Next Article in Journal
Synthesis, Electrochemical and Photochemical Properties of Sulfanyl Porphyrazine with Ferrocenyl Substituents
Previous Article in Journal
Platycodonis Radix Alleviates LPS-Induced Lung Inflammation through Modulation of TRPA1 Channels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Performance of Magnetic Covalent Organic Framework Composites for Bisphenol A and Ibuprofen

Institute of Environmental Science, School of Chemistry and Materials Science, Ludong University, Yantai 264025, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 5214; https://doi.org/10.3390/molecules28135214
Submission received: 17 June 2023 / Revised: 3 July 2023 / Accepted: 3 July 2023 / Published: 5 July 2023
(This article belongs to the Section Analytical Chemistry)

Abstract

:
As typical environmental endocrine disruptors and nonsteroidal anti-inflammatory drugs, bisphenol A and ibuprofen in water supplies can cause great harm to the ecological environment and human health. In this study, magnetic covalent organic framework composites Fe3O4@COF-300 were synthesized by the hydrothermal method and used to remove bisphenol A and ibuprofen from water. Fe3O4@COF-300 could be rapidly separated from the matrix by external magnetic fields, and could selectively adsorb bisphenol A and ibuprofen in the presence of coexisting compounds such as phenol, Congo red, and amino black 10B. The removal efficiency of ibuprofen was 96.12–98.52% at pH in the range of 2–4 and that of bisphenol A was 92.18–95.62% at pH in the range of 2–10. The adsorption of bisphenol A and ibuprofen followed a pseudo-second-order kinetic and Langmuir model, and was a spontaneous endothermic process with the maximum adsorption amounts of 173.31 and 303.03 mg∙g−1, respectively. The material presented favorable stability and reusability, and the removal efficiency of bisphenol A and ibuprofen after 5 cycles was still over 92.15% and 89.29%, respectively. Therefore, the prepared composite Fe3O4@COF-300 exhibited good performance in the adsorption of bisphenol A and ibuprofen in water.

1. Introduction

With the rapid development of economies, the content of organic pollutants in the water environment is constantly increasing and gradually becoming an important global problem that threatens the ecological environment and human health. Among these, synthetic organic pollutants are the main source of organic pollutants. These compounds are of an anthropogenic origin from agriculture and industry and can be divided into several main categories according to their origin: pharmaceuticals and personal care products, steroid hormones, surfactants, and industrial chemicals and pesticides [1]. The synthetic organic pollutants generally undergo decomposition at relatively slow rates, and their removal efficiency using traditional sewage treatment methods is low. Thus, organic contaminants entering water bodies pose a serious threat to plant and animal health through bioconcentration [2].
Environmental endocrine disruptors and nonsteroidal anti-inflammatory drugs are two important synthetic organic contaminants derived primarily from industrial chemicals and pharmaceuticals. Environmental endocrine disruptors are substances that can interfere with the endocrine system of humans and other animals and increase the risk of endocrine disorders, leading to reproductive disorders, metabolic diseases, and various cancers. Bisphenol A is a typical environmental endocrine disruptor commonly used in the production of polycarbonate plastics, epoxy linings for canned food and beverage containers, dental sealants, and thermal paper [3]. At present, sewage treatment plants can only remove bisphenol A to a limited extent, so this compound is frequently detected in many water environments. Nonsteroidal anti-inflammatory drugs are widely used, and ibuprofen is one of the most prescribed nonsteroidal anti-inflammatory drugs in the contemporary era. Owing to its extensive global demand, its production reaches more than 30,000 tons per year [4]. The primary source of ibuprofen contamination in the environment is the excretion of nonmetabolic and metabolic drugs in the urine of humans and animals after medical treatment, followed by drug use and disposal in the aquaculture and pharmaceutical industries [5]. Upon reaching our bodies through contaminated water bodies, ibuprofen can induce serious side effects, such as vomiting, gastric ulcers, bleeding, indigestion, intestinal inflammation, mucosal damage, kidney problems, cardiovascular risks, and central nervous system problems [6].
Certain measures must be taken to effectively remove bisphenol A and ibuprofen from natural water bodies and wastewater to avoid their harmful effects on biological health. At present, the main removal methods are adsorption, photocatalysis, advanced oxidation, and biological techniques [7]. Among them, adsorption is widely adopted for the removal of bisphenol A and ibuprofen due to its simplicity and efficiency, as well as repeatability [8]. The performance of an adsorbent medium has an important effect on the economic feasibility of adsorption. Zeolite, activated carbon, silicate, metal–organic frameworks, metal–organic nanotube materials, covalent organic frameworks (COFs), and some other materials are widely used as adsorbents for contaminants owing to their special structure and physicochemical properties [9].
Among these adsorption materials, COFs possess the advantages of adjustable pore size, large specific surface area, diverse synthesis methods, and easy functional modification, and, thus, they are now increasingly used in the field of adsorption. COFs composed of layered stacking of two-dimensional polymers have a large number of ordered pore columns. These pore columns are ideal channels for ion migration and diffusion, which can greatly promote the adsorption of pollutants [10]. The COFs with imine bonds have excellent stability in water bodies, can be used for acid and alkali wastewater treatment, and have the characteristic of preventing secondary pollution [11]. Compared with the simple ones, COFs composited with other materials demonstrate additional physical and chemical properties and are thus extremely popular. For example, a new-generation COF material named SNW–1 was composited with montmorillonite for the removal of anionic and cationic dyes from aqueous solutions [12]. TpBD was composited with graphene oxide for the extraction of bisphenol A [13]. TpPa–1 composited with bimetallic oxide MnFe2O4 was employed to remove UO22+ from aqueous solutions [14]. Especially, Fe3O4@COF material has many advantages, such as excellent adsorption capacity, significant superparamagnetism, stability, and biocompatibility. These characteristics of Fe3O4@COF enable rapid separation of materials and matrix in the presence of an external magnetic field, thereby solving the problems of other non-magnetic COFs/composite materials that are difficult to separate from water and difficult to reuse.
In this work, COF-300 was used as substrate due to its mild synthesis conditions, short time consumption, and strong adsorption capacity. The new Fe3O4@COF-300 composite was prepared and applied for the adsorption of bisphenol A and ibuprofen in water. This study aimed: (1) To prepare and characterize the Fe3O4@COFs composites; (2) To analyze the effects of solution pH and ionic strength on the adsorption performance of Fe3O4@COF-300; (3) To explore the adsorption isotherms, adsorption kinetics, thermodynamics, and adsorption mechanism of bisphenol A and ibuprofen by Fe3O4@COFs.

2. Results and Discussion

2.1. Characterization of Fe3O4@COF-300

The surface morphology of Fe3O4, COF-300, and Fe3O4@COF-300 materials was investigated using scanning electron microscopy (SEM). As shown in Figure 1a,b, Fe3O4 is a spherical particle with a diameter of approximately 300 nm, and COF-300 is an oval particle, which is consistent with previous literature [15]. Figure 1c demonstrates that many white spherical bright spots are loaded on the surface of COF-300, indicating that Fe3O4 and COF-300 are successfully compounded.
Figure 2 shows the X-ray diffraction (XRD) patterns of the Fe3O4, COF-300, and Fe3O4@COF-300 samples. A single strong diffraction peak at around 2θ = 8.2° was observed for COF-300, indicating that the as-prepared COF-300 possessed high crystallinity. This result was in accordance with that previously reported [16]. Compared with the XRD pattern of COF-300, the main diffraction peaks of Fe3O4@COF-300 appeared in addition to 2θ = 8.6°, another 3 new diffraction peaks located at 2θ = 30.3°, 35.3°, and 42.9° corresponding to the (220), (311), and (400) planes of Fe3O4, respectively [17]. This result indicates the successful preparation of the composites.
The chemical composition and elemental bonding structure of Fe3O4@COF-300 were explored via X-ray photoelectron spectra (XPS). Figure 3a presents the full-scan XPS spectrograms of Fe3O4@COF-300, COF-300, and Fe3O4. The main constituent elements of Fe3O4@COF-300 include carbon, iron, oxygen, nitrogen, and silicon; those of COF-300 include carbon, oxygen, and nitrogen; those of Fe3O4 include iron and oxygen. Figure 3b–f illustrate the high-resolution XPS spectrograms of C1s, Fe2p, O1s, N1s, and Si2p. Fe3O4@COF-300 has a C1s photoelectron peak at 284.8 eV, corresponding to the C=C bond [18]. The high-resolution Fe2p spectrogram of Fe3O4@COF-300 shows 2 peaks with binding energies of 711.0 and 725.1 eV attributed to the 2p3/2 and 2p1/2 peaks of Fe (III), respectively [19]. An energy of 399.3 eV of Fe3O4@COF-300 can be assigned to C=N–C bonds, while an energy of 104.0 eV can be attributed to Si–O bonds [20,21]. Fe3O4 exhibits 2 peaks at 529.9 and 531.5 eV in O1s, attributed to the Fe–O group in Fe3O4 and OH in water, respectively. However, the peak in Fe3O4@COF-300 moved to a higher banding energy of 534.3 eV. The reason may be that the existence of the C=O bond in COF-300 could shift the OH in water to the direction of high binding energy [22]. Comparison of high-resolution XPS spectrograms showed that the binding energies of N1s (399.3 eV), O1s (534.3 eV), and Fe2p (711.0 and 725.1 eV) of Fe3O4@COF-300 are slightly higher than those of COF-300 (N1s: 399.2, O1s: 531.1 eV) and Fe3O4 (O1s: 529.9 and 531.5 eV, Fe2p: 710.3 and 724.1 eV). These results can be attributed to the interaction of Fe3O4 with COF-300 that resulted in the inner shift of N1s, O1s, and Fe2p orbits. XPS findings showed that Fe3O4 and COF-300 are combined by chemical bonds in the prepared composites rather than being simply physically mixed.
Thermogravimetric analysis (TGA) of Fe3O4, COF-300, and Fe3O4@COF-300 was carried out under nitrogen atmosphere with heating from 25 °C to 800 °C at a heating rate of 10 °C/min. As shown in Figure 4, the initial decrease in Fe3O4 was mainly due to the volatilization of water on its surface. Meanwhile, the mass loss of COF-300 was 10.0% as the temperature rose to 450 °C. As the temperature continued to rise to 600 °C, the mass loss of COF-300 reached 47.2%. This is mainly due to the fracture of the imine bond and the disintegration of the diamond network structure in the COF-300 material. When the temperature is higher than 600 °C, the mass loss of COF-300 tends to be stable, indicating that the material has been completely decomposed. The TGA curve of Fe3O4@COF-300 showed a consistent trend with that of COF-300 and lost less than 10% of its weight at 480 °C, indicating that Fe3O4@COF-300 has good stability.
Fourier transform infrared (FTIR) characterization of Fe3O4, COF-300, and Fe3O4@COF-300 composites is shown in Figure 5. The main absorption peak of Fe3O4 is located at 554 cm−1; those of COF-300 are at 1697, 1619, 1493, and 1173 cm−1; and those of Fe3O4@COF-300 are at 1697, 1619, 1493, 1173, 1088, 839, and 554 cm−1. The absorption peak at 1697 cm−1 is caused by the stretching vibration of the C=O group, and that at 1619 cm−1 is caused by the tensile vibration of the imine C=N formed in COF-300 and Fe3O4@COF-300. The absorption peak at 1493 cm−1 is caused by the stretching vibration of the aromatic ring, which is consistent with the conclusion that the molecular structure of COF-300 contains an aromatic ring [23]. The new material Fe3O4@COF-300 is also absorbed at 1493 cm−1, indicating that COF-300 does not destroy the aromatic structure of COF-300 after recombining with Fe3O4. This finding is consistent with the XPS characterization results. The absorption peaks at 1173 and 1088 cm−1 correspond to the C–C telescopic vibration and Si–O–C stretching, respectively, and that at 839 cm−1 is an aromatic ring extended by tetraphenylmethane [16,21]. Fe3O4@COF-300 exhibits a significant absorption peak at 554 cm−1, which is mainly caused by the telescopic vibration of the Fe–O bond [24].

2.2. Adsorption Experiment

2.2.1. Effect of pH on Removal Efficiency

The pH of the solution was adjusted between 2 and 10 to investigate the effect of pH on the adsorption properties of bisphenol A and ibuprofen by Fe3O4@COF-300. As shown in Figure 6, the removal efficiency of bisphenol A in the pH range of 2–10 is basically unchanged and can reach more than 92.18%. For ibuprofen, when the pH was 2–4, Fe3O4@COF-300 showed good adsorption performance (≥96.12%) with a maximum removal efficiency of 98.52%. However, with the increase in pH, the removal efficiency of ibuprofen decreased continuously until it was close to zero. The reason may be that when the solution pH was higher than the acidity coefficient (pKa = 4.91) of ibuprofen, both ibuprofen and Fe3O4@COF-300 were negatively charged (Figure S1). At such pH conditions, electrostatic repulsion occurred between the ibuprofen and the adsorption material, and the phenomenon increased with the increase in pH, thus resulting in a decrease in ibuprofen adsorption. When the solution pH was lower than the pKa of ibuprofen, the surface of Fe3O4@COF-300 remained negatively charged (Figure S1), whereas ibuprofen was a charge-neutral molecule in water. However, the dominant dispersive interactions explained the satisfactory adsorption capacity [25]. Therefore, the adsorption of bisphenol A by Fe3O4@COF-300 was largely unaffected by pH, but the suitable pH range for ibuprofen was between 2 and 4.

2.2.2. Adsorption Kinetics

The initial concentrations of bisphenol A and ibuprofen were set to 10 and 50 mg/L to study the adsorption kinetics. Adsorption was conducted at a certain time (5, 10, 20, 40, 60, 120, 180, 240, 300, 480, and 600 min) to determine the contaminant concentration. As shown in Figure 7, the adsorption capacity of bisphenol A and ibuprofen by Fe3O4@COF-300 increased rapidly within 60 min, reaching 69.4% and 98.2% of the equilibrium adsorption capacity, respectively. After a slow ascension, bisphenol A reached equilibrium at 120 min and ibuprofen reached equilibrium at 80 min.
The adsorption kinetic processes of bisphenol A and ibuprofen were investigated using a pseudo-second-order kinetic model (supplied in the supporting information). As shown in Figure S2 and Table S1, the adsorption processes of bisphenol A and ibuprofen were consistent with the pseudo-second-order kinetic, and there was a good linear correlation (R2 > 0.999).

2.2.3. Adsorption Isotherms

The adsorption isotherms were studied at 15, 25, and 35 °C with the initial concentrations of bisphenol A and ibuprofen at 5–100 and 5–75 mg/L, respectively (Figure 8). The adsorption capacity of Fe3O4@COF-300 rapidly increases with the raising of the concentrations of bisphenol A and ibuprofen. At 15 °C, the adsorption capacity increases from 20.82 to 80.74 mg∙g−1 as the concentration of bisphenol A increases from 5 to 100 mg∙L−1, and it increases from 20.83 to 255.58 mg∙g−1 with the concentration of ibuprofen increased from 5 to 75 mg∙L−1. This may be due to the fact that high concentration gradients can accelerate the diffusion of pollutants into the adsorbent [26]. Meanwhile, the adsorption process is favored by high temperatures. When the concentration of bisphenol A is 40 mg∙L−1, the adsorption capacity of Fe3O4@COF-300 is 130.72 mg∙g−1 at 35 °C, and it is 104.04 and 71.87 mg∙g−1 at 25 and 15 °C, respectively. For ibuprofen (50 mg∙L−1), the adsorption capacity is 248.40 mg∙g−1 at 35 °C, and 240.20 and 216.26 mg∙g−1 at 25 and 15 °C, respectively. This result indicates that adsorption is an endothermic process.
The adsorption of bisphenol A and ibuprofen by Fe3O4@COF-300 was further analyzed using Langmuir and Freundlich models (supplied in the supporting information). Table S2 shows that for the adsorption of bisphenol A and ibuprofen, the linear correlation coefficients (R2) of the Langmuir model are 0.9937–0.9986 and 0.9963–0.9991, respectively, and those of the Freundlich model are 0.8860–0.9738 and 0.8360–0.9646, respectively. Therefore, the adsorption is consistent with the Langmuir model. Meanwhile, the maximum adsorption of bisphenol A and ibuprofen by Fe3O4@COF-300 is 173.31 and 303.03 mg·g–1, respectively, which is basically higher than the adsorption of other materials in the literature (Table 1). This finding shows that Fe3O4@COF-300 is of great application potential in environmental pollution control and water purification.
The thermodynamic parameters (supplied in the supporting information) of the adsorption’s progress are shown in Table 2. The adsorption capacity and Gibbs free energy change (ΔGθ) both decrease with the increase in temperature, indicating that the adsorption is an endothermic process. This result is consistent with that of the adsorption isotherm, and the positive values of enthalpy change (ΔHθ) and entropy change (ΔSθ) further confirm this conclusion. Finally, the negative value of ΔGθ indicates that the adsorption of bisphenol A and ibuprofen by Fe3O4@COF-300 is a spontaneous process.

2.2.4. Effect of Ionic Strength

The effect of ionic strength on the removal efficiency of bisphenol A and ibuprofen was analyzed by adding NaCl to control the salt concentration from 0% to 35% (w/v). Figure 9 shows that the removal efficiency of bisphenol A and ibuprofen kept stable with the increase in ionic strength. Therefore, the adsorption capacity of Fe3O4@COF-300 is unaffected by the ionic strength.

2.3. Selective Adsorption

The selective adsorption properties of Fe3O4@COF-300 for bisphenol A and ibuprofen were investigated by using other coexisting pollutants commonly found in water bodies, such as phenol, Congo red, and amino black 10B. The concentrations of the coexisting pollutants were 50 mg·L−1, and pH was adjusted to 4.0 with HCl (0.1 mol·L−1). As shown in Figure 10, the adsorption of bisphenol A and ibuprofen can reach 116.68 and 240.19 mg·g−1, respectively, which is higher than that of phenol (18.12 mg·g−1), Congo red (4.59 mg·g−1), and amino black 10B (2.89 mg·g−1).
Further calculations indicate that the selectivity coefficients (supplied in the supporting information) of bisphenol A, ibuprofen, phenol, Congo red, and amino black 10B are 2.26, 93.14, 0.11, 0.01, and 0.03, respectively. These results demonstrate that Fe3O4@COF-300 exhibits high adsorption selectivity for bisphenol A and ibuprofen.

2.4. Environmental Implications

The adsorption behavior of Fe3O4@COF-300 on bisphenol A and ibuprofen in actual water was investigated using lake and tap water samples. The concentrations of bisphenol A and ibuprofen added to the actual water were 5–100 and 5–75 mg·L−1, respectively (Figure S3 and Table S3). Similar to that in pure water, the adsorption of bisphenol A and ibuprofen still conformed to the Langmuir model in lake and tap water samples. The maximum adsorption capacity of bisphenol A in lake water and tap water slightly increased by 7.47 (5.76%) and 3.46 mg·g–1 (2.67%), and that of ibuprofen decreased by 5.71 (2.13%) and 2.61 mg·g–1 (0.97%), respectively, which was less variable than that in pure water. This finding implies that Fe3O4@COF-300 has certain anti-interference characteristics and can be applied in the removal of bisphenol A and ibuprofen in actual water bodies.

2.5. Regeneration Property of Fe3O4@COF-300

The regeneration property of an adsorbent is a key factor in its practical application. Hence, cyclic experiments were carried out to investigate the regeneration property of Fe3O4@COF-300. Here, bisphenol A (10 mg·L−1) and ibuprofen (50 mg·L−1) were adsorbed by Fe3O4@COF-300 (200 mg·L−1), then the used Fe3O4@COF-300 was collected and washed 3 times with a mixture of ethanol and water (7:3, v/v), and continuously used for the removal. As shown in Figure S4, the removal efficiency of bisphenol A and ibuprofen decreased slightly with the increase in cycles. The removal efficiency of bisphenol A was 96.16% after the first cycle, and slightly decreased to 94.46%, 93.19%, and 93.02%, respectively, after the second, third, and fourth cycles. The removal efficiency could still be maintained at 92.15% after the fifth cycle. For ibuprofen, the removal efficiency was 93.37%, 92.18%, 91.24%, 89.07%, and 89.29% for the 5 cycles, respectively. Therefore, Fe3O4@COF-300 has satisfactory stability and regeneration property.

2.6. Adsorption Mechanism

The adsorption mechanisms of COFs for organic pollutants mainly include electrostatic and non-electrostatic interactions. Therefore, the adsorption mechanism of Fe3O4@COF-300 for bisphenol A and ibuprofen is summarized as follows.
When the solution pH is lower than the pKa of ibuprofen, ibuprofen exist under their neutral form and Fe3O4@COF-300 is negatively charged (Figure S1). Hence ibuprofen can be adsorbed by Fe3O4@COF-300 through dispersive interactions (non-electrostatic interactions). However, when the solution pH is higher than the pKa of ibuprofen, both ibuprofen and Fe3O4@COF-300 are negatively charged (Figure S1), and electrostatic repulsion (electrostatic interactions) occurs between them [25].
Bisphenol A is minimally affected by the solution pH, so the main adsorption mechanism is non-electrostatic. (1) As indicated by the FTIR results, Fe3O4@COF-300 has an aromatic structure, while bisphenol A also contains aromatic structures as well as abundant π electrons. Due to the π-π bond interactions between aromatic compounds, bisphenol A can undergo π-π interactions with the rich conjugated π domains in Fe3O4@COF-300 [37]. (2) Both Fe3O4@COF-300 and bisphenol A are hydrophobic, hence, hydrophobic interactions may also occur between them [38]. (3) Fe3O4@COF-300 composites are rich in N-containing functional groups (C=N−C) that may form hydrogen bonds with hydroxyl groups in bisphenol A, thus increasing the intermolecular forces and facilitating the adsorption [39].

3. Materials and Methods

3.1. Chemicals and Reagents

Bisphenol A, ibuprofen, phenol, Congo red, and amino black 10B were purchased from J&K Scientific Co., Ltd., Beijing, China. Tetrakis (4-aminophenyl) methane was purchased from Maclean Biochemical Technology Co. Ltd., China. All other reagents were of reagent grade and obtained from Bodi Chemical Industry Co., Ltd., Chongqing, China.

3.2. Preparation of Fe3O4@COF-300

Crushed FeCl3·6H2O (1.39 g) was dissolved in 75 mL of ethylene glycol. Crushed sodium acetate solids (3.56 g) were then added to the solution, sealed with fresh-keeping film, and continuously blended with a magnetic stirrer at room temperature for 1 h to form a uniform mixed solution. The mixed solution was then sonicated for 15 min to dissolve all the solids, transferred to a Teflon-lined stainless autoclave and heated at 200 °C for 8 h. The product was washed with ethanol 3 times after cooling down and maintained at 60 °C for 6 h. The final Fe3O4 product was obtained.
In order to prevent the reaction between Fe3O4 and acetic acid, as well as to prevent the oxidation and aggregation of Fe3O4, Fe3O4 was first compounded with SiO2 (Fe3O4@SiO2). The prepared Fe3O4 (100 mg) dispersed in ethanol (150 mL), ammonia (3 mL) and deionized water (47 mL) was added sequentially to the suspension, and the mixture (ethanol:deionized water:ammonia, 75.0%:23.5%:1.5%, v:v:v) was sonicated for 40 min. Ethyl orthosilicate (0.6 mL) was then added at room temperature, stirring was continued for 8 h. Finally, the product was collected by magnet, washed several times with ethanol and dried under vacuum at 60 °C.
Fe3O4@SiO2 (0.1 g), 1,4-phthalaldehyde (0.25 g), tetrakis (4-aminophenyl) methane (0.4 g), 1,4-dioxane (25 mL) and acetic acid solution (4 mL, 3 mol/L) were mixed. After sonication, the above mixture was transferred to a Teflon-lined autoclave and then heated at 120 °C for 72 h. Finally, the composite was washed alternately with 1,4-dioxane and tetrahydrofuran 3 times, and Fe3O4@COF-300 was obtained after drying at 60 °C.

3.3. Adsorption Experiment

Batch adsorption experiments were performed in a glass reactor with a volume of 25 mL. The volume of the reaction solution was 15 mL with a certain amount of bisphenol A and ibuprofen and 0.2 g·L−1 of Fe3O4@COF-300. The pH was adjusted with HCl and NaOH (0.1 mol/L). The mixture was placed in a water bath shaker at 25 °C for 12 h to completely balance the reaction. The reaction solution was placed on a strong magnet for 5 min after adsorption, and 1 mL of the supernatant was then extracted for contaminant concentration analysis. All the experiments were conducted twice to ensure the reproducibility and reliability of the results.

3.4. Analysis Methods

SEM images were acquired using a SUPPA 55 scanning electron microscope (Zeiss, Germany). XRD patterns were conducted with a D/max-r8 diffractometer (Rigaku, Akishima, Japan). XPS were obtained by an ESCALAB Xi+ system (Thermo Fisher Scientific, Waltham, MA, USA). FTIR spectra were recorded by Mangna–550 (Thermo Nicolet Corporation, Edina, MN, USA). TGA curves were obtained using an SDT Q600 thermogravimeter at 0–800 °C (TA Instruments, New Castle, DE, USA). Zeta potential analysis was measured by a potential analyzer (Malvern Zetasizer Nano ZS90, Marvin, UK). Bisphenol A, ibuprofen, and phenol concentrations were determined by an UltiMate3000 high-performance liquid chromatograph (HPLC; Thermo Fisher Scientific, Waltham, MA, USA). Congo red and amino black 10B concentrations were determined by a T6 UV-vis spectrophotometer (Beijing Purkinje, Beijing, China).

3.5. Preparation of Simulated Water Samples

A certain amount of bisphenol A and ibuprofen was added to the actual water sample to prepare the simulated water sample and explore the adsorption performance of Fe3O4@COF-300. All actual water samples were obtained from Yantai in Shandong Province. Tap water samples were taken from a laboratory of the Chemistry Building at Ludong University (Yantai, China). Lake water samples were acquired from the campus of Ludong University (Yantai, China). All samples were randomly collected and filtered to remove suspended particles and then used immediately within 24 h.

4. Conclusions

In this study, magnetic composites Fe3O4@COF-300 were successfully prepared and adopted for the adsorption of bisphenol A and ibuprofen in water. The composites could not only be rapidly separated from the matrix by external magnetic fields, but also showed excellent selective adsorption capacity for bisphenol A and ibuprofen. The adsorption process conformed to the pseudo-second-order kinetic and Langmuir model. Fe3O4@COF-300 presented satisfactory anti-interference characteristics, stability, and regeneration property and realized the adsorption of ibuprofen through electrostatic and non-electrostatic interactions, and bisphenol A through π–π interaction, hydrophobic interaction, and the formation of hydrogen bond. Fe3O4@COF-300 could be applied for the efficient removal of bisphenol A and ibuprofen in actual water bodies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28135214/s1, Figure S1: Zeta potential analysis of Fe3O4@COF-300; Figure S2: Adsorption kinetics of bisphenol A and ibuprofen by Fe3O4@COF-300; Figure S3: Adsorption isotherms of bisphenol A and ibuprofen in actual water samples; Figure S4: Regeneration property of Fe3O4@COF-300; Table S1: Pseudo-second-order reaction kinetics fitting of bisphenol A and ibuprofen removal with Fe3O4@ COF-300; Table S2: Parameters of Langmuir and Freundlich models for bisphenol A and ibuprofen with Fe3O4@COF-300; Table S3: Langmuir and Freundlich isotherms parameters of bisphenol A and ibuprofen in actual water samples.

Author Contributions

Conceptualization, B.Z. and J.L.; methodology, B.Z., J.L., Y.N. and Y.T.; investigation, Y.T., H.Z. and X.G.; resources, J.L.; writing—original draft preparation, Y.T.; writing—review and editing, B.Z., Y.N., H.Z., X.G. and J.L.; supervision, B.Z. and J.L.; project administration, B.Z., J.L. and Y.N.; funding acquisition, B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Shandong Province, grant number ZR2021MB044.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Acknowledgments

The authors gratefully acknowledge financial support from Natural Science Foundation of Shandong Province.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of all compounds are available from the authors.

References

  1. Bandura, L.; Białoszewska, M.; Malinowski, S.; Franus, W. Adsorptive performance of fly ash-derived zeolite modified by β-cyclodextrin for ibuprofen, bisphenol A and caffeine removal from aqueous solutions-equilibrium and kinetic study. Appl. Surf. Sci. 2021, 562, 150160–150170. [Google Scholar] [CrossRef]
  2. Alharbi, O.M.L.; Basheer, A.A.; Khattab, R.A.; Ali, I. Health and environmental effects of persistent organic pollutants. J. Mol. Liq. 2018, 263, 442–453. [Google Scholar] [CrossRef]
  3. Seachrist, D.D.; Bonk, K.W.; Ho, S.-M.; Prins, G.S.; Soto, A.M.; Keri, R.A. A review of the carcinogenic potential of bisphenol A. Reprod. Toxicol 2016, 59, 167–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tanveer, M.; Guyer, G.T.; Abbas, G. Photocatalytic degradation of ibuprofen in water using TiO2 and ZnO under artificial UV and solar irradiation. Water Environ. Res. 2019, 91, 822–829. [Google Scholar] [CrossRef] [PubMed]
  5. Brillas, E. A critical review on ibuprofen removal from synthetic waters, natural waters, and real wastewaters by advanced oxidation processes. Chemosphere 2022, 286, 131849–131876. [Google Scholar] [CrossRef]
  6. Ali, I.; Al-Othman, Z.A.; Alwarthan, A. Synthesis of composite iron nano adsorbent and removal of ibuprofen drug residue from water. J. Mol. Liq. 2016, 219, 858–864. [Google Scholar] [CrossRef]
  7. Bilgin Simsek, E.; Kilic, B.; Asgin, M.; Akan, A. Graphene oxide based heterojunction TiO2–ZnO catalysts with outstanding photocatalytic performance for bisphenol-A, ibuprofen and flurbiprofen. J. Ind. Eng. Chem. 2018, 59, 115–126. [Google Scholar] [CrossRef]
  8. Oba, S.N.; Ighalo, J.O.; Aniagor, C.O.; Igwegbe, C.A. Removal of ibuprofen from aqueous media by adsorption: A comprehensive review. Sci. Total Environ. 2021, 780, 146608–146630. [Google Scholar] [CrossRef]
  9. Kharrazi, S.M.; Soleimani, M.; Jokar, M.; Richards, T.; Pettersson, A.; Mirghaffari, N. Pretreatment of lignocellulosic waste as a precursor for synthesis of high porous activated carbon and its application for Pb (II) and Cr (VI) adsorption from aqueous solutions. Int. J. Biol. Macromol. 2021, 180, 299–310. [Google Scholar] [CrossRef]
  10. Huang, W.; Xu, H.; Liu, X.; Wang, L.; Li, S.; Ji, L.; Qu, Z.; Yan, N. Surface protection method for the magnetic core using covalent organic framework shells and its application in As(III) depth removal from acid wastewater. J. Environ. Sci. 2022, 115, 1–9. [Google Scholar] [CrossRef]
  11. Wang, L.; Xu, H.; Qiu, Y.; Liu, X.; Huang, W.; Yan, N.; Qu, Z. Utilization of Ag nanoparticles anchored in covalent organic frameworks for mercury removal from acidic waste water. J. Hazard. Mater. 2020, 389, 151824. [Google Scholar] [CrossRef] [PubMed]
  12. Khataei, M.M.; Yamini, Y.; Asiabi, H.; Shamsayei, M. Covalent organic framework and montomorillonite nanocomposite as advanced adsorbent: Synthesis, characterization, and application in simultaneous adsorption of cationic and anionic dyes. J. Environ. Health Sci. Eng. 2020, 18, 1555–1567. [Google Scholar] [CrossRef] [PubMed]
  13. Gao, W.; Cheng, J.; Yuan, X.; Tian, Y. Covalent organic framework-graphene oxide composite: A superior adsorption material for solid phase microextraction of bisphenol A. Talanta 2021, 222, 121501–121507. [Google Scholar] [CrossRef] [PubMed]
  14. Zhong, X.; Lu, Z.; Liang, W.; Hu, B. Incorporating bimetal oxide MnFe2O4 onto covalent organic frameworks for the removal of UO22+ ion from aqueous solution. Appl. Surf. Sci. 2021, 556, 149581–149592. [Google Scholar] [CrossRef]
  15. Ma, T.; Kapustin, E.A.; Yin, S.X.; Liang, L.; Zhou, Z.; Niu, J.; Li, L.-H.; Wang, Y.; Su, J.; Li, J.; et al. Single-crystal x-ray diffraction structures of covalent organic frameworks. Science 2018, 361, 48–52. [Google Scholar] [CrossRef] [Green Version]
  16. Uribe-Romo, F.J.; Hunt, J.R.; Furukawa, H.; Klöck, C.; O’Keeffe, M.; Yaghi, O.M. A Crystalline Imine-Linked 3-D Porous Covalent Organic Framework. J. Am. Chem. Soc. 2009, 131, 4570–4571. [Google Scholar] [CrossRef]
  17. Moradi, S.E.; Haji Shabani, A.M.; Dadfarnia, S.; Emami, S. Effective removal of ciprofloxacin from aqueous solutions using magnetic metal–organic framework sorbents: Mechanisms, isotherms and kinetics. J. Iran. Chem. Soc. 2016, 13, 1617–1627. [Google Scholar] [CrossRef]
  18. Lin, X.; Wang, X.; Wang, J.; Yuan, Y.; Di, S.; Wang, Z.; Xu, H.; Zhao, H.; Qi, P.; Ding, W. Facile synthesis of a core-shell structured magnetic covalent organic framework for enrichment of organophosphorus pesticides in fruits. Anal. Chim. Acta 2020, 1101, 65–73. [Google Scholar] [CrossRef]
  19. Sajjadi, S.; Khataee, A.; Bagheri, N.; Kobya, M.; Şenocak, A.; Demirbas, E.; Karaoğlu, A.G. Degradation of diazinon pesticide using catalyzed persulfate with Fe3O4@MOF-2 nanocomposite under ultrasound irradiation. J. Ind. Eng. Chem. 2019, 77, 280–290. [Google Scholar] [CrossRef]
  20. Niu, X.; Lv, W.; Sun, Y.; Dai, H.; Chen, H.; Chen, X. In situ fabrication of 3D COF-300 in a capillary for separation of aromatic compounds by open-tubular capillary electrochromatography. Microchim. Acta 2020, 187, 233–241. [Google Scholar] [CrossRef]
  21. Chen, X.; Hu, Y.; Jiao, C.; Song, L. Preparation and thermal properties of a novel flame-retardant coating. Polym. Degrad. Stabil. 2007, 92, 1141–1150. [Google Scholar] [CrossRef]
  22. Yang, D.; Velamakanni, A.; Bozoklu, G.; Park, S.; Stoller, M.; Piner, R.D.; Stankovich, S.; Jung, I.; Field, D.A.; Ventrice, C.A., Jr.; et al. Chemical analysis of graphene oxide films after heat and chemical treatments by X-ray photoelectron and Micro-Raman spectroscopy. Carbon 2009, 47, 145–152. [Google Scholar] [CrossRef]
  23. Li, S.; Li, P.; Cai, D.; Shan, H.; Zhao, J.; Wang, Z.; Qin, P.; Tan, T. Boosting pervaporation performance by promoting organic permeability and simultaneously inhibiting water transport via blending PDMS with COF-300. J. Membr. Sci. 2019, 579, 141–150. [Google Scholar] [CrossRef]
  24. Azari, A.; Gharibi, H.; Kakavandi, B.; Ghanizadeh, G.; Javid, A.; Mahvi, A.H.; Sharafi, K.; Khosravia, T. Magnetic adsorption separation process: An alternative method of mercury extracting from aqueous solution using modified chitosan coated Fe3O4 nanocomposites. J. Chem. Technol. Biotechnol. 2017, 92, 188–200. [Google Scholar] [CrossRef]
  25. Guedidi, H.; Reinert, L.; Lévêque, J.-M.; Soneda, Y.; Bellakhal, N.; Duclaux, L. The effects of the surface oxidation of activated carbon, the solution pH and the temperature on adsorption of ibuprofen. Carbon 2013, 54, 432–443. [Google Scholar] [CrossRef]
  26. Javaheri, F.; Kheshti, Z.; Ghasemi, S.; Altaee, A. Enhancement of Cd2+ removal from aqueous solution by multifunctional mesoporous silica: Equilibrium isotherms and kinetics study. Sep. Purif. Technol. 2019, 224, 199–208. [Google Scholar] [CrossRef]
  27. Dehghani, M.H.; Karri, R.R.; Alimohammadi, M.; Nazmara, S.; Zarei, A.; Saeedi, Z. Insights into endocrine-disrupting Bisphenol-A adsorption from pharmaceutical effluent by chitosan immobilized nanoscale zero-valent iron nanoparticles. J. Mol. Liq. 2020, 311, 113317–113329. [Google Scholar] [CrossRef]
  28. Kuo, C.-Y. Comparison with as-grown and microwave modified carbon nanotubes to removal aqueous bisphenol A. Desalination 2009, 249, 976–982. [Google Scholar] [CrossRef]
  29. Chen, Z.H.; Liu, Z.; Hu, J.Q.; Cai, Q.W.; Li, X.Y.; Wang, W.; Faraj, Y.; Ju, X.J.; Xie, R.; Chu, L.Y. β-Cyclodextrin-modified graphene oxide membranes with large adsorption capacity and high flux for efficient removal of bisphenol A from water. J. Membr. Sci. 2020, 595, 117510–117517. [Google Scholar] [CrossRef]
  30. Liu, C.; Wu, P.; Zhu, Y.; Tran, L. Simultaneous adsorption of Cd2+ and BPA on mphoteric surfactant activated montmorillonite. Chemosphere 2016, 144, 1026–1032. [Google Scholar] [CrossRef] [PubMed]
  31. Usman, M.; Ahmed, A.; Ji, Z.; Yu, B.; Shen, Y.; Cong, H. Environmentally friendly fabrication of new β-Cyclodextrin/ZrO2 nanocomposite for simultaneous removal of Pb(II) and BPA from water. Sci. Total Environ. 2021, 784, 147207–147218. [Google Scholar] [CrossRef] [PubMed]
  32. Bhadra, B.N.; Ahmed, I.; Kim, S.; Jhung, S.H. Adsorptive removal of ibuprofen and diclofenac from water using metal-organic framework-derived porous carbon. Chem. Eng. J. 2017, 314, 50–58. [Google Scholar] [CrossRef]
  33. Fröhlich, A.C.; Foletto, E.L.; Dotto, G.L. Preparation and characterization of NiFe2O4/activated carbon composite as potential magnetic adsorbent for removal of ibuprofen and ketoprofen pharmaceuticals from aqueous solutions. J. Clean. Prod. 2019, 229, 828–837. [Google Scholar] [CrossRef]
  34. Zhang, G.; Li, S.; Shuang, C.; Mu, Y.; Li, A.; Tan, L. The effect of incorporating inorganic materials into quaternized polyacrylic polymer on its mechanical strength and adsorption behaviour for ibuprofen removal. Sci. Rep. 2020, 10, 5188–5198. [Google Scholar] [CrossRef] [Green Version]
  35. Phasuphan, W.; Praphairaksit, N.; Imyim, A. Removal of ibuprofen, diclofenac, and naproxen from water using chitosan-modified waste tire crumb rubber. J. Mol. Liq. 2019, 294, 111554–111561. [Google Scholar] [CrossRef]
  36. Kollarahithlu, S.C.; Balakrishnan, R.M. Adsorption of ibuprofen using cysteine-modified silane-coated magnetic nanomaterial. Environ. Sci. Pollut. Res. 2019, 26, 34117–34126. [Google Scholar] [CrossRef] [PubMed]
  37. Wei, H.; Deng, S.; Huang, Q.; Nie, Y.; Wang, B.; Huang, J.; Yu, G. Regenerable granular carbon nanotubes/alumina hybrid adsorbents for diclofenac sodium and carbamazepine removal from aqueous solution. Water Res. 2013, 47, 4139–4147. [Google Scholar] [CrossRef]
  38. Duo, H.; Tang, H.; Ma, J.; Lu, X.; Wang, L.; Liang, X. Iron-based metal–organic framework as an effective sorbent for the rapid and efficient removal of illegal dyes. New J. Chem. 2019, 43, 15351–15358. [Google Scholar] [CrossRef]
  39. Wang, F.; Zeng, Q.; Su, W.; Zhang, M.; Hou, L.; Wang, Z.L. Adsorption of Bisphenol A on Peanut Shell Biochars: The Effects of Surfactants. J. Chem. 2019, 2019, 2428505. [Google Scholar] [CrossRef]
Figure 1. SEM images of Fe3O4 (a), COF-300 (b), and Fe3O4@COF-300 (c).
Figure 1. SEM images of Fe3O4 (a), COF-300 (b), and Fe3O4@COF-300 (c).
Molecules 28 05214 g001
Figure 2. XRD patterns of Fe3O4, COF-300, and Fe3O4@COF-300.
Figure 2. XRD patterns of Fe3O4, COF-300, and Fe3O4@COF-300.
Molecules 28 05214 g002
Figure 3. XPS spectra of survey scan of Fe3O4@COF-300, COF-300, and Fe3O4 (a), C1s (b), Fe2p (c), O1s (d), N1s (e), and Si2p (f).
Figure 3. XPS spectra of survey scan of Fe3O4@COF-300, COF-300, and Fe3O4 (a), C1s (b), Fe2p (c), O1s (d), N1s (e), and Si2p (f).
Molecules 28 05214 g003
Figure 4. TGA curves of Fe3O4, COF-300, and Fe3O4@COF-300.
Figure 4. TGA curves of Fe3O4, COF-300, and Fe3O4@COF-300.
Molecules 28 05214 g004
Figure 5. Infrared spectra of Fe3O4, COF-300 and Fe3O4@COF-300.
Figure 5. Infrared spectra of Fe3O4, COF-300 and Fe3O4@COF-300.
Molecules 28 05214 g005
Figure 6. Effect of pH on bisphenol A and ibuprofen removal.
Figure 6. Effect of pH on bisphenol A and ibuprofen removal.
Molecules 28 05214 g006
Figure 7. Adsorption kinetics of bisphenol A and ibuprofen by Fe3O4@COF-300.
Figure 7. Adsorption kinetics of bisphenol A and ibuprofen by Fe3O4@COF-300.
Molecules 28 05214 g007
Figure 8. Adsorption isotherms of bisphenol A and ibuprofen.
Figure 8. Adsorption isotherms of bisphenol A and ibuprofen.
Molecules 28 05214 g008
Figure 9. Effect of ionic strength on bisphenol A and ibuprofen removal.
Figure 9. Effect of ionic strength on bisphenol A and ibuprofen removal.
Molecules 28 05214 g009
Figure 10. Selective adsorption capacities of Fe3O4@COF-300.
Figure 10. Selective adsorption capacities of Fe3O4@COF-300.
Molecules 28 05214 g010
Table 1. Adsorption capacity of bisphenol A and ibuprofen on reported sorbents.
Table 1. Adsorption capacity of bisphenol A and ibuprofen on reported sorbents.
AdsorbatesAdsorbentsQmax (mg·g–1)References
Bisphenol AnZVI–chitosan65.16[27]
modified carbon nanotubes69.93[28]
β-Cyclodextrin-modified graphene oxide membranes25.50[29]
modified montmorillonite (BS–Mt)80.77[30]
β-Cyclodextrin/ZrO2174.90[31]
IbuprofenZIF–8 derived porous carbon320.00[32]
NiFe2O4/AC261.35[33]
MAP 269.61[34]
chitosan modified rubber70.00[35]
cysteine-modified silane-coated
magnetic nanomaterial
138.10[36]
Table 2. Thermodynamic parameters of bisphenol A and ibuprofen.
Table 2. Thermodynamic parameters of bisphenol A and ibuprofen.
PollutantC0
(mg/L)
Ce (mg/L)ΔSθ
(J/mol/K)
ΔHθ
(kJ/mol)
ΔGθ (kJ/mol)
288 K298 K308 K288 K298 K308 K
Bisphenol A10081.6571.8664.99120.6234.69−0.07−1.27−2.48
Ibuprofen7523.9021.3915.6892.0720.96−24.43−25.35−26.28
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, B.; Tian, Y.; Gao, X.; Zheng, H.; Niu, Y.; Liu, J. Adsorption Performance of Magnetic Covalent Organic Framework Composites for Bisphenol A and Ibuprofen. Molecules 2023, 28, 5214. https://doi.org/10.3390/molecules28135214

AMA Style

Zhang B, Tian Y, Gao X, Zheng H, Niu Y, Liu J. Adsorption Performance of Magnetic Covalent Organic Framework Composites for Bisphenol A and Ibuprofen. Molecules. 2023; 28(13):5214. https://doi.org/10.3390/molecules28135214

Chicago/Turabian Style

Zhang, Beibei, Ye Tian, Xuezhen Gao, Hui Zheng, Yuzhong Niu, and Junshen Liu. 2023. "Adsorption Performance of Magnetic Covalent Organic Framework Composites for Bisphenol A and Ibuprofen" Molecules 28, no. 13: 5214. https://doi.org/10.3390/molecules28135214

Article Metrics

Back to TopTop