Abstract

The present study aims at investigating the performance of Friedel–Crafts acylation of phenols and acyl chlorides over the PTA@MIL-53 (Fe) catalyst (phosphotungstic acid encapsulated in MIL-53 (Fe) via the one-pot method) under ultrasound irradiation. PTA@MIL-53 (Fe) was synthesized using ultrasound irradiation at ambient temperature and pressure. Moreover, X-ray diffraction, energy-dispersive X-ray, Fourier transform infrared, scanning electron microscopy, N2 physisorption, and inductively coupled plasma optical emission spectrometry characterization analyses were performed to characterize the prepared composite. The reactions took place under an ultrasonic environment at ambient temperature and pressure. The obtained results reveal that the synthesized catalyst is efficient in an ultrasound irradiation environment for Friedel–Crafts reactions. The regeneration experiments indicated no significant change in the catalyst performance even after being reused for five times. Furthermore, the response surface methodology was applied to find the relationships between some specific variables and para-hydroxy acetophenone yield as a response. Finally, the results were validated with the experimental data, with which the model results were in good agreement.
Keywords: metal−organic framework, Friedel−Crafts, MIL-53, phosphotungstanic acid, RSM
1. Introduction
The Friedel–Crafts of aromatics serve as valuable synthesis steps in manufacturing fine chemicals and pharmaceuticals. The Friedel–Crafts acylation of phenols and acyl chlorides yields ortho- and para-hydroxy acetophenones (2-HAP and 4-HAP), which are very valuable precursors in the pharmaceutical industry.1−3 Lewis acids (AlCl3, BF3) or mineral acids (HF or H2SO4) as conventional homogeneous catalysts, which have been widely used in such reactions, have serious drawbacks.2,4 In homogeneous catalyst systems, the purification of products and separation and recovery of the catalyst at the end of the reaction are very difficult, time-consuming, and costly.5−8 It is worth mentioning that the replacement of these conventional catalysts by heterogeneous reusable acid catalysts is promising.9 The heterogeneous catalysts have advantages such as easy separation of the catalyst from the reaction mixture, reusability of the catalyst, environmental protection, and economic efficiency.10−13 Zeolite,14,15 sulfonic acid-modified mesostructured SBA-152 materials, and other acidic heterogeneous catalysts such as the sulfonic resin Nafion16 have been studied as catalytic systems in the Friedel–Crafts acylation. In this sense, rapid deactivation of zeolites and limited specific surface areas for sulfonic resin Nafion and SBA-15 encourage researchers to find more applicable catalysts. Therefore, finding a catalyst having the following characteristics, that is, non-toxicity, stability in solvents, easy synthesis, high activity, easy separation from the reaction mixture, and reusability, is still a challenge.
Heteropoly acids are complex Bronsted acids consisting of heteropoly anions, which possess metal–oxygen octahedra as the basic structural units.17−20 Heteropoly acid catalysts such as the Silica-coated magnetic NiFe2O4 nanoparticle-supported phosphomolybdic acid are reported as non-toxic solid acids with high acidity and, therefore, significant catalytic activity, are a very good alternative to liquid mineral acids in industry.21−24 Due to the very high solubility in polar solvents, these materials are mainly used heterogeneously (deposited on the porous solid matrix).25,26 Phosphotungstic acid (PTA), as a heteropoly acid (HPA), has been regarded as a novel efficient reusable catalyst for Fries rearrangement of phenyl acetate in homogeneous or heterogeneous liquid-phase systems, the synthesis of amidoalkyl naphthols, which are generally synthesized via the three-component reaction of β-naphthol with an aldehyde and an amide in the presence of various promoting agents, and the synthesis of N,N′-alkylidene bisamidesa using M-Fe3O4@HAL-SO3H.27,28 Moreover, metal organic frameworks (MOFs) are a class of organic–inorganic structural porous nanomaterials whose crystal structure consists of metal groups or metal oxides coordinated into organic ligands to form regular lattice frameworks.29,30 They have tunable open-metal centers and are able to hold active guest species in their pores.31,32 MIL-53 is a typical class of MOFs, which is built up from Fe(III) cations with 1,4-dicarboxylate as the linker molecule.27 In particular, the presence of iron has many advantages over other metals due to its low cost and non-toxicity, making it interesting as a catalyst. Regarding the one-pot method, PTA molecules are encapsulated in MIL-53 pores and grow bigger in size than the windows of the cavities of MIL-53 (Fe), according to which PTA leaching becomes less probable.33 In conclusion, the PTA encapsulation in the MOF structure is a good approach in immobilizing PTA.34 There are several studies which have used PTA@MOF for catalytic purpose.8,35,36 Friedel–Crafts acylation processes have been traditionally performed using heat, which is a very time-consuming process and needs a large amount of energy at high cost. Ultrasound is a new green, rapid, economic, environment-friendly, and safe technique that can be employed for the acceleration of chemical processes.37 This method can also be employed for the synthesis of porous materials such as MOFs.38−41 Herein and in connection with an ongoing synthesis program of nanoparticles,42−46 PTA was encapsulated in MIL-53 (Fe) with the assistance of ultrasound waves. Afterward, the synthesized material was employed as an efficient heterogeneous catalyst for Friedel–Crafts acylation of phenols and acyl chlorides under ultrasound irradiation at room temperature.
2. Materials and Methods
2.1. Materials
Hexahydrate iron(III)-chloride (FeCl3·6H2O), terephthalic acid (1,4-BDC), dimethylformamide (DMF), PTA, and methanol were purchased from Merck Millipore company. All materials were used without further purification.
2.2. Synthesis
PTA@MIL-53 (Fe) was prepared using the one-pot ultrasound irradiation process based on the reported research.8 In a typical procedure, 1 mmol of FeCl3·6H2O, 1 mmol of 1,4-BDC, and various amounts of PTA were dissolved in 5 mL of dimethylformamide (DMF). Subsequently, the mixture was transferred into the ultrasonic bath (37 kHz frequency and 240 W output power) in an ambient temperature for 15 min. Afterward, the orange product was filtered, washed with DMF, and then, dried in vacuum overnight at 160 °C. Regarding the MIL-53 (Fe) synthesis, the same procedure was used, but without PTA addition.
2.3. Characterization Analysis
The powder X-ray diffraction patterns (XRD) were recorded using a PANalytica, Netherlands with monochromated Cu Kα radiation (λ = 1.54056 Å) at room temperature. Fourier transform infrared spectroscopy (FT-IR) spectra were obtained using a Vertex 70 at room temperature. Thermogravimetric analysis (TGA) was conducted applying an NETZSCH STA 449F3 analyzer. Nitrogen adsorption isotherms were conducted applying a Belsorp-mini II. Moreover, an inductively coupled plasma optical emission spectrometry (ICP-OES) was done using a Varian 730-ES spectrometer.
2.4. Catalytic Reaction Studies
The Friedel–Crafts reaction was performed in a 25 mL round-bottom flask containing 3 mL of acetonitrile that was placed in an ultrasonic bath. Typically, 20–200 mg of PTA@MIL-53 (Fe) as the catalyst containing different amounts of PTA (5–50%), 6 mmol phenol, and 18 mmol acyl chloride was exposed to ultrasound radiation for 5–45 min at ambient temperature. After a predetermined time of the experiment had passed, a sample of suspension was taken, and then, the catalyst was separated by centrifugation. Finally, the product was analyzed applying a GC–MS (7890B Agilent instrument using Rtx@5MS with 30 m length and 0.25 mm inner diameter).
3. Results and Discussion
3.1. Characterization
Herein, we aimed at characterizing PTA (12.09 wt %) @MIL-53 (Fe) as the sample. It is worth mentioning that PTA@MIL-53 (Fe) is used instead of PTA (12.09 wt %) @MIL-53 (Fe) throughout the rest of the text. Moreover, the amount of PTA was evaluated using the ICP-OES technique.
The X-ray diffraction patterns of PTA, MIL-53 (Fe), and PTA@MIL-53 (Fe) are illustrated in Figure 1. Regarding the MIL-53 (Fe) pattern, main peaks appear at 2θ values of 9, 12.45, 17.52, 18.45, and 25.68, which are in good agreement with the reported patterns in the previous literature,47 confirming that MIL-53 (Fe) was successfully synthesized. The similar patterns for PTA@MIL-53 (Fe) show that the structure of MIL-53 (Fe) is not destroyed during the encapsulation process and the structural properties of the framework are fully preserved. Also, the XRD pattern of PTA@MIL-53 (Fe) shows new peaks corresponding to the registered standard (JCPDF; 96-406-8822) that are related to the addition of W and P to the structure. The comparison of XRD patterns of the PTA@MIL-53 (Fe) and MIL-53 (Fe) pattern shows some redistribution, which can be attributed to the interaction of the PTA clusters with the MIL-53 (Fe) frameworks and/or changes in the symmetry of PTA clusters in the cages of MOF.48
Figure 1.

XRD patterns of PTA, PTA@MIL-53 (Fe), and MIL-53 (Fe).
The FT-IR spectra of the PTA, MIL-53 (Fe), and PTA@MIL-53 (Fe) are shown in Figure 2. FT-IR spectra of the MIL-53 (Fe) illustrated the identical peaks as those reported in the literature.43 By comparing the spectra of these three materials, it can be recognized that the PTA@MIL-53 (Fe) has similar spectrum features to that of the MIL-53 (Fe), while the peaks corresponding to asymmetric υ (P–O), υ̅ (W–Ot), υ̅ (W–Oc–W), and υ̅ (W–Oe–W) in the PTA structure at 1018, 972, 883, and 815 cm–1, respectively, have affected the PTA@MIL-53 (Fe) spectrum,8 where Ot, Oc, and Oe refer to terminal, corner, and edge of the oxygen’s atoms, respectively.49
Figure 2.
FT–IR spectra of PTA, MIL-53 (Fe), and PTA@MIL-53 (Fe).
TGA analysis was applied to study the thermal stability of the synthesized MIL-53 (Fe) and PTA@MIL-53 (Fe). In this sense, the sample weight changes were measured from room temperature to 850 °C in an air atmosphere. As indicated in Figure 3, PTA is almost stable even in 800 °C; but H2BDC starts to degrade in 300 °C. In the first step (<300 °C), the weight loss was due to the evaporation of water and DMF molecules inside the pores of the materials. In the second step, the degradation of H2BDC (terephthalic acid) occurs for MIL-53 (Fe) and for PTA@MIL-53 (Fe), and finally, complete demolition of MOF happens and only iron oxide and PTA remain. It can be observed that the weight loss of bare MOF is higher relative to the PTA-supported MOF because the cages of the composite were occupied by PTA molecules.
Figure 3.
TGA profile of PTA, PTA@MIL-53 (Fe), MIL-53 (Fe), and H2BDC.
The surface morphologies of the PTA@MIL-53 (Fe) composite are shown in Figure 4. Clearly, after encapsulating PTA into MIL-53 (Fe), the surface of composite MOF becomes uneven, consisting of small sphere-like particles (Figure 4), but retains the original structure.
Figure 4.
FE-SEM images (a) with a scale bar of 10 μm, (b) with a scale bar of 1 μm, and (c,d) with a scale bar of 200 nm of PTA@MIL-53 (Fe).
As displayed in Figure 5a,b, the N2 adsorption of both MIL-53 (Fe) and PTA@MIL-53 (Fe) represents type (II) adsorption isotherms. Based on the results of Brunauer Emmett–Teller (BET), the total pore volume (Vtotal) and mean pore diameter (Dmean) decreased from 0.089 to 0.057 cm3 g–1 and 23.028 to 10.102 nm, respectively. The obtained results from Table 1 indicate that the BET surface area decreases from 293.26 to 40.70 m2 g–1 for MIL-53 (Fe) and PMA (50 wt %) @ MIL-53 (Fe) due to the confinement of PTA inside the porous of MIL-53 (Fe).
Figure 5.
N2 adsorption–desorption isotherms of MIL-53 (Fe) (a) and PTA@MIL-53 (Fe) (b).
Table 1. Characteristics of PTA@MIL-53 (Fe) Matrix Catalysts.
| catalyst | aW (ppm) | bactual loading of PTA (%) | cSBET (m2/g) |
|---|---|---|---|
| MIL-53 (Fe) | 293.26 | ||
| PTA (5 wt %)@MIL-53 (Fe) | 37303 | 4.87 | 210.37 |
| PTA (14.12 wt %)@MIL-53 (Fe) | 104716 | 13.67 | 167.23 |
| PTA (27.5 wt %)@MIL-53 (Fe) | 192804 | 25.17 | 116.09 |
| PTA (40.88 wt %)@MIL-53 (Fe) | 69019 | 35.12 | 83.90 |
| PTA (50 wt %)@MIL-53 (Fe) | 317277 | 41.42 | 40.70 |
5 mg of PTA@MIL-53 (Fe) was dissolved in 2 mL of 65% aqueous HNO3. The sample was diluted to a 10 mL volumetric flask with deionized water and then performed by the ICP-OES instrument.
In order to investigate the wt % of PTA in the matrix, the amount of W, which was performed by ICP-OES, was divided to 76.6 multiply by 100 (W content of H3PW12O40 = 76.6 wt %).
Specific surface area.
The compositional analysis of the synthesized (a) MIL-53 (Fe) and (b) PTA@MIL-53 (Fe) composites carried out using energy-dispersive X-ray (EDX) analysis (Figure 6). The map spectrum indicates the presence of C (Kα = 0.277 keV), O (Kα = 0.525 keV), and Fe (Lα = 0.705, Kα = 6.398 and Kβ = 7.07 keV) elements and also shows the expected elemental composition of MIL-53 (Fe). Additionally, the EDX measurements on PMA@MIL-53 (Fe) samples demonstrate that all of the X-ray signals of its parent MOF material and the position of a sharp peaks of W (Lα = 8.396 keV, M = 1.774 keV) and P (Kα = 2.013 keV) elements can be ascribed to PTA molecules. These results indicate the expected element composition of C, O, P, Fe, and W with the corresponding mass percentage of 40.46, 27.56, 0.07, 12.80, and 17.16 wt %, respectively, which further demonstrate that the PTA molecules are protected in the MIL-53 (Fe) solid support.
Figure 6.
EDS analysis of (a) MIL-53 (Fe) and (b) PTA@MIL-53 (Fe).
To validate this encapsulation, the hybrid samples were characterized using ICP-OES. The average of PTA hybrids in all PTA@MIL-53 (Fe) samples, which is depicted in Table 1, accords with their theoretically calculated data in Table 1.
3.2. Optimization
Design expert provides powerful tools that can facilitate the research process and also reduce time and cost. The independent variables for this study are reaction time (min), catalyst loading (mg), and PTA (wt %), and the response is the yield of product. Regarding the results of design expert, a total of 17 experiments were designed, which can be seen in Table 2.
Table 2. Experimental Data of Yield (%) of Friedel–Crafts Reaction are Used in Design Expert Analysis.
| independent
variables |
response | |||
|---|---|---|---|---|
| run | reaction time (min) | catalyst loading (mg) | PTA (wt %) | yield (%) |
| 1 | 25 | 110 | 27.50 | 80.8a |
| 2 | 25 | 20.00 | 27.50 | 28.0a |
| 3 | 36.89 | 56.49 | 40.88 | 66.0a |
| 4 | 13.11 | 163.51 | 14.12 | 28.0a |
| 5 | 13.11 | 56.49 | 40.88 | 14.5a |
| 6 | 13.11 | 56.49 | 14.12 | 8.0a |
| 7 | 25.00 | 200 | 27.50 | 72.1a |
| 8 | 36.89 | 56.49 | 14.12 | 33.0a |
| 9 | 45.00 | 110 | 27.50 | 80.8a |
| 10 | 36.89 | 163.51 | 40.88 | 95.3a |
| 11 | 25.00 | 110 | 27.50 | 78.0a |
| 12 | 25.00 | 110 | 50.00 | 80.0a |
| 13 | 25.00 | 110 | 27.50 | 75.0a |
| 14 | 5.00 | 110 | 27.50 | 25.0a |
| 15 | 13.11 | 163.51 | 40.88 | 58.0a |
| 16 | 36.89 | 163.51 | 14.12 | 17.0a |
| 17 | 25.00 | 110 | 5.00 | 14.0a |
| 18 | 160.00 | 163.51 | 40.88 | 25.1b |
Reaction was stirred at ultrasound irradiation.
Reaction was stirred at a refluxing temperature without sonication (conventional thermal catalysis condition).
In order to extract the mathematical model, assessment of the effect of variables, and their interactions on the response, the response surface methodology (RSM) was used. A central composite design with three independent variables was applied to obtain the effect of variables on the yield of para-hydroxy acetophenones. The proposed equation used in predicting the yield is as follows
where A is the reaction time (min), B is the catalyst loading (mg), and C is PTA wt %. The validity of the proposed models must be assessed with the experimental data. The R-squared (R2) and the adjusted R-squared (adj R2) of the proposed model are shown in Table 3. According to the reported values, the model results are in good agreement with the experimental data.
Table 3. Values of R2 and adj R2 for the Proposed Model.
| model | R2 | adj R2 |
|---|---|---|
| yield (%) in Friedel–Crafts reaction | 0.9717 | 0.9353 |
Moreover, the analysis of variance (ANOVA) was carried out to demonstrate the significance of the main and interacting effects of parameters on the product yield. Analysis of variance for the quadratic equation for the yield of Friedel–Crafts reaction is tabulated in Table 4. The significance of the coefficient term is proved by the values of F and p.47
Table 4. ANOVA for the Regression Model and Respective Model Terms for the Product Yield of Friedel–Crafts Reactiona,b,c.
| source | sum of squares | Df | mean square | F-value | p-value(prob > F) | |
|---|---|---|---|---|---|---|
| model | 13969.26 | 9 | 1552.14 | 28.00 | 0.0001 | significant |
| A—reaction time | 2831.46 | 1 | 2831.46 | 51.08 | 0.0002 | |
| B—catalyst loading | 1668.84 | 1 | 1668.84 | 30.10 | 0.0009 | |
| C—PTA | 4904.25 | 1 | 4904.25 | 88.47 | <0.0001 | |
| AB | 315.01 | 1 | 315.01 | 5.68 | 0.0486 | |
| AC | 699.38 | 1 | 699.38 | 12.62 | 0.0093 | |
| BC | 591.68 | 1 | 591.68 | 10.67 | 0.0137 | |
| A2 | 1289.76 | 1 | 1289.76 | 23.27 | 0.0019 | |
| B2 | 1544.21 | 1 | 1544.21 | 27.86 | 0.0012 | |
| C2 | 1841.87 | 1 | 1841.87 | 33.23 | 0.0007 | |
| residual | 388.04 | 7 | 55.43 | |||
| lack of fit | 371.22 | 5 | 74.24 | 8.82 | 0.1049 | not significant |
| pure error | 16.83 | 2 | 8.41 | |||
| cor. total | 14357.31 | 16 |
Degree of freedom.
Test for comparing the model with residual (error) variance.
Probability of finding the observed F value when the null hypothesis is true (significant at p < 0.05 and not significant at p > 0.05).
The p-value of less than 0.0500 represents that the model terms are significant, whereas the values greater than 0.1000 are not significant.48 According to Table 4, the model and all their terms are significant. As we can see in Table 4, the “Lack-of-Fit” is not significant, demonstrating the accuracy of the model.
The following plots were drawn to present the effects of two variables simultaneously on the response, while the other one is constant. In Figure 7a, by increasing the reaction time in a constant catalyst loading, the yield increases until it becomes constant. The reason for this is that, the longer the time, the higher the possibility of interaction of reactant molecules and active sites would be. On the contrary, by increasing the time, more active sites are involved, and therefore, no further change will be observed in the yield. In a constant time, the same trend is seen for the effect of increasing catalyst loading on the yield. This is due to the fact that, in a specific amount of time, increasing the catalyst loading will increase the availability of the active sites for the reactants. However, by increasing it too much, the yield decreases due to the decrease in the total sorption surface area available to adsorbates resulting from the overlapping of active sites. Regarding Figure 7b, we can see two different trends in high and low wt % of PTA. Considering the lower wt % PTA, the increment in time promotes the yield. Afterward, it becomes almost constant, due to no more active sites being available. In higher wt % of PTA, there is a permanent increase in the progress of reaction because of a higher number of active sites being available, although in higher reaction time, the rate of increasing yield decreased due to fewer active sites. Finally, the reaction is performed under the conventional thermal catalysis condition to compare the results of the synthesis of the desired products sonication method (Table 2, entry 19). It could be observed that the presence of ultrasound irradiation is essential for the reaction performance, while performing the reaction without sonication has not led to the product in excellent yield.
Figure 7.
Effects of (a–c) variables on Friedel–Crafts reaction yield.
As illustrated in Figure 7c, in high weight percent of PTA, by increasing the catalyst loading, the yield increases even in high catalyst loading (163.51 mg). The reason is attributed to the amount of available active sites and their overlapping, which was mentioned above in part b and c.
3.3. Mechanism
The plausible mechanism of the phenol Friedel–Crafts Acylation over the catalysis of PTA@MIL-53 (Fe) is shown in Scheme 1. In this process, a resonance-stabilized acylium ion with a positive charge on the carbon is formed as the result of C–Cl bond cleavage. This intermediate acts as an electrophile and reacts with the phenol to yield the corresponding product.
Scheme 1. Plausible Mechanism for the Friedel–Crafts Acylation of Phenol Catalyzed by PTA@MIL-53 (Fe) Solid Acid.
3.4. Reusability
The possibility of catalyst regeneration and the long-term catalyst stability is important for the commercial feasibility of any catalyst process. In this study, the catalyst was recovered through centrifugation, washed once with DMF, twice with deionized water, and twice with methanol, and then, employed in both of the consecutive reactions. According to Figure 8, the yield drops slightly by 8% after five times of recycling. Therefore, the suitable recyclability makes it a promising catalyst for the reactions. In Figure 9, comparing the spectra of the catalyst after three and five times of use with the fresh catalyst showed slight shifting of peak positions and also some changes in peak intensity because of the presence of reactants and the product. It is worth mentioning that the chemical composition of the catalyst does not change, confirming that the catalyst has remained almost unchanged after five times of reuse. The PTA leaching of composite MOF was studied using ICP-OES analysis. The results showed ca. 0.17 wt % PTA leaching for 12.09 wt % PTA loading. Also, the hot-filtration test confirms the heterogeneous nature of the catalyst in reaction medium. This finding along with recycling performance proved the appropriate reusability of the catalyst.
Figure 8.

Catalyst reusability in the Friedel–Crafts reaction.
Figure 9.

FT–IR spectra of fresh PTA@MIL-53 (Fe), three times recycled PTA@MIL-53 (Fe) and five times recycled PTA@MIL-53 (Fe) in Friedel–Crafts.
3.5. Comparison
We have also compared the efficiency of our catalyst with some literature reports. The catalyst is effective against the acylation reaction under milder conditions at shorter reaction time and higher yields with more regioselectivity and reusability (Table 5).
Table 5. Comparison of the Catalytic Activity of PTA@MIL-53 (Fe) in the Acylation of Phenol with Other Reported Catalyst Systems in the Literature.
4. Conclusions
A metal organic framework, MIL-53 (Fe) composite, based on PTA was synthesized through the ultrasound method and was used as a catalyst in Friedel–Crafts reactions. Characterizations of the catalyst were achieved applying XRD, EDX, FT–IR, TGA, FE-SEM, nitrogen adsorption–desorption, and ICP-OES analysis. The batch experiments were performed at different reaction times, catalyst loading, and different amounts of PTA. The maximum yield (95.3) was obtained for para-hydroxy acetophenone in Friedel–Crafts reaction. The regeneration experiments revealed that the yield of the reactions was not changed significantly even after being reused up to seven times, which proved the catalyst stability. The RSM was applied to find the relationships between several explanatory variables and yield of the product as a response.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmaterialsau.2c00062.
Copies of GC chromatogram of standard samples and Friedel–Crafts acylation of phenol reactions (PDF)
Author Contributions
A.N.: project administration, funding acquisition, supervision, and review and editing. H.R.A.: investigation; validation, and formal analysis. N.M.: investigation and writing—original draft preparation. S.M.: investigation, validation, and formal analysis. CRediT: Nasim Mirzaei investigation (equal), writing-original draft (equal); Sara Masoumi formal analysis (equal), investigation (equal), validation (equal).
The authors declare no competing financial interest.
Supplementary Material
References
- Shirini F.; Zolfigol M. A.; Abedini M. Saccharinsulfonic Acid: An Efficient and Recyclable Catalyst for Acetylation of Alcohols, Phenols, and Amines. Monatsh. Chem. 2009, 140, 1495–1498. 10.1007/s00706-009-0214-7. [DOI] [Google Scholar]
- Grieken R. v.; Melero J. A.; Morales G. Fries Rearrangement of Phenyl Acetate over Sulfonic Modified Mesostructured SBA-15 Materials. Appl. Catal., A 2005, 289, 143–152. 10.1016/j.apcata.2005.04.059. [DOI] [Google Scholar]
- Harvey G.; Mäder G. The Shape-Selective Acylation of 2-Methoxynaphthalene, Catalyzed by Zeolites Y, Beta and ZSM-12. Collect. Czech. Chem. Commun. 1992, 57, 862–868. 10.1135/cccc19920862. [DOI] [Google Scholar]
- Gore P. H. Friedel-Crafts and Related Reactions. J. Franklin Inst. 1965, 280, 459. 10.1016/0016-0032(65)90548-x. [DOI] [Google Scholar]
- Kazemi M.; Mohammadi M. Magnetically Recoverable Catalysts: Catalysis in Synthesis of Polyhydroquinolines. Appl. Organomet. Chem. 2020, 34, e5400 10.1002/aoc.5400. [DOI] [Google Scholar]
- Koolivand M.; Nikoorazm M.; Ghorbani-Choghamarani A.; Mohammadi M. A Novel Cubic Zn-citric Acid-based MOF as a Highly Efficient and Reusable Catalyst for the Synthesis of Pyranopyrazoles and 5-substituted 1H-tetrazoles. Appl. Organomet. Chem. 2022, 36, 2641–2663. 10.1002/aoc.6656. [DOI] [Google Scholar]
- Mohammadi M.; Ghorbani-Choghamarani A. L-Methionine-Pd Complex Supported on Hercynite as a Highly Efficient and Reusable Nanocatalyst for C-C Cross-Coupling Reactions. New J. Chem. 2020, 44, 2919–2929. 10.1039/c9nj05325e. [DOI] [Google Scholar]
- Nikseresht A.; Daniyali A.; Ali-Mohammadi M.; Afzalinia A.; Mirzaie A. Ultrasound-Assisted Biodiesel Production by a Novel Composite of Fe(III)-Based MOF and Phosphotangestic Acid as Efficient and Reusable Catalyst. Ultrason. Sonochem. 2017, 37, 203–207. 10.1016/j.ultsonch.2017.01.011. [DOI] [PubMed] [Google Scholar]
- Mohammadi M.; Khodamorady M.; Tahmasbi B.; Bahrami K.; Ghorbani-Choghamarani A. Boehmite Nanoparticles as Versatile Support for Organic–Inorganic Hybrid Materials: Synthesis, Functionalization, and Applications in Eco-Friendly Catalysis. J. Ind. Eng. Chem. 2021, 97, 1–78. 10.1016/j.jiec.2021.02.001. [DOI] [Google Scholar]
- Esmaili S.; Khazaei A.; Ghorbani-Choghamarani A.; Mohammadi M. Silica Sulfuric Acid Coated on SnFe2O4 MNPs: Synthesis, Characterization and Catalytic Applications in the Synthesis of Polyhydroquinolines. RSC Adv. 2022, 12, 14397–14410. 10.1039/d2ra01202b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohammadi M.; Ghorbani-Choghamarani A. Synthesis and characterization of novel hercynite@sulfuric acid and its catalytic applications in the synthesis of polyhydroquinolines and 2,3-dihydroquinazolin-4(1H)-ones. RSC Adv. 2022, 12, 2770–2787. 10.1039/d1ra07381h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rahimzadeh G.; Tajbakhsh M.; Daraie M.; Mohammadi M. Dysprosium–Balsalazide Complex Trapped between the Functionalized Halloysite and g-C3N4: A Novel Heterogeneous Catalyst for the Synthesis of Annulated Chromenes in Water. Appl. Organomet. Chem. 2022, 36, e6829 10.1002/aoc.6829. [DOI] [Google Scholar]
- Tamoradi T.; Mousavi S. M.; Mohammadi M. Praseodymium(Iii) Anchored on CoFe2O4 MNPs: An Efficient Heterogeneous Magnetic Nanocatalyst for One-Pot, Multi-Component Domino Synthesis of Polyhydroquinoline and 2,3-Dihydroquinazolin-4(1: H)-One Derivatives. New J. Chem. 2020, 44, 3012–3020. 10.1039/c9nj05468e. [DOI] [Google Scholar]
- Heidekum A.; Harmer M. A.; Hoelderich W. F. Highly Selective Fries Rearrangement over Zeolites and Nafion in Silica Composite Catalysts: A Comparison. J. Catal. 1998, 176, 260–263. 10.1006/jcat.1998.2034. [DOI] [Google Scholar]
- Coq B.; Gourves V.; Figuéras F. Benzylation of Toluene by Benzyl Chloride over Protonic Zeolites. Appl. Catal., A 1993, 100, 69–75. 10.1016/0926-860X(93)80116-8. [DOI] [Google Scholar]
- Deng T.; Li Y.; Zhao G.; Zhang Z.; Liu Y.; Lu Y. Catalytic Distillation for Ethyl Acetate Synthesis Using Microfibrous-Structured Nafion–SiO 2 /SS-Fiber Solid Acid Packings. React. Chem. Eng. 2016, 1, 409–417. 10.1039/C6RE00088F. [DOI] [Google Scholar]
- Heravi M.; Faghihi Z. Applications of Heteropoly Acids in Multi-Component Reactions. J. Iran. Chem. Soc. 2014, 11, 209–224. 10.1007/s13738-013-0291-8. [DOI] [Google Scholar]
- Heravi M. M.; Sadjadi S. Recent Developments in Use of Heteropolyacids, Their Salts and Polyoxometalates in Organic Synthesis. J. Iran. Chem. Soc. 2009, 6, 1–54. 10.1007/BF03246501. [DOI] [Google Scholar]
- Sánchez-Velandia J. E.; Baldoví H. G.; Sidorenko A. Y.; Becerra J. A.; Martínez O F. Synthesis of Heterocycles Compounds from Condensation of Limonene with Aldehydes Using Heteropolyacids Supported on Metal Oxides. Mol. Catal. 2022, 528, 112511. 10.1016/j.mcat.2022.112511. [DOI] [Google Scholar]
- Jamshidi A.; Maleki B.; Zonoz F. M.; Tayebee R. HPA-Dendrimer Functionalized Magnetic Nanoparticles (Fe3O4@D-NH2-HPA) as a Novel Inorganic-Organic Hybrid and Recyclable Catalyst for the One-Pot Synthesis of Highly Substituted Pyran Derivatives. Mater. Chem. Phys. 2018, 209, 46–59. 10.1016/j.matchemphys.2018.01.070. [DOI] [Google Scholar]
- Sun Z.; Duan X. X.; Gnanasekarc P.; Yan N.; Shi J. Review: Cascade Reactions for Conversion of Carbohydrates Using Heteropolyacids as the Solid Catalysts. Biomass Convers. Biorefin. 2022, 12, 2313–2331. 10.1007/s13399-020-00802-1. [DOI] [Google Scholar]
- Ono Y.; Baba T. Unique Properties of Silver Cations in Solid-Acid Catalysis by Zeolites and Heteropolyacids. Phys. Chem. Chem. Phys. 2015, 17, 15637–15654. 10.1039/c5cp01839k. [DOI] [PubMed] [Google Scholar]
- Maleki B.; Eshghi H.; Khojastehnezhad A.; Tayebee R.; Ashrafi S. S.; Kahoo G. E.; Moeinpour F. Silica Coated Magnetic NiFe2O4 Nanoparticle Supported Phosphomolybdic Acid; Synthesis, Preparation and Its Application as a Heterogeneous and Recyclable Catalyst for the One-Pot Synthesis of Tri- and Tetra-Substituted Imidazoles under Solvent Free Conditi. RSC Adv. 2015, 5, 64850–64857. 10.1039/C5RA10534J. [DOI] [Google Scholar]
- Maleki B.; Baghayeri M. Synthesis of Symmetrical N,N′-Alkylidene Bis-Amides Catalyzed by Silica Coated Magnetic NiFe2O4 Nanoparticle Supported Polyphosphoric Acid (NiFe2O4@SiO2-PPA) and Its Application toward Silver Nanoparticle Synthesis for Electrochemical Detection of Glucose. RSC Adv. 2015, 5, 79746–79758. 10.1039/C5RA16481H. [DOI] [Google Scholar]
- Escobar A. M.; Blustein G.; Luque R.; Romanelli G. P. Recent Applications of Heteropolyacids and Related Compounds in Heterocycle Synthesis. Contributions between 2010 and 2020. Catalysts 2021, 11, 291. 10.3390/catal11020291. [DOI] [Google Scholar]
- Maleki B.; Sheikh E.; Seresht E. R.; Eshghi H.; Ashrafi S. S.; Khojastehnezhad A.; Veisi H. One-Pot Synthesis of 1-Amidoalkyl-2-Naphthols Catalyzed by Polyphosphoric Acid Supported on Silica-Coated NiFe 2 O 4 Nanoparticles. Org. Prep. Proced. Int. 2016, 48, 37–44. 10.1080/00304948.2016.1127098. [DOI] [Google Scholar]
- Kozhevnikova E. Fries Rearrangement of Aryl Esters Catalysed by Heteropoly Acid: Catalyst Regeneration and Reuse. Appl. Catal., A 2004, 260, 25–34. 10.1016/j.apcata.2003.10.008. [DOI] [Google Scholar]
- Maleki B.; Eshghi H.; Barghamadi M.; Nasiri N.; Khojastehnezhad A.; Sedigh Ashrafi S.; Pourshiani O. Silica-Coated Magnetic NiFe2O4 Nanoparticles-Supported H3PW12O40; Synthesis, Preparation, and Application as an Efficient, Magnetic, Green Catalyst for One-Pot Synthesis of Tetrahydrobenzo[b]Pyran and Pyrano[2,3-c]Pyrazole Derivatives. Res. Chem. Intermed. 2016, 42, 3071–3093. 10.1007/s11164-015-2198-8. [DOI] [Google Scholar]
- Koolivand M.; Nikoorazm M.; Ghorbani-Choghamarani A.; Mohammadi M. A Novel Cubic Zn-Citric Acid-Based MOF as a Highly Efficient and Reusable Catalyst for the Synthesis of Pyranopyrazoles and 5-Substituted 1H-Tetrazoles. Appl. Organomet. Chem. 2022, 36, e6656 10.1002/aoc.6656. [DOI] [Google Scholar]
- Ghobakhloo F.; Azarifar D.; Mohammadi M.; Keypour H.; Zeynali H. Copper(II) Schiff-Base Complex Modified UiO-66-NH2(Zr) Metal-Organic Framework Catalysts for Knoevenagel Condensation-Michael Addition-Cyclization Reactions. Inorg. Chem. 2022, 61, 4825–4841. 10.1021/acs.inorgchem.1c03284. [DOI] [PubMed] [Google Scholar]
- Ramish S. M.; Ghorbani-Choghamarani A.; Mohammadi M. Microporous Hierarchically Zn-MOF as an Efficient Catalyst for the Hantzsch Synthesis of Polyhydroquinolines. Sci. Rep. 2022, 12, 1479. 10.1038/s41598-022-05411-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghorbani-Choghamarani A.; Taherinia Z.; Mohammadi M. Facile Synthesis of Fe3O4@GlcA@Ni-MOF Composites as Environmentally Green Catalyst in Organic Reactions. Environ. Technol. Innovation 2021, 24, 102050. 10.1016/j.eti.2021.102050. [DOI] [Google Scholar]
- Masoumi S.; Farshchi Tabrizi F.; Sardarian A. R. Efficient Tetracycline Hydrochloride Removal by Encapsulated Phosphotungstic Acid (PTA) in MIL–53 (Fe): Optimizing the Content of PTA and Recycling Study. J. Environ. Chem. Eng. 2020, 8, 103601. 10.1016/j.jece.2019.103601. [DOI] [Google Scholar]
- Sartori G.; Maggi R. Update 1 of: Use of Solid Catalysts in Friedel–Crafts Acylation Reactions. Chem. Rev. 2011, 111, PR181–PR214. 10.1021/cr100375z. [DOI] [PubMed] [Google Scholar]
- Afzalinia A.; Mirzaie A.; Nikseresht A.; Musabeygi T. Ultrasound-Assisted Oxidative Desulfurization Process of Liquid Fuel by Phosphotungstic Acid Encapsulated in a Interpenetrating Amine-Functionalized Zn(II)-Based MOF as Catalyst. Ultrason. Sonochem. 2017, 34, 713–720. 10.1016/j.ultsonch.2016.07.006. [DOI] [PubMed] [Google Scholar]
- Juan-Alcañiz J.; Ramos-Fernandez E. V.; Lafont U.; Gascon J.; Kapteijn F. Building MOF Bottles around Phosphotungstic Acid Ships: One-Pot Synthesis of Bi-Functional Polyoxometalate-MIL-101 Catalysts. J. Catal. 2010, 269, 229–241. 10.1016/j.jcat.2009.11.011. [DOI] [Google Scholar]
- Nikseresht A.; Ghasemi S.; Parak S. [Cu3(BTC)2]: A Metal–Organic Framework as an Environment-Friendly and Economically Catalyst for the Synthesis of Tacrine Analogues by Friedländer Reaction under Conventional and Ultrasound Irradiation. Polyhedron 2018, 151, 112–117. 10.1016/j.poly.2018.05.018. [DOI] [Google Scholar]
- Bagheri S.; Pazoki F.; Heydari A. Ultrasonic Synthesis and Characterization of 2D and 3D Metal–Organic Frameworks and Their Application in the Oxidative Amidation Reaction. ACS Omega 2020, 5, 21412–21419. 10.1021/acsomega.0c01773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gordon J.; Kazemian H.; Rohani S. MIL-53(Fe), MIL-101, and SBA-15 Porous Materials: Potential Platforms for Drug Delivery. Mater. Sci. Eng., C 2015, 47, 172–179. 10.1016/j.msec.2014.11.046. [DOI] [PubMed] [Google Scholar]
- Yao B.; Lua S.-K.; Lim H.-S.; Zhang Q.; Cui X.; White T. J.; Ting V. P.; Dong Z. Rapid Ultrasound-Assisted Synthesis of Controllable Zn/Co-Based Zeolitic Imidazolate Framework Nanoparticles for Heterogeneous Catalysis. Microporous Mesoporous Mater. 2021, 314, 110777. 10.1016/j.micromeso.2020.110777. [DOI] [Google Scholar]
- Ghasemi S.; Yousefi M.; Nikseresht A.; Omidi H. Covalent Binding and In-Situ Immobilization of Lipases on a Flexible Nanoporous Material. Process Biochem. 2021, 102, 92–101. 10.1016/j.procbio.2020.12.013. [DOI] [Google Scholar]
- Veisi H.; Nikseresht A.; Mohammadi S.; Hemmati S. Facile In-Situ Synthesis and Deposition of Monodisperse Palladium Nanoparticles on Polydopamine-Functionalized Silica Gel as a Heterogeneous and Recyclable Nanocatalyst for Aerobic Oxidation of Alcohols. Cuihua Xuebao 2018, 39, 1044–1050. 10.1016/S1872-2067(18)63049-9. [DOI] [Google Scholar]
- Veisi H.; Nikseresht A.; Ahmadi N.; Khosravi K.; Saeidifar F. Suzuki–Miyaura Reaction by Heterogeneously Supported Pd Nanoparticles on Thio-Modified Multi Walled Carbon Nanotubes as Efficient Nanocatalyst. Polyhedron 2019, 162, 240–244. 10.1016/j.poly.2019.01.070. [DOI] [Google Scholar]
- Veisi H.; Nikseresht A.; Rostami A.; Hemmati S. Fe3O4@PEG Core/Shell Nanoparticles as Magnetic Nanocatalyst for Acetylation of Amines and Alcohols Using Ultrasound Irradiations under Solvent-Free Conditions. Res. Chem. Intermed. 2019, 45, 507–520. 10.1007/s11164-018-3614-7. [DOI] [Google Scholar]
- Nikseresht A.; Aderang E. Encapsulation of Phosphotungstic Acid in the Nanostructure of Metal-Organic Framework as a Heterogonous Catalyst Used for Fries Rearrangement of O-Acyloxy Benzenes in Para-Situation. Appl. Chem. 2021, 16, 25–38. 10.22075/chem.2021.22125.1934. [DOI] [Google Scholar]
- Lotfi S.; Nikseresht A.; Rahimi N. Synthesis of Fe3O4@SiO2/Isoniazid/Cu(II) Magnetic Nanocatalyst as a Recyclable Catalyst for a Highly Efficient Preparation of Quinolines in Moderate Conditions. Polyhedron 2019, 173, 114148. 10.1016/j.poly.2019.114148. [DOI] [Google Scholar]
- Esfandiar N.; Nasernejad B.; Ebadi T. Removal of Mn(II) from Groundwater by Sugarcane Bagasse and Activated Carbon (a Comparative Study): Application of Response Surface Methodology (RSM). J. Ind. Eng. Chem. 2014, 20, 3726–3736. 10.1016/j.jiec.2013.12.072. [DOI] [Google Scholar]
- Ahmadi M.; Masoumi S.; Hassanajili S.; Esmaeilzadeh F. Modification of PES/PU Membrane by Supercritical CO2 to Enhance CO2/CH4 Selectivity: Fabrication and Correlation Approach Using RSM. Chin. J. Chem. Eng. 2018, 26, 2503–2515. 10.1016/j.cjche.2017.11.018. [DOI] [Google Scholar]
- Cui Z.; Liu C.; Lu T.; Xing W. Polyelectrolyte Complexes of Chitosan and Phosphotungstic Acid as Proton-Conducting Membranes for Direct Methanol Fuel Cells. J. Power Sources 2007, 167, 94–99. 10.1016/j.jpowsour.2006.12.112. [DOI] [Google Scholar]
- Liu H.; Chen S.; Zhang X.; Dong C.; Chen Y.; Liu Z.; Tan H.; Zhang W. Structural Elucidation, Total Synthesis, and Cytotoxic Activity of Effphenol A. Org. Biomol. Chem. 2020, 18, 9035–9038. 10.1039/D0OB01985B. [DOI] [PubMed] [Google Scholar]
- Murashige R.; Hayashi Y.; Ohmori S.; Torii A.; Aizu Y.; Muto Y.; Murai Y.; Oda Y.; Hashimoto M. Comparisons of O-Acylation and Friedel–Crafts Acylation of Phenols and Acyl Chlorides and Fries Rearrangement of Phenyl Esters in Trifluoromethanesulfonic Acid: Effective Synthesis of Optically Active Homotyrosines. Tetrahedron 2011, 67, 641–649. 10.1016/j.tet.2010.11.047. [DOI] [Google Scholar]
- Liuxin Y.; Hao T.; Haijuan Z.; Dasheng W.; Nianhai C.. Preparation Method of High-Purity p-Hydroxyacetophenone. CN 106916060 A, CN 106916060 B, 2017.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








