Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access November 23, 2023

Synthesis, characterization, and application of the novel nanomagnet adsorbent for the removal of Cr(vi) ions

  • Norah Salem Alsaiari EMAIL logo , Majed Salem Alsaiari , Fatimah Mohammed Alzahrani , Abdelfattah Amari EMAIL logo and Mohamed A. Tahoon

Abstract

The synthesis of an efficient adsorbent to remove chromium ions from water is challenging. Therefore, in this study, a new nanomagnet composite (Fe3O4/biochar/ZIF-8) was synthesized by a one-pot hydrothermal method using a metal–organic framework (MOF, ZIF-8) as a sacrificial template, citrus peels as a source of biochar, and iron oxide nanoparticles for magnetization. The synthesized nanocomposite showed a high efficiency toward the adsorption of Cr(vi) ions. The adsorption study showed that the experimental data were well-described using the Langmuir isotherm model and pseudo-second-order model. According to the Langmuir model, the adsorption capacities toward Cr(vi) adsorption were 77 and 125 mg·g−1 for Fe3O4/biochar and Fe3O4/biochar/ZIF-8, respectively, indicating the role of MOF in improving the adsorption performance. The Fe3O4/biochar/ZIF-8 showed an excellent adsorption performance in the presence of coexisting ions at a wide pH range using different eluents to study reusability up to five successive cycles. We can conclude from this study that this nanoadsorbent is a promising material for removing pollutants from environmental water samples.

Graphical abstract

1 Introduction

The heavy metals, dyes, and pharmaceuticals pollution released from industrial effluents easily penetrate the water environment and cause a long-term threat to human and aquatic life. Particularly, dangerous heavy metals such as hexavalent chromium ions Cr(vi) cause severe problems [1]. Chromium (Cr) is a crucial element that is extensively used in a variety of sectors, including alloy production, leather tanning, and electroplating [2]. Naturally, chromium has two oxidation states, hexavalent and trivalent, of which the hexavalent state shows a higher toxicity due to its mutagenicity, carcinogenicity, and greater mobility [3]. On the other hand, the trivalent state can be utilized by the cell for many biological processes [4]. Therefore, the World Health Organization and environmental protection organizations in various nations have imposed tight restrictions on the amount of Cr ions in surface or drinking water [5,6]. As a result, it is now of utmost importance to effectively remove Cr(vi) ions from aqueous solutions. Several techniques have been applied for the removal of Cr(vi) ions from aqueous solutions such as ion exchange [7,8], membrane separation [9], phytoremediation [10], microbial remediation [11], and electrochemical reduction [12]. Adsorption technology is crucial in the removal of Cr(vi) ions owing to its exceptional advantages such as simplicity, high efficiency, cost-effectiveness, and flexibility [13] and uses a range of adsorbents, including carbon nanotubes [14,15], activated carbon [16], metal oxides [17], graphene oxide [18], and clay minerals [19]. However, the majority of adsorbents are not suitable for practical applications due to their poor selective characteristics and limited adsorption capabilities in hostile environments [20]. Therefore, the need for novel nanomaterials that can sustain strong adsorption performance in challenging practical environments is critical.

Metal–organic frameworks (MOFs) have attracted attention over the last few years due to their intriguing structural properties, such as variable pore size, programmable functionalities, abundant functional groups, and high specific surface areas [21]. These properties allowed their wide applications in different fields including environmental remediation [22], catalysis [23], supercapacitors [24], sensors [25], energy storage [26], and adsorption [27,28,29,30]. The zeolitic imidazolate framework-8 (ZIF-8) is a member of the MOF family of materials and has comparatively high thermal stability than other MOFs [31]. Through chelation interactions, the metal coordination site (Zn2+) in ZIF-8 may create metal complexes with the adsorbate, and the N heteroatom might enhance adsorption chelation sites [32]. However, the application of MOFs in water treatment is still restricted due to many factors such as the difficult and complicated recyclability of most MOFs, poor stability, and weak selectivity [33]. By encapsulating functional nanoparticles into MOFs, researchers can create MOF-containing nanocomposites that combine the benefits of two different components and improve synergistic effects in the resulting hybrid products. Subsequently, MOFs must be modified to be widely used. Having the potential for regeneration and reuse, biochar is a novel type of environmentally friendly material that is rich in carbon [34]. It can be made from a variety of substances, including industrial bio-wastes, marine and aquatic species, and agricultural and forestry leftovers (such as wood chips, rice straw/husk, fruit peels, leafy litter, etc.) [35,36,37,38]. In recent years, biochar has attracted considerable interest as an adsorbent from environmental media because of its inexpensive cost, relatively high porosity, porous qualities, possibility of regeneration, and reuse. For instance, biochar from crab shells could efficiently remove dye from water [39], human hair-derived biochar is utilized to adsorb tetracycline antibiotics [40], and biochar from coconut shells acts as an efficient adsorbent for the removal of methylene blue [41]. However, using traditional separation techniques like centrifugation, filtering, and gravity sedimentation to separate biochar is laborious, time-consuming, and expensive [42]. Subsequently, there is an urgent need for the magnetization of biochar to produce magnetic biochar adsorbent that exhibits a predominant future due to the easy separation using an external magnet instead of traditional methods [43]. For eliminating inorganic and organic contaminants from aquatic environments, magnetic biochar has been widely used. In general, ferric and ferrous compounds are used as precursors to precipitate and link magnetic media, such as zero-valent iron, magnetite (Fe3O4), and maghemite (γ-Fe2O3), onto biochar matrix [44,45]. With the addition of the biochar matrix and magnetic iron, the magnetic biochar typically displays a relatively high adsorption capability and stability [46]. Since iron oxide is more oxidatively stable than zero-valent iron, iron oxide-doped magnetic biochar has many advantages for real-world use. Due to their strong magnetic and adsorptive properties, Fe3O4 and γ-Fe2O3 are excellent candidates for the synthesis of iron oxide-doped magnetic biochar. Therefore, the preparation of MOF-magnetic biochar-based composites is a promising approach due to the ability to attain complementary effects between the 3-D MOFs and carbon-based materials. Due to the abundance of active sites, hierarchical porosity, and increased specific surface area, MOFs conjugated with biochar formed from naturally renewable biomass might significantly improve the physical and chemical performances of the biochar. The addition of biochar to the MOF can significantly enhance MOF agglomeration and increase the number of exposed active sites, which ultimately results in MOFs with a relatively high adsorption capacity and stability. Additionally, the recycling of adsorbents was aided by magnetic modification, and the ability of carbon-based materials to successfully prevent secondary contamination brought on by the dissolution of metal ions made them suitable for use in genuine large-scale applications.

The novelty of this study is the synthesis of a new magnetically modified MOF-based nanocomposite from three different components having a porous architecture that was characterized and investigated for efficient water treatment. Herein, a novel magnetic ZIF-8/biochar nanocomposite was prepared and investigated for the removal of Cr(vi) ions from an aqueous solution. To assess the removal efficiencies under various conditions, batch adsorption tests were carried out, and the mechanisms were also investigated. The particular points were to synthesize and characterize magnetic ZIF-8/biochar nanocomposite, assess and compare the adsorption capacities toward Cr(vi) ions using different adsorbents by studying the effect of key factors (pH and ionic strength) on adsorption capacity, explore the adsorption mechanisms through a series of experiments, such as kinetics, isotherms, and physicochemical characterization analysis, and investigate the regeneration and reusability of the adsorbent. The findings of this work would serve as a guide for the synthesis of functionalized biochar and the removal of Cr(vi) ions from contaminated water. It will also provide an effective approach for constructing new MOF-based adsorbents with excellent adsorption capacity, and attract MOF’s attention for the water treatment. Furthermore, it will help researchers to overcome the limitations of using MOF-based materials as adsorbents such as poor mechanical stability, difficulty of connecting MOFs with matrix, poor reusability, and high water solubility.

2 Materials and methods

2.1 Chemicals

The citrus fruits as carbon precursors were bought from local supermarkets (Mansoura, Egypt). Zinc nitrate hexahydrate (Zn(NO3)2·6H2O), 2-methylimidazole, and potassium hydroxide (KOH) were purchased from Sigma-Aldrich. Ammonium iron(iii) sulfate dodecahydrate (NH4Fe(SO4)2·12H2O) was supplied by Shanghai Chemical Reagent Co., Ltd. (Shanghai, China). Ammonium ferrous sulfate ((NH4)2Fe(SO4)2·6H2O) was obtained from Chongqing Beibei Chemical Reagent (Chongqing, China). Sodium hydroxide (NaOH), hydrochloric acid (HCl), potassium dichromate (K2Cr2O7), etc. were obtained from Sinopharm Chemical Reagent Co., Ltd. China. All chemicals used were of analytical grade and used without any additional purifications. A stock solution of K2Cr2O7 was prepared using distilled water and diluted for further experimental purposes.

2.2 Synthesis of ZIF-8

To synthesize MOFs (ZIF-8), the steps in Liu et al. [47] were followed with some modifications. Briefly, two separate volumes (40 mL) of methanol were used to dissolve Zn(NO3)2·6H2O (0.587 g) and 2-methylimidazole (1.298 g). Then, the metal solution was stirred for 1 h under sonication. Finally, the ligand solution was added to the metal solution at room temperature with sonication.

2.3 KOH-activated-citrus peel powder

The carbon precursor was extracted from citrus peels. First, citrus peels that had been cleaned were chopped into pieces and dried at 80°C before being processed into powder. Then, KOH (30%) was used to activate the citrus peel powder in a ratio of 1:2 (W/V) to KOH. It was allowed to stand for 30 min before being dried for 3 h at 70°C in an oxygen-free environment.

2.4 Synthesis of Fe3O4 nanoparticles

The magnetic Fe3O4 nanoparticles were synthesized by the co-precipitation method as described in the literature [48]. Briefly, 300 mL of distilled water was used to disperse 3.36 g of (NH4)2Fe(SO4)2·6H2O and 8.1 g of NH4Fe(SO4)2·12H2O at a temperature of 50°C under an N2 atmosphere. Then, ammonium solution (20 mL, 8 mol·L−1) was added drop by drop to the solution reaction to precipitate the magnetic particles. After that, the pH of the solution reaction was adjusted to 11 using NaOH solution under an ultrasonicator for 15 min. Using an external magnet, the precipitated product was separated and washed several times using distilled water until the pH of the filtrate became 7. Finally, the magnetic nanoparticles were dried at 45°C under a vacuum.

2.5 Synthesis of the magnetic ZIF-8/biochar nanocomposite

The nanocomposite of magnetic ZIF-8/biochar was synthesized by a one-pot hydrothermal method. Briefly, 100 mL of ZIF-8 suspensions were used to dissolve 10 g of Fe3O4 and KOH-activated citrus peel powder while stirring at room temperature for 13 h. After that, using an external magnet, the precipitate was collected and dried under a vacuum at 65°C. Then, the resulting product was inserted into a tube furnace for slow continuous pyrolysis under an N2 atmosphere at 600°C for 2 h using a rate of 5°C⋅min−1. The product was dipped in HCl solution (1 M) after cooling to room temperature to remove the impurities and alkali. Then, the product was washed using ethyl alcohol and water several times until the pH of the filtrate reached 7. The resulting composite was then dried and ground. The synthesized nanocomposite magnetic ZIF-8/biochar was then stored for the next experiments. Additionally, the composite Fe3O4/biochar was synthesized using the typical steps but without adding the ZIF-8 MOF during the synthesis.

2.6 Batch adsorption experiments

The efficiency of synthesized adsorbents for the removal of Cr(vi) ions was evaluated using batch adsorption experiments. In brief, 20 mL of Cr(vi) solution in the concentration range of 0–500 mg·L−1 was added to 0.05 g of adsorbent in a flask while the solution pH was 2. Using a shaking water bath, the solution was shaken for 24 h at 180 rpm. To study the kinetics of the adsorption process, 200 mL of 300 mg·L−1 Cr(vi) solution was added to 0.5 g of the adsorbent in a flask at pH 2 and the solution was shaken for 24 h at 180 rpm. Then, at different time intervals, part of the solution was withdrawn and analyzed for the presence of Cr(vi) ions and Cr using UV-vis spectrophotometer at λ = 540 nm after a complex formation with 1,5-diphenylcarbazide with a flame atomic absorption spectrometer, respectively. The difference between total Cr and Cr(vi) ions is used to calculate the concentration of Cr(iii) ions. Before each analysis and after the completion of the adsorption process, the adsorbent was collected using an external magnet. Additionally, the effect of different parameters on adsorption was investigated including ionic strength (0–0.1 M), pH (1:10), and 0.01 M competing ions ( PO 4 3 , SO 4 2 , Cl, Fe3+, Ca2+, and Na+). During these adsorption experiments, the solid/liquid ratio was fixed at 2.5 g·L−1 (20 mL solution). Moreover, desorption and reusability were investigated using different eluents including NaOH, NaHCO3, Na2EDTA, HNO3, and NaNO3 at a concentration of 0.01 M. In this experiment, 20 mL of Cr(vi) solution at pH 2 and a concentration of 50 mg·L−1 was mixed with the adsorbent, which was then collected with a magnet, washed with distilled water, and dipped in an eluent solution for 24 h to study the desorption efficiency. In the reusability experiment, the desorbed adsorbents using NaOH were used to adsorb Cr(vi) ions for five successive cycles to determine the adsorption capacity. All previously described adsorption experiments were performed in triplicate to approve the quality of results.

2.7 Characterization of synthesized adsorbents

To characterize the synthesized materials, different techniques have been used. The essential functional groups on the surface of synthesized materials were investigated using Fourier transform infrared spectroscopy (FT-IR; IRTracer-100 Shimadzu, Japan) in the range of 400–4,000 cm−1. The magnetic behavior of materials was identified using a vibrating sample magnetometer (VSM; Lake Shore 7404, USA). The thermal gravimetric analysis (TGA; STA409PC, DSC404F3, Germany) was used to investigate the thermal stability of the synthesized materials under an N2 environment in the range of 30–1,000°C at a heating rate of 10°C⋅min−1. Moreover, the surface morphology was investigated using a scanning electron microscope (SEM, Zeiss Genimi500, Zeiss, Germany). The Raman spectra were analyzed using a Raman spectrometer (Renishaw, UK). The specific surface area (S BET) was determined using the Brunauer–Emmett–Teller method via the adsorption/desorption isotherms of N2 at a temperature of 77 K. The pH drift method was used to calculate the point of zero charge (pHPZC) at which ΔpH = 0 (where ΔpH is the difference between the initial and final pH values).

3 Results and discussion

3.1 Characterization of the synthesized materials

To achieve the successful characterization of the synthesized materials, several common techniques have been applied. In this section, we used different techniques to verify the successful fabrication of the desired materials. First, the scanning electron microscope (SEM) was used to detect the morphologies of the synthesized materials and follow the surface change resulting from the added materials. Figure 1a shows the SEM image of the citrus peel-derived biochar that appeared to have several pores on its rough structure surface. Figure 1b shows the SEM image of the Fe3O4/biochar composite in which the composite has a stacked and smooth structure surface holding several nanoparticles with diameters in the range of 25–30 nm. The appeared nanoparticles to confirm the successful precipitation of magnetic Fe3O4 nanoparticles on the surface of the biochar.

Figure 1 
                  SEM images of the biochar (a), Fe3O4/biochar (b), and Fe3O4/biochar/ZIF-8 (c).
Figure 1

SEM images of the biochar (a), Fe3O4/biochar (b), and Fe3O4/biochar/ZIF-8 (c).

Figure 1c shows the SEM image of the Fe3O4/biochar/ZIF-8 composite. Figure 1c shows that the introduction of the ZIF-8 MOF to the composite has a great effect on the structure, which is clearly shown via the morphology change. The Fe3O4/biochar/ZIF-8 composite appeared to have wrinkled edges with smooth surfaces and a multi-layer structure, which could be the result of ZIF-8 decomposing at a higher pyrolysis temperature. Additionally, to sustain the framework, the partially decomposed ZIF-8 was distributed between the sheets that resemble graphene as a sacrifice template. Also, the crystallinity of the synthesized materials was studied using XRD patterns as shown in Figure 2a. According to the XRD patterns, the synthesized magnetic Fe3O4 nanoparticles showed the appearance of characteristic planes (440), (511), (422), (400), (311), and (220) at 2θ = 62.9°, 57.2°, 53.7°, 43.4°, 35.5°, and 30.2°, respectively [49]. When the XRD patterns of the Fe3O4 nanoparticles and Fe3O4/biochar composite were compared, it was noticed that the crystallinity remained unchanged, which indicated the stability of the Fe3O4/biochar composite crystal structure even after loading of magnetic nanoparticles on the biochar surface. After the addition of the ZIF-8 MOF to the Fe3O4/biochar composite, several new peaks appeared, which were attributed to the addition of the MOF. The results of the XRD patterns revealed the crystal structures of the Fe3O4 nanoparticles and Fe3O4/biochar composite, suggesting that the synthesized Fe3O4/biochar/ZIF-8 composite was well fabricated. The presence of functional groups and information about chemical bonding can be obtained using FT-IR spectra, as shown in Figure 2b. Generally, the characteristic peak at 1,621 cm−1 was attributed to the C═N bond in the ZIF-8 MOF or to the C═C bond in the aromatic ring. Also, the broad peaks at 3,421 cm−1 were attributed to the N–H stretching vibration or the O–H stretching vibration [50,51]. Additionally, the C–O stretching vibration of carboxylic acids and aromatic ethers is usually represented by the band at 1,095 cm−1. These oxygen-containing functional groups and double bonds of aromatic rings are very effective in pollutant removal via hydrogen bonding and π–π interactions, respectively [50,52,53]. Moreover, that ZIF-8 MOF was well introduced into the nanocomposite was confirmed from the appearance of the peak at 480 cm−1 that was attributed to Zn–N vibrations [54]. Also, that Fe3O4 was well introduced into the nanocomposite was confirmed from the appearance of the peak at 568 cm−1 that was attributed to Fe–O vibrations.

Figure 2 
                  XRD patterns (a), FT-IR spectra (b), and Raman spectra (c) of the synthesized materials.
Figure 2

XRD patterns (a), FT-IR spectra (b), and Raman spectra (c) of the synthesized materials.

Furthermore, the Raman spectra of the synthesized materials are shown in Figure 2c. In Raman spectra, there are two bands G- and D-bands, which are attributed to the ordered graphitic sp2 hybridized carbon and disordered sp3 hybridized carbon, respectively. The degree of graphitization can be obtained from the total peak area of the Raman spectra [55]. As a result, it is possible to use the peak area rate from the D-band to the G-band (I D/I G) as a guideline when determining the degree of structural disorder in carbon materials [56]. According to Figure 2c, the I D/I G of the Fe3O4/biochar is higher than that of the Fe3O4/biochar/ZIF-8 nanocomposite, indicating the increase of sp2-hybridized domains. The Raman spectra results indicated the successful introduction of the ZIF-8 MOF to the composite Fe3O4/biochar. Additionally, the magnetic properties of the synthesized materials were investigated using VSM, as shown in Figure 3a. According to Figure 3a, Fe3O4, Fe3O4/biochar, and Fe3O4/biochar/ZIF-8 have excellent magnetic properties. The saturation magnetization values were 24.53, 32.30, and 67 emu·g−1 for Fe3O4/biochar/ZIF-8, Fe3O4/biochar, and Fe3O4, respectively. According to the results, the saturation magnetization of pure Fe3O4 nanoparticles is the highest and this value was decreased after the incorporation of non-magnetic biochar and ZIF-8. However, by using an external magnet, these composites still exhibited magnetic properties and could be easily collected from the solution.

Figure 3 
                  Magnetic hysteresis loops (a), the N2 adsorption–desorption isotherms (b), and thermogravimetric analysis (TGA) (c) of the synthesized materials.
Figure 3

Magnetic hysteresis loops (a), the N2 adsorption–desorption isotherms (b), and thermogravimetric analysis (TGA) (c) of the synthesized materials.

As the specific surface area is very important in determining the adsorption properties of the materials, the N2 adsorption/desorption isotherm method was used for the investigation of the S BET surface area of synthesized materials [57] as shown in Figure 3b. According to Figure 3b, the synthesized biochar showed an S BET of 8.97 m2⋅g−1. The addition of magnetic nanoparticles to the biochar showed an increase in the S BET to reach 280.55 m2⋅g−1. Additionally, further addition of MOF ZIF-8 to the magnetic biochar increased the S BET to reach 940.25 m2⋅g−1. The S BET of Fe3O4/biochar/ZIF-8 is about two times higher than that of Fe3O4/biochar. It appears that the addition of ZIF-8 facilitated the expansion of the surface area. The Fe3O4/biochar/ZIF-8 nanocomposite showed a combined feature of type I and IV phenomena, indicating the co-occurrence of micropores and mesopores that interpret the enhanced surface area. Additionally, the Fe3O4/biochar/ZIF-8 nanocomposite shows an H4-type hysteresis loop at a relative pressure p/p 0 > 0.4, indicating the higher content of mesopores. In contrast, at a relative pressure p/p 0 < 0.01, the importance of micropores is very clear as the adsorption capacity was significantly increased. The mesoporous and microporous construction will aid in the improved migration of the adsorbate particles via the porous arrangement, increasing the capillary effect and improving the adsorption behavior of nanocomposites [58]. Therefore, the N2 adsorption/desorption isotherms showed the significant effect of ZIF-8 MOF on the nanocomposite. The thermal stability of synthesized materials was determined using TGA analysis in the temperature range from 30 to 1,000°C, as shown in Figure 3c. According to the TGA curves of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar composites, the adsorbed surface water was evaporated in the temperature range from 30 to 100°C; however, the decomposition of functional groups in the composites was observed in the temperature range from 100 to 780°C. It is clear that the weight. loss of the Fe3O4/biochar/ZIF-8 nanocomposite was higher than that of the Fe3O4/biochar in the temperature range from 100 to 780°C by about 11%, which was attributed to the evaporation of the ZIF-8 MOF from the surface of the nanocomposite. The Fe3O4/biochar composite showed a rapid weight loss at 780°C, resulting from the decomposition of its carbon backbone. At a temperature higher than 900°C, the weight loss in the Fe3O4/biochar/ZIF-8 nanocomposite was attributed to the decomposition of the metallic Zn from the MOF. Therefore, the Fe3O4/biochar/ZIF-8 nanocomposite has a higher thermal stability than the Fe3O4/biochar composite, which was attributed to the ZIF-8 MOF acting as a sacrificial template. All characterization results indicated the good fabrication of the nanocomposite Fe3O4/biochar/ZIF-8.

3.2 Adsorption study

3.2.1 Adsorption isotherm and kinetics

To understand the adsorption behavior of Cr(vi) ions on the surface of the synthesized nanocomposites, the adsorption isotherm and kinetics were studied. Figure 4a shows the adsorption isotherms in which the initial concentration of Cr(vi) ions was varied with the adsorption capacities of the prepared adsorbents. For both nanocomposites, Fe3O4/biochar/ZIF-8 and Fe3O4/biochar as adsorbents, the increase in the Cr(vi) initial concentration caused a rapid increase of the adsorption capacities until an equilibrium was reached at which the adsorption remains unchanged. Figure 4a shows the fitting of experimental adsorption data using the Freundlich equation – Eq. (1) – and Langmuir equation – Eq. (2) – for the better understanding of isotherms:

(1) q e = K F C e 1 / 2 ,

(2) q e = Q max K L C e 1 + K L C e ,

where q e is the adsorption capacity (mg·g−1) at equilibrium, K F is the Freundlich equilibrium adsorption constant (mg1−n ·L n ·g−1), K L is the Langmuir equilibrium adsorption constant (L·mg−1), and Q max is the maximum adsorption capacity (mg·g−1). The Freundlich and Langmuir isotherm coefficients are presented in Table 1.

Figure 4 
                     Adsorption isotherms models (a) and adsorption kinetics models (b) for the removal of Cr(vi) ions on the surface of the synthesized materials (initial concentration = 300 mg·L−1, pH = 2, temperature = 25°C, adsorbent dose = 0.5 g, and contact time = 24 h).
Figure 4

Adsorption isotherms models (a) and adsorption kinetics models (b) for the removal of Cr(vi) ions on the surface of the synthesized materials (initial concentration = 300 mg·L−1, pH = 2, temperature = 25°C, adsorbent dose = 0.5 g, and contact time = 24 h).

Table 1

Adsorption isotherm and kinetic parameters for the adsorption of Cr(vi) ions on the surface of the synthesized materials

Adsorbent Langmuir Freundlich Pseudo-first order Pseudo-second order
Parameters q m (mg·g−1) K L (L·mg−1) R 2 N K F (mg·g−1) (L·mg)−1/n R 2 K 1 (min−1) q e R 2 K 2 (g·mg.h−1) q e R 2
Fe3O4/biochar 77.0 0.15 0.998 4.77 25.6 0.844 0.03 67.7 0.930 0.0005 68.7 0.997
Fe3O4/biochar/ZIF-8 125.0 0.54 0.998 6.50 61.0 0.768 0.07 111.0 0.981 0.0010 111.0 0.999

According to Figure 4a, the adsorption isotherms of Cr(vi) on the surface of both nanocomposites were better defined using the Langmuir model than the Freundlich model, which is also from the R 2 values presented in Table 1. This indicated that the adsorption of metal on the surface of adsorbents occurred as a monolayer over the identical adsorption sites [59,60]. The coefficients presented in Table 1 indicate that the maximum adsorption capacities according to the Langmuir isotherm were 125 and 77 mg·g−1 for Fe3O4/biochar/ZIF-8 and Fe3O4/biochar, respectively. The results of adsorption experiments show that even without ZIF-8 doping, the Fe3O4/biochar nanocomposite has a good adsorption capacity, which is attributed to the porous structure, large specific surface area, abundant mineral components, and surface functional groups. For comparison, Table 2 shows the maximum adsorption capacities of different MOF adsorbents in the Cr(vi) removal from water. The adsorption capacity of Fe3O4/biochar/ZIF-8 toward the removal of Cr(vi) ions is higher than many previously prepared adsorbents. The adsorption of Cr(vi) ions on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites was favorable, as indicated by the separation factor (R L) values (from 0–1) [61,62]. Moreover, the stronger adsorption affinity of Fe3O4/biochar/ZIF-8 toward Cr(vi) ions than Fe3O4/biochar was evident from the higher R L values. This was because the introduction of the MOF (ZIF-8) to the nanocomposite provided more adsorption active sites. Additionally, the adsorption kinetics was investigated using pseudo-first- (Eq. (3)) and pseudo-second-order kinetics (Eq. (4)):

(3) q t = q e ( 1 e k 1 t ) ,

(4) q t = k 2 q e 2 t 1 + k 2 q e t ,

where q t (mg·g−1) represents the adsorption capacity of the adsorbent at the reaction time t, k 1 (min−1) represents the rate constant of the pseudo-first-order kinetics, and k 2 (g·mg−1·min−1) represents the rate constant of pseudo-second-order kinetics. The pseudo-first- and pseudo-second-order kinetic model coefficients are presented in Table 1.

Table 2

Comparison of the Fe3O4/biochar/ZIF-8 nanocomposite with previously reported adsorbents for the adsorption of Cr(vi) ions

Adsorbent pH Removal capacity (mg·g−1) Reference
Fe3O4/biochar/ZIF-8 2.0 125.0 This study
Fe3O4@MIL-100 2.0 18.0 [67]
MIL-100 2.0 30.4 [68]
Amine-MIL-101 4.0 44.0 [69]
Mn-UiO-66 32.77 [70]
MOF BUC-17 4.0 121.0 [71]
GO-CS@MOF 3.0 144.9 [72]
ZIF-67 5.0 13.3 [73]
ZIF-8 7.0 0.15 [74]
Cu-BTC 7.0 40.0 [75]

The effect of time on the removal of Cr(vi) ions on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites is shown in Figure 4b. Similar adsorption behavior was observed for both adsorbents. According to Figure 4b, the adsorption equilibrium for the removal of Cr(vi) ions on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites was achieved at 300 and 500 min, respectively. Under high concentrations of Cr(vi) ions, it was a quick initial stage of adsorption aided by ion diffusion. After this stage, the rate of adsorption was decreased due to the saturation of adsorption sites with metal adsorbate ions. The effect of time indicated a faster adsorption of Cr(vi) ions on the Fe3O4/biochar/ZIF-8 than Fe3O4/biochar because the presence of a higher number of abundant functional groups facilitated the rapid uptake of ions. According to Table 1, the adsorption of Cr(vi) ions on Fe3O4/biochar/ZIF-8 than the Fe3O4/biochar adsorbent was best described using a pseudo-second-order kinetic model as indicated by the correlation coefficient (R 2) values. This indicated that the adsorption of both adsorbents was achieved using chemical and physical methods [63]. In the physical method, the diffusion of ions was achieved within the pores of nanocomposites; whereas, in the chemical method, by complexation of ions was achieved by the functional groups and reduction. This kinetic behavior of Cr(vi) adsorption was observed with previously reported adsorbents [64,65,66].

3.2.2 Effect of competing ions and the ionic strength

To understand the efficiency of any adsorbent for the removal of contaminants, the effect of competing ions should be investigated to determine the selectivity of the adsorbent toward targeted ions. Therefore, the competition between Cr(vi) ions and competing ions for the adsorption active sites on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar was investigated, as shown in Figure 5a. According to Figure 5a, the adsorption of Cr(vi) ions on the surface of the Fe3O4/biochar nanocomposite was slightly affected by the presence of Ca2+, Fe3+, and Cl ions, as well as slightly affected by the presence of PO 4 3 and SO 4 2 .

Figure 5 
                     The effect of 0.01 M competing ions (a) and the effect of the ionic strength (b) on the adsorption of Cr(vi) ions on the surface of the synthesized materials.
Figure 5

The effect of 0.01 M competing ions (a) and the effect of the ionic strength (b) on the adsorption of Cr(vi) ions on the surface of the synthesized materials.

Also, the adsorption of Cr(vi) ions on the surface of Fe3O4/biochar/ZIF-8 was significantly affected by the competing ions and the effect was in the order SO 4 2 > Fe3+ > PO 4 3 > Ca2+ > Cl. The SO 4 2 ions had a higher effect on the removal of Cr(vi) ions than PO 4 3 ions. This is attributed to the less net charge of PO 4 3 than Cr2O7 2− at pH 2, as it exists as HPO4 2− and H3PO4 while SO 4 2 carries two negative charges, such as Cr2O7 2−, with the same molecular structure producing a higher competition for adsorption sites [76,77]. The complexation of Cr(vi) oxyanions with Fe3+ and Ca2+ ions was the reason for the high effect on the adsorption process [78]. Another important factor affecting the adsorption of Cr(vi) ions is the ionic strength, as shown in Figure 5b, in which the adsorption was studied on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites at different concentrations of NO3 ions. According to Figure 5b, the increase of NO3 concentration caused a decrease in the adsorption efficiency of the Fe3O4/biochar/ZIF-8 nanocomposite. The increase of NO3 concentration to 0.1 mg·L−1 caused a decrease in adsorption efficiency by 39%. However, for the Fe3O4/biochar nanocomposite, the reduction in the adsorption efficiency was only 11%. The main cause of the decreased Cr(vi) removal at an increased ionic strength was the competition between Cr(vi) and NO3 ions for the adsorption sites [79].

3.2.3 Effect of pH values

It is well-known that one of the key elements that greatly affects the removal of Cr(vi) from aqueous solution is the pH value [80] because the solution pH also influences the charge of the adsorbent surface in addition to the species and redox potential of Cr(vi) that are already present [81,82]. Therefore, the effect of solution pH on the removal of Cr(vi) from aqueous solution on the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites was investigated, as shown in Figure 6a.

Figure 6 
                     The effect of pH (a) and the point of zero charge (pHpzc) (b) for the adsorption of Cr(vi) ions on the surface of the synthesized materials.
Figure 6

The effect of pH (a) and the point of zero charge (pHpzc) (b) for the adsorption of Cr(vi) ions on the surface of the synthesized materials.

While the adsorption behavior between Fe3O4/biochar/ZIF-8 and Fe3O4/biochar nanocomposites was different, the solution pH clearly had a great effect on the removal of Cr(vi) ions. According to Figure 6a, Fe3O4/biochar nanocomposites exhibited a limited pH range for the removal of Cr(vi) ions (pH = 1–4) while the Fe3O4/biochar/ZIF-8 nanocomposite exhibited a wide pH range (pH = 2–10) for the Cr(vi) adsorption. For the Fe3O4/biochar nanocomposite, the increased pH value from 1 to 3 caused an increased adsorption efficiency toward Cr(vi) ions, and this adsorption efficiency was decreased at pH 7. For the Fe3O4/biochar/ZIF-8 nanocomposite, the adsorption capacity was still high even after the pH increased up to 10, and higher than that of the Fe3O4/biochar nanocomposite by about 31%. However, the increased pH value also reduced the adsorption efficiency of the pollutant over both the adsorbents. To better understand the pH effect, the point of zero charges (pHpzc) of the two prepared adsorbents was investigated, as shown in Figure 6b. According to Figure 6b, the pHpzc values were 9.15 and 5.76 for Fe3O4/biochar and Fe3O4/biochar/ZIF-8, respectively. The higher value of Fe3O4/biochar/ZIF-8 was attributed to the introduced MOF. After adsorption, pH variations in the solution were also noticed. The pH of the solution after adsorption on the surface of Fe3O4/biochar was a little lower than the pHpzc value, thereby reducing its adsorption efficiency. Apart from the changes in the Cr(vi) percentage with the increasing solution pH, the electrostatic repulsion between the negatively charged Fe3O4/biochar/ZIF-8 surface and Cr(vi) anions was the primary cause of the decrease in the adsorption efficiency at pH > 6. Additionally, the higher adsorption efficiency of Fe3O4/biochar/ZIF-8 at low pH values than Fe3O4/biochar was attributed to the protonated functional groups on its surface. When the adsorption solution was analyzed after adsorption at low pH values, the presence of Cr(iii) ions was approved, indicating the reduction of Cr(vi) ions. This reduction of Cr(vi) ions to Cr(iii) ions is well known with the biochar adsorbents due to their role as electron donors and electron acceptors as discussed in the literature [83,84]. Subsequently, the removal of Cr(vi) ions on the surfaces of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar was achieved via different mechanisms.

3.2.4 Regeneration and reusability study

Regeneration and reusability study for any adsorbent is very important economically as it decreases the overall cost of the treatment process [85,86,87]. Before reusing the adsorbents, the desorption of adsorbed pollutants should be investigated first. Therefore, many eluents have been explored to desorb Cr(vi) ions from the surface of the synthesized Fe3O4/biochar/ZIF-8 and Fe3O4/biochar as adsorbents, as shown in Figure 7a.

Figure 7 
                     The desorption of total Cr ions using different eluents (a) and the reusability study (b) for the adsorption of Cr(vi) ions on the surface of the synthesized materials.
Figure 7

The desorption of total Cr ions using different eluents (a) and the reusability study (b) for the adsorption of Cr(vi) ions on the surface of the synthesized materials.

According to Figure 7a, the desorption efficiency toward Cr(vi) ions from the surface of Fe3O4/biochar/ZIF-8 was in the order of NaNO3 < NaHCO3 < NaOH < HNO3 < Na2EDTA, whereas for Fe3O4/biochar, the desorption efficiency toward Cr(vi) ions was in the order of NaOH < HNO3 < NaNO3 < NaHCO3 < Na2EDTA, as shown in Figure 7a. It is clear that the sodium EDTA salt has the highest desorption capacity toward Cr(vi) ions. The acids HCl and HNO3 showed good desorption capacities toward Cr(vi) ions from the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar, which was similar to the literature [88,89]. However, the commonly used desorption eluent for Cr(vi) ions is the basic solution of NaOH [90,91]. According to the desorption experimental results, the Na2EDTA solution was used as an eluent to desorb Cr(vi) ions from the surface of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar during the reusability study. The reusability of Fe3O4/biochar/ZIF-8 and Fe3O4/biochar for the removal of Cr(vi) ions was investigated for five successive adsorption–desorption cycles as shown in Figure 7b. According to Figure 7b, the Fe3O4/biochar/ZIF-8 showed higher reusability results than Fe3O4/biochar. For both adsorbents, the removal efficiency was the highest in the first cycle due to the availability of vacant adsorption sites. After that, the removal efficiency was decreased cycle by cycle until it reached its minimum in the fifth cycle due to the destroyed adsorption sites resulting after each cycle. The removal efficiency of Fe3O4/biochar/ZIF-8 toward Cr(vi) removal remained higher than 90.0% even after the last cycle, whereas for Fe3O4/biochar, the removal efficiency was decreased gradually from 90.5 to 45% in the last cycle. The reusability results indicated that Fe3O4/biochar/ZIF-8 is a promising adsorbent to eliminate pollutants from water with cost-effective properties.

3.2.5 Suggested mechanism of adsorption

To study the mechanism of adsorption of Cr(vi) ions on the surface of Fe3O4/biochar/ZIF-8, XPS spectra of Fe3O4/biochar/ZIF-8 before and after adsorption were recorded, as shown in Figure 8a and b, respectively, to provide information regarding the surface chemistry of the adsorbent. Comparing Figure 8a and b showed a change in the metal (Zn) binding energy after and before adsorption, indicating that metal sites were involved in the chelation process. Also, the XPS spectra after adsorption showed the appearance of a new binding energy peak at 580 eV, which belongs to Cr(vi) ions, confirming their presence on the surface of the adsorbent. Additionally, the identification of the adsorbed species of chromium ions was achieved using high-resolution XPS spectra as shown in Figure 8c. According to Figure 8c, the high-resolution XPS spectra showed the presence of two satellite bands of Cr(iii) and Cr(vi) ions at 587 and 577 eV, respectively. Subsequently, the presence of two chromium species Cr(iii) and Cr(vi) on the surface of Fe3O4/biochar/ZIF-8 indicated the removal of Cr(vi) ions on the adsorbent surface by adsorption and by the chemical reduction to Cr(iii). In this regard, the mechanism of Cr(vi) removal can be explained as follows: the protonated nitrogen and oxygen functional groups on the surface of adsorbent under acidic conditions can adsorb Cr(vi) species (HCrO4 ) via electrostatic interactions and chemical bonding. This chemical bonding on the surface of adsorbent was also associated with the chemical reduction of Cr(vi) to Cr(iii). The chemical bonding also occurred through the metal sites (Zn) of the MOFs. The resulting Cr(iii) ions could be removed by complexation with functional groups on the adsorbent surface and others may be dissolved in water.

Figure 8 
                     The XPS survey spectra of Fe3O4/biochar/ZIF-8 before adsorption (a) and after adsorption (b). XPS high-resolution Cr 2p spectra of Fe3O4/biochar/ZIF-8 after the removal of Cr(vi) ions (c), and XRD patterns of Fe3O4/biochar/ZIF-8 after adsorption (d).
Figure 8

The XPS survey spectra of Fe3O4/biochar/ZIF-8 before adsorption (a) and after adsorption (b). XPS high-resolution Cr 2p spectra of Fe3O4/biochar/ZIF-8 after the removal of Cr(vi) ions (c), and XRD patterns of Fe3O4/biochar/ZIF-8 after adsorption (d).

The study of adsorbent stability is very important. Therefore, the stability of Fe3O4/biochar/ZIF-8 as an adsorbent was determined by XRD after adsorption, as shown in Figure 8d. According to Figure 8d and comparing with the results of XRD before adsorption, there is no change in the XRD patterns before and after adsorption. These results indicate the good stability of Fe3O4/biochar/ZIF-8 as an adsorbent. To confirm this stability of Fe3O4/biochar/ZIF-8 as an adsorbent, the concentration of Fe and C in the solution after equilibrium was measured. Carbon was found to be 0.51–0.71% and Fe was found to be 0.2–0.3% under acidic conditions. These results also confirmed the high stability of Fe3O4/biochar/ZIF-8 and the gained stability by MOF modification.

The present MOF-based adsorbent has several advantages such as excellent designability with the control of pore size, pore type, adsorption sites, and surface area that affect the adsorption capacity. This MOF-based adsorbent with a controlled chemical design and pore structure allows its application in water treatment. Another advantage is the reusability with high adsorption efficiency, which reduces the overall cost of treatment so it is cost-effective. Also, it is very stable and easily collected from the reaction medium due to its magnetic properties. It can capture pollutants with different mechanisms due to the diversity of functional groups on its surface. It is very selective and can be used in the presence of competing ions and at a wide pH range. However, like any adsorbent material, it has some disadvantages such as the difficulty of connecting the MOFs with the matrix, high cost of synthesis, long time synthesis, and high accuracy and skill during the synthesis.

4 Conclusions

In summary, Fe3O4/biocha, and Fe3O4/biochar/ZIF-8 nanocomposites were synthesized by the one-pot hydrothermal method. The citrus fruits as a source of biochar were mixed with magnetic nanoparticles and MOFs to synthesize a novel nanocomposite of Fe3O4/biochar/ZIF-8. The nanocomposite was well characterized using different techniques including SEM, XRD, FT-IR, Raman spectroscopy, the N2 adsorption–desorption isotherm, TGA, and magnetization curves. All these results showed an excellent combination between the three components of the nanocomposite. The nanocomposite was employed as a recyclable ecological adsorbent for the effective removal of Cr(vi) ions from the aqueous solution. The Fe3O4/biochar/ZIF-8 nanocomposite showed improved adsorption capacities toward the examined pollutant. The adsorption data were well-described using the Langmuir isotherm model and pseudo-second-order kinetic model, indicating monolayer and physicochemical adsorption. The adsorption capacities were found to be 77 and 125 mg·g−1 for Fe3O4/biochar and Fe3O4/biochar/ZIF-8, respectively, according to the Langmuir model. The results indicated the role of MOFs as a matrix that holds the nanoparticles and improves their properties. The effect of different parameters on adsorption has been investigated, including pH, ionic strength, and competing ions. The results showed that the studied material has a good adsorption capacity at different ionic strength values and in the presence of different competing ions, indicating its applicability in real water treatment. Moreover, the Fe3O4/biochar/ZIF-8 nanocomposite showed excellent reusability toward Cr(vi) removal. The removal mechanism of Cr(vi) ions on the surface of the nanocomposite was studied and the results indicated the removal of ions by chemical bonding with the zinc metal, nitrogen, and oxygen centers, chemical reduction of Cr(vi) ions to Cr(iii) ions, and electrostatic interaction. The XRD results of the nanocomposite after adsorption showed its good stability as an adsorbent. We can conclude that the Fe3O4/biochar/ZIF-8 nanocomposite is a promising recyclable adsorbent for the removal of pollutants from water.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the research groups program (grant no. RGP.2/34/44). This research was also funded by the Princess Nourah bint Abdulrahman University Researchers Supporting Project (number PNURSP2023R19), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

  1. Funding information: This research was funded by the Deanship of Scientific Research at King Khalid University (grant no. RGP.2/34/44). Additionally, this research was funded by the Princess Nourah bint Abdulrahman University Researchers Supporting Project (number PNURSP2023R19), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

  2. Author contributions: Conceptualization, N.S.A. and F.M.A.; methodology, K.M.K.; software, M.S.A.; validation, F.M.A., K.M.K. and M.S.A.; formal analysis, F.M.A.; investigation, K.M.K.; resources, K.M.K.; data curation, F.M.A.; writing – original draft preparation, F.M.A.; writing – review and editing, K.M.K.; visualization, A.A.; supervision, M.A.T.; project administration, A.A.; funding acquisition, N.S.A. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Data availability statement: The datasets generated and/or analyzed during the current study are available from the corresponding author upon reasonable request.

References

[1] Cherdchoo, W., S. Nithettham, and J. Charoenpanich. Removal of Cr (VI) from synthetic wastewater by adsorption onto coffee ground and mixed waste tea. Chemosphere, Vol. 221, 2019, pp. 758–767.10.1016/j.chemosphere.2019.01.100Search in Google Scholar PubMed

[2] Niu, J., P. Ding, X. Jia, G. Hu, and Z. Li. Study of the properties and mechanism of deep reduction and efficient adsorption of Cr (VI) by low-cost Fe3O4-modified ceramsite. Science of the Total Environment, Vol. 688, 2019, pp. 994–1004.10.1016/j.scitotenv.2019.06.333Search in Google Scholar PubMed

[3] Han, S., Y. Zang, Y. Gao, Q. Yue, P. Zhang, W. Kong, et al. Co-monomer polymer anion exchange resin for removing Cr (VI) contaminants: Adsorption kinetics, mechanism and performance. Science of The Total Environment, Vol. 709, 2020, id. 136002.10.1016/j.scitotenv.2019.136002Search in Google Scholar PubMed

[4] Coetzee, J. J., N. Bansal, and E. M. Chirwa. Chromium in environment, its toxic effect from chromite-mining and ferrochrome industries, and its possible bioremediation. Exposure and Health, Vol. 12, 2020, pp. 51–62.10.1007/s12403-018-0284-zSearch in Google Scholar

[5] Mortada, W. I., A. El-Naggar, A. Mosa, K. N. Palansooriya, B. Yousaf, R. Tang, et al. Biogeochemical behaviour and toxicology of chromium in the soil-water-human nexus: A review. Chemosphere, Vol. 331, 2023, id. 138804.10.1016/j.chemosphere.2023.138804Search in Google Scholar PubMed

[6] Mahmoud, M. E., S. M. Elsayed, S. E. M. Mahmoud, R. O. Aljedaani, and M. A. Salam. Recent advances in adsorptive removal and catalytic reduction of hexavalent chromium by metal–organic frameworks composites. Journal of Molecular Liquids, Vol. 347, 2022, id. 118274.10.1016/j.molliq.2021.118274Search in Google Scholar

[7] Xie, Y., J. Lin, J. Liang, M. Li, Y. Fu, and H. Wang. Hypercrosslinked mesoporous poly (ionic liquid) s with high density of ion pairs: Efficient adsorbents for Cr (VI) removal via ion-exchange. Chemical Engineering Journal, Vol. 378, 2019, id. 122107.10.1016/j.cej.2019.122107Search in Google Scholar

[8] Tavakoli, O., V. Goodarzi, M. R. Saeb, N. M. Mahmoodi, and R. Borja. Competitive removal of heavy metal ions from squid oil under isothermal condition by CR11 chelate ion exchanger. Journal of Hazardous Materials, Vol. 334, 2017, pp. 256–266.10.1016/j.jhazmat.2017.04.023Search in Google Scholar PubMed

[9] He, P. Y., Y. J. Zhang, H. Chen, Z. C. Han, and L. C. Liu. Low-cost and facile synthesis of geopolymer-zeolite composite membrane for chromium (VI) separation from aqueous solution. Journal of Hazardous Materials, Vol. 392, 2020, id. 122359.10.1016/j.jhazmat.2020.122359Search in Google Scholar PubMed

[10] Masinire, F., D. O. Adenuga, S. M. Tichapondwa, and E. M. Chirwa. Phytoremediation of Cr (VI) in wastewater using the vetiver grass (Chrysopogon zizanioides). Minerals Engineering, Vol. 172, 2021, id. 107141.10.1016/j.mineng.2021.107141Search in Google Scholar

[11] Banerjee, S, B. Kamila, S. Barman, S. R. Joshi, T. Mandal, G. Halder. Interlining Cr (VI) remediation mechanism by a novel bacterium Pseudomonas brenneri isolated from coalmine wastewater. Journal of Environmental Management, Vol. 233, 2019, pp. 271–282.10.1016/j.jenvman.2018.12.048Search in Google Scholar PubMed

[12] Yao, F., M. Jia, Q. Yang, K. Luo, F. Chen, Y. Zhong, et al. Electrochemical Cr (VI) removal from aqueous media using titanium as anode: Simultaneous indirect electrochemical reduction of Cr (VI) and in-situ precipitation of Cr (III). Chemosphere, Vol. 260, 2020, id. 127537.10.1016/j.chemosphere.2020.127537Search in Google Scholar PubMed

[13] Damiri, F., S. Andra, N. Kommineni, S. K. Balu, R. Bulusu, A. A. Boseila, et al. Recent advances in adsorptive nanocomposite membranes for heavy metals ion removal from contaminated water: A comprehensive review. Materials, Vol. 15, 2022, id. 5392.10.3390/ma15155392Search in Google Scholar PubMed PubMed Central

[14] Mpouras, T., A. Polydera, D. Dermatas, N. Verdone, and G. Vilardi. Multi wall carbon nanotubes application for treatment of Cr (VI)-contaminated groundwater; Modeling of batch & column experiments. Chemosphere, Vol. 269, 2021, id. 128749.10.1016/j.chemosphere.2020.128749Search in Google Scholar PubMed

[15] Mousavi, S. R., M. Asghari, and N. M. Mahmoodi. Chitosan-wrapped multiwalled carbon nanotube as filler within PEBA thin film nanocomposite (TFN) membrane to improve dye removal. Carbohydrate Polymers, Vol. 237, 2020, id. 116128.10.1016/j.carbpol.2020.116128Search in Google Scholar PubMed

[16] Tu, B., R. Wen, K. Wang, Y. Cheng, Y. Deng, W. Cao, et al. Efficient removal of aqueous hexavalent chromium by activated carbon derived from Bermuda grass. Journal of Colloid and Interface Science, Vol. 560, 2020, pp. 649–658.10.1016/j.jcis.2019.10.103Search in Google Scholar PubMed

[17] Cheng, Q., C. Wang, K. Doudrick, and C. K. Chan. Hexavalent chromium removal using metal oxide photocatalysts. Applied Catalysis B: Environmental, Vol. 176, 2015, pp. 740–748.10.1016/j.apcatb.2015.04.047Search in Google Scholar

[18] Mahmoodi, N. M., M. Ghezelbash, M. Shabanian, F. Aryanasab, and M. R. Saeb. Efficient removal of cationic dyes from colored wastewaters by dithiocarbamate-functionalized graphene oxide nanosheets: From synthesis to detailed kinetics studies. Journal of the Taiwan Institute of Chemical Engineers, Vol. 81, 2017, pp. 239–246.10.1016/j.jtice.2017.10.011Search in Google Scholar

[19] Ashour, E. A. and M. A. Tony. Eco-friendly removal of hexavalent chromium from aqueous solution using natural clay mineral: activation and modification effects. SN Applied Sciences, Vol. 2, 2020, pp. 1–13.10.1007/s42452-020-03873-xSearch in Google Scholar

[20] Geng, Z., Y. Lin, X. Yu, Q. Shen, L. Ma, Z. Li, et al. Highly efficient dye adsorption and removal: a functional hybrid of reduced graphene oxide–Fe3O4 nanoparticles as an easily regenerative adsorbent. Journal of Materials Chemistry, Vol. 22, 2012, pp. 3527–3535.10.1039/c2jm15544cSearch in Google Scholar

[21] Zhu, L., D. Shen, C. Xu, X. Dai, M. A. Silver, J. Li, et al. Identifying the recognition site for selective trapping of 99TcO4–in a hydrolytically stable and radiation resistant cationic metal–organic framework. Journal of the American Chemical Society, Vol. 139, 2017, pp. 14873–14876.Search in Google Scholar

[22] Wang, Q., Q. Gao, A. M. Al-Enizi, A. Nafady, and S. Ma. Recent advances in MOF-based photocatalysis: environmental remediation under visible light. Inorganic Chemistry Frontiers, Vol. 7, 2020, pp. 300–339.10.1039/C9QI01120JSearch in Google Scholar

[23] Goetjen, T. A., J. Liu, Y. Wu, J. Sui, X. Zhang, J. T. Hupp, et al. Metal–organic framework (MOF) materials as polymerization catalysts: a review and recent advances. Chemical Communications, Vol. 56, 2020, pp. 10409–10418.10.1039/D0CC03790GSearch in Google Scholar

[24] Ajdari, F. B., E. Kowsari, M. N. Shahrak, A. Ehsani, Z. Kiaei, H. Torkzaban, et al. A review on the field patents and recent developments over the application of metal organic frameworks (MOFs) in supercapacitors. Coordination Chemistry Reviews, Vol. 422, 2020, id. 213441.10.1016/j.ccr.2020.213441Search in Google Scholar

[25] Koo, W.-T., J.-S. Jang, and I.-D. Kim. Metal-organic frameworks for chemiresistive sensors. Chem, Vol. 5, 2019, pp. 1938–1963.10.1016/j.chempr.2019.04.013Search in Google Scholar

[26] Ren, J., Y. Huang, H. Zhu, B. Zhang, H. Zhu, S. Shen, et al. Recent progress on MOF‐derived carbon materials for energy storage. Carbon Energy, Vol. 2, 2020, pp. 176–202.10.1002/cey2.44Search in Google Scholar

[27] Chen, B., Y. Li, M. Li, M. Cui, W. Xu, L. Li, et al. Rapid adsorption of tetracycline in aqueous solution by using MOF-525/graphene oxide composite. Microporous and Mesoporous Materials, Vol. 328, 2021, id. 111457.10.1016/j.micromeso.2021.111457Search in Google Scholar

[28] Alsaiari, N. S., H. Osman, A. Amari, and M. A. Tahoon. The synthesis of metal–organic-framework-based ternary nanocomposite for the adsorption of organic dyes from aqueous solutions. Magnetochemistry, Vol. 8, 2022, id. 133.10.3390/magnetochemistry8100133Search in Google Scholar

[29] Alzahrani, F. M., N. S. Alsaiari, K. M. Katubi, A. Amari, and M. A. Tahoon. Synthesis, characterization, and application of magnetized lanthanum (III)-based metal-organic framework for the organic dye removal from water. Adsorption Science & Technology, Vol. 2022, 2022, id. 3513829.10.1155/2022/3513829Search in Google Scholar

[30] Amari, A., F. M. Alzahrani, N. S. Alsaiari, K. M. Katubi, F. B. Rebah, and M. A. Tahoon. Magnetic metal organic framework immobilized laccase for wastewater decolorization. Processes, Vol. 9, 2021, id. 774.10.3390/pr9050774Search in Google Scholar

[31] Li, K., N. Miwornunyuie, L. Chen, H. Jingyu, P. S. Amaniampong, D. Ato Koomson, et al. Sustainable application of ZIF-8 for heavy-metal removal in aqueous solutions. Sustainability, Vol. 13, 2021, id. 984.10.3390/su13020984Search in Google Scholar

[32] Yu, F., X. Bai, M. Liang, and J. Ma. Recent progress on metal-organic framework-derived porous carbon and its composite for pollutant adsorption from liquid phase. Chemical Engineering Journal, Vol. 405, 2021, id. 126960.10.1016/j.cej.2020.126960Search in Google Scholar

[33] Wu, Y., B. Li, X. Wang, S. Yu, H. Pang, Y. Liu, et al. Magnetic metal-organic frameworks (Fe3O4@ ZIF-8) composites for U (VI) and Eu (III) elimination: simultaneously achieve favorable stability and functionality. Chemical Engineering Journal, Vol. 378, 2019, id. 122105.10.1016/j.cej.2019.122105Search in Google Scholar

[34] de Almeida, L. S., E. Q. Oreste, J. V. Maciel, M. G. Heinemann, and D. Dias. Electrochemical devices obtained from biochar: advances in renewable and environmentally-friendly technologies applied to analytical chemistry. Trends in Environmental. Analytical Chemistry, Vol. 26, 2020, id. e00089.10.1016/j.teac.2020.e00089Search in Google Scholar

[35] Zhou, Y., S. Qin, S. Verma, T. Sar, S. Sarsaiya, B. Ravindran, et al. Production and beneficial impact of biochar for environmental application: A comprehensive review. Bioresource Technology, Vol. 337, 2021, id. 125451.10.1016/j.biortech.2021.125451Search in Google Scholar PubMed

[36] Zhang, A., X. Li, J. Xing, and G. Xu. Adsorption of potentially toxic elements in water by modified biochar: A review. Journal of Environmental Chemical Engineering, Vol. 8, 2020, id. 104196.10.1016/j.jece.2020.104196Search in Google Scholar

[37] Xiang, W., X. Zhang, J. Chen, W. Zou, F. He, X. Hu, et al. Biochar technology in wastewater treatment: A critical review. Chemosphere, Vol. 252, 2020, id. 126539.10.1016/j.chemosphere.2020.126539Search in Google Scholar PubMed

[38] Oladipo, A. A., E. O. Ahaka, and M. Gazi. High adsorptive potential of calcined magnetic biochar derived from banana peels for Cu2+, Hg2+, and Zn2+ ions removal in single and ternary systems. Environmental Science and Pollution Research, Vol. 26, 2019, pp. 31887–31899.10.1007/s11356-019-06321-5Search in Google Scholar PubMed

[39] Dai, L., W. Zhu, L. He, F. Tan, N. Zhu, Q. Zhou, et al. Calcium-rich biochar from crab shell: an unexpected super adsorbent for dye removal. Bioresource Technology, Vol. 267, 2018, pp. 510–516.10.1016/j.biortech.2018.07.090Search in Google Scholar PubMed

[40] Ahmed, M., M. A. Islam, M. Asif, and B. Hameed. Human hair-derived high surface area porous carbon material for the adsorption isotherm and kinetics of tetracycline antibiotics. Bioresource Technology, Vol. 243, 2017, pp. 778–784.10.1016/j.biortech.2017.06.174Search in Google Scholar PubMed

[41] Islam, M. A., M. Ahmed, W. Khanday, M. Asif, and B. Hameed. Mesoporous activated coconut shell-derived hydrochar prepared via hydrothermal carbonization-NaOH activation for methylene blue adsorption. Journal of Environmental Management, Vol. 203, 2017, pp. 237–244.10.1016/j.jenvman.2017.07.029Search in Google Scholar PubMed

[42] Li, R., J. J. Wang, B. Zhou, M. K. Awasthi, A. Ali, Z. Zhang, et al. Recovery of phosphate from aqueous solution by magnesium oxide decorated magnetic biochar and its potential as phosphate-based fertilizer substitute. Bioresource Technology, Vol. 215, 2016, pp. 209–214.10.1016/j.biortech.2016.02.125Search in Google Scholar PubMed

[43] Zhou, Y., S. Cao, C. Xi, X. Li, L. Zhang, G. Wang, et al. A novel Fe3O4/graphene oxide/citrus peel-derived bio-char based nanocomposite with enhanced adsorption affinity and sensitivity of ciprofloxacin and sparfloxacin. Bioresource Technology, Vol. 292, 2019, id. 121951.10.1016/j.biortech.2019.121951Search in Google Scholar PubMed

[44] Han, Y., X. Cao, X. Ouyang, S. P. Sohi, and J. Chen. 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, Vol. 145, 2016, pp. 336–341.10.1016/j.chemosphere.2015.11.050Search in Google Scholar PubMed

[45] Thines, K., E. Abdullah, N. Mubarak, and M. Ruthiraan. Synthesis of magnetic biochar from agricultural waste biomass to enhancing route for waste water and polymer application: a review. Renewable and Sustainable Energy Reviews, Vol. 67, 2017, pp. 257–276.10.1016/j.rser.2016.09.057Search in Google Scholar

[46] Zhou, Y., G. Liu, J. Liu, Y. Xiao, T. Wang, and Y. Xue. Magnetic biochar prepared by electromagnetic induction pyrolysis of cellulose: Biochar characterization, mechanism of magnetization and adsorption removal of chromium (VI) from aqueous solution. Bioresource Technology, Vol. 337, 2021, id. 125429.10.1016/j.biortech.2021.125429Search in Google Scholar PubMed

[47] Liu, Q., B. Zhou, M. Xu, and G. Mao. Integration of nanosized ZIF-8 particles onto mesoporous TiO 2 nanobeads for enhanced photocatalytic activity. RSC Advances, Vol. 7, 2017, pp. 8004–8010.10.1039/C6RA28277FSearch in Google Scholar

[48] Chen, J.-Y., S.-R. Cao, C.-X. Xi, Y. Chen, X.-L. Li, L. Zhang, et al. A novel magnetic β-cyclodextrin modified graphene oxide adsorbent with high recognition capability for 5 plant growth regulators. Food Chemistry, Vol. 239, 2018, pp. 911–919.10.1016/j.foodchem.2017.07.013Search in Google Scholar PubMed

[49] Deng, H., X. Li, Q. Peng, X. Wang, J. Chen and Y. Li. Monodisperse magnetic single‐crystal ferrite microspheres. Angewandte Chemie International Edition, Vol. 44, 2005, pp. 2782–2785.10.1002/anie.200462551Search in Google Scholar PubMed

[50] Tran, H. N., F. Tomul, N. T. H. Ha, D. T. Nguyen, E. C. Lima, G. T. Le, et al. Innovative spherical biochar for pharmaceutical removal from water: Insight into adsorption mechanism. Journal of Hazardous Materials, Vol. 394, 2020, id. 122255.10.1016/j.jhazmat.2020.122255Search in Google Scholar PubMed

[51] Liang, C., Y. Tang, X. Zhang, H. Chai, Y. Huang, and P. Feng. ZIF-mediated N-doped hollow porous carbon as a high performance adsorbent for tetracycline removal from water with wide pH range. Environmental Research, Vol. 182, 2020, id. 109059.10.1016/j.envres.2019.109059Search in Google Scholar PubMed

[52] Zhang, Y., T. Zhang, C. Guo, J. Lv, Z. Hua, S. Hou, et al. Drugs of abuse and their metabolites in the urban rivers of Beijing, China: occurrence, distribution, and potential environmental risk. Science of the Total Environment, Vol. 579, 2017, pp. 305–313.10.1016/j.scitotenv.2016.11.101Search in Google Scholar PubMed

[53] Li, K., P. Du, Z. Xu, T. Gao, and X. Li. Occurrence of illicit drugs in surface waters in China. Environmental Pollution, Vol. 213, 2016, pp. 395–402.10.1016/j.envpol.2016.02.036Search in Google Scholar PubMed

[54] Bedin, K. C., A. L. Cazetta, I. P. Souza, O. Pezoti, L. S. Souza, P. S. Souza, et al. Porosity enhancement of spherical activated carbon: Influence and optimization of hydrothermal synthesis conditions using response surface methodology. Journal of Environmental Chemical Engineering, Vol. 6, 2018, pp. 991–999.10.1016/j.jece.2017.12.069Search in Google Scholar

[55] Feng, D., Y. Zhao, Y. Zhang, S. Sun, S. Meng, Y. Guo, et al. Effects of K and Ca on reforming of model tar compounds with pyrolysis biochars under H2O or CO2. Chemical Engineering Journal, Vol. 306, 2016, pp. 422–432.10.1016/j.cej.2016.07.065Search in Google Scholar

[56] Sui, Z.-Y., X. Li, Z.-Y. Sun, H.-C. Tao, P.-Y. Zhang, L. Zhao, et al. Nitrogen-doped and nanostructured carbons with high surface area for enhanced oxygen reduction reaction. Carbon, Vol. 126, 2018, pp. 111–118.10.1016/j.carbon.2017.10.003Search in Google Scholar

[57] Zhu, L., N. Zhao, L. Tong, Y. Lv, and G. Li. Characterization and evaluation of surface modified materials based on porous biochar and its adsorption properties for 2, 4-dichlorophenoxyacetic acid. Chemosphere, Vol. 210, 2018, pp. 734–744.10.1016/j.chemosphere.2018.07.090Search in Google Scholar PubMed

[58] Cazetta, A. L., O. Pezoti, K. C. Bedin, T. L. Silva, A. Paesano Junior, T. Asefa, et al. Magnetic activated carbon derived from biomass waste by concurrent synthesis: efficient adsorbent for toxic dyes. ACS Sustainable Chemistry & Engineering, Vol. 4, 2016, pp. 1058–1068.10.1021/acssuschemeng.5b01141Search in Google Scholar

[59] Zhang, Z., H. Li, J. Li, X. Li, Z. Wang, X. Liu, et al. A novel adsorbent of core-shell construction of chitosan-cellulose magnetic carbon foam: synthesis, characterization and application to remove copper in wastewater. Chemical Physics Letters, Vol. 731, 2019, id. 136573.10.1016/j.cplett.2019.07.001Search in Google Scholar

[60] Kong, D., L. He, H. Li, and Z. Song. Preparation and characterization of graphene oxide/chitosan composite aerogel with high adsorption performance for Cr (VI) by a new crosslinking route. Colloids and Surfaces A: Physicochemical and Engineering Aspects, Vol. 625, 2021, id. 126832.10.1016/j.colsurfa.2021.126832Search in Google Scholar

[61] Gupta, S. and A. Kumar. Removal of nickel (II) from aqueous solution by biosorption on A. barbadensis Miller waste leaves powder. Applied Water. Science, Vol. 9, 2019, pp. 1–11.10.1007/s13201-019-0973-1Search in Google Scholar

[62] Zhao, P., N. Liu, C. Jin, H. Chen, Z. Zhang, L. Zhao, et al. UiO-66: an advanced platform for investigating the influence of functionalization in the adsorption removal of pharmaceutical waste. Inorganic Chemistry, Vol. 58, 2019, pp. 8787–8792.10.1021/acs.inorgchem.9b01172Search in Google Scholar PubMed

[63] Han, Q., J. Wang, B. A. Goodman, J. Xie, and Z. Liu. High adsorption of methylene blue by activated carbon prepared from phosphoric acid treated eucalyptus residue. Powder Technology, Vol. 366, 2020, pp. 239–248.10.1016/j.powtec.2020.02.013Search in Google Scholar

[64] Liu, H., F. Zhang, and Z. Peng. Adsorption mechanism of Cr (VI) onto GO/PAMAMs composites. Scientific Reports, Vol. 9, 2019, id. 3663.10.1038/s41598-019-40344-9Search in Google Scholar PubMed PubMed Central

[65] Mei, J., H. Zhang, Z. Li, and H. Ou. A novel tetraethylenepentamine crosslinked chitosan oligosaccharide hydrogel for total adsorption of Cr (VI). Carbohydrate Polymers, Vol. 224, 2019, id. 115154.10.1016/j.carbpol.2019.115154Search in Google Scholar PubMed

[66] Niu, C., N. Zhang, C. Hu, C. Zhang, H. Zhang, and Y. Xing. Preparation of a novel citric acid-crosslinked Zn-MOF/chitosan composite and application in adsorption of chromium (VI) and methyl orange from aqueous solution. Carbohydrate Polymers, Vol. 258, 2021, id. 117644.10.1016/j.carbpol.2021.117644Search in Google Scholar PubMed

[67] Yang, Q., Q. Zhao, S. Ren, Q. Lu, X. Guo, and Z. Chen. Fabrication of core-shell Fe3O4@ MIL-100 (Fe) magnetic microspheres for the removal of Cr (VI) in aqueous solution. Journal of Solid State Chemistry, Vol. 244, 2016, pp. 25–30.10.1016/j.jssc.2016.09.010Search in Google Scholar

[68] Forghani, M., A. Azizi, M. J. Livani, and L. A. Kafshgari. Adsorption of lead (II) and chromium (VI) from aqueous environment onto metal-organic framework MIL-100 (Fe): Synthesis, kinetics, equilibrium and thermodynamics. Journal of Solid State Chemistry, Vol. 291, 2020, id. 121636.10.1016/j.jssc.2020.121636Search in Google Scholar

[69] Jalayeri, H., P. Aprea, D. Caputo, A. Peluso, and F. Pepe. Synthesis of amino-functionalized MIL-101 (Cr) MOF for hexavalent chromium adsorption from aqueous solutions. Environmental Nanotechnology, Monitoring & Management, Vol. 14, 2020, id. 100300.10.1016/j.enmm.2020.100300Search in Google Scholar

[70] Yang, Z.-h., J. Cao, Y.-p. Chen, X. Li, W.-p. Xiong, Y.-y. Zhou, et al. Mn-doped zirconium metal-organic framework as an effective adsorbent for removal of tetracycline and Cr (VI) from aqueous solution. Microporous and Mesoporous Materials, Vol. 277, 2019, pp. 277–285.10.1016/j.micromeso.2018.11.014Search in Google Scholar

[71] Guo, J., J.-J. Li, and C.-C. Wang. Adsorptive removal of Cr (VI) from simulated wastewater in MOF BUC-17 ultrafine powder. Journal of Environmental Chemical Engineering, Vol. 7, 2019, id. 102909.10.1016/j.jece.2019.102909Search in Google Scholar

[72] Samuel, M. S., V. Subramaniyan, J. Bhattacharya, C. Parthiban, S. Chand, and N. P. Singh. A GO-CS@ MOF [Zn (BDC)(DMF)] material for the adsorption of chromium (VI) ions from aqueous solution. Composites Part B: Engineering, Vol. 152, 2018, pp. 116–125.10.1016/j.compositesb.2018.06.034Search in Google Scholar

[73] Li, X., X. Gao, L. Ai, and J. Jiang. Mechanistic insight into the interaction and adsorption of Cr (VI) with zeolitic imidazolate framework-67 microcrystals from aqueous solution. Chemical Engineering Journal, Vol. 274, 2015, pp. 238–246.10.1016/j.cej.2015.03.127Search in Google Scholar

[74] Niknam Shahrak, M., M. Ghahramaninezhad, and M. Eydifarash. Zeolitic imidazolate framework-8 for efficient adsorption and removal of Cr (VI) ions from aqueous solution. Environmental Science and Pollution Research, Vol. 24, 2017, pp. 9624–9634.10.1007/s11356-017-8577-5Search in Google Scholar PubMed

[75] Maleki, A., B. Hayat, M. Naghizade, and S. W. Joo. Adsorption of hexavalent chromium by metal organic frameworks from aqueous solution. Journal of Industrial and Engineering Chemistry, Vol. 28, 2015, pp. 211–216.10.1016/j.jiec.2015.02.016Search in Google Scholar

[76] Zhang, X., X. Lin, Y. He, Y. Chen, J. Zhou, and X. Luo. Adsorption of phosphorus from slaughterhouse wastewater by carboxymethyl konjac glucomannan loaded with lanthanum. International Journal of Biological Macromolecules, Vol. 119, 2018, pp. 105–115.10.1016/j.ijbiomac.2018.07.140Search in Google Scholar PubMed

[77] Lin, H., S. Han, Y. Dong, and Y. He. The surface characteristics of hyperbranched polyamide modified corncob and its adsorption property for Cr (VI). Applied Surface Science, Vol. 412, 2017, pp. 152–159.10.1016/j.apsusc.2017.03.061Search in Google Scholar

[78] Dong, L., J. Liang, Y. Li, S. Hunang, Y. Wei, X. Bai, et al. Effect of coexisting ions on Cr (VI) adsorption onto surfactant modified Auricularia auricula spent substrate in aqueous solution. Ecotoxicology and Environmental Safety, Vol. 166, 2018, pp. 390–400.10.1016/j.ecoenv.2018.09.097Search in Google Scholar PubMed

[79] Hu, Z., L. Cai, J. Liang, X. Guo, W. Li, and Z. Huang. Green synthesis of expanded graphite/layered double hydroxides nanocomposites and their application in adsorption removal of Cr (VI) from aqueous solution. Journal of Cleaner Production, Vol. 209, 2019, pp. 1216–1227.10.1016/j.jclepro.2018.10.295Search in Google Scholar

[80] Jin, W., H. Du, S. Zheng, and Y. Zhang. Electrochemical processes for the environmental remediation of toxic Cr (VI): A review. Electrochimica Acta, Vol. 191, 2016, pp. 1044–1055.10.1016/j.electacta.2016.01.130Search in Google Scholar

[81] Liu, F., S. Hua, C. Wang, M. Qiu, L. Jin, and B. Hu. Adsorption and reduction of Cr (VI) from aqueous solution using cost-effective caffeic acid functionalized corn starch. Chemosphere, Vol. 279, 2021, id. 130539.10.1016/j.chemosphere.2021.130539Search in Google Scholar PubMed

[82] Yang, Q., H. Wang, F. Li, Z. Dang, and L. Zhang. Rapid and efficient removal of Cr (vi) by a core–shell magnetic mesoporous polydopamine nanocomposite: roles of the mesoporous structure and redox-active functional groups. Journal of Materials Chemistry A, Vol. 9, 2021, pp. 13306–13319.10.1039/D1TA02475BSearch in Google Scholar

[83] Xu, X., H. Huang, Y. Zhang, Z. Xu, and X. Cao. Biochar as both electron donor and electron shuttle for the reduction transformation of Cr (VI) during its sorption. Environmental Pollution, Vol. 244, 2019, pp. 423–430.10.1016/j.envpol.2018.10.068Search in Google Scholar PubMed

[84] Liu, L., X. Liu, D. Wang, H. Lin, and L. Huang. Removal and reduction of Cr (Ⅵ) in simulated wastewater using magnetic biochar prepared by co-pyrolysis of nano-zero-valent iron and sewage sludge. Journal of Cleaner Production, Vol. 257, 2020, id. 120562.10.1016/j.jclepro.2020.120562Search in Google Scholar

[85] Baskar, A. V., N. Bolan, S. A. Hoang, P. Sooriyakumar, M. Kumar, L. Singh, et al. Recovery, regeneration and sustainable management of spent adsorbents from wastewater treatment streams: A review. Science of the Total Environment, Vol. 822, 2022, id. 153555.10.1016/j.scitotenv.2022.153555Search in Google Scholar PubMed

[86] Hacıosmanoğlu, G. G., C. Mejías, J. Martín, J. L. Santos, I. Aparicio, and E. Alonso. Antibiotic adsorption by natural and modified clay minerals as designer adsorbents for wastewater treatment: A comprehensive review. Journal of Environmental Management, Vol. 317, 2022, id. 115397.10.1016/j.jenvman.2022.115397Search in Google Scholar PubMed

[87] Praveen, S., R. Gokulan, T. B. Pushpa, and J. Jegan. Techno-economic feasibility of biochar as biosorbent for basic dye sequestration. Journal of the Indian Chemical Society, Vol. 98, 2021, id. 100107.10.1016/j.jics.2021.100107Search in Google Scholar

[88] Dobrzyńska, J., A. Wysokińska, and R. Olchowski. Raspberry stalks-derived biochar, magnetic biochar and urea modified magnetic biochar-Synthesis, characterization and application for As (V) and Cr (VI) removal from river water. Journal of Environmental Management, Vol. 316, 2022, id. 115260.10.1016/j.jenvman.2022.115260Search in Google Scholar PubMed

[89] Bayuo, J., M. A. Abukari, and K. B. Pelig-Ba. Desorption of chromium (VI) and lead (II) ions and regeneration of the exhausted adsorbent. Applied Water Science, Vol. 10, 2020, id. 171.10.1007/s13201-020-01250-ySearch in Google Scholar

[90] Ogata, F., N. Nagai, R. Itami, T. Nakamura, and N. Kawasaki. Potential of virgin and calcined wheat bran biomass for the removal of chromium (VI) ion from a synthetic aqueous solution. Journal of Environmental Chemical Engineering, Vol. 8, 2020, id. 103710.10.1016/j.jece.2020.103710Search in Google Scholar

[91] Bouchoum, H., D. Benmoussa, A. Jada, M. Tahiri, and O. Cherkaoui. Synthesis of amidoximated polyacrylonitrile fibers and its use as adsorbent for Cr (VI) ions removal from aqueous solutions. Environmental Progress & Sustainable Energy, Vol. 38, 2019, id. 13196.10.1002/ep.13196Search in Google Scholar

Received: 2023-02-13
Revised: 2023-08-22
Accepted: 2023-10-24
Published Online: 2023-11-23

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 30.4.2024 from https://www.degruyter.com/document/doi/10.1515/rams-2023-0145/html
Scroll to top button