Next Article in Journal
Heat Transfer Analysis of Nanofluid Flow in a Rotating System with Magnetic Field Using an Intelligent Strength Stochastic-Driven Approach
Next Article in Special Issue
Tetracycline Removal by Hercynite-Biochar from the Co-Pyrolysis of Red Mud-Steel Slag-Sludge
Previous Article in Journal
Special Issue: State-of-the-Art Nanophotonic and Optical Nanomaterials in China
Previous Article in Special Issue
Emerging Trends in the Remediation of Persistent Organic Pollutants Using Nanomaterials and Related Processes: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Nano Iron Oxide–Modified Biochar from Co-Hydrothermal Carbonization of Microalgae and Fe(II) Salt for Efficient Removal of Rhodamine B

Changjiang River Scientific Research Institute, Research Center of Water Engineering Safety and Disaster Prevention of Ministry of Water Resources, Wuhan 430010, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(13), 2271; https://doi.org/10.3390/nano12132271
Submission received: 10 June 2022 / Revised: 23 June 2022 / Accepted: 28 June 2022 / Published: 1 July 2022
(This article belongs to the Special Issue Nanomaterials for Detection and Removal of Organic Pollutants)

Abstract

:
Dye adsorption by magnetic modified biochar has now received growing interest due to its excellent adsorption performance and facile separation for recycling. In this study, nano iron oxide–modified biochar was fabricated via the successive hydrothermal-pyrolyzing method using Chlorella vulgaris (Cv) and FeSO4·7H2O as raw materials, and its adsorption on Rhodamine B (RhB) in aqueous solution was studied. Multiple techniques such as X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), scanning electron microscopy (SEM), transmission electron microscopy (TEM), Brunauer–Emmett–Teller (BET), vibrating sample magnetometry (VSM) and X-ray photoelectron spectroscopy (XPS) were employed to comprehensively characterize the structure, morphology and physicochemical properties of the adsorbent. The as-synthesized nano iron oxide–modified biochar (CBC-Fe(II)) exhibited a large surface area (527.6 m2/g) and high magnetic saturation value (13.7 emu/g) to facilitate magnetic separation. Compared with CBC and CBC-Fe(III), CBC-Fe(II) exhibited superior adsorption ability towards RhB in aqueous solution, with a maximum adsorption capacity of 286.4 mg/g. The adsorption process of RhB onto CBC-Fe(II) was well described by the pseudo-second-order kinetic model and Langmuir isotherm model, indicating monolayer chemisorption behaviors for the adsorption system. Facile preparation, great adsorption performance and magnetic recovery properties endow CBC-Fe(II) to be a promising adsorbent for dye removal.

1. Introduction

Recently, the content of organic pollutants in the environment has been increasing due to the rapid development of industry, leading to severe environmental issues [1,2,3,4]. Specifically, the disposed water from the textile, printing, cosmetics, medicine and other industries contains a significantly high amount of organic dye molecules which are potentially carcinogenic and teratogenic with high toxicity, posing a serious threat to various organisms and human health [5,6]. Therefore, developing an efficient technology to remove those dye contaminates from wastewater is indispensable to prevent environmental pollution.
A variety of treatment techniques have been utilized for the removal of dyes from polluted water, including coagulation [7], flocculation [8], oxidation [9], adsorption [10], electrochemical separation [11], microbiological decomposition [12] and the membrane process [13]. Among all, adsorption has become one of the most widely used methods due to its advantages of low cost, high efficiency, simple operation and environment-friendliness [14,15,16]. Some adsorbents developed for removing dyes from wastewater include bentonite [17], graphene [18], activated carbon [19], etc., and research is still going on to develop more economical, sustainable and efficient materials as new adsorbents for dye removal.
Biochar (BC), a carbon-based porous product obtained from the pyrolysis of biomass and possesses high potential for dye pollution remediation owing to its renewability, stability, its various sources, environmental sustainability and its beneficial physicochemical properties, including high surface area, developed pore structure, and others [20,21]. The renewable resources can be easily obtained from agricultural waste or naturally abundant biomass such as fruit peels, coconut shells, cassava slag, rice husk, sewage sludge, macroalgae, microalgae, etc. [22,23,24]. Wu et al. [25] prepared biochar from agricultural waste cassava slag by hydrothermal carbonization treatment, and the cassava slag biochar had excellent adsorption performance on RhB with a maximum adsorption capacity of 105.3 mg/g. Hou et al. [26] synthesized biochar with well-developed pores using agricultural waste from bamboo shoot shells as the raw materials. The biochar produced at the carbonization temperature of 800 °C could efficiently remove RhB from water. Vigneshwaran et al. [27] fabricated sulfur-doped biochar derived from tapioca peel waste, which had the adsorption capacity of 30.18 and 33.10 mg/g for malachite green and RhB dye, respectively. Nevertheless, the adsorption capacity of the above reported biochar is relatively limited. Moreover, solid–liquid separation often requires complementary steps of centrifugation and filtration, restraining its application potential.
In this case, biochar activated by magnetic modification exhibits controlled movement under an external magnetic field, making it easy to separate [28]. Magnetic biochars are fabricated by chemically combining them with other functional elements such as Fe, Fe3O4, γ-Fe2O3, CoFe2O4, etc., which leads to strong metal binding [29,30,31]. Though magnetic biochar has the advantage of easy separation, the surface pores are prone to be plugged with those functional particles, causing a decrease in its adsorption efficiency [32]. In addition, Fe-modified biochar is widely applied in the advanced oxidation process and the adsorption of heavy metal ions, but there are few studies involving adsorption of dyes by Fe-modified biochar [33]. Therefore, it is essential to explore a facile method to fabricate magnetic biochar adsorbents for dye removal and improve adsorption properties of biochar in the meantime.
Herein, nano iron oxide–modified biochar was successfully prepared via successive hydrothermal and pyrolyzing treatments. One of the easily cultivated microalgae (Cv) was selected as carbonaceous feedstock in this work owing to the wide abundance, high biomass productivity as well as the technical challenges associated with its treatment and disposal [34,35,36]. The obtained nano iron oxide–modified biochar was analyzed via XRD, FTIR, SEM, TEM, BET, VSM and XPS techniques. RhB, as a representative cationic organic dye compound, is considered to be more difficult to remove due to the high solubility in water [37,38]. Thus, RhB dye was chosen as the target contaminant to assess the adsorption efficiency of the nano iron oxide–modified biochar. The influence of adsorbent dosage, initial RhB concentration, adsorption time and temperature on the adsorption performance was also examined. The mechanism of RhB adsorption was further analyzed via kinetic and isothermal models. The results of this study offer a new approach to dye wastewater treatment using nano iron oxide–modified biochar and explore a new way for the resource utilization of Cv.

2. Materials and Methods

2.1. Materials

Cv powder was obtained from Xi’an Youshuo Biotechnology Co., Ltd. (Xi’an, China) Iron(II) sulfate heptahydrate (FeSO4·7H2O), iron(III) nitrate nonahydrate (Fe(NO3)3·9H2O), nitric acid (HNO3), ethanol (C2H5OH) and RhB were all supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) All chemical reagents were of analytical grade and were used without further purification. Ultrapure water (18.25 MΩ·cm) was used in all the experiments.

2.2. Preparation of Adsorbents

2.2.1. Preparation of Biochar

The Chlorella-derived biochar (CBC) was prepared through pyrolyzing the carbonaceous precursor (Cv). First, the Cv powder was dried at 60 °C overnight. Then the dried precursor was heated at 5 °C/min to 900 °C in a tubular furnace under nitrogen atmosphere and carbonized for 1 h. After cooling to room temperature, the obtained product was ground in a ball mill for 8 h. Finally, the powder was washed separately with HNO3, absolute ethyl alcohol and ultrapure water and then dried at 60 °C to obtain pure CBC.

2.2.2. Preparation of Nano Iron Oxide–Modified Biochar

Nano iron oxide–modified biochar was prepared via successive hydrothermal-pyrolyzing method. Two kinds of iron salts with different valences (FeSO4·7H2O and Fe(NO3)3·9H2O) were selected as iron sources. First, a certain amount of Cv powder was dispersed in a FeSO4·7H2O solution under continuous stirring in nitrogen atmosphere. The mass ratio of Fe element to Cv was fixed at 1:6. The mixture was then placed into a Teflon-lined autoclave and hydrothermally treated at 150 °C for 24 h. After the resulting product was dried at 60 °C, it was pyrolyzed at 900 °C using a similar procedure as preparing CBC. Then the obtained product was ball milled for 8 h. Additionally, to remove metallic and organic impurities, the powder was washed separately with HNO3, absolute ethyl alcohol and ultrapure water. At last, the sample was dried at 60 °C to yield nano iron oxide–modified biochar, which was denoted as CBC-Fe(II). For comparison, CBC-Fe(III) was also prepared with the similar procedure using the iron salt of Fe(NO3)3·9H2O instead of FeSO4·7H2O.

2.3. Characterizations

The obtained samples were characterized by XRD analysis on D8 advance X-ray diffractometer (Bruker AXS, Karlsruhe, Germany) with filtered Cu Kα radiation. The functional groups of the samples were analyzed by FTIR (Thermo Scientific, Nicolet 67000, Madison, WI, USA). The detailed microstructure and morphology of samples were observed using SEM (ZEISS Gemini SEM 300, Oberkochen, Germany) and TEM (FEI Tecnai F20, Hillsboro, OR, USA). Nitrogen adsorption–desorption isotherms were recorded by Micromeritics ASAP 2460 adsorption apparatus at 77 K (ASAP 2460, Micromeritics, Norcross, GA, USA). The magnetic property of the sample was measured using VSM (LakeShore, Westerville, OH, USA). XPS was conducted at room temperature on an ESCALAB Xi+ instrument (Thermo Fisher Scientific, Waltham, MA, USA) equipped with an Al Kα source.

2.4. Adsorption of RhB

The adsorption experiments were carried out in a thermostated mechanical shaker and the speed was set at 200 rpm. A certain amount of the prepared adsorbents was added to 25 mL of RhB dye solution in a 50mL centrifuge tube. The effects of adsorbent dosage (0–50 mg), adsorption time (0–180 min), temperature (15–35 °C) and initial concentration (100–1500 mg/L) on adsorption were examined. After the adsorption experiment, about 3 mL of the mixture was sampled and filtered through a 0.45 mm membrane filter. The concentration of RhB was determined using Lambda 365 UV/Vis spectrophotometer (PerkinElmer, Waltham, MA, USA) at a wavelength of 554 nm. The removal efficiency R (%), the quantity of RhB adsorbed on the adsorbent Qt (mg/g) and adsorption capacity Qe (mg/g) were calculated using Equations (1)–(3).
R % = C 0 C t C 0 × 100 %
Q t ( mg / g ) = C 0 C t m × V
Q e ( mg / g ) = C 0 C e m × V
where C0 (mg/L) is the initial concentration, Ct (mg/L) is residual concentration at certain time and Ce (mg/L) is the equilibrium concentration; V (L) is the volume of dye solution and m (g) is the mass of adsorbent. To guarantee the reliability of data, all of the above experiments were performed in triplicate.

3. Results and Discussion

3.1. Characterizations of the Synthesized Biochars

The XRD patterns of CBC, CBC-Fe(III) and CBC-Fe(II) are shown in Figure 1a. The broad diffraction peak at about 25° confirmed the presence of amorphous biochar in the three samples [39]. However, only the sample of CBC-Fe(II) showed the peaks related to iron oxide, which were identified as Fe3O4 and α-Fe2O3. Specifically, the diffraction peaks at 2θ = 35.1°, 41.4°, 50.4°, 62.9°, 67.2° and 74.1° corresponded to the (220), (311), (400), (422), (511) and (440) planes of Fe3O4, respectively. The diffraction peaks at 2θ = 38.7°, 41.4°, 47.8°, 58.2° and 63.7° corresponded to the (104), (110), (113), (024) and (116) planes of α-Fe2O3, respectively. The results implied that during the process of successive hydrothermal-pyrolyzing, the addition of Fe(II) salt promoted the formation of iron oxide particles, whereas the addition of Fe(III) salt had no positive effect on the formation of iron species in the biochar.
The functional groups of CBC, CBC-Fe(III) and CBC-Fe(II) were determined by FTIR spectra and are shown in Figure 1b. For the three samples, the peaks at 1592 cm−1 were attributed to the aromatic C=C groups stretching variations [27,40]. The peaks at 1230 cm−1 corresponded to the C–O stretching vibration [41,42]. Notably, the representative absorption bands for the stretching Fe–O band at around 660 cm−1 appeared for the sample of CBC-Fe(II) [43,44], further confirming that iron oxides had been successfully modified onto biochar.
Figure 2 shows the typical morphologies of CBC, CBC-Fe(III) and CBC-Fe(II). It can be seen that the surface morphologies of biochars were obviously different. Compared with CBC-Fe(III) and CBC-Fe(II), CBC exhibited larger carbon particles. It was due to the fact that the modification of Fe could prevent the aggregation of carbon particles during biomass carbonization process. It can be seen that some nanostructured particles were observed for the sample of CBC-Fe(II), which could be attributed to the formation of nan iron oxides particles. TEM images of CBC-Fe(II) (Figure 3a,b) revealed that Fe3O4 and α-Fe2O3 nanoparticles were presented in biochar. Those nano iron oxides could give more active sites to the prepared biochar, which was in favor of dye adsorption. Figure 3b shows the exhibited characteristic fringes with lattice spacing of 0.253 nm and 0.270 nm, which could be indexed as the (311) plane of Fe3O4 and (104) plane of α-Fe2O3, respectively. The result was consistent with the results of XRD analysis as mentioned above.
N2 adsorption-desorption isotherms of CBC, CBC-Fe(III) and CBC-Fe(II) were detected and are shown in Figure 4a, and calculated pore parameters are summarized in Table 1. According to IUPAC classification, N2 adsorption–desorption isotherms of the three samples could be classified as type IV, which indicated that the mesoporous structure existed in the biochars. BET surface areas calculated were 338.6, 350.2 and 527.6 m²/g for CBC, CBC-Fe(III) and CBC-Fe(II), respectively. Notably, compared with CBC and CBC-Fe(III), the BET surface area of CBC-Fe(II) significantly increased. It was due to the fact that the pore structure was increased by the iron oxides formed on CBC-Fe(II). The high specific surface area for CBC-Fe(II) suggested the abundant accessibility of active sites on the adsorbent surface, which could be accountable for an increase in the adsorption efficiency for dye removal. As seen in Table 1, the average pore sizes of CBC, CBC-Fe(III) and CBC-Fe(II) were 4.855, 5.702 and 3.614 nm, respectively, which implied that the prepared samples were mesoporous materials with a certain amount of microporous structure. The result was consistent with the pore size distributions in Figure 4b. It was also observed in Table 1 that CBC-Fe(II) possessed the highest total pore volume and micropore volume compared with the other two samples, further confirming the porous property of CBC-Fe(II).
The magnetic behavior of the CBC-Fe(II) adsorbent was studied. The magnetic hysteresis curves (Figure 5) show that the saturation magnetization (Ms) of CBC-Fe(II) was determined to be 13.7 emu/g. It was concluded that the magnetization of CBC-Fe(II) was high enough to be magnetically collected simply by a magnet.
The XPS spectra of CBC-Fe(II) adsorbent were obtained and are shown in Figure 6. All of the peaks of the scan survey spectra in Figure 6a were ascribed to C, O and Fe elements. The deconvolutions of the C1s spectra of CBC-Fe(II) are shown in Figure 6b. The strongest peak located at 284.8 eV could be assigned to C–C. The other peaks at 285.2 and 285.9 eV corresponded to C–O and C=O functional groups, respectively [45]. Figure 6c represents the XPS spectra of Fe 2p. The peaks at 710.9 and 724.6 eV were respectively ascribed to 2p3/2 and 2p1/2 of Fe(II), whereas the peaks at 712.6 and 726.3 eV corresponded to 2p3/2 and 2p1/2 of Fe(III) [46]. The peak at 719.5 eV was attributed to the binding energy of the shake-up satellite [47]. The above results verified the existence of Fe2O3 and Fe3O4 in the sample of CBC-Fe(II), which was consistent with the XRD results. Figure 6d shows that the O1s peak was asymmetric and can be fitted into three peaks. The lowest peak, located at 530.3 eV, was ascribed to the lattice oxygen (O2−) of Fe2O3 or Fe3O4 in the CBC-Fe(II) sample [48]. The peak centered at 531.8 eV was attributed to C=O in the composite, while the highest peak, at 533.0 eV, was associated with C–O in the composite [45].

3.2. Batch Adsorption Studies

3.2.1. Effect of Adsorbent Dosage on Adsorption of RhB

To find the optimal adsorbent dosage, the effect of adsorbent dosage on removal efficiency of RhB by CBC-Fe(II) was investigated using 100, 200 and 600 mg/L as the initial dye concentrations. From the results presented in Figure 7, it can be seen that for various initial dye concentrations, the removal efficiency increased rapidly with increasing amounts of CBC-Fe(II) from 10 mg to 30 mg. It was because more accessible reactive sites are available to facilitate the dye adsorption process. However, when further increasing the amount of adsorbent from 30 mg to 50 mg, the required adsorption sites were getting closer to saturation, thus the removal efficiency only increased slightly. Thus, considering both the efficient removal of dye molecules and economic benefits, the optimal amount of the CBC-Fe(II) dosage was 30 mg.

3.2.2. Effect of Contact Time on Adsorption of RhB

Figure 8 shows the influence of reaction time on the removal efficiency of RhB by CBC, CBC-Fe(III) and CBC-Fe(II) under the same reaction temperature. The results obtained revealed that the removal efficiency of CBC-Fe(II) to RhB was much higher than that of CBC and CBC-Fe(III), demonstrating that CBC-Fe(II) possessed a relatively strong driving force on RhB. It was evidently observed that the removal efficiency of the three samples to RhB increased quickly during the initial stage, which was attributed to the free adsorption sites available for a large number of dye molecules. However, with the increase in contact time, the number of adsorption positions gradually decreased, leading to a slower adsorption rate. Finally, the adsorption equilibrium was achieved after 120 min, when the removal efficiency became almost constant.

3.2.3. Effect of Temperature on Adsorption of RhB

Figure 9 exhibited the effect of temperature on removal efficiency of RhB using CBC-Fe(II). It could be facilely observed that the removal efficiency of CBC-Fe(II) towards RhB remained the same at different temperatures (15 °C, 25 °C and 35 °C). Therefore, for the sake of saving energy and costs, 25 °C was chosen as the adsorption temperature throughout the whole experiment.

3.3. Kinetic and Isotherm Adsorption

3.3.1. Adsorption Kinetic Models

The adsorption kinetics of RhB onto CBC, CBC-Fe(III) and CBC-Fe(II) were investigated using pseudo-first-order and pseudo-second-order kinetic models, and the results are shown in Figure 10a. The adsorption kinetic models can be expressed as Equations (4) and (5), respectively:
Q t =   Q e ( 1 exp ( K 1 t ) )
Q t = K 2 Q e 2 t 1 + K 2 Q e t
where Qe (mg/g) and Qt (mg/g) represent the quantity of RhB adsorbed on the adsorbent at the equilibrium and arbitrarily given time t (min), respectively. K1 (1/min) and K2 (g/(mg∙min)) are adsorption rate constants of the pseudo-first-order and pseudo-second-order, respectively.
The related kinetic parameters of the above two models were obtained and are listed in Table 2. It was observed that for the three samples, the coefficients of correlation (R2) were higher for the pseudo-second-order than that of the pseudo-first-order kinetic model, indicating that the pseudo-second-order kinetic model was more appropriate to describe the adsorption behavior towards RhB. The result implied that the dye adsorption process by the biochars was mainly dominated by chemical adsorption. Moreover, the theoretical adsorption capacity of CBC-Fe(II) was 284.9 mg/g, which was much higher than that of CBC (186.3 mg/g) and CBC-Fe(III) (184.1 mg/g), showing the best adsorption performance of CBC-Fe(II).
Additionally, the intra-particle diffusion model (Figure 10b) was used to further evaluate the different stages of adsorption rate, which can be expressed as Equation (6).
Q t = K id t 1 2 + c
where Qt (mg/g) represents the quantity of RhB adsorbed on the adsorbent at arbitrarily given time t (min). Kid (mg/(g∙min1/2)) is the diffusion rate constant. Constant c is the thickness of boundary layer.
From Figure 10b and Table 3, it is shown that for the three samples, all the lines did not pass through the origin, indicating that the intra-particle diffusion process was not the only rate-controlling mechanism for the adsorption system. The RhB adsorption process was comprised by three stages. In the first stage, the sharp rise was due to the liquid film diffusion process, while the second stage ascended more slowly, which represented the intra-particle diffusion process. The last equilibrium stage was attributed to the adsorption of dye molecules on the interior surface of adsorbents.

3.3.2. Adsorption Isotherm Models

In order to study the interactive behavior between adsorbates and adsorbents, the equilibrium adsorption isotherms of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II) were described by the Langmuir and Freundlich isotherm models, and the results are presented in Figure 11. The two isotherm models can be expressed as Equations (7) and (8).
Q e = Q m K l C e 1 + K L C e
Q e = K F C e 1 / n
where Ce (mg/L) is the equilibrium concentration of dye solution, Qe (mg/g) and Qm (mg/g) represent equilibrium adsorption capacity and maximum adsorption capacity of CBC-Fe(II), respectively. KL (L/mg) and KF [(mg/g)(mg/L)1/n] represent the constants of the Langmuir and Freundlich isotherms and n is the linearity index of adsorption intensity.
It was observed from Figure 11 that the quantity of RhB adsorbed by CBC, CBC-Fe(III) and CBC-Fe(II) increased rapidly at low concentrations, then gradually approached the saturated adsorption capacity. The experimental data were fitted by the Langmuir and the Freundlich models, and the calculated parameters are described in Table 4. Apparently, for the three samples, the significant correlation coefficient value of Langmuir isotherm model was higher than that of the Freundlich isotherm model. The result indicated that the Langmuir isotherm model was superior for revealing the adsorption process, suggesting the equilibrium adsorption mechanism of RhB by CBC, CBC-Fe(III) and CBC-Fe(II) was more inclined to monolayer adsorption. Furthermore, the maximum adsorption capacity (Qm) for CBC, CBC-Fe(III) and CBC-Fe(II) obtained by the Langmuir model was 179.7 mg/g, 185.1 mg/g and 289.6 mg/g, respectively, which were close to the experimental values of 167.4 mg/g, 179.7 mg/g and 286.4 mg/g, respectively. It is worth noting that the CBC-Fe(II) adsorbent prepared in this study was found to show obvious advantages compared with biochars for RhB removal systhesized in other studies (Table 5).

3.4. Regeneration and Reusability of CBC-Fe(II)

To investigate the regeneration and reusability of CBC-Fe(II), the adsorption experiment was conducted in five consecutive cycles. After each adsorption experiment, the adsorbent was recollected by a magnet and washed with absolute ethyl alcohol and ultrapure water, respectively. For the next adsorption cycle, 25 mL of RhB solution was added to CBC-Fe(II). The results of repetitive use of the adsorbent for RhB removal are shown in Figure 12. It can be seen that the regeneration efficiency of the CBC-Fe(II) decreased with each recycling process, which might be due to the increased RhB residual on the adsorbent surface. However, CBC-Fe(II) still possessed a certain adsorption capacity towards RhB after five adsorption cycles.

4. Conclusions

In summary, a magnetically separable nano iron oxide–modified biochar (CBC-Fe(II)) was fabricated via a facile hydrothermal-pyrolyzing method. The synthesized CBC-Fe(II) adsorbent with a high surface area exhibited excellent magnetic properties and could be simply collected by a magnet. Compared with CBC and CBC-Fe(III), CBC-Fe(II) showed powerful adsorption performance for removing RhB from water. The RhB was optimally adsorbed at a CBC-Fe(II) dosage of 30 mg, with a contact time of 120 min and temperature of 25ºC. The adsorption process of RhB onto CBC-Fe(II) was found to obey the pseudo-second-order model and the Langmuir isotherm model well. Furthermore, the fitting results of the intra-particle diffusion model indicated that intraparticle diffusion was not the sole rate-controlling step as liquid film diffusion and interior surface adsorption also regulated the adsorption rate. All the results revealed that the prepared CBC-Fe(II) adsorbent possessed high potential in environmental remediation for dye removal.

Author Contributions

Conceptualization, X.Z.; validation, X.C.; formal analysis, Z.F.G.; investigation, S.W. and S.D.; data curation, Z.F. and X.L.; writing—original draft preparation, Z.P.; writing—review and editing, Z.P., Z.F. and X.C.; supervision, W.H.; project administration, Y.S.; funding acquisition, X.C., X.Z., Z.F.G. and Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 51979011 and 52179122; the Central Non-Profit Scientific Research Fund for Institutes, grant numbers CKSF2021483/CL, CKSF2021437/CL and CKSF2021430/CL; the Knowledge Innovation Program of Wuhan-Basic Research, grant number CKSD2022360/CL; and the Jiangxi Academy of Qater Science and Engineering Open Project Fund, grant number 2021SKSG04.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors gratefully acknowledge the support from the National Natural Science Foundation of China (No. 51979011 and 52179122), the Central Non-Profit Scientific Research Fund for Institutes (No. CKSF2021483/CL, CKSF2021437/CL and CKSF2021430/CL), the Knowledge Innovation Program of Wuhan-Basic Research (No. CKSD2022360/CL) and the Jiangxi Academy of Qater Science and Engineering Open Project Fund (No. 2021SKSG04).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, J.; Liu, H.; Gao, W.; Cheng, D.; Tan, F.; Wang, W.; Wang, X.; Qiao, X.; Wong, P.K.; Yao, Y. In situ zinc cyanamide coordination induced highly N-rich graphene for efficient peroxymonosulfate activation. J. Mater. Chem. A 2022, 10, 12016–12025. [Google Scholar] [CrossRef]
  2. Chen, P.; Dong, N.; Zhang, J.; Wang, W.; Tan, F.; Wang, X.; Qiao, X.; Wong, P.K. Investigation on visible-light photocatalytic performance and mechanism of zinc peroxide for tetracycline degradation and Escherichia coli inactivation. J. Colloid Interface Sci. 2022, 624, 137–149. [Google Scholar] [CrossRef] [PubMed]
  3. Dong, N.; Ge, L.; Chen, P.; Wang, W.; Tan, F.; Wang, X.; Qiao, X.; Wong, P.K. Non-radical activation of CaO2 nanoparticles by MgNCN/MgO composites for efficient remediation of organic and heavy metal-contaminated wastewater. Sep. Purif. Technol. 2022, 285, 120334. [Google Scholar] [CrossRef]
  4. Dong, N.; Wu, D.; Ge, L.; Wang, W.; Tan, F.; Wang, X.; Qiao, X.; Wong, P.K. Constructing a brand-new advanced oxidation process system composed of MgO2 nanoparticles and MgNCN/MgO nanocomposites for organic pollutant degradation. Environ. Sci. Nano 2022, 9, 335–348. [Google Scholar] [CrossRef]
  5. Ouachtak, H.; El Haouti, R.; El Guerdaoui, A.; Haounati, R.; Amaterz, E.; Addi, A.A.; Akbal, F.; Taha, M.L. Experimental and molecular dynamics simulation study on the adsorption of Rhodamine B dye on magnetic montmorillonite composite γ-Fe2O3@Mt. J. Mol. Liq. 2020, 309, 113142. [Google Scholar] [CrossRef]
  6. Li, Y.; Du, Q.; Liu, T.; Sun, J.; Wang, Y.; Wu, S.; Wang, Z.; Xia, Y.; Xia, L. Methylene blue adsorption on graphene oxide/calcium alginate composites. Carbohyd. Polym. 2013, 95, 501–507. [Google Scholar] [CrossRef]
  7. Hussein, T.K.; Jasim, N.A. A comparison study between chemical coagulation and electro-coagulation processes for the treatment of wastewater containing reactive blue dye. Mater. Today 2021, 42, 1946–1950. [Google Scholar] [CrossRef]
  8. Li, S.; Kang, Y. Effect of PO43− on the polymerization of polyferric phosphatic sulfate and its flocculation characteristics for different simulated dye wastewater. Sep. Purif. Technol. 2021, 276, 119373. [Google Scholar] [CrossRef]
  9. Paździor, K.; Bilińska, L.; Ledakowicz, S. A review of the existing and emerging technologies in the combination of AOPs and biological processes in industrial textile wastewater treatment. Chem. Eng. J. 2019, 376, 120597. [Google Scholar] [CrossRef]
  10. Shrestha, D. Efficiency of wood-dust of dalbergia sisoo as low-cost adsorbent for Rhodamine-B dye removal. Nanomaterials 2021, 11, 2217. [Google Scholar] [CrossRef]
  11. Sun, L.; Mo, Y.; Zhang, L. A mini review on bio-electrochemical systems for the treatment of azo dye wastewater: State-of-the-art and future prospects. Chemosphere 2022, 294, 133801. [Google Scholar] [CrossRef] [PubMed]
  12. de Almeida, E.J.R.; Mazzeo, D.E.C.; Sommaggio, L.R.D.; Marin-Morales, M.A.; de Andrade, A.R.; Corso, C.R. Azo dyes degradation and mutagenicity evaluation with a combination of microbiological and oxidative discoloration treatments. Ecotox. Environ. Saf. 2019, 183, 109484. [Google Scholar] [CrossRef]
  13. Nozad, E.; Poursattar Marjani, A.; Mahmoudian, M. A novel and facile semi-IPN system in fabrication of solvent resistant nano-filtration membranes for effective separation of dye contamination in water and organic solvents. Sep. Purif. Technol. 2022, 282, 120121. [Google Scholar] [CrossRef]
  14. Zhang, J.; Hu, X.; Yan, X.; Feng, R.; Zhou, M.; Xue, J. Enhanced adsorption of Rhodamine B by magnetic nitrogen-doped porous carbon prepared from bimetallic ZIFs. Colloid. Surf. A 2019, 575, 10–17. [Google Scholar] [CrossRef]
  15. Zhang, H.; Shi, X.; Li, J.; Kumar, P.; Liu, B. Selective dye adsorption by zeolitic imidazolate framework-8 loaded UiO-66-NH2. Nanomaterials 2019, 9, 1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Elkartehi, M.E.; Mahmoud, R.; Shehata, N.; Farghali, A.; Gamil, S.; Zaher, A. LDH nanocubes synthesized with zeolite templates and their high performance as adsorbents. Nanomaterials 2021, 11, 3315. [Google Scholar] [CrossRef]
  17. Ain, Q.U.; Rasheed, U.; Yaseen, M.; Zhang, H.; He, R.; Tong, Z. Fabrication of magnetically separable 3-acrylamidopropyltrimethylammonium chloride intercalated bentonite composite for the efficient adsorption of cationic and anionic dyes. Appl. Surf. Sci. 2020, 514, 145929. [Google Scholar] [CrossRef]
  18. Paton-Carrero, A.; Sanchez, P.; Sánchez-Silva, L.; Romero, A. Graphene-based materials behaviour for dyes adsorption. Mater. Today Commun. 2022, 30, 103033. [Google Scholar] [CrossRef]
  19. Wang, H.; Li, Z.; Yahyaoui, S.; Hanafy, H.; Seliem, M.K.; Bonilla-Petriciolet, A.; Luiz Dotto, G.; Sellaoui, L.; Li, Q. Effective adsorption of dyes on an activated carbon prepared from carboxymethyl cellulose: Experiments, characterization and advanced modelling. Chem. Eng. J. 2021, 417, 128116. [Google Scholar] [CrossRef]
  20. Chabi, N.; Baghdadi, M.; Sani, A.H.; Golzary, A.; Hosseinzadeh, M. Removal of tetracycline with aluminum boride carbide and boehmite particles decorated biochar derived from algae. Bioresour. Technol. 2020, 316, 123950. [Google Scholar] [CrossRef]
  21. Binda, G.; Spanu, D.; Bettinetti, R.; Magagnin, L.; Pozzi, A.; Dossi, C. Comprehensive comparison of microalgae-derived biochar from different feedstocks: A prospective study for future environmental applications. Algal Res. 2020, 52, 102103. [Google Scholar] [CrossRef]
  22. Yu, K.L.; Lee, X.J.; Ong, H.C.; Chen, W.-H.; Chang, J.-S.; Lin, C.-S.; Show, P.L.; Ling, T.C. Adsorptive removal of cationic methylene blue and anionic Congo red dyes using wet-torrefied microalgal biochar: Equilibrium, kinetic and mechanism modeling. Environ. Pollut. 2021, 272, 115986. [Google Scholar] [CrossRef] [PubMed]
  23. Fan, Z.; Li, J.; Zhou, Y.; Fu, Q.; Yang, W.; Zhu, X.; Liao, Q. A green, cheap, high-performance carbonaceous catalyst derived from Chlorella pyrenoidosa for oxygen reduction reaction in microbial fuel cells. Int. J. Hydrogen Energ. 2017, 42, 27657–27665. [Google Scholar] [CrossRef]
  24. Fan, Z.; Li, J.; Yang, W.; Fu, Q.; Sun, K.; Song, Y.-C.; Wei, Z.; Liao, Q.; Zhu, X. Green and facile synthesis of iron oxide nanoparticle-embedded N-doped biocarbon as an efficient oxygen reduction electrocatalyst for microbial fuel cells. Chem. Eng. J. 2020, 385, 123393. [Google Scholar] [CrossRef]
  25. Wu, J.; Yang, J.; Huang, G.; Xu, C.; Lin, B. Hydrothermal carbonization synthesis of cassava slag biochar with excellent adsorption performance for Rhodamine B. J. Clean. Prod. 2020, 251, 119717. [Google Scholar] [CrossRef]
  26. Hou, Y.; Huang, G.; Li, J.; Yang, Q.; Huang, S.; Cai, J. Hydrothermal conversion of bamboo shoot shell to biochar: Preliminary studies of adsorption equilibrium and kinetics for Rhodamine B removal. J. Anal. Appl. Pyrol. 2019, 143, 104694. [Google Scholar] [CrossRef]
  27. Vigneshwaran, S.; Sirajudheen, P.; Karthikeyan, P.; Meenakshi, S. Fabrication of sulfur-doped biochar derived from tapioca peel waste with superior adsorption performance for the removal of Malachite green and Rhodamine B dyes. Surf. Interfaces 2021, 23, 100920. [Google Scholar] [CrossRef]
  28. Qu, J.; Shi, J.; Wang, Y.; Tong, H.; Zhu, Y.; Xu, L.; Wang, Y.; Zhang, B.; Tao, Y.; Dai, X.; et al. Applications of functionalized magnetic biochar in environmental remediation: A review. J. Hazard. Mater. 2022, 434, 128841. [Google Scholar] [CrossRef]
  29. Geng, J.; Gu, F.; Chang, J. Fabrication of magnetic lignosulfonate using ultrasonic-assisted in situ synthesis for efficient removal of Cr(Ⅵ) and Rhodamine B from wastewater. J. Hazard. Mater. 2019, 375, 174–181. [Google Scholar] [CrossRef]
  30. Lin, C.-R.; Ivanova, O.S.; Edelman, I.S.; Knyazev, Y.V.; Zharkov, S.M.; Petrov, D.A.; Sokolov, A.E.; Svetlitsky, E.S.; Velikanov, D.A.; Solovyov, L.A.; et al. Carbon double coated Fe3O4@C@C nanoparticles: Morphology features, magnetic properties, dye adsorption. Nanomaterials 2022, 12, 376. [Google Scholar] [CrossRef]
  31. Trinh, B.-S.; Le, P.T.K.; Werner, D.; Phuong, N.H.; Luu, T.L. Rice husk biochars modified with magnetized iron oxides and nano zero valent iron for decolorization of dyeing wastewater. Processes 2019, 7, 660. [Google Scholar] [CrossRef] [Green Version]
  32. Singh, A.; Sharma, R.; Pant, D.; Malaviya, P. Engineered algal biochar for contaminant remediation and electrochemical applications. Sci. Total Environ. 2021, 774, 145676. [Google Scholar] [CrossRef]
  33. Lalhmunsiama; Gupta, P.L.; Jung, H.; Tiwari, D.; Kong, S.-H.; Lee, S.-M. Insight into the mechanism of Cd(II) and Pb(II) removal by sustainable magnetic biosorbent precursor to Chlorella vulgaris. J. Taiwan Inst. Chem. Eng. 2017, 71, 206–213. [Google Scholar] [CrossRef]
  34. Yoon, K.; Cho, D.-W.; Wang, H.; Song, H. Co-pyrolysis route of chlorella sp. and bauxite tailings to fabricate metal-biochar as persulfate activator. Chem. Eng. J. 2022, 428, 132578. [Google Scholar] [CrossRef]
  35. Amin, M.; Chetpattananondh, P.; Khan, M.N. Ultrasound assisted adsorption of reactive dye-145 by biochars from marine Chlorella sp. extracted solid waste pyrolyzed at various temperatures. J. Environ. Chem. Eng. 2020, 8, 104403. [Google Scholar] [CrossRef]
  36. Amin, M.; Chetpattananondh, P. Biochar from extracted marine Chlorella sp. residue for high efficiency adsorption with ultrasonication to remove Cr(VI), Zn(II) and Ni(II). Bioresour. Technol. 2019, 289, 121578. [Google Scholar] [CrossRef]
  37. Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. Adsorption and photocatalytic removal of Rhodamine B from wastewater using carbon-based materials. FlatChem 2021, 29, 100277. [Google Scholar] [CrossRef]
  38. Xue, S.; Tu, B.; Li, Z.; Ma, X.; Xu, Y.; Li, M.; Fang, C.; Tao, H. Enhanced adsorption of Rhodamine B over Zoysia sinica Hance-based carbon activated by amminium chloride and sodium hydroxide treatments. Colloid. Surf. A 2021, 618, 126489. [Google Scholar] [CrossRef]
  39. Guimarães, T.; Luciano, V.A.; Silva, M.S.V.; de Carvalho Teixeira, A.P.; da Costa, M.M.; Lopes, R.P. Biochar-iron composites: An efficient material for dyes removal. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100645. [Google Scholar] [CrossRef]
  40. Lu, Y.; Chen, J.; Zhao, L.; Zhou, Z.; Qiu, C.; Li, Q. Adsorption of Rhodamine B from aqueous solution by goat manure biochar: Kinetics, isotherms, and thermodynamic studies. Pol. J. Environ. Stud. 2020, 29, 2721–2730. [Google Scholar] [CrossRef]
  41. Ye, L.; Wang, R.; Ji, G.; Wu, H.; Qu, H.; Wang, L.; Liu, J. From green tide to biochar: Thermal decomposition kinetics and TG-FTIR study of microalgae from Chaohu Lake. Int. J. Energy Res. 2020, 45, 8083–8090. [Google Scholar] [CrossRef]
  42. Xu, S.; Chen, J.; Peng, H.; Leng, S.; Li, H.; Qu, W.; Hu, Y.; Li, H.; Jiang, S.; Zhou, W.; et al. Effect of biomass type and pyrolysis temperature on nitrogen in biochar, and the comparison with hydrochar. Fuel 2021, 291, 120128. [Google Scholar] [CrossRef]
  43. Fouad, D.E.; Zhang, C.; El-Didamony, H.; Yingnan, L.; Mekuria, T.D.; Shah, A.H. Improved size, morphology and crystallinity of hematite (α-Fe2O3) nanoparticles synthesized via the precipitation route using ferric sulfate precursor. Results Phys. 2019, 12, 1253–1261. [Google Scholar] [CrossRef]
  44. Zhu, Z.; Huang, C.P.; Zhu, Y.; Wei, W.; Qin, H. A hierarchical porous adsorbent of nano-α-Fe2O3/Fe3O4 on bamboo biochar (HPA-Fe/C-B) for the removal of phosphate from water. J. Water Process Eng. 2018, 25, 96–104. [Google Scholar] [CrossRef]
  45. Bogireddy, N.K.R.; Cruz Silva, R.; Valenzuela, M.A.; Agarwal, V. 4-nitrophenol optical sensing with N doped oxidized carbon dots. J. Hazard. Mater. 2020, 386, 121643. [Google Scholar] [CrossRef]
  46. Chenakin, S.; Kruse, N. XPS characterization of transition metal oxalates. Appl. Surf. Sci. 2020, 515, 146041. [Google Scholar] [CrossRef]
  47. Peng, Z.; Wu, D.; Wang, W.; Tan, F.; Ng, T.W.; Chen, J.; Qiao, X.; Wong, P.K. Fabrication of magnetic Fe@ZnO0.6S0.4 nanocomposite for visible-light-driven photocatalytic inactivation of Escherichia coli. Appl. Surf. Sci. 2017, 396, 19–25. [Google Scholar] [CrossRef]
  48. Chen, D.; He, D.; Lu, J.; Zhong, L.; Liu, F.; Liu, J.; Yu, J.; Wan, G.; He, S.; Luo, Y. Investigation of the role of surface lattice oxygen and bulk lattice oxygen migration of cerium-based oxygen carriers: XPS and designed H2-TPR characterization. Appl. Catal. B Environ. 2017, 218, 249–259. [Google Scholar] [CrossRef]
  49. Wang, Z.; Shen, D.; Shen, F.; Wu, C.; Gu, S. Kinetics, equilibrium and thermodynamics studies on biosorption of Rhodamine B from aqueous solution by earthworm manure derived biochar. Int. Biodeter. Biodegr. 2017, 120, 104–114. [Google Scholar] [CrossRef]
  50. Adekola, F.A.; Ayodele, S.B.; Inyinbor, A.A. Activated biochar prepared from plaintain peels: Characterization and Rhodamine B adsorption data set. Chem. Data Collect. 2019, 19, 100170. [Google Scholar] [CrossRef]
  51. Ajayi, O.A.; Nisar, N.; Ifebajo, A.O.; Oladipo, A.A. High-performance magnetic chicken bone-based biochar for efficient removal of rhodamine-B dye and tetracycline: Competitive sorption analysis. Water Sci. Technol. 2017, 76, 373–385. [Google Scholar]
  52. Li, X.; Shi, J.; Luo, X. Enhanced adsorption of Rhodamine B from water by Fe-N co-modified biochar: Preparation, performance, mechanism and reusability. Bioresour. Technol. 2022, 343, 126103. [Google Scholar] [PubMed]
  53. Ren, Z.; Chen, F.; Wang, B.; Song, Z.; Zhou, Z.; Ren, D. Magnetic biochar from alkali-activated rice straw for removal of Rhodamine B from aqueous solution. Environ. Eng. Res. 2019, 25, 536–544. [Google Scholar]
Figure 1. (a) XRD patterns and (b) FTIR spectrum of CBC, CBC-Fe(III) and CBC-Fe(II).
Figure 1. (a) XRD patterns and (b) FTIR spectrum of CBC, CBC-Fe(III) and CBC-Fe(II).
Nanomaterials 12 02271 g001
Figure 2. SEM images of (a) CBC, (b) CBC-Fe(III) and (c) CBC-Fe(II).
Figure 2. SEM images of (a) CBC, (b) CBC-Fe(III) and (c) CBC-Fe(II).
Nanomaterials 12 02271 g002
Figure 3. (a) TEM and (b) HRTEM images of CBC-Fe(II).
Figure 3. (a) TEM and (b) HRTEM images of CBC-Fe(II).
Nanomaterials 12 02271 g003
Figure 4. (a) N2 adsorption–desorption isotherms and (b) pore size distributions of CBC, CBC-Fe(III) and CBC-Fe(II).
Figure 4. (a) N2 adsorption–desorption isotherms and (b) pore size distributions of CBC, CBC-Fe(III) and CBC-Fe(II).
Nanomaterials 12 02271 g004
Figure 5. Magnetic hysteresis curves of CBC-Fe(II).
Figure 5. Magnetic hysteresis curves of CBC-Fe(II).
Nanomaterials 12 02271 g005
Figure 6. (a) XPS survey spectrum; high-resolution XPS spectra of (b) C 1s; (c) Fe 2p and (d) O 1s for CBC-Fe(II).
Figure 6. (a) XPS survey spectrum; high-resolution XPS spectra of (b) C 1s; (c) Fe 2p and (d) O 1s for CBC-Fe(II).
Nanomaterials 12 02271 g006
Figure 7. The influence of adsorbent dosage on removal efficiency of RhB by CBC-Fe(II).
Figure 7. The influence of adsorbent dosage on removal efficiency of RhB by CBC-Fe(II).
Nanomaterials 12 02271 g007
Figure 8. The influence of contact time on removal efficiency of RhB by CBC, CBC-Fe(III) and CBC-Fe(II) (adsorbent dosage: 30 mg; initial dye concentration: 600 mg/L).
Figure 8. The influence of contact time on removal efficiency of RhB by CBC, CBC-Fe(III) and CBC-Fe(II) (adsorbent dosage: 30 mg; initial dye concentration: 600 mg/L).
Nanomaterials 12 02271 g008
Figure 9. The influence of temperature on removal efficiency of RhB by CBC-Fe(II) (adsorbent dosage: 30 mg).
Figure 9. The influence of temperature on removal efficiency of RhB by CBC-Fe(II) (adsorbent dosage: 30 mg).
Nanomaterials 12 02271 g009
Figure 10. (a) Pseudo-first-order and pseudo-second-order and (b) intra-particle diffusion models of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Figure 10. (a) Pseudo-first-order and pseudo-second-order and (b) intra-particle diffusion models of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Nanomaterials 12 02271 g010
Figure 11. Langmuir and Freundlich isotherms of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Figure 11. Langmuir and Freundlich isotherms of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Nanomaterials 12 02271 g011
Figure 12. Recycling test of RhB adsorption by CBC-Fe(II) (adsorbent dosage: 30 mg, initial dye concentration: 600 mg/L).
Figure 12. Recycling test of RhB adsorption by CBC-Fe(II) (adsorbent dosage: 30 mg, initial dye concentration: 600 mg/L).
Nanomaterials 12 02271 g012
Table 1. BET surface area and pore parameters of CBC, CBC-Fe(III) and CBC-Fe(II).
Table 1. BET surface area and pore parameters of CBC, CBC-Fe(III) and CBC-Fe(II).
SampleSBET (m2/g)Smicro (m2/g)Vtotal (cm3/g)Vmicro (cm3/g)Daverage (nm)
CBC338.633.60.4110.0104.855
CBC-Fe(III)350.299.90.4770.0575.702
CBC-Fe(II)527.6169.30.4990.0733.614
Table 2. Kinetic models parameters for RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Table 2. Kinetic models parameters for RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
SampleQe, Experimental
(mg/g)
Pseudo-First-OrderPseudo-Second-Order
K1
(1/min)
Qe, Model
(mg/g)
R2K2
(g/(mg∙min))
Qe, Model
(mg/g)
R2
CBC153.20.028151.10.9850.000154186.30.992
CBC-Fe(III)162.40.044158.50.9930.000273184.10.998
CBC-Fe(II)258.30.053250.70.9900.000226284.90.999
Table 3. Intra-particle diffusion model parameters for RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Table 3. Intra-particle diffusion model parameters for RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
SampleFirst Linear SegmentSecond Linear SegmentThird Linear Segment
Kid (mg/(g∙min1/2))R2Kid (mg/(g∙min1/2))R2Kid (mg/(g∙min1/2))R2
CBC22.3390.9958.9650.9990.0740.905
CBC-Fe(III)26.9620.9988.7830.9940.3480.994
CBC-Fe(II)43.3370.99612.8310.9810.2240.998
Table 4. Parameters derived from isotherms of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
Table 4. Parameters derived from isotherms of RhB adsorption by CBC, CBC-Fe(III) and CBC-Fe(II).
SampleLangmuirFreundlich
Qm (mg/g)KL (L/mg)R2KF [(mg/g)(mg/L)1/n]nR2
CBC179.70.01450.99851.7855.8080.972
CBC-Fe(III)185.10.02500.99869.7327.2120.988
CBC-Fe(II)289.60.03430.99296.3846.1440.982
Table 5. Comparison of the adsorption capacity of RhB with other biochar adsorbents reported in the literature.
Table 5. Comparison of the adsorption capacity of RhB with other biochar adsorbents reported in the literature.
BiocharBiomassQm (mg/g)References
biocharcassava slag105.3[25]
BHC-800bamboo shoot shell85.8[26]
sulfur-doped biochartapioca peel waste33.1[27]
earthworm manure derived biocharearthworm manure21.6[49]
plantain peel activated biocharplantain peel84.4[50]
magnetic chicken bone-based biocharchicken bone96.5[51]
Fe–N co-modified biocharcoconut shell12.4[52]
magnetic biochar (AMBC)Rice straw53.7[53]
CBC-Fe(II)Chlorella vulgaris286.4This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, Z.; Fan, Z.; Chen, X.; Zhou, X.; Gao, Z.F.; Deng, S.; Wan, S.; Lv, X.; Shi, Y.; Han, W. Fabrication of Nano Iron Oxide–Modified Biochar from Co-Hydrothermal Carbonization of Microalgae and Fe(II) Salt for Efficient Removal of Rhodamine B. Nanomaterials 2022, 12, 2271. https://doi.org/10.3390/nano12132271

AMA Style

Peng Z, Fan Z, Chen X, Zhou X, Gao ZF, Deng S, Wan S, Lv X, Shi Y, Han W. Fabrication of Nano Iron Oxide–Modified Biochar from Co-Hydrothermal Carbonization of Microalgae and Fe(II) Salt for Efficient Removal of Rhodamine B. Nanomaterials. 2022; 12(13):2271. https://doi.org/10.3390/nano12132271

Chicago/Turabian Style

Peng, Ziling, Zeyu Fan, Xia Chen, Xian Zhou, Zhuo Fan Gao, Shanshan Deng, Sha Wan, Xingdong Lv, Yan Shi, and Wei Han. 2022. "Fabrication of Nano Iron Oxide–Modified Biochar from Co-Hydrothermal Carbonization of Microalgae and Fe(II) Salt for Efficient Removal of Rhodamine B" Nanomaterials 12, no. 13: 2271. https://doi.org/10.3390/nano12132271

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