Introduction

Graphene oxide (GO) is a compound with surface or edge functional groups and superior chemical structures and features, making it an excellent support for various nanostructures1,2,3,4. As graphene oxide has oxygen atoms in diverse isomers such as epoxide, carbonyl, carboxyl, and hydroxyl on its basal planes, various NPs can interact with GO sheets through charge-transfer interactions, physisorption, or electrostatic binding. This biodegradable nanostructure helps enhance the bioavailability of the nanocomposites and has some other advantages like high levels and thermal stability. GO acts as an anchoring support enabling uniform dispersion of the nanocomposites5.

In the last decade, nanoporous metal–organic frameworks (MOFs) comprising organic linkers and inorganic building units have been the subject of intense research6. This novel hybrid material includes a crystalline two or three and 1D open framework7,8,9. MOF-based structures have unique features such as structural diversity, high porosity, high pore volume, and higher capacity for guest molecular acceptance10. MOFs are used in various fields like separation and storage of gases and as a catalyst11,12,13. MOFs represent higher flexibility for designing catalytic centers than purely inorganic porous materials14. This flexibility makes them attractive substances for heterogeneous catalysts. There are several methods to modulate MOFs owing to the significant variety of metal ions and organic linkers. The appearance and size of the framework can be altered via different linkers and altering their connectivity by adding various functional groups on the linkers15,16. Using metal–organic frameworks as green catalysts accelerates the organic reactions. As mentioned in the previous authoritative articles, the metal ions in the structure of MOFs act as a Lewis acid and the carboxylate anion as a Lewis base. In this respect, the IRMOF-3 structure has been used as a bifunctional catalyst in the IRMOF-3/GO/CuFe2O4 structure due to the presence of metal cation sites (Zn2+) and also carboxylate anions17. The existence of non-coordinated amino groups structure in IRMOF-3 demonstrates that the basicity of the aniline-like amino group is enhanced when incorporated inside the pores of MOF channels.

Using multi-component reactions (MCRs) as an organized and efficient method has caused a revolution in designing novel approaches and creating multifarious chemical libraries with molecular diversity in the last decade. They have specific features such as higher atomic economy, straightforward reaction model, shorter reaction time, and conformity with other green chemistry principles. Hence, these chemical reactions have become the most potent and prevailing approach to synthesizing heterocyclic compounds18. Recently, numerous studies have been conducted on developing novel synthetic techniques to synthesize quinoline and pyridine compounds. This high attention is attributed to their promising and notable diversity in pharmacological and therapeutic activities like geroprotective, antitubercular, anticancer, antidiabetic, and analgesic activities19. Different developed synthetic approaches and catalysts, such as LiBr, CTAB, CuBr, salicylic acid, L-proline, and GO nanoparticles20,21,22,23,24,25, have been proposed to synthesize these biological and chemical molecules. However, they mostly encounter some disadvantages and negative points such as applying impotent and inefficient catalysts, low yield of the products, difficult reaction conditions, and tedious work-up procedures. Thus, designing and providing novel synthetic methods for these biological molecules is challenging.

The present study aims to enhance more effective synthetic procedures, lower the number of isolated reaction stages, and minimalize wastes by preparing the heterocyclic compounds26,27,28. Next, we report a new and mild route for synthesizing heterocyclic compounds through a one-pot reaction between various aromatic aldehydes, diverse primary amines, malononitrile, and dimedone under ultrasound irradiations in the attendance of IRMOF-3/GO/CuFe2O4 nanocatalyst (Fig. 1).

Figure 1
figure 1

Synthesis of pyridines and quinolines using IRMOF-3/GO/CuFe2O4.

Experimental

Preparing CuFe2O4 nanoparticles

A solution of FeCl2.4H2O and CuCl2.2H2O (molar ratio, 2:1) in deionized water (30 mL) was added to a concentrated solution of NaOH to reach a pH of 13. Then, the mixture was stirred vigorously at 70 °C until to formation of a black residue, including CuFe2O4. The precipitate was rinsed several times after cooling and dried overnight at 80 °C. Finally, it was calcinated for 5 h in a furnace at 700 °C.

Preparation of IRMOF-3/CuFe2O4 nanocomposite

A mixture of CuFe2O4 (0.05 g), Zn (NO3)2·6H2O (0.066 g) and NH2-BDC (0.016 g) was dissolved in DMF (10 mL) under vigorous stirring. Next, the solution was transferred to an autoclave and heated at 120 °C for 10 h. The precipitate was separated after cooling, by a simple filtration. The obtained crystals were then soaked in 20 mL of DMF at 80 °C for 12 h. Lastly, the obtained IRMOF-3/CuFe2O4 was dried for 24 h at 40 °C29.

Synthesis of IRMOF-3/GO/CuFe2O4 nanocatalyst

In this research, our catalyst including IRMOF-3/GO/CuFe2O4 was successfully prepared according to the modified Hummer's method and our previously published paper30,31, (See Supplementary Information, Fig. S1).

General synthesis of 2-amino-4-aryl-6-substituted pyridine-3,5-dicarbonitrile (4a-r) using IRMOF-3/GO/CuFe2O4 nanocatalyst

IRMOF-3/GO/CuFe2O4 nanocomposite (0.003 g) was added to a solution of aromatic amine (1 mmol), malononitrile (2 mmol), and aldehyde (1 mmol) in 5 mL of EtOH. The mixture was sonicated under ultrasonic irradiations (25 kHz frequency) for 10–20 min. After completing the reaction, the obtained solid was filtered, and the contents were dissolved in acetone, and the insoluble IRMOF-3/GO/CuFe2O4 nanocomposite was separated by centrifuge. In the last step, the solvent was evaporated under a vacuum, and the precipitate was recrystallized using ethanol to provide the corresponding 2-amino-4-aryl-6-substituted pyridine-3,5-dicarbonitrile.

General synthesis of hydroquinoline-3-carbonitrile derivatives (6a-d) in the presence of IRMOF-3/GO/CuFe2O4 nanostructure

A mixture of anilines (1 mmol), malononitrile (1.1 mmol), aldehyde (1 mmol), dimedone (2 mmol), and IRMOF-3/GO/CuFe2O4 (0.005 g) was added to a flask containing ethanol (5 mL). The mixture was sonicated for the appropriate times. The catalyst was separated over completion (monitored by TLC), by dissolving the obtained heterocyclic compounds in dichloromethane. Finally, the corresponding products in high purity were collected by evaporating the dichloromethane.

The scanned original spectral data of new compounds are provided in Supporting Information.

2-amino-4-(4-cyanophenyl)-6-((4-methoxyphenyl) amino) pyridine-3,5-dicarbonitrile 4j

Yellow solid; m.p. 223–225 °C. IR spectrum ν, cm−1: 3365, 3172, 3051, 2932, 2370, 1705,1674, 1591, 1458, 1205; 1H NMR (250 MHz, DMSO-d6): 3.84 (s, 3H, CH3), 5.43 (s, 2 H, NH2), 6.76–7.04 (d, 2H, J = 8.2 Hz, ArH), 7.28–7.36 (d, 2H, J = 8.3 Hz, ArH), 7.65–7.75 (m, 4H, ArH), 9.85 (s, 1H, NH); 13C NMR (62.9 MHz, DMSO-d6) δ: 27.33, 31.39, 68.62, 76.82, 87.26, 101.03, 112.53, 123.45, 131.34, 133.43, 134.25, 138.56, 144.43, 146.33, 148.62, 156.47, 165.38, 168.28, 170.21, 183.31, 188.22; Anal. Calcd. For: C21H14N6O: C 68.84, H 3.85, N 22.94. O 4.37. Found: C 76.88, H 3.82, N 22.92 O 4.35; MS (EI) (m/z): 366.12 (M+).

2-amino-6-((4-methoxyphenyl)amino)-4-(4-(methylthio)phenyl)pyridine-3,5 dicarbonitrile 4k

Yellow solid; m.p. 210–212 °C. IR spectrum ν, cm–1: 3363, 3170, 2555, 21,381, 1701, 1678, 1593, 1458, 1377, 1207; 1H NMR (250 MHz, DMSO-d6): 2.36 (s, 3H, CH3), 3.72 (s, 3H, CH3), 5.48 (s, 2 H, NH2), 6.76–7.04 (d, 2H, J = 8.2 Hz, ArH), 7.27–7.36 (d, 2H, J = 8.3 Hz, ArH), 7.65–7.78 (m, 4H, ArH), 9.82 (s, 1H, NH); 13C NMR (62.9 MHz, DMSO-d6) δ: 28.39, 32.33, 67.62, 77.82, 86.26, 102.03, 113.52, 125.44, 132.34, 134.43, 136.26, 139.56, 146.44, 148.33, 149.69, 154.47, 162.34, 164.28, 171.21, 181.32, 187.21; Anal. Calcd. For: C21H17N5OS: C 65.10, H 4.42, N 18.08. O 4.13, S 8.27. Found: C 65.14, H 4.45, N 18.04, O 4.316, S 8.25.; MS (EI) (m/z): 387.12 (M+).

2-amino-4-(4-cyanophenyl)-1-(4-methoxyphenyl)-7,7-dimethyl-5-oxo-1,4,5,6,7,8 hexahydroquinoline-3-carbonitrile 6c

Yellow solid; m.p. 242–243 °C. IR spectrum ν, cm–1: 3321, 3062, 2985, 2233, 1608, 1716, 1678, 1381, 1284; 1H NMR (250 MHz, DMSO-d6) δ: 0.82 (s, 3H, CH3), 0.95 (s, 3H, CH3), 2.29 (d, 2H, J = 8.4 Hz, 2CH), 2.44 (d, 2H, J = 8.2 Hz, 2CH), 3.82 (s, 3H, CH3), 4.33 (s, 1 H, CH), 5.33 (s, 2 H, NH2), 6.74–7.05 (d, 2H, J = 8.6 Hz, ArH), 7.28–7.32 (d, 2H, J = 8.3 Hz, ArH), 7.62–7.78 (m, 4H, ArH); 13C NMR (62.9 MHz, DMSO-d6) δ: 28.33, 30.79, 78.61, 88.62, 92.24, 102.06, 117.46, 118.41, 124.54, 128.61, 130.54, 132.93, 133.22, 137.51, 142.63, 144.01, 149.32, 159.27, 166.88, 173.46, 178.81, 180.74, 183.31, 192.12, 194.81, 196.71; Anal. Calcd. For: C26H24N4O2: C 73.56, H 5.70, N 13.20, O 7.54. Found: C 73.54, H 5.73, N 13.24, O 7.52; MS (EI) (m/z): 424.19 (M+).

2-amino-1-(4-methoxyphenyl)-7,7-dimethyl-4-(4-(methylthio)phenyl)-5-oxo-1,4,5,6,7,8 hexahydroquinoline-3-carbonitrile 6d

Yellow solid; m.p. 256–258 °C. IR spectrum ν, cm–1: 3352, 3186, 2962, 2191, 1685, 1651, 1604, 1369, 1211; 1H NMR (250 MHz, DMSO-d6): 0.84 (s, 3H, CH3), 0.97 (s, 3H, CH3), 2.28 (d, 2H, J = 7.3 Hz, 2CH), 2.42 (d, 2H, J = 7.8 Hz, 2CH), 2.78 (s, 3H, SCH3), 3.83 (s, 3H, CH3), 4.32 (s, 1 H, CH), 5.33 (s, 2 H, NH2), 6.75–7.03 (d, 2H, J = 8.3 Hz ArH), 7.27–7.38 (d, 2H, J = 8.6 Hz, ArH), 7.64–7.75 (m, 4H, ArH); 13C NMR (62.9 MHz, DMSO-d6) δ: 27.33, 31.72, 76.26, 85.62, 94.22, 104.06, 116.41, 119.46, 123.54, 126.81, 131.54, 133.93, 135.22, 138.51, 141.62, 143.01, 148.32, 157.24, 164.87, 175.44, 176.86, 182.71, 185.31, 193.11, 193.82, 195.71; Anal. Calcd. For: C26H27N3O2S: C 70.09, H 6.11, N 9.43. O 7.18, S, 7.20. Found: C 70.06, H 6.16, N 9.48 O 7.12, S, 7.24; MS (EI) (m/z): 455.18 (M+).

Results and discussion

IRMOF-3/GO/CuFe2O4 was successfully synthesized as mentioned in our previous study31. The structure of this catalyst was determined using EDX/Mapping, IR, and TGA analysis (See Supplementary Information, Figs. S2–S5)31, and also with other spectral techniques including SEM (Fig. 2), XRD (Fig. 3), BET (Fig. 4), and VSM (Fig. 5).

Figure 2
figure 2

SEM analysis of (a) CuFe2O4, (b) IRMOF-3, (c) GO, and (d) IRMOF-3/GO/CuFe2O4.

Figure 3
figure 3

XRD patterns of (a) pure CuFe2O4, (b) IRMOF-3, (c) GO, and (d) IRMOF-3/GO/CuFe2O4.

Figure 4
figure 4

BET-plot of CuFe2O4 (a), adsorption/desorption of CuFe2O4 (b), BJH-plot of CuFe2O4 (c); BET-plot of GO (d), adsorption/desorption of GO (e), BJH-plot of GO (f); BET-plot of IRMOF-3 (g), adsorption/desorption of IRMOF-3 (h), BJH-plot of IRMOF-3 (i), BET-plot of IRMOF-3/GO/CuFe2O4 (j), adsorption/desorption of IRMOF-3/GO/CuFe2O4 (k), BJH-plot of IRMOF-3/GO/CuFe2O4 (l).

Figure 5
figure 5

Magnetization curves of IRMOF-3/CuFe2O4 and IRMOF-3/GO/CuFe2O4.

The microscopic morphology of the products was visualized by scanning electron microscopy (SEM). This analysis showed that CuFe2O4 is composed of relatively uniform quasispherical particles (Fig. 2a)32. SEM images of IRMOF-3 showed that this structure has a crystalline and rod shape, per previous reports (Fig. 2b)33. SEM image of the GO shows the particles of GO look very dense with the layers stacked together due to dispersive forces and strong specific interactions between the surface groups on the graphene-like layers (Fig. 2c)34. SEM images for the IRMOF-3/GO/CuFe2O4 nanocomposite are shown in Fig. 3d. As can be seen, the particles have regular and spherical shapes (Fig. 2d).

XRD pattern of the CuFe2O4 nanoparticles is shown in Fig. 3a. The peaks at 29.80°, 32.21°, 35.71°, 43.31°, 48.40°, 57.25°, 59.66° and 63.62° (2θ°) indicate the formation of the CuFe2O4 nanoparticles35. Also the peaks at 8.4°, 9.6°, 13.7°, 15.35°,19.39° and 24.62° (2θ°) show the presence of the IRMOF-3 frameworks (Fig. 3b)36. Here, the peak at 2θ = 11.1° indicates the creation of the GO (Fig. 3c)37,38. Fig. 2d presents the XRD pattern of IRMOF-3/GO/CuFe2O4. The peaks observed at 2θ = 29.80°, 32.21°, 35.71°, 43.31°, 48.40°, 57.25°, 59.66° and 63.62° are related to the CuFe2O4. Also, the peaks at 2θ of 8.4°, 9.6°, 13.7°, 15.35°, 19.39°, and 24.62° show the presence of the IRMOF-3 frameworks, while the peak at 2θ = 11.1° indicates the creation of the GO, confirming the formation of the nanocomposite. These results demonstrate that the crystalline structure of the CuFe2O4 and IRMOF-3 materials was unchanged and remained intact during the catalyst preparation.

Nitrogen adsorption/desorption isotherms were used to study the specific surface area and pore volume distribution of nanostructures including CuFe2O4, GO, IRMOF-3, and IRMOF-3/GO/CuFe2O4 by the Brunauer–Emmett–Teller (BET) approach (Fig. 4). The CuFe2O4 represented a BET surface area, total pore volume, and the average pore diameter of 34.65 m2 g−1, 0.138 cm3 g−1, and 1.17 nm respectively. In comparison, these analyses for GO, IRMOF-3 and IRMOF-3/GO/CuFe2O4 gave values of 48.51 m2 g−1, 0.182 cm3 g−1, and 1.24 nm, 884.82 m2 g−1, 0.887 cm3 g−1, and 1.42 nm, and 456.29 m2 g−1, 0.435 cm3 g−1 and, 1.65 nm, respectively (Fig. 4 and Table 1). The adsorption–desorption isotherm of all the four samples exhibits a reversible type-II adsorption isotherm, indicating the presence of micro-and macro-pores39. The decreasing BET surface area and total pore volume can be attributed to incorporating of IRMOF-3/CuFe2O4 groups inside the pure GO. However, the open cavities and high surface areas were retained, which benefited the free dispersal of the reactant and product.

Table 1 Measured BET surface areas and pore volumes of CuFe2O4, GO, IRMOF-3 and IRMOF-3/GO/CuFe2O4.

The so-called VSM analysis was further conducted to investigate the magnetic behavior of the IRMOF-3/CuFe2O4, and IRMOF-3/GO/CuFe2O4 nanocomposite, with the outcomes presented in Fig. 5. According to the results, the value of magnetic saturation was measured at 67.45, and 38.45 emu/g for the IRMOF-3/CuFe2O4 and IRMOF-3/GO/CuFe2O4 nanocomposite, respectively40.

The performance of the prepared IRMOF-3/GO/CuFe2O4 nanocomposite was evaluated using the catalyst in the preparation of 2-amino-4-aryl-6-substituted pyridine-3,5-dicarbonitrile and hydroquinoline-3-carbonitrile derivatives. To this end, a three-component reaction of 4-bromobenzaldehyde, p-toluidine, and malononitrile was chosen as a model study for synthesizing 2-amino-4-(4-bromophenyl)-6-(p-tolylamino)pyridine-3,5-dicarbonitrile (4e). In addition, the reaction of p-toluidine, 4-bromobenzaldehyde, dimedone, and malononitrile was considered a model reaction for the preparation of 2-amino-4-(4-bromophenyl)-7,7-dimethyl-5-oxo-1-(p-tolyl)-1,4,5,6,7,8-hexahydroquinoline-3-carbonitrile (6b) (Table 2). The model reactions were investigated by different conditions, including catalyst, solvent, and amount of catalyst.

Table 2 The influence of various catalysts and solvents in the preparation of 4e and 6b compounds.

Firstly, the preparation of compounds 4e and 6p was considered in the occurrence of different catalysts like Fe3O4, CuFe2O4, MgO, SiO2, IRMOF-3/CuFe2O4, CuI, GO, and IRMOF-3/GO/CuFe2O4 nanostructures under ultrasound irradiations in EtOH as solvent (0.01 g of each catalyst was used). The best results were obtained (Table 2), when using IRMOF-3/GO/CuFe2O4 nanocomposite as catalyst (Table 2, entries 8 and 16). Afterward, it was tried to investigate the influence of various solvents like water, chloroform, dimethylformamide, ethanol, toluene, and solvent-free conditions in synthesizing of 2-amino-4-(4-bromophenyl)-6-(p-tolylamino)pyridine-3,5-dicarbonitrile (4e) in the presence of IRMOF-3/GO/CuFe2O4 (0.01 g) (Table 2, entries 17–22). Table 2 shows no product was achieved without any solvent under ultrasonic irradiation. Furthermore, it was concluded that the presence of ethanol as solvent provided the best outcome considering reaction time and yield of the corresponding product. Subsequently, the preparation of hexahydroquinoline-3-carbonitrile (6b) was also examined using different polar and nonpolar solvents (Table 2, entries 23–28). As indicated in this Table 2, using EtOH as solvent provided the best reaction conditions (Table 2, entries 20 and 26).

In the next stage, it was decided to assess the optimum amount of the nanocatalyst for preparing of 4e and 6b. Different amounts of the catalyst were used for preparing 2-amino-4-(4-bromophenyl)-6-(p-tolylamino)pyridine-3,5-dicarbonitrile and 2-amino-4-(4-bromophenyl)-7,7-dimethyl-5-oxo-1-(p-tolyl)-1,4,5,6,7,8-hexahydroquinoline-3-carbonitrile (Table 3). The model was run using various amounts of the catalyst from 0.001 to 0.007 g. Eventually, the optimal amounts were found to be 0.003 and 0.005 g for compounds 4e and 6p, respectively.

Table 3 The effect of different quantities of the catalyst in model studies.

Some aromatic aldehydes and amines were used to investigate the catalytic efficiency and performance of the IRMOF-3/GO/CuFe2O4 on synthesizing various pyridine and hydroquinoline derivatives under optimized reaction conditions. After conducting some experiments, we obtained some 2-amino-4-aryl-6-substituted-pyridine-3,5-dicarbonitriles and hydroquinolone-3-carbonitriles in good to excellent yields within short reaction times. As illustrated in Table 4, different aldehydes and anilines with various substituents could participate in the multi-component synthesis of corresponding heterocyclic compounds. However, as indicated, benzaldehydes with electron-withdrawing groups reacted faster than electron-donating ones. Moreover, the aromatic amines having electron-releasing groups reacted faster than anilines containing electron-withdrawing groups. According to the mechanism pathway, the first step of the mechanism is the Knoevenagel condensation between malononitrile and aromatic aldehyde. Obliviously, the presence of electron-withdrawing groups on the benzaldehydes leads to a faster reaction due to more electrophilicity of the carbon group of the carbonyl41,42. While using IRMOF-3/GO/CuFe2O4 as a catalyst serves as a Lewis acid catalyst and increases the electrophilicity of the carbonyl groups of aldehydes. The synergic effects of both electron-withdrawing groups and Lewis acidity of IRMOF-3/GO/CuFe2O4 increase the reaction rate. Nevertheless, against the electron-withdrawing groups, the existence of electron-donating groups lowers the speed of the reaction, because of a slower nucleophilic attack on carbonyl groups of aldehydes.

Table 4 Preparation of 1,8-dioxo-decahydro-acridines (4a–r) and 1,8-dioxo-octahydro-xanthenes (6a–d) using IRMOF-3/GO/CuFe2O4 under ultrasound irradiations.

The catalyst efficiency and usability for synthesizing 2-amino-4-aryl-6-substituted pyridine-3,5-dicarbonitrile and hydroquinoline-3-carbonitrile derivatives were compared with those of some previously reported catalysts. As presented in Table 5, IRMOF-3/GO/CuFe2O4 is better than previously reported catalysts in saving time, and energy, and excellent yields of the products (Table 5).

Table 5 Comparison of the outcomes of the production of 2-amino-4-aryl-6-substituted pyridine-3,5-dicarbonitrile and hydroquinoline-3-carbonitrile derivatives by means of different catalysts.

Reuse of catalyst

In this research, the reusability of IRMOF-3/GO/CuFe2O4 composite was investigated for the three-component reaction of 4-bromobenzaldehyde, p-toluidine, and malononitrile (compound 4e). Upon completion of the reaction, the IRMOF-3/GO/CuFe2O4 nanostructure was separated using a centrifuge instrument and washed with dichloromethane. The obtained recovered catalyst was dried at 40 °C for 24 h. As indicated in Fig. 6, the recovered catalyst could be used 7 times. Furthermore, the XRD technique was used to demonstrate the efficiency and consistency of the rescue IRMOF-3/GO/CuFe2O4 after seven runs uses. As indicated, there is no considerable differences between X-ray diffraction of the fresh IRMOF-3/GO/CuFe2O4 and the recovered catalyst, confirming the good stability of the prepared nanocatalyst after 7 runs (Fig. 7).

Figure 6
figure 6

Reusability of the IRMOF-3/GO/CuFe2O4 catalyst.

Figure 7
figure 7

XRD pattern of the recovered IRMOF-3/GO/CuFe2O4.

Figure 8 shows a reasonable mechanism for synthesizing pyridine derivatives (4a–4r) catalyzed by IRMOF-3/GO/CuFe2O4, This result is, supported by previous studies41,42. IRMOF-3/GO/CuFe2O4 was assumed to serve as Lewis acids that causes increasing electrophilicity of the carbonyl groups, double and triple bonds in substrates and intermediate via a strong coordination bond17,49. Initially, the mechanistic approach includes the Knoevenagel condensation of aromatic aldehyde and malononitrile, creating adduct I. Then, adding anilines to intermediate I was followed by involving malononitrile to make intermediate II. Tautomerization and intramolecular cyclization on intermediate II produced intermediate III with a dihydropyridine moiety. Finally, 2-amino-6-alkylamino-3,5-dicyanopyridines 4 was prepared by aromatization of intermediate III. Additionally, the catalytic behavior of IRMOF-3/GO/CuFe2O4 on the quinoline synthesis was identical to the described mechanism for synthesizing pyridine derivatives.

Figure 8
figure 8

The proposed mechanism to synthesize pyridine derivatives using IRMOF-3/GO/CuFe2O4.

Conclusion

The present work aimed to present a convenient and facile technique to prepare magnetic copper ferrite nanoparticles supported on IRMOF-3/GO. This magnetic heterogeneous and reusable nano catalyst to prepare heterocyclic compounds via the reactions between of various aromatic aldehydes, malononitrile, diverse primary amines and dimedone under ultrasound irradiations. The methods for characterizing the IRMOF-3/GO/CuFe2O4 were SEM, BET, FT-IR, XRD, TGA, and EDX. It was indicated that this catalyst is effective for synthesis of pyridines and quinoline derivatives. Moreover, this technique has good benefits like eco-friendly, higher performance and a very convenient working technique.