Next Article in Journal
Symmetry of the Non-Analytic Solution of the Dirac Equation Inside the Proton of Hydrogen Atoms
Previous Article in Journal
Improving Domain Transfer with Consistency-Regularized Joint Distribution Alignment for Medical Image Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Application for Pb2+ Removal of a Novel Magnetic Biochar Embedded with FexOy Nanoparticles

1
School of Materials Engineering, Wuhu Institute of Technology, Wuhu 241003, China
2
College of Chemistry and Chemical Engineering, Anhui University, Hefei 230601, China
3
School of Chemistry and Materials Science, Anhui Normal University, Wuhu 241002, China
*
Author to whom correspondence should be addressed.
Symmetry 2025, 17(4), 516; https://doi.org/10.3390/sym17040516
Submission received: 16 February 2025 / Revised: 14 March 2025 / Accepted: 22 March 2025 / Published: 28 March 2025
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
Biochar (BC) is a widely studied economic and environment-friendly material. However, its application is limited by its underdeveloped pore structure, small specific surface area, low degree of graphitization, and difficulty in being separated from liquids during the application process. Raw cotton contains almost 100% cellulose and has a high yield for preparing biochar. A novel type of magnetic biochar composite was prepared by the impregnation–calcination method using cotton and iron nitrate nonahydrate(Fe(NO3)3·9H2O). The material was characterized using scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), transmission electron microscopy (TEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, thermogravimetric analysis (TG), and specific surface area testing (BET). The results show that FexOy nanoparticles with a uniform morphology and an average particle size of less than 20 nm are embedded in the composite; the saturation magnetization strength of the composite material reaches 21.6 emu/g; and compared to the original biochar, the composite material has a larger specific surface area (326 m2/g). As an adsorbent, the composite material has a fast removal rate for Pb2+ of 95% in 50 min. The Langmuir model calculation results show that the maximum adsorption capacity of the composite for Pb2+ is 252.7 mg/g. FexOy-BC easily achieves solid–liquid separation and can be recycled for Pb2+ wastewater treatment through adsorption–desorption–regeneration.

1. Introduction

Heavy metal pollution in soil and water poses a serious threat to ecosystems and human health. At present, heavy metal treatment technologies for wastewater mainly include chemical precipitation [1], membrane separation [2], electrochemical treatment [3], coagulation [4], biodegradation [5], solvent extraction [6], and adsorption methods [7,8]. Among them, adsorption technology is recognized as one of the most promising environmental remediation methods due to its advantages of low cost, easy operation, high efficiency, and application of renewable materials [9,10,11,12].
In the adsorption material system, biochar (BC) has become a research hotspot due to its easy availability of raw materials, structural stability, and efficient adsorption performance. Biochar is an aromatic solid product formed by the limited oxygen pyrolysis and carbonization of biomass. Its structural characteristics are characterized by a developed multi-stage pore system, high specific surface area, and abundant surface functional groups (carboxyl, carbonyl, phenolic hydroxyl, etc.) [13]. It can capture heavy metal ions effectively through multiple mechanisms, such as ion exchange, chemical precipitation, surface complexation, and physical adsorption [14]. However, the low density and micrometer particle size characteristics of the material severely limit its further application due to its solid–liquid separation efficiency. To this end, the researchers have proposed a magnetic functionalization modification strategy: loading magnetic components (such as Fe3O4) onto a biochar matrix to construct composite materials, and using an external magnetic field to achieve rapid solid–liquid separation, which can significantly improve the recyclability of the material [15].
The current preparation technology of magnetic biochar mainly includes chemical methods (such as co-precipitation [16] and hydrothermal synthesis [17]) and physical methods (such as mechanical ball milling [18,19]). Zhang et al. used the co-precipitation method to load nano-Fe3O4 onto rice husk biochar under oxygen limited conditions at 400 °C, resulting in a composite material with a specific surface area of 101.9 m2/g [20]. The result shows that the BC-Fe composite material can effectively reduce the bioavailability of cadmium in soil, inhibit the absorption and transport of cadmium by rice, and provide a theoretical basis for the remediation of heavy metal passivation in farmland.
In response to the demand for the collaborative treatment of heavy metal pollution, Liu et al. developed a composite adsorbent based on palm fiber-derived graphene biochar (GB) loaded with nano zerovalent iron (nZVI) [21]. The material was prepared by the liquid-phase reduction method (pyrolysis temperature: 700 °C), achieving the synchronous removal of Cd (II)-As (III): the adsorption capacity of As (III) in an acidic environment (pH = 4) reached 363 mg/g and the removal efficiency of Cd (II) in neutral conditions (pH = 7) was 92.8 mg/g. It is worth noting that this system overcomes the technical bottleneck of traditional zerovalent iron being prone to oxidation. Recently, Peng et al. designed Fe3O4-modified corn stover biochar, which achieved a maximum lead adsorption capacity of 113.70 mg/g through the co-precipitation pyrolysis coupling process, demonstrating its excellent cationic heavy metal fixation ability [22].
Although the abovementioned modification strategies significantly improve the adsorption performance of biochar, its application still faces multiple constraints: firstly, the preparation process relies on highly corrosive precursors (such as concentrated hydrochloric acid) and toxic reducing agents (such as NaBH4), which poses a risk of secondary environmental pollution. Secondly, the multi-step synthesis process involves complex operations, such as high-temperature carbonization, liquid-phase reduction, and surface modification, resulting in high material production costs. In addition, the insufficient bonding strength at heterogeneous interfaces may lead to the leaching of active components. Therefore, the development of a new magnetic biochar material system that combines efficient adsorption performance, environment-friendly characteristics, and simple preparation processes remains a key scientific problem that urgently needs to be addressed in the field of environmental functional materials. It needs to solve fundamental problems, such as green synthesis process development, interface stability enhancement, and full lifecycle cost control.
Lead (Pb), as a typical heavy metal pollutant, poses great harm to human health, and low concentrations can also cause irreversible organic damage to the kidney and liver organs [23,24,25]. In aquatic ecosystems, the high toxicity and difficult degradation characteristics of lead, combined with its bioaccumulation effect, make its efficient removal one of the core challenges in the field of environmental pollution control [26].
In view of the abovementioned issues and challenges, the purpose of this work is as follows:
(i)
Develop new biochar composites. To our knowledge, cotton, as one of the abundant, low-cost, and easily accessible raw materials, has not been reported for the synthesis of magnetic composites. Meanwhile, the cellulose content of cotton is almost 100%, much higher than other wood materials. Under the same quality conditions, the yield of biochar prepared from cotton is higher.
(ii)
Explore simple synthesis methods for magnetic composites. BC-FexOy was prepared by impregnation calcination using cotton and Fe(NO3)3·9H2O as precursors without the addition of other auxiliary reagents. The process is simple and easy. Characterization shows that FexOy nanoparticles are embedded in BC, indicating the structural stability of the composite.
(iii)
Investigate the application of composites. Taking the removal of Pd (II) from wastewater as an example, this study explores the adsorption capacity and recycling characteristics of the composite material, proving that BC-FexOy is an efficient and low-cost new magnetic adsorption composite.

2. Materials and Methods

2.1. Synthesis of Materials

In order to avoid environmental pollution caused by waste, the cotton used in the experiment was sourced from a school’s waste collection station. Although raw cotton is indeed a high-cost commodity, utilizing waste cotton can significantly reduce material costs and environmental burdens. The waste cotton, which was cleaned, shredded, and dried, primarily consisted of cellulose and associated organic matter, comprising approximately 98% of its weight, with an ash content of approximately 2%. The main constituents of ash are sodium, magnesium, and calcium salts. The chemicals used in the experiments, including ethanol (95%), NaOH, CH3COOH, HCl, Fe(NO3)3, and Pb(NO3)2, were analytically pure and obtained from Aladdin Reagent (Shanghai, China) Co., Ltd. Each experiment used 18.2 MΩ·cm−1 of deionized water. The main process involved weighing 6.4368 g of cotton, compacting it into a crucible made of quartz (SJM-40, Hanhui Technology Co., Ltd., Shenzhen, China), covering it, and then placing it in a muffle furnace for calcination (JC-MF10-16, Juchuang Environmental Protection Group Co., Ltd., Qingdao, China). The starting temperature was set to be 30 °C. The heating rate was increased by 5 °C per min, and the heating lasted for 110 min. The insulation was maintained at 600 °C for 60 min. The final mass of the biochar was 2.6391 g, which was about 41% of the initial mass, higher than the yield of biochar prepared from rice and corn straw [27].
For the synthesis of magnetic biochar, in a typical experiment, 0.4046 g of Fe(NO3)3·9H2O is weighed and put into a small beaker. Then, 5 mL of deionized water was slowly added with ultrasonic assistance for dissolution. Next, 1.200 g of biochar was placed in another beaker. The Fe(NO3)3 solution was then applied dropwise onto the biochar, and subsequently, 2 mL of ultrapure water was added. During this time, a glass rod was used to stir the mixture to guarantee complete impregnation. It was then left to stand at ambient temperature for approximately 2 h. Subsequently, it was dried in an oven at 120 °C. The dried precursor was placed in a crucible and calcined in a muffle furnace for the calcination process. The starting temperature was 30 °C. The heating rate was set to be 5 °C/min, until the temperature reached to 550 °C, while keeping the insulation duration for 60 min. A slightly brownish–black flocculent solid (BC-FexOy) was obtained. The synthetic process is illustrated in Figure 1.
The following are the chemical equations that are involved in the whole preparation process:
Waste   cotton   ( cellulose )   oxygen   deficit > 500   ° C   Biochar
2 F e ( N O 3 ) 3 · 9 H 2 O = = = = = = < 300   ° C 6 H N O 3 + 2 F e ( O H ) 3
F e ( O H ) 3 = = = = = = Δ F e 2 O 3 + 3 H 2 O
4 F e ( N O 3 ) 3 = = = = = = > 300   ° C 2 F e 2 O 3 + 12 N O 2 + 3 O 2
6 F e 2 O 3 + C = = = = = = = = = high   temperature 4 F e 3 O 4 + C O 2

2.2. Method of Analysis

SEM equipped with an energy spectrometer (TESCAN S8000, Tescan Trading (Shanghai) Co., Ltd., Prague, Czech Republic) and TEM (FEI—TALOS—F200X, Thermo Fisher Scientific (China) Co., Ltd., Shanghai, China) were used to observe the surface morphology. The phase structure was analyzed using XRD (D8 ADVANCE X, Bruker Corporation, Karlsruhe, Germany) with Cu-Kα radiation (λ = 0.154060 nm) at a scanning rate of 0.05 s−1 in the 2θ range from 10° to 80°. The chemical state and mass of the elements were characterized using XPS Escalab 250xi, Thermo Fisher Scientific, Massachusetts, WA, USA). Raman spectroscopy (Raman DXR, Thermo Fisher Scientific, Massachusetts, WA, USA), with a radiation wavelength of 785 nm, was used to analyze the molecular structure and properties.
The mass ratio of the biomass charcoal obtained after pyrolysis to the original biomass raw materials was defined as the yield of biomass charcoal. For BC and BC-FexOy, a TG analyzer (TGA101, Guangdong Kexiao Instrument Co., Ltd., Guangzhou, China) was used to determine the carbon and FexOy content. To study the thermal stability of the adsorbents, a thermogravimetric analysis of the two materials was performed in an air atmosphere. The heating rate was 10 °C/min and the temperature range was 50–850 °C. The specific surface area and pore volume were measured using a surface characteristic analyzer (ASAP 2460, McMurdik (Shanghai) Instrument Co., Ltd., Shanghai, China), and a magnetometer (MicroSenseEZ-VSM, MicroSense, LLC, Fremont, CA, USA) was used to test the magnetic field strength.
An atomic absorption spectrometer (TAS-990, Jinpeng Analytical Instrument Co., Ltd., Shanghai, China) was used to detect the concentration of heavy metal ions. FT-IR spectroscopy (Nicolet iS5, Thermo Fisher Scientific, Carlsbad, CA, USA) was employed to test the surface functional groups, and the absorbance of the sample was measured using a KBr tablet.

2.3. Inquiry into Adsorption Capacity

To imitate wastewater containing Pb ions, a lead solution was prepared using lead nitrate. First, an electronic balance was utilized to weigh 500 mg of lead nitrate, which was then diluted in a 1000 mL volumetric flask. Deionized water was added to create a reserve of lead solutions with other concentrations for the adsorption process. In the adsorption experiment, the standard solution of Pb ions should not have an extremely low concentration. If the concentration is too low, it is fully adsorbed by the adsorbent and the calibration curve will become non-straight. This situation may lead to the inaccuracy of the results in general. The adsorption kinetics of Pb2+ on BC and BC-FexOy were studied in 250 mL conical flasks. Some conical flasks contained 100 mL of 200 mg/L Pb2+ solution added to them. The pH was adjusted to 5.5 (See Supplementary Materials Figure S1 for the pH effect). A dosage of 0.1 g of BC and BC-FexOy was added and the conical flasks were sealed. These containers were placed on a 160 rpm constant-temperature oscillator at 25 °C, and 10 mL of suspension was taken at 5, 10, 20, 30, 40, 60, 80, 120, and 240 min and treated through magnetic force. Permanent magnets were placed around the glass tube, and the adsorbed wastewater flowed through the tube at a speed of approximately 10 mL/min. The samples were treated 3–5 times like this. The absorbance of the samples was measured using an atomic absorption spectrometer to determine the concentration of Pb ions after adsorption.
As shown in Equation (1), based on the difference between the initial solution and the equilibrium concentration of Pb2+, the adsorption capacity q (mg/g) of Pb2+ was calculated in line with a previous study [28].
q = C 0 C e m × V
In the formula, the initial concentration and equilibrium concentration (mg/L) of Pb2+ are represented by C0 and Ce, respectively; the volume of the solution is represented by V (L); and the mass of the added biomass charcoal is represented by m (g).
Equations (2) and (3) are used to process the adsorption data.
The pseudo-first-order equation is:
ln q e q t = l n q e k 1 · t
This shows the diffusion-controlled adsorption process.
The chemical adsorption process is illustrated by the pseudo-second-order equation:
t q t = 1 k 2 · q e 2 + 1 q e · t
The quantity of Pb2+ adsorbed is represented by qt (mg/g). The adsorption time is represented by t (min); qe (mg/g) represents the equilibrium adsorption capacity. The pseudo-first-order kinetic rate constant (min) is represented by k1. The pseudo-second-order kinetic rate is represented by k2, in units of g/(mg·min).

2.4. Exploring Isothermal Adsorption Kinetics

Pb2+ solutions (100 mL) of different concentrations (100, 150, 200, 250, and 300 mg/L) were poured into several conical flasks. The pH was adjusted to 5.5. Then, 0.1 g of BC and BC-FexOy were added, and the containers were sealed. They were then placed at 25 °C on a 160 rpm constant-temperature oscillator for 120 min. Subsequently, 10 mL of the suspension was removed and treated with magnetic force. An atomic absorption spectrometer was used to measure the absorbance of the samples and determine the concentration of lead ions after adsorption.
The adsorption process was analyzed by means of Langmuir and Freundlich isotherm models.
The equation for the Langmuir adsorption isotherm model is:
C e q e = C e q m + 1 q m K L
The equation for the Freundlich adsorption isotherm model is:
ln q e = ln K F + 1 n ln C e
The equilibrium adsorption capacity is represented by qe (mg/g); the maximum adsorption capacity is represented by qm (mg/g); the adsorption equilibrium concentration is represented by Ce (mg/L); and the Langmuir model has a constant KL (L/mg), which is associated with the affinity of the binding. KF, which pertains to the adsorption capacity, is the Freundlich model constant (mg1−n·Ln/g). The Freundlich model has a constant related to the adsorption strength, which is n.

3. Results and Discussion

3.1. Characterization of Biochar and Biochar–FexOy

SEM analysis of the biochar reveals (as depicted in Figure 2a) that its surface is relatively smooth and flat. EDS analysis indicates that the original biochar primarily consists of carbon, as illustrated in Figure 2b. Following magnetization treatment, numerous regular particles are observed adhering to the surface of the biochar, as shown in Figure 2c. Additionally, Figure 2d demonstrates that the modified biochar contains not only carbon but also iron and oxygen elements.
Selecting a micron-scale region, TEM reveals that the biochar appears sponge-like (Figure 3a). Upon further magnification, as depicted in Figure 3b, the biochar’s surface is devoid of attached particles and features numerous pore structures (blank area). Figure 3c displays the structure of magnetized biochar, markedly distinct from that shown in Figure 3a. Following liquid impregnation and subsequent calcination, nanoparticles were dispersed and embedded onto the biochar’s surface, forming spherical nanoparticles (blank area) with an average size of less than 20 nm (Figure 3d) (refer to Supplementary Materials Figure S2 for a particle size distribution diagram). The morphology of nanoparticles is uniform, with most of them having an elliptical granular structure. Moreover, the formation of nanoparticles did not compromise the pore structure of the biochar.
Further analysis of BC and BC-FexOy was conducted using high-resolution transmission electron microscopy (HRTEM) and selected area electron diffraction (SAED). In Figure 4a, no extensive crystal planes are evident on the biochar, primarily composed of amorphous carbon, with minor regions displaying graphene-like structures. The interplanar spacing of carbon atoms is 0.248 nm (with the lattice constant of graphene being 0.246 nm). Some areas, although graphenized, have obvious defects. Figure 4b displays the HRTEM of BC-FexOy, where the spacing between adjacent lattice fringes indicates the presence of Fe2O3 and Fe3O4 on the BC’s surface, with lattice fringe spacings of 0.221 nm (320), 0.251 nm (110), and 0.269 nm (104) [29]. Notably, a lattice spacing value of 0.485 nm is attributed to the crystal plane distance of the Fe3O4 (111) plane [30]. These results are consistent with the XRD display. Similarly, interspaces are observed in BC-FexOy. The presence of defects on the adsorbent surface enhances its adsorption capabilities [31]. The SAED patterns clearly demonstrate that the biomass charcoal is not in a monocrystalline form (Figure 4c). The SAED pattern of FexOy nanoparticles has two clear diffraction rings, corresponding to the (104) and (300) crystal planes. The irregular distribution of other diffraction points indicates that the nanoparticles have a polycrystalline structure (Figure 4d).

3.2. XRD, XPS, and Raman Analysis

XRD is displayed in Figure 5. The broad peak of BC between 15° and 30° indicates that biochar is mainly composed of amorphous carbon, consistent with the results in Figure 4a. After magnetization, the peak intensity of BC at the corresponding position significantly decreases, which is attributed to the surface of biomass charcoal being heavily loaded with iron oxide [32]. At 2θ, values of 24.1°, 33.2°, 35.6°, 40.8°, 49.5°, 54.2°, 62.5°, and 64.1° diffraction peaks corresponding to the (012), (104), (110), (113), (024), (116), (214), and (300) crystal planes of Fe2O3 (JCPDS 89-0597) were displayed. Fe3O4 (JCPDS 19-0629) shows characteristic peaks of iron oxide at 2θ of 35.6° and 62.5°. The peak appearing at 26.1° originates from the graphene (002) crystal plane. The crystal planes corresponding to typical characteristic peaks are consistent with those shown in Figure 4a,b.
The surface composition and chemical states of major elements of BC and BC-FexOy were detected by XPS spectroscopy, as shown in Figure 6a (the blue line represents BC-FexOy, and the red line represents BC). In BC, the atomic content of carbon and oxygen is 83.28% and 15.35%, while in BC-FexOy, the content of carbon, oxygen, and iron is 73.02%, 21.54%, and 4.08%, respectively. High-resolution C 1s spectra of BC-FexOy show that C-C, O-C-O, and O-C=O appear at 284.2 eV, 285.5 eV, and 288.2 eV, respectively (Figure 6b). The high-resolution O 1s spectrum of BC-FexOy shows peaks at 532.8 eV, 531.5 eV, and 529.2 eV, which are attributed to C=O, C-O, and Fe-O (Figure 6c). The corresponding high-resolution O 1s spectrum of BC only shows two peaks: C=O and C-O (inset of Figure 6c). The peak of the high-resolution spectrum of Fe 2p at 709.8 eV binding energy is attributed to the peak at Fe 2p3/2, 723.1 eV corresponding to the peak at Fe 2p1/2, as shown by the green line in Figure 6d. After peak fitting using Gaussian and Tougaard type baselines, the red line in Figure 6d represents the characteristic peak of divalent iron, and the blue line represents the characteristic peak of trivalent iron. The peak values of 711.9 eV and 724.7 eV are derived from the Fe 2p3/2 and Fe 2p1/2 peaks of iron oxide. Values of 718.1 eV and 728.7 eV are typical Fe 2p3/2 satellite peaks and Fe 2p1/2 peaks of divalent iron, while 719.1 eV and 732.5 eV peaks belong to Fe 2p3/2 and Fe 2p1/2 satellite peaks of trivalent iron [33]. The XPS results indicate that the nanoparticle elements embedded in BC are mainly C, O, and Fe, with iron exhibiting two valence states: divalent and trivalent.
As illustrated in Figure 7, the two prominent peaks at approximately 1342 cm−1 and 1602 cm−1 are identified as the D and G bands. The D band reflects the characteristics of disordered sp3 carbon with structural defects, suggesting the presence of defects near the material’s edges. The G band arises from hybrid sp2 carbon within the biochar structure, significantly influencing the extent of defects and disorder within the carbon layer [34]. The ID/IG ratio is commonly employed to characterize the degree of graphitization in carbon materials [35]. The measured ID/IG value for BC-FexOy is 0.94, higher than that of the BC (0.75). The elevated ID/IG value suggests that the introduction of FexOy nanoparticles has enhanced the graphitization of biochar and increased the defects and disorder, thereby activating more active sites within the BC and facilitating the adsorption of Pb2+ [36]. Furthermore, the peaks at 150 cm−1, 400 cm−1, and 634 cm−1 are assignable to Fe2O3 and Fe3O4 [37]. The Raman characterization findings align with the XRD results.

3.3. Physicochemical Properties of BC and BC-FexOy

Upon heating from 30 °C to 850 °C at a rate of 10 °C/min, the thermal stability and composition of BC and BC-FexOy were studied by thermogravimetric analysis. As shown in Figure 8a, at low temperatures (usually below 100 °C), the changes in the curve arise from the weight loss of a small amount of water in the sample and the volatilization of adsorbed gases, which are shown as endothermic processes on the DSC chart. As the temperature increased, the biochar began to react with oxygen. Above 370 °C, the reaction intensified, and this exothermic process can be observed in the DSC curve. Due to the decomposition of hydroxyl groups, there were changes in the two curves of biochar and FexOy at 410 °C and 442 °C. BC and BC-FexOy were heated above 553 °C and 579 °C, respectively, and the thermogravimetric and DSC curves did not show a decreasing trend, indicating that the biochar in BC and BC-FexOy was completely removed. The residual substances of BC and BC-FexOy are 12.5 wt% and 23.8 wt%, respectively, with FexOy accounting for 11.3 wt% of the total weight of BC-FexOy.
Figure 8b shows the magnetization curve of BC-FexOy. The saturation magnetization is 21.6 emu/g. The superparamagnetic behavior and high magnetic saturation strength exhibited by the magnetic biochar at room temperature indicate its susceptibility to external magnetic fields, which can be used to achieve solid–liquid separation and adsorbent recovery.
N2 adsorption/desorption analysis was carried out to show the specific surface area and pore structure. Typically, a greater specific surface area and pore volume usually lead to better adsorption efficiency. Figure 9a presents the N2 adsorption/desorption isotherms of BC and BC-FexOy. BC has a specific surface area of 219 m2/g; when doped with magnetic particles, BC-FexOy has a specific surface area of 326 m2/g. Nanoparticles of iron oxide possess a large specific surface area (around 66 m2/g), which might result in an increase in the specific surface area of biochar [38]. As can be seen from Figure 9b, the pore volumes of BC and BC-FexOy are 0.126 cm3/g and 0.232 cm3/g as per order. This may be because, for the same mass, the smaller the nanoparticles are, the larger their specific surface area will be, and the uniformly dispersed magnetic particles hardly block the pore structure on the BC surface [39]. In addition, the figure shows that when P/P0 > 0.5, the isotherm shows an IV curve with a hysteresis loop, which suggests the existence of mesoporous channels that are advantageous for the adsorption of heavy metal ions. Table 1 shows the physicochemical properties of the two materials.

3.4. Exploration of Adsorption Kinetics

The change in the adsorption capacity of the two materials for Pb2+ over time is presented in Figure 10, and it can be seen that BC-FexOy has a much higher adsorption capacity than BC. Pb2+ adsorption has two distinct phases: fast and less rapid. In the initial stage (0–50 min), the biochar’s adsorption ability for Pb2+ rises steeply, attaining approximately 95% of the saturated adsorption capacity. Subsequently, the adsorption enters a slower stage and attains equilibrium at approximately 120 min. The main reason for the occurrence of the two stages is that, in the initial stage of adsorption, biochar indicates the presence of a large number of holes and oxygen-containing functional groups that rapidly bind to lead ions, leading to a rapid decrease in the lead ion concentration. As the time increased, the number of holes and functional groups that could adsorb lead ions decreased, resulting in a decrease in the adsorption rate [40]. When the adsorption of Pb ions by biochar is saturated, the system reaches an equilibrium state.
The experimental data were linearly fitted with ln(qeqt) against t. Figure 11a shows the fitting results of the pseudo-first-order kinetic models for lead ion adsorption by BC and BC-FexOy. The curves obtained from the fitting show that the linear regression coefficients R2 for the adsorption of the two substances are 0.3939 and 0.4972. Figure 11b shows the t and t/q linear fit curves for the pseudo-second-order kinetic models regarding the adsorption of Pb2+ in wastewater by BC and BC-FexOy, and the linear correlation coefficients R2 of this model for the data from the adsorption experiments are 0.967 and 0.997, respectively, indicating that this model can represent the adsorption processes of BC and BC-FexOy more precisely during the whole adsorption process. It is indicated that chemisorption rather than physisorption controls the adsorption of Pb2+ by these two biochar types [41]. The chemisorption process mainly consists of the sharing or swapping of surface functional groups and electrons, which leads to the creation of valence forces for adsorbing metal ions [42]. The relevant parameters of pseudo-first-order and pseudo-second-order kinetics after linear fitting are presented in Table 2 and Table 3. The small standard estimation error (SE) of the utilized model indicates a high fitting accuracy.

3.5. Exploration of Adsorption Isotherm

The adsorption capabilities and affinities of adsorbents for metal ions can be evaluated using the equilibrium adsorption isotherm [39]. According to the experimental findings, at 25 °C, as the concentration of Pb2+ increased, the adsorption capacity of both normal and magnetized biomass charcoal increased, whereas the adsorption efficiency decreased (refer to Supplementary Materials Figure S3). The adsorption isotherms show that both materials have strong adsorption capabilities for Pb2+.
Linear fits of the Langmuir (Figure 12a) and Freundlich (Figure 12b) adsorption isotherms for Pb2+ adsorption in wastewater by biochar and magnetized biochar are shown in Figure 12. Using Equations (4) and (5), the slopes and intercepts were calculated to obtain the crucial parameters for the Langmuir and Freundlich adsorption isotherms, which are presented in detail in Table 4.
The R2 values of BC and BC-FexOy in the Langmuir models were 0.9935 and 0.9963, respectively, exceeding the R2 values of 0.6266 and 0.9063 in their Freundlich models, which implies that the Langmuir model can represent the adsorption process of heavy metal lead ions in wastewater by BC and BC-FexOy more precisely, with the characteristic of single-layer adsorption. The BC-FexOy has a maximum adsorption capacity of 252.7 mg/g, which is 1.69 times that of BC (149.8 mg/g). The term 1/n in Equation (5) denotes the adsorbent surface’s heterogeneity and adsorption intensity, 1/n > 1 indicates difficult adsorption, while 1/n < 0.5 signifies ease of adsorption [43]. The 1/n values for BC and BC-FexOy were 0.367 and 0.346, respectively, suggesting both materials effectively adsorbed Pb2+.
Additionally, as shown in Table 5, when comparing the adsorption efficacy of various modified biochar types for Pb2+, BC-FexOy exhibits an advantageous adsorption performance for Pb2+. Peng et al. prepared the Fe3O4-loaded biochar adsorbent using agricultural waste corn straw [22]. Perhaps due to its larger surface area and more functional groups, BC-FexOy has a better adsorption performance than Peng’s results. This underscores the significant influence of raw materials, synthesis processes, and methods on the performance of the adsorbents.

3.6. Characterization of FT-IR Spectroscopy

FT-IR spectroscopy can be used to characterize adsorbents and reveal functional groups in the adsorbent structure. This is important for understanding how adsorbents remove metals [49,50]. Figure 13 displays the FT-IR spectra of BC and BC-FexOy before and after Pb2+ adsorption. The two samples possessed comparable functional groups. The -OH groups showed stretching vibrations in the range of 3424–3461 cm−1. In comparison to BC, the functional group location of magnetized biochar has a red shift and its absorption peak is strengthened, which implies an increase in hydroxyl groups during the magnetization process [51]. At 2920 cm−1~2922 cm−1, both BC and BC-FexOy showed -CH2 vibrational bands. The stretching vibration peaks of C=O and C=C in the carboxyl, hydroxyl, and lactone groups are observed between 1570 and 1587 cm−1 [52]. The peak of the C-O stretching vibration can be found in the range from 1300 to 1000 cm−1. The C-H bending and C-C bond vibrations are represented in the 900–600 cm−1 range. A new characteristic peak at 569 cm−1 was observed for BC-FexOy, which was due to the stretching vibration of Fe-O groups. This indicates that, on the surface of the modified biochar, iron oxide forms chelation or bidentate bridge coordination bonds [53].
The shifting of the peaks within the ranges of 1570–1587 cm−1 and 1300–1000 cm−1 after Pb2+ adsorption indicates a complexation reaction between Pb2+ and oxygen-containing groups on the biochar. For BC and BC-FexOy, upon Pb2+ adsorption, the O-H stretching vibration peak intensity at 3424~3461 cm−1 increased markedly, which shows that ion-exchange and complex adsorption reactions took place during the adsorption process [54,55]. Furthermore, it was discovered that the alterations in the vibration peaks on the surface of the magnetic biochar suggest that FexOy might play a role in Pb2+ adsorption.

3.7. Absorption Mechanism

Figure 14 shows the adsorption mechanism of composite materials for Pb2+ according to the above research. The primary adsorption procedure could be like this:
(1)
During physical adsorption, lead ions diffused on the surface of the composite material. As they move, these lead ions are gradually adsorbed into the micropores of the composite material and effectively fixed. This process utilizes the adsorption capacity of the composite material micropores to achieve the effective removal of lead ions.
(2)
The oxygen-containing functional groups that are abundant in biochar play a crucial role. These functional groups have unique chemical properties and can undergo a series of physical and chemical reactions, such as complexation and ion exchange, with heavy metal ions. Specifically, they form stable complexes with heavy metal ions through chelation, effectively binding these ions. Through ion exchange, the ions in the functional groups are replaced with heavy metal ions, further promoting the removal of heavy metal ions. Through the abovementioned mechanism, Pb2+ was effectively removed from wastewater.
(3)
The embedding of FexOy nanoparticles into biochar not only increased the specific surface area and porosity of the composite, improving the adsorption efficiency, but also removed Pb2+ through precipitation/co-precipitation by interacting with the hydroxyl functional groups (Fe-OH) on the surface of Fe3O4 [56].

3.8. Reusability Needs to Be Evaluated

It is crucial to evaluate the reusability of ideal adsorption materials. BC-FexOy was regenerated with eluent 0.5 mol·L−1 hydrochloric acid and ethanol in a volume ratio of 1:1. After the ionization of hydrochloric acid, a large amount of H+ is generated, which can undergo ion exchange with Pb2+, causing Pb2+ to desorb from BC-FexOy. The detached material can be recycled [57]. Ethanol can remove organic impurities adhered to the adsorbent surface, enhancing elution effectiveness. Under optimal experimentation (0.1 g BC-FexOy, 100 mL of 200 mg/L Pb2+ simulant wastewater, 25 °C, PH 5.5), adsorption and desorption were used to evaluate the reusability of BC-FexOy. It was found that after 5 cycles, only 12.16% of Pb2+ could be adsorbed (Figure 15). This might be due to the fact that a small number of adsorption sites were destroyed or blocked [53].

4. Conclusions

The research results show that biochar prepared from cotton as a precursor exhibits significant preparation advantages and adsorption characteristics. During the pyrolysis preparation process, cotton biochar showed a high yield of 41%, which is significantly advantageous over traditional wood materials. Through various characterization methods, it was confirmed that FexOy nanoparticles (<20 nm) were successfully loaded onto the obtained magnetic composite biochar (BC-FexOy), with a magnetization strength of 21.6 emu/g and an excellent magnetic recovery performance. The specific surface area (326 m2/g) and pore volume (0.232 cm3/g) of the material were increased by approximately 1.49 times and 1.84 times, respectively, compared to unmodified BC, indicating that the introduction of nanoparticles effectively optimized the pore structure.
The lead adsorption experiment showed that the adsorption process of BC-FexOy on Pb2+ followed the Langmuir isotherm model (R2 > 0.99), with a maximum adsorption capacity of 252.7 mg/g, which was 68.8% higher than the original BC (149.8 mg/g). FT-IR characterization reveals that the abundant oxygen-containing functional groups, such as hydroxyl and carboxyl groups on the surface of the material, can serve as active sites, achieving efficient adsorption through coordination and ion exchange with Pb2+. The pore structure generated by magnetic composite provides more diffusion channels for heavy metal ions.
Dynamic simulation experiments have confirmed that BC FexOy maintains high adsorption efficiency after five consecutive adsorption desorption cycles. Combining the renewable characteristics of the raw materials indicates that the material has the potential for large-scale application in the field of wastewater treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym17040516/s1, Figure S1: Pb2+ adsorption onto two materials at different pH values; Figure S2: Diagram of particle size distribution; Figure S3: Effect of different initial concentrations.

Author Contributions

Y.Y.: conceptualization, data curation, formal analysis, investigation, methodology, project administration, validation, visualization, writing—original draft, and writing—review and editing; T.Y.: investigation and methodology; C.Q.: resources and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Wuhu High Efficiency Environmental Protection Equipment Manufacturing and Operation Technology Research and Development Center (WHSYFZX202307) and the Major Project of Scientific Research in Higher Education Institutions of Anhui Provincial Department of Education (2024AH040254).

Institutional Review Board Statement

This work did not require ethical approval from a human subject or animal welfare committee.

Data Availability Statement

All data are provided in the electronic Supplementary Materials accompanying this article.

Conflicts of Interest

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

References

  1. Junuzovic, H.; Selimovic, A.; Begic, S.; Dozic, A.; Cvrk, R.; Ahmetovic, M.; Alibasic, H. Chemical precipitation of cations from aqueous solutions using waste sludge from the solway process as a potential agent. Int. J. Res. Appl. Sci. Biotechnol. 2020, 7, 329–337. [Google Scholar] [CrossRef]
  2. Efome, J.E.; Rana, D.; Matsuura, T.; Lan, C.Q. Effects of operating parameters and coexisting ions on the efficiency of heavy metal ions removal by nano-fibrous metal-organic framework membrane filtration process. Sci. Total Environ. 2019, 674, 355–362. [Google Scholar] [CrossRef]
  3. Biesheuvel, P.M.; Porada, S.; Elimelech, M.; Dykstra, J.E. Tutorial review of reverse osmosis and electrodialysis. J. Membr. Sci. 2022, 647, 120221. [Google Scholar]
  4. Kumar, A.; Nidheesh, P.V.; Kumar, M.S. Composite wastewater treatment by aerated electrocoagulation and modified peroxi-coagulation processes. Chemosphere 2018, 205, 587–593. [Google Scholar] [PubMed]
  5. Carolin, C.F.; Kumar, P.S.; Saravanan, A.; Joshiba, G.J.; Naushad, M. Efficient techniques for the removal of toxic heavy metals from aquatic environment: A review. J. Environ. Chem. Eng. 2017, 5, 2782–2799. [Google Scholar]
  6. Ezugwu, C.I.; Ujam, O.T.; Ukoha, P.O.; Ukwueze, N.N. Complex formation and extraction studies of N, N-Bis(salicylidene)-3,5-diaminobenzoic Acid on Hg(II) and Ag(I). Chem. Sci. Trans. 2013, 2, 1118–1125. [Google Scholar]
  7. Alghamdi, A.A.; Al-Odayni, A.B.; Saeed, W.S.; AlKahtani, A.; Alharthi, F.A.; Aouak, T. Efficient adsorption of lead (II) from aqueous phase solutions using polypyrrole-based activated carbon. Materials 2019, 12, 2020. [Google Scholar] [CrossRef] [PubMed]
  8. Ali, I.; Neskoromnaya, E.A.; Melezhik, A.V.; Babkin, A.V.; Kulnitskiy, B.A.; Burakov, A.E.; Burakova, I.V.; Tkachev, A.G.; Almalki, A.S.A.; Alsubaie, A. Magnetically active nanocomposite aerogels: Preparation, characterization and application for water treatment. J. Porous Mat. 2022, 29, 5450–5457. [Google Scholar]
  9. Bolisetty, S.; Peydayesh, M.; Mezzenga, R. Sustainable technologies for water purification from heavy metals: Review and analysis. Chem. Soc. Rev. 2019, 48, 463–487. [Google Scholar]
  10. Zhao, P.; Jian, M.; Zhang, Q.; Xu, R.; Liu, R.; Zhang, X.; Liu, H. A new paradigm of ultrathin 2D nanomaterial adsorbents in aqueous media: Graphene and GO, MoS2, MXenes, and 2D MOFs. J. Mater. Chem. A 2019, 7, 16598–16621. [Google Scholar]
  11. Huang, L.; Liu, R.; Yang, J.; Shuai, Q.; Yuliarto, B.; Kaneti, Y.V.; Yamauchi, Y. Nanoarchitectured porous organic polymers and their environmental applications for removal of toxic metal ions. Chem. Eng. J. 2021, 408, 127991. [Google Scholar] [CrossRef]
  12. He, W.; Ai, K.; Ren, X.; Wang, S.; Lu, L. Inorganic layered ion-exchangers for decontamination of toxic metal ions in aquatic systems. J. Mater. Chem. A. 2017, 5, 19593–19606. [Google Scholar]
  13. Dong, S.; Xu, W.; Wu, F.; Yan, C.; Li, D.; Jia, H. Fe-modified biochar improving transformation of arsenic form in soil and inhibiting its absorption of plant. Trans. Chin. Soc. Agric. Eng. 2016, 32, 204–212. [Google Scholar]
  14. Kumar, S.; Masto, R.E.; Ram, L.C.; Sarkar, P.; George, J.; Selvi, V.A. Biochar preparation from Parthenium hysterophorus and its potential use in soil application. Ecol. Eng. 2013, 55, 67–72. [Google Scholar]
  15. Dinesh, M.; Kumar, H.; Sarswat, A. Cadmium and lead remediation using magnetic oak wood and oak bark fast pyrolysis biochars. Chem. Eng. J. 2014, 236, 513–528. [Google Scholar]
  16. Ahmad, S.; Liu, X.; Tang, J.; Zhang, S. Biochar-supported nanosized zero-valent iron (nZVI/BC) composites for removal of nitro and chlorinated contaminants. Chem. Eng. J. 2022, 431, 133187. [Google Scholar] [CrossRef]
  17. Wei, J.; Liu, Y.; Li, J.; Zhu, Y.; Yu, H.; Peng, Y. Adsorption and co-adsorption of tetracycline and doxycycline by one-step synthesized iron loaded sludge biochar. Chemosphere 2019, 236, 124254. [Google Scholar]
  18. Li, X.; Qin, Y.; Jia, Y.; Li, Y.; Zhao, Y.; Pan, Y.; Sun, J. Preparation and application of Fe/biochar (Fe-BC) catalysts in wastewater treatment: A review. Chemosphere 2021, 274, 129766. [Google Scholar]
  19. Amusat, S.O.; Kebede, T.G.; Dube, S.; Nindi, M.M. Ball-milling synthesis of biochar and biochar-based nanocomposites and prospects for removal of emerging contaminants: A review. J. Water Process Eng. 2021, 41, 101993. [Google Scholar]
  20. Zhang, J.-Y.; Zhou, H.; Gu, J.-F.; Huang, F.; Yang, W.-J.; Wang, S.-L.; Yuan, T.-Y.; Liao, B.-H. Effects of nano-Fe3O4-modified biochar on iron plaque formation and Cdaccumulation in rice (Oryza sativa L.). Environ. Pollut. 2020, 260, 113970. [Google Scholar]
  21. Liu, K.; Li, F.; Cui, J.; Yang, S.; Fang, L. Simultaneous removal of Cd(II) and As(III) by graphene-like biochar-supported zero-valent iron from irrigation waters under aerobic conditions: Synergistic effects and mechanisms. J. Hazard. Mater. 2020, 395, 122623. [Google Scholar]
  22. Peng, J.; Zhang, Z.Y.; Wang, Z.W.; Zhou, F.; Yu, J.X.; Chi, R.; Xiao, C.Q. Adsorption of Pb2+ in solution by phosphate-solubilizing microbially modified biochar loaded with Fe3O4. J. Taiwan Inst. Chem. E 2024, 156, 105363. [Google Scholar]
  23. Wani, A.L.; Ara, A.; Usmani, J.A. Lead toxicity: A review. Interdiscipl. Toxicol. 2015, 8, 55–64. [Google Scholar]
  24. Zhang, M.; Jia, F.F.; Dai, M.; Son, S.X. Combined electrosorption and chemisorption of low concentration Pb(II) from aqueous solutions with molybdenum disulfifide as electrode. Appl. Surf. Sci. 2018, 455, 258–266. [Google Scholar]
  25. Chen, J.; Liao, C.; Guo, X.X.; Hou, S.C.; He, W.D. PAAO cryogels from amidoximated P(acrylic acid-co-acrylonitrile) for the adsorption of lead ion. Eur. Polym. J. 2022, 171, 111192. [Google Scholar]
  26. Nie, G.Z.; Qiu, S.J.; Wang, X.; Du, Y.; Zhang, Q.R.; Zhang, Y.; Zhang, H.L. A millimeter-sized negatively charged polymer embedded with molybdenum disulfide nanosheets for efficient removal of Pb(II) from aqueous solution. Chin. Chem. Lett. 2021, 32, 2342–2346. [Google Scholar]
  27. Zhu, Q.L.; Cao, M.; Zhang, X.B.; Tao, K.; Ke, Y.C.; Meng, L. Physicochemical and Infrared Spectroscopic Properties of Gramineae Plants Biochar at Different Pyrolysis Temperatures. Biomass Chem. Eng. 2021, 55, 21–28. [Google Scholar]
  28. Yao, Y.-Z.; Shi, Y.-J.; Hu, K.-H. Preparation of Molybdenum Disulfide with Different Nanostructures and Its Adsorption Performance for Copper (II) Ion in Water. Nanomaterials 2023, 13, 1194. [Google Scholar]
  29. Shi, C.J.; Wang, Y.; Zhang, K.; Lichtfouse, E.; Li, C.; Zhang, Y.H. Fe-biochar as a safe and efficient catalyst to activate peracetic acid for the removal of the acid orange dye from water. Chemosphere 2022, 307, 135686. [Google Scholar] [PubMed]
  30. Ye, Z.; Qie, Y.; Fan, Z.; Liu, Y.; Shi, Z.; Yang, H. Soft magnetic Fe5C2-Fe3C@C as an electrocatalyst for the hydrogen evolution reaction. Dalton T 2019, 48, 4636–4642. [Google Scholar]
  31. Chen, B.; Liu, E.; He, F.; Shi, C.; He, C.; Li, J.; Zhao, N. 2D sandwich-like carboncoated ultrathin TiO2@defect-rich MoS2 hybrid nanosheets: Synergistic-effectpromoted electrochemical performance for lithium ion batteries. Nano Energy 2016, 26, 541–549. [Google Scholar] [CrossRef]
  32. Gao, R.; Fu, Q.; Hu, H.; Wang, Q.; Liu, Y.; Zhu, J. Highly-effective removal of Pb by co-pyrolysis biochar derived from rape straw and orthophosphate. J. Hazard. Mater. 2019, 371, 191–197. [Google Scholar] [CrossRef]
  33. Wang, C.; Sun, R.; Huang, R. Highly dispersed iron-doped biochar derived from sawdust for Fenton-like degradation of toxic dyes. J. Clean. Prod. 2021, 297, 126681. [Google Scholar] [CrossRef]
  34. Ren, F.; Zhu, W.; Zhao, J.; Liu, H.; Zhang, X.; Zhang, H.; Zhu, H.; Peng, Y.; Wang, B. Nitrogendoped graphene oxide aerogel anchored with spinel CoFe2O4 nanoparticles for rapid degradation of tetracycline. Sep. Purif. Technol. 2020, 241, 116690. [Google Scholar] [CrossRef]
  35. Cheng, P.; Li, T.; Yu, H.; Zhi, L.; Liu, Z.H.; Lei, Z. Biomass-derived carbon fiber aerogel as binder-free electrode for high-rate supercapacitor. J. Phys. Chem. C 2016, 120, 2079–2086. [Google Scholar] [CrossRef]
  36. Qin, Y.; Li, X.; Wang, L.; Luo, J.; Li, Y.; Yao, C.; Xiao, Z.; Zhai, S.; An, Q. Valuable cobalt/biochar with enriched surface oxygen-containing groups prepared from bio-waste shrimp shell for efficient peroxymonosulfate activation. Sep. Purif. Technol. 2022, 281, 119901. [Google Scholar] [CrossRef]
  37. Testa-Anta, M.; Ramos-Docampo, M.A.; Comesaña-Hermo, M.; Rivas-Murias, B.; Salgueiriño, V. Raman spectroscopy to unravel the magnetic properties ofiron oxide nanocrystals for bio-related applications. Nanoscale Adv. 2019, 1, 2086–2103. [Google Scholar] [CrossRef]
  38. Lai, L.; Xie, Q.; Fang, W.K.; Xing, M.C.; Wu, D.Y. Removal and recycle of phosphor from water using magnetic core /shell structured Fe3O4@SiO2 nanoparticles functionalized with hydrous aluminum oxide. Environ. Sci. 2016, 37, 1444–1450. [Google Scholar]
  39. Wang, Y.; Yu, S.; Yuan, H.W.; Zhang, L. Constructing N,S co-doped network biochar confined CoFe2O4 magnetic nanoparticles adsorbent: Insights into the synergistic and competitive adsorption of Pb2+ and ciprofloxacin. Environ. Pollut. 2024, 343, 123178. [Google Scholar] [CrossRef]
  40. Hu, X.Y.; Chen, Y.J.; Zhang, S.H.; Wang, X.Q.; Li, C.C.; Guo, X. Cd removal from aqueous solution using magnetic biochar derived from maize straw and its recycle. Trans. Chin. Soc. Agric. Eng. 2018, 34, 208–218. [Google Scholar]
  41. Guo, W.; Ai, Y.; Men, B.; Wang, S. Adsorption of phenanthrene and pyrene by biochar produced from the excess sludge: Experimental studies and theoretical analysis. Int. J. Environ. Sci. Technol. 2017, 14, 1889–1896. [Google Scholar]
  42. Wang, T.; Sun, H.; Ren, X.; Li, B.; Mao, H. Adsorption of heavy metals from aqueous solution by UV-mutant Bacillus subtilis loaded on biochars derived from different stock materials. Ecotoxicol. Environ. Saf. 2018, 148, 285–292. [Google Scholar]
  43. Pei, H.N.; Wang, J.; Yang, Q.F.; Yang, W.X.; Hu, N.; Suo, Y.R.; Zhang, D.H.; Li, Z.H.; Wang, J.L. Interfacial growth of nitrogen-doped carbon with multi-functional groups on the MoS2 skeleton for efficient Pb(II) removal. Sci. Total Environ. 2018, 632, 912–920. [Google Scholar] [CrossRef]
  44. Li, S.; Yang, F.; Li, J.; Cheng, K. Porous biochar-nanoscale zero-valent iron composites: Synthesis, characterization and application for lead ion removal. Sci. Total. Environ. 2020, 746, 141037. [Google Scholar]
  45. Chen, H.; Zhang, J.; Tang, L.; Su, M.; Tian, D.; Zhang, L.; Li, Z.; Hu, S. Enhanced Pb immobilization via the combination of biochar and phosphate solubilizing bacteria. Environ. Int. 2019, 127, 395–401. [Google Scholar] [CrossRef]
  46. Chen, H.; Li, W.; Wang, J.; Xu, H.; Liu, Y.; Zhang, Z.; Li, Y.; Zhang, Y. Adsorption of cadmium and lead ions by phosphoric acid-modified biochar generated from chicken feather: Selective adsorption and influence of dissolved organic matter. Bioresour. Technol. 2019, 292, 121948. [Google Scholar]
  47. Dewage, N.B.; Fowler, R.E.; Pittman, C.U.; Mohan, D.; Mlsna, T. Lead (Pb2+) sorptive removal using chitosan-modified biochar: Batch and fixed-bed studies. RSC Adv. 2018, 8, 25368–25377. [Google Scholar]
  48. Li, Y.C.; Yin, H.; Cai, Y.H.; Luo, H.Y.; Yan, C.Y.; Dang, Z. Regulating the exposed crystal facets of α-Fe2O3 to promote Fe2O3-modified biochar performance in heavy metals adsorption. Chemosphere 2023, 311, 136976. [Google Scholar]
  49. Hamadeen, H.M.; Elkhatib, E.A.; Moharem, M.L. Optimization and mechanisms of rapid adsorptive removal of chromium (VI) from wastewater using industrial waste derived nanoparticles. Sci. Rep. 2022, 12, 14174. [Google Scholar] [CrossRef]
  50. El-Kammah, M.; Elkhatib, E.; Aboukila, E. Ecofriendly nanoparticles derived from water industry byproducts for effective removal of Cu (II) from wastewater: Adsorption isotherms and kinetics. Inorg. Chem. Commun. 2022, 146, 110062. [Google Scholar]
  51. Liu, X.; Lai, D.; Wang, Y. Performance of Pb(II) removal by an activated carbon supported nanoscale zero-valent iron composite at ultralow iron content. J. Hazard. Mater. 2019, 361, 37–48. [Google Scholar]
  52. Jin, M.L.J.H.; Deng, C.; Wang, S. Preparation and characterization of nanoparticles containing Fe3O4 cores in biochar. J. Agro-Environ. Sci. 2018, 37, 592–597. [Google Scholar]
  53. Dong, X.W.; He, L.Z.; Hu, H.; Liu, N.; Gao, S.; Piao, Y.X. Removal of 17β-estradiol by using highly adsorptive magnetic biochar nanoparticles from aqueous solution. Chem. Eng. J. 2018, 352, 371–379. [Google Scholar]
  54. Feng, Z.; Chen, N.; Feng, C.; Gao, Y. Mechanisms of Cr (VI) removal by FeCl3-modified lotus stem-based biochar (FeCl3@LSBC) using mass-balance and functional group expressions. Colloid Surface A 2018, 551, 17–24. [Google Scholar]
  55. Zhao, Y.; Zhang, R.; Liu, H.; Li, M.; Chen, T.; Chen, D.; Zou, X.; Frost, R.L. Green preparation of magnetic biochar for the effective accumulation of Pb (II): Performance and mechanism. Chem. Eng. J. 2019, 375, 122011. [Google Scholar]
  56. Wei, S.X.; Li, C.H.; Ma, W.Q.; Zhang, L.Y.; Li, J.M.; Mao, A.Q.; He, C.; Li, M.H.; Zhu, W.C. Study on the removal mechanism of Zn(II) and Pb(II) by magnetic flake nZVI-Fe3O4. Chin. J. Process Eng. 2024, 24, 346–359. (In Chinese) [Google Scholar]
  57. Kataria, N.; Garg, V.K. Green synthesis of Fe3O4 nanoparticles loaded sawdust carbon for cadmium (II) removal from water: Regeneration and mechanism. Chemosphere 2018, 208, 818–828. [Google Scholar]
Figure 1. Illustration of fabrication strategy.
Figure 1. Illustration of fabrication strategy.
Symmetry 17 00516 g001
Figure 2. The SEM and EDS images of biochar (a,b); biochar–FexOy (c,d).
Figure 2. The SEM and EDS images of biochar (a,b); biochar–FexOy (c,d).
Symmetry 17 00516 g002
Figure 3. TEM images of biochar (a,b); biochar–FexOy (c,d).
Figure 3. TEM images of biochar (a,b); biochar–FexOy (c,d).
Symmetry 17 00516 g003
Figure 4. HRTEM and SAED images of biochar (a,c); biochar–FexOy (b,d).
Figure 4. HRTEM and SAED images of biochar (a,c); biochar–FexOy (b,d).
Symmetry 17 00516 g004
Figure 5. XRD images of biochar and biochar–FexOy.
Figure 5. XRD images of biochar and biochar–FexOy.
Symmetry 17 00516 g005
Figure 6. (a) Full range of XPS spectra of BC and BC-FexOy, (b) high-resolution C 1s spectra of BC-FexOy, (c) high-resolution O 1s spectra of BC-FexOy (Inset: high-resolution O 1s spectrum of BC), and (d) high-resolution Fe 2p spectra of BC-FexOy.
Figure 6. (a) Full range of XPS spectra of BC and BC-FexOy, (b) high-resolution C 1s spectra of BC-FexOy, (c) high-resolution O 1s spectra of BC-FexOy (Inset: high-resolution O 1s spectrum of BC), and (d) high-resolution Fe 2p spectra of BC-FexOy.
Symmetry 17 00516 g006
Figure 7. Raman spectra of BC and BC-FexOy.
Figure 7. Raman spectra of BC and BC-FexOy.
Symmetry 17 00516 g007
Figure 8. (a) Thermogravimetric curves of BC, (b) BC-FexOy and magnetization curve of BC-FexOy.
Figure 8. (a) Thermogravimetric curves of BC, (b) BC-FexOy and magnetization curve of BC-FexOy.
Symmetry 17 00516 g008
Figure 9. (a) Nitrogen adsorption/desorption isotherms and (b) Pore volumes of BC and BC-FexOy.
Figure 9. (a) Nitrogen adsorption/desorption isotherms and (b) Pore volumes of BC and BC-FexOy.
Symmetry 17 00516 g009
Figure 10. Adsorption kinetics of Pb2+ removal by BC and BC-FexOy (pH 5.5).
Figure 10. Adsorption kinetics of Pb2+ removal by BC and BC-FexOy (pH 5.5).
Symmetry 17 00516 g010
Figure 11. Pseudo-first-order (a) and pseudo-second-order (b) kinetic models for adsorption of Pb2+.
Figure 11. Pseudo-first-order (a) and pseudo-second-order (b) kinetic models for adsorption of Pb2+.
Symmetry 17 00516 g011
Figure 12. Langmuir (a) and Freundlich (b) adsorption isotherm models for Pb2+ adsorption.
Figure 12. Langmuir (a) and Freundlich (b) adsorption isotherm models for Pb2+ adsorption.
Symmetry 17 00516 g012
Figure 13. (a) FT-IR FTIR spectra of BC and BC-FexOy; (b) FT-IR FTIR spectra of BC and BC-FexOy after the adsorption of lead ions.
Figure 13. (a) FT-IR FTIR spectra of BC and BC-FexOy; (b) FT-IR FTIR spectra of BC and BC-FexOy after the adsorption of lead ions.
Symmetry 17 00516 g013
Figure 14. Proposed mechanism of Pb2+ adsorption.
Figure 14. Proposed mechanism of Pb2+ adsorption.
Symmetry 17 00516 g014
Figure 15. The recycling performances of adsorption (replicates n = 3).
Figure 15. The recycling performances of adsorption (replicates n = 3).
Symmetry 17 00516 g015
Table 1. Physical and chemical properties of BC and BC-FexOy.
Table 1. Physical and chemical properties of BC and BC-FexOy.
MaterialsMagnetization Strength
(emu/g)
Specific Surface
(m2/g)
Pore Volume
(m3/g)
BC-2190.126
BC-FexOy21.63260.232
Table 2. The relevant parameters of pseudo-first-order kinetics after linear fitting.
Table 2. The relevant parameters of pseudo-first-order kinetics after linear fitting.
Pseudo-First-OrderInterceptSlopeStatistics
ValueStandard ErrorValueStandard ErrorAdj. R2
BC-FexOy4.392510.27467−0.009190.003080.4972
BC4.889140.13414−0.003840.001540.39394
Table 3. The relevant parameters of pseudo-second-order kinetics after linear fitting.
Table 3. The relevant parameters of pseudo-second-order kinetics after linear fitting.
Pseudo-Second-OrderInterceptSlopeStatistics
ValueStandard ErrorValueStandard ErrorAdj. R2
BC-FexOy0.011160.002420.001482.8696 × 10−50.99699
BC0.036950.011030.002161.408 × 10−40.96704
Table 4. Related parameters of the Langmuir adsorption isotherm model and Freundlich adsorption isotherm model.
Table 4. Related parameters of the Langmuir adsorption isotherm model and Freundlich adsorption isotherm model.
MaterialsLangmuir ModelFreundlich Model
KLqm (mg/g)R2nKFR2
Biochar0.57149.80.99352.7252.290.6266
Biochar–FexOy0.69252.70.99632.89112.30.9063
Table 5. Comparison of Pb2+ adsorption by different modified biochar types.
Table 5. Comparison of Pb2+ adsorption by different modified biochar types.
Biochar MaterialsAdsorption Capacity (mg/g)Refs.
Microbial modification of iron-carrying biochar113.70[22]
Alginate-modified rice husk biochar112.30[29]
Acid ammonium persulfate oxidize biochar135.40[44]
PSB-modified sludge biochar12.17[45]
H3PO4-modified chicken feather biochar24.41[46]
Chitosan-modified pine biochar134.00[47]
α-Fe2O3-modified biochar (Fe2O3/BC)390.60 [48]
Biochar embedded FexOy nanoparticle composite252.7This work
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

Yao, Y.; Yao, T.; Qian, C. Synthesis and Application for Pb2+ Removal of a Novel Magnetic Biochar Embedded with FexOy Nanoparticles. Symmetry 2025, 17, 516. https://doi.org/10.3390/sym17040516

AMA Style

Yao Y, Yao T, Qian C. Synthesis and Application for Pb2+ Removal of a Novel Magnetic Biochar Embedded with FexOy Nanoparticles. Symmetry. 2025; 17(4):516. https://doi.org/10.3390/sym17040516

Chicago/Turabian Style

Yao, Youzhi, Tiancheng Yao, and Cheng Qian. 2025. "Synthesis and Application for Pb2+ Removal of a Novel Magnetic Biochar Embedded with FexOy Nanoparticles" Symmetry 17, no. 4: 516. https://doi.org/10.3390/sym17040516

APA Style

Yao, Y., Yao, T., & Qian, C. (2025). Synthesis and Application for Pb2+ Removal of a Novel Magnetic Biochar Embedded with FexOy Nanoparticles. Symmetry, 17(4), 516. https://doi.org/10.3390/sym17040516

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