Next Article in Journal
Association between Sleep Onset Problem and Subjective Cognitive Complaints among Japanese Older Adults during the Coronavirus Disease 2019 Pandemic
Next Article in Special Issue
Fast Capture and Efficient Removal of Bloom Algae Based on Improved Dielectrophoresis Process
Previous Article in Journal
Profiling Physical Fitness of Physical Education Majors Using Unsupervised Machine Learning
Previous Article in Special Issue
Carbon Emission Accounting and the Carbon Neutralization Model for a Typical Wastewater Treatment Plant in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent-Free Synthesis of Magnetic Sewage Sludge-Derived Biochar for Heavy Metal Removal from Wastewater

1
College of Life and Environmental Sciences, Minzu University of China, Beijing 100081, China
2
Technical Centre for Soil, Agriculture and Rural Ecology and Environment, Ministry of Ecology and Environment, Beijing 100012, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Environ. Res. Public Health 2023, 20(1), 155; https://doi.org/10.3390/ijerph20010155
Submission received: 30 November 2022 / Revised: 15 December 2022 / Accepted: 20 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Wastewater Pollution and Control)

Abstract

:
The commonly used two-step and one-pot synthesis methods for producing biochar require the use of iron salt solutions, resulting in the undesirable consequences of energy consumption for dewatering and potential pollution risks. To address this drawback, a magnetic sewage sludge-derived biochar (MSBC-2) was synthesized by a solvent-free method in this study. The pseudo-second-order kinetic model and Langmuir model provided the best fit to the experimental data, implying a monolayered chemisorption process of Pb2+, Cd2+and Cu2+ onto MSBC-2. As the reaction temperature increased from 25 °C to 45 °C, the maximum adsorption capacities increased from 113.64 mg·g−1 to 151.52 mg·g−1 for Pb2+, from 101.01 mg·g−1 to 109.89 mg·g−1 for Cd2+ and from 57.80 mg·g−1 to 74.07 mg·g−1 for Cu2+, respectively. Thermodynamic parameters (ΔG0 < 0, ΔS0 > 0, ΔH0 > 0) revealed that the adsorption processes of all three metals by MSBC-2 were favourable, spontaneous and endothermic. Surface complexation, cation-π interaction, ion exchange and electrostatic attraction mechanisms were involved in the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2. Overall, this study will provide a new perspective for the synthesis of magnetic biochar and MSBC-2 shows great potential as an adsorbent for heavy metal removal.

1. Introduction

Industrial wastewater often contains various heavy metals, including Pb2+, Cd2+ and Cu2+, etc. Heavy metals are not biodegradable and can be enriched in humans through the food chain, leading to a serious impact on the aquatic environment and public health [1,2]. Current heavy metal removal approaches include ion exchange, co-precipitation, membrane filtration and adsorption, etc. [3,4,5,6,7]. Among them, adsorption is regarded as one of the most simple, cost-effective and efficient techniques, and has attracted increasing attention. As adsorbents are essential for adsorption applications, various adsorbents have been developed, such as activated carbon, biochar, metal oxides in nanoscale, natural minerals, and polymers [8,9]. Generally, biochar is a carbon-based adsorbent which is generated by the pyrolysis of biomass under anaerobic conditions or in the presence of limited oxygen [10]. Due to its large specific surface area, abundant oxygen-containing functional groups and low cost, biochar has been considered as a promising adsorbent [11,12].
Sewage sludge is a by-product of wastewater treatment plants, which contains various organic pollutants and pathogens [13]. Traditional treatment methods for sewage sludge (e.g., incineration and sanitary landfill) often have high energy consumption, low efficiency and a tendency to cause secondary pollution, thus it is imperative to develop environmentally friendly sludge treatment approaches [14,15]. Lately, sewage sludge has been converted into biochar by pyrolysis and applied to remove heavy metals from wastewater as the adsorbent or to mitigate greenhouse gas emissions as the soil amendment, which has received increasing interest [16]. Previous studies reported that sewage sludge-derived biochar (SBC) was employed as an efficient adsorbent for the removal of Pb, Zn, Cd and Cu from aqueous solution [17,18]. Compared to other biochar, SBC is demonstrated to have more oxygen-containing functional groups (e.g., carboxyl, hydroxyl, and carbonyl groups) and larger specific surface area, because of the large number of microorganisms and organic matters in the sludge [19]. However, it is difficult to separate SBC adsorbent from the aqueous solution after adsorption due to the tiny particle size, restricting its practical application in wastewater treatment.
Despite the difficulty in separating from water, raw biochar has other disadvantages such as small particle size and low density. To eliminate the above shortcomings and enhance the metal removal efficiency, various modification methods (e.g., surface oxidation, impregnation of metal oxides and functionalization) have been used to modify biochar [20,21,22]. Among them, the magnetic modification of biochar offers the potential for rapid separation and recovery in the presence of an external magnetic field. For instance, previous studies have reported that FeCl3-modified biochar had excellent magnetic sensitivity and could be separated from the aqueous solution rapidly [23]. In addition, magnetic modification can enhance the metal adsorption performance [19,24,25]. Currently, two-step and one-pot synthesis methods are the most used to produce magnetic biochar [26,27]. For example, the former study successfully used a two-step method to prepare a magnetic tea-based biochar with an iron-containing solution [27]. In terms of the one-pot method, the substrate is usually placed in a magnetic solution containing metal ions for impregnation loading and then pyrolysis is conducted to obtain biochar [28]. However, both require the use of iron salt solutions as post- and pre-treatment reagents, respectively. This caused the inevitable consequences of energy consumption for dewatering, potential pollution risks and operational burden [29,30]. To address this drawback, the development of solvent-free synthesis methods for magnetic biochar gradually gains the attention of researchers.
In this study, a magnetic sludge-derived biochar (MSBC-2) was synthesized based on a solvent-free method. The as-prepared material was characterized by scanning electron microscopy (SEM), Fourier transform infrared spectrometer (FTIR), X-ray photoelectron spectroscopy (XPS), vibrating sample magnetometer (VSM) and Raman spectroscopy. The Pb2+, Cd2+ and Cu2+ were selected as representative heavy metal ions and the effects of pH, temperature, background ionic strength and adsorbent dosage on the metal removal efficiency of MSBC-2 were determined. Combining adsorption kinetics, isotherms, thermodynamics analysis and further characterization results, the adsorption mechanisms of three heavy metals onto MSBC-2 were systematically revealed. This work will provide a new perspective for the synthesis of magnetic biochar with excellent metal adsorption capacities.

2. Materials and Methods

2.1. Materials

Cadmium (II) nitrate tetrahydrate (Cd(NO3)2·4H2O, >99.0% purity), copper (II) nitrate trihydrate (Cu(NO3)2·3H2O, >99.0% purity) and Fe3O4 nanoparticles (99.5% purity) were purchased from Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China). Lead (II) nitrate (Pb(NO3)2, >99.0% purity), sodium nitrate (NaNO3, >99.0% purity) and nitric acid (65–68% purity) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ultra-pure water was prepared by the Milli-Q water purification system.

2.2. Preparation of Magnetic Sewage Sludge-Derived Biochar

The sewage sludge used in this study was obtained from the No.2 wastewater treatment plant of Kunming City, with a moisture content of 80.20 ± 0.07%, volatile solid of 43.99 ± 0.05% (dry basis) and C/N of 6.48. The moisture content of the sewage sludge decreased to approximately 30% after several days of drying. Sewage sludge was further air-dried to reduce moisture content (<5%), and ground and passed through a 60-mesh sieve.
Subsequently, 5 g of air-dried sewage sludge was mixed with a certain mass of Fe3O4 nanoparticles and mechanically ground to make a uniform mixture. The mixture was placed in a quartz boat and pyrolyzed at 800 °C for 2 h with a heating rate of 5 °C/min under a N2 atmosphere in a tube furnace (SK-G06123K, Tianjin Zhonghuan Furnace Corp, China). The obtained material was cooled to room temperature, then ground and passed through the 60-mesh sieve. In our previous study, the Fe3O4 nanoparticle content (based on Fe element) was set to 0.07%, 0.15%, 0.36%, 0.58%, 0.72%, 1%, 2% and 5%, respectively [31]. The results showed that the Pb2+, Cd2+ and Cu2+ adsorption capacities were highest when the Fe content was set to 2%, thus magnetic sludge-based biochar with a Fe content of 2% (MSBC-2) was used in this study.

2.3. Characterization of Magnetic Sludge-Based Biochar

The scanning electron microscope (SEM) was used to observe the microscopic morphology (Gemini 300, Zeiss, Birmingham, UK). The X-ray photoelectron spectrometer (XPS) was used to characterize the elemental composition and chemical state of the sample surface before and after adsorption (ESCALAB 250Xi, Thermo Fisher, Waltham, MA, USA). The vibrating Sample Magnetometer (VSM) was used to determine saturation magnetization (LakeShore 7404, Westerville, OH, USA). Raman spectroscopy was measured by the Micro-Raman System 2000 (Renishaw, Wotton-under-Edge, UK). The Fourier transform infrared spectrometer (FTIR) was applied to characterize the functional groups (PerkinElmer Frontier, Waltham, MA, USA).

2.4. Batch Adsorption Experiments

The effects of pH (2, 3, 4, 5, and 6), temperature (25, 35 and 45 °C), ionic strength (0.005, 0.01, 0.05, 0.1 and 0.5 mol·L−1) and adsorbent dosage (0.1, 0.5, 1, 2 and 4 g·L−1) on the metal removal efficiency were determined by batch adsorption experiments. Typically, a certain amount of MSBC-2 was added to the metal ion-containing solution (Pb2+, Cd2+ and Cu2+) of a certain concentration and shaken orbitally at 250 rpm for 24 h at a certain temperature in the incubator. The initial solution pH was adjusted by adding either NaOH or HNO3. After adsorption, the mixtures were separated by centrifugation at 4000 rpm for 20 min. Then the supernatant was filtered by using a 0.45 μM PES syringe filter and diluted in 5% HNO3. The metal concentration was measured by inductively coupled plasma-optical emission spectroscopy (ICP-OES, Prodigy7, LeemanLabs, Hudson, NH, USA). All the experiments were conducted in triplicate.
The adsorption capacity (Qe) and removal efficiency (E) of metal ions onto MSBC-2 were calculated according to Equations (1) and (2).
Q e = ( C 0 C e ) × V 0 m
E = C 0 C e C 0 × 100 %
where Qe is the equilibrium adsorption capacity of adsorbent for heavy metal ions, mg·g−1; C0 and Ce are the initial and equilibrium concentrations of heavy metal ions, mg·g−1; V0 is the volume of the reaction solution, L; m is the addition dosage of adsorbent, g; and E is the removal efficiency of heavy metal ions, %.
As for the kinetics experiments, samples were taken at 1, 3, 5, 10, 20, 30, 60, 90, 120, 180, 240, 360, 720, and 1440 min. Other procedures were the same as above. The results were fitted to pseudo-first-order (Equation (3)), pseudo-second-order (Equation (4)), Elovich (Equation (5)) and intraparticle diffusion (Equation (6)) models.
ln   ( Q e Q t ) = ln   ( Q e ) k 1 t
t Q t = 1 k 2 Q e 2 + t Q e
Q t = 1 b ln ( a b ) + 1 b ln ( t )
Q t = K i d t 0.5 + C
where Qt and Qe are the metal concentrations adsorbed at equilibrium and time t, mg·g−1; k1 is the rate constant of pseudo-first-order model, min−1; k2 is the rate constant of pseudo-second-order, g·(mg·min)−1; a and b represent the initial sorption rate constants (mg·(g·min)−1) and desorption constant (g·mg-1), respectively; Kid is the intraparticle diffusion rate constant, mg/(g·min0.5); and C reflects the boundary layer effect, mg·g−1.
As for the adsorption isotherm, experimental data were fitted to Langmuir (Equation (7)) and Freundlich (Equation (8)) models.
C e Q e = 1 Q m K L + C e   Q m
ln ( Q e ) = ln ( K F ) + 1 n ln ( C e )
R L = 1 1 + K L C 0
where Qm is the maximum adsorption capacity, mg·g−1; KL is the Langmuir adsorption constant, L·mg−1; KF is the Freundlich constant, mg·g−1·(L·mg−1)1/n; n is the empirical heterogeneity factor; and RL is the separation factor.
Based on KL obtained from the Langmuir model, thermodynamic parameters were calculated by Equations (10)–(12).
Δ G 0 = R T ln K L
Δ G 0 = Δ H 0 T Δ S 0
ln K L = Δ S 0 R Δ H 0 R T
where ΔG0 is the free energy change, kJ·mol−1; ΔS0 is the entropy change, kJ·(mol·K)−1; ΔH0 is the enthalpy change, kJ·mol−1; R is the ideal gas constant, 8.314 J·(mol·K)−1; and T is the thermodynamic temperature, K.

2.5. Statistical Analysis

The adsorption capacity (Qe) and removal efficiency (E) data were analyzed by using a one-way ANOVA with a significance set at p < 0.05 (SPSS 26.0, IBM, Armonk, NY, USA).

3. Results and Discussion

3.1. Characterization of MSBC-2

According to Figure 1a,b, the morphology of SBC was smoother, less porous and cleaner than that of MSBC-2. After magnetic modification, various nano-sized particles were exposed on the surface of the carbon skeleton. The specific surface area and pore structure of MSBC-2 and unmodified sewage sludge-derived biochar (SBC) are shown in Table 1. The specific surface area of MSBC-2 was 63.68 m2·g−1, which was higher than that of SBC (59.38 m2·g−1), due to the larger specific surface area of loaded Fe3O4 nanoparticles and the pore expansion effect brought by magnetic modification. The total pore volume of MSBC-2 (0.089 cm3·g−1) increased slightly compared to SBC (0.073 cm3·g−1). An increase in pore volume would accelerate the adsorbate entering the inner pore system of the adsorbent. The average pore diameter of MSBC-2 was 5.96 nm, implying that MSBC-2 had a typical mesoporous structure. Thus, MSBC-2 had a higher specific surface area and more abundant adsorption sites for binding metal ions.
Figure 1c indicates that the saturation magnetization value of MSBC-2 was 5.07 emu·g−1 and MSBC-2 had a typical superparamagnetic behaviour and high saturation magnetization value. Hence, MSBC-2 could be easily recycled from solutions by using an external magnetic field (Figure 1d).
Raman spectroscopy was used to explore the carbon structure (Figure 2a). As-prepared samples exhibited two peaks at approximately 1360 and 1590 cm−1, corresponding to the D band (disordered band) and G band (graphite band). The D-band/G-band (ID/IG) value of SBC was 1.45, while the value of MSBC-2 decreased to 1.42, indicating that magnetic modification would reduce the graphitization degree of biochar. The FTIR spectra of SBC and MSBC-2 before and after adsorption are shown in Figure 2b. The band at 3646 cm−1 was ascribed to the hydroxyl stretching vibration peak, and the peak intensity of the MSBC-2 was stronger than that of SBC, implying that the magnetic modification had loaded more hydroxyl groups onto MSBC-2 [32]. The appearance of a peak around 1796 cm−1 was observed after loading Fe, which corresponded to the stretching vibration of the carboxyl or lactone group. The result indicates that the modification process introduced more aromatic groups in biochar, which might enhance the metal adsorption capacity [33]. The band at 1498 cm−1 was attributed to the deformation vibration of the C-N [34].

3.2. The Influence of Environmental Factors

The impacts of pH, temperature, adsorbent addition dose and ionic strength on metal removal efficiency and adsorption capacity were determined (Figure 3).

3.2.1. pH

The pH value has a critical effect on the surface charge of biochar, thus significantly affecting the adsorption performance. As a higher pH would lead to precipitation, the pH range in this study was set to 2–6 for Cd2+, Pb2+ and Cu2+ (Figure 1a–c) [35]. For Pb2+ and Cd2+, the adsorption efficiency and capacity increased rapidly from pH 2 to 4, and then remained stable from pH 4 to 6. The maximum removal efficiencies were 98.9% for Pb2+ and 99.5% for Cd2+, respectively. For Cu2+, the removal efficiency did not change apparently from pH 4 to 6, which increased from 42.9% to 51.1%. The poor adsorption performance of all three metals at the pH of 2–3 was due to the higher concentration of H+ which would compete and occupy binding sites with heavy metal ions [36]. In addition, several functional groups such as carboxylic groups in MSBC-2 were protonated, resulting in a positive charge on the MSBC-2 and electrostatic repulsion between the heavy metal ions and MSBC-2 [37].
As the pH increased, the deprotonation process occurred and negatively charged carboxyl groups and free hydroxyl groups on the surface would provide more adsorption sites to enhance heavy metal removal [38]. The initial pH was set to 6 for the following batch adsorption experiments.

3.2.2. Temperature

Figure 3e,f presents the effect of temperature on the adsorption of Pb2+, Cd2+, and Cu2+ by MSBC-2. At an initial heavy metal concentration of 100 mg·L−1, the removal efficiency of Pb2+ and Cd2+ by MSBC-2 was almost not affected by temperature due to the excessive adsorption sites, while the copper removal efficiency increased as temperature increased. When the initial Pb2+/Cd2+ concentration increased to 250 mg·L−1, the removal efficiency increased by 14.5% for Pb2+ and 18.3% for Cd2+, respectively, as the temperature increased from 25 °C to 45 °C. The varied influence of temperatures might be due to the different affinities of Pb2+, Cd2+, and Cu2+ to MBSC-2 [39,40].

3.2.3. Adsorbent Dosage

The dosage of adsorbent is another important factor affecting removal efficiency. Insufficient amounts cannot achieve the purpose of treating water pollution caused by heavy metals, while excessive amounts do not make full use of resources and increase costs. The removal efficiency of Pb2+ and Cd2+ increased sharply when the MSBC-2 addition dose was increased from 0.1 g·L−1 to 1 g·L−1, then remained stable, while the inflection point occurred at 2 g·L−1 for Cu2+ (Figure 3g–i). For all three metal ions, adsorption capacities decreased as the adsorbent dosage increased. With the increasing MSBC-2 addition dose, more available adsorption sites were provided, which led to higher metal removal efficiency but reduced the adsorption per unit mass of adsorbent as well [41].

3.2.4. Ionic Strength

The background ionic strengths were set to 0.005–0.5 mol·L−1 NaNO3 (Figure 3i–l). The ionic strength had a negligible effect on the removal of Pb2+ and Cd2+, suggesting that the inner complex might be generated [42]. However, the removal efficiency and adsorption capacity of Cu2+ by MSBC-2 decreased significantly with the increasing ionic strengths (p < 0.05), probably due to competition between Na+ and Cu2+ for adsorption sites [43].

3.3. Adsorption Kinetics

The adsorption capacity of Pb2+, Cd2+ and Cu2+ at different contact time is shown in Figure 4. The adsorption kinetics curves of Pb2+, Cd2+ and Cu2+ onto MSBC-2 were similar and the adsorption process could be divided into three stages: fast adsorption stage, slow adsorption stage and adsorption equilibrium stage. In the fast adsorption stage, the adsorption capacity and metal removal efficiency of MSBC-2 for all three metals increased rapidly, due to abundant adsorption sites on the surface of MSBC-2. Gradually, MSBC-2 adsorption sites were occupied, and then the adsorption capacity increased slowly and finally reached equilibrium. The equilibrium time was 120 min for Pb2+ and Cd2+ adsorption systems, while the Cu2+ adsorption onto MSBC-2 reached equilibrium at about 360 min.
To further explore the adsorption process of Pb2+, Cd2+ and Cu2+ onto MSBC-2, the experimental data were fitted to pseudo-first-order, pseudo-second-order, Elovich and intraparticle diffusion models (Table 2).
As for Pb2+, Cd2+ and Cu2+, the correlation coefficients (R2) of the intraparticle model (R2 = 0.4–0.74) were lower than the fits to the other three models (R2 = 0.74–1.00), indicating that intraparticle diffusion was not the only rate-limiting step during the adsorption process [44]. All kinetics data could be described best by the pseudo-second-order model with the R2 over 0.99, indicating that chemisorption was the most responsible rate-limiting step for adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2 [45,46]. Moreover, the adsorption rates for the three metals were various, and the rate constant sequence was: Pb2+ > Cd2+ > Cu2+. The Elovich model generally presents the heterogeneous chemisorption process [44]. The experimental data for the Elovich models gave high correlation coefficients for Cd2+ (R2 = 0.9124, p < 0.0001) and Cu2+ (R2 = 0.9794, p < 0.0001), implying the adsorption systems were highly heterogeneous.

3.4. Adsorption Isotherms

Figure 5 shows the adsorption isotherms of Pb2+, Cd2+ and Cu2+ onto MSBC-2 at 25 °C, 35 °C and 45 °C. As the initial metal concentrations increased, the adsorption capacities of Pb2+, Cd2+ and Cu2+ by MSBC-2 increased gradually and then tended to reach equilibrium. The adsorption capacities of the three metals were greater at a higher temperature. To further illustrate the adsorption mechanism of Pb2+, Cd2+ and Cu2+ onto MSBC-2, Langmuir and Freundlich models were used to fit the experimental data (Table 3). Compared to Freundlich model, the Langmuir model showed better fits to the experimental data with R2 of 0.9958–0.9999. The results revealed that adsorption processes were monolayer adsorption between the metal ions and oxygen-containing functional groups distributed homogeneously on the MSBC-2 surface [47,48]. When the reaction temperature increased from 25 °C to 45 °C, the maximum adsorption capacities increased from 113.64 mg·g−1 to 151.52 mg·g−1 for Pb2+, from 101.01 mg·g−1 to 109.89 mg·g−1 for Cd2+ and from 57.80 mg·g−1 to 74.07 mg·g−1 for Cu2+, respectively. The adsorption processes were endothermic for all three metal ions.
The separation factor RL (Equation (9)) was calculated from the Langmuir model, which can indicate whether adsorption is favourable (0 < RL < 1), linear (RL = 1), unfavourable (RL > 1) or irreversible (RL = 0) [49]. As shown in Table 3, RL values ranged from 0.0011 to 0.2278, thus the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2 at various temperatures was favourable. The 1/n value represents the degree of heterogeneity [50]. The 1/n values were all less than 1, proving that the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2 was favourable, which was in accordance with RL values [51].

3.5. Adsorption Thermodynamic Analysis

Based on the KL values achieved from Langmuir model, the thermodynamic parameters were calculated and presented in Table 4. Negative ΔG0 values implied the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2 was spontaneous [52]. With the increasing temperatures, ΔG0 decreased for all three metals illustrating that the adsorption process was more spontaneous and favourable at higher temperatures [53]. This could be due to the enhancement of the adsorbate molecules mobility, indicating greater affinity at the higher temperature. The sorting of ΔG0 values was: Cu2+ > Pb2+ > Cd2+, implying that the adsorption of Cd2+ onto MSBC-2 had the highest spontaneity and largest feasibility. All the ΔH0 values were positive, indicating that the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2 was endothermic in nature, which was consistent with the adsorption isotherms results. Typically, the value of ΔH0 in a range of 80–200 kJ·mol−1 suggests chemisorption [54,55]. The calculated ΔH0 values were 91.41 kJ·mol−1, 186.89 kJ·mol−1 and 46.28 kJ·mol−1 for Pb2+, Cd2+ and Cu2+, respectively. The results imply that the adsorption of the three metal ions by MSBC-2 mainly depended on chemical adsorption, which agreed with the adsorption kinetics and adsorption isotherm results. The positive ΔS values exhibited that the randomness on the metal solution-MSBC interface increased and the affinity of metal ions was adequate to adhere to the adsorbent surface during the sorption process [56].

3.6. Adsorption Mechanisms

To further explore the adsorption mechanism, FTIR and XPS analyses were performed by identifying the changes in functional groups before and after adsorption. According to FTIR, the peak intensity of the –OH group was weakened after the adsorption of heavy metals, probably due to the involvement of the –OH group in surface complexation and ion exchange. The intensity of the peak at 1042 cm−1, which was ascribed to C–O stretching vibration, reduced after metal adsorption, suggesting that C–O was involved in the formation of chelates with metals [57].
The XPS spectroscopy was used to investigate the elemental composition and chemical state of MSBC-2 before and after adsorption (Figure 6). The Fe 2p spectrum show two peaks representing Fe 2p3/2 (711.3 eV) and Fe 2p1/2 (725.3 eV), proving that iron was loaded onto biochar successfully [58]. The characteristic binding energies ascribed to Pb 4f (139.3 and 14.1 eV), Cd 3d (412.9 and 406.2 eV) and Cu 2p (411.9 and 405.1 eV) occurred after adsorption, indicating that all three metal ions were adsorbed by MSBC-2 [59]. Specifically, the peaks with the binding energies of 139.3 and 144.1 eV in the Pb 4f were attributed to Pb2+ and Pb-O, and the existence of Pb2+ species exhibited that electrostatic attraction occurred between Pb and MSBC-2 [60]. The Cd 3d could be divided into Cd-π (412.9 eV) and Cd-O binding (406.2 eV), suggesting that coordination with π electrons and electrostatic attraction were involved in the Cd2+ adsorption process [61,62]. The spectra of Cu 2p could be divided into three main peaks Cu 2p1/2, Cu 2p3/2, and shake-up satellites. The peaks with the binding energy of 934.0 eV in the Cu 2d were ascribed to (–COO)2Cu and (–O)2Cu, representing that the carboxyl and hydroxyl groups reacted with the Cu2+ [63]. The appearance of the shake-up satellites demonstrates that an ion exchange existed in the Cu2+ adsorption process [64]. The O 1s XPS spectra of the MSBC-2 was decomposed into three components: Fe–O–H, Fe–O–Pb, Fe–O–Cd, or Fe–O–Cu at ~531.5 eV, C–O at 532.4 eV and O=C–O at 533.6 eV, respectively. After metal ions were adsorbed by MSBC-2, the intensity of the peaks at ~531.5 eV increased from 7.5% to 16.1% for the Pb2+ system, to 24.2% for the Cd2+ system and to 21.0% for the Cu2+ system, due to the interaction of the surface of Fe–O–H with Pb2+/Cd2+/Cu2+ through ligand exchange (O-Pb, O-Cd or O-Cd) [62,65]. The O=C-O peak decreased from 47.9% to 30.4% (for Pb2+), 35.4% (for Cd2+) and 29.5% (for Cu2+), respectively, implying that −COOH groups were involved in surface complexation [65].
In conclusion, surface complexation, cation-π interaction, ion exchange and electrostatic attraction were involved in the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2.

3.7. Comparison with Other Relevant Adsorbents

Table 5 summarizes several adsorption parameters (e.g., Qm and k2) of as-prepared material and other relevant adsorbents. The maximum adsorption capacities of MSBC-2 for Cd2+, Pb2+ and Cu2+ were greater than those of most resembling adsorbents, thus MSBC-2 could be used as a potential adsorbent for Pb2+, Cd2+ and Cd2+ removal.

4. Conclusions

To address the drawback of two-step and one-pot synthesis methods for producing biochar, a magnetic sewage sludge-derived biochar (MSBC-2) was synthesized by a solvent-free method in this study. The adsorption performance of three metal ions (Pb2+, Cd2+ and Cu2+) onto the MSBC-2 was investigated in detail together with adsorption mechanisms. The pseudo-second-order kinetic model (R2 = 0.9971–0.9999, p< 0.0001) and Langmuir model (R2 = 0.9958–0.9999, p < 0.0001) provided the best fit to the experimental data, implying a monolayered chemisorption process of Pb2+, Cd2+and Cu2+ onto MSBC-2. As the temperature increased from 25 °C to 45 °C, the maximum adsorption capacities increased from 113.64 mg·g−1 to 151.52 mg·g−1 for Pb2+, from 101.01 mg·g−1 to 109.89 mg·g−1 for Cd2+ and from 57.80 mg·g−1 to 74.07 mg·g−1 for Cu2+, respectively. Thermodynamic parameters (ΔG0 < 0, ΔS0 > 0, ΔH0 > 0) demonstrated that the adsorption processes of all three metals by MSBC-2 were favourable, spontaneous and endothermic. The adsorption mechanisms involved surface complexation, cation-π interaction, ion exchange and electrostatic attraction for the adsorption of Pb2+, Cd2+ and Cu2+ onto MSBC-2. Overall, this study will provide a new perspective for the synthesis of magnetic biochar and MSBC-2 presents a significant potential as an adsorbent for heavy metal removal.

Author Contributions

Methodology, K.G. and Y.L.; Validation, B.Z.; Formal analysis, J.T. and Y.S.; Investigation, Y.S. and R.L.; Writing—original draft, J.T. and K.G.; Writing—review & editing, T.Y.; Supervision, T.C. and T.Y.; Funding acquisition, T.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 42007128), the Fundamental Research Funds for the Central Universities (2022QNYL27), and the Research Fund of Tsinghua University-China Huaneng Group Co., Ltd. Joint Institute for Base Energy (HNKJ20-H50). The authors would like to thank Yueru Wu from the Shiyanjia Lab (www.shiyanjia.com) for the XPS analysis.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Järup, L. Hazards of heavy metal contamination. Br. Med. Bull. 2003, 68, 167–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, P.; Zhang, J.; Xie, H.; Liu, C.; Liang, S.; Ren, Y.; Wang, W. Heavy metal bioaccumulation and health hazard assessment for three fish species from Nansi Lake, China. Bull. Environ. Contam. Toxicol. 2015, 94, 431–436. [Google Scholar] [CrossRef] [PubMed]
  3. Motsi, T.; Rowson, N.A.; Simmons, M.J.H. Adsorption of heavy metals from acid mine drainage by natural zeolite. Int. J. Miner. Process. 2009, 92, 42–48. [Google Scholar] [CrossRef]
  4. Wang, L.K.; Vaccari, D.A.; Li, Y.; Shammas, N.K. Chemical precipitation. In Physicochemical Treatment Processes; Springer: Berlin/Heidelberg, Germany, 2005; pp. 141–197. [Google Scholar]
  5. Zhang, Z.; Wu, Y.; Luo, L.; Li, G.; Li, Y.; Hu, H. Application of disk tube reverse osmosis in wastewater treatment: A review. Sci. Total Environ. 2021, 792, 148291. [Google Scholar] [CrossRef]
  6. Suhas; Carrott, P.J.M.; Carrott, M.M.L.R. Lignin–from natural adsorbent to activated carbon: A review. Bioresour. Technol. 2007, 98, 2301–2312. [Google Scholar] [CrossRef]
  7. Saleh, T.A.; Mustaqeem, M.; Khaled, M. Water treatment technologies in removing heavy metal ions from wastewater: A review. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100617. [Google Scholar] [CrossRef]
  8. Bartoli, M.; Arrigo, R.; Malucelli, G.; Tagliaferro, A.; Duraccio, D. Recent advances in biochar polymer composites. Polymers 2022, 14, 2506. [Google Scholar] [CrossRef]
  9. Arif, M.; Liu, G.; Yousaf, B.; Ahmed, R.; Irshad, S.; Ashraf, A.; Zia-ur-Rehman, M.; Rashid, M.S. Synthesis, characteristics and mechanistic insight into the clays and clay minerals-biochar surface interactions for contaminants removal-A review. J. Clean. Prod. 2021, 310, 127548. [Google Scholar] [CrossRef]
  10. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  11. Inyang, M.I.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.; Mosa, A.; Pullammanappallil, P.; Ok, Y.S.; Cao, X. A review of biochar as a low-cost adsorbent for aqueous heavy metal removal. Crit. Rev. Environ. Sci. Technol. 2016, 46, 406–433. [Google Scholar] [CrossRef]
  12. Qiu, B.; Tao, X.; Wang, H.; Li, W.; Ding, X.; Chu, H. Biochar as a low-cost adsorbent for aqueous heavy metal removal: A review. J. Anal. Appl. Pyrolysis 2021, 155, 105081. [Google Scholar] [CrossRef]
  13. Zhang, Q.; Hu, J.; Lee, D.-J.; Chang, Y.; Lee, Y.-J. Sludge treatment: Current research trends. Bioresour. Technol. 2017, 243, 1159–1172. [Google Scholar] [CrossRef] [PubMed]
  14. Kelessidis, A.; Stasinakis, A.S. Comparative study of the methods used for treatment and final disposal of sewage sludge in European countries. Waste Manag. 2012, 32, 1186–1195. [Google Scholar] [CrossRef] [PubMed]
  15. Feng, L.; Luo, J.; Chen, Y. Dilemma of sewage sludge treatment and disposal in China. Environ. Sci. Technol. 2015, 49, 4781–4782. [Google Scholar] [CrossRef]
  16. Ghorbani, M.; Konvalina, P.; Walkiewicz, A.; Neugschwandtner, R.W.; Kopecký, M.; Zamanian, K.; Chen, W.-H.; Bucur, D. Feasibility of Biochar Derived from Sewage Sludge to Promote Sustainable Agriculture and Mitigate GHG Emissions—mdash;A Review. Int. J. Environ. Res. Public Health 2022, 19, 12983. [Google Scholar]
  17. Ifthikar, J.; Wang, J.; Wang, Q.; Wang, T.; Wang, H.; Khan, A.; Jawad, A.; Sun, T.; Jiao, X.; Chen, Z. Highly efficient lead distribution by magnetic sewage sludge biochar: Sorption mechanisms and bench applications. Bioresour. Technol. 2017, 238, 399–406. [Google Scholar] [CrossRef] [Green Version]
  18. Zahedifar, M.; Seyedi, N.; Shafiei, S.; Basij, M. Surface-modified magnetic biochar: Highly efficient adsorbents for removal of Pb (ΙΙ) and Cd (ΙΙ). Mater. Chem. Phys. 2021, 271, 124860. [Google Scholar] [CrossRef]
  19. Li, X.; Wang, C.; Zhang, J.; Liu, J.; Liu, B.; Chen, G. Preparation and application of magnetic biochar in water treatment: A critical review. Sci. Total Environ. 2020, 711, 134847. [Google Scholar] [CrossRef]
  20. Ghorbani, M.; Konvalina, P.; Kopecký, M.; Kolář, L. A meta-analysis on the impacts of different oxidation methods on the surface area properties of biochar. Land Degrad. Dev. 2022, 1–14. [Google Scholar] [CrossRef]
  21. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Chen, M. Progress in the preparation and application of modified biochar for improved contaminant removal from water and wastewater. Bioresour. Technol. 2016, 214, 836–851. [Google Scholar] [CrossRef]
  22. Zhang, A.; Li, X.; Xing, J.; Xu, G. Adsorption of potentially toxic elements in water by modified biochar: A review. J. Environ. Chem. Eng. 2020, 8, 104196. [Google Scholar] [CrossRef]
  23. Wang, Y.; Miao, J.; Saleem, M.; Yang, Y.; Zhang, Q. Enhanced adsorptive removal of carbendazim from water by FeCl3-modified corn straw biochar as compared with pristine, HCl and NaOH modification. J. Environ. Chem. Eng. 2022, 10, 107024. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Zhang, R.; Liu, H.; Li, M.; Chen, T.; Chen, D.; Zou, X.; Frost, R. Green preparation of magnetic biochar for the effective accumulation of Pb (II): Performance and mechanism. Chem. Eng. J. 2019, 375, 122011. [Google Scholar] [CrossRef]
  25. Dong, J.; Shen, L.; Shan, S.; Liu, W.; Qi, Z.; Liu, C.; Gao, X. Optimizing magnetic functionalization conditions for efficient preparation of magnetic biochar and adsorption of Pb(II) from aqueous solution. Sci. Total Environ. 2022, 806, 151442. [Google Scholar] [CrossRef] [PubMed]
  26. Liang, S.; Shi, S.; Zhang, H.; Qiu, J.; Yu, W.; Li, M.; Gan, Q.; Yu, W.; Xiao, K.; Liu, B.; et al. One-pot solvothermal synthesis of magnetic biochar from waste biomass: Formation mechanism and efficient adsorption of Cr(VI) in an aqueous solution. Sci. Total Environ. 2019, 695, 133886. [Google Scholar] [CrossRef]
  27. Inbaraj, B.S.; Sridhar, K.; Chen, B.-H. Removal of polycyclic aromatic hydrocarbons from water by magnetic activated carbon nanocomposite from green tea waste. J. Hazard. Mater. 2021, 415, 125701. [Google Scholar] [CrossRef]
  28. Qu, J.; Dong, M.; Wei, S.; Meng, Q.; Hu, L.; Hu, Q.; Wang, L.; Han, W.; Zhang, Y. Microwave-assisted one pot synthesis of β-cyclodextrin modified biochar for concurrent removal of Pb (II) and bisphenol a in water. Carbohydr. Polym. 2020, 250, 117003. [Google Scholar] [CrossRef]
  29. Jack, J.; Huggins, T.M.; Huang, Y.; Fang, Y.; Ren, Z.J. Production of magnetic biochar from waste-derived fungal biomass for phosphorus removal and recovery. J. Clean. Prod. 2019, 224, 100–106. [Google Scholar] [CrossRef]
  30. Yi, Y.; Huang, Z.; Lu, B.; Xian, J.; Tsang, E.P.; Cheng, W.; Fang, J.; Fang, Z. Magnetic biochar for environmental remediation: A review. Bioresour. Technol. 2020, 298, 122468. [Google Scholar] [CrossRef]
  31. Guo, K.; Tian, J.; Sun, Y.; Lin, R.; Gong, Y.; Yang, T.; Chen, T.; Zhuo, Y. Adsorption performance of magnetic sludge-derived biochar towards Pb2+. Chin. J. Environ. Eng. 2022, 16, 1416–1428. [Google Scholar]
  32. Yin, W.; Zhao, C.; Xu, J. Enhanced adsorption of Cd (II) from aqueous solution by a shrimp bran modified Typha orientalis biochar. Environ. Sci. Pollut. Res. 2019, 26, 37092–37100. [Google Scholar] [CrossRef] [PubMed]
  33. Zhou, X.; Zhou, J.; Liu, Y.; Guo, J.; Ren, J.; Zhou, F. Preparation of iminodiacetic acid-modified magnetic biochar by carbonization, magnetization and functional modification for Cd(II) removal in water. Fuel 2018, 233, 469–479. [Google Scholar] [CrossRef]
  34. Li, M.; Wei, D.; Liu, T.; Liu, Y.; Yan, L.; Wei, Q.; Du, B.; Xu, W. EDTA functionalized magnetic biochar for Pb(II) removal: Adsorption performance, mechanism and SVM model prediction. Sep. Purif. Technol. 2019, 227, 115696. [Google Scholar] [CrossRef]
  35. Wang, Q.; Wang, B.; Lee, X.; Lehmann, J.; Gao, B. Sorption and desorption of Pb (II) to biochar as affected by oxidation and pH. Sci. Total Environ. 2018, 634, 188–194. [Google Scholar] [CrossRef] [Green Version]
  36. Shan, R.; Shi, Y.; Gu, J.; Wang, Y.; Yuan, H. Single and competitive adsorption affinity of heavy metals toward peanut shell-derived biochar and its mechanisms in aqueous systems. Chin. J. Chem. Eng. 2020, 28, 1375–1383. [Google Scholar] [CrossRef]
  37. Khan, Z.H.; Gao, M.; Qiu, W.; Islam, M.S.; Song, Z. Mechanisms for cadmium adsorption by magnetic biochar composites in an aqueous solution. Chemosphere 2020, 246, 125701. [Google Scholar] [CrossRef] [PubMed]
  38. Yuan, S.; Hong, M.; Li, H.; Ye, Z.; Gong, H.; Zhang, J.; Huang, Q.; Tan, Z. Contributions and mechanisms of components in modified biochar to adsorb cadmium in aqueous solution. Sci. Total Environ. 2020, 733, 139320. [Google Scholar] [CrossRef] [PubMed]
  39. Elaigwu, S.E.; Rocher, V.; Kyriakou, G.; Greenway, G.M. Removal of Pb2+ and Cd2+ from aqueous solution using chars from pyrolysis and microwave-assisted hydrothermal carbonization of Prosopis africana shell. J. Ind. Eng. Chem. 2014, 20, 3467–3473. [Google Scholar] [CrossRef]
  40. Semerciöz, A.S.; Göğüş, F.; Celekli, A.; Bozkurt, H. Development of carbonaceous material from grapefruit peel with microwave implemented-low temperature hydrothermal carbonization technique for the adsorption of Cu (II). J. Clean. Prod. 2017, 165, 599–610. [Google Scholar] [CrossRef]
  41. Gao, L.; Li, Z.; Yi, W.; Li, Y.; Zhang, P.; Zhang, A.; Wang, L. Impacts of pyrolysis temperature on lead adsorption by cotton stalk-derived biochar and related mechanisms. J. Environ. Chem. Eng. 2021, 9, 105602. [Google Scholar] [CrossRef]
  42. Xiao, J.; Hu, R.; Chen, G.; Xing, B. Facile synthesis of multifunctional bone biochar composites decorated with Fe/Mn oxide micro-nanoparticles: Physicochemical properties, heavy metals sorption behavior and mechanism. J. Hazard. Mater. 2020, 399, 123067. [Google Scholar] [CrossRef] [PubMed]
  43. Gong, J.-L.; Wang, X.-Y.; Zeng, G.-M.; Chen, L.; Deng, J.-H.; Zhang, X.-R.; Niu, Q.-Y. Copper (II) removal by pectin–iron oxide magnetic nanocomposite adsorbent. Chem. Eng. J. 2012, 185, 100–107. [Google Scholar] [CrossRef]
  44. Pholosi, A.; Naidoo, E.B.; Ofomaja, A.E. Intraparticle diffusion of Cr(VI) through biomass and magnetite coated biomass: A comparative kinetic and diffusion study. South. Afr. J. Chem. Eng. 2020, 32, 39–55. [Google Scholar] [CrossRef]
  45. Abdelhafez, A.A.; Li, J. Removal of Pb(II) from aqueous solution by using biochars derived from sugar cane bagasse and orange peel. J. Taiwan Inst. Chem. Eng. 2016, 61, 367–375. [Google Scholar] [CrossRef]
  46. Kołodyńska, D.; Wnętrzak, R.; Leahy, J.J.; Hayes, M.H.B.; Kwapiński, W.; Hubicki, Z. Kinetic and adsorptive characterization of biochar in metal ions removal. Chem. Eng. J. 2012, 197, 295–305. [Google Scholar] [CrossRef]
  47. Allen, S.J.; McKay, G.; Porter, J.F. Adsorption isotherm models for basic dye adsorption by peat in single and binary component systems. J. Colloid Interface Sci. 2004, 280, 322–333. [Google Scholar] [CrossRef]
  48. Zeng, H.; Qi, W.; Zhai, L.; Wang, F.; Zhang, J.; Li, D. Preparation and Characterization of Sludge-Based Magnetic Biochar by Pyrolysis for Methylene Blue Removal. Nanomaterials 2021, 11, 2473. [Google Scholar] [CrossRef]
  49. Shirvanimoghaddam, K.; Czech, B.; Tyszczuk-Rotko, K.; Kończak, M.; Fakhrhoseini, S.M.; Yadav, R.; Naebe, M. Sustainable synthesis of rose flower-like magnetic biochar from tea waste for environmental applications. J. Adv. Res. 2021, 34, 13–27. [Google Scholar] [CrossRef]
  50. Pezoti Junior, O.; Cazetta, A.L.; Gomes, R.C.; Barizão, É.O.; Souza, I.P.A.F.; Martins, A.C.; Asefa, T.; Almeida, V.C. Synthesis of ZnCl2-activated carbon from macadamia nut endocarp (Macadamia integrifolia) by microwave-assisted pyrolysis: Optimization using RSM and methylene blue adsorption. J. Anal. Appl. Pyrolysis 2014, 105, 166–176. [Google Scholar] [CrossRef]
  51. Truong, Q.-M.; Ho, P.-N.-T.; Nguyen, T.-B.; Chen, W.-H.; Bui, X.-T.; Patel, A.K.; Singhania, R.R.; Chen, C.-W.; Dong, C.-D. Magnetic biochar derived from macroalgal Sargassum hemiphyllum for highly efficient adsorption of Cu(II): Influencing factors and reusability. Bioresour. Technol. 2022, 361, 127732. [Google Scholar] [CrossRef]
  52. Dou, D.; Wei, D.; Guan, X.; Liang, Z.; Lan, L.; Lan, X.; Liu, P.; Mo, H.; Lan, P. Adsorption of copper (II) and cadmium (II) ions by in situ doped nano-calcium carbonate high-intensity chitin hydrogels. J. Hazard. Mater. 2022, 423, 127137. [Google Scholar] [CrossRef] [PubMed]
  53. Song, Y.-X.; Chen, S.; You, N.; Fan, H.-T.; Sun, L.-N. Nanocomposites of zero-valent Iron@Activated carbon derived from corn stalk for adsorptive removal of tetracycline antibiotics. Chemosphere 2020, 255, 126917. [Google Scholar] [CrossRef] [PubMed]
  54. Khosa, M.A.; Ullah, A. Mechanistic insight into protein supported biosorption complemented by kinetic and thermodynamics perspectives. Adv. Colloid Interface Sci. 2018, 261, 28–40. [Google Scholar] [CrossRef] [PubMed]
  55. Sag, Y.; Kutsal, T. Determination of the biosorption heats of heavy metal ions on Zoogloea ramigera and Rhizopus arrhizus. Biochem. Eng. J. 2000, 6, 145–151. [Google Scholar] [CrossRef] [PubMed]
  56. Özsin, G.; Kılıç, M.; Apaydın-Varol, E.; Pütün, A.E. Chemically activated carbon production from agricultural waste of chickpea and its application for heavy metal adsorption: Equilibrium, kinetic, and thermodynamic studies. Appl. Water Sci. 2019, 9, 56. [Google Scholar] [CrossRef] [Green Version]
  57. Khan, Z.H.; Gao, M.; Qiu, W.; Song, Z. Properties and adsorption mechanism of magnetic biochar modified with molybdenum disulfide for cadmium in aqueous solution. Chemosphere 2020, 255, 126995. [Google Scholar] [CrossRef]
  58. Wen, T.; Wang, J.; Li, X.; Huang, S.; Chen, Z.; Wang, S.; Hayat, T.; Alsaedi, A.; Wang, X. Production of a generic magnetic Fe3O4 nanoparticles decorated tea waste composites for highly efficient sorption of Cu (II) and Zn (II). J. Environ. Chem. Eng. 2017, 5, 3656–3666. [Google Scholar] [CrossRef]
  59. Zhang, L.; Guo, J.; Huang, X.; Wang, W.; Sun, P.; Li, Y.; Han, J. Functionalized biochar-supported magnetic MnFe2O4 nanocomposite for the removal of Pb (ii) and Cd (ii). RSC Adv. 2019, 9, 365–376. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, L.; Huang, Y.; Zhang, S.; Gong, Y.; Su, Y.; Cao, J.; Hu, H. Adsorption characteristics and mechanism of Pb (II) by agricultural waste-derived biochars produced from a pilot-scale pyrolysis system. Waste Manag. 2019, 100, 287–295. [Google Scholar] [CrossRef]
  61. Chen, Y.; Li, M.; Li, Y.; Liu, Y.; Chen, Y.; Li, H.; Li, L.; Xu, F.; Jiang, H.; Chen, L. Hydroxyapatite modified sludge-based biochar for the adsorption of Cu2+ and Cd2+: Adsorption behavior and mechanisms. Bioresour. Technol. 2021, 321, 124413. [Google Scholar] [CrossRef]
  62. Sun, Y.; Bai, D.; Lu, L.; Li, Z.; Zhang, B.; Liu, Y.; Zhuang, L.; Yang, T.; Chen, T. The potential of ferrihydrite-synthetic humic-like acid composite to remove metal ions from contaminated water: Performance and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2023, 658, 130771. [Google Scholar] [CrossRef]
  63. Feng, Z.; Feng, C.; Chen, N.; Lu, W.; Wang, S. Preparation of composite hydrogel with high mechanical strength and reusability for removal of Cu(II) and Pb(II) from water. Sep. Purif. Technol. 2022, 300, 121894. [Google Scholar] [CrossRef]
  64. Wang, W.; Wen, T.; Bai, H.; Zhao, Y.; Ni, J.; Yang, L.; Xia, L.; Song, S. Adsorption toward Cu(II) and inhibitory effect on bacterial growth occurring on molybdenum disulfide-montmorillonite hydrogel surface. Chemosphere 2020, 248, 126025. [Google Scholar] [CrossRef]
  65. Xue, S.; Fan, J.; Wan, K.; Wang, G.; Xiao, Y.; Bo, W.; Gao, M.; Miao, Z. Calcium-Modified Fe3O4 Nanoparticles Encapsulated in Humic Acid for the Efficient Removal of Heavy Metals from Wastewater. Langmuir 2021, 37, 10994–11007. [Google Scholar] [CrossRef]
  66. Islam, M.S.; Kwak, J.-H.; Nzediegwu, C.; Wang, S.; Palansuriya, K.; Kwon, E.E.; Naeth, M.A.; El-Din, M.G.; Ok, Y.S.; Chang, S.X. Biochar heavy metal removal in aqueous solution depends on feedstock type and pyrolysis purging gas. Environ. Pollut. 2021, 281, 117094. [Google Scholar] [CrossRef]
  67. Neeli, S.T.; Ramsurn, H.; Ng, C.Y.; Wang, Y.; Lu, J. Removal of Cr (VI), As (V), Cu (II), and Pb (II) using cellulose biochar supported iron nanoparticles: A kinetic and mechanistic study. J. Environ. Chem. Eng. 2020, 8, 103886. [Google Scholar] [CrossRef]
  68. Li, A.Y.; Deng, H.; Jiang, Y.H.; Ye, C.H.; Yu, B.G.; Zhou, X.L.; Ma, A.Y. Superefficient removal of heavy metals from wastewater by Mg-loaded biochars: Adsorption characteristics and removal mechanisms. Langmuir 2020, 36, 9160–9174. [Google Scholar] [CrossRef]
  69. Xiang, J.; Lin, Q.; Cheng, S.; Guo, J.; Yao, X.; Liu, Q.; Yin, G.; Liu, D. Enhanced adsorption of Cd (II) from aqueous solution by a magnesium oxide–rice husk biochar composite. Environ. Sci. Pollut. Res. 2018, 25, 14032–14042. [Google Scholar] [CrossRef]
  70. Jain, M.; Yadav, M.; Kohout, T.; Lahtinen, M.; Garg, V.K.; Sillanpää, M. Development of iron oxide/activated carbon nanoparticle composite for the removal of Cr (VI), Cu (II) and Cd (II) ions from aqueous solution. Water Resour. Ind. 2018, 20, 54–74. [Google Scholar] [CrossRef]
  71. Mallakpour, S.; Khadem, E. Chitosan/CaCO3-silane nanocomposites: Synthesis, characterization, in vitro bioactivity and Cu (II) adsorption properties. Int. J. Biol. Macromol. 2018, 114, 149–160. [Google Scholar] [CrossRef]
  72. Mohan, D.; Singh, P.; Sarswat, A.; Steele, P.H.; Pittman, C.U., Jr. Lead sorptive removal using magnetic and nonmagnetic fast pyrolysis energy cane biochars. J. Colloid Interface Sci. 2015, 448, 238–250. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM surface morphology of SBC (a), ×5000 and MSBC-2 (b), ×5000, magnetization curve of MSBC-2 (c) and actual separation performance (d).
Figure 1. SEM surface morphology of SBC (a), ×5000 and MSBC-2 (b), ×5000, magnetization curve of MSBC-2 (c) and actual separation performance (d).
Ijerph 20 00155 g001
Figure 2. Raman spectrum (a) and FTIR spectrum (b) of SBC and MSBC-2.
Figure 2. Raman spectrum (a) and FTIR spectrum (b) of SBC and MSBC-2.
Ijerph 20 00155 g002
Figure 3. Effect of pH (ac), temperature (df), adsorbent dosage (gi) and ionic strength (jl) on removal efficiency of Pb2+, Cd2+ and Cu2+ by MSBC-2.
Figure 3. Effect of pH (ac), temperature (df), adsorbent dosage (gi) and ionic strength (jl) on removal efficiency of Pb2+, Cd2+ and Cu2+ by MSBC-2.
Ijerph 20 00155 g003
Figure 4. Adsorption kinetics curve of Cd2+, Pb2+ and Cu2+ onto MSBC-2 and SBC.
Figure 4. Adsorption kinetics curve of Cd2+, Pb2+ and Cu2+ onto MSBC-2 and SBC.
Ijerph 20 00155 g004
Figure 5. The adsorption curves of Cd2+, Pb2+ and Cu2+ onto MSCB-2 at different temperatures.
Figure 5. The adsorption curves of Cd2+, Pb2+ and Cu2+ onto MSCB-2 at different temperatures.
Ijerph 20 00155 g005
Figure 6. XPS analysis of MSBC-2 before and after Cd2+, Pb2+ and Cu2+ adsorption.
Figure 6. XPS analysis of MSBC-2 before and after Cd2+, Pb2+ and Cu2+ adsorption.
Ijerph 20 00155 g006
Table 1. Specific surface area and pore structure characteristics of MSBC-2 and SBC [31].
Table 1. Specific surface area and pore structure characteristics of MSBC-2 and SBC [31].
SampleBET
(m2·g−1)
Total Pore Volume
(cm3·g−1)
Average Pore Diameter (nm)
MSBC-263.680.0895.96
SBC59.380.0734.56
Table 2. Adsorption kinetics parameters for various adsorption kinetics models.
Table 2. Adsorption kinetics parameters for various adsorption kinetics models.
MetalQe,exp
(mg·g−1)
Pseudo-First-Order ModelPseudo-Second-Order ModelElovich ModelIntraparticle Diffusion Model
Qe,cal
(mg·g−1)
k1(min−1)R2pQe,cal
(mg·g−1)
k2
(g·(mg·min)−1)
R2pa
(mg·(g·min)−1)
b
(g·mg−1)
R2pKid
(mg·(g·min0.5)−1)
C
(mg·g−1)
R2p
Pb95.7544.354.80 × 10−30.8788 0.0006 95.24 3.33 × 10−30.9999<0.00016.75 × 1020.1040.8392<0.00011.33361.940.42630.0213
Cd93.2244.334.60 × 10−30.78960.000693.461.27 × 10−30.9998<0.00011.53 × 1020.095 0.9124<0.00011.63849.410.57500.0043
Cu53.1724.393.00 × 10−30.73980.000252.918.60 × 10−30.9971<0.00016.75 × 1020.201 0.9794<0.00010.86226.870.73500.0002
Table 3. Adsorption isotherm parameters of heavy metals onto MSBC-2 at different temperatures for Langmuir and Freundlich models.
Table 3. Adsorption isotherm parameters of heavy metals onto MSBC-2 at different temperatures for Langmuir and Freundlich models.
MetalTemperature
(°C)
LangmuirFreundlich
Qm
(mg·g−1)
KL
(L·mg−1)
R2pRLKf
(mg·g−1·(L·mg−1)1/n)
nR2p
Pb25113.641.0730.9990<0.00010.0011–0.045839.9584.0240.75270.0052
35131.581.3100.9999<0.00010.0020–0.076844.2283.9860.67090.0129
45151.520.6170.9997<0.00010.0042–0.150149.5113.8490.60360.0233
Cd25101.010.3390.9992<0.00010.0073–0.227834.034.400.53790.0384
35106.380.3550.9997<0.00010.0070–0.219833.084.060.58900.0262
45109.890.3870.9996<0.00010.0064–0.205335.084.130.56010.0327
Cu2557.800.4130.9968<0.00010.0060–0.194927.1835.890.76950.0095
3566.230.5570.9958<0.00010.0045–0.152232.1226.030.72580.0149
4574.070.8080.9966<0.00010.0031–0.110141.046.890.72310.0153
Table 4. Thermodynamic parameters of Cd2+, Pb2+ and Cu2+ onto MSBC-2.
Table 4. Thermodynamic parameters of Cd2+, Pb2+ and Cu2+ onto MSBC-2.
MetalT (K)ΔG0 (kJ·mol−1)ΔS0 (kJ·(mol·K)−1)ΔH0 (kJ·mol−1)R2p
Pb298.15−4.83210.321891.41390.9631<0.05
308.15−7.0649
318.15−11.3108
Cd298.15−4.97830.6430186.88690.9973<0.05
308.15−10.8810
318.15−17.8607
Cu298.15−1.43160.159446.28070.9464<0.05
308.15−2.4190
318.15−4.6460
Table 5. Comparison of Qm and k2 between MSBC-2 and various magnetic biochars.
Table 5. Comparison of Qm and k2 between MSBC-2 and various magnetic biochars.
Raw Materialsk2
(g·(mg·min)−1)
Qm
(mg·g−1)
Reference
Rice straw2.4 × 10−2 (Pb)
9.00 × 10−2 (Cd)
3.90 × 10−2 (Cu)
133.3 (Pb)
42.7 (Cd)
19.6 (Cu)
[66]
Date leaves and stalks3.98 × 10−3 (Pb)
2.70 × 10−3 (Cd)
103.1 (Pb)
106.4 (Cd)
[18]
Cellulose5.00 × 10−3 (Pb)
5.00 × 10−4 (Cu)
17.3 (Pb)
42.2 (Cu)
[67]
Corn straw3.2 × 10−4 (Pb)
1.30 × 10−4 (Cd)
6.10 × 10−4 (Cu)
54.5 (Pb)
66.2 (Cd)
84.8 (Cu)
[68]
Rice husk5.00 × 10−2 (Cd)21.7 (Cd)[69]
Sunflower2.8 × 10−2 (Cd)
2.3 × 10−2 (Cu)
2.9 (Cd)
2.7 (Cu)
[70]
Chitosan4.5 × 10−3 (Cu)33.9 (Cu)[71]
Cane6.33 × 10−4 (Pb)40.6 (Pb)[72]
Sludge3.33 × 10−3 (Pb)
1.27 × 10−3 (Cd)
8.60 × 10−4 (Cu)
113.6 (Pb)
101.0 (Cd)
57.8 (Cu)
This 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

Tian, J.; Guo, K.; Sun, Y.; Lin, R.; Chen, T.; Zhang, B.; Liu, Y.; Yang, T. Solvent-Free Synthesis of Magnetic Sewage Sludge-Derived Biochar for Heavy Metal Removal from Wastewater. Int. J. Environ. Res. Public Health 2023, 20, 155. https://doi.org/10.3390/ijerph20010155

AMA Style

Tian J, Guo K, Sun Y, Lin R, Chen T, Zhang B, Liu Y, Yang T. Solvent-Free Synthesis of Magnetic Sewage Sludge-Derived Biochar for Heavy Metal Removal from Wastewater. International Journal of Environmental Research and Public Health. 2023; 20(1):155. https://doi.org/10.3390/ijerph20010155

Chicago/Turabian Style

Tian, Jiayi, Kexin Guo, Yucan Sun, Ruoxi Lin, Tan Chen, Bing Zhang, Yifei Liu, and Ting Yang. 2023. "Solvent-Free Synthesis of Magnetic Sewage Sludge-Derived Biochar for Heavy Metal Removal from Wastewater" International Journal of Environmental Research and Public Health 20, no. 1: 155. https://doi.org/10.3390/ijerph20010155

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