Next Article in Journal
Predicting Parameters of Heat Transfer in a Shell and Tube Heat Exchanger Using Aluminum Oxide Nanofluid with Artificial Neural Network (ANN) and Self-Organizing Map (SOM)
Next Article in Special Issue
Application of a Partial Nitrogen Lab-Scale Sequencing Batch Reactor for the Treatment of Organic Wastewater and Its N2O Production Pathways, and the Microbial Mechanism
Previous Article in Journal
Structure and Properties of New Antifriction Composites Based on Tool Steel Grinding Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Remediation of Emerging Heavy Metals from Water Using Natural Adsorbent: Adsorption Performance and Mechanistic Insights

1
State Key Laboratory of Advanced Special Steel, School of Materials Science and Engineering, International Joint Laboratory of Catalytic Chemistry, College of Sciences, Shanghai University, Shanghai 200444, China
2
Institute of Chemical Sciences, Gomal University, Dera Ismail Khan 29220, Pakistan
3
Department of Pharmaceutical Chemistry, College of Pharmacy, King Khalid University, Abha 62529, Saudi Arabia
4
Ministry of Education (MOE) Key Laboratory of Environment Remediation and Ecology Health, College of Environmental and Resources Science, Zhejiang University, Hangzhou 310058, China
5
Department of Computer Science, Federal Urdu University, Islamabad 44000, Pakistan
6
School of Chemical and Biomolecular Engineering, The University of Sydney, Camperdown, NSW 2006, Australia
*
Authors to whom correspondence should be addressed.
Sustainability 2021, 13(16), 8817; https://doi.org/10.3390/su13168817
Submission received: 5 July 2021 / Revised: 27 July 2021 / Accepted: 3 August 2021 / Published: 6 August 2021
(This article belongs to the Special Issue Advances in Technologies for Wastewater Treatment and Reuse)

Abstract

:
The presence of potentially toxic metals in water causes a strong impact on environment and human health. In this study, activated biochar was produced by using chemical oxidation method from wheat straw as natural adsorbent and was employed for heavy metals competitive remediation. The morphology, structure, and chemical properties of biochar before and after adsorption were characterized by FTIR, XRD, SEM and EDX mapping techniques. The competitive adsorption efficiency of adsorbent for divalent cadmium (Cd) and lead (Pb) from contaminated water was investigated by using wide range of several initial metal concentration, contact time and pH. Maximum adsorption of Cd(II) and Pb(II) was found in the pH range of 6–8. The adsorption capacity for Cd(II) and Pb(II) was 8.85 and 9.03 mg/g, respectively. Thermodynamics parameters and kinetic models were applied to adsorption data. The isotherm data followed Langmuir model, corresponding to monolayer adsorption of the two ions in the contaminated water. The kinetic data followed the pseudo 2nd order kinetics model, which authenticates the chemisorption nature. The thermodynamic study indicated that Cd adsorption is a spontaneous exothermic process while Pb adsorption is an endothermic process. Mineral precipitation, surface complexation, and cation-π interactions are the major remediation strategies for Cd(II) and Pb(II).

1. Introduction

The occurrence of heavy metals in industrial wastes, pesticides, fertilizers, metal mining, and energy utilization process are among the most hazardous pollutants to human beings due to their toxicity and non-degradable nature [1]. Water containing heavy metals causes serious irritation to the respiratory tract, and damage the liver, kidney and human olfactory sense through prolonged contact. Cr(V), Fe(III), Ni(II), Zn(II), Cd(II), Hg(II), and Pb(II) are the most well-known toxic heavy metal pollutants in wastewater, which need to be removed by appropriate methods before water recycling and reuse [2]. Cd(II) and Pb(II) have been considered most critical environmental pollutants due to their carcinogenic nature and longer persistent in the environment. It has been found that the excessive intake of Cd(II) can cause kidney, lungs, and bone problems, whereas Pb(II) can potentially damage the genital and nervous systems [3]. Thus, cost-effective methods to remove Cd(II) and Pb(II) from wastewater are urgently demanded. Up to now, several nominal approaches have been developed for heavy metals removal [4], including ion exchange [5], electrochemical [6], neutralization precipitation [7], membrane filtration [8], nanofiltration [9], flotation [10], and adsorption [11], of which adsorption is considered as one of the most environment-friendly and highly-efficient option [12]. To achieve good adsorption performance, the selection of adsorbents is very crucial. Activated carbon [13], and molecular sieves [14] were widely used because of their distinctive adsorption properties, resulting from firm micro and mesoporous structures. However, their high cost, complex preparation and recycling methods make it difficult to be conducted on a large-scale.
Natural adsorbent as biochar is a non-toxic material synthesized from agricultural wastes by the pyrolysis of biomass during carbonization and is considered as a good candidate to supplant the high-cost adsorbents for the expulsion of heavy metals from wastewater [15]. The biochar poses high organic carbon contents and aromatic structures that intensify its adsorption capacity [16]. Moreover, the higher oxygen content and acidic groups of biochar also plays a part in increasing its metal sorption efficiency [17]. Biomass type and carbonization conditions are important parameters that affect the pore size and chemical properties of biochar. The world annual production of wheat straw is about 529 million tons/year [18], and it contains high values of cellulose (34–40%), hemicellulose (20–25%) and lignin (20%) [19]. The low cost and abundance of organic functional groups in the wheat straw make it a potential adsorbent for the binding of heavy metals after carbonization. Usually, the pristine biochar derived from wheat-straw has a lower surface area due to the narrow microspores and poor heavy metals removal efficiency. Hence, an activation step is required to enlarge the porosity and enhancement of mass transfer fluxes and active loadings [20].
The activation process is classified into physical and chemical methods based on the activation mechanism [21]. Physical activation is the process in which oxidizing agents (H2O, O2, or CO2) are used to develop porosity by controlled carbon gasification, while chemical activation uses chemical agents to enhance the gasification rate of biochar [22]. The frequently used chemicals for chemical activation are HCl [23], NaOH [24], KOH [25], H3PO4 [26], and ZnCl2 [27]. Ahmed et al. [28] revealed that chemical modification of biochar can enhance its surface properties and produce a higher number of oxygenated functional groups where alkaline modification improves the surface aromaticity and removes the organic and inorganic contaminants efficiently. Naeem et al. [29] investigated wheat straw biochar (WSB) and activated wheat straw biochar (AWSB) by using batch and column methods for the removal of Cd(II) and found that AWSB removes Cd(II) more efficiently from contaminated water. Stasi et al. [20] reported that physically activated wheat straw by CO2 has good catalytic ability and stability. Recently, Yu et al. [3] explored the sulfonated biochar for the removal of heavy metals. Previously, Inyang et al. [30] investigated the anaerobically digested biomass-derived biochar for the remediation of heavy metals. Although it is meaningful and economical to transform agricultural wastes into biochar for sorting the environmental risks, however the understanding of heavy metals adsorption by biochar remains challenging, due to the interaction between biochar and heavy metals, the long-term attachment of heavy metals, and the release of heavy metals from biochar. Furthermore, activation through the chemical oxidation process is less/never investigated for competitive bi-heavy metals remediation.
In this study, AWSB was prepared by activating the WSB through the chemical oxidation process which increases the surface functionalities like oxygen-containing groups and hydrophilicity of WSB after activation. Activation by chemical oxidation can also increase the pore size, resulting in improved competitive adsorption capacity of polar adsorbates, i.e., Cd(II) and Pb(II). Batch experiments and characterizations were performed to evaluate the remediation and competitive adsorption of Cd(II) and Pb(II) by using AWSB. The uptake ability of biochar as well as the effect of metal concentration, contact time, and pH on Cd and Pb adsorption was evaluated. Moreover, the adsorption isotherms, thermodynamics, kinetics, and the adsorption mechanism were investigated in detail.

2. Materials and Methods

2.1. Preparation and Characterization of AWSB

Wheat straw was used as a feedstock for biochar preparation and was collected from the KOND farm of the Pir Mehr Ali Shah (PMAS) Arid Agriculture University, Rawalpindi, Pakistan. Firstly, wheat straw was washed several times with deionized water to remove all impurities and then dried in an oven at 333 K for 24 h. Then, 15 g of dried wheat straw was transferred to a ceramic crucible and pyrolyzed in a muffle furnace at 673 K for 2 h. The sample was cooled to room temperature and crushed into fine powder. The WSB was activated with 1 mol/L HCl solution for 1 h and then washed with deionized water several times until pH 7. After drying at 333 K for 4 h, the resultant powder was obtained and named as AWSB.
The AWSB was characterized before and after adsorption by several innovative techniques including scanning electron microscopy (SEM; Tescan Vega 3), energy dispersive X-ray (EDX) analysis, Fourier transforms infrared spectroscopy (FT-IR) (Bruker IFS-66), and X-ray Diffraction (XRD) analysis (X’Pert, DY-3805).

2.2. Adsorption Experiments

The stock solution of the bimetallic equimolar aqueous solution of Cd and Pb was prepared by dissolving cadmium nitrate (A.R., Cd(NO3)2; 8.90 mM) and lead nitrate (A.R., Pb(NO3)2; 4.80 mM) in deionized water. Standard solutions (2, 4, 8, 15, 30, and 50 ppm) were then precisely prepared from the stock solution and kept at room temperature for further tests. To evaluate the adsorption of Cd and Pb on AWSB (0.5 g), 50 mL of standard solutions of various concentrations were used. The mixture was then incubated at 180 rpm in a shaker for 2 h and later filtered. The Cd and Pb concentrations before and after adsorption were analysed by atomic absorption spectrophotometer (AAS; model AA-6300, GFA-Ex-71 Shimadzu). The pH of standard solutions was adjusted to 2, 4, 6, 8, 10, and 12 by 1 mol/L HCl and 1 mol/L NaOH solutions to study the influence of pH on Cd and Pb adsorption. The impact of Cd and Pb concentrations on adsorption was studied employing standard solutions with metal concentrations of 2, 4, 8, 15, 30, and 50 ppm. The effect of contact time was analysed after shaking for 0–16 h. The removal efficiency of each AWSB for Cd and Pb was calculated using the following Equation (1):
Removal   Efficiency   % = C 0 C e C 0 × 100

2.3. Adsorption Isotherms

Isotherm modelling has vital significance while planning a sorption-based system which identifies the metal division as a function of the metal concentration between solid adsorbent and aqueous media. The concentration of metals over the adsorbent increases till equilibrium and thereafter the metal ion concentration constantly distributes between the solid-liquid phase. Three models were frequently used to fit the heavy metal adsorption isotherms. The Langmuir model shows the perfect monolayer adsorption which demonstrates that there is no mutual interaction between adsorption of metals because of equivalent adsorption energy of all adsorption sites [31]. The Freundlich model explains the heterogeneous absorption with continuously increasing adsorption degree at higher concentrations [32]. The Temkin model is suitable for the system in which there is a linear decline in the adsorption heat by the coverage of adsorption because of the interaction between adsorbate and adsorbent [33]. The Langmuir, Freundlich, and Temkin isotherm’s linear forms are mentioned in Equations (2)–(5) [34]:
1 Q e = 1 Q m + 1 K L   Q m   1 C e
lnQ e = lnK F + 1 n   lnC e
Q e = R T b   ln AC e
B = RT b
R L = 1 1 + K L C 0
where “Qe” is the total Cd and Pb adsorbed at equilibrium on AWSB (mg/g); “Qm” is the maximum adsorption capacity of Langmuir (mg/g); “KL” is the affinity constant of Langmuir (L/mg); “Ce” is the equilibrium concentration of Cd and Pb in solution (mg/L); “KF” is the Freundlich constant (mg/g)·(L/g)n; “n” is the dimensionless constant; “R” is the constant of universal gas (8.314 J/mol); “T” is the temperature (K); “b” is the constant of Temkin isotherm; “A” is the constant of Temkin equilibrium binding (L/g); “RL” is the separation parameter as calculated by Equation (6), which might be 0 (irreversible), 1 (linear), 0–1 (favorable) or >0 (unfavorable); C0 is the initial concentration of Cd and Pb [35]. Freundlich dimensionless constant describes the favorable adsorption and strong interaction between heavy metal and AWSB when its value varies between 1 to 10 [33].

2.4. Thermodynamics

Thermodynamic investigations can explain the effect of temperature on adsorption performance. The equilibrium constant (K) is determined by Equation (7) [34], while Gibbs free energy (ΔG°), entropy (ΔS°), and enthalpy (ΔH°) are calculated by the Equations (8)–(10) [34].
K = K F n × M × 1000 × 55.5
G = RT   lnK
G = H + T S
lnK = H R   1 T + S R
where “M” is the metal’s molecular weight (Cd: 112.41 g/mol; Pb: 207.2 g/mol); “55.5” is the molar concentration of water; “R” is the constant of the universal gas (8.314 J/mol K); “T” is the temperature (K).

2.5. Kinetics

Adsorption kinetics was studied to evaluate the mechanism of metal adsorption on the surface of AWSB. The rate of reaction and degree of adsorption was well explained by kinetic equations. The pseudo 1st order kinetics, pseudo 2nd order kinetics, power function, and intraparticle diffusion are mentioned in Equations (11)–(14).
log q e q t = log q e K l t 2.303
t q t = 1 K 2 q 2 e + t q e
Lnq t = Lnb   +   k f Lnt
  q t = k dif   t + C
where “qe” shows the amount of metal adsorbed by the adsorbent; “qt” indicates the quantity of Pb and Cd adsorbed at a standardized time (mg/g); “K1” and “K2” corresponds to the pseudo 1st order kinetic and pseudo 2nd order kinetic models; “t” represents time (h); “kf” represents power function rate constants, “ki” represents constant intraparticle diffusion (mg/g·min0.5); “C” is a constant related to the border layer thickness (mg/g).

3. Results and Discussion

3.1. Characterizations of AWSB

The surface morphology and porous structure of pristine/pre-adsorbed WSB and AWSB were studied by SEM technique. As shown in Figure 1A and Figure S1, the pre-adsorbed AWSB has a heterogeneous structure with large bulk and smaller cracks as compared to pristine WSB. The bulk cracks contain parallel pores with a pore size of ~5 μm with a smooth pore wall. SEM-EDX mapping was performed to study the composition of pristine WSB and AWSB (Figure S2 and Figure 1B). The elemental percentage is shown in Table 1.
As a result of pyrolysis treatment, the carbon content of the biochar increases. FTIR spectrum was used to investigate the functional groups on the surface of pristine WSB and AWSB. As shown in Figure 2A and Figure S3A, the—OH stretching peaks at 3373 cm−1 and 3356 cm−1 and C=C stretching vibrations of aromatic rings at 1577 cm−1 and 1637 cm−1 were observed on the pristine WSB and AWSB [33]. The O—H stretching peak becomes stronger after activation. The peak at 1380 cm−1 is allotted to CH3 bending for pristine WSB where 1602 cm−1 is assigned to C=C stretching [29,31,33], while the peak around 650 cm−1 corresponds to CO32− [36] for pre-adsorbed AWSB. The number and classes of functional groups can decide the sorption capacity of the biochar and govern its adoption mechanism [31]. The C=C functional groups of biochar were usually considered favorable for metal adsorption by forming a surface complex through inter-metal-π-interactions [37]. XRD technique was used to study the crystallographic structure of the AWSB sample. The XRD pattern in Figure 2B and Figure S3B shows inorganic moieties with the sharp and strongest peak of SiO2 on 10° for AWSB and about 28° for pristine WSB at 2-theta degree [29]. The biochar attains a well-arranged crystal face, shape, and degree of graphitization after activating. The XRD peak around 28° on pristine WSB shifts to the lower 2-theta degree around 10° on AWSB, indicating the changes in the structural morphology which is in well accordance with SEM micrographs. Hence, the XRD results of pristine WSB and pre-adsorbed AWSB well explained the change in the structural morphology of biochar. The surface of pristine WSB and pre-adsorbed AWSB might be heterogeneous, according to XRD findings.

3.2. Adsorption Performance

3.2.1. Effect of the Initial Metal Concentration

The important factor in the study of contaminant removal from the aqueous media is the removal efficiency of biochar with respect to initial heavy metals concentration. The Cd and Pb concentrations were increased from 2 mg/L to 50 mg/L with a constant AWSB dosage of 2 g. The Figure 3A presented a surge in the removal efficiency with an increase in the concentration of heavy metals until 30 mg/L. AWSB showed maximum removal efficiency (98.3% and 98.7%) for both Cd and Pb at 30 mg/L, respectively. Further increase in the metals concentration reduces the removal efficiency for Cd, which could be due to the saturation of AWSB active sites at higher metal concentrations [31,38]. The adsorption sites after a certain time get to be depleted and reaches an equilibrium which encourages adsorption from aqueous media is not conceivable. Initial metal ion accumulation contributes to the conquest of all metal transfer resistances within solid and liquid, which means a higher probability of collision between Cd, Pb, and active sites of AWSB. At the higher metal concentration of 50 mg/L, the Pb adsorption still maintains at high level (98.2%), whereas the Cd adsorption has decreased to 83.5%, indicating that competitive adsorption of Cd and Pb occurred over the active sites of AWSB.

3.2.2. Effect of Contact Time

For remediation of Cd and Pb from contaminated water using AWSB, contact time also plays a crucial role. The impact of contact time is studied by fixing the usage of AWSB (0.05 g in 20 mL solution), and Cd and Pb concentrations (15 mg/L). As shown in Figure 3B, in the first hour, the removal rate rapidly increases to the maxima and then gradually slows down and attains equilibrium. The Cd and Pb maximum adsorption after 2 h were 1.9 and 5.6 mg/g, respectively. Rapid adsorption of Cd and Pb indicates a fast reaction rate within a short contact time due to the high availability of active sites on AWSB [34]. The efficiency of the material is shown by the accomplishment of the adsorption equilibrium in a small contact time. Small contact time is prioritized in wastewater treatment, as the strength of the procedure proficiency can be improved with less operating time and the total cost of the metal evacuation procedure can be reduced [39]. Hence, for the removal of Cd and Pb, AWSB proved to be a cost and time-effective adsorbent. The quick initial adsorption rate also proposes that adsorbent and adsorbate have high reactivity, thus defining the interaction between the active surface and the metal ions, which is suggested to be chemisorption rather than the slow physisorption process.

3.2.3. Effect of pH

pH is another essential parameter that affects the acid-base property, degree of ionization, and surface charge, which directly affects the ionic strength of metal ions in the solution [40]. The pH of the aqueous media varied between 2 to 12 in this study. The Figure 3C showed that as the pH of the aqueous solution increases from 2 to 8, the Cd and Pb adsorption capacities of AWSB also increases. Maximum Cd and Pb adsorption of 98.0% is obtained by AWSB at pH = 8, while remediation efficiency decreases with further increasing the pH. The poor adsorption efficiency at low pH may be caused by the repulsion between the Cd and Pb on the surface of biochar and acidic functional groups. In contrast, the competition between Cd and Pb and hydronium ions decreases at high pH, and the biochar surface gradually turned to be negatively charged, which provides an active surface for Cd and Pb adsorption. Our findings are in good agreement with the literature study, which reported that there is a direct relation between heavy metal adsorption and pH, i.e., an increase in pH improves the adsorption of heavy metals [41]. Notably, AWSB shows high removal efficiency for Pb as compared to Cd, indicating that Cd adsorption is more sensitive to pH.

3.2.4. Adsorption Isotherms

Adsorption isotherm was studied to evaluate the mechanisms and total adsorption capacity of AWSB. The equilibrium isotherm is the foremost noteworthy phy-chem- method in a solid-liquid framework, which is mostly used to investigate the adsorption behavior [42]. The adsorption equilibrium isotherm is studied using the linear form of Langmuir, Freundlich, and Temkin models. Linear correlation coefficients and constant parameters are elucidated in Table 2. The Langmuir model suits Cd and Pb adsorption on AWSB better than Freundlich and Temkin model at temperatures of 298, 310, and 322 K (Figure 4A–C and Figures S4–S9). The adsorption behavior of AWSB fitted well with the Langmuir model suggests a monolayer adsorption process rather than heterogeneous coverage on homogenous sites [43]. For Cd and Pb adsorption, the Langmuir maximum monolayer adsorption capacity (Qm) value is 9.3 and 0.6 mg/g at 298 K, respectively. Moreover, the RL value for Cd and Pb adsorption with a concentration of 15 ppm at 298 K is 0.940 and 0.953, explaining that both the heavy metals favorably adsorbed on AWSB. Freundlich dimensionless constant n values for Cd (1.949) and Pb (2.865) are greater than 1, which indicates their favorable adsorption on AWSB.

3.2.5. Adsorption Thermodynamics

The thermodynamic analysis is examined to explore the nature and energetic changes in Cd and Pb adsorption over AWSB. The obtained thermodynamic parameters are tabulated in Table 3. Negative ΔG° indicates that Cd and Pb adsorption on AWSB is a spontaneous process. The adsorption becomes more spontaneous and favorable at low temperatures with the decline in ΔG° under promoting temperature. The ΔG° value of 4–10 kJ/mol indicates the existence of Van der Waal forces of interaction, hydrogen bonding from 2–40 kJ/mol, dipole force of interaction from 2–29 kJ/mol and >60 kJ/mol chemical bonding is the dominant force of interaction [34]. In this case, ΔG° ranges from −35.44 to −37.67 kJ/mol for Cd and −32.04 to −34.47 kJ/mol illustrates hydrogen bonding as the major interactive force between heavy metals and AWSB. The graph between lnK and 1/T is plotted to extract the values of ΔH° and ΔS° as shown in Figure 5. Positive ΔS° value (3.49 J/mol K for Cd and 119.26 J/mol K for Pb) expresses the good affinity between AWSB and Cd and Pb with the disordered solid-liquid interface. The disorder-ness between solid-liquid might be because of the discharge of water molecules due to the migration of positive and negative ions generated by the transfer of heavy metals on the surface of AWSB. Compared to Cd, the higher ΔS° for Pb indicates its stronger affinity with AWSB. Negative ΔH° value for Cd (ΔH° = −1.90 kJ/mol) reveals that the adsorption process is exothermic while a positive amount of ΔH° for Pb (ΔH° = 112.42 kJ/mol) illustrates its adsorption on AWSB is endothermic. Generally, ΔH° value for physical adsorption ranges from 2.1 to 20.9 kJ/mol where the chemical adsorption process ranges from 80 to 200 kJ·mol−1, respectively [44], thus ΔH° values for Cd and Pb adsorption well fits in physical adsorption process.

3.2.6. Adsorption Kinetics

Adsorption kinetics governs a crucial role in designing the adsorption process and forecast the solution rejected adsorbate. Adsorption kinetics was examined for the removal of heavy metals based on the adsorption time. Pseudo 1st order and pseudo 2nd order are two renowned equations used on the adsorption data to achieve the kinetic process presented in Figure S10 and Figure 6and the measured kinetic parameters are presented in Table 4. For Cd and Pb, the quantity of correlation coefficient (R2) for pseudo 2nd order is higher than pseudo 1st order, power function (Figure S11), and intraparticle diffusion (Figure S12). Thus, Cd and Pb adsorption on AWSB fits best on pseudo 2nd order kinetics. It is described that for the adsorption process of Cd and Pb, the chemisorption is the rate-controlling mechanism on AWSB by sharing electrons between heavy metals and biochar [41]. The qe experimental and calculated values for pseudo 2nd order are very close to each other expressing that the mechanism of adsorption depends upon the limited rate of chemisorption, such as precipitation and complexation [45]. The maximum Cd and Pb adsorbed on AWSB is 8.9 and 9.0 mg/g at 298 K, respectively. AWSB showed superior adsorption capacity in comparison to other adsorbents in literature (Table 5) [14,30,41,46,47,48,49,50]. The lower R2 value of power function and intraparticle diffusion reveals that Cd and Pb adsorption never progress through power function and intraparticle diffusion model.

3.3. Insights of the Adsorption Mechanism

Generally, the adsorption mechanism of Cd and Pb on biochar governs surface complexation, mineral precipitation, and metal-π interaction. The key method for the removal of heavy metals by biochar is precipitation with minerals. Biochar possesses CaCO3, Ca3(PO4)2, a variety of additional substances and inorganic anions like OH, CO32− and PO43− were released when Cd and Pb contacted the biochar within the aqueous environment, facilitating their precipitation. The precipitates from Pb2+ and inorganic anion reactions can be observed after adsorption in the XRD study according to previous literature. Therefore, a possible mechanism for Cd and Pb adsorption on AWSB could be their surface mineral precipitation [8,51]. The proof of biochar-derived precipitation with minerals is as follows.
(a)
SEM images (Figure 7A) for post-adsorption of AWSB indicated the adsorption of Cd and Pb ions in the internal pores ultimately making the surface area of biochar well-arranged and least porous. The increased carbon content of the biochar due to pyrolysis thereby increases the adsorption area for the metals. The Cd and Pb adsorption on the AWSB surface were confirmed by using SEM-EDX (Figure 7B). The EDX results revealed that the Pb percentage on AWSB of the pre-adsorbed sample was 0% and increased to 21.52% after adsorption (Table 1). Similarly, the percentage of adsorption for Cd on AWSB surface for pre and post adsorption samples was increased from 8.56% to 19.18%. A high percentage of Si, K, and Ca was also observed on biochar surfaces in EDX analysis. These results strongly evidenced and confirmed the adsorption of heavy metals on AWSB surface where Pb results illustrate the strong adsorption on biochar surface. When Cd and Pb are adsorbed, the AWSB surface has certain particles or minerals crystal on it, showing that Cd and Pb related compounds are produced between Cd, Pb, and biochar, according to SEM spectra.
(b)
XRD pattern of post-adsorbed Cd and Pb on AWSB is presented in Figure 8A. Post-adsorption of Cd(NO3)2, Pb(NO3)2, and pH controlling agent (HCl and NaOH), these XRD peaks become broader and stronger at about 20° which corresponds to SiO2 which are in-line with the EDX results [29,33]. The change in the structure of biochar after Cd and Pb adsorption accelerates the precipitation process at the time of adsorption and facilitates the surface precipitation which is an important adsorption mechanism for these metals. The new crystal peak was emerged at 42° after Cd and Pb adsorption on AWSB corresponds to CdCO3 and PbCO3 presence. The shape and characterization of CO32− was consistent with the previous study [52].
(c)
Biochar metal cations (Ca2+, Mg2+, K+) are stored by complexation (e.g., —COOM, —R—O—M with —OH and COOH) or precipitation or by electrostatic attraction. During the adsorption process, Cd and Pb are exchanged through these cations in solution through surface complexation by metal exchange process, co-precipitation, or by an electrostatic exchange. Our findings are also supported by previous literature that the concentration of cations and anions decreased after heavy metal adsorption on biochar [53].
A significant mechanism for metal sorption by biochar is also surface complexation with organic functional groups (OFGs). There are many types of OFGs found in biochar. The OFGs are affected greatly by pyrolytic temperatures. The types and content of OFGs have reduced markedly with increasing pyrolytic temperature. Interaction between Cd/Pb and OFGs occurred in this research due to a higher number of oxygen-containing functional groups, especially at low temperatures. The FTIR spectra (Figure 8B) showed the variation in absorbance peaks responsible for competitive adsorption between the two heavy metals eliminated from the self-contaminated water. The adsorption of Cd and Pb might occur by π-electrons offered by the aromatic ring. The post-adsorption peak for -OH stretching ranges from 3300 to 3500 cm−1 emerged due to water after physio-sorbed adsorption [29,36]. The C—C and C—O stretching vibrations of alcohols, phenols, ether, or ester groups refer to the peak at 1070.5 cm−1 where the peak around 650 cm−1 corresponds to CO32− 36. According to the FTIR analysis, some peaks moved or vanished when Cd and Pb adsorbed onto AWSB, including C—C, C=C, C—O, and —OH. They followed the mechanism of surface complexation that occurred between Cd, Pb, and AWSB. AWSB was partially carbonized at low temperatures and some OFGs were reserved in the AWSB.
The cation-π interaction is another well-known mechanism for metal adsorption by biochars. Zhang et.al., reported that the interaction of cation-π played a significant role during biochar heavy metal sorption [54]. The cyclic aromatic π-system’s π-electrons can act as a donor of the electrons, such donation becomes increased with the increase of the number of associated rings [55]. The peak for C=C groups shifted in the post-adsorption FTIR analysis, which confirms that the cation-π interaction indeed makes contributions to the Cd and Pb adsorption in our experiment.

4. Conclusions

Activated biochar was synthesized by chemical oxidation through HCl as an activating agent. A strong competitive Cd and Pb adsorption efficiency from aqueous media with an adsorption value of 8.9 and 9.0 mg/g was shown by the AWSB. The AWSB has a good porous structure and optimal adsorption was achieved with the concentration of 30 mg/L, pH = 8, and contact time of 2 h. Equilibrium, thermodynamic and kinetics were studied to investigate the feasibility and adsorption mechanism. Equilibrium isotherm studies revealed that all isotherms match well with experimental results, whereas amongst them for Cd and Pb Langmuir isotherm best fitted with highest fitting R2 values of 0.939 and 0.943 at 298 K. The thermodynamic investigation exposed the spontaneous adsorption process by ΔG° values. Cd governs exothermic where the endothermic pathway was followed by Pb which was revealed by ΔH° values, respectively. As well, the entropy values reflected a decent affinity of AWSB for Pb and showed a decreased affinity for Cd. Furthermore, thermodynamic investigations also proved that adsorption was primarily a result of Van der Waal’s force interaction between heavy metals and AWSB. The adsorption kinetics confirmed the well-fitting of pseudo 2nd order kinetics. The mechanism regulating the AWSB for the remediation of Cd and Pb is minerals precipitation, surface complexation, and cation-π interaction.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/su13168817/s1, Figure S1: Scanning electron micrographs (SEM) of pristine WSB and pre-adsorbed AWSB (pyrolyzed at 673 K), Figure S2: EDX mapping of the pristine WSB, Figure S3: (A) FTIR spectra and (B) XRD pattern of the pristine WSB, Figure S4: Langmuir isotherm for the adsorption of Cd(II) and Pb(II) AWSB at 310K, Figure S5: Langmuir isotherm for the adsorption of Cd(II) and Pb(II) on AWSB at 322K, Figure S6: Freundlich isotherm for the adsorption of Cd(II) and Pb(II) on AWSB at 310K, Figure S7: Freundlich isotherm for the adsorption of Cd(II) and Pb(II) on AWSB at 322K, Figure S8: Temkin isotherm for the adsorption of Cd(II) and Pb(II) on AWSB at 310K, Figure S9: Temkin isotherm for the adsorption of Cd(II) and Pb(II) on AWSB at 322K, Figure S10: Pseudo 1st order fittings of Cd and Pb adsorption data, Figure S11. Power function fittings of Cd(II) and Pb(II) adsorption data for AWSB, Figure S12. Intraparticle diffusion fittings of Cd(II) and Pb(II) adsorption data for AWSB.

Author Contributions

Conceptualization, M.N.K. and H.U.; methodology, M.N.K., H.U.; software, M.N.K. and W.A.; validation, M.N.K. and H.U.; formal analysis, M.N.K. and S.N.; investigation, M.N.K.; resources, M.N.K. and S.N.; data curation, M.N.K.; writing—original draft preparation, M.N.K.; writing—review and editing, M.N.K., H.U., J.U., Y.H., J.D.; visualization, M.N.K.; supervision, H.U. and J.U.; project administration, H.U. and J.U. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We thank the Soil Science & Water Conservation Department, PMAS Arid Agriculture University, Pakistan for providing facilities of biochar characterizations. Authors also extend their appreciation to the deanship of scientific research at King Khalid University for supporting this work through prolific research group No. GRP/92/41.
Conflicts of interest: There are no conflicts of interest.

References

  1. Li, A.Y.; Deng, H.; Jiang, Y.H.; Ye, C.H.; Yu, B.G.; Zhou, X.L. Superefficient Removal of Heavy Metals from Wastewater by Mg- Loaded Biochars: Adsorption Characteristics and Removal Mechanisms. Langmuir 2020, 36, 9160–9174. [Google Scholar] [CrossRef]
  2. Chai, J.; Au, P.; Mubarak, N.M.; Khalid, M.; Ng, W.P. Adsorption of Heavy Metal from Industrial Wastewater onto Low-Cost Malaysian Kaolin Clay–Based Adsorbent. Environ. Sci. Pollut. Res. 2020, 27, 13949–13962. [Google Scholar] [CrossRef]
  3. Yu, W.; Hu, J.; Yu, Y.; Ma, D.; Gong, W.; Qiu, H.; Hu, Z.; Gao, H. Facile Preparation of Sulfonated Biochar for Highly Efficient Removal of Toxic Pb(II) and Cd(II) from Wastewater. Sci. Total Environ. 2021, 750, 141545. [Google Scholar] [CrossRef]
  4. Hassan, A.; Khan, A.; Nawaz, I.; Yousaf, S.; Sabir, A.; Iqbal, M. Soil Amendments Enhanced the Growth of Nicotiana Alata L. and Petunia Hydrida L by Stabilizing Heavy Metals from Wastewater. J. Environ. Manag. 2019, 242, 46–55. [Google Scholar] [CrossRef]
  5. Poo, K.; Son, E.; Chang, J.; Ren, X.; Choi, Y.; Chae, K. Biochars Derived from Wasted Marine Macro-Algae (Saccharina Japonica and Sargassum Fusiforme) and Their Potential for Heavy Metal Removal in Aqueous Solution. J. Environ. Manag. 2018, 206, 364–372. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, C.; Wu, T.; Hsu, P.; Xie, J.; Zhao, J.; Liu, K.; Sun, J.; Xu, J.; Tang, J.; Ye, Z.; et al. Direct/Alternating Current Electrochemical Method for Removing and Recovering Heavy Metal from Water Using Graphene Oxide Electrode. ACS Nano 2019, 13, 6431–6437. [Google Scholar] [CrossRef] [PubMed]
  7. Fan, Z.; Zhang, Q. Investigating the Sorption Behavior of Cadmium from Aqueous Solution by Potassium Permanganate-Modified Biochar : Quantify Mechanism and Evaluate the Modification Method. Environ. Sci. Pollut. Res. 2018, 25, 8330–8339. [Google Scholar] [CrossRef] [PubMed]
  8. Liu, L.; Fan, S. Removal of Cadmium in Aqueous Solution Using Wheat Straw Biochar : Effect of Minerals and Mechanism. Environ. Sci. Pollut. Res. 2018, 25, 8688–8700. [Google Scholar] [CrossRef] [PubMed]
  9. Pino, L.; Vargas, C.; Schwarz, A.; Borquez, R. Influence of Operating Conditions on the Removal of Metals and Sulfate from Copper Acid Mine Drainage by Nanofiltration. Chem. Eng. J. 2018, 345, 114–125. [Google Scholar] [CrossRef]
  10. Gay, A.; Pal, P.; Banat, F.; Abu, M. Separation and Purification Technology Enhanced Removal of Mixed Metal Ions from Aqueous Solutions Using Flotation by Colloidal Gas Aphrons Stabilized with Sodium Alginate. Sep. Purif. Technol. 2018, 202, 103–110. [Google Scholar] [CrossRef]
  11. Yang, X.; Zhang, S.; Ju, M. Preparation and Modification of Biochar Materials and Their Application in Soil Remediation. Appl. Sci. 2019, 9, 1365. [Google Scholar] [CrossRef] [Green Version]
  12. Technol, W.S.; Heidelberg, S.B. Biosorption (Removing) of Cd(II), Cu(II) and Methylene Blue Using Biochar Produced by Different Pyrolysis Conditions of Beech and Spruce Sawdust. Wood Sci. Technol. 2017, 51, 1321–1338. [Google Scholar] [CrossRef]
  13. Zhang, S.; Zhang, H.; Cai, J.; Zhang, X.; Zhang, J.; Shao, J. Evaluation and Prediction of Cadmium Removal from Aqueous Solution by Phosphate-Modified Activated Bamboo Biochar. Energy Fuels 2018, 23, 4467–4477. [Google Scholar] [CrossRef]
  14. Ni, B.; Huang, Q.; Wang, C.; Ni, T.; Sun, J.; Wei, W. Competitive Adsorption of Heavy Metals in Aqueous Solution onto Biochar Derived from Anaerobically Digested Sludge. Chemosphere 2019, 219, 351–357. [Google Scholar] [CrossRef]
  15. Li, C.; Zhang, L.; Gao, Y.; Li, A. Facile Synthesis of Nano ZnO/ZnS Modified Biochar by Directly Pyrolyzing of Zinc Contaminated Corn Stover for Pb(II), Cu(II) and Cr(VI) Removals. Waste Manag. 2018, 79, 625–637. [Google Scholar] [CrossRef] [PubMed]
  16. Nartey, O.D.; Zhao, B. Biochar Preparation, Characterization, and Adsorptive Capacity and Its Effect on Bioavailability of Contaminants: An Overview. Adv. Mater. Sci. Eng. 2014, 2014, 12. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, S.; Gao, B.; Zimmerman, A.R.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K.W. Chemosphere Physicochemical and Sorptive Properties of Biochars Derived from Woody and Herbaceous Biomass. Chemosphere 2015, 134, 257–262. [Google Scholar] [CrossRef] [PubMed]
  18. Govumoni, S.P.; Koti, S.; Kothagouni, S.Y.; Venkateshwar, S. Evaluation of Pretreatment Methods for Enzymatic Saccharification of Wheat Straw for Bioethanol Production. Carbohydr. Polym. 2013, 91, 646–650. [Google Scholar] [CrossRef]
  19. Rodriguez-gomez, D.; Lehmann, L.; Schultz-jensen, N.; Bjerre, A.B.; Hobley, T.J. Examining the Potential of Plasma-Assisted Pretreated Wheat Straw for Enzyme Production by Trichoderma Reesei. Appl. Biochem. Biotechnol. 2012, 166, 2051–2063. [Google Scholar] [CrossRef] [PubMed]
  20. Stasi, C. Di; Alvira, D.; Greco, G.; González, B.; Manyà, J.J. Physically Activated Wheat Straw-Derived Biochar for Biomass Pyrolysis Vapors Upgrading with High Resistance against Coke Deactivation. Fuel 2019, 255, 115807. [Google Scholar] [CrossRef]
  21. Anto, S.; Sudhakar, M.P.; Shan, T.; Samuel, M.S.; Mathimani, T.; Brindhadevi, K.; Pugazhendhi, A. Activation Strategies for Biochar to Use as an Efficient Catalyst in Various Applications. Fuel 2021, 285, 119205. [Google Scholar] [CrossRef]
  22. Plaza, M.G.; González, A.S.; Pis, J.J.; Rubiera, F.; Pevida, C. Production of Microporous Biochars by Single-Step Oxidation: Effect of Activation Conditions on CO2 Capture. Appl. Energy 2014, 114, 551–562. [Google Scholar] [CrossRef] [Green Version]
  23. Liu, H.; Xu, F.; Xie, Y.; Wang, C.; Zhang, A.; Li, L.; Xu, H. Science of the Total Environment Effect of Modi Fi Ed Coconut Shell Biochar on Availability of Heavy Metals and Biochemical Characteristics of Soil in Multiple Heavy Metals Contaminated Soil. Sci. Total Environ. 2018, 645, 702–709. [Google Scholar] [CrossRef]
  24. Zhi-liang, C.; Jian-qiang, Z.; Ling, H.; Zhi-hui, Y.; Zhao-jun, L.I.; Min-chao, L.I.U. Removal of Cd and Pb with Biochar Made from Dairy Manure at Low Temperature. J. Integr. Agric. 2019, 18, 201–210. [Google Scholar] [CrossRef] [Green Version]
  25. Ellis, N.; Dehkhoda, A.M. A Novel Method to Tailor the Porous Structure of KOH-Activated Biochar and Its Application in Capacitive Deionization and Energy Storage. Biomass Bioenergy 2016, 87, 107–121. [Google Scholar] [CrossRef] [Green Version]
  26. Chen, T.; Luo, L.; Deng, S.; Shi, G.; Zhang, S. Sorption of Tetracycline on H3PO4 Modified Biochar Derived from Rice Straw and Swine Manure. Bioresour. Technol. 2018, 267, 431–437. [Google Scholar] [CrossRef] [PubMed]
  27. Sun, K.; Huang, Q.; Chi, Y.; Yan, J. Effect of ZnCl2-Activated Biochar on Catalytic Pyrolysis of Mixed Waste Plastics for Producing Aromatic-Enriched Oil. Waste Manag. 2018, 81, 128–137. [Google Scholar] [CrossRef]
  28. Boshir, M.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Chen, M. Bioresource Technology 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]
  29. Naeem, M.A.; Imran, M.; Amjad, M.; Abbas, G.; Tahir, M.; Murtaza, B.; Zakir, A.; Shahid, M.; Bulgariu, L.; Ahmad, I. Batch and Column Scale Removal of Cadmium from Water Using Raw and Acid Activated Wheat. Water 2019, 11, 1438. [Google Scholar] [CrossRef] [Green Version]
  30. 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] [PubMed]
  31. Batool, S.; Idrees, M.; Al-wabel, M.I.; Ahmad, M.; Hina, K. Sorption of Cr(III) from Aqueous Media via Naturally Functionalized Microporous Biochar : Mechanistic Study. Microchem. J. 2019, 144, 242–253. [Google Scholar] [CrossRef]
  32. Chen, Y.; Ho, S.; Wang, D.; Wei, Z.; Chang, J.; Ren, N. Lead Removal by a Magnetic Biochar Derived from Persulfate-ZVI Treated Sludge Together with One-Pot Pyrolysis. Bioresour. Technol. 2018, 247, 463–470. [Google Scholar] [CrossRef]
  33. Wu, Q.; Xian, Y.; He, Z.; Zhang, Q.; Wu, J.; Yang, G.; Zhang, X. Adsorption Characteristics of Pb(II) Using Biochar Derived from Spent Mushroom Substrate. Sci. Rep. 2019, 9, 15999. [Google Scholar] [CrossRef] [Green Version]
  34. Idrees, M.; Batool, S.; Ullah, H.; Hussain, Q.; Al-wabel, M.I.; Ahmad, M.; Hussain, A.; Riaz, M.; Sik, Y.; Kong, J. Adsorption and Thermodynamic Mechanisms of Manganese Removal from Aqueous Media by Biowaste-Derived Biochars. J. Mol. Liq. 2018, 266, 373–380. [Google Scholar] [CrossRef]
  35. Hamza, I.A.A.; Martincigh, B.S.; Ngila, J.C.; Nyamori, V.O. Adsorption Studies of Aqueous Pb(II) onto a Sugarcane Bagasse/Multi-Walled Carbon Nanotube Composite. Phys. Chem. Earth 2013, 66, 157–166. [Google Scholar] [CrossRef]
  36. Shi, L.; Zhang, G.; Wei, D.; Yan, T.; Xue, X.; Shi, S.; Wei, Q. Preparation and Utilization of Anaerobic Granular Sludge-Based Biochar for the Adsorption of Methylene Blue from Aqueous Solutions. J. Mol. Liq. 2014, 198, 334–340. [Google Scholar] [CrossRef]
  37. Bashir, S.; Zhu, J.; Fu, Q.; Hu, H. Comparing the Adsorption Mechanism of Cd by Rice Straw Pristine and KOH-Modified Biochar. Environ. Sci. Pollut. Res. 2018, 25, 11875–11883. [Google Scholar] [CrossRef] [PubMed]
  38. Dong, X.; Ma, L.Q.; Li, Y. Characteristics and Mechanisms of Hexavalent Chromium Removal by Biochar from Sugar Beet Tailing. J. Hazard. Mater. 2011, 190, 909–915. [Google Scholar] [CrossRef] [PubMed]
  39. Logam, P.; Kumbahan, A. Removal of Heavy Metals from Wastewater Using Date Palm as a Biosorbent: A Comparative Review. Sains Malays. 2018, 47, 35–49. [Google Scholar]
  40. Forghani, M.; Azizi, A.; Jamal, M.; Asadi, L. Adsorption of Lead(II) and Chromium(VI) from Aqueous Environment onto Metal-Organic Framework MIL-100 (Fe): Synthesis, Kinetics, Equilibrium and Thermodynamics. J. Solid State Chem. 2020, 291, 121636. [Google Scholar] [CrossRef]
  41. Usman, A.; Sallam, A.; Zhang, M.; Vithanage, M. Sorption Process of Date Palm Biochar for Aqueous Cd(II) Removal : Efficiency and Mechanisms. Water Air Soil Pollut. 2016, 227, 449. [Google Scholar] [CrossRef]
  42. El-yazeed, W.S.A.; El-reash, Y.G.A.; Elatwy, L.A.; Ahmed, A.I. Facile Fabrication of Bimetallic Fe-Mg MOF for the Synthesis of Xanthenes and Removal of Heavy Metal. RSC Adv. 2020, 10, 9693–9703. [Google Scholar] [CrossRef] [Green Version]
  43. Coşkun, Y.İ.; Aksuner, N.; Yanik, J. Sandpaper Wastes as Adsorbent for the Removal of Brilliant Green and Malachite Green Dye. Acta Chimica Slovenica 2019, 66, 402–413. [Google Scholar] [CrossRef]
  44. Mittal, H.; Babu, R.; Dabbawala, A.A.; Alhassan, S.M. Low-Temperature Synthesis of Magnetic Carbonaceous Materials Coated with Nanosilica for Rapid Adsorption of Methylene Blue. ACS Omega 2020, 5, 6100–6112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Wne, R.; Leahy, J.J.; Hayes, M.H.B.; Kwapin, W. Kinetic and Adsorptive Characterization of Biochar in Metal Ions Removal. Chem. Eng. J. 2012, 197, 295–305. [Google Scholar] [CrossRef]
  46. Yanagisawa, H.; Matsumoto, Y.; Machida, M. Adsorption of Zn(II) and Cd(II) Ions onto Magnesium and Activated Carbon Composite in Aqueous Solution. Appl. Surf. Sci. 2010, 256, 1619–1623. [Google Scholar] [CrossRef]
  47. Xu, D.; Tan, X.; Chen, C.; Wang, X. Removal of Pb(II) from Aqueous Solution by Oxidized Multiwalled Carbon Nanotubes. J. Hazard. Mater. 2008, 154, 407–416. [Google Scholar] [CrossRef]
  48. Mohan, D.; Pittman, C.U.; Bricka, M.; Smith, F.; Yancey, B.; Mohammad, J.; Steele, P.H.; Alexandre-Franco, M.F.; Gómez-Serrano, V.; Gong, H. Sorption of Arsenic, Cadmium, and Lead by Chars Produced from Fast Pyrolysis of Wood and Bark during Bio-Oil Production. J. Colloid Interface Sci. 2007, 310, 57–73. [Google Scholar] [CrossRef]
  49. Liu, Z.; Zhang, F.S. Removal of Lead from Water Using Biochars Prepared from Hydrothermal Liquefaction of Biomass. J. Hazard. Mater. 2009, 167, 933–939. [Google Scholar] [CrossRef]
  50. Park, J.; Sik, Y.; Kim, S.; Cho, J.; Heo, J.; Delaune, R.D.; Seo, D. Competitive Adsorption of Heavy Metals onto Sesame Straw Biochar in Aqueous Solutions. Chemosphere 2016, 142, 77–83. [Google Scholar] [CrossRef]
  51. Chi, T.; Zuo, J.; Liu, F. Performance and Mechanism for Cadmium and Lead Adsorption from Water and Soil by Corn Straw Biochar. Front. Environ. Sci. Eng. 2017, 11, 15. [Google Scholar] [CrossRef]
  52. Shan, R.; Yan, L.; Yang, K.; Hao, Y.; Du, B. Adsorption of Cd(II) by Mg–Al–CO3- and Magnetic Fe3O4/Mg–Al–CO3-Layered Double Hydroxides: Kinetic, Isothermal, Thermodynamic and Mechanistic Studies. J. Hazard. Mater. 2015, 299, 42–49. [Google Scholar] [CrossRef]
  53. Cui, X.; Fang, S.; Yao, Y.; Li, T.; Ni, Q.; Yang, X.; He, Z. Science of the Total Environment Potential Mechanisms of Cadmium Removal from Aqueous Solution by Canna Indica Derived Biochar. Sci. Total Environ. 2016, 562, 517–525. [Google Scholar] [CrossRef]
  54. Zhang, T.; Zhu, X.; Shi, L.; Li, J.; Li, S.; Lü, J.; Li, Y. Efficient Removal of Lead from Solution by Celery-Derived Biochars Rich in Alkaline Minerals. Bioresour. Technol. 2017, 235, 185–192. [Google Scholar] [CrossRef]
  55. 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]
Figure 1. (A) SEM and (B) EDX mapping of the pre-adsorbed AWSB.
Figure 1. (A) SEM and (B) EDX mapping of the pre-adsorbed AWSB.
Sustainability 13 08817 g001
Figure 2. (A) FTIR spectra and (B) XRD pattern of the pre-adsorbed AWSB.
Figure 2. (A) FTIR spectra and (B) XRD pattern of the pre-adsorbed AWSB.
Sustainability 13 08817 g002
Figure 3. (A). Effects of initial metal concentration on Cd(II) and Pb(II) removal efficiency of AWSB. (B). Effects of contact time on Cd(II) and Pb(II) removal efficiency of AWSB. (C). Effects of pH on Cd(II) and Pb(II) removal efficiency of AWSB.
Figure 3. (A). Effects of initial metal concentration on Cd(II) and Pb(II) removal efficiency of AWSB. (B). Effects of contact time on Cd(II) and Pb(II) removal efficiency of AWSB. (C). Effects of pH on Cd(II) and Pb(II) removal efficiency of AWSB.
Sustainability 13 08817 g003
Figure 4. (A) Langmuir, (B) Freundlich and (C) Temkin Isotherms for the adsorption of Cd(II) and Pb(II) on AWSB at 298 K.
Figure 4. (A) Langmuir, (B) Freundlich and (C) Temkin Isotherms for the adsorption of Cd(II) and Pb(II) on AWSB at 298 K.
Sustainability 13 08817 g004
Figure 5. Thermodynamic constraints for Cd(II) and Pb(II) adsorption: plot of lnK vs. 1/T for AWSB.
Figure 5. Thermodynamic constraints for Cd(II) and Pb(II) adsorption: plot of lnK vs. 1/T for AWSB.
Sustainability 13 08817 g005
Figure 6. Pseudo 2nd order fittings of Cd(II) and Pb(II) adsorption data for AWSB.
Figure 6. Pseudo 2nd order fittings of Cd(II) and Pb(II) adsorption data for AWSB.
Sustainability 13 08817 g006
Figure 7. (A) SEM and (B) EDX mapping of the AWSB post-adsorption of 4 ppm Cd(II) and Pb(II).
Figure 7. (A) SEM and (B) EDX mapping of the AWSB post-adsorption of 4 ppm Cd(II) and Pb(II).
Sustainability 13 08817 g007
Figure 8. (A) XRD spectra of post-adsorbed Cd(II) and Pb(II) on activated wheat straw biochar, (B) FTIR spectra of post-adsorbed Cd(II) and Pb(II) on activated wheat straw biochar.
Figure 8. (A) XRD spectra of post-adsorbed Cd(II) and Pb(II) on activated wheat straw biochar, (B) FTIR spectra of post-adsorbed Cd(II) and Pb(II) on activated wheat straw biochar.
Sustainability 13 08817 g008
Table 1. Elemental analysis of the pristine WSB and AWSB and post-adsorbed AWSB (2 ppm).
Table 1. Elemental analysis of the pristine WSB and AWSB and post-adsorbed AWSB (2 ppm).
ElementWSB AWSBPost-Adsorption
Weight %Atomic %Weight %Atomic %Weight %Atomic %
K35.2030.5335.8437.00--
Mg--5.749.54--
Si42.7551.6232.7447.05--
Ca20.8517.65--8.629.09
Al----50.6779.32
Cd--8.563.0719.187.21
Pb1.210.20--21.524.39
Table 2. Constant parameters and correlation coefficients of isothermal models for Cd(II) and Pb(II) adsorption onto AWSB.
Table 2. Constant parameters and correlation coefficients of isothermal models for Cd(II) and Pb(II) adsorption onto AWSB.
Heavy MetalTemperature
(K)
Langmuir ModelFreundlich ModelTemkin Model
KL
(L/mg)
Qm
(mg/g)
R2KF [(mg/g) (L/g)n]nR2B
(J/mol)
A
(L/g)
R2
Cd2981.0489.2850.9390.7251.9490.900318.10.1880.811
3101.1039.0090.9280.7021.9540.913335.80.1760.807
3220.6353.7240.9110.7051.9540.909348.60.1770.804
Pb2980.7360.6210.9430.4982.8650.930169.90.2690.923
3100.1610.1910.9350.5132.8840.924179.60.2670.903
3220.0280.1720.9260.5192.8950.919196.50.2690.815
Table 3. Thermodynamic parameters of Cd(II) and Pb(II) adsorption on AWSB.
Table 3. Thermodynamic parameters of Cd(II) and Pb(II) adsorption on AWSB.
Heavy MetalTemperature (K)ΔG° (kJ/mol)ΔH° (kJ/mol)ΔS° (J/mol K)
Cd298−35.44−1.903.49
310−36.46
322−37.67
Pb298−32.04112.42119.26
310−33.28
322−34.47
Table 4. Kinetic parameters of Cd(II) and Pb(II)adsorption on activated wheat straw biochar at 298K.
Table 4. Kinetic parameters of Cd(II) and Pb(II)adsorption on activated wheat straw biochar at 298K.
AdsorbentKinetic ModelsParametersCdPb
AWSBPseudo 1st orderqe (Exp) mg/g6.0646.921
qe (Cal) mg/g7.6447.872
K10.0050.004
h0.2680.267
R20.9830.993
Pseudo 2nd orderqe (Exp) mg/g7.538.142
qe (Cal) mg/g8.859.03
K20.0040.003
h0.2770.272
R20.9840.996
Power functionkf−0.233−0.090
b1.5301.893
R20.7820.839
Intraparticle diffusionkdif0.6200.988
c0.2690.262
R20.8140.839
Table 5. Comparison of efficiency of adsorbents for remediation of Cd(II) and Pb(II) from polluted water.
Table 5. Comparison of efficiency of adsorbents for remediation of Cd(II) and Pb(II) from polluted water.
AdsorbentsAdsorbentsQuantity Adsorbed (mg/g)Reference
CdOut-grassed magnesium composite (Mg-AC)0.094Yanagisawa et al. [42]
BC-3004.41Usman et al. [48]
Dairy manure0.28Xu et al. [49]
Oak0.05Mohan et al. [50]
Pine0.01Mohan et al. [50]
Anaerobically digested sludge
Adsorption
0.44Ni et al. [51]
AWSB8.85Present work
PbOxidized MWCNTs Bagasse2.06Xu et al. [49]
Digested whole sugar beet0.20Inyang et al. [52]
Rice husk0.01Liu and Zhang [53]
Sesame straw0.49Park et al. [54]
Pinewood0.02Liu and Zhang [53]
Anaerobically digested sludge
Adsorption
0.61Ni et al. [51]
AWSB9.06Present work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, M.N.; Ullah, H.; Naeem, S.; Uddin, J.; Hamid, Y.; Ahmad, W.; Ding, J. Remediation of Emerging Heavy Metals from Water Using Natural Adsorbent: Adsorption Performance and Mechanistic Insights. Sustainability 2021, 13, 8817. https://doi.org/10.3390/su13168817

AMA Style

Khan MN, Ullah H, Naeem S, Uddin J, Hamid Y, Ahmad W, Ding J. Remediation of Emerging Heavy Metals from Water Using Natural Adsorbent: Adsorption Performance and Mechanistic Insights. Sustainability. 2021; 13(16):8817. https://doi.org/10.3390/su13168817

Chicago/Turabian Style

Khan, Mehak Nawaz, Hidayat Ullah, Sundas Naeem, Jalal Uddin, Yasir Hamid, Waqar Ahmad, and Jia Ding. 2021. "Remediation of Emerging Heavy Metals from Water Using Natural Adsorbent: Adsorption Performance and Mechanistic Insights" Sustainability 13, no. 16: 8817. https://doi.org/10.3390/su13168817

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