Next Article in Journal
Virulence Factors of Candida spp. and Host Immune Response Important in the Pathogenesis of Vulvovaginal Candidiasis
Next Article in Special Issue
Theory-Guided Design of a Method to Obtain Competitive Balance between U(VI) Adsorption and Swaying Zwitterion-Induced Fouling Resistance on Natural Hemp Fibers
Previous Article in Journal
Ameliorative Effects of Lactobacillus plantarum HAC01 Lysate on 3T3-L1 Adipocyte Differentiation via AMPK Activation and MAPK Inhibition
Previous Article in Special Issue
Template-Free Preparation of a Mesopore-Rich Hierarchically Porous Carbon Monolith from a Thermally Rearrangeable Polyurea Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Waste-to-Resource Strategy to Fabricate Functionalized MOFs Composite Material Based on Durian Shell Biomass Carbon Fiber and Fe3O4 for Highly Efficient and Recyclable Dye Adsorption

1
Co-Innovation Center for the Sustainable Forestry in Southern China, College of Biology and the Environment, Nanjing Forestry University, Nanjing 210037, China
2
Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensification Technology, School of Chemistry and Chemical Engineering, Guangxi University, Nanning 530004, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(11), 5900; https://doi.org/10.3390/ijms23115900
Submission received: 25 March 2022 / Revised: 18 May 2022 / Accepted: 23 May 2022 / Published: 24 May 2022
(This article belongs to the Special Issue Adsorption Materials and Adsorption Behavior)

Abstract

:
Recently, metal–organic frameworks (MOFs), which are porous inorganic–organic hybrid materials consisting of metal ions (clusters or secondary building units) and organic ligands through coordination bonds, have attracted wide attention because of their high surface area, huge ordered porosity, uniform structural cavities, and excellent thermal/chemical stability. In this work, durian shell biomass carbon fiber and Fe3O4 functionalized metal–organic framework composite material (durian shell fiber-Fe3O4-MOF, DFM) was synthesized and employed for the adsorption removal of methylene blue (MB) from wastewater. The morphology, structure, and chemical elements of the DFM material were characterized by scanning electron microscope (SEM), X-ray diffraction (XRD), transmission electron microscope (TEM), and X-ray photoelectron spectroscope (XPS) techniques. Adsorption conditions such as pH, adsorption time, and temperature were optimized. The adsorption isotherm and kinetics results show that the adsorption process of DFM material to MB is more in line with the Freundlich model and pseudo-second-order kinetic model. Using these models, the maximum adsorption capacity of 53.31 mg/g was obtained by calculation. In addition, DFM material could be easily reused through an external magnet and the removal rate of MB was still 80% after five adsorption cycles. The obtained results show that DFM composite material, as an economical, environmentally friendly, recyclable new adsorbent, can simply and effectively remove MB from wastewater.

1. Introduction

With the development of industry, the increase of organic pollutants in wastewater will have various adverse effects on the environment, so more effective treatment methods need to be explored [1,2]. Methylene blue (MB) is an organic dye commonly used in the textile, leather, printing, and dyeing industries. Because of its high chroma and large organic content, if it is directly discharged into the surroundings without treatment, it will endanger the physical and mental health of the human body, causing people to have rapid heartbeat, nausea, and vomiting and, in serious cases, causing cancer. [3]. I methylene blue enters water, it will lead to poor light transmittance and affect the growth of plants and animals in water [4]. Common water-treatment technologies include flocculation precipitation, membrane separation, photocatalysis, biological methods, and ion exchange [5,6,7,8,9,10,11,12,13,14]. Compared with several other treatment technologies, the adsorption method has the advantages of high treatment efficiency, simple operation, no by-products, and recyclability, making it one of the important methods in pollutant treatment [15]. Commonly used adsorbents include activated carbon, molecular sieves, silicone, and zeolite [16], however, these adsorbents are difficult to recycle and the effect of reuse is poor [17,18]. So, designing a new adsorbent with a simple method, easy separation, and good chemical/physical stability is important.
Biomass is a general term for organisms that come directly or indirectly from various plants, with the most extensive source being agricultural and forestry waste [19,20,21].The common biomass treatment methods include incineration and burial. These treatment methods not only waste resources, but also cause other environmental pollution problems [22,23]. Durian shell is a type of agricultural and forestry waste. The main components are cellulose, hemicellulose, and lignin. It is a good material for the preparation of biomass adsorbents [24]. However, most of the common adsorbents prepared from durian shells have the disadvantages of large particle size and difficulty in recycling after adsorption [25,26,27]. Moreover, Fe3O4 shows excellent separate characteristics due to its magnetic properties [28]. Therefore, it has been widely used in wastewater treatment in recent years [29,30,31]. In previous studies, Sun et al. [32] prepared Fe3O4@C submicron rods by hydrothermal carbonization to remove MB in aqueous solution, but the larger particle size affects the comparison area and reduces the adsorption efficiency of MB. Rsjput et al. [33] synthesized Fe3O4 nanoparticles by co-precipitation to adsorb Pb2+. However, Fe3O4 nanoparticles are easy to agglomerate and the maximum adsorption capacity is only 53.11 mg/g. Therefore, it is necessary to study and prepare an adsorbent that can prevent Fe3O4 nanoparticles from agglomerating and has good pollutant removal effect.
Metal–organic framework materials (MOFs) based on their huge porosity, large specific surface area, adjustable pore size, etc. [34,35], are very broad prospects for various applications [36]. In recent years, various MOF materials have been applied to wastewater treatment. Abdi et al. [37] reported zeolite imidazolate framework (ZIF-8) nanocomposites prepared with graphene oxide (GO) and carbon nanotubes (CNT) as substrates to remove malachite green (MG) from colored wastewater. Lin et al. [38] synthesized a zirconium-based MOF (UiO-66-NH2) to remove fluoride in wastewater and the crystal structure of UiO-66-NH2 remained intact during the adsorption process. Efome et al. [39] synthesized water-based metal organic framework particles (Zr-808) on the surface of polyacrylonitrile (PAN) nanofibers, and the maximum adsorption capacities for heavy metal ions (Cd2+ and Zn2+) in aqueous solution were 225.05 and 287.06 mg/g, respectively. Although the MOF material can effectively remove pollutants in water, it is difficult to separate and recover quickly at the end of adsorption, which limits its practical application in wastewater treatment. In our work, fibers were extracted from durian shell and Fe3O4 nanoparticles were loaded onto its surface, which not only realized the resource reuse of durian shell, but also effectively reduced the agglomeration of Fe3O4 nanoparticles. At the same time, the combination of Fe3O4 nanoparticles and MOF materials can effectively solve the problem that MOF materials are difficult to separate from aqueous solution. In this paper, the durian shell fiber–Fe3O4 MOF composite was synthesized by a step-by-step method to remove methylene blue (MB) from wastewater. The structures were characterized by SEM, TEM, XRD, and XPS techniques. The effects of adsorption time, initial concentration, temperature, and pH value on the adsorption process, and the adsorption kinetics, isotherm, and thermodynamics of MB were studied.

2. Results and Discussion

2.1. SEM and TEM Analysis

The morphology and structure of DF, DFM-10, DFM-20, and DFM-40 samples at different magnifications were observed by SEM and TEM and the result are shown in Figure 1, Figure 2 and Figure S1. The width of durian shell fiber was around 10 µm and the surface of durian shell fiber was relatively smooth (Figure S1). As can be seen from DF images, the particle size of Fe3O4 was uniform and the magnetic Fe3O4 nanoparticles were successfully loaded onto the surface of durian shell fibers. From SEM results of DFM-10, DFM-20, and DFM-40, the magnetic Fe3O4 nanoparticles were successfully wrapped with a layer of MOF shell. After 10 cycles, sample DFM-10 showed a rough surface. The thickness of the MOF shell gradually increased with the increase in cycle numbers. However, after 20 and 40 cycles, the thickness of the MOF shell had a very significant increase compared with that after 10 cycles and the magnetic Fe3O4 nanoparticles appeared cross-linked and agglomerated. From TEM images, we found that the Fe3O4 nanoparticles were surrounded by MOF shells. As the number of cycles increased, the thickness of the MOF shell gradually increased.
The XRD peaks of the durian shell fiber and the prepared DF, DFM-10, DFM-20, and DFM-40 are shown in Figure 3a. As can be seen from (A), the crystal lattice of the durian shell fibers had peaks of 15.9°and 21.88°, which is consistent with previous literature reports [40]. The X-ray diffraction pattern of DF in (B) shows the result of diffraction peaks at 2θ values of 30.1°(220), 35.5°(311), 43.2°(400), 57.1°(511), and 62.8°(440), which is consistent with the standard XRD (JCPDS card, file No. 89-4319) data of Fe3O4 with a face-centered cubic structure reported in [41]. DFM-10, DFM-20, and DFM-40 also had the same diffraction peaks (C–E), indicating that the lattice structure of Fe3O4 will not be destroyed after many cycles. At the same time, with the loading of Fe3O4 and the MOF structure of the synthetic durian shell fiber, the diffraction peaks at 13° to 25° were significantly weakened. FTIR spectra of durian shell fiber (A), DF (B), and DFM-10 (C) materials are shown in Figure 3b.The peaks at 560 and 630 cm−1 could be ascribed to the stretching vibration of Fe-O in Fe3O4, indicating the existence of Fe3O4 in the materials. Additionally, peaks at 1634 and 1421 cm−1 were observed in the products of DFM-10, which, respectively, corresponded to C=C and C-O vibrations, confirming the presence of carboxylate groups. The carboxylate group in H3BTC combines with Fe3+ ions to coordinate to form MOF. In this way, the growth of MOF shell proceeded on the DF. The vibration bands corresponded to the -O-C-O- groups around 1574 cm−1 and 1384 cm−1, indicating the presence of the dicarboxylate within the compound. The strong stretching-vibration bands in the range of 3050–3500 cm−1 belonged to -OH and -NH2.
XPS was used to observe the changes in the chemical valence state of the elements before and after the adsorption of MB by DFM-10, and the spectra of C 1s, O 1s, Fe 2p, and broad scans are showed in Figure 4. Figure 4a shows the C 1s spectrum, which had three main peaks, namely C-C (284.8 eV), C-O-C (268.1 eV), and O-C= O (287.7 eV). Figure 4b shows the O 1s spectrum. The peak value was at 532.64 eV binding energy and the peak value did not change before and after adsorption, indicating that no oxidation–reduction reaction was involved in the adsorption process. Figure 4c shows the spectrum of Fe2p. It can be seen that the peaks at 711.05 eV and 724.85 eV represent Fe2p3/2 and Fe2p1/2, respectively, which indicates that Fe2+ and Fe3+ co-exist in the synthesized DFM-10. The total distribution energy spectrum of each element is shown in Figure 4d, from which it can be seen that there was little change before and after adsorption, indicating that the synthesized DFM-10 is a stable material. there was no obvious change in the characteristic peaks of the material before and after adsorption (A and B), indicating that DFM-10 was a stable material.

2.2. Adsorption Experiments

2.2.1. Effect of Different Concentrations

To select a more suitable adsorbent, the adsorption performance of DF, DFM-10, DFM-20, and DFM-40 were evaluated. The adsorbent (5 mg) was mixed with MB solutions of different initial concentrations (5 mg/L, 10 mg/L, and 15 mg/L), and placed in constant-temperature oscillator at 25 °C. After a reaction time of 3 h, the adsorbent material was collected with a magnet. As shown in Figure 5, compared with DFM-0, the adsorption capacities of DFM-10, DFM-20, and DFM-40 were significantly improved, indicating that the introduction of a MOF structure significantly improve the adsorption performance of the material. The adsorption capacities of DFM-10 and DFM-20 to different concentrations of MB were basically the same, but the adsorption capacity of DFM-40 to MB was slightly lower than those of DFM-10 and DFM-20. Combined with SEM and TEM characterization, the reason may be that as the number of cycles increased, the thickness of the MOF shell gradually increased, and the MOF shell outside the adjacent magnetic Fe3O4 nanoparticles underwent crosslinking and agglomeration, which prevented the MB molecules from diffusing into the material, resulting in a decrease in adsorption effect. Therefore, DFM-10 was selected as the adsorbent, and its adsorption isotherm model, adsorption kinetics, and adsorption thermodynamics for MB were studied.

2.2.2. Effect of pH and Ionic Strength

pH plays a significant role in the process of adsorption. Therefore, the adsorption capacity of DF, DFM-10, DFM-20, and DFM-40 under different pH was studied, and the results can be seen in Figure 6a. It can be seen from the figure that as the pH value rose from 2 to 12, the adsorption capacity of DFM-0 on MB gradually increased. This was because of the presence of a large amount of H+ in the particles at low pH solutions, which led to the protonation of the Fe3O4 nanoparticle surface. however, the MB molecule was positively charged, so electrostatic repulsion occurred between the two, resulting in a small amount of adsorption With the gradual increase of pH, the surface of nano-Fe3O4 gradually deprotonated, so the adsorption capacity of MB gradually increased. However, the adsorption amount of MB on DFM-10, DFM-20, and DFM-40 did not change significantly with the increase of pH. This shows that the electrostatic interaction between DFM-10 (or DFM-20, DFM-40) and MB during the adsorption process was not the main factor affecting the adsorption effect. The larger adsorption capacity of DFM-10, DFM-20, or DFM-40 on MB may be attributed to the porous structure of the adsorbent itself, the van der Waals force and hydrogen bonding between the -OH and -COOH groups on the surface of DFM-10 (or DFM-20, DFM-40) and the MB molecule. Studying the influence of pH on the adsorption capacity provides an important reference for the practical application of materials. Therefore, considering comprehensive considerations, pH = 9 was selected in the next experiment. As can be seen from Figure 6b, the isoelelctric point(pHpzc) of DFM-10 was at 5.9, suggesting that the surfaces of the prepared DFM-10 adsorbents were positively charged at pH < 5.9 and the negatively charged at pH > 5.9.
In this work, NaCl and CaCl2, two common salts, were selected to judge the effects of ionic strength on the process of MB adsorption on DFM-10. As shown in Figure 6c, with the increase of salt concentration from 0 to 0.1 mol/L, the adsorption capacity decreased initially. This is because ionic strength reduced the surface charge of DFM-10 and MB, resulting in the decrease of adsorption capacity. When the concentration increased from 0.1 to 0.3 mol/L, the adsorption capacity increased because the salt may have reduced the solubility of MB in water. The decrease of solubility promoted the diffusion of more MB molecules to the adsorbent surface, thus improving the adsorption capacity. When the concentration of NaCl and CaCl2 was at 0.3mol/L, these two salts showed the same effect on adsorption capacity. From 0.3 mol/L to 0.5mol/L, the adsorption capacity showed a slight fluctuation. Although the adsorption capacity was reduced, the reduction was only 23.12%. Therefore, the results showed that the prepared DFM-10 adsorbent had potential application prospects in the treatment of wastewater containing MB.
The competitor experiments were performed towards humic acids under the optimal adsorption condition (adsorbent dosage: 5 mg; initial concentration of MB: 20 mg/L (10 mL); pH: 6; temperature: 25 °C; adsorption time: 3 h), and the results were shown in Figure 6d. As can be seen from Figure 6d, with increasing concentration of humic acids, the MB removal percent decreased. This could have been the result of the competitive effect between humic acids and MB molecule by electrostatic adsorption. However, the MB removal percent decreased by only 18.09% although the concentration of 100 mg/L for humic acids was far higher than that in a realistic water environment, indicating that the adsorbents have the potential to be employed for MB adsorption.

2.2.3. Effect of Time and Initial Concentration of MB

We added 5 mg of adsorbent to 15 mL of MB solution at concentrations of 5 mg/L, 10 mg/L, and 15 mg/L, and, after ultrasonic dispersion, placed the solutions in a constant-temperature shaker at 25 °C and vibrated for adsorption. The effects of different initial concentrations and adsorption time on the adsorption performance of DFM-10 were studied with time-interval sampling, and the results are shown in Figure 7.
It can be seen that with the increase of initial concentration, the adsorption amount of MB by DFM-10 also gradually increased. The adsorption process of MB on DFM-10 can be divided into two stages. The first stage was within 0–15 min after the beginning of adsorption. When the adsorbent is added, MB molecules quickly combined with the adsorption sites on the surface of DFM-10, and the adsorption capacity increased rapidly. With the passage of time, the adsorption sites were gradually occupied, which slowed down the diffusion rate of MB molecules into the adsorbent. Therefore, the increase of adsorption amount gradually slowed until the adsorption sites were completely occupied, an adsorption saturation state was reached, and the adsorption amount did not increase.

2.3. Adsorption Isotherms

To study the adsorption performance of DFM-10 to MB and their interaction pathways, Langmuir, Freundich, Temkin, and Dubinin–Radushkevich adsorption isotherm models were used. [42]. The Langmuir isotherm model assumes that only a single layer of adsorption occurs on the limited active sites on the surface of the adsorbent, and there is no interaction between adjacent sites [43]. The Freundich isotherm model assumes that the adsorption process is a multilayer adsorption, and there are interactions between adsorbate molecules [44]. The Temkin isotherm model considers the interaction between adsorbate and adsorbent. The Temkin model assumes that the adsorption heat of all molecules in the layer will decrease linearly with the accumulation of adsorbed molecules on the adsorbent surface. The Dubinin–Radushkevich isotherm model is an empirical model initially conceived for the adsorption onto micropore solids following a pore-filling mechanism. It is generally applied to express the adsorption mechanism with a Gaussian energy distribution onto a heterogeneous surface [21].
The Langmuir, Freundic, Temkin, and Dubinin–Radushkevich isotherm models are represented by the following equations:
The   Langmuir   isotherm   model   equation :   C e Q e = 1 Q L K L + C e Q L
The   Freundich   isotherm   model   equation :   ln Q e = ln K F + 1 n ln C e
The   Temkin   isotherm   model   equation :   Q e = ( R T b T ) ln K T + ( R T b T ) ln C e
The   Dubinin Radushkevich   isotherm   model   equation :   ln Q e = ln Q m K D R ε 2
ε = R T ln [ 1 + 1 C e ]
where Qe (mg/g) shows the adsorption capacity at equilibrium and QL (mg/g) is the theoretical maximum adsorption capacity. Ce (mg/L) represents the concentration when the adsorption equilibrium is reached and KL is the Langmuir constant (L/mg). KF is the Freundlich constant (mg1−1/n · L1/n/g). 1/n is the parameter characterizing the energy heterogeneity of the adsorption surface. bT is the Temkin constant related to heat of adsorption (J·g/mol·mg). KT is the Temkin isotherm equilibrium binding-energy constant (L/mg). R is the gas constant (8.314 J/mol·K). T is the absolute temperature (K). KDR is the constant related to the adsorption energy (mol2·J2). ε is the adsorption potential (J/mol) [45,46].
The Langmuir, Freundich, Temkin, and Dubinin–Radushkevich isotherm models of MB adsorption by DFM-10 at different temperatures (298 K, 308 K, and 318 K) are shown in Figure 8, and the corresponding fitting parameters are listed in Table 1. It can be seen that the R2 of the Langmuir isotherm fitting model was in the range 0.9791–0.9948, and the fitting curves was basically consistent with the experimental plots. The R2 of Freundich, Temkin, and Dubinin–Radushkevich isotherm models were in the ranges 0.9570–0.9766, 0.8447–0.8910, and 0.8567–0.8811, respectively, and the fitting curves were inconsistent with the experimental data. The correlation coefficients of the isotherms were in the order Langmuir > Freundich > Temkin > Dubinin–Radushkevich. The above results indicate that the Langmuir isotherm model was more appropriate for fitting the experimental data than other three models. In addition, with the increase of temperature, the adsorption amount also gradually increased, indicating that heating was conducive to the adsorption process. Monomolecular layer adsorption could have played a vital role, and the maximum adsorption capacities were 53.31 mg/g at temperature of 318 K. In the case of the Freundich isotherm model, the constant of Freundlich adsorption (KF) increased with increasing adsorption temperatures. High values of KF indicated that there was very good interaction between DFM-10 and MB.

2.4. Adsorption Kinetics

To calculate the adsorption kinetics of DFM-10 adsorbing MB, the experimental data were analyzed through pseudo-first-order, pseudo-second-order, and intraparticle diffusion kinetic models. The pseudo-first-order, pseudo-second-order, and intraparticle diffusion kinetic models are expressed by the following equations [47]:
ln ( Q e Q t ) = ln Q e K 1 t
t Q t = 1 K 2 Q e 2 + t Q e
Q t = K p t 1 2 + C
where Qe is adsorbed by adsorbents at equilibrium and Qt (mg/g) is the amount of MB adsorbed at time t. K1 (min−1), K2 (g mg−1 min−1), and Kp (mg g−1 min1/2) are the rate constants of adsorption for the pseudo-first-order, pseudo-second-order, and intraparticle diffusion kinetic models.
Under the same temperature and dosage, different concentrations (5, 10, and 15 mg/L) of MB solution were selected to study the adsorption effect of DFM-10 on MB at different times. The fitting curves of pseudo-first-order (a), pseudo-second-order (b), and intraparticle diffusion (c) kinetic models of MB were obtained by data analysis and fitting as shown in Figure 9, and the corresponding fitting parameters are listed in Table 2. It can be seen from the fitting results that at different concentrations, the R2 (0.9998, 0.9998, 0.9998) of the pseudo-second-order kinetic fitting model of DMF-10 for MB adsorption was greater than that of the pseudo-first-order kinetic fitting model (0.6817, 0.8401, 0.9523) and intraparticle diffusion kinetic model. This indicates that the adsorption process of DMF-10 to MB is more complex in the pseudo-second-order kinetic model, and the adsorption process is dominated by chemical adsorption. With the increase of initial concentration, the rate constant K2 of the pseudo-second-order kinetic model decreased gradually. This shows that with the increase of initial dye concentration, the adsorption rate gradually slowed down, and the adsorption rate was faster at low concentration. The intraparticle diffusion kinetic curves were treated to test whether internal particle diffusion or external diffusion was the rate-limiting step. The intraparticle diffusion kinetic plot did not match the origin of coordinates (Figure 9c). This result indicated that the adsorption mechanism was influenced by multi-mechanisms, and that internal particle diffusion may be not the main element.

2.5. Thermodynamic Study

In order to understand the effect of temperature on the adsorption of MB by DFM-10 and affirm the adsorption process, thermodynamic results from the adsorption data of MB were calculated at three different temperatures by using DFM-10 (298 K, 308 K, and 318 K): Gibbs free energy ΔG (KJ/mol), enthalpy change ΔH (KJ/mol)) and entropy change ΔS (J/(mol·K)) [48].The calculation formula of each thermodynamic parameter is as follows:
K e = Q e C e
Δ G = - R T ln K e
ln K e = Δ S R Δ H R T
where R is the ideal gas constant (8.314 J/mol−1 K−1). T acted as the absolute temperature (K). Ke is the thermodynamic equilibrium constant, Qe (mg/g) is the amount of adsorption at adsorption equilibrium, and Ce (mg/L) is the equilibrium concentration [49]. We plotted lnKe versus 1/T and calculated ΔS from the intercept and ΔH from the slope. The calculation results are listed in Table 3.
It can be seen from Table 4 that ΔH was positive, indicating that the adsorption of MB in aqueous solution by DFM-10 was an endothermic process. The higher the temperature, the better the adsorption effect. When ΔH was less than 40 kJ/mol, the adsorption process was dominated by physical adsorption [50], and the ΔH value of the adsorption of MB in aqueous solution by DFM-10 was 4.6747 kJ/mol within this range, indicating that the adsorption process was dominated by physical adsorption. ΔG was negative, indicating that the adsorption process of DFM-10 to MB was spontaneous, and as the temperature increased, the value of ΔG gradually decreased, which shows that the spontaneity of the reaction was gradually increasing as the temperature rose. When the value of ΔG is in the range of −20 to 0 kJ/mol, the adsorption process is physical adsorption. Obviously, the ΔG value of the adsorption of MB in the aqueous solution by DFM-10 was within this range, indicating that the adsorption process was physical adsorption, which is consistent with the discussion of the ΔH value. In addition, ΔS was 16.8767 kJ/mol greater than 0, indicating that the disorder of the solid−liquid interface increased during the process of adsorption and the adsorption process was irreversible. Thermodynamic studies have shown that the adsorption of MB in aqueous solution by DFM-10 is a spontaneous endothermic physical adsorption process.
Based on all the discussions above, the interaction mechanism when DFM-10 adsorbs MB is proposed, as shown in Figure 10. It can be seen from Table 4 that the specific surface area of DFM-10 is 226.88 m2/g, which is significantly higher than that of DFM-10. At the same time, DFM-10 has smaller mean pore radius and larger total pore volume, which provides a good structural basis for MB adsorption. Meanwhile, in the metal–organic framework (MOF) synthesis process, H3BTC and MB molecules have a weak interaction and can be adsorbed by hydrogen bonds and other forces [51]. The -OH structure in DFM-10 can be used as a π electron donor, and the conjugated unsaturated carbonyl structure in MB molecule can be used as a π electron acceptor. Therefore, the π−π interaction in the adsorption process may be a potential mechanism.
Figure 11 shows the separation performance of DFM-10 in aqueous solution. It can be seen from the same figure that under the action of a strong magnet, DFM-10 can be basically separated from aqueous solution within 9 s, and can be completely separated from aqueous solution within 30 s. This is because the load of Fe3O4 provides the ability for rapid separation of DFM-10 from aqueous solution.

2.6. Total Organic Carbon (TOC) Analysis

The total organic carbon (TOC) is an indicator for evaluating the removal of organic pollutants in water. Figure 12 showed the TOC amount of the natural river water (Hung-tse Lake in Jiangsu Province, Hongze) before and after adsorption by DFM-10. The initial amount of TOC was 51.22 mg/L before adding DFM-10. After the absorption reaction, the TOC amount decreased to 4.39 mg/L. The results indicated that the TOC removal rate reached 91.44%, indicating that the adsorbent had excellent purification performance.

2.7. Recycling Ability

Through cyclic experiments, we studied whether DFM-10 can be recycled, which has an important impact on the practicality of the adsorbent. We conducted MB adsorption-cycle experiments on DFM-10 five times. At 298 K, the initial concentration of MB was 10 mg/L. After 3 h of adsorption, the adsorption was carried out with HCl/ethanol (0.5%, v/v). The obtained result is shown in Figure 13. It can be seen from Figure 13a that the removal rate of MB by DFM-10 was 90.50% in the first cycle. After five cycles, the removal rate of MB by DFM-10 remained at 83.15% without significant decrease. From Figure 13b, the crystal structure of DFM-10 had no apparent change after four consecutive cycle reactions, suggesting that DFM-10 could be used as an efficient and recyclable adsorbent to remove MB from wastewater. The removal efficiency of DFM-10 for MB was also higher than reported adsorbents, as mentioned in Table S1 [52,53,54,55,56,57].

3. Materials and Methods

3.1. Chemicals and Reagents

Durian shells were purchased from supermarket in Nanjing. Ferric chloride hexahydrate (FeCl3•6H2O) and sodium chlorite (NaClO2) were from Shanghai Macklin Biochemical Co., Ltd.(Shanghai, China) Benzene-1, 3, 5-tricarboxylic acid (H3BTC) and urea (CH4N2O) were purchased from Aladdin Reagent Co., Ltd. (Shanghai, China). Sodium hydroxide (NaOH), sodium bicarbonate (NaHCO3), and methylene blue (MB) were purchased from Sinopharm Chemical Regent Co., Ltd. (Shanghai, China). Vitamin C (C6H8O6) was purchased from Yuanye Biological Technology Co., Ltd. (Shanghai, China). All the above chemical reagents were used without further purification.

3.2. Synthesis of Adsorbent

3.2.1. Extraction of Durian Shell Fiber

Firstly, 25 g of durian shell was dried at 60 °C for 6 h, cut into pieces, and placed in a NaOH solution (1 M, 500 mL). The solid was then stirred at 85 °C for 4 h to remove cellulose and impurities. Subsequently, the durian shell fibers were put into a 5 wt% NaClO2 solution for decolorization, and magnetically stirred at 80 °C for 2h, filtered, separated, washed five times with deionized water, and dried at 60 °C for 24 h to get durian shell fibers.

3.2.2. Preparation of Durian Shell Fiber–Fe3O4 (DF)

Firstly, 0.2 g durian husk fiber, 2.0 g NaOH, and 4.0 g CH4N2O were dissolved in 40 mL deionized water, stirred to form a transparent solution, and flocculent fibers were collected. Secondly, 0.2 mol/L FeCl3•6H2O solution (10 mL) was added to 0.45 mol/L NaHCO3 solution (10 mL), and then vigorously stirred for 30 min. Then, 0.03 mol/L vitamin C solution (10 mL) was added to above-mentioned flocculent durian shell fiber and stirred for 20 min at 25 ℃. Then, the material was put into a 50 mL Teflon kettle and heated at 180 °C for 8 h. The final products could then be separated with ethanol and washed with deionized water three times, collected by centrifugation (8000 rpm) and dried at 80 °C for 12 h to obtain durian shell fiber–Fe3O4 (DF).

3.2.3. Preparation of Durian Shell Fiber–Fe3O4 MOF (DFM)

The preparation of durian shell fiber–Fe3O4 MOF composite material was carried out by the self-assembly method. Firstly, 0.05g of the prepared DF was dispersed ultrasonically and added to 4 mL FeCl3•6H2O ethanol solution (10 mmol/L) for 15 min, then dispersed in 4 mL of H3BTC ethanol solution (10 mmol/L), and heated at 70 °C for 30 min. Each step was separated with a strong permanent magnet and washed with ethanol. After a certain number of FeCl3•6H2O/H3BTC cycles, the sample was recovered with a permanent magnet and washed with ethanol/water, and then vacuum dried at 60 °C for 12 h to obtain DFM. The synthetic process of DFM preparation can be seen in Scheme 1. The obtained samples at different cycles (0, 10, 20, and 40) were named DF, DFM-10, DFM-20, and DFM-40, respectively.

3.3. Adsorption Properties

In the adsorption experiment, the MB stock solution had a concentration of 100 mg/L. It was stored in a dark place and dilute to 2.5–20 mg/L when used. The MB adsorption experiment was carried out in a constant-temperature shaker. The specific experiment was as follows: The adsorbent (5 mg) was mixed with 15mL MB solutions of different initial concentrations and placed in constant-temperature oscillator at 25 °C. After a reaction time of 3 h, the adsorbent material was collected with a magnet. MB concentration in the supernatants was measured at 664 nm by using a UV–Vis spectrophotometer. The quantity of adsorption (Qe (mg/g)) and rate of removing MB (R (%)) were calculated based on the following equations:
Q e = ( C 0 C e ) × V m
R = C 0 - C e C 0 × 100 %
where C0 (mg/L) is the initial concentrations of MB and Ce (mg/L) is the equilibrium concentrations of MB; m is determined to the weight of adsorbent (g), and V is defined as the volume of the liquid solution (mL) [58,59].

3.4. Characterization

FTIR spectra of the sample was performed with an FTIR spectrometer (AVATAR 360, Nicole company, MA, USA). A minimum of 32 scans was signal-averaged with a resolution of 2 cm−1 in the 4000–400 cm−1 ranges. X-ray diffraction (XRD) analysis of the material was conducted on X-ray powder diffraction (XRD-6100, Shimadzu company, Kyoto, Japan) with a scanning rate of 5 °/min. The morphology of the as-prepared adsorbent was characterized by scanning electron microscope (SEM, S-4800, Hitachi company, Kyoto, Japan). Transmission electron microscopy (TEM) analysis was undertaken on a JEM-1400 UV–Vis spectrophotometer (UV-2450, Shimadzu Company, Kyoto, Japan). The zeta potential of material was tested by using a Zeta potential analyzer (Zetasizer Nano ZS90, Malvern Instruments Ltd., Malvern, UK). Total organic carbon(TOC) was obtained by total organic carbon analyzer (Shimadzu, Kyoto, Japan).

4. Conclusions

In this work, durian shell biomass carbon fiber and Fe3O4 functionalized metalorganic framework composite material (DFM) was prepared and the structure was characterized. Compared with conventional adsorbents, it can effectively remove MB from wastewater. Then, comparing the adsorption capacity with different adsorbents to MB, it can be seen that DFM-10 had the best effect and the maximum adsorption capacity for MB was 53.31mg/g. Adsorption isotherms studies showed that DFM-10 is more in line with the Langmuir isotherm model, adsorption kinetics studies showed that DFM-10 is more in line with the pseudo-second-order kinetic model, and thermodynamic analysis showed that DFM-10 adsorbs MB in an aqueous spontaneous endothermic physical adsorption process. This study proposes an effective method to prepare DFM composite material, which has high adsorption efficiency and can be recycled, and is expected to be widely used to removing MB in wastewater.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms23115900/s1. References [52,53,54,55,56,57] are cited in supplementary materials.

Author Contributions

Conceptualization, methodology, and formal analysis were done by Z.C., Q.L. and H.L.; material characterization were done by J.W., G.T. and F.W.; writing—original draft preparation, resources, supervision, and funding acquisition were done by J.H., Y.Z. and G.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by The Natural Science Foundation of the Jiangsu Higher Education Institutions of China (21KJB530010), Jiangsu Planned Projects for Postdoctoral Research Funds (2020Z098), Opening Project of Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensification Technology (2021K006), Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology (JLCT202107), and Start-up Fund from Nanjing Forestry University and the Analysis and Test Center of Nanjing Forestry University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We thank “National Positioning Observation Station of Hung-tse Lake Wetland Ecosystem in Jiangsu Province, Hongze 223100, China” for their generous assistance with the support of the experiment.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Qiu, J.; Feng, Y.; Zhang, X.; Jia, M.; Yao, J. Acid-promoted synthesis of UiO-66 for highly selective adsorption of anionic dyes: Adsorption performance and mechanisms. J. Colloid Interface Sci. 2017, 499, 151–158. [Google Scholar] [CrossRef] [PubMed]
  2. Reynel, H.E.; Mendoza, D.I.; Bonilla, A. Relevance of anionic dye properties on water decolorization performance using bone char: Adsorption kinetics, isotherms and breakthrough curves. J. Mol. Liq. 2016, 219, 425–434. [Google Scholar] [CrossRef]
  3. Atrous, M.; Sellaoui, L.; Bouzid, M.; Lima, E.C.; Thue, P.S.; Bonilla-Petriciolet, A.; Ben Lamine, A. Adsorption of dyes acid red 1 and acid green 25 on grafted clay: Modeling and statistical physics interpretation. J. Mol. Liq. 2019, 294, 111610. [Google Scholar] [CrossRef]
  4. Rashid, J.; Tehreem, F.; Rehman, A.; Kumar, R. Synthesis using natural functionalization of activated carbon from pumpkin peels for decolourization of aqueous methylene blue. Sci. Total Environ. 2019, 671, 369–376. [Google Scholar] [CrossRef]
  5. Wu, G.; Xing, W. Fabrication of ternary visible-light-driven semiconductor photocatalyst and its effective photocatalytic performance. Mater. Technol. 2018, 34, 292–300. [Google Scholar] [CrossRef]
  6. Hassan, M.M.; Carr, C.M. A critical review on recent advancements of the removal of reactive dyes from dyehouse effluent by ion-exchange adsorbents. Chemosphere 2018, 209, 201–219. [Google Scholar] [CrossRef]
  7. Wu, G.; Liu, Q.; Wang, J.; Cai, Z.; Li, H.; Zhang, T.; Lu, R.; Li, P.; Han, J.; Xing, W. Synthesis of silver-based composite photocatalysis material and its visible-light-driven photocatalytic degradation of dye pollutants. Fresenius Environ Bull. 2021, 30, 9696–9706. [Google Scholar]
  8. Huang, X.; Wei, D.; Zhang, X.; Fan, D.; Sun, X.; Du, B.; Wei, Q. Synthesis of amino-functionalized magnetic aerobic granular sludge-biochar for Pb(II) removal: Adsorption performance and mechanism studies. Sci. Total Environ. 2019, 685, 681–689. [Google Scholar] [CrossRef]
  9. Wu, G.; Ma, F.; Xiong, R.; Xu, Z.; Li, P.; Han, J.; Xing, W. Fabrication of semiconductor photocatalyst and its high photo-catalytic performance under visble light irradiation. Fresenius Env. Bull. 2021, 30, 10482–10491. [Google Scholar]
  10. AliYounis, S.; Maitlo, H.A.; Lee, J.; Kim, K.-H. Nanotechnology-based sorption and membrane technologies for the treatment of petroleum-based pollutants in natural ecosystems and wastewater streams. Adv. Colloid Interface Sci. 2019, 275, 102071. [Google Scholar] [CrossRef]
  11. Wu, G.; Liu, Q.; Wang, J.; Zhang, Y.; Yu, C.; Bian, H.; Hegazy, M.; Han, J.; Xing, W. Facile fabrication of Bi2WO6/biochar composites with enhanced charge carrier separation for photodecomposition of dyes. Colloids Surf. A Physicochem. Eng. Asp. 2021, 634, 127945. [Google Scholar] [CrossRef]
  12. Wu, G.; Xing, W.; Han, J.; Li, P. Fabrication of silver-based visible-light-driven photocatalyst and degradation of organic pollutions in wastewater. Fresenius Env. Bull. 2020, 29, 445–453. [Google Scholar]
  13. Filote, C.; Volf, I.; Santos, S.C.R.; Botelho, C.M.S. Bioadsorptive removal of Pb(II) from aqueous solution by the biore-finery waste of Fucus spiralis. Sci. Total Environ. 2019, 648, 1201–1209. [Google Scholar] [CrossRef]
  14. Xing, W.; Zhang, Y.; Cheng, K.; Zou, J.; Wu, G. Fabrication of novel carbon species into porous g-C3N4 nanosheet frameworks with enhanced photocatalytic performance. New J. Chem. 2021, 45, 10589–10593. [Google Scholar] [CrossRef]
  15. Jawad, A.H.; Mubarak, N.S.A.; Abdulhameed, A.S. Hybrid Crosslinked Chitosan-Epichlorohydrin/TiO2 Nanocomposite for Reactive Red 120 Dye Adsorption: Kinetic, Isotherm, Thermodynamic, and Mechanism Study. J. Polym. Environ. 2020, 28, 624–637. [Google Scholar] [CrossRef]
  16. Tran, T.H.; Le, A.H.; Pham, T.H.; Nguyen, D.T.; Chang, S.W.; Chung, W.J. Adsorption isotherms and kinetic modeling of methylene blue dye onto a carbonaceous hydrochar adsorbent derived from coffee husk waste. Sci. Total Environ. 2020, 725, 138325. [Google Scholar] [CrossRef]
  17. Maksoud, M.I.A.A.; Elgarahy, A.M.; Farrell, C.; Al-Muhtaseb, A.H.; Rooney, D.W.; Osman, A.I. Insight on water remediation application using magnetic nanomaterials and biosorbents. Coord. Chem. Rev. 2020, 403, 213096. [Google Scholar] [CrossRef]
  18. Hasanzadehab, M.; Simchi, A.; Far, H.S. Nanoporous composites of activated carbon-metal organic frameworks for organic dye adsorption: Synthesis, adsorption mechanism and kinetics studies. J. Ind. Eng. Chem. 2019, 81, 405–414. [Google Scholar] [CrossRef]
  19. Thines, K.; Abdullah, E.C.; Mubarak, N.; Ruthiraan, M. Synthesis of magnetic biochar from agricultural waste biomass to enhancing route for waste water and polymer application: A review. Renew. Sustain. Energy Rev. 2017, 67, 257–276. [Google Scholar] [CrossRef]
  20. Lai, F.; Miao, Y.-E.; Zuo, L.; Lu, H.; Huang, Y.; Liu, T. Biomass-Derived Nitrogen-Doped Carbon Nanofiber Network: A Facile Template for Decoration of Ultrathin Nickel-Cobalt Layered Double Hydroxide Nanosheets as High-Performance Asymmetric Supercapacitor Electrode. Small 2016, 12, 3235–3244. [Google Scholar] [CrossRef]
  21. Xing, W.; Liu, Q.; Wang, J.; Xia, S.; Ma, L.; Lu, R.; Zhang, Y.; Huang, Y.; Wu, G. High Selectivity and Reusability of Bi-omass-Based Adsorbent for Chloramphenicol Removal. Nanomaterials 2021, 11, 2950. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, C.; Zeng, G.; Huang, D.; Lai, C.; Chen, M.; Cheng, M.; Tang, W.; Tang, L.; Dong, H.; Huang, B.; et al. Biochar for environmental management: Mitigating greenhouse gas emissions, contaminant treatment, and potential negative impacts. Chem. Eng. J. 2019, 373, 902–922. [Google Scholar] [CrossRef]
  23. Son, E.-B.; Poo, K.-M.; Chang, J.-S.; Chae, K.-J. Heavy metal removal from aqueous solutions using engineered magnetic biochars derived from waste marine macro-algal biomass. Sci. Total Environ. 2018, 615, 161–168. [Google Scholar] [CrossRef]
  24. Luo, M.-T.; Zhao, C.; Huang, C.; Chen, X.-F.; Huang, Q.-L.; Qi, G.-X.; Tian, L.-L.; Xiong, L.; Li, H.-L.; Chen, X.-D. Efficient Using Durian Shell Hydrolysate as Low-Cost Substrate for Bacterial Cellulose Production by Gluconacetobacter xylinus. Indian J. Microbiol. 2017, 57, 393–399. [Google Scholar] [CrossRef]
  25. Yazidi, A.; Atrous, M.; Soetaredjo, F.E.; Sellaoui, L.; Ismadji, S.; Erto, A.; Bonilla-Petriciolet, A.; Dotto, G.L.; Ben Lamine, A. Adsorption of amoxicillin and tetracycline on activated carbon prepared from durian shell in single and binary systems: Experimental study and modeling analysis. Chem. Eng. J. 2019, 379, 122320. [Google Scholar] [CrossRef]
  26. Laysandra, L.; Santosa, F.H.; Austen, V.; Soetaredjo, F.E.; Foe, K.; Putro, J.N.; Ju, Y.H.; Ismadji, S. Rarasapo-nin-bentonite-activated biochar from durian shells composite for removal of crystal violet and Cr(VI) from aqueous solution. Environ. Sci. Pollut. Res. 2018, 25, 30680–30695. [Google Scholar] [CrossRef]
  27. Kurniawan, A.; Sisnandy, V.O.A.; Trilestari, K.; Sunarso, J.; Indraswati, N.; Ismadji, S. Performance of durian shell waste as high capacity biosorbent for Cr(VI) removal from synthetic wastewater. Ecol. Eng. 2011, 37, 940–947. [Google Scholar] [CrossRef] [Green Version]
  28. Yang, Q.; Zhao, Q.; Ren, S.; Lu, Q.; Guo, X.; Chen, Z. Fabrication of core-shell Fe3O4@MIL-100(Fe) magnetic microspheres for the removal of Cr(VI) in aqueous solution. J. Solid State Chem. 2016, 244, 25–30. [Google Scholar] [CrossRef]
  29. Zheng, X.; Wang, J.; Xue, X.; Liu, W.; Kong, Y.; Cheng, R.; Yuan, D. Facile synthesis of Fe3O4@MOF-100(Fe) magnetic microspheres for the adsorption of diclofenac sodium in aqueous solution. Environ. Sci. Pollut. Res. 2018, 25, 31705–31717. [Google Scholar] [CrossRef]
  30. Luu, Q.S.; Do, U.T.; Kim, D.; Kim, J.; Jo, D.; Nguyen, Q.T.; Lee, Y. Enhancing adsorption efficiencies of organic molecules through covalently bonded structures of magnetic carbon nanoparticles. J. Ind. Eng. Chem. 2021, 105, 74–82. [Google Scholar] [CrossRef]
  31. Wu, G.; Liu, Q.; Wang, J.; Xia, S.; Wu, H.; Zong, J.; Han, J.; Xing, W. Facile fabrication of rape straw biomass fi-ber/β-CD/Fe3O4 as adsorbent for effective removal of ibuprofen. Ind. Crops. Prod. 2021, 173, 114150. [Google Scholar] [CrossRef]
  32. Sun, Y.; Ni, P.Y.; Zhu, M.Y.; Yao, Y.L.; Fu, S.W. Preparation of Fe3O4@C submicron rods for adsorption of methylene blue and fast separation from water. Micro. Nano Lett. 2019, 14, 962–966. [Google Scholar] [CrossRef]
  33. Rajput, S.; Pittman, C.U., Jr.; Mohan, D. Magnetic magnetite (Fe3O4) nanoparticle synthesis and applications for lead (Pb2+) and chromium (Cr6+) removal from water. J. Colloid Interface Sci. 2016, 468, 334–346. [Google Scholar] [CrossRef]
  34. Wang, J.; Zhao, G.; Yu, F. Facile preparation of Fe3O4@MOF core-shell microspheres for lipase immobilization. J. Taiwan Inst. Chem. Eng. 2016, 69, 139–145. [Google Scholar] [CrossRef]
  35. Su, M.; Fang, Y.; Li, B.; Yin, W.; Gu, J.; Liang, H.; Li, P.; Wu, J. Enhanced hexavalent chromium removal by activated carbon modified with micro-sized goethite using a facile impregnation method. Sci. Total Environ. 2018, 647, 47–56. [Google Scholar] [CrossRef]
  36. Sun, X.; Gao, G.; Yan, D.; Feng, C. Synthesis and electrochemical properties of Fe3O4@MOF core-shell microspheres as an anode for lithium ion battery application. Appl. Surf. Sci. 2017, 405, 52–59. [Google Scholar] [CrossRef]
  37. Abdi, J.; Vossoughi, M.; Mahmoodi, N.M.; Alemzadeh, I. Synthesis of metal-organic framework hybrid nanocomposites based on GO and CNT with high adsorption capacity for dye removal. Chem. Eng. J. 2017, 326, 1145–1158. [Google Scholar] [CrossRef]
  38. Lin, K.Y.A.; Liu, Y.T.; Chen, S.Y. Adsorption of fluoride to UiO-66-NH2 in water: Stability, kinetic, isotherm and thermodynamic studies. J. Colloid Interf. Sci. 2016, 461, 79–87. [Google Scholar] [CrossRef]
  39. Efome, J.E.; Rana, D.; Matsuura, T.; Lan, C.Q. Insight Studies on Metal-Organic Framework Nanofibrous Membrane Adsorption and Activation for Heavy Metal Ions Removal from Aqueous Solution. ACS Appl. Mater. Interfaces 2018, 10, 18619–18629. [Google Scholar] [CrossRef]
  40. Xia, L.J.; Zhou, S.J.; Zhang, C.H.; Fu, Z.A.; Wang, A.M.; Zhang, Q.; Wang, Y.L.; Liu, X.; Wang, X.G.; Xu, W.L. En-vironment-friendly Juncus effusus-based adsorbent with a three-dimensional network structure for highly efficient removal of dyes from wastewater. J. Clean Prod. 2020, 259, 120812. [Google Scholar] [CrossRef]
  41. Liang, S.; Shi, S.; Zhang, H.; Qiu, J.; Yu, W.; Li, M.; Gan, Q.; Yu, W.; Xiao, K.; Liu, B.; et al. One-pot solvothermal synthesis of magnetic biochar from waste biomass: Formation mechanism and efficient adsorption of Cr(VI) in an aqueous solution. Sci. Total Environ. 2019, 695, 133886. [Google Scholar] [CrossRef]
  42. Zhou, Y.; Liu, X.; Xiang, Y.; Wang, P.; Zhang, J.; Zhang, F.; Wei, J.; Luo, L.; Lei, M.; Tang, L. Modification of biochar derived from sawdust and its application in removal of tetracycline and copper from aqueous solution: Adsorption mechanism and modelling. Bioresour. Technol. 2017, 245, 266–273. [Google Scholar] [CrossRef]
  43. Dong, H.; Deng, J.; Xie, Y.; Zhang, C.; Jiang, Z.; Cheng, Y.; Hou, K.; Zeng, G. Stabilization of nanoscale zero-valent iron (nZVI) with modified biochar for Cr(VI) removal from aqueous solution. J. Hazard. Mater. 2017, 332, 79–86. [Google Scholar] [CrossRef]
  44. Han, Y.; Cao, X.; Ouyang, X.; Sohi, S.; Chen, J. Adsorption kinetics of magnetic biochar derived from peanut hull on removal of Cr (VI) from aqueous solution: Effects of production conditions and particle size. Chemosphere 2016, 145, 336–341. [Google Scholar] [CrossRef]
  45. Fan, S.; Wang, Y.; Wang, Z.; Tang, J.; Tang, J.; Li, X. Removal of methylene blue from aqueous solution by sewage sludge-derived biochar: Adsorption kinetics, equilibrium, thermodynamics and mechanism. J. Environ. Chem. Eng. 2017, 5, 601–611. [Google Scholar] [CrossRef]
  46. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Sornalingam, K. Single and competitive sorption properties and mechanism of functionalized biochar for removing sulfonamide antibiotics from water. Chem. Eng. J. 2017, 311, 348–358. [Google Scholar] [CrossRef]
  47. Yap, M.W.; Mubarak, N.M.; Sahu, J.N.; Abdullah, E.C. Microwave induced synthesis of magnetic biochar from agri-cultural biomass for removal of lead and cadmium from wastewater. J. Ind. Eng. Chem. 2017, 45, 287–295. [Google Scholar] [CrossRef]
  48. Yuan, J.; Zhu, Y.; Wang, J.; Gan, L.; He, M.; Zhang, T.; Li, P.; Qiu, F. Preparation and application of Mg–Al composite oxide/coconut shell carbon fiber for effective removal of phosphorus from domestic sewage. Food Bioprod. Process. 2021, 126, 293–304. [Google Scholar] [CrossRef]
  49. Sewu, D.D.; Boakye, P.; Woo, S.H. Highly efficient adsorption of cationic dye by biochar produced with Korean cabbage waste. Bioresour. Technol. 2017, 224, 206–213. [Google Scholar] [CrossRef]
  50. Eren, E.; Cubuk, O.; Ciftci, H.; Eren, B.; Caglar, B. Adsorption of basic dye from aqueous solutions by modified sepiolite: Equilibrium, kinetics and thermodynamics study. Desalination 2010, 252, 88–96. [Google Scholar] [CrossRef]
  51. Xu, Y.; Jin, J.; Li, X.; Song, C.; Meng, H.; Zhang, X. Adsorption behavior of methylene blue on Fe3O4-embedded hybrid magnetic metal–organic framework. Desalination Water Treat. 2016, 57, 25216–25225. [Google Scholar] [CrossRef]
  52. Cai, X.; Li, J.; Liu, Y.; Hu, X.; Tan, X.; Liu, S.; Wang, H.; Gu, Y.; Luo, L. Design and Preparation of Chitosan-Crosslinked Bismuth Ferrite/Biochar Coupled Magnetic Material for Methylene Blue Removal. Int. J. Environ. Res. Public Health. 2020, 17, 6. [Google Scholar] [CrossRef] [Green Version]
  53. Li, Y.; Zhang, Y.; Zhang, Y.; Wang, G.; Li, S.; Han, R.; Wei, W. Reed biochar supported hydroxyapatite nanocomposite: Characterization and reactivity for methylene blue removal from aqueous media. J. Mol. Liq. 2018, 263, 53–63. [Google Scholar] [CrossRef]
  54. Aslam Malana, M.; Parveen, S.; Beenish Qureshi, R. Adsorptive removal of organic dyes from aqueous solutions using acrylic acid-acrylonitrile-N-isopropylacrylamide polymeric gels as adsorbents: Linear and non linear isotherms. Desalination Water Treat. 2016, 57, 22543–22550. [Google Scholar] [CrossRef]
  55. Hoslett, J.; Ghazal, H.; Mohamad, N.; Jouhara, H. Removal of methylene blue from aqueous solutions by biochar prepared from the pyrolysis of mixed municipal discarded material. Sci. Total Environ. 2020, 714, 136832. [Google Scholar] [CrossRef]
  56. Gao, Y.; Sun, D.; Han, C.; Huang, J. Comprehensive Utilization of Phosphogypsum: Adsorption of Methylene Blue and its Application in Bricks. Surf. Rev. Lett. 2021, 28, 2150075. [Google Scholar] [CrossRef]
  57. Wanyonyi, W.C.; Onyari, J.M.; Shiundu, P.M. Adsorption of methylene blue dye from aqueous solutions using eichhornia crassipes. Bull. Environ. Contam. Toxicol. 2013, 91, 362–366. [Google Scholar] [CrossRef]
  58. Liu, Z.; Yang, W.; Xu, W.; Liu, Y. Removal of elemental mercury by bio-chars derived from seaweed impregnated with potassium iodine. Chem. Eng. J. 2018, 339, 468–478. [Google Scholar] [CrossRef]
  59. Selvaraju, G.; Abu Bakar, N.K. Production of a new industrially viable green-activated carbon from Artocarpus integer fruit processing waste and evaluation of its chemical, morphological and adsorption properties. J. Clean. Prod. 2016, 141, 989–999. [Google Scholar] [CrossRef]
Figure 1. SEM images of DF (a,b), DFM-10 (c,d), DFM-20 (e,f), and DFM-40, and (g,h) samples at different magnifications.
Figure 1. SEM images of DF (a,b), DFM-10 (c,d), DFM-20 (e,f), and DFM-40, and (g,h) samples at different magnifications.
Ijms 23 05900 g001
Figure 2. TEM images of DF (a,b), DFM-10 (c,d), DFM-20 (e,f), and DFM-40, and (g,h) samples at different magnifications.
Figure 2. TEM images of DF (a,b), DFM-10 (c,d), DFM-20 (e,f), and DFM-40, and (g,h) samples at different magnifications.
Ijms 23 05900 g002
Figure 3. (a) XRD patterns of durian shell fiber (A), DF (B), DFM-10 (C), DFM-20 (D), and DFM-40 (E) and (b) FTIR spectra of durian shell fiber (A), DF (B), and DFM-10 (C).
Figure 3. (a) XRD patterns of durian shell fiber (A), DF (B), DFM-10 (C), DFM-20 (D), and DFM-40 (E) and (b) FTIR spectra of durian shell fiber (A), DF (B), and DFM-10 (C).
Ijms 23 05900 g003
Figure 4. XPS spectra of C 1s (a), O 1s (b), Fe 2p (c), and broad scans (d) of DFM-10 before (A) and after (B) the adsorption of MB.
Figure 4. XPS spectra of C 1s (a), O 1s (b), Fe 2p (c), and broad scans (d) of DFM-10 before (A) and after (B) the adsorption of MB.
Ijms 23 05900 g004
Figure 5. Adsorption effects of DF, DFM-10, DFM-20, and DFM-40 on MB at different concentrations.
Figure 5. Adsorption effects of DF, DFM-10, DFM-20, and DFM-40 on MB at different concentrations.
Ijms 23 05900 g005
Figure 6. (a) Adsorption effects of DF, DFM-10, DFM-20, and DFM-40 on MB at different pH values, (b) zeta potential of DFM-10 adsorbent, (c) effect of ionic strength on the adsorption capacity, and (d) effects of humic acids solution concentration on MB removal percent.
Figure 6. (a) Adsorption effects of DF, DFM-10, DFM-20, and DFM-40 on MB at different pH values, (b) zeta potential of DFM-10 adsorbent, (c) effect of ionic strength on the adsorption capacity, and (d) effects of humic acids solution concentration on MB removal percent.
Ijms 23 05900 g006
Figure 7. The relationship between adsorption amount and time of MB adsorbed by DFM-10 at different concentrations.
Figure 7. The relationship between adsorption amount and time of MB adsorbed by DFM-10 at different concentrations.
Ijms 23 05900 g007
Figure 8. Langmuir (a), Freundich (b), Temkin (c), and Dubinin–Radushkevich (d) isotherms of MB adsorption on DFM-10.
Figure 8. Langmuir (a), Freundich (b), Temkin (c), and Dubinin–Radushkevich (d) isotherms of MB adsorption on DFM-10.
Ijms 23 05900 g008
Figure 9. Pseudo-first-order (a), pseudo-second-order (b), and intraparticle diffusion (c) kinetic models fitting curves of MB.
Figure 9. Pseudo-first-order (a), pseudo-second-order (b), and intraparticle diffusion (c) kinetic models fitting curves of MB.
Ijms 23 05900 g009
Figure 10. Possible adsorption mechanism of MB onto DFM-10.
Figure 10. Possible adsorption mechanism of MB onto DFM-10.
Ijms 23 05900 g010
Figure 11. Magnetic separation process of DFM-10 in solution.
Figure 11. Magnetic separation process of DFM-10 in solution.
Ijms 23 05900 g011
Figure 12. Removal effect of DFM-10 on TOC in actual wastewater.
Figure 12. Removal effect of DFM-10 on TOC in actual wastewater.
Ijms 23 05900 g012
Figure 13. (a) Recycling performance of DFM-10 adsorbent and (b) XRD patterns of DFM-10 before and after five reaction cycles.
Figure 13. (a) Recycling performance of DFM-10 adsorbent and (b) XRD patterns of DFM-10 before and after five reaction cycles.
Ijms 23 05900 g013
Scheme 1. The synthetic process of DFM preparation.
Scheme 1. The synthetic process of DFM preparation.
Ijms 23 05900 sch001
Table 1. Langmuir, Freundic, Temkin, and Dubinin–Radushkevich adsorption isotherm model parameters.
Table 1. Langmuir, Freundic, Temkin, and Dubinin–Radushkevich adsorption isotherm model parameters.
ModelParameter298 K308 K318 K
LangmuirQl (mg/g)34.810042.250053.3100
KL (L/mg)1.84271.34021.9642
R20.99310.99480.9791
FreundichKF (mg1−1/n L1/n /g)20.560022.690030.6800
1/n0.20650.26240.2862
R20.97660.96090.9570
TemkinbT (J.g/mol.mg)0.70530.51240.4142
KT (L/mg)1.00651.00511.0054
R20.89100.87350.8447
Dubinin–RadushkevichQl (mg/g)28.373131.875340.1395
KDR (molJ)0.11050.14320.1423
R20.88020.85670.8811
Table 2. Pseudo-first-order, pseudo-second-order, and intraparticle diffusion dynamics-model-related parameters.
Table 2. Pseudo-first-order, pseudo-second-order, and intraparticle diffusion dynamics-model-related parameters.
ModelParameter5 (mg/L)10 (mg/L)15 (mg/L)
Pseudo-first-orderQe (mg/g)1.51004.09006.8100
K1 (/min)0.02250.03010.0438
R20.68170.80410.9523
Pseudo-second-orderQe (mg/g)15.090027.600037.5100
K2 (g/(mg min))0.05810.01710.0101
R20.99980.99980.9998
Kp1 (mg/g/min1/2)0.32370.37670.7017
C1 (mg/g)3.52096.75509.4854
R20. 99610. 99280.9884
Intraparticle diffusionKp2 (mg/g/min1/2)0.01840.09930.1144
C2 (mg/g)4.70228.334312.6905
R20. 82020.74680.5608
Kp3 (mg/g/min1/2)0.02300.13450.0730
C3 (mg/g)4.66167.961513.2741
R20. 96650.95910. 9548
Table 3. Thermodynamic parameters for MB adsorption by DFM-10.
Table 3. Thermodynamic parameters for MB adsorption by DFM-10.
ΔHΔSΔG (kJ/mol)R2
(KJ/mol)(J/(mol·K))298 K308 K318 K
4.674716.8786−3.0115−4.2291−5.82630.9851
Table 4. Porous performance parameters of DF and DFM-10.
Table 4. Porous performance parameters of DF and DFM-10.
SimplesDFDFM-10
Mean pore radius (nm)10.535.45
Total pore volume (cm3 g−1)0.190.24
BET surface area (m2 g−1)60.55226.88
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cai, Z.; Liu, Q.; Li, H.; Wang, J.; Tai, G.; Wang, F.; Han, J.; Zhu, Y.; Wu, G. Waste-to-Resource Strategy to Fabricate Functionalized MOFs Composite Material Based on Durian Shell Biomass Carbon Fiber and Fe3O4 for Highly Efficient and Recyclable Dye Adsorption. Int. J. Mol. Sci. 2022, 23, 5900. https://doi.org/10.3390/ijms23115900

AMA Style

Cai Z, Liu Q, Li H, Wang J, Tai G, Wang F, Han J, Zhu Y, Wu G. Waste-to-Resource Strategy to Fabricate Functionalized MOFs Composite Material Based on Durian Shell Biomass Carbon Fiber and Fe3O4 for Highly Efficient and Recyclable Dye Adsorption. International Journal of Molecular Sciences. 2022; 23(11):5900. https://doi.org/10.3390/ijms23115900

Chicago/Turabian Style

Cai, Zhangzhen, Qi Liu, Haoxin Li, Jingyi Wang, Guoyu Tai, Fan Wang, Jiangang Han, Yongli Zhu, and Guangyu Wu. 2022. "Waste-to-Resource Strategy to Fabricate Functionalized MOFs Composite Material Based on Durian Shell Biomass Carbon Fiber and Fe3O4 for Highly Efficient and Recyclable Dye Adsorption" International Journal of Molecular Sciences 23, no. 11: 5900. https://doi.org/10.3390/ijms23115900

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop