Next Article in Journal
SBMLWebApp: Web-Based Simulation, Steady-State Analysis, and Parameter Estimation of Systems Biology Models
Previous Article in Journal
Alkaline Reduced Water Attenuates Oxidative Stress-Induced Mitochondrial Dysfunction and Innate Immune Response Triggered by Intestinal Epithelial Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Single and Binary Adsorption Behaviour and Mechanisms of Cd2+, Cu2+ and Ni2+ onto Modified Biochar in Aqueous Solutions

1
College of Environment, Zhejiang University of Technology, Hangzhou 310032, China
2
Shandong Provincial Key Laboratory of Applied Microbiology, Ecology Institute, Shandong Academy of Sciences, Qilu University of Technology, Jinan 250103, China
3
Global Centre for Environmental Remediation, College of Engineering, Science and Environment, University of Newcastle, Callaghan, NSW 2308, Australia
*
Authors to whom correspondence should be addressed.
Submission received: 1 September 2021 / Revised: 2 October 2021 / Accepted: 11 October 2021 / Published: 14 October 2021
(This article belongs to the Section Materials Processes)

Abstract

:
The chitosan–EDTA modified magnetic biochar (E–CMBC) was successfully used as a novel adsorbent to remove heavy metals. The adsorption behaviour and mechanisms of E–CMBC to Cd2+, Cu2+ and Ni2+ were performed in single and binary system in aqueous solutions. In single–metal system, the adsorption process of Cd2+, Cu2+ and Ni2+ on E–CMBC fitted well with the Avrami fractional–order kinetics model and the Langmuir isotherm model. The measured maximum adsorption capacities were 61.08 mg g−1, 48.36 mg g1 and 41.17 mg g1 for Cd2+, Cu2+ and Ni2+, respectively. In binary–metal system, coexisting ions have obvious competitive adsorption behaviour on E–CMBC when the concentration of heavy meal beyond 20 mg L−1. The maximum adsorption capacities of the heavy metals were found to be lower than that in single–metal system. The order of the competitive adsorption ability was Cu2+ > Ni2+ > Cd2+. Interestingly, in Cd2+–Cu2+ system the earlier adsorbed Cd2+ could be completely replaced by Cu2+ from the solution. Different competitive adsorption ability of those heavy metal were due to the characteristics of heavy metal and resultant affinity of the adsorption sites on E–CMBC. The adsorption mechanism indicated that chemical adsorption played a dominating role. Therefore, E–CMBC could be a potential adsorbent for wastewater treatment.

Graphical Abstract

1. Introduction

With the rapid development of industry, various organic and inorganic industrial pollutants are discharged into the environment, leading to a series of environmental and ecological problems [1,2,3]. Among them, heavy metal pollution in wastewater has attracted widespread attention for its mobility, non–degradability and toxicity in aquatic environment [4]. Heavy metal ions (such as Cu2+, Cd2+, Ni2+) can accumulate in human organisms, which will have serious impact on human health when its levels exceed the permissible limits [5]. Numerous treatment technologies, such as chemical precipitation [6], membrane filtration [7], electrochemical treatment [8], ion exchange [9], biological treatment [10], have been applied to remove the heavy metal ions from wastewater. However, there still exists many defects in each technology, such as low adsorption efficiency, low selectivity, high cost, and secondary pollution, which restrain its large–scale application [11,12]. Currently, adsorption is considered one of the best choice due to its easy operation, low cost and high efficiency [13]. Therefore, it is particularly important to develop a cheap, efficient and widely available adsorbent.
Biochar as a carbon–rich adsorbent has attracted more attention due to its unique properties, such as high surface area, high porosity and cation exchange capacity [14]. Biochar has been used widely as a multifunctional material in environmental remediation, soil reclamation, and energy recovery [15,16]. The adsorption performance of biochar for heavy metals is greatly affected by feedstock types (pinewood, dairy manure, rice husk, etc.) and pyrolysis conditions (temperature, heating rate, residence time, etc.). However, the relatively low heavy metal adsorption ability and difficulty to separate from aqueous solution are regarded as the major disadvantages which limit its application in the treatment of heavy metal polluted wastewater [17]. Acid/base treatment, surface oxidization, chemical graft, physical modification, and magnetic modification, are the main methods to modify and enhance the adsorption performance of biochar [18]. In particular, chemical graft receives more attention as it can add different functional groups on biochar to meet the individual requirements of environmental restoration [19].
In recent years, plenty of researches have confirmed chemical grafted biochar have promising adsorption properties for heavy metals. Li et al. [20] synthesized an aminothiourea chitosan modified magnetic biochar composite which had higher affinity for Cd2+ than the raw biochar. Further, the composite adsorbent presented an excellent adsorption efficiency for other heavy metal ions from real mining water. Lv et al. [21] found that the maximum adsorption capacities for Cu2+ increased from 6.85 mg g1 to 42.19 mg g1 by the EDTA–-functionalized bamboo activated carbon. Sajjadi et al. [22] developed an urea functionalization of ultrasound–treated biochar and investigated its capability in adsorbing Ni2+. It was found that the modified biochar could remove more Ni2+ in six hours than that using the raw biochar in twelve hours. However, most of the present research focused on the adsorption behaviour of heavy metals by biochar in a single–metal system [23,24,25]. In fact, wastewater often contains more than one type of heavy metal, and the coexist ions will influence the adsorption of target ions by biochar [26]. However, studies focused on the adsorption of biochar in multi–metal systems are still lacking. Deng et al. [27] have synthesized the chitosan–pyromellitic dianhydride modified biochar to remove heavy metal ions (Cd2+, Cu2+ and Pb2+) from mixed–metal aqueous solutions. It focused on the comparison of adsorption capacity and the selectivity sequences of the three metal ions while the competitive absorption behaviour and mechanisms of heavy metals in binary–metal systems is still not clear.
Our previous study found that chitosan–EDTA modified magnetic biochar (E–CMBC) had an excellent adsorption performance for Pb2+ in the single–metal system [28]. However, its efficiency to remove Cd2+, Cu2+ and Ni2+ from single–metal (non–competitive) and binary–metal (competitive) systems are yet to be investigated. The purposes of this study were: (1) to investigate the adsorption capacities of Cd2+, Cu2+ and Ni2+ on E–CMBC in single–metal systems; (2) to evaluate the competitive adsorption behaviour of Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal systems; (3) to elucidate the adsorption mechanisms of Cd2+, Cu2+ and Ni2+ on E–CMBC.

2. Materials and Methods

2.1. Materials

The raw biomass measurements of the peanut shells were obtained from an agricultural field (Taian, Shandong, China). The chemicals used in the present research, including FeCl3·6H2O, copper nitrate trihydrate (Cu(NO3)2·3H2O), cadmium nitrate tetrahydrate (Cd(NO3)2·4H2O), nickel nitrate hexahydrate (Ni(NO3)2·6H2O), 1–ethyl–3–(3–dimethylaminopropyl) carbodiimide hydrochloride (EDAC), and ethylenediaminetetraacetic acid (EDTA), acetic acid, chitosan, NaOH, HCl, were purchased from Aladdin Reagent Co. Ltd. Shanghai, China. All of these chemicals were analytically grade without extra treatment. De–ionized water was used to prepared solutions in all experiments.

2.2. Preparation and Characterization of Modified Biochar

The milled peanut shells were immersed into FeCl3 solution (1 mol L−1). After 24 h, the immersed biomass was oven–dried, then pyrolyzed at 500 °C for 2 h in N2 atmosphere. The biochar obtained were milled and added to the acetic acid solution of chitosan (1%, g v−1) and reacted for 2 h, then the mixture solution was adjusted to pH 10 by adding 1% NaOH solution. After stirring overnight, the filter residue was washed with plenty of de–ionized water, oven–dried, and milled. Equal quality of the sample and EDTA (1.0 g) were stirred at 60 °C in de–ionized water for 4 h. Then NaOH and EDAC were added when the reacted solution cooled down to 40 °C. The solution was stirred for 2 h and then naturally cool down to room temperature and keep the reaction overnight. The filtered solid composite was washed to be neutral, and oven–dried and sieved through a 100–mesh, which was referred to as E–CMBC.
The E–CMBC sample was characterized by SEM (Zeiss, Oberkochen, Germany), BET (Quantachrome, Boynton Beach, FL USA), XPS (Thermo, Waltham, MA USA), FTIR (Thermo, Waltham, MA USA), XRD (Rigaku, Tokyo, Japan),VSM (Lake Shore, Dunkirk, NY, USA), and Zeta potential (Malvern, Worcestershire, UK). Detailed synthesis process and physico–chemical characteristics of the E–CMBC had been reported in our previous study [28].

2.3. Batch Adsorption Experiments

The standard stock solutions (1000 mg L−1) of Cd2+, Cu2+ and Ni2+ ions were prepared by dissolving the exact amounts of respective salts. The standard solutions were diluted to a specific concentration by de–ionized water when used. Both single–metal and binary–metals batch adsorption experiments of Cd2+, Cu2+ and Ni2+ were conducted. In single–metal systems, batch experiments were performed in 50 mL conical flask by mixing 25 mg E–CMBC with 25 mL heavy metal solution. The effect of initial pH on the adsorption capacity of E–CMBC were investigated in the range of 2–6. The pH of the solution was adjusted by adding negligible volumes of either HCl or NaOH. Adsorption kinetics experiments were examined at different time intervals (5 min–1440 min) with heavy metals concentration of 100 mg L1 at pH 6. Adsorption isotherms were determined within different initial concentrations of the heavy metals (10 mg L1–500 mg L1) for 24 h. In binary–metal systems, the initial concentration of the two metals is the same. Batch experiments were conducted with the same procedure of single–metal systems. Kinetics experiments were conducted with the same time intervals as the experiment in single–metal systems. Adsorption isotherms were determined with initial concentrations of the heavy metals varying from 10 mg L1 to 200 mg L1.
After adsorption, the residual heavy metal was detected by atomic absorption spectrometer (Agilent 240FS AA). All adsorption experiments were conducted in triplicates. The adsorption amount of metal qe (mg g1) were evaluated using the following equation:
q e = c 0 c e × V m
where c0 (mg L1) and ce (mg L1) are the initial concentration and the concentration at equilibrium, respectively; m (g) and V (L) are the weight of E–CMBC and the volume of metal solution, respectively.

3. Results and Discussion

3.1. Adsorption of Cd2+, Cu2+ and Ni2+ in the Single–Metal System

Batch adsorption experiments were conducted to investigate the adsorption potential of E–CMBC to Cd2+, Cu2+ and Ni2+ in a single–metals system. The effects of pH on the adsorption capacity of E–CMBC to Cd2+, Cu2+ and Ni2+ were illustrated in Figure 1. To avoid the precipitation of heavy metal ions, the pH range of the solution were selected between 2–6. The adsorption capacities of E–CMBC to Cd2+ and Cu2+ were 39.75 mg g1, 30.72 mg g1 at pH 2, and the adsorption capacity increased with the raise of the pH, finally increased to 60.35 mg g1, 48.05 mg g1 at pH 6. It may be due to the stronger positive charge on the surface of E–CMBC at lower pH which will repel the metal cations [29]. With the pH increases (pHPZC 3.25), the surface of E–CMBC gradually changes from positively charged to negatively charged, which will provide more adsorption sites and improve the adsorption capacity to heavy metal ions [30]. However, the adsorption capacity to Ni2+ exhibited a slight increase from 37.75 mg g1 to 40.7 mg g1 when the pH increased from 2 to 6. It suggested that the adsorption of E–CMBC to Ni2+ is less affected by pH, and the electrostatic interactions is not the key role in the process of Ni2+ adsorption, we found the similar adsorption behaviour of E–CMBC to Pb2+ in single–metal system [28].
The influence of the contacting time on the adsorption capacity of E–CMBC to Cd2+, Cu2+ and Ni2+ were also investigated. As shown in Figure 2a, the adsorption capacity of E–CMBC to Cd2+, Cu2+ and Ni2+ within 60 min accounted for approximately 91.76%, 90.03%, and 91.89% of the equilibrium capacity. It may be attributed to the abundant available adsorption sites on the surface of the adsorbent at the beginning of the adsorption process. With the adsorption time increases, the available adsorption sites become less, therefore the adsorption rate slows down gradually.
To evaluate the adsorption rates and explore the adsorption mechanism, the adsorption kinetics of E–CMBC to Cd2+, Cu2+ and Ni2+ were fitted by three classic kinetic models, namely pseudo–first–order, pseudo–second–order, and Avrami fractional–order models. The formula and parameters are listed in Table S1 and the kinetic models fitting results are list in Table 1. The coefficient of determination (R2) of the Avrami fractional–order model for Cd2+, Cu2+ and Ni2+ was 0.99, 0.95, and 0.99, respectively, which were greater than that of pseudo–second–order and pseudo–first–order model for Cd2+ (0.98, 0.81), Cu2+ (0.88, 0.62), and Ni2+ (0.97, 0.82). According to the Avrami fractional–order model, the theoretical adsorption capacity of Cd2+, Cu2+ and Ni2+ were 59.57 mg g1, 47.64 mg g1 and 40.24 mg g1, which were approximately the experiment results of Cd2+ (60.35 mg g1), Cu2+ (48.05 mg g1) and Ni2+ (40.7 mg g1), respectively. Therefore, the Avrami fractional–order model could more accurately describe the adsorption process. All the above findings suggested that the adsorption of E–CMBC to Cd2+, Cu2+ and Ni2+ on was a multiple kinetics process, and the chemical adsorption played a dominating role [31,32].
The effect of the initial concentration of single Cd2+, Cu2+ and Ni2+ on adsorption capacity was shown in Figure 2b. Two widespread–used nonlinear isotherms including Langmuir and Freundlich were fitted by experimental data to evaluate the adsorption type and the adsorption capacity. The formula and parameters of the two isotherm models were shown in Table S2. Obviously, the adsorption capacity increased with the initial concentration of single Cd2+, Cu2+ and Ni2+ and gradually progresses to equilibrium. From Table 2, the R2 of the Langmuir model for Cd2+, Cu2+ and Ni2+ were higher than that of the Freundlich model. It was clear that the Langmuir model describes the adsorption behaviour better than the Freundlich model. This indicated that the adsorption of E–CMBC to Cd2+, Cu2+ and Ni2+ in the single–metal system mainly occurred in monolayer [25]. Further, the Langmuir maximum adsorption capacities of E–CMBC to Cd2+, Cu2+ and Ni2+ in the single–metal system ( q max 0 ) were higher than that of some previous reported adsorbents [33,34,35,36]. Furthermore, the Langmuir maximum adsorption capacity order is Cd2+ > Cu2+ > Ni2+ in single–metal system, which is different from the results of other researchers [37,38].

3.2. Adsorption of Cd2+, Cu2+ and Ni2+ in the Binary–Metal System

The competitive adsorption behaviour in binary–metal systems (Cd2+–Ni2+, Cu2+–Ni2+, Cd2+–Cu2+) were studied in the same initial concentrations. Some characteristic properties of Cd2+, Cu2+ and Ni2+, such as the hydrated radius, electronegativity, absolute hardness, and binding constant with EDTA, were listed in Table 3. The results showed that the maximum adsorption capacities of the three metals in the binary system ( q max mix ) were all lower than q max 0 . The characteristic properties of E–CMBC and the heavy metal ions may be responsible for the different performances [38].

3.2.1. Cd2+–Ni2+ System

The adsorption kinetics of Cd2+–Ni2+ on E–CMBC were showed in Figure 3a. For both Cd2+ and Ni2+, the adsorption capacities increased with the contact time and the adsorption process included a fast stage and a slow stage. The adsorption capacities of E–CMBC to Cd2+ and Ni2+ increased rapidly in the initial 60 min as there were plenty of binding sites on the surface of the adsorbent. After the binding sites were occupied by heavy metals, the adsorption gradually slows down and finally reaches the adsorption equilibrium. From Table S3, the experiment data fitted the pseudo–second–order and Avrami fractional–order model perfectly. It revealed that the adsorption of E–CMBC to Cd2+ and Ni2+ in Cd2+–Ni2+ systems was a multiple kinetics process too. The adsorption isotherms of E–CMBC to Cd2+ and Ni2+ in binary–metal systems were investigated and shown in Figure 3d, the parameters fitted by the Langmuir and Freundlich models were shown in Table S4. The R2 values (≥0.94) suggested that the Langmuir model provided a better fit than the Freundlich model for both Cd2+ and Ni2+ in Cd2+–Ni2+ systems. It demonstrated that Cd2+ and Ni2+ were absorbed to distinct sites, in which each site could only combine one ion [31]. Competition between Cd2+ and Ni2+ affected the adsorption capacity of each other on E–CMBC. The Langmuir maximum adsorption capacities of Cd2+ and Ni2+ decreased from 61.95 mg g−1 and 40.38 mg g1 in the single–metal systems to 32.63 mg g1 and 26.93 mg g1 in Cd2+–Ni2+ systems. The decrease in the amplitude of Cd2+ adsorption (47.33%) was larger than that of Ni2+ (33.31%), which suggesting that Cd2+ adsorption was more affected by Ni2+ competition. The resultant affinity of adsorption sites and chemical characteristics of the heavy metals may be responsible for this performance. A reasonable explanation for this result is that the hydrated radius of Ni2+ (4.04 Å) is smaller than that of Cd2+ (4.26 Å), that means Ni2+ has greater affinity for most functional groups on surface of E–CMBC (such as amino and carboxylic groups) [41]. Furthermore, the electronegativity of Ni2+ (1.91) is higher than Cd2+ (1.69), which ensure that Ni2+ is more favorably adsorbed than Cd2+. Ni2+ has the larger binding constant with EDTA compared with Cd2+ (Ni2+ logK = 18.7, Cd2+ logK = 16.5). Therefore, Ni2+ is more likely to combine with EDTA on the of the surface of E–CMBC.

3.2.2. Cu2+–Ni2+ System

The relationship between adsorption capacity with contact time was established in Cu2+–Ni2+ system. Figure 3b showed the quantity of Cu2+ and Ni2+ with time. The adsorption capacities of Cu2+ and Ni2+ increases with time, and the adsorption process also includes a fast and slow stage too (Figure 3b). The pseudo–second–order and Avrami fractional–order model fit the Cu2+ and Ni2+ adsorption data well in the Cu2+–Ni2+ system (Table S3). The effect of initial concentration on the adsorption capacity and selectivity was shown in Figure 3e. With the initial heavy metal concentration increased, the adsorption capacity of Cu2+ gradually increased and approached equilibrium. However, the adsorption capacity of Ni2+ increased at first and then decreased to equilibrium. At low initial concentration (≤20 mg L−1), the adsorption capacities of Cu2+ and Ni2+ in Cu2+–Ni2+ system were almost same to those in the single–metal systems. The results suggested that there were excess binding sites on E–CMBC at low heavy metal concentration. Cu2+ and Ni2+ could bind to different sites without competition. With the heavy metal concentration increased (>20 mg L1), the adsorption capacity of Cu2+ gradually increased and approached equilibrium adsorption capacity, but it was still lower than the q max 0 of Cu2+. However, the adsorption capacity of Ni2+ decreased to equilibrium with the increase of the Ni2+ concentrations. Similar results have also been reported by other researchers [38]. These indicated that the adsorption of Cu2+ and Ni2+ could be inhibited by each other at the high initial concentration (>20 mg L1). Although the hydrated radius is larger and the electronegativity and binding constant with EDTA are similar, Cu2+ have stronger competitive adsorption performance than Ni2+. It may be attributed to the lower polarizability and absolute hardness of Cu2+ than that of Ni2+ in the theory of hard and soft acids and bases [38].

3.2.3. Cu2+–Cd2+ System

The competitive adsorption of Cu2+ and Cd2+ within specified intervals were shown in Figure 3c. For Cu2+, adsorption increased rapidly in the initial 60 min, then showed only a small increment over the subsequent period, which fitted the Avrami fractional–order model better (Table S3). In contrast, the adsorption capacity of Cd2+ was initially reached a maximum, then gradually decreased with time and finally became zero. It could be attributed to the abundant sites on the surface of E–CMBC at the beginning for Cu2+ and Cd2+; subsequently, the initially adsorbed Cd2+ was replaced by Cu2+ in the solution. Similar results were reported by other researchers [38,39,41]. Figure 3f showed the competitive adsorption of Cu2+ and Cd2+ in specified concentrations. It was clear that the adsorption capacity of Cu2+ increased with the increasing equilibrium concentration. Compared with the single–metal system, the maximum adsorption capacities of Cu2+ hardly changes in Cu2+–Cd2+ systems. However, the adsorption capacity of Cd2+ increased at first and then gradually decreased to zero. The above phenomenon showed that the competitive adsorption and replacement would occur when the Cu2+ and Cd2+ reached certain concentration in the binary–metal system. Furthermore, the adsorption of Cd2+ was greatly inhibited by the coexisting Cu2+, but the adsorption of Cu2+ was slightly affected by Cd2+. The higher affinity of Cu2+ over Cd2+ on E–CMBC could be attributed to that: Cu2+ has a smaller hydrated radius (4.19 Å) and absolute hardness (8.3) than that of Cd2+ (4.26 Å and 10.3) but a higher electronegativity (1.90) and binding constant with EDTA (18.8) than that of Cd2+ (1.69 and 1.65).
A replacement experiment was carried out to verify the hypothesis that the adsorbed Cd2+ on E–CMBC might be replaced by Cu2+ in the binary–metal systems. The Cd2+ saturated E–CMBC sample was collected and labelled as Cd–loaded sample. The Cd–loaded samples were used to adsorb Cu2+ at the same conditions as batch experiments. A blank control experiment was carried out in de–ionized water with the same pH as Cu2+ solution. After adsorption equilibrium, the concentration of Cu2+ was decreased by 45.89 mg L1, and the concentration of Cd2+ was measured by 55.93 mg L1 (Figure 4). While there was almost no Cd2+ could be detected in the blank control experiment. The XPS analysis detected the of peak of Cu2+, while the peak of Cd2+ disappeared in the Cd–loaded sample after adsorption of Cu2+ (Figure 5b). These results further confirmed that Cd2+ on E–CMBC could be completely replaced by Cu2+ in the binary–metal system.

3.3. Analysis of Adsorption Mechanism

The FTIR was used to characterize the surface’s groups of E–CMBC before and after the adsorption of heavy metal ions (Figure 5a). The overlapping peak around 3416 cm−1 presented in E–CMBC, indicated the –OH and –NH2 stretching vibrational modes, this peak decreased to 3399 cm−1 (Cd2+), 3404 cm−1 (Cu2+) and 3385 cm−1 (Ni2+) after adsorption, which suggested that these groups had interacted with the heavy metal ions. The peaks appeared around 1630 cm−1 and 1403 cm−1 were related to the stretching vibration of C=O (–NHCO, –COOH). After the adsorption of heavy metal ions, these peaks shifted to low wavenumber, too. The results clearly revealed the interaction of the amides and carboxyl groups with the heavy metal ions during the adsorption process [42]. The oxygen–rich functional groups on the surface of E–CMBC, such as hydroxyl and carbonyl functional groups, could form organometallic complexes with heavy metals, through electrostatic surface and via inner–sphere complexation [43]. When the pH value of the solution was greater than Zeta potential, the adsorption capacity of heavy metal increased, which indicated that the electrostatic interaction with negatively charged functional groups on E–CMBC surface was also one of the adsorption mechanisms [31].
In Figure 5b, the heavy metal peaks (Cd3d, Cu2p, Ni2p) appeared on the full–range XPS spectra of E–CMBC after adsorption which indicated that E–CMBC had adequate sorption for heavy metal ions. The typical peak of Na1s at 1072.08 eV disappeared after adsorption, which indicated that the exchange reaction between Na+ and heavy metal ions occurred. In addition, physical adsorption was one of the common mechanisms for porous materials [21].
The possible mechanisms of heavy metal ions adsorption on E–CMBC in single–metal system can be summarized as follows: surface complexation; electrostatic interaction; ion exchange; physical adsorption. The isotherm, kinetic parameters for the adsorption of Cd2+, Cu2+ and Ni2+ indicated that the adsorption process was mainly chemical adsorption [44], which is consistent with the introduction of active groups on the surface of the modified biochar. The adsorption mechanism in the binary–metal system is similar to the single–metal system, while the characteristics of different heavy metal ions will lead to different competitive adsorption ability.

4. Conclusions

In this study, E–CMBC was successfully used as an effective adsorbent to adsorb Cd2+, Cu2+ and Ni2+. The Langmuir isotherm model and the Avrami fractional–order kinetics model provided the better fit for the experimental data of the single–metal system. The Langmuir maximum adsorption capacities of Cd2+, Cu2+ and Ni2+ were 61.95 mg g−1, 48.73 mg g1 and 40.38 mg g1, respectively. In a binary–metal system, the coexistence of the two metal ions will suppress the adsorption ability of each other due to the competitive adsorption. The order of the competitive adsorption ability was Cu2+ > Ni2+ > Cd2+. Moreover, the adsorbed Cd2+ could be replaced completely by Cu2+ in the solution, which shows E–CMBC has the potential to separate and purify Cd2+–Cu2+ mixed ions. These results indicate that the difference in competitive adsorption ability is related to the properties of E–CMBC and also to the characteristics of heavy metal ions. The mechanistic analysis demonstrates that the adsorption is a monolayer and multiple kinetics process, and chemical adsorption was the dominant mechanisms. Our findings suggest that E–CMBC has the potential to remove a variety of heavy metal ions from aqueous solutions.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/pr9101829/s1, Table S1: Adsorption kinetic models, equations and parameters for single and binary metal systems. Table S2: Adsorption isotherm models, equations and parameters for single and binary metal systems. Table S3: Adsorption kinetic model parameters for Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal system. Table S4: Adsorption isotherm model parameters for Cd2+ and Ni2+ on E–CMBC in Cd2+–Ni2+ system.

Author Contributions

Conceptualization, L.Z.; Data curation, J.D. and Y.H.; Formal analysis, Y.L.; Funding acquisition, Y.G.; Investigation, W.Z.; Methodology, L.Z., L.D. and X.P.; Project administration, L.Z.; Validation, Q.Z.; Writing—original draft, L.Z.; Writing—review & editing, Y.G., J.D., Y.L., R.N. and X.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Joint Funds of the National Natural Science Foundation of China (No. U1806217) and the Shandong Provincial Natural Science Foundation (No. ZR2019BD012).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, Z.; Wang, Z.; Wu, X.; Li, M.; Liu, X. Competitive adsorption of tylosin, sulfamethoxazole and Cu(II) onnano-hydroxyapatitemodified biochar in water. Chemosphere 2020, 240, 124884. [Google Scholar] [CrossRef]
  2. Munusamy, T.; Ya-Ting, J.; Jiunn-Fwu, L. Enhanced adsorption of inorganic and organicpollutants by amine modified sodium montmorillonite nanosheets. RSC Adv. 2015, 5, 20583. [Google Scholar] [CrossRef]
  3. Peiris, C.; Gunatilake, S.R.; Mlsna, T.E.; Mohan, D.; Vithanage, M. Biochar based removal of antibiotic sulfonamides and tetracyclines inaquatic environments: A critical review. Bioresour. Technol. 2017, 246, 150–159. [Google Scholar] [CrossRef] [PubMed]
  4. Fu, F.; Wang, Q. Removal of heavy metal ions fromwastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef] [PubMed]
  5. Mansoor, S.; Kour, N.; Manhas, S.; Zahid, S.; Wani, O.A.; Sharma, V.; Wijaya, L.; Alyemeni, M.N.; Alsahli, A.A.; El-Serehy, H.A.; et al. Biochar as a tool for effective management of drought and heavymetal toxicity. Chemosphere 2021, 271, 129458. [Google Scholar] [CrossRef] [PubMed]
  6. Pohl, A. Removal of heavy metal ions from water and wastewaters by sulfur-containing precipitation agents. Water Air Soil Pollut. 2020, 231, 503. [Google Scholar] [CrossRef]
  7. Pei, X.; Gan, L.; Tong, Z.; Gao, H.; Meng, S.; Zhang, W.; Wang, P.; Chen, Y. Robust cellulose-based composite adsorption membrane for heavymetal removal. J. Hazard. Mater. 2021, 406, 124746. [Google Scholar] [CrossRef]
  8. Yu, Y.; Zhong, Y.; Wang, M.; Guo, Z. Electrochemical behavior of aluminium anode in super-gravity field and its application in copper removal from wastewater by electrocoagulation. Chemosphere 2021, 272, 129614. [Google Scholar] [CrossRef]
  9. Wu, J.; Wang, T.; Wang, J.; Zhang, Y.; Pan, W. A novel modified method for the efficient removal of Pb and Cd from wastewater by biochar: Enhanced the ion exchange and precipitation capacity. Sci. Total Environ. 2021, 754, 142150. [Google Scholar] [CrossRef]
  10. Abbas, S.Z.; Rafatullah, M. Recent advances in soil microbial fuel cells for soil contaminants remediation. Chemosphere 2021, 272, 129691. [Google Scholar] [CrossRef]
  11. Won, S.W.; Kotte, P.; Wei, W.; Lim, A.; Yun, Y.S. Biosorbents for recovery of precious metals. Bioresour. Technol. 2014, 160, 203–212. [Google Scholar] [CrossRef]
  12. Zhao, F.; Repo, E.; Yin, D.; Sillanpää, M.E.T. Adsorption of Cd(II) and Pb(II) by a novel EGTA-modified chitosan material: Kinetics and isotherms. J. Colloid Interface Sci. 2013, 409, 174–182. [Google Scholar] [CrossRef]
  13. Wu, D.; Hu, L.; Wang, Y.; Wei, Q.; Yan, L.; Yan, T.; Li, Y.; Du, B. EDTA modified ß-cyclodextrin/chitosan for rapid removal of Pb(II) and acid red from aqueous solution. J. Colloid Interface Sci. 2018, 523, 56–64. [Google Scholar] [CrossRef]
  14. Oliveira, F.R.; Patel, A.K.; Jaisi, D.P.; Adhikari, S.; Lu, H.; Khanal, S.K. Environmental application of biochar: Current status and perspectives. Bioresour. Technol. 2017, 246, 110–222. [Google Scholar] [CrossRef]
  15. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U., Jr. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent—A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef]
  16. Wei, D.; Li, B.; Huang, H.; Lin, L.; Zhang, J.; Yang, Y.; Guo, J.; Tang, L.; Zeng, G.; Zhou, Y. Biochar-based functional materials in the purification of agricultural wastewater: Fabrication, application and future research needs. Chemosphere 2018, 197, 165–180. [Google Scholar] [CrossRef]
  17. Nguyen, T.A.; Ngo, H.H.; Guo, W.S.; Zhang, J.; Liang, S.; Yue, Q.Y. Applicability of agricultural waste and by-products for adsorptive removal of heavy metals from wastewater. Bioresour. Technol. 2013, 148, 574–585. [Google Scholar] [CrossRef]
  18. Sizmur, T.; Fresno, T.; Akgül, G.; Frost, H.; Moreno-Jiménez, E. Biochar modification to enhance sorption of inorganics from water. Bioresour. Technol. 2017, 246, 34–47. [Google Scholar] [CrossRef]
  19. Rajapaksha, A.U.; Chen, S.S.; Tsang DC, W.; Zhang, M.; Vithanage, M.; Mandal, S.; Gao, B.; Bolan, N.; Ok, Y.S. Engineered/designer biochar for contaminant removal/immobilization from soil and water: Potential and implication of biochar modification. Chemosphere 2016, 148, 276–291. [Google Scholar] [CrossRef]
  20. Li, R.; Liang, W.; Huang, H.; Jiang, S.; Guo, D.; Li, M.; Zhang, Z.; Ali, A.; Wang, J.J. Removal of cadmium(II) cations from an aqueous solution with aminothiourea chitosan strengthened magnetic biochar. J. Appl. Polym. Sci. 2018, 135, 46239. [Google Scholar] [CrossRef]
  21. Lv, D.; Liu, Y.; Zhou, J.; Yang, K.; Lou, Z.; Baig, S.A.; Xu, X. Application of EDTA-functionalized bamboo activated carbon (BAC) for Pb(II) and Cu(II) removal from aqueous solutions. Appl. Surf. Sci. 2018, 428, 648–658. [Google Scholar] [CrossRef]
  22. Sajjadi, B.; Broome, J.W.; Chen, W.Y.; Mattern, D.L.; Egiebor, N.O.; Hammer, N.; Smith, C.L. Urea functionalization of ultrasound-treated biochar: A feasible strategy for enhancing heavy metal adsorption capacity. Ultrason. Sonochem. 2019, 51, 20–30. [Google Scholar] [CrossRef] [PubMed]
  23. Guo, Z.; Chen, R.; Yang, R.; Yang, F.; Chen, J.; Li, Y.; Zhou, R.; Xu, J. Synthesis of amino-functionalized biochar/spinel ferrite magnetic composites for low-cost and efficient elimination of Ni(II) from wastewater. Sci. Total Environ. 2020, 722, 137822. [Google Scholar] [CrossRef] [PubMed]
  24. Song, Z.; Lian, F.; Yu, Z.; Zhu, L.; Xing, B.; Qiu, W. Synthesis and characterization of a novel MnOx-loaded biochar and its adsorption properties for Cu2+ in aqueous solution. Chem. Eng. J. 2014, 242, 36–42. [Google Scholar] [CrossRef]
  25. Zhou, X.; Zhou, J.; Liu, Y.; Guo, J.; Ren, J.; Zhou, F. Preparation of iminodiacetic acid-modified magnetic biochar by carbonization, magnetization and functional modification for Cd(II) removal in water. Fuel 2018, 233, 469–479. [Google Scholar] [CrossRef]
  26. Ni, B.J.; Huang, Q.S.; Wang, C.; Ni, T.Y.; Sun, J.; Wei, W. Competitive adsorption of heavy metalsin aqueous solution onto biochar derived from anaerobically digested sludge. Chemosphere 2019, 219, 351–357. [Google Scholar] [CrossRef]
  27. Deng, J.; Liu, Y.; Liu, S.; Zeng, G.; Tan, X.; Huang, B.; Tang, X.; Wang, S.; Hua, Q.; Yan, Z. Competitive adsorption of Pb(II), Cd(II) and Cu(II) onto chitosan-pyromellitic dianhydride modified biochar. J. Colloid Interface Sci. 2017, 506, 355–364. [Google Scholar] [CrossRef]
  28. Zheng, L.; Gao, Y.; Du, J.; Zhang, W.; Huang, Y.; Wang, L.; Zhao, Q.; Pan, X. A novel, recyclable magnetic biochar modified by chitosan-EDTA for the effective removal of Pb(II) from aqueous solution. RSC Adv. 2020, 10, 40196. [Google Scholar] [CrossRef]
  29. Liu, Y.; Wang, L.; Wang, X.; Jing, F.; Chang, R.; Chen, J. Oxidative ageing of biochar and hydrochar alleviating competitive sorption of Cd(II) and Cu(II). Sci. Total Environ. 2020, 725, 138419. [Google Scholar] [CrossRef]
  30. 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]
  31. Chen, Y.; Li, M.; Li, Y.; Liu, Y.; Chen, Y.; Li, H.; Li, L.; Xu, F.; Jiang, H.; Chen, L. Hydroxyapatite modified sludge-based biochar for the adsorption of Cu2+and Cd2+: Adsorption behavior and mechanisms. Bioresour. Technol. 2021, 321, 124413. [Google Scholar] [CrossRef]
  32. Qu, J.; Liu, Y.; Cheng, L.; Jiang, Z.; Zhang, G.; Deng, F.; Wang, L.; Han, W.; Zhang, Y. Green synthesis of hydrophilic activated carbon supported sulfide nZVI for enhanced Pb(II) scavenging from water: Characterization, kinetics, isotherms and mechanisms. J. Hazard. Mater. 2021, 403, 123607. [Google Scholar] [CrossRef]
  33. Shi, Y.; Zhang, T.; Ren, H.; Kruse, A.; Cui, R. Polyethylene imine modified hydrochar adsorption for chromium (VI) and nickel (II) removal from aqueous solution. Bioresour. Technol. 2018, 247, 370–379. [Google Scholar] [CrossRef]
  34. Wang, Y.; Liu, R. H2O2 treatment enhanced the heavy metals removal by manure biocharin aqueous solutions. Sci. Total Environ. 2018, 628–629, 1139–1148. [Google Scholar] [CrossRef]
  35. Xu, X.; Cao, X.; Zhao, L.; Wang, H.; Yu, H.; Gao, B. Removal of Cu, Zn, and Cd from aqueous solutions by the dairy manure-derived biochar. Environ. Sci. Pollut. Res. 2013, 20, 358–368. [Google Scholar] [CrossRef]
  36. Yang, G.; Jiang, H. Amino modification of biochar for enhanced adsorption of copper ions from synthetic wastewater. Water Res. 2014, 48, 396–405. [Google Scholar] [CrossRef]
  37. Liu, R.; Lian, R. Non-competitive and competitive adsorption of Cd2+, Ni2+, and Cu2+ by biogenic vaterite. Sci. Total Environ. 2019, 659, 122–130. [Google Scholar] [CrossRef]
  38. Tan, P.; Hu, Y.; Bi, Q. Competitive adsorption of Cu2+, Cd2+and Ni2+from an aqueous solution on graphene oxide membranes. Colloids Surf. A Physicochem. Eng. Asp. 2016, 509, 56–64. [Google Scholar] [CrossRef]
  39. Ahmad, Z.; Gao, B.; Mosa, A.; Yu, H.; Yin, X.; Bashir, A.; Ghoveisi, H.; Wang, S. Removal of Cu(II), Cd(II) and Pb(II) ions from aqueous solutions by biochars derived from potassium-rich biomass. J. Clean. Prod. 2018, 180, 437–449. [Google Scholar] [CrossRef]
  40. Nightingale, E.R., Jr. Phenomenological theory of ion solvation. Effective radii of hydratedions. J. Phys. Chem. 1959, 63, 1381–1387. [Google Scholar] [CrossRef]
  41. Park, J.H.; Ok, Y.S.; Kim, S.H.; Cho, J.S.; Heo, J.S.; Delaune, R.D.; Seo, D.C. Competitive adsorption of heavy metals onto sesame straw biochar in aqueous solutions. Chemosphere 2016, 142, 77–83. [Google Scholar] [CrossRef]
  42. Liu, Y.; Fu, R.; Sun, Y.; Zhou, X.; Baig, S.A.; Xu, X. Multifunctional nanocomposites Fe3O4@SiO2-EDTA for Pb (II) and Cu (II) removal from aqueous solutions. Appl. Surf. Sci. 2016, 369, 267–276. [Google Scholar] [CrossRef]
  43. Shan, R.; Shi, Y.; Gu, Y.; Wang, Y.; Yuan, H. Single and competitive adsorption affinity of heavy metals toward peanut shell-derived biochar and its mechanisms in aqueous systems. Chin. J. Chem. Eng. 2020, 28, 1375–1383. [Google Scholar] [CrossRef]
  44. Wu, S.; Wang, F.; Yuan, H.; Zhang, L.; Mao, S.; Liu, X.; Alharbi, N.S.; Rohani, S.; Lu, J. Fabrication of xanthate-modifiedchitosan/poly (N-isopropylacrylamide) composite hydrogel for the selective adsorption of Cu (II), Pb (II)and Ni (II) metal ions. Chem. Eng. Res. Des. 2018, 139, 197–210. [Google Scholar] [CrossRef]
Figure 1. Effect of initial solution pH and zeta potential on Cd2+, Cu2+ and Ni2+ for E–CMBC.
Figure 1. Effect of initial solution pH and zeta potential on Cd2+, Cu2+ and Ni2+ for E–CMBC.
Processes 09 01829 g001
Figure 2. (a) Adsorption kinetics of Cd2+, Cu2+ and Ni2+ on E–CMBC in single–metal system; (b) adsorption isotherms of Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Figure 2. (a) Adsorption kinetics of Cd2+, Cu2+ and Ni2+ on E–CMBC in single–metal system; (b) adsorption isotherms of Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Processes 09 01829 g002
Figure 3. Adsorption kinetics of Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal system: (a) Cd2+–Ni2+ system; (b) Cu2+–Ni2+ system; (c) Cd2+–Cu2+ system. Adsorption isotherms of Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal system; (d) Cd2+–Ni2+ system; (e) Cu2+–Ni2+ system; (f) Cd2+–Cu2+ system.
Figure 3. Adsorption kinetics of Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal system: (a) Cd2+–Ni2+ system; (b) Cu2+–Ni2+ system; (c) Cd2+–Cu2+ system. Adsorption isotherms of Cd2+, Cu2+ and Ni2+ on E–CMBC in binary–metal system; (d) Cd2+–Ni2+ system; (e) Cu2+–Ni2+ system; (f) Cd2+–Cu2+ system.
Processes 09 01829 g003
Figure 4. Replacement adsorption between Cu2+ and Cd2+ on Cd–loaded E–CMBC in Cu2+ solution.
Figure 4. Replacement adsorption between Cu2+ and Cd2+ on Cd–loaded E–CMBC in Cu2+ solution.
Processes 09 01829 g004
Figure 5. (a) FTIR spectra of E–-CMBC before and after adsorption of Cd2+, Cu2+ and Ni2+; (b) XPS spectra of E–CMBC before and after adsorption of Cd2+, Cu2+, Ni2+ and replacement adsorption of Cd2+–Cu2+.
Figure 5. (a) FTIR spectra of E–-CMBC before and after adsorption of Cd2+, Cu2+ and Ni2+; (b) XPS spectra of E–CMBC before and after adsorption of Cd2+, Cu2+, Ni2+ and replacement adsorption of Cd2+–Cu2+.
Processes 09 01829 g005
Table 1. Adsorption kinetic model parameters for Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Table 1. Adsorption kinetic model parameters for Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Metal IonPseudo–First–Order ModelPseudo–Second–Order ModelAvrami Fractional–Order Model
Cd2+qe (mg g−1)57.0qe (mg g−1)59.0qe (mg g−1)59.75
k1 (min−1)0.17k2 (g mg−1 min−1)0.0048k3 (min−1)0.17
R20.81R20.98R20.99
Cu2+qe (mg g−1)45.0qe (mg g−1)46.3qe (mg g−1)47.64
k1 (min−1)0.22k2 (g mg−1 min−1)0.0089k3 (min−1)0.36
R20.62R20.88R20.95
Ni2+qe (mg g−1)38.48qe (mg g−1)39.84qe (mg g−1)40.24
k1 (min−1)0.17k2 (g mg−1 min−1)0.0072k3 (min−1)0.18
R20.82R20.97R20.99
Table 2. Adsorption isotherm model parameters for Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Table 2. Adsorption isotherm model parameters for Cd2+, Cu2+ and Ni2+ on E–CMBC in the single–metal system.
Metal IonLangmuir ModelFreundlich Model
qm (mg g−1)KL (L mg−1)R2KF (L mg−1)nR2
Cd2+61.953.240.9836.350.110.62
Cu2+48.731.720.9727.890.110.68
Ni2+40.386.610.9626.150.0940.73
Table 3. The properties of Cd2+, Cu2+ and Ni2+ [38,39,40].
Table 3. The properties of Cd2+, Cu2+ and Ni2+ [38,39,40].
Metal IonHydrated Radius (Å)Electronegativity (Pauling)Absolute HardnessBinding Constant with EDTA
Cd2+4.261.6910.316.5
Cu2+4.191.908.318.8
Ni2+4.041.918.518.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, L.; Gao, Y.; Du, J.; Zhang, W.; Huang, Y.; Zhao, Q.; Duan, L.; Liu, Y.; Naidu, R.; Pan, X. Single and Binary Adsorption Behaviour and Mechanisms of Cd2+, Cu2+ and Ni2+ onto Modified Biochar in Aqueous Solutions. Processes 2021, 9, 1829. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101829

AMA Style

Zheng L, Gao Y, Du J, Zhang W, Huang Y, Zhao Q, Duan L, Liu Y, Naidu R, Pan X. Single and Binary Adsorption Behaviour and Mechanisms of Cd2+, Cu2+ and Ni2+ onto Modified Biochar in Aqueous Solutions. Processes. 2021; 9(10):1829. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101829

Chicago/Turabian Style

Zheng, Liwen, Yongchao Gao, Jianhua Du, Wen Zhang, Yujie Huang, Qingqing Zhao, Luchun Duan, Yanju Liu, Ravi Naidu, and Xiangliang Pan. 2021. "Single and Binary Adsorption Behaviour and Mechanisms of Cd2+, Cu2+ and Ni2+ onto Modified Biochar in Aqueous Solutions" Processes 9, no. 10: 1829. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9101829

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