C) 30 50 70 90 110
X3: time (min) 4 8 12 16 20


The experimental design matrix and the corresponding experimental parameters and response values are shown in Table 2.

Table 2 Conditions of predicted experiments in CCD
Run X1 X2 X3 Yield (%)
1 −1.68 0 0 45.09
2 0 0 0 85.75
3 0 0 0 81.08
4 0 −1.68 0 39.45
5 0 0 +1.68 86.02
6 +1 +1 −1 86.77
7 0 0 0 81.23
8 +1 −1 −1 64.22
9 −1 +1 −1 72.65
10 0 0 0 84.67
11 −1 −1 −1 40.10
12 −1 +1 +1 94.76
13 0 0 −1.68 59.10
14 0 0 0 86.48
15 +1.68 0 0 81.87
16 +1 +1 +1 96.33
17 0 0 0 80.01
18 +1 −1 +1 58.98
19 −1 −1 +1 52.00
20 0 +1.68 0 94.30


The analysis of variance (ANOVA) was used to determine the statistical significance of the constructed models and to evaluate which factor(s) significantly affect the response. From the ANOVA, as shown in Table 3, the p-values of the model and the lack of fit are lower and higher than 0.05, respectively, implying that the fitted model is significant, with a confidence level of 95%. Additionally, the results imply that all the independent variables (X1, X2 and X3) and their quadratic terms (X12, X22 and X32) significantly affect the response value (yield), and a significant interaction between the catalyst amount (X1) and time (X3) is obvious (P < 0.05). Also, the R-squared and adj R-squared values are close to 1.0, suggesting a strong correlation between the observed and the predicted values and indicating that the regression model affords an excellent explanation of the relationship between the independent variables and the response.

Table 3 Analysis of variance for the response surface quadratic model for yield
Source Sum of squares df Mean square F value p-Value prob > F
Model 6247.37 9 694.15 32.50 <0.0001
X1: catalyst amount 864.33 1 864.33 40.46 <0.0001
X2: temperature 3788.31 1 3788.31 177.35 <0.0001
X3: time 511.80 1 511.80 23.96 0.0006
X1X2 29.68 1 29.68 1.39 0.2658
X1X3 110.19 1 110.19 5.16 0.0465
X2X3 78.19 1 78.19 3.66 0.0848
X12 540.19 1 540.19 25.29 0.0005
X22 349.14 1 349.14 16.34 0.0024
X32 122.21 1 122.21 5.72 0.0378
Residual 213.61 10 21.36    
Lack of fit 175.63 5 35.13 4.63 0.0591
Pure error 37.97 5 7.59    
Cor total 6460.97 19      
R2 = 0.97 Adj-R2 = 0.94 Pred-R2 = 0.78


The polynomial response surface models for yield, based on significant levels and actual values were obtained from the ANOVA:

Y (yield) = −123.45 + 10.06X1 + 2.28X2 + 5.02X3 − 0.02X1X2 − 0.18X1X3 + 0.04X2X3 − 0.24X12 − 0.01X22 − 0.18X32

In this equation, the coefficients show the effect of the parameters. Therefore, the amount of catalyst can be considered as the main factor influencing the progression of the reaction.

Response surface plots facilitate the study of how process variables interact and affect the response by keeping one variable at zero and varying the other variables within the experimental range. 3D response plots of yield versus a pair of parameters are shown in Fig. 6. According to their p-values, the interaction of X1X3 is significant, indicating that a simultaneous increase in the amount of catalyst and the length of time cause an increase in the product yield.


image file: c6ra21048a-f6.tif
Fig. 6 3-D response surfaces for the effect of factors on the reaction yield.

The main objective of this design is to determine the best reaction conditions. Based on the experimental design obtained using Design-Expert 7.0.0, the model predicted that the yield could reach 96.47% under the optimal process conditions of 12 mg of catalyst and a temperature of 85 °C, with a reaction time of 13 min.

With the optimum conditions in hand, in order to examine the generality of the reaction conditions, the study was extended to different aromatic aldehydes and 1,3-cyclic diketones as shown in Table 4.

Table 4 Synthesis of benzimidazoloquinazolines using n-Fe3O4@TDI@TiO2-SO3Ha
Entry R1 R2 R3 Time (min) Yieldb (%) Melting point (°C)
a Reaction conditions: 1 (1 mmol), 2 (1 mmol), 3 (1 mmol), Fe3O4@TDI@TiO2–SO3H (12 mg) under solvent-free condition at 85 °C.b Isolated yield.
1 CH3 CH3 –H 13 96 318–320 (ref. 41)
2 CH3 CH3 4-CH3 15 91 330–332 (ref. 28)
3 CH3 CH3 4-OCH3 17 88 317–318 (ref. 41)
4 CH3 CH3 4-OH 15 90 329–331 (ref. 41)
5 CH3 CH3 4-NO2 13 92 332–334 (ref. 28)
6 CH3 CH3 4-F 12 95 325–326 (ref. 28)
7 CH3 CH3 4-Cl 12 91 339–341 (ref. 41)
8 CH3 CH3 4-Br 14 93 314–315 (ref. 28)
9 CH3 CH3 4-N(CH3)2 15 90 >300 (ref. 42)
10 H H –H 15 93 310–312
11 H H 4-CH3 15 90 >300 (ref. 39)
12 H H 4-OCH3 18 90 >300 (ref. 39)
13 H H 4-OH 16 92 >300 (ref. 39)
14 H H 4-NO2 13 91 >300 (ref. 39)
15 H H 4-F 13 94 >300 (ref. 39)
16 H H 4-Cl 14 94 >300 (ref. 39)
17 H H 4-Br 15 90 >300 (ref. 39)
18 H H 4-N(CH3)2 13 92 >300 (ref. 39)
19 H Ph –H 15 95 332–335 (ref. 28)


The results illustrate that the reaction proceeded very well within a relatively short reaction time for 1,3-cyclic diketones and both electron-releasing and electron-withdrawing substituents of aromatic aldehydes. This confirms the high efficiency of the nanocatalyst for the synthesis of a series of structurally diverse benzimidazoloquinazolines in high purity. The formation and purity of the products were confirmed by comparing their melting points with the literature values, and they were in good accordance with them. The structures of some of them were well characterized using FT-IR, 1H NMR and 13C NMR spectral data.

The suggested mechanism for the synthesis of benzimidazoloquinazolines is outlined in Scheme 3. At the beginning, 1,3-cyclic ketone (I) turns into its enol form (II) and reacts with the carbonyl group of aldehyde (III), which becomes protonated and activated due to the solid acid nanocatalyst and affords (IV). Then, an α,β-unsaturated carbonyl compound forms upon the loss of water (V). Next, the α,β-unsaturated carbonyl compound undergoes a Michael addition reaction with the endocyclic nitrogen in 3-aminobenzimidazole (VI), yielding the corresponding intermediate (VII). Finally, intra-molecular cyclization and loss of a water molecule (VIII) leads to the formation of the desired enamine product (VIIII).


image file: c6ra21048a-s3.tif
Scheme 3 Probable mechanism for the synthesis of benzimidazoloquinazolines in the presence of n-Fe3O4@TDI@TiO2–SO3H.

The promising results mentioned above encouraged us to extend further the scope of the catalytic performance of n-Fe3O4@TDI@TiO2–SO3H to the synthesis of polyhydroquinolines, through the one-pot four-component Hantzsch condensation of 1,3-cyclic diketones, aromatic aldehydes, ethyl acetoacetate and ammonium acetate under the optimized reaction conditions for benzimidazoquinazolinones (Scheme 4).


image file: c6ra21048a-s4.tif
Scheme 4 Synthesis of polyhydroquinolines using n-Fe3O4@TDI@TiO2–SO3H as a nanocatalyst.

As shown in Table 5, different aromatic aldehydes bearing electron-withdrawing or electron-releasing substituents and various 1,3-cyclic diketones gave the desired product in good to excellent yields in a short reaction time.

Table 5 Synthesis of polyhydroquinoline derivatives using n-Fe3O4@TDI@TiO2–SO3Ha
Entry R1 R2 R3 Time (min) Yieldb (%) Melting point (°C)
a Reaction conditions: 1 (1 mmol), 2 (1 mmol), 5 (1 mmol), 6 (2.5 mmol), n-Fe3O4@TDI@TiO2–SO3H (12 mg) under solvent-free condition at 85 °C.b Isolated yield.
1 CH3 CH3 –H 15 95 200–202 (ref. 43)
2 CH3 CH3 4-CH3 16 92 284–286 (ref. 44)
3 CH3 CH3 4-OCH3 18 90 256–258 (ref. 43)
4 CH3 CH3 4-OH 18 92 232–235 (ref. 43)
5 CH3 CH3 2-NO2 18 88 201–203 (ref. 44)
6 CH3 CH3 3-NO2 15 91 170–172 (ref. 44)
7 CH3 CH3 4-NO2 14 94 209–211 (ref. 44)
8 CH3 CH3 4-F 10 95 181–183 (ref. 44)
9 CH3 CH3 4-Cl 12 92 245–247 (ref. 43)
10 H H –H 13 94 240–241 (ref. 43)
11 H H 4-CH3 17 90 240–242 (ref. 45)
12 H H 4-OCH3 16 88 193–194 (ref. 45)
13 H H 4-OH 20 89 222–223 (ref. 45)
14 H H 2-NO2 20 90 189–191 (ref. 44)
15 H H 3-NO2 15 92 200–202 (ref. 44)
16 H H 4-NO2 12 94 204–205 (ref. 45)
17 H H 4-F 10 90 243–245 (ref. 45)
18 H H 4-Cl 10 93 233–234 (ref. 45)
19 H Ph –H 18 87 213–214 (ref. 44)


A possible mechanism for the synthesis of polyhydroquinolines is outlined in Scheme 5. First of all, the 1,3-cyclic dicarbonyl compound (I) converts into its enol form after tautomerisation (II) and reacts with the carbonyl group of aldehyde which is activated by the catalyst (III), to afford a knoevenagel intermediate (V). On the other side, ammonia resulting from ammonium acetate (VI) and the β-ketoester activated by the catalyst (VII) yields enamine (VIII). Afterward, the reaction between the intermediates (V) and (VIII) gives the intermediate (IX). Subsequent tautomerization, intramolecular N-cyclization and dehydration yields the hexahydroquinoline.


image file: c6ra21048a-s5.tif
Scheme 5 Probable mechanism for the synthesis of polyhydroquinolines in the presence of n-Fe3O4@TDI@TiO2–SO3H.

To highlight the advantages and effectiveness of the prepared catalyst, a comparison of its catalytic activity with some previous methods is presented (Table 6). This shows that n-Fe3O4@TDI@TiO2–SO3H is more or less superior compared with previously reported catalytic systems in terms of easy separability, high activity and recyclability, in addition to high product yields in short reaction times under neat conditions. These results considerably verify the high efficiency of the present catalyst.

Table 6 Comparison of n-Fe3O4@TDI@TiO2–SO3H with other reported catalysts for the synthesis of benzimidazoquinazolinones
Entry Catalyst Reaction condition Time (min) Yield (%) Ref.
1 I2 (10 mol%) CH3CN/reflux 10 85 46
2 p-TsOH–H2O (15 mol%) CH3CN/50 °C 25 95 47
3 [bmim+][BF4] (3 mL) 90 °C 420 86 48
4 Fe–chitosan (20 mg) Ethanol/40 °C 105 82 49
5 MCM-41-SO3H (20 mol%) Ethanol/reflux 30 97 50
6 n-Fe3O4@TDI@TiO2–SO3H (12 mg) Solvent-free/85 °C 13 96 This work


Recycling of the catalyst

In order to develop a greener organic synthesis, the capability of the catalyst to be recovered and reused should be seriously considered. In other words, the catalyst should be able to be separated from the reaction mixture, and be reused several times. Therefore, the recoverability of n-Fe3O4@TDI@TiO2–SO3H was investigated for the model reaction of benzimidazoloquinazoline. As shown in Fig. 7, the catalyst can be reused at least six times and a conversion of 89% was still attained in the 6th run.
image file: c6ra21048a-f7.tif
Fig. 7 Reusability of n-Fe3O4@TDI@TiO2–SO3H for the synthesis of benzimidazoloquinazoline.

After the fifth cycle, the recovered catalyst had a similar morphology to the fresh one (Fig. 8) and no noticeable change in structure was observed, by comparing its FT-IR spectrum to that of the fresh catalyst (Fig. 9).


image file: c6ra21048a-f8.tif
Fig. 8 FE-SEM images of (a) fresh n-Fe3O4@TDI@TiO2–SO3H, (b) n-Fe3O4@TDI@TiO2–SO3H after being reused six times.

image file: c6ra21048a-f9.tif
Fig. 9 FT-IR spectra of n-Fe3O4@TDI@TiO2–SO3H before use (a) and after being reused six times (b).

3. Experimental

Materials and instruments

All chemicals and reagents were purchased commercially from the Aldrich and Merck chemical companies. Solvents were dried by conventional methods. TDI was industrial grade, containing an 80[thin space (1/6-em)]:[thin space (1/6-em)]20 mixture of 2,4 and 2,6 isomers and was used as received. The progress of the reactions was monitored using thin layer chromatography (TLC) with silica gel 60 F254 glass plates. Melting points were measured on a Thermo scientific 9100 apparatus. FT-IR were recorded on a Shimadzu-8400 spectrometer in the range of 400–4000 cm−1 using KBr pellets. Du Pont 2000 thermal analysis apparatus was used for TGA, heated from 25 °C to 800 °C with a heating rate of 10 °C min−1 under an air atmosphere. XRD patterns were recorded at room temperature on a Siemens D5000 (Siemens AG, Munich, Germany) using Cu-Kα radiation. FE-SEM images were collected using a Philips XL30 instrument (Royal Philips Electronics, Amsterdam, The Netherlands) at 30 kV. The magnetite measurement was performed on a vibrating sample magnetometer (4 inch, Daghigh Meghnatis Kashan Co., Kashan, Iran) at room temperature. The NMR spectra were recorded on a Bruker Avance 300 MHz instrument.

Preparation of the magnetic Fe3O4 nanoparticles

The magnetic Fe3O4 nanoparticles were synthesized using co-precipitation method according to reported literature.32 FeCl3·6H2O (5.838 g, 0.0216 mol) and FeCl2·4H2O (2.147 g, 0.0108 mol) were dissolved in 100 mL deionized water at 80 °C under vigorous stirring and N2 atmosphere. Then, 25% aqueous ammonia (10 mL) was added into the reaction mixture, which resulted in an immediate black precipitation of MNPs. After 30 min, the reaction was cooled to room temperature and the precipitate was separated by applying an external magnet. It was then washed several times with distilled water and dried for 48 h.

Functionalization of Fe3O4 nanoparticles with TDI (n-Fe3O4@TDI)

Both Fe3O4 nanoparticles (1.0 g) and TDI (1.40 g) were dissolved in 50 mL dried toluene in an ultrasonic bath for 10–15 min. Next, the reaction mixture was stirred magnetically at 95 °C for 20 h under an atmosphere of nitrogen. The powder product was separated magnetically and was carefully washed with dry toluene to remove the unreacted and physically absorbed TDI. It was then dried under vacuum at 100 °C for 4 h.

Preparation of TiO2 nanoparticles

Nano-TiO2 was synthesized according to the previously reported procedure, using a hydrothermal method.36 NH3·H2O was added to a solution of TiCl4 until the pH value reached 1.8. After 2 h stirring at 70 °C, the final solution pH was adjusted to 6 and was aged for 24 h at ambient temperature. The final powder was filtered, and rinsed with NH4Ac–HOAc until no Cl was detected. The precipitate was separated by centrifugation, washed with ethanol and dried in a vacuum. Treatment at 650 °C for 2 h attained the TiO2 nanoparticles.

Preparation of n-Fe3O4@TDI@TiO2

1.0 g n-Fe3O4@TDI was dispersed in 100 mL dried toluene. Next, 0.3 g n-TiO2 was added and the dispersion was heated at 110 °C for 48 h under constant stirring. The product, n-Fe3O4@TDI@TiO2, was separated using a permanent magnet, washed with acetone and dried at 100 °C for 4 h.

Preparation of n-Fe3O4@TDI@TiO2–SO3H

1.0 g Fe3O4@TDI@TiO2 nanoparticles were dispersed in 15 mL dry CH2Cl2. Chlorosulfonic acid (0.25 mL, 3.75 mmol) was added drop wise at room temperature over a period of 30 min. Upon completing the addition, the mixture was continuously stirred for 1 h to allow complete evolution of HCl from the reaction mixture. Then, the solvent was removed under reduced pressure and the powder was washed three times with ethanol (10 mL) to remove the unattached acid and was dried at 100 °C for 5 h.

General procedure for the synthesis of benzimidazoloquinazoline derivatives (4)

1,3-Cyclic diketone (1 mmol), aromatic aldehyde (1 mmol), 2-aminobenzimidazole (1 mmol) and catalyst (12 mg) were stirred under solvent free conditions at 85 °C. Upon completion of the reaction (monitored by TLC), the catalyst was separated by applying an external magnet (within 5 s). The reaction mixture was decanted an eluted using hot ethanol (5 mL). The products were obtained by recrystallization using ethanol solution. Percentage yields of the products were calculated as follows:
% yield = actual yield (g)/theoretical yield (g) × 100

General procedure for the synthesis of polyhydroquinoline derivatives (7)

1,3-Cyclic diketone (1 mmol), aromatic aldehyde (1 mmol), ethyl acetoacetate (1 mmol), ammonium acetate (2.5 mmol) and catalyst (12 mg) were mixed under neat condition at 85 °C in an oil bath. Upon completion of the reaction (monitored by TLC), the catalyst was separated by employing an external magnet (within 5 s). The reaction mixture was decanted and eluted using hot ethanol (5 mL). The products were obtained by recrystallization using ethanol solution. Percentage yields of the products were calculated as follows:
% yield = actual yield (g)/theoretical yield (g) × 100

General procedure for recycling n-Fe3O4@TDI@TiO2–SO3H

For evaluating the recoverability of the catalyst, the condensation of dimedone (1 mmol), benzaldehyde (1 mmol) and 2-aminobenzimidazole (1 mmol) under the optimized conditions was chosen as a model reaction. After completion of the reaction, the catalyst was separated using an external magnet, rinsed thoroughly with ethanol (2 × 5 mL) and acetone (2 × 5 mL) and dried at 70 °C for 2 h to be used for subsequent runs.

4. Conclusions

SO3H-functionalized magnetic-titania nanoparticles (n-Fe3O4@TDI@TiO2–SO3H) were successfully synthesized and well characterized using FT-IR, TGA, XRD, FE-SEM, VSM and acid–base titration. FT-IR and TGA confirmed the successful immobilization of sulfonic acid groups. XRD showed the crystallinity retention of Fe3O4 and TiO2 nanoparticles after treatment with an acidic reagent. The FE-SEM results revealed the nanometer-sized particles of the catalyst and VSM demonstrated their magnetic behavior, useful for quick magnetic separation using a conventional magnet. Moreover, according to the acid–base titration, the sulfonic acid group loading was 2.70 mmol g−1. The catalyst exhibited excellent activity in the synthesis of benzimidazoquinazolinones and polyhydroquinolines under solvent-free conditions, which was optimized by CCD and afforded high to excellent yields of products. Furthermore, the catalyst could be easily separated using an external magnet, and be used repeatedly six times, which makes it a potent acidic heterogeneous nanocatalyst for chemical transformations.

Acknowledgements

We gratefully acknowledge the Faculty of chemistry of Semnan University for supporting this work.

References

  1. A. Corma and H. Garcia, Adv. Synth. Catal., 2006, 348, 1391–1412 CrossRef CAS.
  2. F.-C. Zheng, Q.-W. Chen, L. Hu, N. Yan and X.-K. Kong, Dalton Trans., 2014, 1220–1227 RSC.
  3. R. H. Vekariya, K. D. Patel and H. D. Patel, RSC Adv., 2015, 5, 90819–90837 RSC.
  4. X. Zhang, Y. Zhao, S. Xu, Y. Yang, J. Liu, Y. Wei and Q. Yang, Nat. Commun., 2014, 5 DOI:10.1038/ncomms4170.
  5. M. A. Zolfigol, Tetrahedron, 2001, 57, 9509–9511 CrossRef CAS.
  6. J. M. Thomas and W. J. Thomas, Principles and practice of heterogeneous catalysis, John Wiley & Sons, 2014 Search PubMed.
  7. Y. Zhang and S. N. Riduan, Chem. Soc. Rev., 2012, 41, 2083–2094 RSC.
  8. P. Puthiaraj and K. Pitchumani, Green Chem., 2014, 16, 4223–4233 RSC.
  9. M. M. Heravi, E. Hashemi, Y. S. Beheshtiha, K. Kamjou, M. Toolabi and N. Hosseintash, J. Mol. Catal. A: Chem., 2014, 392, 173–180 CrossRef CAS.
  10. S. Kawamorita, R. Murakami, T. Iwai and M. Sawamura, J. Am. Chem. Soc., 2013, 135, 2947–2950 CrossRef CAS PubMed.
  11. H. Naeimi and D. Aghaseyedkarimi, Dalton Trans., 2016, 1243–1253 RSC.
  12. K. Swapna, S. N. Murthy, M. T. Jyothi and Y. V. D. Nageswar, Org. Biomol. Chem., 2011, 9, 5978–5988 CAS.
  13. E. Tabrizian, A. Amoozadeh and S. Rahmani, RSC Adv., 2016, 6, 21854–21864 RSC.
  14. M. Hajjami and B. Tahmasbi, RSC Adv., 2015, 5, 59194–59203 RSC.
  15. E. Kolvari, N. Koukabi and M. M. Hosseini, J. Mol. Catal. A: Chem., 2015, 397, 68–75 CrossRef CAS.
  16. H. Naeimi and Z. S. Nazifi, Appl. Catal., A, 2014, 477, 132–140 CrossRef CAS.
  17. G. Ngnie, G. K. Dedzo and C. Detellier, Dalton Trans., 2016, 9065–9072 RSC.
  18. A. Ghorbani-Choghamarani and B. Tahmasbi, New J. Chem., 2016, 40, 1205–1212 RSC.
  19. E. Lucas, S. Decker, A. Khaleel, A. Seitz, S. Fultz, A. Ponce, W. Li, C. Carnes and K. J. Klabunde, Chem.–Eur. J., 2001, 7, 2505–2510 CrossRef CAS.
  20. J. Che, Y. Xiao, X. Wang, A. Pan, W. Yuan and X. Wu, Surf. Coat. Technol., 2007, 201, 4578–4584 CrossRef CAS.
  21. R. Ghahremanzadeh, Z. Rashid, A.-H. Zarnani and H. Naeimi, Dalton Trans., 2014, 15791–15797 RSC.
  22. N. T. S. Phan and C. W. Jones, J. Mol. Catal. A: Chem., 2006, 253, 123–131 CrossRef CAS.
  23. S. Shylesh, V. Schünemann and W. R. Thiel, Angew. Chem., Int. Ed., 2010, 49, 3428–3459 CrossRef CAS PubMed.
  24. Y. Zhu, L. P. Stubbs, F. Ho, R. Liu, C. P. Ship, J. A. Maguire and N. S. Hosmane, ChemCatChem, 2010, 2, 365–374 CrossRef CAS.
  25. M. B. Gawande, P. S. Branco and R. S. Varma, Chem. Soc. Rev., 2013, 42, 3371–3393 RSC.
  26. F. Nemati, A. Elhampour, H. Farrokhi and M. B. Natanzi, Catal. Commun., 2015, 66, 15–20 CrossRef CAS.
  27. E. Tabrizian, A. Amoozadeh and T. Shamsi, React. Kinet., Mech. Catal., 2016, 119, 245–258 CrossRef CAS.
  28. A. Amoozadeh and S. Rahmani, J. Mol. Catal. A: Chem., 2015, 306, 96–107 CrossRef.
  29. S. Rahmani, A. Amoozadeh and E. Kolvari, Catal. Commun., 2014, 56, 184–188 CrossRef CAS.
  30. A. Amoozadeh, S. Golian and S. Rahmani, RSC Adv., 2015, 5, 45974–45982 RSC.
  31. E. Tabrizian and A. Amoozadeh, Catal. Sci. Technol., 2016, 6, 6267–6276 CAS.
  32. X. Liu, Z. Ma, J. Xing and H. Liu, J. Magn. Magn. Mater., 2004, 270, 1–6 CrossRef CAS.
  33. H. Xia and M. Song, J. Mater. Chem., 2006, 16, 1843–1851 RSC.
  34. Y. S. Chun, K. Ha, Y.-J. Lee, J. S. Lee, H. S. Kim, Y. S. Park and K. B. Yoon, Chem. Commun., 2002, 1846–1847,  10.1039/B205046C.
  35. S. Li and Y. Liu, Polyurethane Adhesive, Chemical Industry Press, Beijing, China, 1998, pp. 13–20 Search PubMed.
  36. J. Chen, M. Liu, L. Zhang, J. Zhang and L. Jin, Water Res., 2003, 37, 3815–3820 CrossRef CAS PubMed.
  37. B. Ou, D. Li, Q. Liu, Z. Zhou and B. Liao, Mater. Chem. Phys., 2012, 135, 1104–1107 CrossRef CAS.
  38. P. Maloo, T. K. Roy, D. M. Sawant, R. T. Pardasani and M. M. Salunkhe, RSC Adv., 2016, 6, 41897–41906 RSC.
  39. Q. Shao, S. Tu, C. Li, L. Cao, D. Zhou, B. Jiang, Y. Zhang, W. Hao and Q. Wang, J. Heterocycl. Chem., 2008, 45, 411–416 CrossRef CAS.
  40. M. M. Heravi, L. Ranjbar, F. Derikvand, B. Alimadadi, H. A. Oskooie and F. F. Bamoharram, Mol. Diversity, 2008, 12, 181–185 CrossRef CAS PubMed.
  41. G. M. Ziarani, A. Badiei, Z. Aslani and N. Lashgari, Arabian J. Chem., 2015, 8, 54–61 CrossRef.
  42. F. Y. a. C. Y. L. Wu, Bull. Chem. Soc. Ethiop., 2010, 14, 417–423 Search PubMed.
  43. K. N. S. Otokesh, E. Kolvari, A. Amoozadeh, M. Malmir and S. Azhari, S. Afr. J. Chem., 2015, 68, 15–20 CrossRef.
  44. A. Amoozadeh, S. Rahmani, M. Bitaraf, F. B. Abadi and E. Tabrizian, New J. Chem., 2016, 40, 770–780 RSC.
  45. S. Ko, M. N. V. Sastry, C. i. Lin and C.-F. Yao, Tetrahedron Lett., 2005, 46, 5771–5774 CrossRef CAS.
  46. R. G. Puligoundla, S. Karnakanti, R. Bantu, K. Nagaiah, S. B. Kondra and L. Nagarapu, Tetrahedron Lett., 2013, 54, 2480–2483 CrossRef CAS.
  47. M. R. Mousavi and M. T. Maghsoodlou, Monatsh. Chem., 2014, 145, 1967–1973 CrossRef CAS.
  48. C. Yao, S. Lei, C. Wang, T. Li, C. Yu, X. Wang and S. Tu, J. Heterocycl. Chem., 2010, 47, 26–32 CAS.
  49. A. Maleki and N. Ghamari, A three-component one-pot procedure for the synthesis benzimidazolo-quinazolinone derivatives in the presence of chitosan-supported metal nanocomposite as a green and reusable catalyst, The 17th International Electronic Conference on Synthetic Organic Chemistry (ECSOC-17), Spain, 2013 Search PubMed.
  50. M. R. Naimi-Jamal and M. Karimi, The 19th International Electronic Conference on Synthetic Organic Chemistry (ECSOC-19), 1–30 November 2015 Search PubMed.

Footnote

Electronic supplementary information (ESI) available: FT-IR, 1H NMR and 13C NMR spectra. See DOI: 10.1039/c6ra21048a

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.