Next Article in Journal
Advances in Small Molecular Agents against Oral Cancer
Previous Article in Journal
Organophosphates as Versatile Substrates in Organic Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cobalt/Iron Bimetallic Biochar Composites for Lead(II) Adsorption: Mechanism and Remediation Performance

1
Key Laboratory of Water Pollution Treatment & Resource Reuse of Hainan Province, Key Laboratory of Natural Polymer Function Material of Haikou City, College of Chemistry and Chemical Engineering, Hainan Normal University, No. 99 Longkunnan Road, Haikou 571158, China
2
Hainan Pujin Environmental Technology Co., Ltd., Haikou 570125, China
3
Hainan Huantai Environmental Resources Co., Ltd., Haikou 571158, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(7), 1595; https://doi.org/10.3390/molecules29071595
Submission received: 4 March 2024 / Revised: 18 March 2024 / Accepted: 21 March 2024 / Published: 3 April 2024
(This article belongs to the Topic Technologies for Wastewater and Sludge Treatment)

Abstract

:
The performance of nano-zero-valent iron for heavy metal remediation can be enhanced via incorporation into bimetallic carbon composites. However, few economical and green approaches are available for preparing bimetallic composite materials. In this study, novel Co/Fe bimetallic biochar composites (BC@Co/Fe-X, where X = 5 or 10 represents the CoCl2 concentration of 0.05 or 0.1 mol L−1) were prepared for the adsorption of Pb2+. The effect of the concentration of cross-linked metal ions on Pb2+ adsorption was investigated, with the composite prepared using 0.05 mol L−1 Co2+ (BC@Co/Fe-5) exhibiting the highest adsorption performance. Various factors, including the adsorption period, Pb2+ concentration, and pH, affected the adsorption of Pb2+ by BC@Co/Fe-5. Further characterisation of BC@Co/Fe-5 before and after Pb2+ adsorption using methods such as X-ray diffraction and X-ray photoelectron spectroscopy suggested that the Pb2+ adsorption mechanism involved (i) Pb2+ reduction to Pb0 by Co/Fe, (ii) Co/Fe corrosion to generate Fe2+ and fix Pb2+ in the form of PbO, and (iii) Pb2+ adsorption by Co/Fe biochar. Notably, BC@Co/Fe-5 exhibited excellent remediation performance in simulated Pb2+-contaminated water and soil with good recyclability.

Graphical Abstract

1. Introduction

Currently, Pb pollution is among the most serious environmental issues. Pb enters the human body through the food chain and can detrimentally affect the central nervous system, liver, and kidneys [1]. Common methods for Pb removal include chemical precipitation, ion exchange, membrane filtration, adsorption, and electrochemical treatment [2]. Adsorption is the most widely used treatment technology owing to its low cost and high efficiency. Typical adsorbents consist of carboxymethyl cellulose/graphene oxide/Fe nanoparticles [2], modified chitosan hydrogels [3], analcime-activated carbon composites [4], nano-zero-valent iron (nZVI) [5], and amino-carboxyl cellulose [6]. nZVI has been extensively used in the remediation of hazardous groundwater, permeable reactive barriers, and contaminated subsoil [5,7]. Although nZVI has many benefits for treating heavy metal pollution [8], including high mobility, high reactivity, excellent reducibility, and a large specific surface area, issues such as easy passivation in air and easy agglomeration limit its applicability to environmental remediation [9]. Enhanced stability and higher adsorption capacities have frequently been achieved via composite formation or surface modification. The combination of nZVI with porous carbon materials such as carbon nanotubes [10], activated carbon [11], and biochar [12,13], which have unique pore structures, good chemical stability, and high electrical conductivity, can effectively reduce agglomeration and increase pollutant removal efficiency [14]. In particular, bimetallic carbon composites are considered promising for improving the reactivity of nZVI.
Xing et al. [15] prepared a novel mesoporous Santa Barbara Amorphous-15-supported Fe/Ni bimetallic composite (SBA-15@Fe/Ni) to remove Cr(VI). Their results showed that the removal efficiency of SBA-15@Fe/Ni is better than two separate systems (SBA-15 and Fe/Ni). Introducing Ni can promote Cr(VI) reduction through electron transfer and catalytic hydrogenation. Yuan et al. [16] prepared microscale iron–copper (mFe/Cu) bimetals wherein copper coated the surface of mFe0 particles. Electrochemical analysis revealed that copper plating promotes the release of electrons from mFe0 and reduces the impedance of mFe0. Fe and Co reportedly have synergistic catalytic effects; a low percentage of Co can improve the reduction activity of Fe [17]. However, research on contaminant removal by Fe/Co composites has rarely been reported [18]. Qin et al. [19] prepared FeCo bimetallic nanoparticles via a hydrothermal reduction method, which can achieve higher activity and significantly improve the reaction kinetics for removing Cr(VI) compared with those of Fe0. Hong et al. [20] prepared a novel 2D metal–organic complex that exhibits highly efficient removal performance for trace Pb2+ in neutral aqueous solutions (pH 7).
The primary technique for preparing bimetallic composites is liquid-phase reduction, in which metal ions such as Cu2+ and Ni2+ are reduced using NaBH4 and loaded onto the surface of Fe. However, the NaBH4 reduction process is expensive and produces numerous secondary pollutants [21]. In addition, Fe/Co bimetals prepared by hydrothermal methods generally exist as alloys and have low activity when treating pollutants. Therefore, developing economical and green methods for preparing bimetallic composite materials is highly desirable. The carbothermal method has recently gained popularity as an environmentally friendly approach for preparing Fe/C composites [22]; zero-valent metals were produced during the carbothermic reduction process and loaded onto biochar [23]. Nonetheless, carbon-supported bimetallic nanoparticles synthesised using this method have rarely been reported.
Sodium alginate (SA), a renewable natural resource, contains a large number of carboxyl and hydroxyl groups capable of interacting with divalent (e.g., Ca2+) or trivalent (e.g., Fe3+) ions to form hydrogels [24]. Motivated by this property, SA-Fe3+/Co2+ gels were prepared and used as precursors for carbothermal reduction synthesis, with Co/Fe-embedded biochar (BC@Co/Fe-X, where X represents the Co2+ concentration during gel preparation) produced via one-step pyrolysis. This synthesis has several advantages: (ⅰ) compared with the traditional metal salt impregnation method, SA-Fe3+/Co2+ gels, as precursors, can better ensure the dispersion of metal ions; (ⅱ) Fe3+(Co2+)–O clusters were isolated by ligands in the complex and then reduced in situ to bimetallic Co/Fe nanoparticles, which ensured their stability; and (ⅲ) ligands were pyrolysed to biochar, which provides conductive and large surface area support to immobile nanoparticles.
Therefore, the objective of this study was to prepare efficient bimetallic composite materials using a hydrothermal reduction method and investigate the adsorption mechanisms for Pb2+. The effect of Co doping on the Pb2+ adsorption performance of the composites was analysed. The obtained BC@Co/Fe-X composites exhibited a large surface area, well-designed structure, and efficient adsorption performance. Furthermore, the Pb2+ adsorption mechanism of BC@Co/Fe-5 and its remediation performance in Pb2+-contaminated water and soil were investigated.

2. Results and Discussion

2.1. Characterisation

From the SEM images of BC@Co/Fe-5 shown in Figure 1a,b, the surface of BC@Co/Fe-5 is irregular and porous, and the metals are distributed nonuniformly. As shown in the HRTEM images of BC@Co/Fe-5 (Figure 1c,d), the metal particles have irregular shapes with sizes of 50–100 nm. The lattice fringe spacing of 0.203 nm (Figure 1d) corresponds to the (110) crystal plane of Fe [25]. Furthermore, an oxide film was observed on the outer layer of Fe [26]. The HRTEM mapping images of BC@Co/Fe-5 (Figure 1e–h) reveal that C and O have dense distributions, whereas Fe and Co have relatively disperse and uneven distributions.
XRD was used to investigate the crystal structures of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10 (Figure 2a). All these materials exhibit two relatively strong diffraction peaks at 2θ = 44.7° and 65.0°, which correspond to the (110) and (200) crystal planes, respectively, of Fe0 (PDF#06-0696) [27]. No characteristic Co peaks were observed, which was consistent with the relatively low Co contents of BC@Co/Fe-5 and BC@Co/Fe-10. Hysteresis loops were observed in the N2 adsorption–desorption isotherms of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10 (Figure 2b). According to the IUPAC classification, the N2 adsorption–desorption isotherms are type IV and the hysteresis loops are type H2 [28]. The specific surface areas of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10 are 464, 429, and 547 m2 g−1, respectively. The variation in the specific surface area is related to the type of aperture formed. The relatively high specific surface area of BC@Co/Fe-10 is related to its mainly microporous structure. The total pore volume and average pore size parameters of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10 are listed in Table 1. The pore size distributions (Figure 2c) reveal that BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10 exhibit both microporous and mesoporous structures. Moreover, the proportion of micropores in the pore size distribution of BC@Co/Fe-10 is relatively high. The diameter of Pb2+ is 0.266 nm [29]. The average pore size parameters of BC@Co/Fe-10 is the smallest, about 2.65 nm, which inevitably increases the difficulty of Pb2+ diffusion. The magnetisation curves of BC@Co/Fe-5 and BC@Co/Fe-5-Pb (Figure 2d) revealed that the coercivity and remanence values of these materials are very small, which is typical for superparamagnetic materials [30]. The saturation magnetisations of BC@Co/Fe-5 and BC@Co/Fe-5-Pb are 90.2 and 70.4 emu g−1, respectively. As BC@Co/Fe-5-Pb has high magnetisation, after Pb2+ adsorption, the composite material can be easily collected and separated using a magnet.

2.2. Performance of BC@Fe and BC@Co/Fe-X

The Fe and Co contents of BC@Fe and BC@Co/Fe-X were determined following digestion, as shown in Figure 3a. BC@Fe has the highest Fe content of 59.5% and C and O contents of 32% and 5.2%, respectively. The content of Fe in BC@Fe is higher than that of zero-valent iron/biochar (Fe0-BC) prepared via carbothermal reduction using wood waste and iron mud as raw materials [31]. The Fe contents of BC@Co/Fe-5 and BC@Co/Fe-10 decrease significantly as the Co content increases, attributed to the complexing ability of different metal ions with alginate, which follows the sequence Fe3+ > Cu2+ > Co2+ [32]. Furthermore, Co2+ can significantly inhibit alginate cross-linking Fe3+. The Pb2+ adsorption properties of BC@Fe and BC@Co/Fe-X are compared in Figure 3b. The Pb2+ adsorption capacity of BC@Fe is the lowest. With increasing Co doping, the Pb2+ adsorption capacity of BC@Co/Fe-X first increases and then decreases. The decreased Pb2+ adsorption capacity of BC@Co/Fe-10 may be related to its lower Fe loading. The mass fractions of Fe and Co in BC@Co/Fe-5 are 45.3% and 0.35%, respectively.

2.3. Effect of pH on Pb2+ Adsorption

The effect of pH on the Pb2+ adsorption performance of BC@Co/Fe-5 was investigated (Figure 4a). At pH 2, the Pb2+ adsorption capacity of BC@Co/Fe-5 is low, mainly because excess H+ can compete with Pb2+ for the adsorption sites. In addition, the higher H+ concentration at low pH favours the formation of Fe2+ from Fe0. The formation of iron oxides is difficult under low pH conditions, contributing to the low adsorption performance of BC@Co/Fe-5 [8]. As the H+ concentration decreases (pH 3–5.5), the Pb2+ adsorption capacity of BC@Co/Fe-5 increases [8]. After Pb2+ adsorption by BC@Co/Fe-5, the pH of the solution slightly increases (Figure 4b). Pb2+ forms hydrated ions at pH values higher than 6 and precipitates at pH values higher than 8. Therefore, Pb2+ adsorption experiments are generally conducted at pH values less than 7 [33]. Considering that Pb2+ remains in the form of ions and does not precipitate at pH values less than 5 [34], pH 5 was selected as optimal for Pb2+ adsorption in this study.

2.4. Adsorption Kinetics

The kinetic data for Pb2+ adsorption by BC@CoFe-5 were fitted using pseudo-first-order and pseudo-second-order kinetic models. The obtained fitting parameters are summarised in Table 2, and Figure 5 shows the Pb2+ adsorption capacity of BC@Co/Fe-5 over time. As shown in Figure 5, the adsorption rate of Pb2+ on BC@Co/Fe-5 is very high over the first 6 h, likely because of a redox reaction between Fe and Pb2+ on the surface. Subsequently, the adsorption rate gradually decreases. During this stage, internal Fe may corrode to form Fe2+, which would convert Pb2+ into PbO.
As shown in Table 2, the correlation coefficient (R2 = 0.98) of the pseudo-second-order kinetic model is higher than that of the pseudo-first-order kinetic model (R2 = 0.95) for Pb2+ adsorption by BC@Co/Fe-5. However, the theoretical adsorption capacity (128.7 mg g−1) estimated from the pseudo-second-order equation was slightly higher compared to the experimental results (116.5 mg g−1). In addition, the theoretical adsorption capacity (111.9 mg g−1) obtained from the pseudo-first-order kinetic model was more consistent with the experimental results, suggesting that the rate-limiting step was governed via diffusion while the ions were mainly retained on the surface via chemical forces [35,36].

2.5. Adsorption Isotherms

The experimental data for Pb2+ adsorption by BC@Co/Fe-5 were fitted using the Langmuir and Freundlich models. The fitted curves for Pb2+ adsorption by BC@Co/Fe-5 are shown in Figure 6, with the fitting parameters listed in Table 3.
While the correlation coefficient (R2 = 0.98) of the Langmuir model is higher than that of the Freundlich model (R2 = 0.967) for Pb2+ adsorption by BC@Co/Fe-5, the theoretical adsorption capacity of the Langmuir model (2076.6 mg g−1) is higher than the actual adsorption capacity (1240 mg g−1), possibly because the Langmuir isotherm model assumes that active sites with identical energy exist on the surface of the adsorbent, which facilitates homogenous or surface-specific monolayer adsorption typical of chemical adsorption processes [37]. However, redox reactions and corrosion adsorption are the main adsorption mechanisms of Pb2+ by BC@Co/Fe-5. The Pb2+ adsorption process may be multilayer adsorption, confirmable via cyclic adsorption experiments. The values of n, which represent the favourability of the adsorption, were more than 1.0, indicating that the adsorption of Pb2+ by BC@Co/Fe-5 was favourable [38]. Table 4 compares the Pb2+ adsorption performance of BC@Co/Fe-5 and those of previously reported materials. Notably, BC@Co/Fe-5 exhibits superior Pb2+ adsorption performance.

2.6. Pb2+ Adsorption Mechanisms

The FTIR spectra of BC@Co/Fe-5 before and after Pb2+ adsorption are shown in Figure 7a. For BC@Co/Fe-5 before Pb2+ adsorption, the strong, wide absorption peak at 3440.49 cm−1 corresponds to the stretching vibration of O–H [44], the absorption peak at 1625.77 cm−1 corresponds to the asymmetric stretching vibration of COO [44], the absorption peaks at 1571.03 and 1549.02 cm−1 are attributed to the vibration of the benzene ring skeleton [45], the absorption peak at 1118.05 cm−1 is attributed to the C–O–C stretching vibration [44], and the absorption peak at 565.17 cm−1 corresponds to the stretching vibration of Fe–O [44]. The Fe–O absorption peak appears because BC@Co/Fe-5 is oxidised to form ferrite during grinding. After Pb2+ adsorption by BC@Co/Fe-5, the characteristic absorption peaks at 3440.49, 1625.77, 1571.03, 1549.02, and 1118.05 cm−1 are blue-shifted to 3343.49, 1630.33, 1583.27, 1568.35, and 1135.30 cm−1, respectively [46]. This blue shift of the peaks corresponding to O–H, COO, the benzene ring skeleton, and C–O–C indicates the formation of a complex with Pb2+. The increased ionic volume weakened the stretching and bending vibrations of the functional groups, resulting in a blue shift [47].
The XRD pattern of BC@Co/Fe-5-Pb (Figure 7b) exhibits two strong diffraction peaks at 2θ = 44.7° and 65.0°, characteristic of Fe0 (PDF#06-0696) [27], indicating that the content of Fe0 in BC@Co/Fe-5 remains high after Pb2+ adsorption. In addition, new diffraction peaks appear at 2θ = 31.3°, 36.26°, 52.2°, and 62.1°, corresponding to the (111), (200), (220), and (311) crystal planes, respectively, of Pb (JCPDS no. 04-0686) [48]. The diffraction peaks at 2θ = 28.6°, 31.8°, 35.7°, and 48.5° correspond to the (101), (110), (002), and (112) crystal planes, respectively, of PbO (JCPDS no. 65-0399) [49]. The diffraction peak at 2θ = 57.1° corresponds to the (600) crystal plane of FeOOH (JCPDS no. 22-0353) [50]. These results indicate that Pb, PbO, and FeOOH formed on the surface of BC@Co/Fe-5 after Pb2+ adsorption.
The XPS survey spectra of BC@Co/Fe-5 before and after Pb2+ adsorption are shown in Figure 8a. An obvious Pb absorption peak appears in the XPS spectrum of BC@Co/Fe-5-Pb. The Pb XPS spectrum after Pb2+ adsorption by BC@Co/Fe-5 is shown in Figure 8b. The Pb 4f7/2 and Pb 4f5/2 peaks can each be fitted to three peaks [51]. The peak areas of the fitted peaks at 136.7 and 141.8 eV account for 35.8% of the total peak area, suggesting that some Pb2+ is reduced to Pb and adsorbed in this form [52]. The peak areas of the fitted peaks at 137.4 and 142.45 eV, which may correspond to PbO [52], account for 38.5% of the total peak area. The peak areas of the fitted peaks at 138.15 and 143.2 eV account for 25.7% of the total peak area, suggesting that Pb2+ could form a complex with active groups and be adsorbed as a Pb2+ complex [52].
The Fe XPS spectrum of BC@Co/Fe-5 (Figure 8c) exhibits Fe 2p3/2 and Fe 2p1/2 peaks. The fitted peaks at 707.19, 711.20, 714.50, 719.00, 724.30, and 730.40 eV correspond to Fe0 2p3/2, Fe2+ 2p3/2, Fe3+ 2p3/2, Fe0 2p1/2, Fe2+ 2p1/2, and Fe3+ 2p1/2, respectively [53]. Fitting of the Fe 2p3/2 and Fe 2p1/2 peaks of BC@Co/Fe-5-Pb led to peaks at 710.7, 712.1, 719.00, 724.30, and 726.5 eV, corresponding to Fe2+ 2p3/2, Fe3+ 2p3/2, Fe0 2p1/2, Fe2+ 2p1/2, and Fe3+ 2p1/2, respectively. The disappearance of the Fe0 2p3/2 peak at 707.19 eV indicates that Fe is consumed during Pb2+ adsorption.
The Co XPS spectrum of BC@Co/Fe-5 is shown in Figure 8d. Fitting yielded four peaks at 780.5, 782.2, 785.9, and 791.5 eV, corresponding to Co0, Co2+, Co3+, and satellites, respectively. Fitting of the Co XPS spectrum of BC@Co/Fe-5-Pb also afforded four peaks [54]; the peaks at 780.5, 782.4, 786.4, and 791.5 eV correspond to Co0, Co2+, Co3+, and satellites, respectively [54,55]. The changes in the binding energies of the characteristic Fe and Co peaks before and after adsorption may be due to the reactions with Pb2+. The Fe morphologies in BC@Co/Fe-5 revealed by the XPS (Figure 8c) and XRD (Figure 2a) analyses differ slightly. As the depth of the XPS analysis is limited to within 5 nm of the surface of the material, the results are affected by Fe oxidation that occurs on the surface layer during transportation and testing. In contrast, the depth of the XRD analysis is generally several to dozens of micrometres, and surface oxidation has little effect on the results.
The C XPS spectrum of BC@Co/Fe-5 (Figure 8e) exhibits major absorption peaks corresponding to C–C/C=C (284.70 eV), C–O (285.15 eV), C=O (286.2 eV), and COO (289.15 eV) [56]. The change in the binding energies of C–C/C=C, C–O, C=O, and COO after Pb2+ adsorption by BC@Co/Fe-5 suggests that these functionalities form complexes with Pb2+ during the adsorption process, consistent with the FTIR spectroscopy results (Figure 7a).
The FTIR, XRD, and XPS results before and after Pb2+ adsorption by BC@Co/Fe-5 indicate that most Pb2+ was converted into Pb and PbO, as shown in Equations (1)–(5), while the remainder was adsorbed as Pb2+ [57]. The proposed adsorption mechanism is illustrated in Figure 9.
2 F e 0 + 3 P b 2 + + 4 H 2 O C o 3 P b 0 + 2 F e O O H + 6 H +
2 F e 0 + O 2 + 4 H + 2 F e 2 + + 2 H 2 O
F e 2 + + H 2 O F e O H + + H +
F e O H + + P b 2 + P b O H + + F e 2 +
P b O H + P b O + H +

2.7. Recyclability of BC@Co/Fe-5

Figure 10 shows the Pb2+ adsorption performance of BC@Co/Fe-5 over four cycles. The adsorption capacities of BC@Co/Fe-5 for Pb2+ in the second, third, and fourth cycles are 29.8%, 39.6%, and 51.3%, respectively, lower than the initial adsorption capacity. Thus, BC@Co/Fe-5 maintains 48.7% of the initial adsorption capacity after four cycles. The cyclic performance likely deteriorates because the adsorption of Pb2+ by Co and Fe inside BC@Co/Fe-5 is relatively difficult. The newly added high concentration of Pb2+ increases the chance of contact between Fe and Pb2+ and the adsorption amount of Pb2+. In addition, the presence of H+ in the Pb2+ solution during subsequent cycles can promote the corrosion of Fe inside BC@Co/Fe-5 to form Fe2+, which can participate in Pb2+ adsorption [58].

2.8. Pb2+ Removal from Contaminated Wastewater and Soil

The Pb2+ adsorption results for BC@Co/Fe-5 in simulated wastewater are shown in Figure 11a. Upon the addition of 0.05 and 0.15 g BC@Co/Fe-5, the Pb2+ removal rates reach 62.9% and 91.7%, respectively, under the condition of background ions at five times their standard concentrations. When the concentrations of the background ions are increased to ten times their standard values, the Pb2+ removal rates of 0.05 and 0.15 g BC@Co/Fe-5 decrease to 56.7% and 73.2%, respectively.
The results for the remediation of simulated Pb2+-contaminated soil are shown in Figure 11b. When 1, 2, or 3 wt.% BC@Co/Fe-5 is added to the soil, the Pb2+ content decreases by 30.2%, 41.7%, and 58%, respectively, after 7 d, and by 59.6%, 68.9%, and 80.7% after 21 d. The remediation effect does not change significantly after 28 d relative to that after 21 d, suggesting that the system was close to equilibrium after 21 d. Upon increasing the BC@Co/Fe-5 dosage from 1 wt.% to 3 wt.%, the remediation effect on Pb2+-contaminated soil increases gradually.

3. Materials and Methods

3.1. Materials

The materials used in this study included SA, FeCl3·6H2O, and CoCl2 (AR, Aladdin, Shanghai, China); NaOH (AR, Shanghai Yi En Chemical Technology Co., Ltd., Shanghai, China); PbCl2 and NaNO3 (AR, Shanghai Baishun Biotechnology Co., Ltd., Shanghai, China); Pb and Co standard solutions (China National Academy of Metrology Sciences, Beijing, China); and pH buffer reagent (Apure, Shanghai, China).

3.2. Preparation of BC@Co/Fe-X

In a beaker, 30 g SA was dissolved in 1 L of distilled water by stirring (B90-SH, Guangzhou Gangran Electromechanical Equipment Co., Ltd., Guangzhou, China). The SA solution was added dropwise to a solution containing 0.3 mol L−1 FeCl3 and 0.05 or 0.1 mol L−1 CoCl2, with the mixture allowed to stand for 24 h to undergo cross-linking. The obtained gel was recovered via filtration, with metal ions on the surface removed by washing with distilled water (6–8 times). The gel was dried in a vacuum oven (JHG-9053A, Shanghai Jinghong Experimental Equipment Co., Ltd., Shanghai, China) at 60 °C and then at 100 °C for 3 h. The dried SA-Fe3+/Co2+ gel was placed in a crucible and inserted into the quartz tube of a tube furnace (OTF1200X, Hefei Kejing Material Technology Co., Ltd., Hefei, China). After evacuation for 15 min, high-purity nitrogen was introduced at a flow rate of 200 mL min−1 for 20 min. Under nitrogen at a flow rate of 100 mL min−1, the temperature was increased to 200 °C at a heating rate of 5 °C min−1 and then to 900 °C at a heating rate of 10 °C min−1. After pyrolysis at 900 °C for 3 h, the furnace was cooled to room temperature at a cooling rate of 10 °C min−1. The prepared Co/Fe biochar composites (BC@Co/Fe-X, where X = 5 or 10 represents the CoCl2 concentration of 0.05 or 0.1 mol L−1) were stored under vacuum [36].

3.3. Characterisation

Scanning electron microscopy (SEM) images were obtained using a JSM-7400F microscope to observe the basic morphology of the as-prepared samples. Transmission electron microscopy (TEM), high-resolution TEM (HRTEM), and elemental mapping analyses were conducted on a JEM-2100Plus microscope (JEOL Ltd., Akishima-shi, Japan). X-ray diffraction (XRD) measurements were performed using an UItima IV diffractometer (Rigaku, Tokyo Metropolis, Japan) with a Cu Kα light source (λ = 0.15418 nm) at an operating voltage of 40 kV and an operating current of 10 mA in the scanning range of 2θ = 10–80°. Magnetisation curves were obtained using a SQUID-VSM magnetic measuring system (Quantum Design, San Diego, CA, USA). N2 adsorption–desorption isotherms were obtained using a Tristar II 3020 surface area and porosity analyser (Micromeritics, Atlanta, GA, USA). X-ray photoelectron spectroscopy (XPS) measurements were conducted using an ESCALAB 250Xi spectrometer (Thermo Fisher, Waltham, MA, USA) with an Al Kα light source (hv = 1486.6 eV) at a power of 150 W and a 500 µm spot energy analyser at a transmission energy of 30 eV. Fourier transform infrared (FTIR) spectra were recorded using a Thermo Fisher 6700 FTIR spectrometer (Thermo Fisher, Waltham, MA, USA). The elemental analysis of carbon and oxygen was conducted using a Vario EL cube elemental analyser (Elementar, Frankfurt, Germany).

3.4. Batch Sorption Experiments

To compare the Pb2+ adsorption performance of different materials, 0.05 g of BC@Fe, BC@Co/Fe-5, or BC@Co/Fe-10 was added to 100 mL of a 100 mg L−1 Pb2+ solution (pH 5) in a conical flask. The conical flasks were placed in a constant temperature shaker at 25 °C for 24 h. To compare the effect of pH on Pb2+ adsorption by BC@Co/Fe-5, 0.05 g BC@Co/Fe-5 was added to 100 mL of a Pb2+ solution (100 mg L−1) with pH values of 2, 3, 4, 5, or 5.5 in a conical flask. The conical flasks were placed in a constant temperature shaker at 25 °C for 24 h. To analyse the kinetics of Pb2+ adsorption by BC@Co/Fe-5, 0.05 g BC@Co/Fe-5 was added to 100 mL of Pb2+ solution (100 mg L−1) in a conical flask [9]. The conical flask was placed in a constant temperature shaker at 25 °C. At fixed times (1–24 h), 1 mL of the Pb2+ solution was removed and diluted for analysis [59]. For thermodynamic experiments, 0.05 g BC@Co/Fe-5 was added to 100 mL of a Pb2+ solution (pH 5) with concentrations of 300–1000 mg L−1 in a conical flask. The conical flasks were placed in a constant temperature shaker at 25 °C for 24 h, and then 1 mL of the Pb2+ solution was removed and diluted for analysis [59]. To investigate the recyclability, 0.05 g BC@Co/Fe-5 was added to 100 mL of a Pb2+ solution (100 mg L−1), with the mixture shaken at 25 °C for 24 h. After the adsorption process was complete, the solution was removed. Then, fresh Pb2+ solution was added, with the mixture shaken again for 24 h under the same conditions [58]. The adsorption performance of BC@Co/Fe-5 was investigated over four cycles.
The data for the kinetic adsorption of Pb2+ by BC@Co/Fe-5 were fitted using pseudo-first-order and pseudo-second-order kinetic models (Equations (6) and (7), respectively) [60]. From a mathematical point of view, the pseudo-first-order model well describes the kinetic processes limited from beginning to end by film diffusion at the phase boundary. It also describes the final stages (the so-called regular modes) of particle diffusion processes with a good approximation [2]. The pseudo-second-order model explains the chemisorption of heavy metal ions onto adsorbents [61].
q t = q e ( 1 - e   k 1 t )
q t = k 2 q e 2 t 1 + k 2 q e t
where qt is the adsorption capacity at time t (mg g−1), qe is the adsorption capacity at equilibrium (mg g−1), k1 is the first-order rate constant (min−1), k2 is the second-order rate constant (g mg−1 min−1), and t is time (min).
The thermodynamic data for Pb2+ adsorption by BC@Co/Fe-5 were fitted using the Langmuir and Freundlich models (Equations (8) and (9), respectively) [62].
q e = q m K L c e 1 + K L c e
q e = K F c e 1 / n
where qe is the adsorption capacity at adsorption equilibrium (mg g−1), qm is the adsorption capacity at saturation (mg g−1), ce is the concentration of Pb2+ in solution at adsorption equilibrium (mg L−1), KL is the Langmuir adsorption constant (L mg−1), and KF is the Freundlich adsorption constant (mg1−n Ln g−1).

3.5. Pb2+ Adsorption by BC@Co/Fe-5 in Simulated Samples

To investigate the ability of BC@Co/Fe-5 to remove Pb2+ from sewage, Class V water from the China groundwater standard (GB/T 14848-2017) [63], containing 350 mg L−1 Cl, 350 mg L−1 SO42−, 350 mg L−1 CO32−, 300 mg L−1 NO3, 1.5 mg L−1 Cu2+, and 5 mg L−1 Zn2+, was used as the background [64]. The Pb2+ (100 mg L−1) removal efficiency was investigated by adding 0.05 or 0.15 g BC@Co/Fe-5 to 100 mL of a background solution containing Cl, SO42−, CO32−, NO3, Cu2+, and Zn2+ at concentrations five and ten times higher than those in the Class V groundwater standard.
To investigate the ability of BC@Co/Fe-5 to remove Pb2+ from the soil, surface soil (0–20 cm) was collected from Guilinyang Farm, Haikou City, and dried in the shade. The soil was crushed using a soil crusher and passed through a 60-mesh sieve. Simulated contaminated soil with a Pb content of 800 mg kg−1 was prepared by adding a PbCl2 solution to the soil. The contaminated soil was dried in the shade, crushed, and passed through a 60-mesh sieve. Then, 25 g simulated contaminated soil was mixed with BC@Co/Fe-5 (1, 2, and 3 wt.%), and deionised water was added to achieve a soil moisture content of 75%. The soil was stored in the dark and weighed every 2 d, after which water was added to maintain a constant soil moisture content. Soil samples were removed and freeze-dried on days 7, 14, 21, and 28 [65].
The Pb2+ content was measured using toxicity characteristic leaching procedure extraction and atomic absorption methods, as regulated by the United States Environmental Protection Agency (1992); these methods have been widely used to analyse the toxicity characteristics of toxic elements in polluted soil [66].

4. Conclusions

In this study, well-dispersed and polyporous BC@Co/Fe-X was prepared via carbothermal reduction. Co-doping improved the Pb2+ adsorption capacity of BC@Co/Fe-X, with the highest adsorption capacity observed for BC@Co/Fe-5, which had Fe and Co mass fractions of 45.3% and 0.35%, respectively. Notably, such a low Co content is not harmful to the environment. The mechanism of Pb2+ adsorption on BC@Co/Fe-5 comprised the following processes: (i) Co/Fe reduction of Pb2+ to Pb0, (ii) Co/Fe corrosion to form Fe2+ and fix Pb2+ as PbO, and (iii) Pb2+ adsorption by the Co/Fe biochar. The redox reaction and corrosion adsorption are the main adsorption mechanisms. Kinetic fitting to the result indicated that the rate-limiting step was governed via diffusion while the ions were mainly retained on the surface via chemical forces. BC@Co/Fe-5 maintains 48.7% of its initial Pb2+ adsorption capacity after four cycles. Furthermore, the acidic and high concentration of Pb2+ facilitated the iron corrosion continuously generated, resulting in highly effective adsorption of Pb2+. Isotherms fitting to the result indicated that the Pb2+ adsorption process may be multilayer adsorption, confirmable via cyclic adsorption experiments. Reasonable Pb2+ removal effects were observed in simulated wastewater, with 0.15 g BC@Co/Fe-5 achieving Pb2+ removal rates of up to 91.7% and 73.2% under the condition of background ions at five and ten times their standard concentrations, respectively. For soil remediation, the addition of 3 wt.% BC@Co/Fe-5 reduced the Pb2+ content of the soil by 83.5% after 28 d. Overall, BC@Co/Fe-5 is a cost-effective and environmentally sustainable composite, holding promise for the remediation and treatment of Pb2+ pollution.

Author Contributions

Validation, C.Y.; Investigation, J.Z. (Jingyu Zhao), Y.Q., Y.L., Y.S., Q.L., Z.J., A.S., Y.F. and J.Z. (Jiayi Zhang); Resources, M.C.; Data curation, C.Y.; Writing—original draft, J.Z. (Jingyu Zhao); Writing—review & editing, C.Y.; Project administration, M.C.; Funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology special fund of the Hainan Province of China (Project No. ZDYF2022SHFZ279; ZDYF2020079).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

We gratefully acknowledge the Engineering Technology Research Center of Soild Waste Treatment & Disposal and Soil Remediation of Haikou City for providing the necessary equipment for this study.

Conflicts of Interest

Author Miao Cai was employed by the Hainan Pujin Environmental Technology Co., Ltd., Author Zhenya Jia was employed by the Hainan Huantai Environmental Resources Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

BC@Co/Fe-XCo/Fe biochar composite
CBCcorn cob biochar
ceconcentration at adsorption equilibrium (mg L−1)
FTIRFourier transform infrared
HRTEMhigh-resolution transmission electron microscopy
k1first-order rate constant (min−1)
k2second-order rate constant (g mg−1 min−1)
KLLangmuir adsorption constant (L g−1)
KFFreundlich adsorption constant (mg g−1)
nZVInano-zero-valent iron
qeadsorption capacity at equilibrium (mg g−1)
qmadsorption capacity at saturation (mg g−1)
qtadsorption capacity at time t (mg g−1)
SExstandard error
RSSresidual sum of squares
SAsodium alginate
SEMscanning electron microscopy
ttime (min)
XPSX-ray photoelectron spectroscopy
XRDX-ray diffraction

References

  1. Zhang, X.; Lin, J.H.; Li, T.T.; Wang, Z.K.; Shiu, B.C.; Lou, C.W. A study on modified fabric-contained melamine/bentonite polyurethane foam: Pb2+ adsorption and mechanical properties. Polym. Compos. 2023, 44, 1877–1888. [Google Scholar] [CrossRef]
  2. Neskoromnaya, E.A.; Khamizov, R.K.; Melezhyk, A.V.; Memetova, A.E.; Mkrtchan, E.S.; Babkin, A.V. Adsorption of lead ions (Pb2+) from wastewater using effective nanocomposite GO/CMC/FeNPs: Kinetic, isotherm, and desorption studies. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130224. [Google Scholar] [CrossRef]
  3. Hao, D.; Liang, Y. Adsorption of Cu2+, Cd2+ and Pb2+ in wastewater by modified chitosan hydrogel. Environ. Technol. 2022, 43, 876–884. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Q.; Lv, L.; Zhao, X.; Wang, Y.; Wang, Y. Cost-effective microwave-assisted hydrothermal rapid synthesis of analcime-activated carbon composite from coal gangue used for Pb2+ adsorption. Environ. Sci. Pollut. Res. Int. 2022, 29, 77788–77799. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, Y.-Z.; Zhou, W.-H.; Liu, F.; Yi, S.; Geng, X. Microstructure and morphological characterization of lead-contaminated clay with nanoscale zero-valent iron (nZVI) treatment. Eng. Geol. 2019, 256, 84–92. [Google Scholar] [CrossRef]
  6. Gou, J.; Zhang, W.; Wang, X.-F.; Hao, D.; Shen, H.; You, N.; Long, W.-Y. Amino-carboxyl cellulose for adsorption of Cd2+ and Pb2+. Chemosphere 2023, 339, 139705. [Google Scholar] [CrossRef] [PubMed]
  7. Bagbi, Y.; Sarswat, A.; Tiwari, S.; Mohan, D.; Pandey, A.; Solanki, P.R. Synthesis of l -cysteine stabilized zero-valent iron (nZVI) nanoparticles for lead remediation from water. Environ. Nanotechnol. Monit. Manag. 2017, 7, 34–45. [Google Scholar] [CrossRef]
  8. Jiao, W.; Song, Y.; Zhang, D.; Chang, G.; Fan, H.; Liu, Y. Nanoscale zero-valent iron modified with carboxymethyl cellulose in an impinging stream-rotating packed bed for the removal of lead(II). Adv. Powder Technol. 2019, 30, 2251–2261. [Google Scholar] [CrossRef]
  9. Wang, Z.; Wu, X.; Luo, S.; Wang, Y.; Tong, Z.; Deng, Q. Shell biomass material supported nano-zero valent iron to remove Pb2+ and Cd2+ in water. R. Soc. Open Sci. 2020, 7, 201192. [Google Scholar] [CrossRef]
  10. Vilardi, G.; Mpouras, T.; Dermatas, D.; Verdone, N.; Polydera, A.; Di Palma, L. Nanomaterials application for heavy metals recovery from polluted water: The combination of nano zero-valent iron and carbon nanotubes. Competitive adsorption non-linear modeling. Chemosphere 2018, 201, 716–729. [Google Scholar] [CrossRef]
  11. Rashtbari, Y.; Sher, F.; Afshin, S.; Hamzezadeh, A.; Ahmadi, S.; Azhar, O.; Rastegar, A.; Ghosh, S.; Poureshgh, Y. Green synthesis of zero-valent iron nanoparticles and loading effect on activated carbon for furfural adsorption. Chemosphere 2022, 287, 132114. [Google Scholar] [CrossRef] [PubMed]
  12. Chouli, F.; Ezzat, A.O.; Sabantina, L.; Benyoucef, A.; Zehhaf, A. Optimization Conditions of Malachite Green Adsorption onto Almond Shell Carbon Waste Using Process Design. Molecules 2023, 29, 54. [Google Scholar] [CrossRef] [PubMed]
  13. Gu, S.; Cui, J.; Liu, F.; Chen, J. Biochar loaded with cobalt ferrate activated persulfate to degrade naphthalene. RSC Adv. 2023, 13, 5283–5292. [Google Scholar] [CrossRef]
  14. Vega, R.; Rong, R.; Dai, M.; Ali, I.; Naz, I.; Peng, C. Fe-C-based materials: Synthesis modulation for the remediation of environmental pollutants-a review. Environ. Sci. Pollut. Res. Int. 2022, 29, 64345–64369. [Google Scholar] [CrossRef]
  15. Xing, X.; Ren, X.; Alharbi, N.S.; Chen, C. Efficient adsorption and reduction of Cr(VI) from aqueous solution by Santa Barbara Amorphous-15 (SBA-15) supported Fe/Ni bimetallic nanoparticles. J. Colloid Interface Sci. 2023, 629, 744–754. [Google Scholar] [CrossRef]
  16. Yuan, Y.; Zhou, Z.; Zhang, X.; Li, X.; Liu, Y.; Yang, S.; Lai, B. Efficient reduction of hexavalent chromium with microscale Fe/Cu bimetals: Efficiency and the role of Cu. Chin. Chem. Lett. 2023, 34, 107932. [Google Scholar] [CrossRef]
  17. Sun, Y.; Wang, K.; Chen, D.; Xu, Q.; Li, N.; Li, H.; Lu, J. Activation of persulfate by highly dispersed FeCo bimetallic alloy for in-situ remediation of polycyclic aromatic hydrocarbon-contaminated soil. Sep. Purif. Technol. 2023, 317, 123781. [Google Scholar] [CrossRef]
  18. Gao, J.F.; Wu, Z.L.; Duan, W.J.; Zhang, W.Z. Simultaneous adsorption and degradation of triclosan by Ginkgo biloba L. stabilized Fe/Co bimetallic nanoparticles. Sci. Total Environ. 2019, 662, 978–989. [Google Scholar] [CrossRef] [PubMed]
  19. Qin, N.; Zhang, Y.; Zhou, H.; Geng, Z.; Liu, G.; Zhang, Y.; Zhao, H.; Wang, G. Enhanced removal of trace Cr(VI) from neutral and alkaline aqueous solution by FeCo bimetallic nanoparticles. J. Colloid Interface Sci. 2016, 472, 8–15. [Google Scholar] [CrossRef]
  20. Hong, J.; Kang, L.; Shi, X.; Wei, R.; Mai, X.; Pan, D.; Naik, N.; Guo, Z. Highly efficient removal of trace lead (II) from wastewater by 1,4-dicarboxybenzene modified Fe/Co metal organic nanosheets. J. Mater. Sci. Technol. 2022, 98, 212–218. [Google Scholar] [CrossRef]
  21. Bakshi, S.; Banik, C.; Rathke, S.J.; Laird, D.A. Arsenic sorption on zero-valent iron-biochar complexes. Water Res. 2018, 137, 153–163. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Q.; Ye, X.; Chen, D.; Xiao, W.; Zhao, S.; Li, J.; Li, H. Chromium(VI) removal from synthetic solution using novel zero-valent iron biochar composites derived from iron-rich sludge via one-pot synthesis. J. Water Process Eng. 2022, 47, 102720. [Google Scholar] [CrossRef]
  23. Hu, Z.; Rao, J.; Xie, Z.; Liu, M.; Su, L.; Chen, Y.; Gao, W.; Tan, Y.; Zhou, Z.; Zhou, N. A green in-situ synthesis of biochar-supported Fe0/Cu0 bimetallic catalyst for the efficient oxidation antibacterial in water: Performance and mechanism analysis. J. Water Process Eng. 2023, 51, 103424. [Google Scholar] [CrossRef]
  24. Hu, X.; Long, L.; Gong, T.; Zhang, J.; Yan, J.; Xue, Y. Enhanced alginate-based microsphere with the pore-forming agent for efficient removal of Cu2+. Chemosphere 2020, 240, 124860. [Google Scholar] [CrossRef] [PubMed]
  25. Arafat, Y.; Azhar, M.R.; Zhong, Y.; O’Hayre, R.; Tadé, M.O.; Shao, Z. Organic ligand-facilitatedin situexsolution of CoFe alloys over Ba0.5Sr0.5Co0.8Fe0.2O3−δperovskite toward enhanced oxygen electrocatalysis for rechargeable Zn-air batteries. J. Mater. Chem. A 2023, 11, 12856–12865. [Google Scholar] [CrossRef]
  26. Lari, L.; Steinhauer, S.; Lazarov, V.K. In situ TEM oxidation study of Fe thin-film transformation to single-crystal magnetite nanoparticles. J. Mater. Sci. 2020, 55, 12897–12905. [Google Scholar] [CrossRef]
  27. Jiang, S.-F.; Ling, L.-L.; Chen, W.-J.; Liu, W.-J.; Li, D.-C.; Jiang, H. High efficient removal of bisphenol A in a peroxymonosulfate/iron functionalized biochar system: Mechanistic elucidation and quantification of the contributors. Chem. Eng. J. 2019, 359, 572–583. [Google Scholar] [CrossRef]
  28. Huang, W.; Tang, Y.; Zhang, X.; Luo, Z.; Zhang, J. nZVI-biochar derived from Fe3O4-loaded rabbit manure for activation of peroxymonosulfate to degrade sulfamethoxazole. J. Water Process Eng. 2022, 45, 102470. [Google Scholar] [CrossRef]
  29. Fansuri, H.; Supriadi, W.; Ediati, R.; Utomo, W.P.; Hidayati, R.E.; Iqbal, R.M.; Sulistiono, D.O.; Abdullah, M.M.A.B.; Subaer. Immobilization and leaching behavior of Cd2+ and Pb2+ heavy metal ions in Indonesian fly ash-based geopolymers. Environ. Adv. 2024, 15, 100510. [Google Scholar] [CrossRef]
  30. Saeed, A.A.H.; Harun, N.Y.; Sufian, S.; Bilad, M.R.; Zakaria, Z.Y.; Jagaba, A.H.; Ghaleb, A.A.S.; Mohammed, H.G. Pristine and Magnetic Kenaf Fiber Biochar for Cd2+ Adsorption from Aqueous Solution. Int. J. Environ. Res. Public Health 2021, 18, 7949. [Google Scholar] [CrossRef]
  31. Chen, C.; Liu, J.; Gen, C.; Liu, Q.; Zhu, X.; Qi, W.; Wang, F. Synthesis of zero-valent iron/biochar by carbothermal reduction from wood waste and iron mud for removing rhodamine B. Environ. Sci. Pollut. Res. Int. 2021, 28, 48556–48568. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, M.; Luo, L.; Fu, L.; Yang, H. Ion responsiveness of polyacrylamide/sodium alginate (PAM/SA) shape memory hydrogel. Soft Mater. 2019, 17, 418–426. [Google Scholar] [CrossRef]
  33. Xu, L.; Liu, Y.; Wang, J.; Tang, Y.; Zhang, Z. Selective adsorption of Pb2+ and Cu2+ on amino-modified attapulgite: Kinetic, thermal dynamic and DFT studies. J. Hazard. Mater. 2021, 404, 124140. [Google Scholar] [CrossRef] [PubMed]
  34. Wei, E.; Wang, K.; Muhammad, Y.; Chen, S.; Dong, D.; Wei, Y.; Fujita, T. Preparation and conversion mechanism of different geopolymer-based zeolite microspheres and their adsorption properties for Pb2+. Sep. Purif. Technol. 2022, 282, 119971. [Google Scholar] [CrossRef]
  35. Xu, L.; Bai, T.; Yi, X.; Zhao, K.; Shi, W.; Dai, F.; Wei, J.; Wang, J.; Shi, C. Polypropylene fiber grafted calcium alginate with mesoporous silica for adsorption of Bisphenol A and Pb2+. Int. J. Biol. Macromol. 2023, 238, 124131. [Google Scholar] [CrossRef]
  36. Shi, Y.; Yu, C.; Liu, M.; Lin, Q.; Lei, M.; Wang, D.; Yang, M.; Yang, Y.; Ma, J.; Jia, Z. One-pot synthesis of spherical nanoscale zero-valent iron/biochar composites for efficient removal of Pb(ii). RSC Adv. 2021, 11, 36826–36835. [Google Scholar] [CrossRef] [PubMed]
  37. Akinterinwa, A.; Oladele, E.; Adebayo, A.; Adamu, M. Characterization of aqueous Pb2+ adsorption onto cross-linked-carboxymethyl legume starch phosphate using FTIR and SEM-EDX. Biomass Convers. Biorefinery 2023. [Google Scholar] [CrossRef]
  38. Sun, J.; Chen, Y.; Yu, H.; Yan, L.; Du, B.; Pei, Z. Removal of Cu2+, Cd2+ and Pb2+ from aqueous solutions by magnetic alginate microsphere based on Fe3O4/MgAl-layered double hydroxide. J. Colloid Interface Sci. 2018, 532, 474–484. [Google Scholar] [CrossRef]
  39. Li, J.; Wang, J.; Yuan, X.; Wang, Z.; Sun, S.; Lyu, Q.; Hu, S. Efficient adsorption of BPA and Pb2+ by sulfhydryl-rich β-cyclodextrin polymers. Sep. Purif. Technol. 2023, 309, 122913. [Google Scholar] [CrossRef]
  40. Jiang, J.; Li, R.; Yang, K.; Li, Y.; Deng, L.; Che, D. Investigation on Pb2+ adsorption characteristics by AAEMs-rich biochar in aqueous solution: Performance and mechanism. Environ. Res. 2023, 236 Pt 1, 116731. [Google Scholar] [CrossRef]
  41. Farghali, M.A.; Abo-Aly, M.M.; Salaheldin, T.A. Modified mesoporous zeolite-A/reduced graphene oxide nanocomposite for dual removal of methylene blue and Pb2+ ions from wastewater. Inorg. Chem. Commun. 2021, 126, 108487. [Google Scholar] [CrossRef]
  42. Abukhadra, M.R.; Bakry, B.M.; Adlii, A.; Yakout, S.M.; El-Zaidy, M.E. Facile conversion of kaolinite into clay nanotubes (KNTs) of enhanced adsorption properties for toxic heavy metals (Zn2+, Cd2+, Pb2+, and Cr6+) from water. J. Hazard. Mater. 2019, 374, 296–308. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, H.; Zhao, B.; Wang, L. Adsorption/desorption performance of Pb2+ and Cd2+ with super adsorption capacity of PASP/CMS hydrogel. Water Sci. Technol. 2021, 84, 43–54. [Google Scholar] [CrossRef]
  44. Zhang, J.; Li, X.; Xu, H.; Zhang, W.; Feng, X.; Yao, Y.; Ma, Y.; Su, L.; Ren, S.; Li, S. Removal of Cd2+, Pb2+ and Ni2+ from water by adsorption onto magnetic composites prepared using humic acid from waste biomass. J. Clean. Prod. 2023, 411, 137237. [Google Scholar] [CrossRef]
  45. Samaraweera, H.; Nawalage, S.; Nayanathara, R.M.O.; Peiris, C.; Karunaratne, T.N.; Gunatilake, S.R.; Thirumalai, R.V.K.G.; Zhang, J.; Zhang, X.; Mlsna, T. In Situ Synthesis of Zero-Valent Iron-Decorated Lignite Carbon for Aqueous Heavy Metal Remediation. Processes 2022, 10, 1659. [Google Scholar] [CrossRef]
  46. Fang, Q.; Chen, B.; Lin, Y.; Guan, Y. Aromatic and hydrophobic surfaces of wood-derived biochar enhance perchlorate adsorption via hydrogen bonding to oxygen-containing organic groups. Environ. Sci. Technol. 2014, 48, 279–288. [Google Scholar] [CrossRef] [PubMed]
  47. Yu, C.; Li, H.; Ma, H.; Zhang, L.; Li, Y.; Lin, Q. Characteristics and mechanism of Cu(II) adsorption on prepared calcium alginate/carboxymethyl cellulose@MnFe2O4. Polym. Bull. 2021, 76, 1201–1216. [Google Scholar] [CrossRef]
  48. Wang, Q.; Guan, Y.; Yan, J.; Liu, Y.; Shao, Q.; Ning, F.; Yi, J. Facile synthesis of lead-tin nanoparticles for electrocatalyzing carbon dioxide reduction to formate. Dalton Trans. 2023, 52, 4136–4141. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, G.; Gou, B.; Yang, Y.; Liu, M.; Li, X.; Xiao, L.; Hao, G.; Zhao, F.; Jiang, W. CuO/PbO Nanocomposite: Preparation and Catalysis for Ammonium Perchlorate Thermal Decomposition. ACS Omega 2020, 5, 32667–32676. [Google Scholar] [CrossRef]
  50. Wu, L.; Yu, L.; McElhenny, B.; Xing, X.; Luo, D.; Zhang, F.; Bao, J.; Chen, S.; Ren, Z. Rational design of core-shell-structured CoP @FeOOH for efficient seawater electrolysis. Appl. Catal. B Environ. 2021, 294, 120256. [Google Scholar] [CrossRef]
  51. Li, Z.; Wang, L.; Meng, J.; Liu, X.; Xu, J.; Wang, F.; Brookes, P. Zeolite-supported nanoscale zero-valent iron: New findings on simultaneous adsorption of Cd(II), Pb(II), and As(III) in aqueous solution and soil. J. Hazard. Mater. 2018, 344, 1–11. [Google Scholar] [CrossRef] [PubMed]
  52. Diao, Z.H.; Du, J.J.; Jiang, D.; Kong, L.J.; Huo, W.Y.; Liu, C.M.; Wu, Q.H.; Xu, X.R. Insights into the simultaneous removal of Cr6+ and Pb2+ by a novel sewage sludge-derived biochar immobilized nanoscale zero valent iron: Coexistence effect and mechanism. Sci. Total Environ. 2018, 642, 505–515. [Google Scholar] [CrossRef]
  53. Zhai, Q.; Liu, R.; Wang, C.; Wen, X.; Li, X.; Sun, W. A novel scheme for the utilization of Cu slag flotation tailings in preparing internal electrolysis materials to degrade printing and dyeing wastewater. J. Hazard. Mater. 2022, 424, 127537. [Google Scholar] [CrossRef] [PubMed]
  54. Li, J.C.; Meng, Y.; Zhang, L.; Li, G.; Shi, Z.; Hou, P.X.; Liu, C.; Cheng, H.M.; Shao, M. Dual-Phasic Carbon with Co Single Atoms and Nanoparticles as a Bifunctional Oxygen Electrocatalyst for Rechargeable Zn–Air Batteries. Adv. Funct. Mater. 2021, 31, 2103360. [Google Scholar] [CrossRef]
  55. Zhang, L.; Cai, W.; Bao, N. Top-Level Design Strategy to Construct an Advanced High-Entropy Co–Cu–Fe–Mo (Oxy)Hydroxide Electrocatalyst for the Oxygen Evolution Reaction. Adv. Mater. 2021, 33, 2100745. [Google Scholar] [CrossRef] [PubMed]
  56. Hong, J.; Li, J.; Kang, L.; Gao, L.; Shi, X. Preparation of novel terephthalic acid modified Fe/Ni metal organic nanosheet with high adsorption performance for trace Pb2+. Appl. Surf. Sci. 2022, 579, 152268. [Google Scholar] [CrossRef]
  57. Huang, P.; Ye, Z.; Xie, W.; Chen, Q.; Li, J.; Xu, Z.; Yao, M. Rapid magnetic removal of aqueous heavy metals and their relevant mechanisms using nanoscale zero valent iron (nZVI) particles. Water Res. 2013, 47, 4050–4058. [Google Scholar] [CrossRef]
  58. Zeng, Y.; Walker, H.; Zhu, Q. Reduction of nitrate by NaY zeolite supported Fe, Cu/Fe and Mn/Fe nanoparticles. J. Hazard. Mater. 2017, 324, 605–616. [Google Scholar] [CrossRef] [PubMed]
  59. Xie, H.; Zhang, S.; Zhong, L.; Wang, Q.; Hu, J.; Tang, A. Effect of the occurrence state of magnesium in talc on the adsorption of Pb(II). J. Alloys Compd. 2021, 887, 161288. [Google Scholar] [CrossRef]
  60. Bo, S.; Luo, J.; An, Q.; Xiao, Z.; Wang, H.; Cai, W.; Zhai, S.; Li, Z. Efficiently selective adsorption of Pb(II) with functionalized alginate-based adsorbent in batch/column systems: Mechanism and application simulation. J. Clean. Prod. 2020, 250, 119585. [Google Scholar] [CrossRef]
  61. Joshi, N.C.; Rawat, B.S.; Kumar, P.; Kumar, N.; Upadhyay, S.; Chetana, S.; Gururani, P.; Kimothi, S. Sustainable synthetic approach and applications of ZnO/r-GO in the adsorption of toxic Pb2+ and Cr6+ ions. Inorg. Chem. Commun. 2022, 145, 110040. [Google Scholar] [CrossRef]
  62. Yang, Y.; Xie, Y.; Pang, L.; Li, M.; Song, X.; Wen, J.; Zhao, H. Preparation of Reduced Graphene Oxide/Poly(acrylamide) Nanocomposite and Its Adsorption of Pb(II) and Methylene Blue. Langmuir 2013, 29, 10727–10736. [Google Scholar] [CrossRef] [PubMed]
  63. GB/T 14848-2017; Standard for Groundwater Quality. General Administration of Quality Supervision, Inspection and Quarantine of the People’s Republic of China: Beijing, China.
  64. Zhang, Q.; Wang, L.; Wang, H.; Zhu, X.; Wang, L. Spatio-Temporal Variation of Groundwater Quality and Source Apportionment Using Multivariate Statistical Techniques for the Hutuo River Alluvial-Pluvial Fan, China. Int. J. Environ. Res. Public Health 2020, 17, 1055. [Google Scholar] [CrossRef] [PubMed]
  65. Frick, H.; Tardif, S.; Kandeler, E.; Holm, P.E.; Brandt, K.K. Assessment of biochar and zero-valent iron for in-situ remediation of chromated copper arsenate contaminated soil. Sci. Total Environ. 2019, 655, 414–422. [Google Scholar] [CrossRef] [PubMed]
  66. Aborisade, M.A.; Feng, A.; Oba, B.T.; Kumar, A.; Battamo, A.Y.; Huang, M.; Chen, D.; Yang, Y.; Sun, P.; Zhao, L. Pyrolytic synthesis and performance efficacy comparison of biochar-supported nanoscale zero-valent iron on soil polluted with toxic metals. Arch. Agron. Soil Sci. 2022, 69, 2249–2266. [Google Scholar] [CrossRef]
Figure 1. (a,b) SEM, (c,d) HRTEM, and (eh) HRTEM mapping images of BC@Co/Fe-5.
Figure 1. (a,b) SEM, (c,d) HRTEM, and (eh) HRTEM mapping images of BC@Co/Fe-5.
Molecules 29 01595 g001
Figure 2. (a) XRD patterns, (b) N2 adsorption–desorption isotherms, and (c) pore size distributions of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10. (d) Magnetisation curves of BC@Co/Fe-5 and BC@Co/Fe-5-Pb.
Figure 2. (a) XRD patterns, (b) N2 adsorption–desorption isotherms, and (c) pore size distributions of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10. (d) Magnetisation curves of BC@Co/Fe-5 and BC@Co/Fe-5-Pb.
Molecules 29 01595 g002
Figure 3. (a) Fe and Co contents and (b) Pb2+ adsorption properties of BC@Fe and BC@Co/Fe-X.
Figure 3. (a) Fe and Co contents and (b) Pb2+ adsorption properties of BC@Fe and BC@Co/Fe-X.
Molecules 29 01595 g003
Figure 4. (a) Effect of initial pH on Pb2+ adsorption by BC@Co/Fe-5, and (b) the pH of the Pb2+ solution after adsorption by BC@Co/Fe-5.
Figure 4. (a) Effect of initial pH on Pb2+ adsorption by BC@Co/Fe-5, and (b) the pH of the Pb2+ solution after adsorption by BC@Co/Fe-5.
Molecules 29 01595 g004
Figure 5. Pb2+ adsorption kinetics of BC@Co/Fe-5.
Figure 5. Pb2+ adsorption kinetics of BC@Co/Fe-5.
Molecules 29 01595 g005
Figure 6. Fitted isotherms for Pb2+ adsorption by BC@Co/Fe-5.
Figure 6. Fitted isotherms for Pb2+ adsorption by BC@Co/Fe-5.
Molecules 29 01595 g006
Figure 7. (a) FTIR spectra of BC@Co/Fe-5 before and after Pb2+ adsorption. (b) XRD pattern of BC@Co/Fe-5 after Pb2+ adsorption.
Figure 7. (a) FTIR spectra of BC@Co/Fe-5 before and after Pb2+ adsorption. (b) XRD pattern of BC@Co/Fe-5 after Pb2+ adsorption.
Molecules 29 01595 g007
Figure 8. (a) XPS survey spectra of BC@Co/Fe-5 before and after Pb2+ adsorption. (b) Pb 4f XPS results for BC@Co/Fe-5 after Pb2+ adsorption. (c) Fe 2p, (d) Co 2p, and (e) C 1s XPS results for BC@Co/Fe-5 before and after Pb2+ adsorption.
Figure 8. (a) XPS survey spectra of BC@Co/Fe-5 before and after Pb2+ adsorption. (b) Pb 4f XPS results for BC@Co/Fe-5 after Pb2+ adsorption. (c) Fe 2p, (d) Co 2p, and (e) C 1s XPS results for BC@Co/Fe-5 before and after Pb2+ adsorption.
Molecules 29 01595 g008
Figure 9. Mechanism of Pb2+ adsorption by BC@Co/Fe-5.
Figure 9. Mechanism of Pb2+ adsorption by BC@Co/Fe-5.
Molecules 29 01595 g009
Figure 10. Recyclability of BC@Co/Fe-5.
Figure 10. Recyclability of BC@Co/Fe-5.
Molecules 29 01595 g010
Figure 11. Remediation of (a) simulated Pb2+-containing wastewater and (b) simulated Pb2+-contaminated soil using BC@Co/Fe-5.
Figure 11. Remediation of (a) simulated Pb2+-containing wastewater and (b) simulated Pb2+-contaminated soil using BC@Co/Fe-5.
Molecules 29 01595 g011
Table 1. Textural properties of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10.
Table 1. Textural properties of BC@Fe, BC@Co/Fe-5, and BC@Co/Fe-10.
SampleSurface Area
(m2 g−1)
Total Pore Volume
(cm3 g−1)
Average Pore Size Parameters
(nm)
BC@Fe4640.3033.53
BC@Co/Fe-54290.2943.21
BC@Co/Fe-105470.3432.65
Table 2. Constants and correlation coefficients for the Pb2+ adsorption kinetics of BC@Co/Fe-5.
Table 2. Constants and correlation coefficients for the Pb2+ adsorption kinetics of BC@Co/Fe-5.
SamplePseudo-First-Order ModelPseudo-Second-Order Model
BC@Co/Fe-5qe,cal
(mg g−1)
SExk1 (10−3 min−1)R2RSSReduced chi-squaredqe,cal
(mg g−1)
SExk2 (10−5
g mg−1 min−1)
R2RSSReduced chi-squared
111.93.096.480.9517.22.5128.72.926.360.986.00.87
Table 3. Isotherm fitting parameters for Pb2+ adsorption by BC@Co/Fe-5.
Table 3. Isotherm fitting parameters for Pb2+ adsorption by BC@Co/Fe-5.
SampleLangmuirFreundlich
BC@Co/Fe-5qm,cal
(mg g−1)
SExKL
(L mg−1)
R2RSSReduced chi-squarednKF
(mg1−n Ln g−1)
SExR2RSSReduced chi-squared
2076.6155.80.00460.9843.37.221.6639.767.630.96791.215.20
Table 4. Pb2+ adsorption properties of various materials.
Table 4. Pb2+ adsorption properties of various materials.
AdsorbentpHConcentration Range (mg L−1)qm (mg g−1)Ref.
Modified chitosan hydrogel5100–1100420.98[3]
Amino-carboxyl cellulose5100–1000117.6[6]
Analcime-activated carbon composite5.5100125.57[4]
Sulfhydryl-rich
β-cyclodextrin polymers
--604.64[39]
Alkali and alkaline earth metal-rich biochar5.510–4000226.64[40]
Mesoporous zeolite-A/reduced graphene oxide6.5-416.7[41]
Carboxymethyl cellulose-nZVI6100–10001376[8]
Geopolymer-based zeolite microspheres5100–600529.7[34]
Carboxymethylcellulose/graphene oxide + Fe 18%655–10501850[2]
Kaolinite nanotubes625–3501428[42]
Polyaspartic acid/carboxymethyl salix psammophila hydrogel5.52000–10,0001954[43]
BC@Co/Fe-55300–10001240This study
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

Zhao, J.; Qin, Y.; Liu, Y.; Shi, Y.; Lin, Q.; Cai, M.; Jia, Z.; Yu, C.; Shang, A.; Fei, Y.; et al. Cobalt/Iron Bimetallic Biochar Composites for Lead(II) Adsorption: Mechanism and Remediation Performance. Molecules 2024, 29, 1595. https://doi.org/10.3390/molecules29071595

AMA Style

Zhao J, Qin Y, Liu Y, Shi Y, Lin Q, Cai M, Jia Z, Yu C, Shang A, Fei Y, et al. Cobalt/Iron Bimetallic Biochar Composites for Lead(II) Adsorption: Mechanism and Remediation Performance. Molecules. 2024; 29(7):1595. https://doi.org/10.3390/molecules29071595

Chicago/Turabian Style

Zhao, Jingyu, Yuhong Qin, Yue Liu, Yunlong Shi, Qiang Lin, Miao Cai, Zhenya Jia, Changjiang Yu, Anqi Shang, Yuxiao Fei, and et al. 2024. "Cobalt/Iron Bimetallic Biochar Composites for Lead(II) Adsorption: Mechanism and Remediation Performance" Molecules 29, no. 7: 1595. https://doi.org/10.3390/molecules29071595

Article Metrics

Back to TopTop