Revolutionizing acridine synthesis: novel core-shell magnetic nanoparticles and Co-Zn zeolitic imidazolate framework with 1-aza-18-crown-6-ether-Ni catalysts | Scientific Reports

Blog

HomeHome / Blog / Revolutionizing acridine synthesis: novel core-shell magnetic nanoparticles and Co-Zn zeolitic imidazolate framework with 1-aza-18-crown-6-ether-Ni catalysts | Scientific Reports

Nov 01, 2024

Revolutionizing acridine synthesis: novel core-shell magnetic nanoparticles and Co-Zn zeolitic imidazolate framework with 1-aza-18-crown-6-ether-Ni catalysts | Scientific Reports

Scientific Reports volume 14, Article number: 25739 (2024) Cite this article 258 Accesses Metrics details Nanoparticles have emerged as a critical catalyst substrate due to their exceptional features,

Scientific Reports volume 14, Article number: 25739 (2024) Cite this article

258 Accesses

Metrics details

Nanoparticles have emerged as a critical catalyst substrate due to their exceptional features, such as catalytic efficiency, high stability, and easy recovery. In our research, we have developed an innovative and environmentally friendly magnetic mesoporous nanocatalyst. Using the co-precipitation method, we produced magnetic nanoparticles (Fe3O4) and coated them with Zeolitic imidazolate frameworks (ZIFs) to enhance their surface area and chemical stability. The resulting substrate was functionalized with 1-aza-18-crown-6-ether and nickel metal. Our prepared catalyst has been rigorously evaluated using advanced techniques, including X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FT-IR), Brunauer-Emmet-Teller (BET), vibrating sample magnetometry (VSM), scanning electron microscopy and energy dispersive X-ray (SEM-EDS), inductively coupled plasma (ICP), elemental mapping analysis (EMA), thermogravimetric analysis (TGA), and transmission electron microscopy (TEM). By synthesizing acridine derivatives, we have demonstrated the exceptional efficiency of our catalyst in organic compound synthesis. Through optimization, we have established the ideal parameters for catalytic processes, including catalyst amount, temperature, time, and ultrasonic use. Our catalyst has been proven to exhibit remarkable physical and chemical properties, such as porosity, temperature resistance, and recyclability. Notably, our heterogeneous nanocatalyst has shown outstanding performance and can be recycled six times without any loss in efficiency, affirming its potential in acridine.

The critical properties of nanomaterials that cause the creation of technology and economic and industrial growth in this field draw attention to engineered nanomaterials and their widespread use1. Numerous diverse nanocatalysts are favored for their non-toxic, eco-friendly, and cost-efficient attributes, resulting in high yields of desired products. Convenient separation and reusability further elevate their appeal over homogeneous catalysts2,3. Homogeneous catalysts are difficult to use because they are homo-phases with the reaction system4. There is great interest in the design of catalysts easily separated from the system of critical organic reactions. Developing recyclable magnetic catalysts could be a solution for proper separation and high recycling ability5. Catalysts are vital for synthesizing chemical compounds, but using them in large quantities can be challenging. Economic and environmental factors drive the need for selective heterogeneous catalysts to reduce production costs and minimize the disposal of catalyst waste in liquid-phase reactions6. The heterogeneous catalyst is environmentally friendly because it is easily separated from the reaction vessel and reused multiple times. And it has attracted a lot of attention7,8.

Magnetic nanoparticles and heterogeneous catalytic systems are important branches in synthesizing organic compounds in the liquid phase9. Due to their excellent catalytic properties, environmental friendliness, and large surface area, nanocatalysts have an essential impact on organic chemistry10. Despite limited atom economy, harsh reaction conditions, and significant chemical waste, sustainability has been overlooked in the face of transformative discoveries. Nonetheless, we are delighted that modern chemistry is evolving, with a growing emphasis on minimizing harmful environmental impacts in reaction designs11,12. Crystalline porous materials made of metal cations and anions of organic ligands are called metal-organic frameworks (MOFs)13,14. Due to their robustness, easy tuneability, exceptional porosity, defined structure, geometry, and architecture control, these structures are becoming one of the newest materials in chemistry15,16. Recently, many MOFs have been synthesized due to the essential features of these materials, which have great potential for use in gas storage17, water purification18, catalysts19, sensors20, biomedical21, and electronics22, energy conversion and storage23. Zeolitic imidazolate frameworks made of cobalt and zinc metal cations and 2-methylimidazole organic ligands are a porous structure of MOFs that have suitable properties for catalysts in producing organic compounds24,25,26. ZIF67 and ZIF8 have comprehensive chemical active sites. For example, catalysts based on ZIF8 have many active sites, high surface area, and high-temperature resistance, while catalysts based on ZIF67 have a higher degree of surface modification and excellent biocompatibility27,28.

As an appropriate and clean method, multicomponent reactions (MCRs) have been readily available to produce a wide range of organic materials in recent years. The reaction occurs in a reaction vessel composed of three or more reactants. Due to the favorable economic conditions and time savings, MCRs are a popular option29,30. Using MCR reactions reduces the number of steps required to reach the product. Therefore, these reactions are compatible with the environment and have high efficiency31. There has been a significant increase in the use of ultrasound technology for modifying chemical reactions, particularly organic reactions. Ultrasonic waves in the liquid phase produce cavitation, including nucleation, growth, and collapse of bubbles. Millions of places across the liquid medium experience bubble bursting, which creates supercritical conditions with temperatures as high as 5000 K and pressures as high as 1000 atm32,33,34. Ultrasonic technology is revolutionizing organic reactions due to its numerous advantages over traditional methods. These include milder reaction conditions, higher yields, faster reaction times, and cleaner reactions35,36.

Acridine derivatives, as an essential class of heterocyclic compounds, have been investigated due to their high potential in biological activities and the treatment of several diseases, such as antimalarial37, antibacterial38, cytotoxic39, antiviral40, and cancer treatment by binding to DNA41 (Fig. 1). Today, many scientists are looking for new and effective drugs that are less toxic because many people in the world have cancer and viral, bacterial, and genetic diseases42. Acridine derivatives are an important class of nitrogen-containing heterocycles. A three-ring aromatic system that is a derivative of anthracene. Acridine forms a wide range of drugs with chemical and physical properties, low toxicity, and practical efficacy43,44. The traditional method of synthesizing organic compounds without catalysts presents several drawbacks, such as low efficiency, prolonged reaction times, increased costs, and the use of hazardous solvents. By employing catalysts in conjunction with microwave and ultrasound waves, a more suitable and efficient approach to organic compound synthesis can be achieved. Notably, recent advancements have shown successful synthesis of acridines using microwaves and ultrasound45,46.

Catalysts based on nanomaterials are one of the important components of green chemistry, especially catalysts based on magnetic nanoparticles, which are used in synthesizing organic compounds such as acridine derivatives due to their efficient recycling. To protect magnetic nanoparticles, the physical and chemical stability of colloidal systems, improve their surface, and increase their dispersion in aqueous environments, they need to modify the surface for proper performance as a catalyst47,48,49. In this research, ZIF67 and ZIF8 were used to modify the surface of magnetic nanoparticles due to their suitable chemical and physical properties and high porosity and contact surface. MOFs were considered suitable coatings for heterogeneous catalysts and magnetic cores that easily separated from the reaction medium. Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni nanocatalyst is of great interest due to its excellent catalytic properties, compatibility with the environment, and large surface area due to the use of separable magnetic. The core of magnetic nanoparticles was used for easy separation of the catalyst, and the zeolite frameworks of imidazolate and 1-aza-18-crown-6-ether with a large surface area and many holes were used to modify the surface of magnetic nanoparticles nickel as an active metal was used l in catalytic processes.

Following up on our research on catalysts50,51,52,53,54,55, we introduce Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni as a novel and green nanocatalyst for the impressive manufacturing of acridines in a one-pot reaction. As a result, this catalyst offers excellent performance, a large surface area, a reusable form, easily removed, and a simple process (Fig. 2).

Examples of acridine structures with medicinal applications.

All materials were acquired from Merck and Aldrich without any purification. Thermogravimetric analysis (TGA) was carried out on a TA Q600. X-ray diffraction (XRD) was assayed on a Philips PW1730. Morphology and elemental analysis were provided via FESEM and EDS using the MIRA III. The BET and BJH analysis were used to measure the pore volume, particle size, and surface area using a BELSORP MINI II. Infra-red spectra were obtained with a Thermo-Nicolet Nexus 670 device. VSM evaluated the magnetic properties using a Lake Shore VSM 7410. TEM micrograph was performed using a Philips CM30 instrument. The NMR spectra were provided using DRX-300 from Bruker company.

The co-precipitation procedure was utilized to produce Fe3O4 in this study: A solution of FeCl3·6H2O (3.0 g) and FeCl2·4H2O (1.5 g) in H2O (50 mL) was provided. Afterward, the mixture was sonicated for 75 min in the ultrasonic processor. NH4OH (25%, 10 mL) was added drop-by-drop into the past solution after 15 min of starting ultrasonication (TOPSONICS ultrasonic liquid processor, made in Iran, 50 W). Subsequently, a uniform black precipitate was obtained, which was separated using a magnet. The obtained compound was rinsed in EtOH/H2O and dried (at 60 °C).

Fe3O4@ZIF67@ZIF8 was prepared using a sonochemical process. Co(NO3)2·6H2O (2.94 g) was dissolved in methanol (35 mL), and the magnetic nanoparticles obtained from the previous step were mixed with ultrasonication for 15 min, followed by Sonication (80 W) for 30 min after the addition of 2-methylimidazole. Zn(NO3)2·6H2O (2.89 g) was added to the ultrasonicated solution, and the Sonication was maintained for 30 min. The resulting product was collected with the help of an external magnet, rinsed several times with methanol, and dried at 60 °C for 6 h56.

Fe3O4@ZIF67@ZIF8 (1.9 g) synthesized in the previous step was added to EtOH (40 mL) with CPTMS (3-chloropropyl-tri-methoxysilane) (2.1 g), followed by the reaction under reflux for 24 h at 60 °C. After that, the obtained product was rinsed with EtOH, and finally, Fe3O4@ZIF67@ZIF8@-(CH2)3-Cl was collected using an external magnet and dried.

Fabrication of preservative with 1-aza-18-crowne-6-ether: The substance obtained from the previous step (1.1 g) was dissolved in ethanol (30 mL). Then, 1-aza-18-crowne-6-ether (1 g) was added to the solution. Finally, triethylamine (2 mL) was added to the reagent and underwent reflux for 20 h. Then dried at 70 °C for 12 h.

Fabrication of metal catalyst: Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether (1.1 g) was dissolved with nickel salt (NiSO4·6H2O) at a ratio of 1 to 0.5 in ethanol (30 mL). It was placed under reflux at 60 °C for 24 h, then collected and dried.

A round-bottom flask with a magnet was used to dissolve 2 mmol dimedone in 20 mL H2O at 75 °C. Then, arylglyoxal (1 mmol) was added to the mixture. Ammonium acetate (1 mmol) and nanocatalyst Fe3O4@ZIF-67@ZIF-8@1-aza-18-crown-6-ether-Ni (10 mg) were added to the reaction medium for several minutes. The obtained solution was subjected to ultrasonication for three minutes. After three minutes of ultrasonication, the reaction medium was allowed to cool to room temperature. TLC was used to determine whether the one-pot reaction was complete. As a black deposit, a magnetic nanocatalyst was made. It was drawn apart by the magnet, yielding acridines 4a–j (75–98%) as the desired product (Table 1; Fig. 2). FT-IR, 1H-NMR, 13C-NMR and the melting temperatures of the end compounds were used to identify them (The FT-IR, 1H-NMR, and 13C-NMR spectra were inserted in a supplementary file).

Schematic of obtaining acridine derivatives via three-component reactions of aryl glyoxal, ammonium acetate, and cyclic 1,3-dione using a novel nanoporous and heterogeneous separable catalyst.

The heterogeneous catalyst was prepared with magnetic nanoparticles coated with Co/Zn-ZIFs and placed 1-aza-18-crown-6-ether-Ni on them. Necessary investigations were done to identify the structure of this catalyst. Then, the prepared catalyst was used to prepare acridines, and various parameters such as solvent effect, amount of catalyst, temperature, and time were investigated and optimized. A brief pathway of the catalyst synthesis state is presented in Fig. 3.

Stepwise synthesis of Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

The morphological and chemical structure, thermal, and magnetic virtues of nanocatalyst were successfully assessed by various techniques, including FT-IR, XRD, TGA, BET, SEM-EDS, EMA, TEM, ICP, and VSM.

First, the type of bonds among components and function groups have been defined by the interaction of IR light with the material. FT-IR spectroscopy is the analysis of infrared light interacting with a molecule. Organic compounds absorb red light radiation that corresponds to these vibrations in terms of energy. Figure 4 is an infrared spectrum obtained, a graph of the intensity of the infrared spectrum measured against the wavelength of light. Information on group function and interactions between molecules has been obtained by measuring the vibrations of atoms. Figure 4a is related to magnetic nanoparticles synthesized in the first step. The vibration in 558 cm−1 is related to Fe–O, and the bands of 1613 and 3414 cm−1 are related to bending and stretching O–H connections, which are placed on the surface of Fe3O4 nanoparticles, confirming the formation of Fe3O4 nanoparticles57,58. According to Fig. 4b of the spectrum after coating magnetic nanoparticles with Co/Zn-ZIFs, a band at 2920 cm−1 can be assigned to the C–H bond vibration characteristic. The characteristic absorption bands at 1570 cm−1 for C = N and 1136, 430, 559, and 685 cm−1 correspond to C–N, Fe–O, N-Zn, and N-Co, respectively, which indicates the existence of Co/Zn-ZIFs compounds59. C–H Stretching vibrations appeared in Fig. 4c at 2927 cm−1. A band relevant to Si–O and C–Cl connections was confirmed at 820 and 681 cm−1, indicating the linker’s successful connection (CPTMO). According to Fig. 4d, related to ligand binding1-aza-18-crown-6-ether, the curve associated with group N–H stretching and bending was seen in 1579 cm−1. Moreover, the peaks observed at wavelengths 1173 and 1209 cm−1 are related to the stretching vibrations of C–N. After connecting the crown ether to carbon, the peak at 680 cm−1, corresponding to C–Cl, was not observed to show that the correct connection of crown ether to carbon was observed. Finally, the peaks observed in Fig. 4e curve Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni at the wavelengths of 598 and 692 cm−1 correspond to Ni–O and N–Ni respectively, confirming the metal’s successful stabilization on the ligand54. The changes in the spectrum during the synthesis of different catalyst stages with the increase of various compounds confirm the change in the structure and proper obtaining of the catalyst.

Infra-red spectra acquired for Fe3O4 (a), Fe3O4@ZIF-67@ZIF-8 (b), Fe3O4@ZIF67@ZIF8@-(CH2)3-Cl (c), Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether (d), and Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni (e).

X-ray diffraction is a non-destructive and widely accepted method of determining the crystal structure of materials, providing vital information on their properties. In the XRD pattern of Fe3O4 (Fig. 5a) structure with inverted spinel structure, the six peaks appeared at around 2θ = 30°, 35°, 43°, 54°, 57°, and 63°, which correspond to diffraction of the (158), (445), (109), (37), (81), and (115). Figure 5b demonstrates the XRD results for the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni, indicating that six peaks at values of 7.5°, 10°, 13°, 14°, 15°, 18°, 22°, 23°, and 35°, refer to (217), (162), (127), (165), (150), (256), (78), (79), (117) crystal surface, respectively. In addition, the presence of Fe3O4@ZIF67@ZIF8 in the compound structure has been confirmed. Collected coherently scattered peak intensities agree with the crystalline spinel form of Fe3O4 nanoparticles (JCPDS Card No. 01-1111) and ZIF67@ZIF8 (JCPDS Card No. 00-062-1030). The results show that the crystal anatomy stable tetrahedral magnetic nanoparticles were maintained after surface modification. With the appearance of the peaks related to ZIF67 and ZIF8, based on the results and the structure of Fe3O4@ZIF67@ZIF8, the crystal structure of the catalyst has been confirmed59,60.

The XRD pattern of Fe3O4 (a) and Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni (b).

Using thermal gravimetric analysis, the changes in the weight of the catalyst were recorded due to temperature changes (Fig. 6). Furthermore, it was presented as two graphs (TGA-DTG) as output. The results of the TGA analysis were reported as a curve of weight changes according to temperature changes (25–800 °C). The results obtained from these curves were expressed in two forms of the initial report. At each stage of the weight change, the initial and final temperatures were determined as the starting and ending temperatures of the thermal event. The amount of weight loss in each event was expressed as a percentage. The first 10% of the weight loss is due to the removal of absorbed organic solvents below 200 °C. 10% mass reduction before 200 degrees indicates the thermal stability of the synthesized catalyst up to 200 degrees. The weight loss (20%) between 200 and 600 degrees is related to the removal of organic substances and the removal of bonds Fe3O4@Co/Zn-ZIFs and nickel complexes. The last weight loss at 605.47 °C can be related to catalytic decomposition and the change of crystal phase probability. Therefore, it can be concluded that organic agents cover the sample’s surface, and a 31% reduction in the weight of the sample up to a temperature of 600 degrees indicates the appropriate temperature resistance that the sample shows.

Thermal gravimetric analysis of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

A method capable of absorption was considered to determine the porosity of the catalyst. The volumetric method was used to measure the adsorption isotherm. The results of this measurement have been announced as a graph showing the relationship between the amount of absorption in terms of pressure at constant temperature. According to the type of graph, the absorption isotherm is type IV (Fig. 7). There is a relationship between the sample and the porosity of the firm and mesoporous adsorbent at (P/Po = 0.02–0.99) with H2-type hysteresis loops. The pore volume (0.1035 cm3.g−1), the surface area (40.411 m2 g−1), and the related pore diameter (10.247 nm) of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni catalyst were obtained from the by BET/BJH analyses. Today, due to the widespread use of porous materials and incredibly porous nanostructures, it has become a necessary tool in the field of catalysts. Due to the substantial porosity in the zeolite imidazolate frameworks and the addition of nanoparticles, the porosity and surface area in the catalyst increased. The corresponding pore diameter distribution in nanometer wavelength was determined, which confirms the nanoporous structure61.

Represents the nitrogen adsorption and desorption isotherms of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

A vibrating sample magnetometer was utilized to estimate the magnetic attributes of the catalyst. The VSM device works based on Faraday’s law of induction. This law states that a change in the magnetic field creates an electric field. By measuring the induced electric field, it is possible to acquire information about the changes in the magnetic field. In the diagram presented, the horizontal axis H is Oe with an Oersted unit, and the vertical axis M is an emu g−1 unit (Fig. 8). According to the results shown in the diagram, the magnetic saturation of Fe3o4 and Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni is evaluated. The value of Ms for Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni is 28 emu g−1. The Ms value for the catalyst was lower than that of the uncoated magnetic nanoparticles (22 emu g−1). This result seems to confirm Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni that surrounds the Fe3O4 surface. Therefore, this reduction is for the Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni coating on the surface of the primary magnetic nanoparticles for the catalyst. According to the magnetic saturation diagram, the hysteresis loop is not formed, and the sample is paramagnetic. In the presence of an external magnet, paramagnetic materials become magnetized on a macroscopic scale due to the alignment of the magnetization of the constituent particles with the direction of the external magnet.

The VSM data for Fe3O4 (a) and Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni (b).

EDS analysis confirmed the presence of Iron, Nickel, Nitrogen, Carbon, Oxygen, Silicon, Cobalt, and Zinc elements in the catalyst structure. In Fig. 9, the presence of elements C (16.34), Fe (25.56), N (3.95), Si (2.48), O (47.44), Ni (2.23), Co (2.30), and Zn (2.17) was confirmed. According to the FTIR analysis, the presence of peaks related to Ni and O bond and nickel on the catalyst’s surface using EDS and ICP-OES determined the characteristic value of nickel to be 3.2%. The catalyst preparation has been completed because nickel metal and other elements have a suitable weight% and proper dispersion.

EDX results of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

A scanning electron microscope was utilized to record the image using electrons, with high magnification and precision, and to determine the morphology and surface characteristics in nanometer dimensions. Images from the external surface of the sample were captured as two-dimensional images. The photos are presented in 195–501 nm and 5 μm scales. Using SEM results, morphological characteristics such as the structure of pores and cavities were investigated. According to Fig. 10a, the samples of newly synthesized magnetic nanoparticles have a uniform spherical shape with a particle size of 15 nm. SEM microscopic images of the catalyst sample (Fig. 10b-d( show the particle size of 23–50 nm in cloud state. For the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni, the spherical particles have increased in size and have a cloud state due to the modification of the surface of the magnetic nanoparticles by binding the ligand Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni to the Fe3O4 nanoparticles. By examining the SEM results of the sample surface with diverse magnifications and comparing them with the image Fe3O4, it can be observed that the contact surface of the catalyst increases with the increase in particle size. The attendance of holes and the spherical structure of the catalyst were confirmed.

SEM image of Fe3O4 (a) and Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni (b–d).

A TEM microscope with an electron beam was used to create high-resolution images at the nanoscale to observe the fine details, precise structure, and different catalyst components. The uniform size and morphology of the prepared substances, as shown in the obtained image (Fig. 11), demonstrate the significance of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni nanocatalyst procedure and its application to advance the fabrication of acridines. It has also shown that, within a reasonable period, the production with homogeneous morphology was successfully achieved. The particle size distribution of Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni shows that the average size is around 10–22 nm (Fig. 11). The similarity between the newly produced catalyst’s TEM images and those obtained after six recycling steps demonstrates the nanocatalyst structure’s durability and the particles’ efficient dispersion. Thus, different structures in Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni can be identified by their dark regions covered in brilliant outer layers and their cross-linking distances, which result in the appearance of holes.

TEM image of synthesized Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

From the structural studies of Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni, we’ve expected that this reagent could serve as a suitable catalyst for the acridine derivatives synthesized using aryl glyoxal, ammonium acetate, and cyclic 1,3-dione.

For this purpose, the model reaction was selected, and its various conditions were optimized under the impact of diverse empirical parameters containing various temperatures, the amount of newly synthesized catalyst, the solvent, and ultrasound. The results obtained are outlined in Table 2. The reaction was conducted without a catalyst at the reaction vessel for 23 h (entry 1, Table 2).

Carrying out the reaction with less than 10 mg (entries 2, Table 2) is associated with a long reaction time and a low yield percentage. Without using ultrasonic, the reaction time is longer than when using ultrasonic (entry 3, Table 2). The best conditions were created for preparing acridines when the reaction was performed in 3 min in the presence of the green solvent of water and at a temperature of 75 oC using 10 mg of catalyst and under ultrasonic radiation (power of 85 W, entry 4, Table 2). According to the reaction conditions and its repetition with a different amount of catalyst, the amount of catalyst more than 10 mg did not increase the reaction yield (entries 5 and 6, Table 2). The reaction with organic solvents has been checked using ethanol, toluene, and n-hexane, respectively (Table 2 and entries 7–9). The temperature increases to 110 degrees, indicating a high reaction time and a low percentage of the product produced for non-aqueous solvents.

By obtaining the right amount of catalyst, the time, temperature of the reaction, and the proper solvent to carry out the reaction, this reaction was carried out with different reagents of aryl glyoxal (electron-donating and electron-drawing substituents) with cyclic 1,3-dione and ammonium acetate. (Table 1)

The reaction mechanism of obtaining acridine derivatives using the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni catalyst is shown in Fig. 12. According to the proposed mechanism, firstly, the C = O of aryl glyoxal 1a–e activated with nanoparticles; after the Knoevenagel condensation among the enolic form of cyclic 1,3-dione 2a, b and activated aryl glyoxal 1a–e by dewateration occurs, and generate I. Afterward, the second enolic form of cyclic 1,3-dione 2a, b via Michael reaction was added to I to produce II. In the next stage, due to the reaction of II with CH3CO2NH4 (3), dehydration generates III. Afterward, the intramolecular nucleophilic assail of the –NH2 group on the carbonyl group with the effect of the catalyst causes the fabrication of a ring, and eventually, intramolecular cyclization with the water-removal was done, and the favorable products 4a–j and catalyst was obtained54,62.

Mechanistic pathway representing the synthesis of compounds 4a–j.

Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni was compared to other published literature to assess the system’s capabilities and efficiencies in the preparation of acridines (Table 3). According to the results obtained from this comparison, Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni catalyst performed better and was more suitable in the presence of green water solvent.

To investigate the role of Ni metal as an active metal in the catalyst structure, the model reaction was carried out in the presence of Fe3O4, Fe3O4@ZIF67@ZIF8, Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether, and Ni(NO3)2.6H2O as a catalyst (Table 4), with Examining the results of Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni containing Ni metal is the best catalyst for the preparation of acridine.

According to the analyses performed by SEM, TEM, FTIR, and XRD on the recovered catalyst (Fig. 13), the similarity of the structure and the results of the studies conducted after the recovery of the catalyst with the newly synthesized catalyst is of interest, and the recovery efficiency of the catalyst after six times of use is 80–82%.

The reuse of catalysts leads to a significant reduction in industrial waste production, thereby mitigating environmental pollutants. This process reduces costs related to preparing and using new catalysts and industrial waste disposal and minimizes the energy required for producing new catalysts. In essence, catalyst recycling is instrumental in cost reduction, environmental protection, and enhancing manufacturing productivity.We investigated the reusability of the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni for the reaction between ammonium acetate, and cyclic 1,3-dione, aryl glyoxal under the selected status. The catalyst was magnetically isolated quickly and easily after the completion of the reaction by a magnet and H2O-rinsed and dried at 75 °C. Finally, six new runs were conducted without losing catalyst properties using the awarded catalyst (Fig. 13). According to the information obtained from the recycling of the catalyst and product yield in the last cycle and the decrease in the percentage of the yield of the product compared to the previous stages, the reduction in the efficiency of the catalyst in the last stages of recycling can be attributed to the decrease in the amount of nickel on the surface of the catalyst due to washing. By examining the ICP analysis, the amount of nickel metal decreased to 2.5 after the recycling steps, which indicates the reduction of nickel metal. The result obtained from the TEM/SEM images, FT-IR spectrum, and XRD pattern showed that the recovered catalyst did not change and was similar to the structure of the new catalyst. (Fig. 14)

Recycling of Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

Image of analysis FTIR (a), XRD (b), SEM (c), and TEM (d) on the recovered Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

During the synthesis of acridines, a hot filtration was performed to evaluate the magnetic nanocatalyst’s heterogeneity and the Ni mixture’s leaching, as arylglyoxal, ammonium acetate, and cyclic 1,3-dione reacted. Furthermore, the product yield (33%) was obtained within 3 min of the reaction. The catalyst was then separated, filtered, and given a period to react. No further reactions were detected after the hot filtration. The fact that the reaction’s efficiency in this instance was 32% confirms minimal nickel leaching. (Fig. 15)

Hot filtration test for Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni.

In this research, we present a new heterogeneous catalyst Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni. XRD, TEM, TGA, FT-IR, BET, SEM-EDS, EDS, ICP, and VSM techniques have successfully evaluated the prepared catalyst. Acridine derivatives were synthesized in the presence of the catalyst to investigate the efficiency of the new catalyst in synthesizing organic compounds with high efficiency. Acridine derivatives were prepared through one-pot, multicomponent reactions of aryl glyoxals, ammonium acetate, and cyclic 1,3-dione in water under mild conditions using ultrasound radiation at 75 °C. Exceptional product yields of 98% are achieved within just 3 min. The Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Nicatalyst exhibits outstanding recyclability, maintaining its catalytic activity even after at least six reuse cycles. Energy sources were used to synthesize acridine derivatives, among which ultrasonic demonstrated the best efficiency. Indeed, ultrasonic irradiation demonstrating a synergistic effect with Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni nanocatalyst accelerates the reaction rate. According to the analyses performed by SEM, TEM, FTIR, and XRD on the recovered catalyst, the similarity of the structure and the results of the studies conducted after the recovery of the catalyst with the newly synthesized catalyst is of interest, and the recovery efficiency of the catalyst after six times of use is 80–82%. Some of the advantages of this catalyst are ease of use, high product yield with low catalyst consumption, environmental friendliness due to easy separation, and recycling several times without reducing efficiency. According to the investigations carried out in this research, the Fe3O4@Co/Zn-ZIFs@1-aza-18-crown-6-ether-Ni catalyst can be a suitable catalyst for obtaining acridines.

All data have been given in the article and supporting information.

Mcintyre, R. A. Common nano-materials and their use in real world applications. Sci. Prog. 95, 1–22 (2012).

Article CAS PubMed PubMed Central Google Scholar

Esmaili, S., Moosavi-Zare, A. R. & Khazaei, A. Nano-[Fe3O4@SiO2/N-propyl-1-(thiophen-2-yl)ethanimine][ZnCl2] as a nano magnetite Schiff base complex and heterogeneous catalyst for the synthesis of pyrimido[4,5-b]quinolones. RSC Adv. 12, 5386–5394 (2022).

Article ADS CAS PubMed PubMed Central Google Scholar

Sonawane, H. R., Deore, J. V. & Chavan, P. N. Reusable nano catalysed synthesis of heterocycles: An overview. ChemistrySelect 7, e202103900 (2022).

Article CAS Google Scholar

Mandari, V. & Devarai, S. K. Biodiesel production using homogeneous, heterogeneous, and enzyme catalysts via transesterification and esterification reactions: A critical review. Bioenergy Res. 15, 935–961 (2022).

Article CAS PubMed Google Scholar

Zhang, T. et al. Co@C nanorods as both magnetic stirring nanobars and magnetic recyclable nanocatalysts for microcatalytic reactions. Appl. Catal. B. 304, 120925 (2022).

Article CAS Google Scholar

Rossi, L. M., Costa, N. J., Silva, F. P. & Wojcieszak, R. Magnetic nanomaterials in catalysis: Advanced catalysts for magnetic separation and beyond. Green Chem. 16, 2906–2933 (2014).

Article CAS Google Scholar

Alshorifi, F. T., Ali, S. L. & Salama, R. S. Promotional synergistic effect of Cs–Au NPs on the performance of Cs–Au/MgFe2O4 catalysts in catalysis 3,4-dihydropyrimidin-2(1H)-ones and degradation of RhB Dye. J. Inorg. Organomet. Polym Mater. 32, 3765–3776 (2022).

Article CAS Google Scholar

Merroun, Y. et al. A new heterogeneous catalyst of triple superphosphate/titanium tetrachloride is used for the synthesis of 1,8-dioxooctahydroxanthenes and tetrahydrobenzo[b]pyrans. J. Mol. Struct. 1294, 136554 (2023).

Article CAS Google Scholar

Zirakjou, S., Dekamin, M. G., Valiey, E. & Dohendou, M. Tannic acid-poly ethyleneimine magnetic nanoparticles: An efficient and recyclable catalyst for green synthesis of 2,3-dihydroquinazole derivatives. Research Square. https://doi.org/10.21203/rs.3.rs-2524709/v1 (2023).

Article Google Scholar

Bilal, M., Zhao, Y., Rasheed, T. & Iqbal, H. M. Magnetic nanoparticles as versatile carriers for enzymes immobilization: A review. Int. J. Biol. Macromol. 120, 2530–2544 (2018).

Article CAS PubMed Google Scholar

Kim, Y. & Li, C. J. Perspectives on green synthesis and catalysis. Green. Synthesis Catal. 1, 1–11 (2020).

Article Google Scholar

Ying, S. et al. Green synthesis of nanoparticles: current developments and limitations. Environ. Technol. Innov. 26, 102336 (2022).

Article CAS Google Scholar

Furukawa, H., Cordova, K. E., O’Keeffe, M. & Yaghi, O. M. The chemistry and applications of metal-organic frameworks. Science 341, 1230444 (2013).

Article PubMed Google Scholar

Kitagawa, S. Metal–organic frameworks (MOFs). Chem. Soc. Rev. 43, 5415–5418 (2014).

Article PubMed Google Scholar

Ashworth, D. J. & Foster, J. A. Metal–organic framework nanosheets (MONs): A new dimension in materials chemistry. J. Mater. Chem. A. 6, 16292–16307 (2018).

Article CAS Google Scholar

Gaillac, R. et al. Liquid metal–organic frameworks. Nat. Mater. 16, 1149–1154 (2017).

Article ADS CAS PubMed Google Scholar

Zhang, H. et al. Green and rapid preparation of hierarchically porous metal–organic zeolites and simulation of their growth. J. Mater. Chem. A 7, 1022–1029 (2019).

Article CAS Google Scholar

Dias, E. M. & Petit, C. Towards the use of metal–organic frameworks for water reuse: A review of the recent advances in the field of organic pollutants removal and degradation and the next steps in the field. J. Mater. Chem. A 3, 22484–22506 (2015).

Article CAS Google Scholar

Yap, M. H., Fow, K. L. & Chen, G. Z. Synthesis and applications of MOF-derived porous nanostructures. Green. Energy Environ. 2, 218–245 (2017).

Article Google Scholar

Fu, L. et al. A novel AgNPs/MOF substrate-based SERS sensor for high-sensitive on-site detection of wheat gluten. Food Sci. Hum. Wellness. 13, 681–687 (2024).

Article CAS Google Scholar

Huang, H. et al. An in situ self-assembly template strategy for the preparation of hierarchical-pore metal-organic frameworks. Nat. Commun. 6, 8847 (2015).

Article ADS CAS PubMed Google Scholar

Ejsmont, A., Darvishzad, T., Słowik, G., Stelmachowski, P. & Goscianska, J. Cobalt-based MOF-derived carbon electrocatalysts with tunable architecture for enhanced oxygen evolution reaction. J. Colloid Interface Sci. 653, 1326–1338 (2024).

Article CAS PubMed Google Scholar

Adel, M., Salama, R. S., Adly, M. S., Ibrahim, A. A. & Ahmed A. I. Enhanced electrochemical performance of the electrodeposited nickel sulfide on ZIF-67@RGO composite for supercapacitor applications. J. Alloys Compd. 991, 174539 (2024).

Article CAS Google Scholar

Ding, M., Flaig, R. W., Jiang, H. L. & Yaghi, O. M. Carbon capture and conversion using metal–organic frameworks and MOF-based materials. Chem. Soc. Rev. 48, 2783–2828 (2019).

Article CAS PubMed Google Scholar

Cheng, W. et al. Lattice-strained metal–organic-framework arrays for bifunctional oxygen electrocatalysis. Nat. Energy. 4, 115–122 (2019).

Article ADS CAS Google Scholar

Wang, Q. & Astruc, D. State of the art and prospects in metal–organic framework (MOF)-based and MOF-derived nanocatalysis. Chem. Rev. 120, 1438–1511 (2019).

Article PubMed Google Scholar

Suo, Y., Zhang, Z., Zhang, Z. & Hu, G. Cobalt and nitrogen-doped carbon with enlarged pore size derived from ZIF-67 by a NaCl-assisted pyrolysis strategy towards oxygen reduction reaction. Ionics 27, 289–303 (2021).

Article CAS Google Scholar

Luo, H. et al. Self-catalyzed growth of Co–N–C nanobrushes for efficient rechargeable Zn–Air batteries. Small. 16, 2001171 (2020).

Article CAS Google Scholar

Veisi, H. et al. Recent advances in the application of magnetic nanocatalysts in multicomponent reactions. RSC Adv. 13, 20530–20556 (2023).

Article ADS CAS PubMed PubMed Central Google Scholar

Verma, D., Sharma, V., Okram, G. S. & Jain, S. Ultrasound-assisted high-yield multicomponent synthesis of triazolo[1,2-a]indazole-triones using silica-coated ZnO nanoparticles as a heterogeneous catalyst. Green Chem. 19, 5885–5899 (2017).

Article CAS Google Scholar

Merroun, Y. et al. Novel heterogeneous catalyst for green synthesis of imidazole derivatives: Environmental sustainability metrics assessment. Res. Chem. Intermed. 50, 3411–3433 (2024).

Article CAS Google Scholar

Srivastava, R. M., Neves Filho, R. A. W., de Silva, C. A. & Bortoluzzi, A. J. First ultrasound-mediated one-pot synthesis of N-substituted amides. Ultrason. Sonochem. 16, 737–742 (2009).

Article CAS PubMed Google Scholar

Tizhoosh, N. Y., Khataee, A., Hassandoost, R., Soltani, R. D. C. & Doustkhah, E. Ultrasound-engineered synthesis of WS2@CeO2 heterostructure for sonocatalytic degradation of tylosin. Ultrason. Sonochem. 67, 105114 (2020).

Article Google Scholar

Rad, T. S., Ansarian, Z., Khataee, A., Vahid, B. & Doustkhah, E. N-doped graphitic carbon as a nanoporous MOF-derived nanoarchitecture for the efficient sonocatalytic degradation process. Sep. Purif. Technol. 256, 117811 (2021).

Article Google Scholar

Xu, H., Zeiger, B. W. & Suslick, K. Sonochemical synthesis of nanomaterials. Chem. Soc. Rev. 42, 2555–2567 (2013).

Article CAS PubMed Google Scholar

Cravotto, G. & Cintas, P. Power ultrasound in organic synthesis: Moving cavitational chemistry from academia to innovative and large-scale applications. Chem. Soc. Rev. 35, 180–196 (2006).

Article CAS PubMed Google Scholar

Azim, M. K., Ahmed, W., Khan, I. A., Rao, N. A. & Khan, K. M. Identification of acridinyl hydrazides as potent aspartic protease inhibitors. Bioorg. Med. Chem. Lett. 18, 3011–3015 (2008).

Article CAS PubMed Google Scholar

Ungar, J. & Robinson, F. Investigation of the antibacterial and toxic action of certain acridine derivatives. J. Pharmacol. Exp. Ther. 80, 217–232 (1944).

CAS Google Scholar

Mohammadi Ziarani, G., Mousavi, S., Rahimifard, M. & Badiei, A. One-pot synthesis of benzo[c]acridine derivatives using SBA-Pr-SO3H as nano catalyst. J. Mex. Chem. Soc. 58, 168–172 (2014).

Google Scholar

Goni, L. K., Mazumder, J., Tripathy, M. A., Quraishi, M. A. & D. B. & Acridine and its derivatives: Synthesis, biological, and anticorrosion properties. Materials 15, 7560 (2022).

Article ADS CAS PubMed PubMed Central Google Scholar

Zhang, B., Li, X., Li, B., Gao, C. & Jiang, Y. Acridine and its derivatives: A patent review (2009–2013). Expert Opin. Ther. Pat. 24, 647–664 (2014).

Article CAS PubMed Google Scholar

Gensicka-Kowalewska, M., Cholewiński, G. & Dzierzbicka, K. Recent developments in the synthesis and biological activity of acridine/acridone analogues. RSC Adv. 7, 15776–15804 (2017).

Article ADS CAS Google Scholar

Geddes, C. D. Optical thin film polymeric sensors for the determination of aqueous chloride, bromide and iodide ions at high pH, based on the quenching of fluorescence of two acridinium dyes. Dyes Pigm. 45, 243–251 (2000).

Article CAS Google Scholar

Sabarees, G. et al. An overview of acridine analogs: Pharmacological significance and recent developments. Curr. Med. Chem. https://doi.org/10.2174/0109298673310987240610091522 (2024).

Article PubMed Google Scholar

Nguyen, P. N. et al. High-yield, fast, and green synthesis of acridine derivatives using a Co/C catalyst from rice husks with a microwave-assisted method. Reaction Chem. Eng. 9, 2034–2049 (2024).

Article CAS Google Scholar

Bayramoğlu, D. & Özgür, M. Ultrasound-assisted one-pot synthesis of 9-(substituted heteroaryl) acridinedione derivatives. Düzce Üniversitesi Bilim ve Teknoloji Dergisi 9, 1610–1620 (2021).

Article Google Scholar

Ebadi, A., Vadie, S. & Shojaei, S. Preparation of core/shell Fe3O4@SiO2⋅DIL as a magnetically separable catalyst for the synthesis of 1,8-dioxodecahydroacridine. ChemistrySelect 9, e202303166 (2024).

Shylesh, S., Schünemann, V. & Thiel, W. R. Magnetically separable nanocatalysts: Bridges between homogeneous and heterogeneous catalysis. Angew. Chem. Int. Ed. 49, 3428–3459 (2010).

Article CAS Google Scholar

Wang, D. & Astruc, D. Fast-growing field of magnetically recyclable nanocatalysts. Chem. Rev. 114, 6949–6985 (2014).

Article CAS PubMed Google Scholar

Payamifar, S. & Poursattar Marjani, A. A new β-cyclodextrin-based nickel as green and water-soluble supramolecular catalysts for aqueous Suzuki reaction. Sci. Rep. 13, 21279 (2023).

Article ADS CAS PubMed PubMed Central Google Scholar

Bibak, S. & Poursattar Marjani, A. Magnetically retrievable nanocatalyst Fe3O4@CPTMO@dithizone-Ni for the fabrication of 4H-benzo[h]chromenes under green medium. Sci. Rep. 13, 17894 (2023).

Article ADS CAS PubMed PubMed Central Google Scholar

Farajollahi, A. & Poursattar Marjani, A. Preparation of MWCNT/CoMn2O4 nanocomposite for effectual degradation of picric acid via peroxymonosulfate activation. Sci. Rep. 14, 11475 (2024).

Article ADS CAS PubMed Google Scholar

Poursattar Marjani, A., Asadzadeh, F. & Danandeh Asl, A. Novel core–shell magnetic nanoparticles@zeolitic imidazolate with glycerol-nickel for the synthesis of dihydropyrimidinones. Appl. Organomet. Chem. 37, e7260 (2023).

Article CAS Google Scholar

Abbaszadehghan, M., Poursattar Marjani, A., Bibak, S. & Sarreshtehdar Aslaheh, H. Nickel-asparagine complex fixed on a magnetic substrate as a precursor for preparing substituted acridines. Appl. Organomet. Chem. 37, e7247 (2023).

Article CAS Google Scholar

Bibak, S., Poursattar Marjani, A. & Sarreshtehdar Aslaheh, H. MCM-41 supported 2-aminothiophenol/Cu complex as a sustainable nanocatalyst for Suzuki coupling reaction. Sci. Rep. 14, 18070 (2024).

Article CAS PubMed PubMed Central Google Scholar

Song, X. et al. Synthesis of magnetic nanocomposite Fe3O4@ZIF-8@ZIF-67 and removal of tetracycline in water. Environ. Sci. Pollut. Res. 29, 35204–35216 (2022).

Article CAS Google Scholar

Ahangaran, F., Hassanzadeh, A. & Nouri, S. Surface modification of Fe3O4@SiO2 microsphere by silane coupling agent. Int. Nano Lett. 3, 1–5 (2013).

Article Google Scholar

Santos, G. T. et al. Hybrid nanocomposites of Fe3O4/SiO2-EDTA: holistic comparison of one-step and two-step modification methods. Powder Technol. 444, 120046 (2024).

Article CAS Google Scholar

Wu, Z. et al. Core-shell magnetic Fe3O4@Zn/Co-ZIFs to activate peroxymonosulfate for highly efficient degradation of carbamazepine. Appl. Catal. B. 277, 119136 (2020).

Article CAS Google Scholar

Li, N., Zhou, L., Jin, X., Owens, G. & Chen, Z. Simultaneous removal of tetracycline and oxytetracycline antibiotics from wastewater using a ZIF-8 metal organic-framework. J. Hazard. Mater. 366, 563–572 (2019).

Article CAS PubMed Google Scholar

Phan, A. et al. Synthesis, structure, and carbon dioxide capture properties of zeolitic imidazolate frameworks. Acc. Chem. Res. 43, 58–67 (2010).

Article CAS PubMed Google Scholar

Yü, S. J., Wu, S., Zhao, X. M. & Lü, C. W. Green and efficient synthesis of acridine-1,8-diones and hexahydroquinolines via a KH2PO4 catalyzed Hantzsch-type reaction in aqueous ethanol. Res. Chem. Intermed. 43, 3121–3130 (2017).

Article Google Scholar

Poursattar Marjani, A., Khalafy, J. & Mahmoodi, S. A simple one-pot synthesis of new 9-aroyl-3,4,6,7,9,10-hexahydro-1,8(2H,5H)-acridinediones. ARKIVOC 3, 262–270 (2016).

Poursattar Marjani, A., Khalafy, J., Chitan, M. & Mahmoodi, S. Microwave-assisted synthesis of acridine-1,8(2H,5H)-diones via a one-pot, three component reaction. Iran. J. Chem. Chem. Eng. 38, 205–211 (2019).

Google Scholar

Li, D. et al. Sustainable Ru-doped ni catalyst derived from hydrotalcite in propane reforming. Appl. Clay Sci. 43, 49–56 (2009).

Article Google Scholar

Mousavi, S. R., Nodeh, R., Foroumadi, A. & H. & Magnetically recoverable graphene-based nanoparticles for the one-pot synthesis of acridine derivatives under solvent-free conditions. Polycycl. Aromat. Compd. 41, 746–760 (2021).

Article CAS Google Scholar

Pamuk, H., Aday, B., Şen, F. & Kaya, M. Retracted article: pt NPs@GO as a highly efficient and reusable catalyst for one-pot synthesis of acridinedione derivatives. RSC Adv. 5, 49295–49300 (2015).

Article ADS CAS Google Scholar

Ghasemzadeh, M. A., Safaei-Ghomi, J. & Molaei, H. Fe3O4 nanoparticles: As an efficient, green and magnetically reusable catalyst for the one-pot synthesis of 1,8-dioxo-decahydroacridine derivatives under solvent-free conditions. Comptes Rendus Chim. 15, 969–974 (2012).

Article CAS Google Scholar

Maripi, S., Korupolu, R. B. & Madasu, S. B. Nano nickel-cobalt ferrite catalysed one-pot multi-component synthesis of xanthenediones and acridinediones. Green Sustain. Chem. 7, 70–84 (2017).

Article CAS Google Scholar

Download references

The authors would like to acknowledge the support from the Research Council of Urmia University.

Department of Organic Chemistry, Faculty of Chemistry, Urmia University, Urmia, Iran

Fatemeh Asadzadeh & Ahmad Poursattar Marjani

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

Fatemeh Asadzadeh: Data curation, Investigation, Methodology, Writing–original draft. Ahmad Poursattar Marjani: Project administration, Conceptualization, Investigation, Supervision, Writing–review & editing.

Correspondence to Ahmad Poursattar Marjani.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Below is the link to the electronic supplementary material.

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

Asadzadeh, F., Poursattar Marjani, A. Revolutionizing acridine synthesis: novel core-shell magnetic nanoparticles and Co-Zn zeolitic imidazolate framework with 1-aza-18-crown-6-ether-Ni catalysts. Sci Rep 14, 25739 (2024). https://doi.org/10.1038/s41598-024-75591-y

Download citation

Received: 16 July 2024

Accepted: 07 October 2024

Published: 28 October 2024

DOI: https://doi.org/10.1038/s41598-024-75591-y

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative