Next Article in Journal
Exploration of the Functional Constituents of the Substrate of Flammulina velutipes
Next Article in Special Issue
Mineral Heterostructures for Simultaneous Removal of Lead and Arsenic Ions
Previous Article in Journal
Solid–Liquid Phase Equilibria of the Aqueous Quaternary System Rb+, Cs+, Mg2+//SO42− - H2O at T = 323.2 K
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Eco-Friendly Green Approach to the Biosorption of Hazardous Dyes from Aqueous Solution on Ragweed (Ambrosia artemisiifolia) Biomass

by
Natalija Nedić
1,
Tamara Tadić
1,
Bojana Marković
1,
Aleksandra Nastasović
1,
Aleksandar Popović
2 and
Sandra Bulatović
1,*
1
Institute of Chemistry, Technology and Metallurgy, University of Belgrade, Njegoševa 12, 11000 Belgrade, Serbia
2
Faculty of Chemistry, University of Belgrade, Studentski Trg 12-16, 11000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Separations 2024, 11(11), 310; https://doi.org/10.3390/separations11110310
Submission received: 1 October 2024 / Revised: 23 October 2024 / Accepted: 25 October 2024 / Published: 28 October 2024
(This article belongs to the Special Issue Materials from Biomass and Waste for Adsorption Applications)

Abstract

:
The presented research includes the preparation, characterization, and implementation of magnetic biosorbent (Fe3O4/RWB), obtained from ragweed (Ambrosia artemisiifolia) biomass. Fe3O4/RWB was examined for the removal of a hazardous dye, malachite green (MG), from an aqueous solution in a batch system. The effects of the experimental parameters—initial dye concentration (10–300 mg/L), contact time (0–120 min), biosorbent dose (1–5 g/L), initial pH (2–10), ionic strength (0–1 mol/L), and temperature (298–318 K) on dye biosorption—were studied. The results showed that increases in biosorbent dose, contact time, and initial pH led to an increase in biosorption efficiency, while the increase in initial dye concentration, the ionic strength, and temperature had the opposite effect. The biosorption kinetics for MG on Fe3O4/RWB were analyzed with pseudo-first-order, pseudo-second-order, and Elovich kinetic models, while the Langmuir, Freundlich and Temkin isotherm models were used for equilibrium data analysis. It was observed that the MG biosorption followed the pseudo-second-order kinetic model, whereas the Langmuir model was the best fit for the equilibrium biosorption data of MG, with a Qmax of 34.1 mg/g. the desorption of MG from Fe3O4/RWB indicated reusability in five adsorption/desorption cycles, good performance, and potential in practical applications.

Graphical Abstract

1. Introduction

Industrialization is associated with the disposal of large amount of toxic pollutants, such as synthetic dyes, which can have negative effects on the environment and human health [1]. Around the world, about 7 × 105 tons of synthetic dyes are produced every year, and 2/3 of this amount are used in the textile industry. During the textile-dyeing process, about 30% of synthetic dyes are wasted through washing and exhaustion procedures and up to 2 × 105 tons of these chemicals are disposed as wastewaters in the environment [2]. Environmental pollution with synthetic dyes has become a focus in recent decades because dyes can have negative ecotoxicological effects, with a tendency to bioaccumulate. Many of these compounds are very toxic and can have negative effects on public health, especially when present in water [3]. Malachite green (MG) is an organic and toxic triphenylmethane compound [4]. It is used in aquaculture, textile, pharmaceutical, and food processing industries, among others. Because of its solubility and water stability, MG is highly resistant to microbial degradation [5]. Due to its application in many industrial areas, MG can enter the environment relatively easily, mostly affecting aquatic ecosystems. Human exposure to MG could result in mutagenesis, carcinogenesis, respiratory diseases, chromosomal fracture, etc. In many developed countries the usage of MG has been banned, or is heavily regulated [4,6]. Conventional methods for removing synthetic dyes from wastewater include adsorption, ion exchange, photocatalysis, membrane filtration, precipitation, and oxidation [7,8]. These methods require the excessive use of chemicals and other resources, and are often expensive. The effectiveness of these methods often depends on many factors, including the secondary treatment of pollutants not removed by the primary treatment [9]. Therefore, it would be desirable to develop an efficient and economical method for the removal of synthetic dyes in order to prevent environmental pollution and health risks.
Many researchers have studied the use of biomass in wastewater treatments. Biomass contains cellulose, lignin, hemicellulose, lipids, and proteins, with functional groups (alcohol, carboxylic, amino, phenol groups, etc.), making biosorbents attractive materials for dye removal [10]. The use of biosorption has rapidly expanded because of its advantages, such as its availability, sustainability, reusability, eco-friendliness, and low cost [11]. Various biosorbents have been applied in the biosorption of dyes, including microorganisms, forestry, wood wastes, etc. [12,13,14].
Ambrosia artemisiifolia is one of the most important invasive plants, and has spread widely in America, Asia, and Europe [15]. In Europe, approximately 13 million people suffer from allergies caused by Ambrosia, with medical treatments costing about EUR seven billion annually [16]. This plant is also widespread in the Republic of Serbia, especially in its northern regions [17]. According to the currently available data, the maximum concentration of Ambrosia pollen in the air in Serbia was 12,356 grain/m3, making it dominant in comparison with the other allergens [18]. Aside from being allergenic, A. artemisiifolia is, as already mentioned, also known as an invasive plant, causing significant yield losses in many crops [15]. Although invasive, the use of A. artemisiifolia as a biosorbent for hazardous chemicals has been a subject of interest by several researchers [19,20,21,22]. Still, there is a deficiency of literature data, leaving open many possibilities for testing the potential of this plant as a biosorbent. In this study, a magnetic biosorbent (Fe3O4/RWB) obtained from a cheap and eco-friendly A. artemisiifolia biomass was tested as an MG biosorbent in an aqueous solution in a batch system (Figure 1).
The characterization of the obtained magnetic biosorbent (Fe3O4/RWB) was carried out using FT-IR, SEM-EDS, XRD, and the determination of pHpzc. The effects of different parameters on the biosorption process’s efficiency, including the initial dye concentration, contact time, biosorbent dose, initial pH, ionic strength, and temperature, were assessed. The biosorption kinetics were determined using the pseudo-first-order (PFO), pseudo-second-order (PSO), and Elovich models. Three isotherm models, Langmuir, Freundlich, and Temkin, were used to examine equilibrium sorption data. Thermodynamic studies were also performed. Finally, the desorption of MG and reusability of the magnetic biosorbent (Fe3O4/RWB) were also analyzed. The ability of a magnetic biosorbent to regenerate and reuse was tested in several cycles of the adsorption/desorption process. This low-cost approach, with the minimal use of chemicals, could lead to significant savings, does not involve the formation of secondary waste, and is in line with the principles of green chemistry. Progress in this research area could be a driving force for further investigations, with an impact on industry (new businesses), economy (circular economy), and society (environmental and health sustainability).

2. Materials and Methods

2.1. Materials and Instrumentation

For the purposes of this research, the following chemicals were used: iron (II) chloride (FeCl2∙4H2O, >98%), iron (III) chloride (FeCl3∙6H2O, ≥97%), Malachite Green (MG, C23H25ClN2), sodium hydroxide (NaOH), hydrochloric acid (HCl), and sodium chloride (NaCl). All chemicals were purchased from Sigma Aldrich (Saint Louis, MO, USA). Fe3O4/RWB was characterized via Nicolet SUMMIT FT-IR Spectrometer (Thermo Fisher Scientific, Waltham, MA, USA), in ATR mode, in the range 4000–400 cm−1, scanning Electron Microscopy with Energy-Dispersive X-ray Spectroscopy (SEM-EDS) (JEOL JSM-6390 LV, JEOL Ltd., Tokyo, Japan), and an EDS detector (Oxford instruments X-MaxN, High Wycombe, UK). The X-ray diffraction (XRD) analysis (Rigaku Ultima IV diffractometer, Tokyo, Japan) of Fe3O4/RWB was also performed in the range of 10–80° 2θ. The Point of Zero Charge (pHpzc) of the Fe3O4/RWB was determined according to the pH drift method. The pH values were measured using a pH meter (Hanna HI 2210, Hanna Instruments, Navi Mumbai, India). UV–Vis spectroscopy (NOVEL-102S, COLOLab Experts, Polje ob Sotli, Slovenia) was used for determination of the dye concentrations. During the experiments, the samples were mixed in an orbital shaker (Orb-Pro Labbox Labware, Barcelona, Spain). All measurements in this research were performed in triplicate.

2.2. Sample Collection and Pretreatment of A. artemisiifolia Biomass

A. artemisiifolia samples were obtained from the outskirts of Belgrade (capital of the Republic of Serbia), by cutting off plant’s parts above the soil surface. After sampling, the plant material was washed with tap water and dried in sunlight for seven days. After drying, the raw plant material was cut into small pieces and ground in an electric mill. The resulting particles were sieved to obtain particles with the size < 300 µm. They were washed with distilled water to remove probable impurities and dried in an oven at 323 K.

2.3. Preparation of Magnetic Biosorbent (Fe3O4/RWB)

The magnetic biosorbent was prepared using the modified coprecipitation method proposed by Medina-Zazueta et al. [23]. The method consists of mixing 3.9 g of FeCl3∙6H2O and 1.7 g of FeCl2∙4H2O in 190 mL of deionized water under a stream of N2, at 353 K, with stirring (450 rpm, 20 min). After that, a few drops of 10 mol/L NaOH were added to the reaction mixture for magnetite precipitation [24], and stirring was continued under the same conditions (for 30 min). A total of 2 g of biomass (RWB) was added to the reaction mixture and stirring was continued (for 30 min). Finally, the prepared Fe3O4/RWB was filtered and washed with deionized water until the pH was neutral and dried in an oven at 323 K.

2.4. Biosorption Studies

The effects of biosorbent dose (1–5 g/L), contact time (0–120 min), initial dye concentration (10–300 mg/L), initial pH (2–10), ionic strength (0–1 mol/L), and temperature (298–318 K) on MG biosorption on Fe3O4/RWB were analyzed in a batch system. During the typical biosorption process, 5 g/L of the Fe3O4/RWB was added to a vial flask with 20 mL of a 50 mg/L MG solution. The samples were shaken for 60 min at 400 rpm. After biosorption, Fe3O4/RWB from the samples was separated using an external magnet, and after decanting, the remaining MG concentration was measured using a UV–Vis spectrophotometer at λmax = 618 nm. MG removal efficiency (R%) and the biosorption capacity (Qt) were calculated according to Equations (1) and (2) [25,26]:
R % = C 0 C e C 0 × 100
Q t   = V ( C 0 C t ) m        
where C0, Ce, and Ct (mg/L) are MG concentrations in the solution at the initial, equilibrium, and any time t, respectively. V (L) is the solution volume and m (g) is the biosorbent mass.
For the regeneration/reusability analysis, Fe3O4/RWB (5 g/L) was added to the MG dye solution (C0 = 50 mg/L; contact time = 60 min; shaking speed = 400 rpm). After MG biosorption, the MG-loaded Fe3O4/RWB was separated using an external magnet. Desorption agents (20 mL), comprising absolute ethanol [27], 0.1 mol/L HCl [28], and 0.1 mol/L NaOH [29], were added to the MG-loaded Fe3O4/RWB and shaken for 60 min at 400 rpm. The samples were separated using an external magnet and the concentrations of MG were measured using a UV-Vis spectrophotometer at λmax = 618 nm. The desorption percentage (D%) of MG was determined following Equation (3) [30]. The best desorbing agent after the first cycle was absolute ethanol, and this was used for the remaining regeneration/reusability analysis.
D % = C d e s C a d s × 100
where Cdes and Cads (mg/L) are the desorbed and adsorbed MG concentration, respectively.

3. Results and Discussion

3.1. Characterization of Fe3O4/RWB

The FT-IR spectra of Fe3O4/RWB before and after MG biosorption are shown in Figure 2. For biosorbent materials, adsorption peaks from 3500 to 3200 cm−1, representing the stretching vibrations of the –NH2 groups, are typical. In the FTIR spectrum of Fe3O4/RWB, the peak from 3100 cm−1 to 3400 cm−1 is also related to –OH stretching vibrations from macromolecules of cellulose, hemicellulose, and lignin [3,31,32]. The peaks at 1641.5 cm−1 and 1419.6 cm−1 could be attributed to (C = C) vibrations in aromatic rings [33]. The peak at 1026.5 cm−1 is distinctive for vibrations of –COH [34,35]. The peak at 579.9 cm−1 refers to magnetite, which is incorporated into the structure of Fe3O4/RWB [36]. Distinctive variations were observed in the FT-IR spectrum of Fe3O4/RWB–MG. In addition to the changes in existing peaks, new specific peaks were also detected in the fingerprint region (500–1600 cm−1) after MG biosorption. It was found that the peak at 3400 cm−1, which corresponds to the –OH stretching vibration, was drastically decreased. The vibration at 1580.1 cm−1 could be a consequence of the C = C stretching of benzene rings in MG [37]. The peak at 1355.2 cm−1 for C–N stretching vibrations corresponds to the aromatic tertiary amine group [38], while the peak at 1162.6 cm−1 is assigned to the C–O-stretching vibrations from –COOH groups [39]. The FT-IR spectra of Fe3O4/RWB indicate that the functional groups on the biosorbent surface, such as –COOH, –OH, and –NH2, could be potential binding sites for MG biosorption.
In order to obtain morphological information, the Fe3O4/RWB surface was observed using an SEM under 100, 10, and 5 µm magnification; the results are displayed in Figure 3. The SEM analysis indicated that the Fe3O4/RWB surface appears rough, which is typical for plant biosorbents [40]. The observed porous structure of Fe3O4/RWB might be able to provide additional active sites for MG biosorption.
EDS analysis revealed carbon (46.6%) and oxygen (31.2%) to be the main elements of Fe3O4/RWB, which is typical for plant materials [41]. The presence of nitrogen, sulfur, and phosphorus was negligible, while the presence of iron (22.1%) suggests Fe3O4/RWB magnetization (Figure 4).
The XRD pattern of Fe3O4/RWB, presented in Figure 5, contains reflections at ~30°, ~36°, ~43°, ~54°, ~57°, and ~63° 2θ, corresponding to the (220), (311), (400), (422), (511), and (440) planes of magnetite [42]. This confirms the presence of magnetite in the biosorbent structure. Also, according to the XRD analysis, Fe3O4/RWB is composed of cellulose. The presence of distinct reflections at ~17°, ~23°, and ~35° 2θ corresponds to the (110), (200), and (004) planes of cellulose I [43].
The point of zero charge (pHpzc) of Fe3O4/RWB was determined by adapting a pH drift method previously reported by Lemos et al. [44]. The pHpzc of the biosorbent allows for the optimal pH conditions under which the best biosorption performances can be obtained to be predicted [45]. The Fe3O4/RWB (0.05 g) was added in glass vials with 20 mL of 0.1 mol/L NaCl, in a pH range from 2 to 10. The pH of the suspensions was measured after 24 h, and the pHpzc determined for Fe3O4/RWB was 4.8 (Figure 6). Based on this result, it can be concluded that, at pH values < 4.8, a positive charge will be dominant, while a negative charge will be dominant at pH values > 4.8. Since MG is a cationic dye, it is expected that more dye molecules will bind to the Fe3O4/RWB when the dye molecules are positively charged and the biosorbent surface is negatively charged.

3.2. Batch Biosorption Study

3.2.1. Effect of pH

The removal of MG from aqueous solutions using the Fe3O4/RWB biosorbent was assessed at different pH values (2, 4, 6, 8, and 10), keeping the other parameters constant (C0 = 50 mg/L; biosorbent dose = 5 g/L; T = 298 K; t = 60 min; shaking speed = 400 rpm). As shown in Figure 7a, the MG removal percentage increased from 63.8% (pH 2) to 86.6% (pH 8). When initial pH values were lower than the pHpzc (4.8), electrostatic repulsions between the positively charged surface of the Fe3O4/RWB biosorbent and the positively charged functional groups of the MG dye hindered biosorption. The efficiency of the removal of MG by the Fe3O4/RWB is attributed to other interactions, such as hydrogen bonding and π-π interactions (Scheme 1). On the other hand, when the initial pH is higher than 4.8, electrostatic attractions between the negatively charged functional groups (–O, –COO) on the Fe3O4/RWB surface and positively charged MG become dominant, promoting MG’s removal from the solution. At pH > 8.0, the decolorization of MG is intensive due to a chemical reaction between the excess of OH ions in the solution and the MG molecules. The OH ions attack the central carbon atom of the MG molecules, quenching the resonance and forming a carbinol base, which is colorless. This could provide an explanation of the seeming increase in removal efficiency at pH > 8, while the colorless dye remains in the solution [46]. Therefore, pH 8 was selected as the optimal pH value for the removal of MG using Fe3O4/RWB.

3.2.2. Effect of Sorbent Dose

The influence of different Fe3O4/RWB doses (1–5 g/L) on the removal efficiency of MG, with the other experimental variables being kept constant (C0 = 50 mg/L; pH 8; T = 298 K; t = 60 min; shaking speed = 400 rpm), is represented in Figure 7b. The results show an increase in the MG removal efficiency from 24.6% to 62.2% in the range of 1–5 g/L of Fe3O4/RWB. Based on the obtained results, the biosorption efficiency of MG increases with an increase in the sorbent dose. Therefore, the optimal dose of biosorbent was 5 g/L.

3.2.3. Effect of Ionic Strength

The influence of ionic strength (Figure 7c) on MG biosorption was examined using different concentrations of NaCl (0.1–1 mol/L), while the other experimental variables were kept constant (C0 = 50 mg/L; biosorbent dose = 5 g/L, pH 8; T = 298 K; t = 60 min; shaking speed = 400 rpm). With the increase in NaCl concentration from 0 to 1 mol/L, the removal efficiency decreased from 62.2% to 4%. With an increase in NaCl concentration, the adverse effect of ionic strength on the MG removal efficiency was higher. A possible explanation for this phenomenon lies in the competition between Na+ ions and MG molecules for the binding sites on the Fe3O4/RWB. The Na+ ions had an advantage compared to the MG molecules, probably due to their increased concentration and smaller size. This influence was most noticeable with the increase in the concentration of NaCl from 0.8 to 1 mol/L, where the efficiency of the MG removal was drastically decreased from 44.8% to 4%. At the same time, Cl ions reduce the electropositivity of the –N(CH3)2+ group in MG, negatively affecting the electrostatic attraction between negatively charged active sites on the biosorbent and positively charged MG [47].

3.2.4. Biosorption Thermodynamics

The thermodynamic parameters, Gibbs free energy (Δ), enthalpy (Δ), and entropy (Δ) for MG biosorption on Fe3O4/RWB were analyzed at three temperatures (298, 308, and 318 K), while the other experimental variables were kept constant (C0 = 50 mg/L; biosorbent dose = 5 g/L, pH 8; t = 60 min; shaking speed = 400 rpm). To evaluate the spontaneity of the biosorption, the Gibbs free energy was determined (Table 1). The decrease in Δ and removal efficiency (Figure 7d) with increasing temperature indicates that the biosorption of MG on Fe3O4/RWB is a spontaneous and thermodynamically stable process. The negative values for Δ and Δ indicate that biosorption is an exothermic process with a reduction in randomness at the interface between the solution phase and the biosorbent. Thermodynamic parameters were determined using the following Van’t Hoff Equations (4)–(7) [48,49]:
G ° = R T   l n   K
K c = C a C e
l n   K = S ° R H ° R T  
G ° = H ° T S °  
where R is the ideal gas constant (8.314 J/mol K), T (K) is the absolute temperature, K is the thermodynamic equilibrium constant, Ca (mg/L) is the amount of dye adsorbed at equilibrium, and Ce (mg/L) is the dye concentration in the solution at equilibrium.

3.2.5. Biosorption Kinetic

The effect of contact time on the removal of MG is shown in Figure 8. As can be seen, the uptake of MG by Fe3O4/RWB reached 91.8% after 50 min of biosorption, indicating relatively high affinity. The obtained results clearly show that biosorption was faster at the beginning, and then the removal efficiency slowly increased until the biosorption reached equilibrium. The fast elimination of MG from the solution reveals that the examined biosorbent could be a good candidate for wastewater treatment. The biosorption mechanism of MG on Fe3O4/RWB was evaluated using the linearized form of pseudo-first-order (PFO), pseudo-second-order (PSO), and Elovich kinetic models (Table S1). The higher R2 value of PSO (0.999) indicates that this model is suitable for describing the biosorption kinetics of MG (Table 2; Figure 9), suggesting that chemisorption limits the rate of the biosorption process. Also, the values of the equilibrium biosorption capacity of the PSO (9.59 mg/g) and the experimentally obtained biosorption capacity (9.3 mg/g) were matched.

3.2.6. Biosorption Isotherms

Figure 10 represents the biosorption of MG at different initial concentrations of the dye solution (10–300 mg/L) with a constant biosorbent dose of 5 g/L as a function of contact time (0–120 min) for a stirring speed of 400 rpm, at 298 K and pH = 8. The obtained results indicate that an increase in initial dye concentration leads to an increase in the biosorption capacity of Fe3O4/RWB. It also can be concluded that the biosorption capacity remains constant at MG concentrations higher than 150 mg/L due to the fact that the biosorbent reaches its maximum capacity at this concentration. The Langmuir, Freundlich, and Temkin isotherm models (Table S2) were used to evaluate the biosorption on Fe3O4/RWB and to predict the maximum efficiency (Table 3; Figure 11). The Langmuir isotherm model assumes a monolayer and homogeneous biosorption, in which each molecule possesses constant enthalpies and a constant biosorption activation energy. A Freundlich model can be applied to multilayer biosorption, with a heterogeneous distribution over the biosorbent surface [50]. The Temkin model assumes that the heat of biosorption would linearly decrease with the increase in the coverage of the biosorbent, and this has been applied to describe biosorption on heterogeneous surfaces [51]. The coefficient of determination value, R2 = 0.9736, indicates that biosorption probably follows the Langmuir isotherm model. Another confirmation of this fact is that the theoretical maximum sorption capacity, calculated from the Langmuir adsorption isotherm model, was 34.1 mg/g, which is very close to the experimentally determined sorption capacity of 31.5 mg/g. The maximum sorption capacity of MG on the pristine biosorbent (RWB) was 33.4 mg/g, and a minimal difference was observed compared with the Fe3O4/RWB.

3.2.7. Comparation Studies

A comparison of the maximum biosorption capacity of MG with other biosorbents reported in the literature is shown in Table 4. The calculated capacities indicate that Fe3O4/RWB has significant potential for dye removal from aqueous solutions and can be considered suitable for the treatment of polluted waters. Compared to other biosorbents used for wastewater treatments, Fe3O4/RWB showed relatively good efficiency, with a biosorption capacity of 34.1 mg/g. However, it is important to note that the studies mentioned in Table 4 were performed under different experimental conditions, and for a more precise evaluation it will be necessary to perform a set of studies under uniform conditions.

3.3. Reusability of Fe3O4/RWB

From the environmental and economic perspectives, it is important that biosorbents have the possibility of regeneration and reuse. A regeneration/reusability analysis of Fe3O4/RWB was carried out with three different desorption agents (0.1 mol/L HCl, 0.1 mol/L NaOH, and absolute ethanol), and the obtained results are presented in Figure 12. Based on the presented results, the best desorbing agent after the first adsorption/desorption cycle was absolute ethanol; as such, this was used for the rest of the regeneration/reusability analyses. As shown in Figure 13, after the first adsorption/desorption cycle, the efficiency of MG removal decreased significantly, from 79% to 49%. This can be explained by the fact that all binding sites of the Fe3O4/RWB were free during the first adsorption/desorption cycle, while during the second cycle, despite the desorption that occurred after the first cycle, a certain portion of the binding sites on the biosorbent remained occupied by MG molecules. After the second cycle, the trend regarding reuse of the Fe3O4/RWB did not change significantly, suggesting that it has significant potential for future commercial use in the treatment of wastewater, and that it is probably usable over a larger number of cycles.

4. Conclusions

This investigation revealed the MG biosorption potential of Fe3O4/RWB (from Ambrosia artemisiifolia biomass) in a batch mode. The results showed that the PSO model was appropriate for modeling MG biosorption using Fe3O4/RWB, indicating that chemisorption limits the biosorption process. An analysis of the adsorption isotherms shows that the Langmuir model matched the experimental data well, which assumes that a monolayer is present, as well as the homogeneous biosorption of MG. Under optimal experimental conditions (C0 = 50 mg/L; Fe3O4/RWB dose = 5 g/L; T = 298 K; t = 60 min; pH = 8) the maximum adsorption capacity of MG was 34.1 mg/g. Based on the obtained results, it can be concluded that Fe3O4/RWB has the potential to eliminate MG from wastewater. This approach offers a cheap, eco-friendly, and effective technological alternative to conventional methods for the removal of MG. Further research directions will focus on the applicability of Fe3O4/RWB in the biosorption of wide range of different contaminants (other synthetic dyes, pesticides, drugs, etc.). Transference of the lab-scale results to industrial-scale applications is also one of the future research perspectives regarding this approach. That will involve determining the technological feasibility, economy, and ecological performance of the application of biosorption at a large scale.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/separations11110310/s1, Table S1: Kinetic models used to analyze MG biosorption; Table S2. Isotherm models used to analyze MG biosorption; Reference [59] is cited in the Supplementary Materials.

Author Contributions

Investigation, and writing—original draft, N.N.; resources, and methodology, T.T.; data curation, and validation, B.M.; visualization, and funding acquisition, A.N.; formal analysis, and project administration, A.P.; conceptualization, supervision, and writing—review and editing, S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been financially supported by the Ministry of Science, Technological Development and Innovation of Republic of Serbia (Contract No: 451-03-66/2024-03/200026 and 451-03-66/2024-03/200168).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. El Hadj Ali, Y.A.; Hejji, L.; Seddik, N.B.; Azzouz, A.; Pérez-Villarejo, L.; Stitou, M.; Sonne, C. Remediation of malachite-green dye from textile wastewater using biosorbent almond shell-based cellulose. J. Mol. Liq. 2024, 399, 124435. [Google Scholar] [CrossRef]
  2. Yunusa, U.; Bishir, U.; Ibrahim, M.B. Kinetic and thermodynamic studies of malachite green adsorption using activated carbon prepared from desert date seed shell. Alger. J. Eng. Technol. 2020, 2, 37–45. [Google Scholar]
  3. Li, S.; Tao, M. Removal of cationic dyes from aqueous solutions by a low-cost biosorbent: Longan shell. Desal. Water Treat. 2016, 57, 5193–5199. [Google Scholar] [CrossRef]
  4. Gao, Y.; Chang, H.; Liang, D.; Yang, X.Y.; Chen, Z.Q.; Liu, X. Harnessing Prunella vulgaris L. residues for enhanced biosorption of dyes in wastewater. Chem. Eng. Res. Des. 2023, 200, 469–480. [Google Scholar] [CrossRef]
  5. Litefti, K.; Freire, M.S.; Stitou, M.; González-Álvarez, J. Biosorption of methylene blue and malachite green from single and binary solutions by Pinus pinaster bark. Maderas-Cienc. Tecnol. 2024, 26, 1–12. [Google Scholar] [CrossRef]
  6. Pathy, A.; Krishnamoorthy, N.; Chang, S.X.; Paramasivan, B. Malachite green removal using algal biochar and its composites with kombucha SCOBY: An integrated biosorption and phytoremediation approach. Surf. Interfaces 2022, 30, 101880. [Google Scholar] [CrossRef]
  7. Qi, Y.; Zhu, L.; Shen, X.; Sotto, A.; Gao, C.; Shen, J. Polyethyleneimine-modified original positive charged nanofiltration membrane: Removal of heavy metal ions and dyes. Sep. Purif. Technol. 2019, 222, 117–124. [Google Scholar] [CrossRef]
  8. Miyar, H.K.; Pai, A.; Goveas, L.C. Adsorption of Malachite Green by extracellular polymeric substance of Lysinibacillus sp. SS1: Kinetics and isotherms. Heliyon 2021, 7, e07169. [Google Scholar] [CrossRef]
  9. El Ouadrhiri, F.; Althomali, R.H.; Adachi, A.; Saleh, E.A.M.; Husain, K.; Lhassani, A.; Lahkimi, A. Nitrogen and phosphorus co-doped carbocatalyst for efficient organic pollutant removal through persulfate-based advanced oxidation processes. J. Saudi Chem. Soc. 2023, 27, 101648. [Google Scholar] [CrossRef]
  10. Perendija, J.; Ljubić, V.; Popović, M.; Milošević, D.; Arsenijević, Z.; Đuriš, M.; Cvetković, S. Assessment of waste hop (Humulus Lupulus) stems as a biosorbent for the removal of malachite green, methylene blue, and crystal violet from aqueous solution in batch and fixed-bed column systems: Biosorption process and mechanism. J. Mol. Liq. 2024, 394, 123770. [Google Scholar] [CrossRef]
  11. Li, Z.; Han, Q.; Zong, Z.; Xu, Q.; Wang, K. Exploring on the optimal preparation conditions of activated carbon produced from solid waste produced from sugar industry and Chinese medicine factory. Waste Dispos. Sustain. Energy. 2020, 2, 65–77. [Google Scholar] [CrossRef]
  12. Nath, J.; Ray, L. Biosorption of Malachite green from aqueous solution by dry cells of Bacillus cereus M116 (MTCC 5521). J. Environ. Chem. Eng. 2015, 3, 386–394. [Google Scholar] [CrossRef]
  13. Deniz, F.; Kepekci, R.A. Bioremoval of Malachite green from water sample by forestry waste mixture as potential biosorbent. Microchem. J. 2017, 132, 172–178. [Google Scholar] [CrossRef]
  14. Chen, L.; Mi, B.; He, J.; Li, Y.; Zhou, Z.; Wu, F. Functionalized biochars with highly-efficient malachite green adsorption property produced from banana peels via microwave-assisted pyrolysis. Bioresour. Technol. 2023, 376, 128840. [Google Scholar] [CrossRef]
  15. Knolmajer, B.; Jócsák, I.; Taller, J.; Keszthelyi, S.; Kazinczi, G. Common Ragweed-Ambrosia artemisiifolia L.: A Review with Special Regards to the Latest Results in Biology and Ecology. Agronomy 2024, 14, 497. [Google Scholar] [CrossRef]
  16. Battlay, P.; Wilson, J.; Bieker, V.C.; Lee, C.; Prapas, D.; Petersen, B.; Hodgins, K.A. Large haploblocks underlie rapid adaptation in the invasive weed Ambrosia artemisiifolia. Nat. Commun. 2023, 14, 1717. [Google Scholar] [CrossRef]
  17. Janaćković, P.; Rajčević, N.; Gavrilović, M.; Novaković, J.; Radulović, M.; Miletić, M.; Marin, P.D. Essential oil composition of Ambrosia artemisiifolia and its antibacterial activity against phytopathogens. Biol. Life Sci. Forum. 2022, 15, 22. [Google Scholar]
  18. The Ministry of Environmental Protection of the Republic of Serbia; Environmental Protection Agency. Air Quality in the Republic of Serbia in 2023; The Ministry of Environmental Protection of the Republic of Serbia, Environmental Protection Agency: Belgrade, Serbia, 2024; p. 98. [Google Scholar]
  19. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U., Jr. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent–A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef]
  20. Feng, Q.; Wang, B.; Chen, M.; Wu, P.; Lee, X.; Xing, Y. Invasive plants as potential sustainable feedstocks for biochar production and multiple applications: A review. Resour. Conserv. Recycl. 2021, 164, 105204. [Google Scholar] [CrossRef]
  21. Nguyen, D.T.C.; Tran, T.V.; Kumar, P.S.; Din, A.T.M.; Jalil, A.A.; Vo, D.V.N. Invasive plants as biosorbents for environmental remediation: A review. Environ. Chem. Lett. 2022, 20, 1421–1451. [Google Scholar] [CrossRef]
  22. Osman, A.I.; Ayati, A.; Farghali, M.; Krivoshapkin, P.; Tanhaei, B.; Karimi-Maleh, H.; Sillanpaä, M. Advanced adsorbents for ibuprofen removal from aquatic environments: A review. Environ. Chem. Lett. 2024, 22, 373–418. [Google Scholar] [CrossRef]
  23. Medina-Zazueta, L.; Miranda-Castro, F.C.; Romo-Garcia, F.; Martínez-Gil, M.; Esparza-Ponce, H.E.; Encinas-Basurto, D.; Ibarra, J. Development of sustainable magnetic biosorbent using aqueous leaf extract of Vallesia glabra for methylene blue removal from wastewater. Sustainability 2023, 15, 4586. [Google Scholar] [CrossRef]
  24. Sharma, R.; Sarswat, A.; Pittman, C.U.; Mohan, D. Cadmium and lead remediation using magnetic and non-magnetic sustainable biosorbents derived from Bauhinia purpurea pods. RSC Adv. 2017, 7, 8606–8624. [Google Scholar] [CrossRef]
  25. Hamri, N.; Imessaoudene, A.; Hadadi, A.; Cheikh, S.; Boukerroui, A.; Bollinger, J.C.; Mouni, L. Enhanced adsorption capacity of methylene blue dye onto kaolin through acid treatment: Batch adsorption and machine learning studies. Water 2024, 16, 243. [Google Scholar] [CrossRef]
  26. Tadić, T.; Vuković, Z.; Pavlović, V.; Nastasović, A.; Onjia, A.; Marković, B. Preparation of magnetic surface molecularly imprinting polymers based on glycidyl methacrylate as selective sorbents for aniline removal from aqueous medium. Sci. Sinter. 2024, 00, 8. [Google Scholar] [CrossRef]
  27. Mallakpour, S.; Sirous, F.; Dinari, M. Bio-sorbent alginate/citric acid-sawdust/Fe3O4 nanocomposite beads for highly efficient removal of malachite green from water. Int. J. Biol. Macromol. 2022, 222, 2683–2696. [Google Scholar] [CrossRef]
  28. Sadiq, A.C.; Rahim, N.Y.; Suah, F.B.M. Adsorption and desorption of malachite green by using chitosan-deep eutectic solvents beads. Int. J. Biol. Macromol. 2020, 164, 3965–3973. [Google Scholar] [CrossRef]
  29. Jasińska, A.; Bernat, P.; Paraszkiewicz, K. Malachite green removal from aqueous solution using the system rapeseed press cake and fungus Myrothecium roridum. Desalination Water Treat. 2013, 51, 7663–7671. [Google Scholar] [CrossRef]
  30. El Messaoudi, N.; El Khomri, M.; El Mouden, A.; Bouich, A.; Jada, A.; Lacherai, A.; Américo-Pinheiro, J.H.P. Regeneration and reusability of non-conventional low-cost adsorbents to remove dyes from wastewaters in multiple consecutive adsorption–desorption cycles: A review. Biomass Convers. Biorefin. 2024, 14, 11739–11756. [Google Scholar] [CrossRef]
  31. Farias, K.C.; Guimarães, R.C.; Oliveira, K.R.; Nazário, C.E.; Ferencz, J.A.; Wender, H. Banana peel powder biosorbent for removal of hazardous organic pollutants from wastewater. Toxics 2023, 11, 664. [Google Scholar] [CrossRef]
  32. Sayğılı, H.; Güzel, F. Usability of activated carbon with well-developed mesoporous structure for the decontamination of malachite green from aquatic environments: Kinetic, equilibrium and regeneration studies. J. Porous Mater. 2018, 25, 477–488. [Google Scholar] [CrossRef]
  33. Melhi, S.; Alqadami, A.A.; Alosaimi, E.H.; Ibrahim, G.M.; El-Gammal, B.; Bedair, M.A.; Elnaggar, E.M. Effective Removal of Malachite Green Dye from Water Using Low-Cost Porous Organic Polymers: Adsorption Kinetics, Isotherms, and Reusability Studies. Water 2024, 16, 1869. [Google Scholar] [CrossRef]
  34. Chen, Z.; Deng, H.; Chen, C.; Yang, Y.; Xu, H. Biosorption of malachite green from aqueous solutions by Pleurotus ostreatus using Taguchi method. J. Environ. Health Sci. Eng. 2014, 12, 63. [Google Scholar] [CrossRef] [PubMed]
  35. Muinde, V.; Onyari, J.M.; Wamalwa, B.M.; Wabomba, J. Adsorption of malachite green from aqueous solutions onto rice husks: Kinetic and equilibrium studies. J. Environ. Prot. 2017, 8, 215–230. [Google Scholar] [CrossRef]
  36. Smiri, M.; Guey, F.; Chemingui, H.; Dekhil, A.B.; Elarbaoui, S.; Hafiane, A. Remove of humic acid from water using magnetite nanoparticles. Eur. J. Adv. Chem. Res. 2020, 1. [Google Scholar] [CrossRef]
  37. Olukanni, O.D.; Adenopo, A.; Awotula, A.O.; Osuntoki, A.A. Biodegradation of malachite green by extracellular laccase producing Bacillus thuringiensis RUN1. J. Basic Appl. Sci. 2013, 9, 543–549. [Google Scholar] [CrossRef]
  38. Merrad, S.; Abbas, M.; Trari, M. Adsorption of malachite green onto walnut shells: Kinetics, thermodynamic, and regeneration of the adsorbent by chemical process. Fibers Polym. 2023, 24, 1067–1081. [Google Scholar] [CrossRef]
  39. Eltaweil, A.S.; Mohamed, H.A.; Abd El-Monaem, E.M.; El-Subruiti, G.M. Mesoporous magnetic biochar composite for enhanced adsorption of malachite green dye: Characterization, adsorption kinetics, thermodynamics and isotherms. Adv. Powder Technol. 2020, 31, 1253–1263. [Google Scholar] [CrossRef]
  40. Zhao, X.; Chen, J.; Chen, F.; Wang, X.; Zhu, Q.; Ao, Q. Surface characterization of corn stalk superfine powder studied by FTIR and XRD. Colloids Surf. B Biointerfaces 2013, 104, 207–212. [Google Scholar] [CrossRef]
  41. Shooto, N.D.; Thabede, P.M. Binary adsorption of chromium and cadmium metal ions by hemp (Cannabis sativa) based adsorbents. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100683. [Google Scholar] [CrossRef]
  42. Ouyang, Z.W.; Chen, E.C.; Wu, T.M. Thermal stability and magnetic properties of polyvinylidene fluoride/magnetite nanocomposites. Materials 2015, 8, 4553–4564. [Google Scholar] [CrossRef] [PubMed]
  43. Das, D.; Hussain, S.; Ghosh, A.K.; Pal, A.K. Studies on cellulose nanocrystals extracted from Musa sapientum: Structural and bonding aspects. Cellul. Chem. Technol. 2018, 52, 729–739. [Google Scholar]
  44. Lemos, E.S.; Fiorentini, E.F.; Bonilla-Petriciolet, A.; Escudero, L.B. Malachite Green removal by grape stalks biosorption from natural waters and effluents. Adsorpt. Sci. Technol. 2023, 2023, 6695937. [Google Scholar] [CrossRef]
  45. Carvalho, J.T.T.; Milani, P.A.; Consonni, J.L.; Labuto, G.; Carrilho, E.N.V.M. Nanomodified sugarcane bagasse biosorbent: Synthesis, characterization, and application for Cu (II) removal from aqueous medium. Environ. Sci. Pollut. Res. 2021, 28, 24744–24755. [Google Scholar] [CrossRef] [PubMed]
  46. De Marco, C.; Mauler, R.S.; Daitx, T.S.; Krindges, I.; Cemin, A.; Bonetto, L.R.; Giovanela, M. Removal of malachite green dye from aqueous solutions by a magnetic adsorbent. Sep. Sci. Technol. 2020, 55, 1089–1101. [Google Scholar] [CrossRef]
  47. Zhang, J.; Zhao, Y.; Wu, S.; Jia, G.; Cui, X.; Zhao, P.; Li, Y. Enhanced adsorption of malachite green on hydroxyl functionalized coal: Behaviors and mechanisms. Process Saf. Environ. Prot. 2022, 163, 48–57. [Google Scholar] [CrossRef]
  48. Nastasović, A.; Marković, B.; Suručić, L.; Onjia, A. Methacrylate-based polymeric sorbents for recovery of metals from aqueous solutions. Metals 2022, 12, 814. [Google Scholar] [CrossRef]
  49. Sorour, F.H.; Aboeleneen, N.M.; Abd El-Monem, N.M.; Ammar, Y.A.; Mansour, R.A. Removal of malachite green from wastewater using date seeds as natural adsorbent; isotherms, kinetics, Thermodynamic, and batch adsorption process design. Int. J. Phytoremediation 2024, 26, 1321–1335. [Google Scholar] [CrossRef]
  50. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  51. Ebrahimian, A.; Saberikhah, E.; Emami, M.S.; Sotudeh, M. Study of biosorption parameters: Isotherm, kinetics and thermodynamics of basic blue 9 biosorption onto foumanat tea waste. Cellulose Chem. Technol. 2014, 48, 735–743. [Google Scholar]
  52. Parlayıcı, Ş.; Pehlivan, E. Biosorption of methylene blue and malachite green on biodegradable magnetic Cortaderia selloana flower spikes: Modeling and equilibrium study. Int. J. Phytoremediation 2021, 23, 26–40. [Google Scholar] [CrossRef] [PubMed]
  53. Abate, G.Y.; Alene, A.N.; Habte, A.T.; Getahun, D.M. Adsorptive removal of malachite green dye from aqueous solution onto activated carbon of Catha edulis stem as a low cost bio-adsorbent. Environ. Syst. Res. 2020, 9, 1–13. [Google Scholar] [CrossRef]
  54. Youssef, E.E.; Beshay, B.Y.; Tonbol, K.; Makled, S.O. Biological activities and biosorption potential of red algae (Corallina officinalis) to remove toxic malachite green dye. Sci. Rep. 2023, 13, 13836. [Google Scholar] [CrossRef] [PubMed]
  55. Bayram, O.; Köksal, E.; Göde, F.; Pehlivan, E. Decolorization of water through removal of methylene blue and malachite green on biodegradable magnetic Bauhinia variagata fruits. Int. J. Phytoremediation 2022, 24, 311–323. [Google Scholar] [CrossRef] [PubMed]
  56. Săcară, A.M.; Indolean, C.; Cristea, V.M.; Mureşan, L.M. Application of adaptive neuro-fuzzy interference system on biosorption of malachite green using fir (Abies nordmanniana) cones biomass. Chem. Eng. Commun. 2019, 206, 1249–1263. [Google Scholar] [CrossRef]
  57. Djelloul, C.; Ghodbane, H. Removal of malachite green using ultrasound-assisted and conventional batch adsorption based on Paracentrotus lividus shells as biosorbent. Desalination Water Treat. 2022, 267, 201–214. [Google Scholar] [CrossRef]
  58. Deniz, F.; Ersanli, E.T. Purification of malachite green as a model biocidal agent from aqueous system by using a natural widespread coastal biowaste (Zostera marina). Int. J. Phytoremediation 2021, 23, 772–779. [Google Scholar] [CrossRef]
  59. Ouettar, L.; Guechi, E.K.; Hamdaoui, O.; Fertikh, N.; Saoudi, F.; Alghyamah, A. Biosorption of triphenyl methane dyes (malachite green and crystal violet) from aqueous media by alfa (Stipa tenacissima L.) leaf powder. Molecules 2023, 28, 3313. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of all stages of this research.
Figure 1. Schematic representation of all stages of this research.
Separations 11 00310 g001
Figure 2. FTIR spectra of Fe3O4/RWB and Fe3O4/RWB–MG.
Figure 2. FTIR spectra of Fe3O4/RWB and Fe3O4/RWB–MG.
Separations 11 00310 g002
Figure 3. The SEM images of Fe3O4/RWB at (a) 100 μm, (b) 10 μm, and (c) 5 μm magnification.
Figure 3. The SEM images of Fe3O4/RWB at (a) 100 μm, (b) 10 μm, and (c) 5 μm magnification.
Separations 11 00310 g003
Figure 4. The EDS analysis of Fe3O4/RWB.
Figure 4. The EDS analysis of Fe3O4/RWB.
Separations 11 00310 g004
Figure 5. XRD pattern of Fe3O4/RWB.
Figure 5. XRD pattern of Fe3O4/RWB.
Separations 11 00310 g005
Figure 6. pHpzc determination using the pH drift method.
Figure 6. pHpzc determination using the pH drift method.
Separations 11 00310 g006
Figure 7. The effects of the experimental parameters (a) initial pH (2–10), (b) biosorbent dose (1–5 g/L), (c) ionic strength (0–1 mol/L), and (d) temperature (298–318 K) on MG biosorption on Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Figure 7. The effects of the experimental parameters (a) initial pH (2–10), (b) biosorbent dose (1–5 g/L), (c) ionic strength (0–1 mol/L), and (d) temperature (298–318 K) on MG biosorption on Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Separations 11 00310 g007
Scheme 1. Possible biosorption mechanism of MG on Fe3O4/RWB.
Scheme 1. Possible biosorption mechanism of MG on Fe3O4/RWB.
Separations 11 00310 sch001
Figure 8. Effect of contact time (0–120 min) on the removal of MG (C0 = 50 mg/L; MB dose = 5 g/L; T = 298 K; pH = 8; shaking speed = 400 rpm). Vertical error bars represent 95% confidence intervals.
Figure 8. Effect of contact time (0–120 min) on the removal of MG (C0 = 50 mg/L; MB dose = 5 g/L; T = 298 K; pH = 8; shaking speed = 400 rpm). Vertical error bars represent 95% confidence intervals.
Separations 11 00310 g008
Figure 9. Linear plots of kinetics models—(a) PFO, (b) PSO, and (c) Elovich—for the biosorption of MG on Fe3O4/RWB (Fitting line = red line; experimental data = blue dots).
Figure 9. Linear plots of kinetics models—(a) PFO, (b) PSO, and (c) Elovich—for the biosorption of MG on Fe3O4/RWB (Fitting line = red line; experimental data = blue dots).
Separations 11 00310 g009
Figure 10. Influence of the initial concentration of MG and contact time on the biosorption capacity. Vertical error bars represent 95% confidence intervals.
Figure 10. Influence of the initial concentration of MG and contact time on the biosorption capacity. Vertical error bars represent 95% confidence intervals.
Separations 11 00310 g010
Figure 11. Linear plots of isotherm models—(a) Langmuir, (b) Freundlich, and (c) Temkin—for the biosorption of MG on Fe3O4/RWB (Fitting line = red line; experimental data = blue dots).
Figure 11. Linear plots of isotherm models—(a) Langmuir, (b) Freundlich, and (c) Temkin—for the biosorption of MG on Fe3O4/RWB (Fitting line = red line; experimental data = blue dots).
Separations 11 00310 g011
Figure 12. Influence of examined desorption agent on the regeneration/reusability of Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Figure 12. Influence of examined desorption agent on the regeneration/reusability of Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Separations 11 00310 g012
Figure 13. Influence of number of desorption cycles on the reusability of Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Figure 13. Influence of number of desorption cycles on the reusability of Fe3O4/RWB. Vertical error bars represent 95% confidence intervals.
Separations 11 00310 g013
Table 1. Thermodynamic parameters for the biosorption of MG on Fe3O4/RWB.
Table 1. Thermodynamic parameters for the biosorption of MG on Fe3O4/RWB.
Temperature (K)ΔG° (J/mol)ΔH° (J/mol)ΔS° (J/mol·K)
298−2010−31,097−98
308−691
318−60
Table 2. Kinetics parameters for the biosorption of MG (C0 = 50 mg/L; MB dose = 5 g/L; T = 298 K; pH = 8; shaking speed = 400 rpm).
Table 2. Kinetics parameters for the biosorption of MG (C0 = 50 mg/L; MB dose = 5 g/L; T = 298 K; pH = 8; shaking speed = 400 rpm).
ParameterValues
Qeexp, mg/g9.3
(PFO)
Qecal, mg/g4.24
k1, min−10.054
R20.960
(PSO)
Qecal, mg/g9.59
k2, g/mg min0.034
R20.999
Elovich model
αE, mg/g72.02
βE, g/mg0.81
R20.960
Table 3. Parameters of the adsorption isotherms of MG biosorption on Fe3O4/RWB.
Table 3. Parameters of the adsorption isotherms of MG biosorption on Fe3O4/RWB.
ParameterValue
Langmuir model
Qm, mg/g34.1
KL, L/mg0.085
R20.997
Freundlich model
nF2.25
KF, L/g4.54
R20.937
Temkin model
AT, L/mg10.31
R20.9187
Table 4. Comparison of MG removal using different biosorbents.
Table 4. Comparison of MG removal using different biosorbents.
BiosorbentConcetration (mg/L)pHBiosorbent Dosage (g/L)Time (min)Temperature (K)Qm (mg/g)Ref.
Fe3O4/RWB 508.05.0012029834.1This study
Mag. Cortaderia selloana flower spikes25–3506.04.004529856.5[52]
Catha edulis stem10–5010.010.0060/5.6[53]
Red algae
Corallina ofcinalis
206.01.00120300101.3[54]
Mag. Bauhinia variagata fruits1508.00.206029830.1[55]
Fir (Abies nordmanniana) cones1103.350.00876029411.0[56]
Paracentrotus lividus shells10initial 4.001802947.2[57]
Coastal biowaste (Zostera marina)1540.0136029897.6[58]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nedić, N.; Tadić, T.; Marković, B.; Nastasović, A.; Popović, A.; Bulatović, S. Eco-Friendly Green Approach to the Biosorption of Hazardous Dyes from Aqueous Solution on Ragweed (Ambrosia artemisiifolia) Biomass. Separations 2024, 11, 310. https://doi.org/10.3390/separations11110310

AMA Style

Nedić N, Tadić T, Marković B, Nastasović A, Popović A, Bulatović S. Eco-Friendly Green Approach to the Biosorption of Hazardous Dyes from Aqueous Solution on Ragweed (Ambrosia artemisiifolia) Biomass. Separations. 2024; 11(11):310. https://doi.org/10.3390/separations11110310

Chicago/Turabian Style

Nedić, Natalija, Tamara Tadić, Bojana Marković, Aleksandra Nastasović, Aleksandar Popović, and Sandra Bulatović. 2024. "Eco-Friendly Green Approach to the Biosorption of Hazardous Dyes from Aqueous Solution on Ragweed (Ambrosia artemisiifolia) Biomass" Separations 11, no. 11: 310. https://doi.org/10.3390/separations11110310

APA Style

Nedić, N., Tadić, T., Marković, B., Nastasović, A., Popović, A., & Bulatović, S. (2024). Eco-Friendly Green Approach to the Biosorption of Hazardous Dyes from Aqueous Solution on Ragweed (Ambrosia artemisiifolia) Biomass. Separations, 11(11), 310. https://doi.org/10.3390/separations11110310

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