Next Article in Journal
The Relationship between Emotional Intelligence and Psychological Well-Being among Male University Students: The Mediating Role of Perceived Social Support and Perceived Stress
Previous Article in Journal
Health Literacy: From a Property of Individuals to One of Communities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Behavior and Relative Distribution of Cd2+ Adsorption Mechanisms by the Magnetic and Nonmagnetic Biochars Derived from Chicken Manure

1
Guangdong Industrial Contaminated Site Remediation Technology and Equipment Engineering Research Center, School of Environmental Science and Engineering, Guangdong University of Technology, Guangzhou 510006, China
2
College of Natural Resources and Environment, South China Agricultural University, Guangzhou 510642, China
3
Guangdong Key Laboratory of Water and Air Pollution Control, South China Institute of Environmental Sciences, Ministry of Ecology and Environment, Guangzhou 510530, China
*
Author to whom correspondence should be addressed.
The authors contributed equally to this study.
Int. J. Environ. Res. Public Health 2020, 17(5), 1602; https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph17051602
Submission received: 21 February 2020 / Revised: 25 February 2020 / Accepted: 26 February 2020 / Published: 2 March 2020
(This article belongs to the Section Environmental Science and Engineering)

Abstract

:
The present study investigated the adsorption of Cd2+ by nonmagnetic and magnetic biochars (CMB and M-CMB) derived from chicken manure, respectively. The adsorption characteristics were investigated as a function of initial pH, contact time, initial Cd2+ concentration and magnetic separation. Adsorption process of both biochars were better described by Pseudo-second-order kinetic equation and Freundlich isotherm model, which were spontaneous and endothermic in nature. It was found that maximum capacities were 60.69 and 41.07 mg/g obtained at the initial Cd2+ concentration of 180 mg/L for CMB and M-CMB, and the turbidity of adsorption-treated solution was reduced from 244.3 to 11.3 NTU after magnetic separation of 0.5 min. These indicated that M-CMB had lower adsorption capacity of Cd2+ than CMB, though it was successfully separated from the treated solutions. Furthermore, both biochars before and after adsorption were analyzed by SEM-EDS, XRD and FTIR. Adsorption mechanisms mainly included precipitation, ion-exchange, complexation and Cπ-coordination, in which precipitation and ion-exchange dominated the adsorption process by CMB, while in M-CMB, precipitation was always predominant mechanism, followed by ion-exchange. The two other mechanisms of complexation and Cπ-coordination were trivial in both biochars, jointly contributing 7.21% for CMB and 5.05% for M-CMB to total adsorption. The findings deepen our understanding of the mechanisms governing the adsorption process, which are also important for future practical applications in the removal of heavy metals from wastewater by the biochars.

1. Introduction

Cadmium (Cd) is one of the most toxic heavy metal pollutants in wastewater, since it accumulates in living organisms and passes through the food chain into human organs, causing serious toxicity even at low concentration of 0.001–0.1 mg/L [1]. Significant quantities of Cd are introduced into the environment by anthropogenic activities, including manufacturing of nickel-cadmium batteries, synthetic pigment production, metal coatings, stabilizers in plastic products, and incineration of solid wastes, in which the worldwide production of Cd in 2005 was estimated to be 20,000 metric tons [2]. Accordingly, the adsorption using agricultural waste materials is one of the most popular and effective processes for the removal of heavy metals from wastewater. Recently, the attention has been diverted towards the biochar that is obtained from the pyrolysis process of biomass including crop residues, manure and sewage sludge under low temperature (<700 °C) and oxygen-limited environment [3].
Among the biochars, magnetic biochar have been extensively used for heavy metal adsorption, since it could remove Cd2+ from aqueous solution to some extent and is easily separated from the wastewater [4]. For example, Li et al. [5] reported the synthesis of magnetic biochar by direct heating of siderite and rice husk under N2 condition, and indicated that the adsorption capacity of U(VI) was significantly enhanced after magnetization. While Son et al. [6] developed an engineered magnetic biochar by pyrolyzing waste marine macro-alga, and found that the adsorption efficiency of heavy metal was partially reduced by magnetic biochar. These investigations and other studies [7,8,9,10] showed that the adsorption capacity of heavy metal by magnetic biochar, might be higher or lower than to that of nonmagnetic biochar. Therefore, it is necessary to evaluate the effect of magnetism loading on biochar in terms of the cationic heavy metal adsorption capacity, before a decision for the potential use of magnetic biochar as an industrial adsorbent is made.
The mechanisms responsible for Cd2+ adsorption by the biochars include: (i) precipitation with minerals (such as PO43−, CO32−, SiO32−) [11,12]; (ii) exchange with cations (e.g., K+, Ca2+, Na+, Mg2+) [13,14]; (iii) surface complexation with oxygen-containing functional groups (e.g., −OH, −COOH, −R−OH) [15,16]; and (iv) coordination with Cπ electrons (e.g., C = C, C ≡ C) [13,17]. Among these mechanisms, Cui et al. [18] reported that the precipitation with minerals was the primary mechanism for Cd2+ adsorption on high-temperature biochars derived from Canna indica. Zhang et al. [12] showed that the Cd2+ adsorption primarily involved coordination with Cπ electrons in the biochar derived from Phyllostachys pubescens. Likewise, Wang et al. [16] indicated that Cπ-coordination was the predominant mechanism for Cd2+ adsorption by the biochars derived from bamboo, straw and pig manure. Despite the previous work, there is still a lack of existing literature on the comparison between magnetic and nonmagnetic biochar, focusing on the relative contribution of different mechanisms to total adsorption.
To summarize, the main objectives of the present study were to examine the adsorption characterization of Cd2+ by magnetic and nonmagnetic biochars derived from chicken manure, respectively, and to understand the contributions of each mechanism to total adsorption on a qualitative and quantitative basis by means of SEM-EDS, XRD and FTIR analysis.

2. Materials and Methods

2.1. Biochar Preparation and Characteristics

Chicken manure was collected from the farmland at South China Agricultural University, Guangzhou, China. The raw material was dried at 80 °C for 48 h until moisture was evaporated completely, and then sieved to <2.0 mm particles by a stainless grinding machine. For the synthesis of magnetic biochar (M-CMB), FeCl3·6H2O (20.0 g) and FeSO4·7H2O (11.1 g) was added to the beaker with 600 mL of deionized water and stirred until they were dissolved completely, after which the powdered biomass was mixed with the solutions and stirred at room temperature (25 ± 2 °C) for 20 min. Thereafter, 10 M-NaOH (aqueous) was added drop wise into the mixed suspension until the pH reached to 10–11. After being stirred for 30 min, the suspension was slowly pyrolyzed in a muffle furnace under N2 atmosphere at 600 °C for 4 h according to the methods of Mohan et al. [19]. As a control, nonmagnetic biochar sample (CMB) was prepared under the same pyrolysis conditions, as previously described by Huang et al. [20].
The pH of biochar in water solution was determined at 1:20 (w/v) ratio after stirring for 1 h, and the content of ash was measured by heating at 800 °C for 4 h in a muffle furnace. To determine the point of zero charge (pHpzc), the zeta potential at various pH values was examined by a Zeta Meter (Malvern, Nano ZS90, UK). The BET surface area was measured by N2 adsorption at liquid nitrogen temperature using a Gemini 2360 Micromeritics surface area analyzer and Brunauer-Emmett-Teller equation, and the content of C, H, O, N and S was determined by an elemental analyzer (Flash EA1112, Thermo Finnigan, Italy). The minerals content of K, Ca, Na and Mg were determined by a flame atomic absorption spectrophotometer (FAAS) (M6, Termo Elemental, Waltham, MA, USA), respectively. The content of Fe and Cd in digestion solution were measured by inductively coupled plasma mass spectrometry (ICP-MS, Thermo Fisher Scientific, Waltham, MA, USA). The concentrations of water-soluble PO43− and CO32− were measured by ion chromatography (IC1000, Dionex Co., Ltd. Sunnyvale, CA, USA). The magnetic properties of biochar were determined by a comprehensive physical measurement system (PPMS-9, Quantum Design, San Diego, CA, USA).

2.2. Adsorption Experiments

Metal stock solutions of 1000 mg/L were prepared by dissolving CdCl2·2.5H2O (GR, guaranteed reagent) in double distilled water, and then diluted to the desired concentration prior to the adsorption experiments. Batch adsorption were performed by adding 0.03 g biochar samples to 30 mL synthetic solutions containing a known initial Cd2+ concentration at room temperature (25 ± 2 °C), and agitated at a speed of 150 rpm. The concentrations of ions including Na+ and Cl were lower than the detection limits, and the initial pH of mixed solutions was adjusted to 6.01 ± 0.2 by 0.1 M NaOH or HCl. The effect of pH on the adsorption was investigated in range of 2.0–8.0, and carried out by mixing 1 g/L adsorbents at the initial Cd2+ concentration of 100 mg/L. For adsorption kinetics, biochar samples were mixed with synthetic water containing 20, 50, and 100 mg Cd2+/L at pH 6.0, respectively, and the residual Cd2+ concentration was determined at various time intervals up to 360 min. After adsorption equilibrium, the final suspensions were centrifuged (8000 rpm, 10 min) and then filtered (0.22 μm Millipore filter), the supernatant was thus prepared to the measurement of Cd2+. Likewise, adsorption isotherm experiments were carried out at initial Cd2+ concentrations in the range of 10–180 mg/L under the temperature of 293, 303 and 313 K, respectively. The magnetic separation of biochar particles from aqueous solution was investigated using a magnetic field of 0.5 T, according to the methods of Wang et al. [21].
Lagergren’s pseudo-first-order equation (Equation (1)), and pseudo-second-order equation (Equation (2)) were used to fit the adsorption kinetics process. Langmuir (Equation (3)), and Freundlich (Equation (4)) isotherm equations were applied to model the experimental data. Adsorption thermodynamic (Equations (5) and (6)) was established by fixing free energy (∆G0), enthalpy (∆H0) and entropy (∆S0), according to the methods described by Milonjić [22].
q t = q e ( 1 e k 1 t )
q t = k 2 q e 2 t 1 + k 2 q e t
q e = q max K L C e 1 + K L C e
q e = K F C e 1 n
Δ G 0 = RTln ( ρ ω K D )
ln ( ρ ω K D ) = Δ S 0 R Δ H 0 RT
where qe and qt are the adsorption capacity at equilibrium and time t (mg/g), respectively. k1 (l/min), and k2 (g/mg min) are the rate constants corresponding to the respective kinetic model. Ce (mg/L) is the residual Cd2+ concentration at equilibrium, qmax (mg/g) is the maximum adsorption capacity, KL (L/mg), KF (L/g) and n are the rate constants corresponding to the respective isotherm model. ∆G0 (kJ/mol), ∆S0 (kJ/mol/K), and ∆H0 (kJ/mol) represent the changes of Gibb’s free energy, entropy and enthalpy, respectively. KD (L/g) is the thermodynamic equilibrium constant (Langmuir isotherm constant) and ρ w (g/L) is the water density. R (8.314 J/K/mol) is the gas constant and T (K) is absolute temperature. In the case when KD is given in L/mg, the constants of ρ wKD can be easily recalculated to become dimensionless by multiplying it by 1,000,000 Milonjić [22].

2.3. The Contribution of Different Adsorption Mechanisms

Both biochars (CMB and M-CMB) were demineralized by rinsing with 1 M HCl and distilled water until the pH of liquid became constant, then untreated and demineralized biochars (0.03 g) were added into 30 mL solutions with a known initial Cd2+ concentration (100 mg/L) and without Cd2+, respectively. After adsorption, the reduced amount of Cd2+ by the biochars before and after demineralization could be considered as the contribution of these removed minerals, because most minerals in the biochars were removed by the acid dipping procedure, but the surface oxygen-containing functional groups were not altered [12,23]. After the adsorption of untreated biochars (CMB and M-CMB), the cations including K+, Na+, Ca2+ and Mg2+ in the solution with and without Cd2+ were determined by ICP-MS. After that, a drop of pH before and after Cd2+ adsorption on both demineralized biochars was merely from the complexation with the oxygen-containing functional groups. In addition, the biochars loaded with and without Cd2+ were prepared for chemical analysis using scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDS, Hitachi, Japan), X-ray diffractometer (XRD-6000, Shimadzu, Japan) and fourier transform infrared spectroscopy (FTIR, PerkinElmer 2000, Waltham, WA, USA).
The adsorption capacity resulting from precipitation (Qpre), ion-exchange (Qexc), complexation (Qcom), Cπ-coordination (Q) and total adsorption (Qct), were calculated by the modified methods of Gao et al. [1] and Wang et al. [23], respectively. The corresponding contribution percentage of each mechanism to total adsorption was determined by the Qpre/Qct, Qexc/Qct, Qcom/Qct and Q/Qct ratio.
(i) The adsorption capacity resulted from precipitation was calculated by the difference in the adsorption amount of Cd2+ between on untreated and demineralized biochars.
Q pre = Q ct Q ca Y
where Qca is the adsorption capacity of demineralized biochars (mg/g), Y is the yield of demineralized biochar from original biochar.
(ii) The adsorption capacity attributed to ion-exchange was obtained by subtracting cations released in the 0 mg Cd2+/L solution as background concentration from total released cations in the100 mg Cd2+/L solution.
Q exc = Q k + Q Ca + Q Na + Q Mg
where Qk, QCa, QNa, QMg are the adsorbed Cd2+ by net release amount of K+, Ca2+, Na+ and Mg2+ during the adsorption process, respectively (mg/g).
(iii) The adsorption capacity resulted from complexation was estimated by the difference in pH change before and after adsorption on the demineralized biochars.
Q com = Q pH Y
where QpH is the adsorbed Cd2+ by complexation with oxygen-containing function groups on demineralized biochars (mg/g).
(iv) The adsorption capacity resulted from Cπ-coordination was calculated by subtracting the sum of Qpre, Qexc and Qcom from total adsorption capacity.
Q c π = Q ct Q pre Q exc Q com

2.4. Statistical Analysis

All adsorption experiments were conducted in triplicate, and the standard deviation was calculated by descriptive statistics. Data were statistically analyzed by one-way analysis of variance (ANOVA) at a 0.05 probability level using SPSS 18.0.

3. Results and Discussion

3.1. Biochar Properties

Both biochars were alkaline with high pH values (Table 1, 11.43 for CMB and 11.55 for M-CMB), suggesting that it would potentially remove heavy metals from acid mine drainage [24]. This was correlated to ash content, since the content of ash showed a positive correlation with the alkalinity of bicohar, as observed by many researchers [20,25,26]. High content of ash was also partly reflected by the SEM observations exhibiting many impurities on the surface of biochars (Figure 1). Moreover, the ash contained a large amount of alkaline mineral elements such as K, Ca, Na and Mg, as demonstrated by EDS and XRD results, which might begin to separate from the organic matrix during the pyrolysis process at high temperature (>300 °C), resulting in the increase of pH to strong basicity [25]. In addition, the surface area decreased from 25.56 m2/g in CMB to 5.44 m2/g in M-CMB, which could be attributed to the synthesis of magnetic biochar, causing many micropores were covered by the ash of M-CMB. Similar phenomenon was obtained in the previous work of some researchers reporting the reduction in the surface area after magnetization, because the Fe-containing particles reduced the surface area by filling micropores [6,10,27].
Compared with CMB, the elemental content of M-CMB including C, H, O, and N decreased obviously, resulting in the increase of atomic ratio of H/C and O/C (Table 1). This suggested the increase in abundance of oxygen-containing functional groups such as hydroxyl and carboxylic, which might be responsible for the Cd2+ adsorption by surface complexation or Cπ-coordination [15,28]. Additionally, CMB contained higher content of alkaline metal cations and soluble anions compared to M-CMB, which could play an important role in the Cd2+ adsorption by ion-exchange and precipitation, respectively [29,30].
After the magnetization, M-CMB showed ferromagnetic properties with the saturation magnetization (Ms) of 64.96 emu/g (Figure S1), which was higher than that of pure Fe3O4 materials (Ms = 58.94 emu/g) [21]. Coincidentally, the SEM-EDS results implied that the iron particles might aggregate to the surface of magnetic biochar (Figure 1B), which was indicated by the FTIR spectra observing the stretching vibrations of Fe-O (around 565 cm−1) [31], and further confirmed by XRD pattern implying the formation of Fe3O4. These suggested that M-CMB presented a superior magnetic response, because of massive coverage of biochar surface by iron particles.

3.2. Adsorption Characteristics

3.2.1. pH Adsorption Edge

Both adsorption capacities increased with increasing pH, and began to decline after reaching maximum adsorption at approximately pH 6.0, at which the maximum adsorption capacities were 51.37 mg/g for CMB and 28.30 mg/g for M-CMB, respectively (Figure 2A). The pHpzc was calculated to be 2.18 and 2.44 for CMB and M-CMB, respectively (Figure 2B). When pH < pHpzc, the surface charge of biochars became positive, causing electrostatic repulsion to positively charged Cd2+, thus low level of adsorption was observed at pH 2.0 for both biochars. While pH > pHpzc, the surface charge of biochars became negative, resulting in the significant increase of adsorption capacity from pH 2.0 to 6.0 [16]. Note that CMB showed higher adsorption capacities of Cd2+ than M-CMB, which may be due to the greater amount of negative charge, or larger of specific surface area of CMB compared to M-CMB under the pH range studied (Figure 2A,B). A similar phenomenon that the biochars have lower pHIEP but higher adsorption capacity, was also reported in many previous works reporting that electrostatic interaction was likely the driving force for the Cd2+ adsorption onto biochar surfaces [1,32,33].
Compared with the initial pH, significant increase of equilibrium pH was observed for both biochars (Figure 2C), which was resulted from the addition of alkaline biochars to neutralize the solution acidity. Compared with the system without Cd2+, equilibrium pH was lower when Cd2+ was adsorbed onto biochars (Figure 2C,D). This reduction could be attributed to surface complexation between Cd2+ and oxygen-containing functional groups such as −COOH and −OH, this was accompanied with the release H+ into the solution, resulting in the decrease of equilibrium pH after adsorption [16,34]. Another explanation may be the formation of the Cd precipitate with PO43− and CO32−, reducing equilibrium pH after adsorption [18,20]. These observations suggested that electrostatic ion-exchange, complexation, and precipitation were involved in the adsorption process.

3.2.2. Adsorption Kinetics

Both adsorption generally exhibited initial high removal efficiency within 30 min, and then reached the equilibrium within almost 150 min (Figure 3). The adsorption capacities of CMB were 15.56 mg/g, 28.20 mg/g, and 37.97 mg/g within 150 min at initial Cd2+ concentrations of 20, 50 and 100 mg/L, accounting for 97%, 99%, and 97% to total capacity, respectively. In contrast, it seemed that the adsorption of M-CMB took longer to reach equilibrium, especially at the initial Cd2+ concentrations of 50 mg/L (Figure 3B). The adsorption capacities of M-CMB were much lower than those of CMB at the experimental conditions.
Adsorption kinetics were assessed by pseudo-first-order and pseudo-second-order kinetic model, respectively (Figure 3). For both biochars, all the regression coefficients (R2) of pseudo-second-order kinetic model were above 0.95, which were much larger than those of pseudo-first-order kinetic model (Table S1). Moreover, the adsorption capacities at equilibrium (Qe,cal) calculated from pseudo-second-order kinetic model were very close to the experimental data (Qe, exp). These indicated the adsorption followed the pseudo-second-order mechanism and corresponded to a chemisorption process [35]. In this regard, many studies indicated that the pseudo-second order model could predict almost all adsorption kinetic processes between metals and biomaterials, because of the basic on the adsorption capacity of solid adsorbents that forecasted the adsorption behavior over the entire process [36,37,38].

3.2.3. Adsorption Isotherms

Adsorption capacities increased rapidly with increasing initial Cd2+ concentration for both biochars, and leveled off approximately when the initial metal concentrations were above 120 mg/L in all cases (Figure 4A–C). These could be explained by a stronger driving force with increasing Cd2+ concentrations, leading to higher probability of collision between Cd2+ and biochars [39]. The maximum capacities were 42.44, 47.45 and 60.69 mg/g for CMB at the temperature of 293 K, 303 K, and 313 K, respectively. These were higher than those of M-CMB, with the respective maximum values of 29.32, 29.39, and 41.07 mg/g. Such difference in the adsorption capacities of both biochars was consistent with the results from adsorption kinetics (Figure 3) and pH adsorption edge (Figure 2), which were correlated with the different physico-chemical characteristics of both biochars.
The experimental data were better fitted with the Freundlich isotherm than Langmuir isotherm under the temperature of 293 K, 303 K, and 313 K, respectively (Figure 4D–F). For both biochars, all the R2 values of Freundlich model were greater than 0.95 (0.95–0.99), which were much higher than those of Langmuir model (0.67–0.90). Meanwhile, all the 1/n values were observed in the ranges of 0.13–0.45 (Table S2), which satisfied the favorable adsorption condition (0 < RL < 1) under the concentration ranges studied [35]. These indicated the chemisorption of Cd2+ onto both biochars had a great heterogeneity of adsorption affinity [40]. These results were in line with published data, describing that heavy metal adsorption were well fitted to Freundlich model by the magnetic biochars derived from rice husk [41], loofah sponges [42], and chlorella vulgaris [43].

3.2.4. Adsorption Thermodynamic

The adsorption process was spontaneous and endothermic in nature, with negative values of free energy (∆G0), positive values of enthalpy (∆H0) and entropy (∆S0) (Table 2). For both biochars, the negative values of ∆G0 suggested that the spontaneous and feasible nature of adsorption process at all the studied temperatures. The negative values of ∆G0 decreased as the rising temperature in CMB, implying the driving force to the adsorption increased with increasing temperature [44]. Meanwhile, the positive values of ∆H0 highlighted the endothermic nature of adsorption by both biochars, and the positive values of ∆S0 indicated randomness slightly increased during metal chemisopriotn from solution onto the surface of biochar [8,45].

3.2.5. Magnetic Separation of M-CMB

The turbidity of solution decreased from 244.3 to 11.3 NTU after magnetic separation within 0.5 min, whereas that of the one without magnetic field was 179.7 NTU (Figure 5). Moreover, the turbidity of solution kept relatively high (38.3 NTU) only by gravity after 50 min, while the magnetic separation rate was very rapid, owing to strong magnetism of M-CMB (Ms = 64.96 emu/g, Figure S1). These suggested that magnetic biochar particles could be successfully separated from the treated-solution with the help of external magnet.
Combined with adsorption experimental data, it suggested that CMB had higher adsorption capacity of Cd2+ than that of M-CMB, this might be largely attributed to higher content of soluble ions for CMB, such as K+, Ca2+, Na+, Mg2+, CO32− and PO43− (Table 1), which was beneficial to the adsorption of Cd2+ through mixed mechanisms involving surface complexation, precipitation, ion-exchange, and Cπ-coordination. Similarly, the finding that the magnetization of biochar significantly reduced the adsorption efficiency, were also reported in the previous works, attributing it to the active sites of adsorbent being occupied by iron oxide particles [24,46].

3.3. Adsorption Mechanisms

3.3.1. Metal Precipitation

To confirm the role of precipitation, the untreated and Cd2+-loaded biochars (150 mg/L) were scanned by XRD (Figure 6A). These precipitates on Cd-loaded biochars were possibly identified with typical peaks as CdCO3, Cd3(PO4)2, CdSiO3, Cd(OH)2, CdS, CdP2, and CdFe2O4, although their presence cannot prove definitely. Coincidentally, some white granular crystals were visible on the biochars surface from the SEM observations (Figure 1), and their elemental composition mainly included Cd, C, O, P, S, Fe and Si, as indicated by EDS spectrum. This might partly support the occurrence of surface precipitation with minerals, or surface complexation with oxygen-containing functional groups [47]. The unfamiliar reduction state of P was detected as CdP2 after adsorption, which might be resulted from the release of CH4, H2 and CO during pyrolysis [48]. Among these precipitates, such as CdCO3, Cd3(PO4)2, CdSiO3, Cd(OH)2 and CdS, were probably due to high concentrations of CO32−, PO43−, S and Si content in the original biochars (Figure 1 and Table 1), which were also observed in the previous studies on the heavy metal adsorption by the biochars [1,49,50]. In particular, the precipitates of CdFe2O4 were detected on the surface of M-CMB after adsorption, which could be attributed to the presence of Fe3O4 particles, generating adsorption sites for metal ions [6,10]. Generally, the intensities of XRD peak were strong in both biochars, suggesting the precipitation played an important role in total adsorption.

3.3.2. Surface Complexation

FTIR analysis was applied to characterize the changes of functional groups before and after adsorption (Figure 6B). For M-CMB, the peaks at 3779 cm−1 and 3405 cm−1 representing O-H vibrations of hydroxyl groups, were disappeared after interaction with Cd2+ [24]. The peak at 1626 cm−1 was attributed to C = O stretching vibration of carboxyl, were significantly decreased after adsorption [47]. The band at 1414 cm−1 was assigned to CO32− and the peak at 1045 cm−1 was due to P-O stretching vibrations of PO43− [51], these were changed after adsorption, supporting the occurrence of the metal-carbonate and metal-phosphate precipitates, respectively (Figure 6A). Note that the bands between 880–780 cm−1, 1100–1000 cm−1, and 700–400 cm−1 were characteristic of Si-O-Si symmetric stretching, asymmetric stretching, and bending, respectively [52]. Considering that low content of Si in the original biochar (Figure 1), it would be more accurate to state that Si element probably participates in the adsorption by forming CdSiO3, as partly supported by XRD analysis (Figure 6A). In particular, the peak at 446 cm−1 could be assigned to the stretching vibrations of Fe-O, testifying the formation of iron oxides [21]. In contrast, the bands of CMB at 2983 cm−1 (-CH stretching), 2513 cm−1 (CO32−), 1799 cm−1, 1037 cm−1 (P-O stretching vibrations of PO43−), 874 cm−1 (Si-O-Si symmetric stretching), and 567 cm−1 (Si-O-Si bending) were all increased slightly after adsorption. These suggested that oxygen-containing functional groups such as -OH, -COOH, as well as C-H aromatic group might be involved in the adsorption. However, the intensities of FTIR spectra in both biochars changed slightly within 1–2 units, implying the complexation was an insignificant mechanism in total adsorption. Indeed, the adsorption capacity resulting from this mechanism were 2.13 mg/g and 1.26 mg/g for CMB and M-CMB, respectively (Table 3).

3.3.3. Cπ-Coordination

The cyclic aromatic π-system formed in the biochars could function as the π-donor in the Cd2+ adsorption [17]. According to the FTIR spectra (Figure 6B), the peaks between 900–700 cm−1 and 700–400 cm−1 representing the aromatic C-H groups changed slightly after adsorption, suggesting aromatic functional groups such as γ-CH of furan and β–ring of pyridine might be involved in the adsorption by both biochars [30]. These heterocyclic compounds were a weak cation-π binder and easily bind with Cd2+ by donating π electrons [53], it was thus reasonable to believe that the coordination with Cπ electrons would occur in the adsorption. Similar results were reported by Harvey et al. [13] and Machida et al. [54], who found that the heavy metal adsorption by biochars were mainly attributed to the coordination with delocalized π electrons from the functional groups such as γ-CH and C=C. The formation of Cd-Cπ bonds between Cd2+ and delocalized lone-pair π electrons could be described by the following reaction:
Ijerph 17 01602 i001

3.3.4. Ion-Exchange

Ion-exchange was considered as the dominant mechanism for heavy metal adsorption by the biochars [13,30]. In parallel with the Cd2+ adsorption, significant amount of K+, Ca2+, Na+ and Mg2+ were released into the solution, equivalent to 22.75 and 8.01 mg Cd2+/g (Table 4), accounting for 47.62% and 23.90% of total adsorption by CMB and M-CMB, respectively (Figure 7). These indicated that cation exchange played an important role in total adsorption. Moreover, the dominance of K+ and Ca2+ among these released cations was observed in CMB, this was seemed fairly comparable to the studies of Flores-Cano et al. [14] and Chen et al. [55], who proposed that Cd2+ substitution for Ca2+ was a major mechanism in the adsorption.

3.4. Relative Distribution of Adsorption Mechanisms

The precipitation formed the biggest fraction in M-CMB, because the Qpre was 23.83 mg/g with relatively high contribution of 71.06% to total adsorption (Figure 7), highlighting that the precipitation was the dominant mechanism in total adsorption. While in CMB, the Qpre was 21.57 mg/g, whose contribution proportion was 45.18% in total adsorption. Similar to the precipitation, ion-exchange produced the largest contribution to total adsorption for CMB (47.61%), though its importance was decreased slightly in M-CMB (23.89%). This situation seemed fairly comparable to the finding, in which the heavy metal adsorption by dairy-manure biochar was mainly attributed to the precipitation with PO43− and CO32− [51], and the adsorption by rice-straw biochar was closely related to the exchangeable cations [50], respectively. Thus, the precipitation and ion-exchange dominated the adsorption for both biochars, jointly contributing 92.79% in CMB and 94.95% in M-CMB to total adsorption. On the other hand, complexation and Cπ-coordination were comparably tiny among these mechanisms, since both Qcom and Q values were relatively low. The values of Qcom were 2.13 and 1.26 mg/g for CMB and M-CMB, with their corresponding contribution proportions were 4.46% and 3.76%, respectively. The Q values were 1.31 mg/g for CMB and 0.43 mg/g for M-CMB, whose contribution proportions were 2.75% and 1.29%, respectively.
Based on the present results, magnetic biochar like M-CMB should be further modified to enhance its adsorption ability for heavy metal ions. As a result, a variety of approaches would be applied to improve the adsorption ability of magnetic biochar, such as oxidation [56], amination [46] and loading of metal nanoparticles [57]. Among these methods, more efforts should be paid to explore the effect of new functional groups-impregnated magnetic biochar after modification, since there is a great potential in the enhancement of adsorption capacity for M-CMB, mainly by improving complexation and Cπ-coordination.

4. Conclusions

Compared with the M-CMB, CMB have greater adsorption capacities, faster adsorption kinetics, but have difficulties in separating from the solution after adsorption. For both biochars, the adsorption process followed chemisorption mechanism, occurring on the heterogeneous surface at tested conditions. Among the four mechanisms, the precipitation and ion-exchange dominated the adsorption, in combination contributing 92.79% for CMB and 94.95% for M-CMB, but complexation and coordination were comparably tiny in total adsorption. These could help to improve the application of biochar in environmental remediation for heavy metal. However, the research regarding the enhancement of adsorption capacity by magnetic biochar should be further investigated in future.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1660-4601/17/5/1602/s1, Fig. S1. Magnetization curves of M-CMB at room temperature. Table S1. Kinetic parameters of adsorption for Cd2+ by both biochars at different initial metal concentrations. Table S2. Isotherm parameters of adsorption for Cd2+ by both biochars at different temperatures.

Author Contributions

R.-B.X. conceived and designed the ideas to develop in the article; F.H. analyzed data and wrote the manuscript with help of L.Z., R.-R.W. provided the experimental sets; S.-M.Z. contributed in editing the manuscript. All authors contributed to the final version of this manuscript.

Funding

This research was financially supported by the National Key Research and Development Program of China (2017YFD0801000), National Natural Science Foundation of China (41977116), Key-Area Research and Development Program of Guangdong Province (2019B110207001), Basic and Applied Research Program of Guangdong Province (2019A1515012187), and the Key Laboratory of Water and Air Pollution Control of Guangdong Province (2017A030314001).

Conflicts of Interest

The authors declare that they have no competing financial interests.

References

  1. Gao, L.Y.; Deng, J.H.; Huang, G.F.; Li, K.; Cai, K.Z.; Liu, Y.; Huang, F. Relative distribution of Cd2+ adsorption mechanisms on biochars derived from rice straw and sewage sludge. Bioresour. Technol. 2019, 272, 114–122. [Google Scholar] [CrossRef] [PubMed]
  2. Joseph, P. Mechanisms of cadmium carcinogenesis. Toxicol. Appl. Pharm. 2009, 238, 272–279. [Google Scholar] [CrossRef] [PubMed]
  3. Lehmann, J.; Joseph, S. Biochar for Environmental Management: Science and Technology; Earthscan: London, UK, 2009. [Google Scholar]
  4. Yi, Y.; Huang, Z.; Lu, B.; Xian, J.; Tsang, P.E.; Wen, C.; Fang, J.; Fang, Z. Magnetic biochar for environmental remediation: A review. Bioresour. Technol. 2019, 298, 122468. [Google Scholar] [CrossRef] [PubMed]
  5. Li, M.; Liu, H.; Chen, T.; Dong, C.; Sun, Y. Synthesis of magnetic biochar composites for enhanced uranium(VI) adsorption. Sci. Total Environ. 2019, 651, 1020–1028. [Google Scholar] [CrossRef] [PubMed]
  6. Son, E.B.; Poo, K.M.; Chang, J.S.; Chae, K.J. Heavy metal removal from aqueous solutions using engineered magnetic biochars derived from waste marine macro-algal biomass. Sci. Total Environ. 2018, 615, 161–168. [Google Scholar] [CrossRef] [PubMed]
  7. 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]
  8. Navarathna, C.M.; Karunanayake, A.G.; Gunatilake, S.R.; Pittman Jr, C.U.; Perez, F.; Mohan, D.; Mlsna, T. Removal of Arsenic(III) from water using magnetite precipitated onto Douglas fit biochar. J. Environ. Manag. 2019, 250, 109429. [Google Scholar] [CrossRef]
  9. Zhao, Y.; Zhang, R.; Liu, H.; Li, M.; Chen TChen, 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] [CrossRef]
  10. Wang, X.; Xu, J.; Liu, J.; Liu, J.; Xia, F.; Wang, C.; Dahlgren, R.A.; Liu, W. Mechanism of Cr(VI) removal by magnetic greigite/biochar composites. Sci. Total Environ. 2020, 700, 134414. [Google Scholar] [CrossRef]
  11. Inyang, M.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.R.; Pullammanappallil, P.; Cao, X. Removal of heavy metals from aqueous solution by biochars derived from anaerobically digested biomass. Bioresour. Technol. 2012, 110, 50–56. [Google Scholar] [CrossRef]
  12. Zhang, C.; Shan, B.; Tang, W.; Zhu, Y. Comparison of cadmium and lead sorption by Phyllostachys pubescens biochar produced under a low-oxygen pyrolysis atmosphere. Bioresour. Technol. 2017, 238, 352–360. [Google Scholar] [CrossRef] [PubMed]
  13. Harvey, O.R.; Herbert, B.E.; Rhue, R.D.; Kuo, L.J. Metal interactions at the biochar-water interface: Engergetics and structure-sorption relationships elucidated by flow adsorption microcalorimetry. Environ. Sci. Technol. 2011, 45, 5550–5556. [Google Scholar] [CrossRef] [PubMed]
  14. Flores-Cano, J.V.; Leyva-Ramos, R.; Mendoza-Barron, J.; Guerrero-Coronado, R.M.; Aragón-Piña, A.; Labrada-Delgado, G.J. Sorption mechanism of Cd(II) from water solution onto chicken eggshell. Appl. Surf. Sci. 2013, 276, 682–690. [Google Scholar] [CrossRef]
  15. Park, J.H.; Choppala, G.K.; Bolan, N.S.; Chung, J.W.; Chuasavathi, T. Biochar reduces the bioavailability and phytotoxicity of heavy metals. Plant Soil 2011, 348, 439–451. [Google Scholar] [CrossRef]
  16. Wang, R.Z.; Huang, D.L.; Liu, Y.G.; Zhang, C.; Lai, C.; Zeng, G.M.; Cheng, M.; Gong, X.M.; Wan, J.; Luo, H. Investigating the adsorption behavior and the relative distribution of Cd2+ sorption mechanisms on biochars by different feedstock. Bioresour. Technol. 2018, 261, 265–271. [Google Scholar] [CrossRef]
  17. Keiluweit, M.; Nigo, P.S.; Johnson, M.G.; Kleber, M. Dynamic molecular structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44, 1247–1253. [Google Scholar] [CrossRef] [Green Version]
  18. Cui, X.; Fang, S.; Yao, Y.; Li, T.; Ni, Q.; Yang, X.; He, Z. Potential mechanisms of cadmium removal from aqueous solution by Canna indica derived biochar. Sci. Total Environ. 2016, 562, 517–525. [Google Scholar] [CrossRef]
  19. Mohan, D.; Kumar, H.; Sarswat, A.; Alexandre-Franco, M.; Pittman, C.U., Jr. Cadmium and lead remediation using magnetic oak wood and oak bark fast pyrolysis bio-chars. Chem. Eng. J. 2014, 236, 513–528. [Google Scholar] [CrossRef]
  20. Huang, F.; Gao, L.Y.; Deng, J.H.; Chen, S.H.; Cai, K.Z. Quantitative contribution of Cd2+ adsorption mechanisms by chicken-manure-derived biochars. Environ. Sci. Pollut. Res. 2018, 25, 28322–28334. [Google Scholar] [CrossRef]
  21. Wang, S.Y.; Tang, Y.K.; Li, K.; Mo, Y.Y.; Li, H.F.; Gu, Z.Q. Combined performance of biochar sorption and magnetic separation processes for treatment of chromium-contained electroplating wastewater. Bioresour. Technol. 2014, 174, 67–73. [Google Scholar] [CrossRef]
  22. Milonjić, S.K. A consideration of the correct calculation of thermodynamic parameters of adsorption. J. Serb. Chem. Soc. 2007, 72, 1363–1367. [Google Scholar] [CrossRef]
  23. Wang, Z.; Liu, G.; Zheng, H.; Li, F.; Ngo, H.H.; Guo, W.; Liu, C.; Chen, L.; Xing, B. Investigating the mechanisms of biochar’s removal of lead from solution. Bioresour. Technol. 2015, 177, 308–317. [Google Scholar] [CrossRef] [PubMed]
  24. Lu, H.; Zhang, W.; Yang, Y.; Huang, X.; Wang, S.; Qiu, R. Relative distribution of Pb2+ sorption mechanisms by sludge-derived biochar. Water Res. 2012, 46, 854–862. [Google Scholar] [CrossRef] [PubMed]
  25. Cao, X.; Harris, W. Properties of dairy-manure-derived biochar pertinent to its potential use in remediation. Bioresour. Technol. 2010, 101, 5222–5228. [Google Scholar] [CrossRef] [PubMed]
  26. Xu, X.; Zhao, Y.; Sima, J.; Zhao, L.; Masek, O.; Cao, X. Indispensable role of biochar-inherent mineral constituents in its environmental applications: A review. Bioresour. Technol. 2017, 241, 887–899. [Google Scholar] [CrossRef] [Green Version]
  27. Dewage, N.B.; Liyanage, A.S.; Pittman, C.U., Jr.; Mohan, D.; Mlsna, T. Fast nitrate and fluoride adsorption and magnetic separation from water on α-Fe2O3 and Fe3O4 dispersed on Douglas fir biochar. Bioresour. Technol. 2018, 263, 258–265. [Google Scholar] [CrossRef]
  28. Cao, X.; Ma, L.; Gao, B.; Harris, W. Dairy-manure derived biochar effectively sorbs lead and atrazine. Environ. Sci. Technol. 2009, 43, 3285–3291. [Google Scholar] [CrossRef]
  29. Uchimiya, M.; Lima, I.M.; Thomas Klasson, K.; Chang, S.; Wartelle, L.H.; Rodgers, J.E. Immobilization of heavy metal ions (CuII, CdII, NiII, and PbII) by broiler litter-derived biochars in water and soil. J. Agric. Food Chem. 2010, 58, 5538–5544. [Google Scholar] [CrossRef]
  30. 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. 2015, 46, 406–433. [Google Scholar] [CrossRef]
  31. Li, G.Y.; Jiang, Y.R.; Huang, K.L.; Ding, P.; Chen, J. Preparation and properties of magnetic Fe3O4-chitosan nanoparticles. J. Alloys Compd. 2008, 466, 451–456. [Google Scholar] [CrossRef]
  32. Huang, F.; Dang, Z.; Guo, C.L.; Lu, G.N.; Gu, R.R.; Liu, H.J.; Zhang, H. Biosorption of Cd(II) by live and dead cells of Bacillus cereus RC-1 isolated from cadmium-contaminated soil. Colloids Surf. B 2013, 107, 11–18. [Google Scholar] [CrossRef]
  33. Çolak, F.; Atar, N.; Yazıcıoǧlu, D.; Olgun, A. Biosorption of lead from aqueous solutions by Bacillus strains possessing heavy-metal resistance. Chem. Eng. J. 2011, 173, 422–428. [Google Scholar] [CrossRef]
  34. Wnetrzak, R.; Leahy, J.J.; Chojnacka, K.W.; Saeid, A.; Novotny, E.; Jensen, L.S.; Kwapinski, W. Influence of pig manure biochar mineral content on Cr(III) sorption capacity. J. Chem. Technol. Biot. 2014, 89, 569–578. [Google Scholar] [CrossRef]
  35. Kołodyńska, D.; Wnȩtrzak, 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]
  36. Vijayaraghavan, V.; Yun, Y.S. Bacterial biosorbents and biosorption. Biotechnol. Adv. 2008, 26, 266–291. [Google Scholar] [CrossRef] [PubMed]
  37. Harikishore Kumar Reddy, D.; Lee, S.M. Magnetic biochar composite: Facile synthesis, characterization and application for heavy metal removal. Colloids Surf. A 2014, 454, 96–103. [Google Scholar] [CrossRef]
  38. Kołodyńska, D.; Krukowska, J.; Thomas, P. Comparison of sorption and desorption studies of heavy metal ions from biochar and commercial active carbon. Chem. Eng. J. 2017, 307, 353–363. [Google Scholar] [CrossRef]
  39. Yahya, S.K.; Zakaria, Z.A.; Samin, J.; Santhana Raj, A.S.; Ahmad, W.A. Isotherm kinetics of Cr(III) removal by non-viable cells of Acinetobacter haemolyticus. Colloids Surf. B 2012, 94, 362–368. [Google Scholar] [CrossRef]
  40. Kim, W.K.; Shim, T.; Kim, Y.S.; Hyun, S.; Ryu, C.; Park, Y.K.; Jung, J. Characterization of cadmium removal from aqueous solution by biochar produced from a giant Miscanthus at different pyrolytic temperatures. Bioresour. Technol. 2013, 138, 266–270. [Google Scholar] [CrossRef]
  41. Agrafioti, E.; Kalderis, D.; Diamadopoulos, E. Ca and Fe modified biochars as adsorbents of arsenic and chromium in aqueous solutions. J. Environ. Manag. 2014, 146, 444–450. [Google Scholar] [CrossRef]
  42. Xiao, F.; Chen, J.; Cao, W.; Yang, C.; Chen, J.; Luo, Z. Removal of heavy metals from aqueous solution using chitosan-combined magnetic biochars. J. Colloid Interface Sci. 2019, 540, 579–584. [Google Scholar] [CrossRef] [PubMed]
  43. Lalhmunsiama Gupta, P.L.; Jung, H.; Tiwari, D.; Kong, S.H.; Lee, S.M. Insight into the mechanism of Cd and Pb removal by sustainable magnetic biosorbent precursor to Chlorella vulgaris. J. Taiwan Inst. Chem. Eng. 2017, 71, 206–213. [Google Scholar] [CrossRef]
  44. Ali, I.; Al-Othman, Z.A.; Alwarthan, A.; Asim, M.; Khan, T.A. Removal of arsenic species from water by batch and column operations on bagasse fly ash. Environ. Sci. Pollut. Res. 2014, 21, 3218–3229. [Google Scholar] [CrossRef] [PubMed]
  45. Frantz, T.S.; Silveira, N.; Quadro, M.S.; Andreazza, R.; Barcelos, A.A.; Cadaval, T.R.; Pinto, L.A. Cu(II) adsorption from copper mine water by chitosan films and the matrix effects. Environ. Sci. Pollut. Res. 2017, 24, 5908–5917. [Google Scholar] [CrossRef]
  46. Huang, X.; Wei, D.; Zhang, X.; Fan, D.; Sun, X.; Du, B.; Wei, Q. Synthesis of amino-functionalized magnetic aerobic granular sludge-biochar for Pb(II) removal: Adsorption performance and mechanism studies. Sci. Total Environ. 2019, 685, 681–689. [Google Scholar] [CrossRef]
  47. Zhang, F.; Wang, X.; Yin, D.; Peng, B.; Tan, C.; Liu, Y.; Wu, S. Efficiency and mechanisms of Cd removal form aqueous solution by biochar derived from water hyacinth (Eichornia crassipes). J. Environ. Manag. 2015, 153, 68–73. [Google Scholar] [CrossRef]
  48. Demirbas, A. Effect of temperature and particle size on bio-char yield from pyrolysis of agricultural residues. J. Anal. Appl. Pyrolysis 2004, 72, 243–248. [Google Scholar] [CrossRef]
  49. Xu, Y.; Chen, B. Organic carbon and inorganic silicon speciation in rice-bran-derived biochars affect its capacity to adsorb cadmium in solution. J. Soils Sediments 2015, 15, 60–70. [Google Scholar] [CrossRef]
  50. Park, J.H.; Wang, J.J.; Kim, S.H.; Cho, J.S.; Kang, S.W.; Delaune, R.D.; Han, K.J.; Seo, D. Recycling of rice straw through pyrolysis and its adsorption behaviors for Cu and Zn ions in aqueous solution. Colloids Surf. A 2017, 533, 330–337. [Google Scholar] [CrossRef]
  51. Xu, X.; Cao, X.; Zhao, L. Comparison of rice husk- and dairy manure-derived biochars for simultaneously removing heavy metals from aqueous solutions: Role of mineral components in biochars. Chemosphere 2013, 92, 955–961. [Google Scholar] [CrossRef]
  52. Xiao, X.; Chen, B.; Zhu, L. Transformation, morphology, and dissolution of silicon and carbon in rice straw-derived biochars under different pyrolytic temperatures. Environ. Sci. Technol. 2014, 48, 3411–3419. [Google Scholar] [CrossRef] [PubMed]
  53. Keiluweit, M.; Kleber, M. Molecular-level interactions in soils and sediments: The role of aromatic π-systems. Environ. Sci. Technol. 2009, 43, 3421–3429. [Google Scholar] [CrossRef] [PubMed]
  54. Machida, M.T.; Mochimaru, T.; Tatsumoto, H. Lead(II) adsorption onto the graphene layer of carbonaceous materials in aqueous solution. Carbon 2006, 44, 2681–2688. [Google Scholar] [CrossRef]
  55. Chen, T.; Zhou, Z.; Han, R.; Meng, R.; Wang, H.; Lu, W. Adsorption of cadmium by biochar derived from municipal sewage sludge: Impact factors and adsorption mechanism. Chemosphere 2015, 134, 286–293. [Google Scholar] [CrossRef]
  56. Sun, C.; Chen, T.; Huang, Q.; Wang, J.; Lu, S.; Yan, J. Enhanced adsorption for Pb(II) and Cd(II) of magnetic rice husk biochar by KMnO4 modification. Environ. Sci. Pollut. Res. 2019, 26, 8902–8913. [Google Scholar] [CrossRef]
  57. Rajput, S.; Pittman, C., Jr.; Mohan, D. Magnetic magnetite (Fe3O4) nanoparticle synthesis and applications for lead (Pb2+) and chromium (Cr6+) removal from water. J. Colloid Interface Sci. 2016, 468, 334–346. [Google Scholar] [CrossRef]
Figure 1. SEM images (left) and corresponding EDS spectra (right) of CMB (A) and M-CMB (B) before and after adsorption for Cd2+.
Figure 1. SEM images (left) and corresponding EDS spectra (right) of CMB (A) and M-CMB (B) before and after adsorption for Cd2+.
Ijerph 17 01602 g001
Figure 2. Effect of pH on the adsorption by both biochars (A). Zeta potential of both biochars at different initial pH values (B). Equilibrium pH of mixed solution without Cd2+ at different initial pH values (C). Equilibrium pH of mixed solution with Cd2+ at different initial pH values (D).
Figure 2. Effect of pH on the adsorption by both biochars (A). Zeta potential of both biochars at different initial pH values (B). Equilibrium pH of mixed solution without Cd2+ at different initial pH values (C). Equilibrium pH of mixed solution with Cd2+ at different initial pH values (D).
Ijerph 17 01602 g002
Figure 3. Effect of contact time on the adsorption at initial Cd2+ concentrations of 20 (A), 50 (B) and 100 mg/L (C), respectively.
Figure 3. Effect of contact time on the adsorption at initial Cd2+ concentrations of 20 (A), 50 (B) and 100 mg/L (C), respectively.
Ijerph 17 01602 g003
Figure 4. Effect of initial Cd2+ concentration on the adsorption under the temperature of 293 (A), 303 (B) and 313 K (C), respectively. Adsorption isotherms of the biochars under the temperature of 293 (D), 303 (E) and 313 K (F), respectively.
Figure 4. Effect of initial Cd2+ concentration on the adsorption under the temperature of 293 (A), 303 (B) and 313 K (C), respectively. Adsorption isotherms of the biochars under the temperature of 293 (D), 303 (E) and 313 K (F), respectively.
Ijerph 17 01602 g004
Figure 5. Effect of time on magnetic separation of M-CMB by a magnetic field of 0.5 T.
Figure 5. Effect of time on magnetic separation of M-CMB by a magnetic field of 0.5 T.
Ijerph 17 01602 g005
Figure 6. XRD patterns of both biochars before and after Cd2+ adsorption (A). FTIR spectra of both biochars before and after Cd2+ adsorption (B).
Figure 6. XRD patterns of both biochars before and after Cd2+ adsorption (A). FTIR spectra of both biochars before and after Cd2+ adsorption (B).
Ijerph 17 01602 g006
Figure 7. Adsorption capacity of Cd2+ adsorption by both biochars (A). Contribution percentage of different mechanisms to total adsorption by both biochars (B).
Figure 7. Adsorption capacity of Cd2+ adsorption by both biochars (A). Contribution percentage of different mechanisms to total adsorption by both biochars (B).
Ijerph 17 01602 g007
Table 1. The main properties of both biochars.
Table 1. The main properties of both biochars.
BiocharpHpHpzcAsh (%)SBET (m2/g)Fe (mg/g)Magnetism (emu/g)Elemental Content (%)Atomic RatioSoluble Ions Content (mg/g)
CHONSH/CO/CK+Ca2+Na+Mg2+CO32−PO43−
CMB11.432.1878.6425.5611.351.0723.070.890.251.131.528.370.303.340.021.190.03
M-CMB11.552.4486.645.4410.8264.963.170.3715.930.720.451.403.775.710.020.520.010.630.01
SBET is the Brunauer-Emmett-Tell (BET) surface area, m2/g. pHpzc is the pH at which the surface charge of biochar is zero.
Table 2. Thermodynamic parameters of adsorption for Cd2+ by both biochars.
Table 2. Thermodynamic parameters of adsorption for Cd2+ by both biochars.
Biochar∆G0 (kJ mol−1)∆H0 (kJ mol−1)∆S0 (kJ mol−1 K−1)
293 K303 K313 K
CMB−32.37−35.63−36.2925.540.20
M-CMB−25.11−36.42−31.6373.050.34
Table 3. Changes in hydrogen ion concentration of adsorption for Cd2+ by both biochars.
Table 3. Changes in hydrogen ion concentration of adsorption for Cd2+ by both biochars.
BiocharInitial pHFinal pH△H+(10−6 mmol)Qe (mg/g)
CMB5.96 ± 0.023.87 ± 0.01134 ± 0.062.13 ± 0.02
M-CMB5.02 ± 0.014.36 ± 0.0234.1 ± 0.181.26 ± 0.01
Table 4. The release of K+, Ca2+, Na+ and Mg2+ during the Cd2+ adsorption by both biochars.
Table 4. The release of K+, Ca2+, Na+ and Mg2+ during the Cd2+ adsorption by both biochars.
BiocharThe Net Amount of Released Cations (mequiv/g)SumOverall Adsorbed Cd2+(mg/g)
K+Ca2+Na+Mg2+
CMB14.43 ± 0.269.58 ± 0.387.10 ± 1.732.41 ± 0.0733.52 ± 1.8322.75 ± 0.88
M-CMB7.90 ± 0.131.74 ± 0.084.11 ± 0.100.26 ± 0.0214.01 ± 0.088.01 ± 0.06

Share and Cite

MDPI and ACS Style

Huang, F.; Zhang, L.; Wu, R.-R.; Zhang, S.-M.; Xiao, R.-B. Adsorption Behavior and Relative Distribution of Cd2+ Adsorption Mechanisms by the Magnetic and Nonmagnetic Biochars Derived from Chicken Manure. Int. J. Environ. Res. Public Health 2020, 17, 1602. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph17051602

AMA Style

Huang F, Zhang L, Wu R-R, Zhang S-M, Xiao R-B. Adsorption Behavior and Relative Distribution of Cd2+ Adsorption Mechanisms by the Magnetic and Nonmagnetic Biochars Derived from Chicken Manure. International Journal of Environmental Research and Public Health. 2020; 17(5):1602. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph17051602

Chicago/Turabian Style

Huang, Fei, Lu Zhang, Ren-Ren Wu, Si-Ming Zhang, and Rong-Bo Xiao. 2020. "Adsorption Behavior and Relative Distribution of Cd2+ Adsorption Mechanisms by the Magnetic and Nonmagnetic Biochars Derived from Chicken Manure" International Journal of Environmental Research and Public Health 17, no. 5: 1602. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph17051602

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